paper_id
stringlengths
9
16
version
stringclasses
26 values
yymm
stringclasses
311 values
created
timestamp[s]
title
stringlengths
6
335
secondary_subfield
sequencelengths
1
8
abstract
stringlengths
25
3.93k
primary_subfield
stringclasses
124 values
field
stringclasses
20 values
fulltext
stringlengths
0
2.84M
1403.4098
2
1403
2015-03-19T08:59:39
Gamma stability in free product von Neumann algebras
[ "math.OA" ]
Let $(M, \varphi) = (M_1, \varphi_1) \ast (M_2, \varphi_2)$ be a free product of arbitrary von Neumann algebras endowed with faithful normal states. Assume that the centralizer $M_1^{\varphi_1}$ is diffuse. We first show that any intermediate subalgebra $M_1 \subset Q \subset M$ which has nontrivial central sequences in $M$ is necessarily equal to $M_1$. Then we obtain a general structural result for all the intermediate subalgebras $M_1 \subset Q \subset M$ with expectation. We deduce that any diffuse amenable von Neumann algebra can be concretely realized as a maximal amenable subalgebra with expectation inside a full nonamenable type III$_1$ factor. This provides the first class of concrete maximal amenable subalgebras in the framework of type III factors. We finally strengthen all these results in the case of tracial free product von Neumann algebras.
math.OA
math
GAMMA STABILITY IN FREE PRODUCT VON NEUMANN ALGEBRAS CYRIL HOUDAYER Abstract. Let (M, ϕ) = (M1, ϕ1) ∗ (M2, ϕ2) be a free product of arbitrary von Neumann algebras endowed with faithful normal states. Assume that the centralizer M ϕ1 is diffuse. We first show that any intermediate subalgebra M1 ⊂ Q ⊂ M which has nontrivial central sequences in M is necessarily equal to M1. Then we obtain a general structural result for all the intermediate subalgebras M1 ⊂ Q ⊂ M with expectation. We deduce that any diffuse amenable von Neumann algebra can be concretely realized as a maximal amenable subalgebra with expectation inside a full nonamenable type III1 factor. This provides the first class of concrete maximal amenable subalgebras in the framework of type III factors. We finally strengthen all these results in the case of tracial free product von Neumann algebras. 1 1. Introduction and statement of the main results A von Neumann algebra M ⊂ B(H) (with separable predual) is amenable if there exists a norm one projection E : B(H) → M . By Connes' celebrated result [Co75b], all the amenable von Neumann algebras are hyperfinite. Moreover, the amenable or hyperfinite factors are completely classified by their flows of weights (see [Co72, Co75b, Co85, Ha84]). In particular, there is a unique amenable II1 factor [Co75b]: it is the hyperfinite II1 factor of Murray and von Neumann [MvN43]. Since the amenable von Neumann algebras form a monotone class, any von Neumann algebra admits maximal amenable subalgebras. The first concrete examples of maximal amenable subalgebras inside II1 factors were obtained by Popa in [Po83]. He showed that any generator masa A in a free group factor L(Fn) with n ≥ 2 is maximal amenable. This result answered in the negative a question raised by Kadison. Indeed, A ⊂ L(Fn) is an abelian subalgebra generated by a selfadjoint operator and yet there is no intermediate hyperfinite subfactor in L(Fn) which contains A as a subalgebra. Popa discovered in [Po83] a powerful method to prove that a given amenable subalgebra is maximal amenable inside an ambient II1 factor. Using this strategy for the generator masa A ⊂ L(Fn), he first showed that A satisfies a certain asymptotic orthogonality property and then deduced that A is maximal amenable in L(Fn) using various mixing techniques. His results actually showed that the generator masa A is maximal Gamma inside L(Fn). Recall that a II1 factor M (with separable predual) has property Gamma of Murray and von Neumann [MvN43] if there exists a sequence of unitaries un ∈ U (M ) such that limn→∞ τ (un) = 0 and limn→∞ kxun − unxk2 = 0 for all x ∈ M . Subsequently, Cameron, Fang, Ravichandran and White proved in [CFRW08] that the radial masa in a free group factor L(Fn) with 2 ≤ n < ∞ is maximal amenable. Recently, the author vastly generalized in [Ho12a, Ho12b] Popa's results from [Po83] and obtained many new examples of maximal amenable subalgebras inside the crossed product II1 factors associated with free Bogoljubov actions of amenable groups. Very recently, Boutonnet and Carderi showed in [BC13] that any infinite maximal amenable subgroup Λ in a Gromov word-hyperbolic group Γ gives rise to a maximal amenable subalgebra L(Λ) inside the group von Neumann algebra 2010 Mathematics Subject Classification. 46L10; 46L54. Key words and phrases. Free product von Neumann algebras; Ultraproduct von Neumann algebras; Amenable von Neumann algebras; Type III factors; Property Gamma; Asymptotic orthogonality property. Research supported by ANR grant NEUMANN. 1 2 CYRIL HOUDAYER L(Γ). For other related results regarding maximal amenability in the framework of II1 factors, we refer the reader to [Br12, Fa06, Ga09, Ge95, Jo10, Po13, Sh05]. In this paper, we obtain new results regarding maximal amenability and Gamma stability for subalgebras of free products of arbitrary von Neumann algebras. We will be particularly interested in the structure of free product type III factors. Before stating our main results, we first introduce some terminology. Recall that a von Neumann algebra M is diffuse if M has no minimal projection. We say that a von Neumann subalgebra Q ⊂ M is with expectation if there exists a faithful normal conditional expectation EQ : M → Q. Let now ω ∈ β(N) \ N be a non-principal ultrafilter. We say that a von Neumann algebra M has property Gamma if the central sequence algebra M′ ∩ M ω is diffuse. Observe that in the case when M is a II1 factor with separable predual, this definition is equivalent to the property Gamma of Murray and von Neumann [MvN43] (see e.g. [Co74, Corollary 3.8]). Our first main result deals with Gamma stability inside arbitrary free product von Neumann algebras (M, ϕ) = (M1, ϕ1) ∗ (M2, ϕ2). We show in Theorem A that the subalgebra M1 ⊂ M sits in a very rigid position with respect to taking central sequences inside M . Theorem A. Let (M1, ϕ1) and (M2, ϕ2) be σ-finite von Neumann algebras endowed with faithful normal states. Assume that the centralizer M ϕ1 is diffuse. Denote by (M, ϕ) = 1 (M1, ϕ1) ∗ (M2, ϕ2) the free product von Neumann algebra. Then the inclusion M1 ⊂ M is Gamma stable in the following sense: for every intermediate von Neumann subalgebra M1 ⊂ Q ⊂ M such that Q′ ∩ M ω is diffuse, we have Q = M1. It is worth noticing that in the statement of Theorem A, the intermediate subalgebra M1 ⊂ Q ⊂ M is not assumed a priori to be with expectation in M . The proof of Theorem A is based on a key result (see Theorem 3.1) which is a generalization of Popa's result [Po83, Lemma 2.1] regarding asymptotic orthogonality for free group factors to arbitrary free product von Neumann algebras. The proof uses Popa's original method together with ε-orthogonality techniques from [Ho12a, Ho12b]. In order to obtain structural results for the intermediate subalgebras M1 ⊂ Q ⊂ M , we will next assume that Q is with expectation in M in the statement of Corollary B. Recall that a factor M (with separable predual) is full if its asymptotic centralizer Mω is trivial (see [Co74]). Observe that by [AH12, Theorem 5.3], this is equivalent to M′ ∩ M ω = C. Corollary B. Let (M1, ϕ1) and (M2, ϕ2) be von Neumann algebras with separable predual endowed with faithful normal states. Assume that the centralizer M ϕ1 is diffuse. Denote by 1 (M, ϕ) = (M1, ϕ1) ∗ (M2, ϕ2) the free product von Neumann algebra. Then any intermediate von Neumann subalgebra M1 ⊂ Q ⊂ M with faithful normal conditional expectation EQ : M → Q is globally invariant under the modular automorphism group (σϕ t ). Moreover, there exists a sequence of pairwise orthogonal projections zn ∈ Q′ ∩ M ⊂ Z(M1) such that Pn zn = 1 and • M1z0 = Qz0 and • Qzn is a full nonamenable factor such that (Qzn)′ ∩ (znM zn)ω = Czn for every n ≥ 1. Corollary B generalizes and strengthens [Po83, Lemma 3.1] and [Ge95, Lemma 4.4]. Corollary B moreover implies that if M1 has property Gamma, then M1 ⊂ M is a maximal Gamma subalgebra with expectation in M . The structural result in Corollary B allows us to obtain a wide range of maximal amenable subalgebras inside nonamenable factors. In particular, Corollary C below provides the first class of concrete maximal amenable subalgebras with expectation in the framework of type III factors. GAMMA STABILITY IN FREE PRODUCT VON NEUMANN ALGEBRAS 3 Corollary C. Any diffuse amenable von Neumann algebra with separable predual can be con- cretely realized as a maximal amenable subalgebra with expectation inside a full nonamenable type III1 factor. Our main last result deals with Gamma stability for subalgebras of tracial free product von Neumann algebras. Theorem D below is a further generalization of Corollary B where the subalgebra Q ⊂ M is only assumed to have a diffuse intersection with M1. Theorem D. Let (M1, τ1) and (M2, τ2) be von Neumann algebras with separable predual en- dowed with faithful normal tracial states. Assume that M1 is diffuse. Denote by (M, τ ) = (M1, τ1) ∗ (M2, τ2) the tracial free product von Neumann algebra. Then for every von Neumann subalgebra Q ⊂ M such that Q ∩ M1 is diffuse, there exists a central projection z ∈ Z(Q′ ∩ M ) ∩ Z(Q′ ∩ M ω) ⊂ M1 such that • Qz ⊂ zM1z and • (Q′ ∩ M ω)(1 − z) = (Q′ ∩ M )(1 − z) is discrete. Theorem D shows in particular that whenever Q ⊂ M is a subalgebra such that both Q ∩ M1 and Q′ ∩ M ω are diffuse, then Q ⊂ M1 (see Theorem 4.1). This is a strengthening of the Gamma stability result in Theorem A. Besides the asymptotic orthogonality property obtained in Theorem 3.1, the proof of Theorem D uses two more ingredients of II1 factors: Popa's intertwining techniques [Po01, Po03] and Peterson's L2-rigidity results for tracial free product von Neumann algebras [Pe06]. In Section 2, we recall a few preliminaires on free product and ultraproduct von Neumann algebras. In Section 3, we prove the key result regarding asymptotic orthogonality inside free products of arbitrary von Neumann algebras. Finally, we prove in Section 4 the main results of the paper. Acknowledgments. I am grateful to R´emi Boutonnet and Sven Raum for their valuable comments regarding a first draft of this manuscript. I especially thank R´emi for pointing out a gap in the initial proof of Proposition 2.5. Finally, I thank the referee for carefully reading the paper and useful remarks. 2. Preliminaries We fix once and for all a non-principal ultrafilter ω ∈ β(N)\ N. All the von Neumann algebras that we consider in this paper are assumed to be σ-finite, that is, countably decomposable. We say that M is a tracial von Neumann algebra if M admits a faithful normal tracial state τ . Background on σ-finite von Neumann algebras. Let M be any σ-finite von Neumann algebra. We denote by Ball(M ) the unit ball of M with respect to the uniform norm k · k∞, U (M ) the group of unitaries in M and Z(M ) the center of M . Let ϕ ∈ M∗ be a faithful normal state. We denote by L2(M, ϕ) (or simply L2(M ) when no confusion is possible) the GNS L2-completion of M with respect to the inner product defined by hx, yiϕ = ϕ(y∗x) for all x, y ∈ M . We denote by Λϕ : M → L2(M ) : x 7→ Λϕ(x) the canonical embedding and by Jϕ : L2(M ) → L2(M ) the canonical conjugation. We have xΛϕ(y) = Λϕ(xy) for all x, y ∈ M . We say that two elements x, y ∈ M are ϕ-orthogonal in M if ϕ(y∗x) = 0 or equivalently if the vectors Λϕ(x) and Λϕ(y) are orthogonal in the Hilbert space L2(M ). For all x ∈ M , write kxkϕ = ϕ(x∗x)1/2 and kxk♯ ϕ = ϕ(x∗x+xx∗)1/2. Recall that the strong (resp. ∗-strong) topology on uniformly bounded subsets of M coincides with the topology defined by k · kϕ (resp. k · k♯ ϕ). 4 CYRIL HOUDAYER An element x ∈ M is said to be analytic with respect to the modular automorphism group (σϕ t ) if the function R → M : t 7→ σϕ t (x) can be extended to an M -valued entire analytic function over C. We will be using the following standard facts. Proposition 2.1. Let (M, ϕ) be any σ-finite von Neumann algebra endowed with a faithful normal state. modular automorphism group (σϕ (1) The subset A ⊂ M of all the elements in M which are analytic with respect to the t ) forms a unital σ-strongly dense ∗-subalgebra of M . (2) For all a ∈ A and all x ∈ M , we have Λϕ(xa) = Jϕσϕ (3) For all a ∈ A and all x ∈ M , we have −i/2(a∗)Jϕ Λϕ(x). ϕ(ax) = ϕ(xσϕ −i(a)). In particular, for all a ∈ A and all x, y ∈ M , we have that xa and y are ϕ-orthogonal in M if and only if x and yσϕ i (a)∗ are ϕ-orthogonal in M . Proof. (1) follows from [Ta03, Lemma VIII.2.3] and (2) follows from [Ta03, Lemma VIII.3.10]. Let us prove (3). For every a ∈ A and every x ∈ M , we have −i(a)), Λϕ(1)iϕ −i(a)) = hΛϕ(xσϕ ϕ(xσϕ i/2(a∗)Jϕ Λϕ(x), Λϕ(1)iϕ −i/2(a)Jϕ Λϕ(1)iϕ = hJϕσϕ = hΛϕ(x), Jϕσϕ = hΛϕ(x), Λϕ(a∗)iϕ = ϕ(ax). In particular, for all a ∈ A and all x, y ∈ M , we have ϕ((yσϕ i (a)∗)∗ x) = ϕ(σϕ i (a) y∗x) = ϕ(y∗ xa). Hence xa and y are ϕ-orthogonal in M if and only if x and yσϕ i (a)∗ are ϕ-orthogonal in M . (cid:3) Proposition 2.2. Let M be any σ-finite von Neumann algebra. (1) We have that M is diffuse if and only if there exists a faithful normal state ϕ ∈ M∗ such that the centralizer M ϕ is diffuse. Moreover in that case, there exists a unitary u ∈ U (M ϕ) such that uk → 0 weakly as k → ∞. (2) Let N ⊂ M be a von Neumann subalgebra with expectation. If N is diffuse, so is M . Proof. (1) Assume first that M is diffuse. There exists a sequence of pairwise orthogonal projections zn ∈ Z(M ) such that Pn zn = 1, M z0 is a von Neumann algebra with a diffuse center and M zn is a diffuse factor for every n ≥ 1. Choose any faithful normal state ϕ0 on M z0. By [HS90, Theorem 11.1], for every n ≥ 1, choose a faithful normal state ϕn on M zn such that the centralizer (M zn)ϕn is diffuse. Let (an)n be a sequence of positive reals so that Pn an = 1. The formula ϕ = Pn anϕn defines a faithful normal state on M such that M ϕ = Mn (M zn)ϕn. Therefore, M ϕ is diffuse. Assume next that M ϕ is diffuse for some faithful normal state ϕ ∈ M∗. Using the above decomposition, for every n ≥ 1 such that zn 6= 0, letting ϕn = 1 ϕ(zn) ϕ(·zn), we have that (M zn)ϕn = M ϕzn is diffuse. Therefore M zn is a non-type I factor and so is diffuse. Thus, M is diffuse. GAMMA STABILITY IN FREE PRODUCT VON NEUMANN ALGEBRAS 5 When M ϕ is diffuse, take A ⊂ M ϕ a maximal abelian subalgebra. Then A is necessarily diffuse. Then choose a diffuse subalgebra B ⊂ A with separable predual. Since B ∼= L∞(T), we can then take a unitary u ∈ U (B) such that uk → 0 weakly as k → ∞. (2) Denote by E : M → N a faithful normal conditional expectation and choose a faithful normal state ψ ∈ N∗ such that N ψ is diffuse. Then ϕ = ψ ◦ E is a faithful normal state on M such that N ψ ⊂ M ϕ. Since N ψ is diffuse and M ϕ is tracial, M ϕ is diffuse and so is M by item (1) of the proposition. (cid:3) Free product von Neumann algebras. For i = 1, 2, let (Mi, ϕi) be any σ-finite von Neu- mann algebra endowed with a faithful normal state. The free product von Neumann algebra (M, ϕ) = (M1, ϕ1) ∗ (M2, ϕ2) is the von Neumann algebra M generated by M1 and M2 where the faithful normal state ϕ satisfies the following freeness condition: ϕ(x1 ··· xn) = 0 whenever xj ∈ Mij ⊖ C and i1 6= ··· 6= in. Here and in what follows, we denote by Mi ⊖ C = ker(ϕi). We refer to the product x1 ··· xn where xj ∈ Mij ⊖ C and i1 6= ··· 6= in as a reduced word in (Mi1 ⊖ C)··· (Min ⊖ C) of length n ≥ 1. The linear span of 1 and of all the reduced words in (Mi1 ⊖ C)··· (Min ⊖ C) where n ≥ 1 and i1 6= ··· 6= in forms a unital σ-strongly dense ∗-subalgebra of M . For all n ≥ 1 and all i1 6= ··· 6= in, the mapping L2((Mi1 ⊖ C)··· (Min ⊖ C), ϕ) → L2(Mi1 ⊖ C, ϕi1) ⊗ ··· ⊗ L2(Min ⊖ C, ϕin) Λϕ(x1 ··· xn) 7→ Λϕi1 defines a unitary operator. Moreover, we have (x1) ⊗ ··· ⊗ Λϕin (xn) L2(M, ϕ) = C ⊕Mn≥1 Mi16=···6=in t ∗σϕ2 t = σϕ1 L2(Mi1 ⊖ C, ϕi1 ) ⊗ ··· ⊗ L2(Min ⊖ C, ϕin). For all t ∈ R, we have σϕ (see [Ba93, Lemma 1] and [Dy92, Theorem 1]). By [Ta03, Theorem IX.4.2], there exists a unique ϕ-preserving faithful normal conditional expectation EM1 : M → M1. Moreover, we have EM1(x1 ··· xn) = 0 for all the reduced words x1 ··· xn which contains at least one letter from M2 ⊖ C (see [Ue11, Lemma 2.1]). For more on free product von Neumann algebras, we refer the reader to [Ue98, Ue11, Vo85, Vo92]. t Ultraproduct von Neumann algebras. Let M be any σ-finite von Neumann algebra. De- fine I ω(M ) = {(xn)n ∈ ℓ∞(N, M ) : xn → 0 ∗ −strongly as n → ω} Mω(M ) = {(xn)n ∈ ℓ∞(N, M ) : (xn)n I ω(M ) ⊂ I ω(M ) and I ω(M ) (xn)n ⊂ I ω(M )} . We have that the multiplier algebra Mω(M ) is a C∗-algebra and I ω(M ) ⊂ Mω(M ) is a norm closed two-sided ideal. Following [Oc85], we define the ultraproduct von Neumann algebra M ω by M ω = Mω(M )/I ω(M ). We denote the image of (xn)n ∈ Mω(M ) by (xn)ω ∈ M ω. For all x ∈ M , the constant sequence (x)n lies in the multiplier algebra Mω(M ). We will then identify M with (M +I ω(M ))/I ω(M ) and regard M ⊂ M ω as a von Neumann subalgebra. The map EM : M ω → M : (xn)ω 7→ σ-weak limn→ω xn is a faithful normal conditional expectation. For every faithful normal state ϕ ∈ M∗, the formula ϕω = ϕ ◦ EM defines a faithful normal state on M ω. Observe that ϕω((xn)ω) = limn→ω ϕ(xn) for all (xn)ω ∈ M ω. Let Q ⊂ M be any von Neumann subalgebra with faithful normal conditional expectation EQ : M → Q. Choose a faithful normal state ϕ on Q and still denote by ϕ the faithful normal state ϕ ◦ EQ on M . We have ℓ∞(N, Q) ⊂ ℓ∞(N, M ), I ω(Q) ⊂ I ω(M ) and Mω(Q) ⊂ Mω(M ). We will then identify Qω = Mω(Q)/I ω(Q) with (Mω(Q) + I ω(M ))/I ω(M ) and regard Qω ⊂ M ω as a von Neumann subalgebra. Observe that the norm k · k(ϕQ)ω on Qω is 6 CYRIL HOUDAYER the restriction of the norm k · kϕω to Qω. Observe moreover that (EQ(xn))n ∈ I ω(Q) for all (xn)n ∈ I ω(M ) and (EQ(xn))n ∈ Mω(Q) for all (xn)n ∈ Mω(M ). Therefore, the mapping EQω : M ω → Qω : (xn)ω 7→ (EQ(xn))ω is a well-defined conditional expectation satisfying ϕω ◦ EQω = ϕω. Hence, EQω : M ω → Qω is a faithful normal conditional expectation. Put H = L2(M, ϕ). The ultraproduct Hilbert space Hω is defined to be the quotient of ℓ∞(N,H) by the subspace consisting in sequences (ξn)n satisfying limn→ω kξnkH = 0. We denote the im- age of (ξn)n ∈ ℓ∞(N,H) by (ξn)ω ∈ Hω. The inner product space structure on the Hilbert space Hω is defined by h(ξn)ω, (ηn)ωiHω = limn→ωhξn, ηniH. The GNS Hilbert space L2(M ω, ϕω) can be embedded into Hω as a closed subspace by Λϕω ((xn)ω) 7→ (Λϕ(xn))ω. For more on ultra- product von Neumann algebras, we refer the reader to [AH12, Oc85]. Put xϕ = ϕ(·x) and ϕx = ϕ(x·) for all x ∈ M and all ϕ ∈ M∗. We will be using the following standard facts. Lemma 2.3. Let (M, ϕ) be any σ-finite von Neumann algebra endowed with a faithful normal state. Then for every x ∈ M , we have kxϕk ≤ kxkϕ, kϕxk ≤ kx∗kϕ and kxϕ − ϕxk = kx∗ϕ − ϕx∗k. Proof. Let x ∈ M . Using the Cauchy-Schwarz inequality, for all y ∈ Ball(M ), we have (xϕ)(y) = ϕ(yx) ≤ ky∗kϕ kxkϕ ≤ kxkϕ and hence kxϕk ≤ kxkϕ. Likewise, for all y ∈ Ball(M ), we have (ϕx)(y) = ϕ(xy) ≤ kx∗kϕ kykϕ ≤ kx∗kϕ and hence kϕxk ≤ kx∗kϕ. Moreover, for all y ∈ Ball(M ), we have (x∗ϕ − ϕx∗)(y) = ϕ(yx∗ − x∗y) = ϕ(yx∗ − x∗y) = ϕ(xy∗ − y∗x) = (xϕ − ϕx)(y∗). This implies that kxϕ − ϕxk = kx∗ϕ − ϕx∗k. Proposition 2.4. Let (M, ϕ) be any σ-finite von Neumann algebra endowed with a faithful normal state. (cid:3) (1) For every (xn)n ∈ Mω(M ) and every (yn)n ∈ ℓ∞(N, M ) such that xn − yn → 0 ∗- (2) For every (xn)n ∈ ℓ∞(N, M ) satisfying limn→ω kxnϕ − ϕxnk = 0, we have (xn)n ∈ (3) For every projection e ∈ M ω, there exists a sequence of projections (en)n ∈ Mω(M ) strongly as n → ω, we have (yn)n ∈ Mω(M ) and (xn)ω = (yn)ω ∈ M ω. Mω(M ) and (xn)ω ∈ (M ω)ϕω such that e = (en)ω. . Proof. (1) Let (xn)n ∈ Mω(M ) and (yn)n ∈ ℓ∞(N, M ) such that xn − yn → 0 ∗-strongly as n → ω. Then (yn− xn)n ∈ I ω(M ) ⊂ Mω(M ) and hence (yn)n = (yn− xn)n + (xn)n ∈ Mω(M ). Moreover, by the definition of the ultraproduct von Neumann algebra M ω, we have (xn)ω = (yn)ω ∈ M ω. (2) Let (xn)n ∈ ℓ∞(N, M ) such that limn→ω kxnϕ − ϕxnk = 0. Let (bn)n ∈ I ω(M ). We may assume that max{kxnk∞,kbnk∞ : n ∈ N} ≤ 1. Using the Cauchy-Schwarz inequality, for all n ∈ N, we have (kxnbnk♯ ϕ)2 = ϕ(b∗n x∗nxnbn) + ϕ(xn bnb∗nx∗n) ≤ kbnkϕ kx∗nxnbnkϕ + (xnϕ − ϕxn)(bnb∗nx∗n) + ϕ(bn b∗nx∗nxn) ≤ kbnkϕ + kxnϕ − ϕxnkkbnb∗nx∗nk∞ + kb∗nkϕ kb∗nx∗nxnkϕ ≤ kbnkϕ + kxnϕ − ϕxnk + kb∗nkϕ. GAMMA STABILITY IN FREE PRODUCT VON NEUMANN ALGEBRAS 7 Therefore, we obtain limn→ω kxnbnk♯ we have ϕ = 0 and so (xnbn)n ∈ I ω(M ). Likewise, for all n ∈ N, (kbnxnk♯ ϕ)2 = ϕ(x∗n b∗nbnxn) + ϕ(bn xnx∗nb∗n) ≤ (x∗nϕ − ϕx∗n)(b∗nbnxn) + ϕ(b∗n bnxnx∗n) + kb∗nkϕ kxnx∗nb∗nkϕ ≤ kx∗nϕ − ϕx∗nkkb∗nbnxnk∞ + kbnkϕ kbnxnx∗nkϕ + kb∗nkϕ ≤ kxnϕ − ϕxnk + kbnkϕ + kb∗nkϕ. ϕ = 0 and so (bnxn)n ∈ I ω(M ). This shows that (xn)n ∈ Therefore, we obtain limn→ω kbnxnk♯ Mω(M ). Moreover, x = (xn)ω ∈ (M ω)ϕω by [AH12, Lemma 4.35]. (3) The proof is identical to the one of [Co75a, Proposition 1.1.3]. Let e ∈ M ω be any projection. We may choose a sequence (xn)n ∈ Mω(M ) such that kxnk∞ ≤ 1 for all n ∈ N and e = (xn)ω. Put yn = x∗nxn for all n ∈ N. Since e = e∗e, we have limn→ω kxn − ynk♯ ϕ = 0, (yn)n ∈ Mω(M ) nk♯ and e = (yn)ω. Since e = e2, we moreover have limn→ω kyn − y2 nkϕ. Letting en = 1[1−√εn,1](yn) ∈ M for all n ∈ N, we have limn→ω kyn − enk♯ ϕ = 0 by [Co75a, It follows that (en)n ∈ Mω(M ) and e = (en)ω ∈ M ω by item (1) of the Lemma 1.1.5]. proposition. ϕ = 0. Put εn = kyn − y2 (cid:3) = Q′ ∩ (M ω)ϕω . is diffuse. t ) and since σϕω t (x) = σϕ = C or Q′ ∩ (M ω)ϕω ). Hence (Q′ ∩ M ω)ϕω = (Q′ ∩ M ω) ∩ (M ω)ϕω The next proposition will be useful to prove Corollary B. Proposition 2.5. Let M be any factor with separable predual and Q ⊂ M any irreducible subfactor with expectation. Then, either Q′ ∩ M ω = C or Q′ ∩ M ω is diffuse. Proof. Denote by EQ : M → Q the faithful normal conditional expectation. Choose a faithful normal state on Q and still denote by ϕ the faithful normal state ϕ ◦ EQ on M . Since Q is globally invariant under the modular automorphism group (σϕ t (x) for all x ∈ M , the relative commutant Q′∩ M ω is globally invariant under the modular automorphism group (σϕω t Claim. Either Q′ ∩ (M ω)ϕω Proof of the Claim. We use the proof of [Io12, Lemma 2.7]. Put Q = Q′ ∩ (M ω)ϕω and denote by e ∈ Z(Q) the maximum central projection in Q such that Qe is discrete. We may represent e = (en)ω by a sequence of projections (en)n ∈ Mω(M ). Put λ = ϕω(e) = limn→ω ϕ(en). Since Q′ ∩ M = C, we have en → λ1 σ-weakly as n → ω. Next, we construct by induction a sequence of projections (fm)m≥1 in Q such that (1) Indeed, assume that f1, . . . , fm ∈ Q have been constructed. For every 1 ≤ j ≤ m, represent fj = (fj,n)ω by a sequence of projections (fj,n)n ∈ Mω(M ). Let (xi)i∈N be a k · k♯ ϕ-dense sequence in Ball(Q). Since e = (en)ω ∈ (M ω)ϕω ϕ = 0 for all i ∈ N and since en → λ1 σ-weakly as n → ω, we can find an increasing sequence (kn)n in N such that for every n ≥ 1, we have (P1) keknϕ − ϕeknk ≤ 1 n , (P2) keknxi − xieknk♯ ϕ ≤ 1 (P3) ϕ(enekn) − λϕ(en) ≤ 1 (P4) ϕ(enfj,nekn) − λϕ(enfj,n) ≤ 1 Property (P1) together with Proposition 2.4 imply that the sequence (ekn)n lies in the multiplier algebra Mω(M ) and f = (ekn)ω ∈ (M ω)ϕω . Property (P2) implies that xif = f xi for all i ∈ N. Since {xi : i ∈ N} is ∗-strongly dense in Ball(Q), we obtain that f ∈ Q′ ∩ (M ω)ϕω = Q. ϕω(efi) = λ2 and ϕω(efifj) = λ3,∀1 ≤ i < j. , since limn→ω kenxi − xienk♯ n for all 1 ≤ j ≤ m. n for all 1 ≤ i ≤ n, n and 8 CYRIL HOUDAYER Finally, Property (P3) implies that ϕω(ef ) = λϕω(e) = λ2 and Property (P4) together with the induction hypothesis imply that ϕω(efjf ) = λϕω(efj) = λ3 for all 1 ≤ j ≤ m. We can now put fm+1 = f . This finishes the proof of the induction. Define pm = fme which is a projection in Qe. Observe that since Qe is a discrete tracial von Neumann algebra, Qe is ∗-isomorphic to a countable direct sum of finite dimensional factors and hence its unit ball Ball(Qe) is k·kϕω ϕω(e) . Thus, we may choose a subsequence (pmk )k≥1 which is k · kϕω e -convergent in Ball(Qe). By Cauchy-Schwarz inequality, for all 1 ≤ j < k, we have e -compact, where ϕω e = ϕω(e·e) ϕω e (pmj pmk ) − ϕω e (pmj ) = ϕω e . e (pmj (pmk − pmj )) ≤ kpmj − pmkkϕω Taking the limit as (j, k) → ∞ and using (1), we obtain λ2 = λ3. Therefore λ ∈ {0, 1} and so e ∈ {0, 1}. This implies that either e = 0 and Q is diffuse or e = 1 and Q is a discrete tracial von Neumann algebra. In the case when Q is a discrete tracial von Neumann algebra, we show that Q = C. Assume by contradiction that Q is a discrete tracial von Neumann algebra and that Q 6= C. Denote by EM : M ω → M the canonical faithful normal conditional expectation. Recall that ϕ ◦ EM = ϕω. Since Q 6= C, we may choose a projection e ∈ Q satisfying ϕω(e) = λ with λ 6= 0, 1. We may represent e = (en)ω ∈ Q by a sequence of projections (en)n ∈ Mω(M ). Observe that EM (e) = λ1 = σ-weak limn→ω en. Then for all y ∈ Ball(M ), we have ke − ykϕω ≥ ke − EM (e)kϕω = pλ − λ2 > 0. √λ−λ2 2 Put ε = . Put e1 = e ∈ Q. Next, we construct by induction a sequence of projections em ∈ Q such that kep − eqkϕω ≥ ε for all p, q ≥ 1 such that p 6= q. Assume that e1, . . . , em ∈ Q have been constructed. For every 1 ≤ j ≤ m, represent ej = (ej,n)ω by a sequence of projections (ej,n)n ∈ Mω(M ). Let (xi)i∈N be a k·k♯ ϕ-dense sequence in Ball(Q). Since e = (en)ω ∈ (M ω)ϕω , since limk→ω kekxi − xiekk♯ ϕ = 0 for all i ∈ N and since limk→ω kek − ej,nkϕ = ke− ej,nkϕω ≥ 2ε for all 1 ≤ j ≤ m and all n ∈ N, we can find an increasing sequence (kn)n in N such that for every n ≥ 1, we have (P1) keknϕ − ϕeknk ≤ 1 n , (P2) keknxi − xieknk♯ ϕ ≤ 1 (P3) kekn − ej,nkϕ ≥ ε for all 1 ≤ j ≤ m. By the same reasoning as before, Properties (P1) and (P2) imply that (ekn)n ∈ Mω(M ) and f = (ekn)ω ∈ Q. Moreover, Property (P3) implies that kf − ejkϕω ≥ ε for all 1 ≤ j ≤ m. We can now put em+1 = f . This finishes the proof of the induction. So, we have constructed a sequence of projections em ∈ Q such that kep − eqkϕω ≥ ε for all p, q ≥ 1 such that p 6= q. This however contradicts the fact that Ball(Q) is k·kϕω -compact and finishes the proof of the Claim. Assume that Q′ ∩ (M ω)ϕω = C. Then by [AH12, Lemma 5.4], we have that Q′ ∩ M ω = C or Q′ ∩ M ω is a type III1 factor. Next, assume that Q′ ∩ (M ω)ϕω is diffuse. Then, using Proposition 2.2, we have that Q′ ∩ M ω is diffuse. Therefore, either Q′ ∩ M ω = C or Q′ ∩ M ω is diffuse. n for all 1 ≤ i ≤ n and (cid:3) (cid:3) Proposition 2.6. For every diffuse amenable von Neumann algebra M with separable predual, the central sequence algebra M′ ∩ M ω is diffuse. Proof. Let M be any diffuse amenable von Neumann algebra with separable predual. There exists a sequence of pairwise orthogonal projections zn ∈ Z(M ) such that Pn zn = 1, M z0 is an amenable von Neumann algebra with a diffuse center and separable predual and M zn is a diffuse GAMMA STABILITY IN FREE PRODUCT VON NEUMANN ALGEBRAS 9 amenable factor with separable predual for every n ≥ 1. It is obvious that (M z0)′ ∩ (M z0)ω is diffuse. By the classification of amenable factors with separable predual (see [Co72, Co74, Co75b, Co85, Ha84]), M zn is hyperfinite and (M zn)′ ∩ (M zn)ω is diffuse for every n ≥ 1. Therefore M′ ∩ M ω = Ln(M zn)′ ∩ (M zn)ω is diffuse. An elementary fact on ε-orthogonality. Let H be a complex Hilbert space and ε ≥ 0. We say that two (not necessarily closed) subspaces K,L ⊂ H are ε-orthogonal and we denote by K ⊥ε L if (cid:3) Define the function hξ, ηiH ≤ εkξkH kηkH, ∀ξ ∈ K,∀η ∈ L. δ : (cid:20)0, 2(cid:19) → R+ : t 7→ 2t 1 p1 − t − √2 t√1 − t . We will be using the following elementary fact regarding ε-orthogonality whose proof can be found in [Ho12a, Proposition 2.3]. Proposition 2.7 ([Ho12a]). Let k ≥ 1. Let 0 ≤ ε < 1 such that δ◦(k−1)(ε) < 1. For all 1 ≤ i ≤ 2k, let pi ∈ B(H) be projections such that piH ⊥ε pjH for all i, j ∈ {1, . . . , 2k} such that i 6= j. Write Pk = W2k 2k i=1 pi. Then for all ξ ∈ H, we have Yj=0(cid:0)1 + δ◦j (ε)(cid:1)2 Xi=1 kpiξk2 H ≤ k−1 kPkξk2 H. 3. Asymptotic orthogonality in the ultraproduct framework The key result of the paper is the following generalization of Popa's result [Po83, Lemma 2.1] regarding asymptotic orthogonality for free group factors to arbitrary free product von Neumann algebras. There are mainly two difficulties that arise in generalizing Popa's result [Po83, Lemma 2.1] to the setting of arbitrary free product von Neumann algebras. The first main difficulty is that the free product von Neumann algebra (M, ϕ) = (M1, ϕ1)∗(M2, ϕ2) is no longer assumed to be tracial. Hence, we need to work in the ultraproduct von Neumann algebra framework and carefully approximate elements in M in the σ-strong topology by finite linear combinations of reduced words which are analytic with respect to the modular automorphism group (σϕ t ) (see also the proof of [Ue11, Proposition 3.5] where a similar method is used). The second main difficulty is that unlike the case of the free group factors, M is no longer assumed to have a nice basis of unitary elements. To circumvent this issue, we will use ε-orthogonality techniques from [Ho12a, Ho12b]. Theorem 3.1. Let (M1, ϕ1) and (M2, ϕ2) be σ-finite von Neumann algebras endowed with faithful normal states. Assume that the centralizer M ϕ1 is diffuse. Denote by (M, ϕ) = 1 (M1, ϕ1) ∗ (M2, ϕ2) the free product von Neumann algebra. Let u ∈ U (M ϕ1 and every y ∈ M ⊖ M1, the elements y(x− EM ω are pairwise ϕω-orthogonal in M ω. 1 ) be any unitary such that uk → 0 weakly as k → ∞. For every x ∈ {u}′ ∩ M ω 1 (x)y 1 (x)), (x− EM ω 1 (x))y and yEM ω 1 (x) − EM ω Proof. For every i ∈ {1, 2}, denote by Ai ⊂ Mi (resp. A ⊂ M ) the unital σ-strongly dense ∗-subalgebra of all the elements in Mi (resp. M ) which are analytic with respect to the modular automorphism group (σϕi t )) (see Proposition 2.1). Observe that for every i ∈ {1, 2}, Ai ⊂ A. Denote by (Ai1 ⊖ C)··· (Ain ⊖ C) the set of all the reduced words of the form a1 ··· an with aj ∈ Aij ⊖ C, n ≥ 1 and i1 6= ··· 6= in. The linear span of t ) (resp. (σϕ forms a unital σ-strongly dense ∗-subalgebra of M . {1, (Ai1 ⊖ C)··· (Ain ⊖ C) : n ≥ 1, i1 6= ··· 6= in} 10 CYRIL HOUDAYER Using the existence of the normal conditional expectation EM1 : M → M1, every y ∈ M ⊖ M1 can be approximated with respect to the σ-strong topology by a net (yα)α∈I of finite linear combinations of reduced words in (Ai1 ⊖ C)··· (Ain ⊖ C) where n ≥ 1, 2 ∈ {i1, . . . , in} and i1 6= ··· 6= in. Assume that for every α ∈ I and every x ∈ {u}′ ∩ M ω, yα(x − EM ω 1 (x)), 1 (x)yα are pairwise ϕω-orthogonal in M ω. Then since (x − EM ω yα → y σ-strongly as α → ∞, it follows that 1 (x))yα and yαEM ω 1 (x) − EM ω yα(x − EM ω (x − EM ω 1 (x) − EM ω 1 (x)) → y(x − EM ω 1 (x))yα → (x − EM ω 1 (x)yα → yEM ω 1 (x)) 1 (x))y 1 (x) − EM ω 1 (x)y yαEM ω 1 (x)y are σ-strongly as α → ∞. Therefore, y(x − EM ω pairwise ϕω-orthogonal in M ω. Using the previous discussion, we infer that it suffices to prove the result for 1 (x)), (x − EM ω 1 (x))y and yEM ω 1 (x) − EM ω y = k Xj=1 wj where wj = aj,1bj,1 ··· bj,nj aj,nj+1 with nj ≥ 1, aj,1 = 1 or aj,1 ∈ A1 ⊖ C, aj,nj+1 = 1 or aj,nj+1 ∈ A1 ⊖ C, aj,2, . . . , aj,nj ∈ A1 ⊖ C and bj,1, . . . , bj,nj ∈ A2 ⊖ C. We fix such an element y ∈ M ⊖ M1 until the end of the proof. Observe that for every 1 ≤ j ≤ k, we have wj ∈ A ⊖ C and −i (b∗j,nj )··· σϕ2 −i(w∗j ) = σϕ1 σϕ −i (a∗j,nj+1)σϕ2 −i (b∗j,1)σϕ1 −i (a∗j,1). −i(w∗j ) is a reduced word containing at least one letter from M2 ⊖ C. It follows that σϕ Denote by V ⊂ M1 the finite dimensional vector subspace generated by 1 and by • the first letters coming from M1 ⊖ C of the reduced words wi, w∗i , σϕ −i(w∗i ) and the first letters coming from M1 ⊖ C of all the reduced words arising in the finite linear decomposition of w∗j wi into reduced words, for all 1 ≤ i, j ≤ k, and • the last letters coming from M1 ⊖ C of the reduced words wi and the last letters coming from M1 ⊖ C of all the reduced words arising in the finite linear decomposition of wiσϕ −i(w∗j ) into reduced words, for all 1 ≤ i, j ≤ k. Let ℓ = dim(V ) and choose elements e1, . . . , eℓ ∈ V so that (Λϕ1(ei))ℓ basis for Λϕ1(V ). By Gram-Schmidt process, choose a vector subspace W ⊂ M1 so that i=1 forms an orthonormal L2(M1) = Λϕ1(V ) ⊕ Λϕ1(W ). We will be using the following notation: • K1 ⊂ L2(M ) is the closed subspace generated by the image under Λϕ of the linear span of all the reduced words in (M2 ⊖ C)··· (M2 ⊖ C), (V ⊖ C)(M2 ⊖ C)··· (M2 ⊖ C), (M2 ⊖ C)··· (M2 ⊖ C)(M1 ⊖ C) and (V ⊖ C)(M2 ⊖ C)··· (M2 ⊖ C)(M1 ⊖ C). Observe that K1 ∼= Λϕ(V ) ⊗ L2((M2 ⊖ C)··· (M2 ⊖ C)M1). • K2 ⊂ L2(M ) is the closed subspace generated by the image under Λϕ of the linear span of all the reduced words in W (M2⊖C)··· (M2⊖C) and W (M2⊖C)··· (M2⊖C)(V ⊖C). Observe that K2 ∼= L2(W (M2 ⊖ C)··· (M2 ⊖ C)) ⊗ Λϕ(V ). • L ⊂ L2(M ) is the closed subspace generated by the image under Λϕ of the linear span of all the reduced words in W (M2 ⊖ C)··· (M2 ⊖ C)W . Observe that L2(M1) ⊕ K1 ⊕ K2 ⊕ L = L2(M ). GAMMA STABILITY IN FREE PRODUCT VON NEUMANN ALGEBRAS 11 1 ) ⊂ U (M ϕ), we have T Λϕ(z) = Λϕ(uzu∗) for all z ∈ M . 1 ) such that uk → 0 weakly as k → ∞ and put T = u JϕuJϕ ∈ U (L2(M )). Let u ∈ U (M ϕ1 Observe that since u ∈ U (M ϕ1 Claim 1. For all ε > 0, there exists k0 ∈ N such that for all i ∈ {1, 2} and all k ≥ k0, we have T kKi ⊥ε Ki. Proof of Claim 1. Let ξ, η ∈ K1 that we write Pℓ with ξi, ηj ∈ L2((M2 ⊖ C)··· (M2 ⊖ C)M1). Observe that kξk2 Pℓ j=1 kηjk2 i=1 Λϕ(ei) ⊗ ξi and η = Pℓ i=1 kξik2 i=1 Λϕ(ukei) ⊗ JϕukJϕξi and hence ϕ(e∗j ukei)kξikϕ kηjkϕ. ϕ. Since u ∈ M ϕ, we have T kξ = Pℓ Xi,j=1 j=1 Λϕ(ej) ⊗ ηj ϕ and kηk2 ϕ = hT kξ, ηiϕ ≤ ϕ = Pℓ ℓ Since uk → 0 weakly as k → ∞, we may choose k1 ∈ N such that for all k ≥ k1 and all 1 ≤ i, j ≤ ℓ, we have ϕ(e∗j ukei) ≤ ε/ℓ. By Cauchy-Schwarz inequality, for all k ≥ k1, we obtain hT kξ, ηiϕ ≤ εkξkϕkηkϕ. Likewise let ξ, η ∈ K2 that we write Pℓ L2(W (M2 ⊖ C)··· (M2 ⊖ C)). Observe that kξk2 u ∈ M ϕ, we have T kξ = Pℓ j=1 ηj ⊗ Λϕ(ej) with ξi, ηj ∈ ϕ. Since i=1 ξi ⊗ Λϕ(ei) and η = Pℓ ϕ and kηk2 ϕ = Pℓ ϕ = Pℓ j=1 kηjk2 i=1 kξik2 i=1 ukξi ⊗ Λϕ(eiu−k) and hence hT kξ, ηiϕ ≤ ϕ(e∗j eiu−k)kξikϕ kηjkϕ. ℓ Xi,j=1 Since uk → 0 weakly as k → ∞, we may choose k2 ∈ N such that for all k ≥ k2 and all 1 ≤ i, j ≤ ℓ, we have ϕ(e∗j eiu−k) ≤ ε/ℓ. By Cauchy-Schwarz inequality, for all k ≥ k2, we obtain hT kξ, ηiϕ ≤ εkξkϕkηkϕ. Put k0 = max(k1, k2). Then for all i ∈ {1, 2} and all k ≥ k0, we have that T kKi ⊥ε Ki. Claim 2. For all i ∈ {1, 2} and all (zn)ω ∈ {u}′ ∩ M ω, we have (cid:3) lim n→ω kPKi(Λϕ(zn))kϕ = 0. Proof of Claim 2. Let i ∈ {1, 2} and z = (zn)ω ∈ {u}′ ∩ M ω. We may assume that kznk∞ ≤ 1 for all n ∈ N. For all n ∈ N and all k ∈ N, we have kPKi(Λϕ(zn))k2 ϕ ϕ = kT kPKi(Λϕ(zn))k2 = kT kPKi(Λϕ(zn)) − PT kKi(Λϕ(zn)) + PT kKi(Λϕ(zn))k2 ≤ 2kT kPKi(Λϕ(zn)) − PT kKi(Λϕ(zn))k2 = 2kPT kKi(Λϕ(ukznu−k − zn))k2 ≤ 2kukznu−k − znk2 ϕ + 2kPT kKi(Λϕ(zn))k2 ϕ. ϕ + 2kPT kKi(Λϕ(zn))k2 ϕ ϕ ϕ + 2kPT kKi(Λϕ(zn))k2 ϕ Fix K ≥ 1. Choose ε > 0 very small according to [Ho12a, Proposition 2.3] so that QK−1 j=0 (1 + δ◦j (ε))2 ≤ 2. Then choose a subset G ⊂ N of 2K integers such that two distinct integers in G are at least at distance k0 from one another. By Claim 1, we obtain T k1Ki ⊥ε T k2Ki for all k1, k2 ∈ G such that k1 6= k2. Thus, we obtain ϕ ≤ 2Xk∈G Since G is finite, we have limn→ω kPKi(Λϕ(zn))k2 limn→ω kPKi(Λϕ(zn))kϕ = 0. ϕ ≤ 22−K for all K ≥ 1. Therefore, we obtain 2KkPKi(Λϕ(zn))k2 kukznu−k − znk2 ϕ + 4kznk2 ϕ. (cid:3) 12 CYRIL HOUDAYER Claim 3. The subspaces y L, Jϕσϕ wise orthogonal in L2(M ). −i/2(y∗)Jϕ L and y L2(M1) + Jϕσϕ −i/2(y∗)Jϕ L2(M1) are pair- Proof of Claim 3. Recall that y = Pk j=1 wj where wj = aj,1bj,1 ··· bj,nj aj,nj+1 with nj ≥ 1, aj,1 = 1 or aj,1 ∈ A1 ⊖ C, aj,nj+1 = 1 or aj,nj+1 ∈ A1 ⊖ C, aj,2, . . . , aj,nj ∈ A1 ⊖ C and bj,1, . . . , bj,nj ∈ A2 ⊖ C. Observe that (2) y L ⊂ span{Λϕ(wjW (M2 ⊖ C)··· (M2 ⊖ C)W ) : 1 ≤ j ≤ k} −i/2(y∗)Jϕ L ⊂ span{Λϕ(W (M2 ⊖ C)··· (M2 ⊖ C)W wj) : 1 ≤ j ≤ k} Jϕσϕ (3) and (4) y L2(M1) + Jϕσϕ −i/2(y∗)Jϕ L2(M1) ⊂ span{Λϕ(wiM1), Λϕ(M1wj) : 1 ≤ i, j ≤ k}. −i/2(y∗)Jϕ L are orthogonal in L2(M ). Let 1 ≤ i ≤ k. Observe that by the choice of the vector subspace W ⊂ M1, any letter v ∈ W is ϕ-orthogonal in M to the first letter of the reduced word w∗i and to the first letter of the reduced word σϕ −i(w∗i ). Hence wiv is a reduced word starting with the first letter of wi and ending with a letter from M1⊖ C and vwi is a reduced word starting with a letter from M1 ⊖ C and ending with the last letter of wi. Moreover both vwi and wiv contain at least one letter from M2 ⊖ C. Let 1 ≤ i, j ≤ k. By the choice of the vector subspace W ⊂ M1 and the remark above, the first letter of any reduced word wiv with v ∈ W is ϕ-orthogonal to W in M . This implies that W (M2 ⊖ C)··· (M2 ⊖ C)W wj and wiW (M2 ⊖ C)··· (M2 ⊖ C)W are ϕ-orthogonal in M . Since this holds for all 1 ≤ i, j ≤ k, using (2) and (3), we obtain that the subspaces y L and Jϕσϕ Let 1 ≤ i, j ≤ k. If ni ≤ nj, then any element in wiM1 is a finite linear combination of reduced words which have at most ni letters from M2⊖C while a reduced word in wjW (M2⊖C)··· (M2⊖ C)W has at least nj + 1 letters from M2 ⊖ C. This implies that wjW (M2 ⊖ C)··· (M2 ⊖ C)W If ni > nj, then w∗j wi is a finite linear combination of and wiM1 are ϕ-orthogonal in M . reduced words whose first letter is ϕ-orthogonal to W in M and which contain at least one It follows that any element in w∗j wiM1 is a finite linear combination letter from M2 ⊖ C. of reduced words whose first letter is ϕ-orthogonal to W in M . Again, this implies that wjW (M2 ⊖ C)··· (M2 ⊖ C)W and wiM1 are ϕ-orthogonal in M . Next, since wi contains at least one letter from M2 ⊖ C and by the choice of the vector subspace W ⊂ M1, any element in M1wi is a finite linear combination of reduced words whose last letter is ϕ-orthogonal to W in M . This implies that wjW (M2 ⊖ C)··· (M2 ⊖ C)W and M1wi are ϕ-orthogonal in M . Since the previous reasoning holds for all 1 ≤ i, j ≤ k, using (2) and (4), we obtain that the subspaces y L and y L2(M1) + Jϕσϕ Let 1 ≤ i, j ≤ k. Since wi contains at least one letter from M2 ⊖ C and by the choice of the vector subspace W ⊂ M1, any element in wiM1 is a finite linear combination of reduced words whose first letter is ϕ-orthogonal to W in M . This implies that W (M2 ⊖ C)··· (M2 ⊖ C)W wj and wiM1 are ϕ-orthogonal in M . Next, if ni ≤ nj, then any element in M1wi is a finite linear combination of reduced words which have at most ni letters from M2 ⊖ C while a reduced word in W (M2 ⊖ C)··· (M2 ⊖ C)W wj has at least nj + 1 letters from M2 ⊖ C. This implies that W (M2 ⊖ C)··· (M2 ⊖ C)W wj and M1wi are ϕ-orthogonal in M . If ni > nj, then wiσϕ −i(w∗j ) is a finite linear combination of reduced words whose last letter is ϕ-orthogonal to W in M and which contain at least one letter from M2 ⊖ C. It follows that any element in M1wiσϕ −i(w∗j ) is a finite linear combination of reduced words whose last letter is ϕ-orthogonal to W in M . Using Proposition 2.1, this implies again that W (M2 ⊖ C)··· (M2 ⊖ C)W wj and M1wi are ϕ-orthogonal in M . Since the previous reasoning holds for all 1 ≤ i, j ≤ k, using (3) and −i/2(y∗)Jϕ L2(M1) are orthogonal in L2(M ). GAMMA STABILITY IN FREE PRODUCT VON NEUMANN ALGEBRAS 13 (4), we obtain that the subspaces Jϕσϕ orthogonal in L2(M ). This finishes the proof of Claim 3. −i/2(y∗)Jϕ L and y L2(M1) + Jϕσϕ −i/2(y∗)Jϕ L2(M1) are (cid:3) We are now ready to finish the proof of Theorem 3.1. Let x ∈ {u}′ ∩ M ω and put z = 1 ). Write x − EM ω z = (zn)ω ∈ {u}′ ∩ (M ω ⊖ M ω 1 ) with zn = xn − EM1(xn). By Claim 2 and since y is analytic with respect to the modular automorphism group (σϕ 1 (x). Observe that since u ∈ M1 ⊂ M ω 1 , we have z ∈ {u}′ ∩ (M ω ⊖ M ω t ), we obtain Λϕω (yz) = (Λϕ(yzn))ω = (y Λϕ(zn))ω = (y PL(Λϕ(zn)))ω ∈ L2(M )ω Λϕω (zy) = (Λϕ(zny))ω = (Jϕσϕ = (Jϕσϕ −i/2(y)∗Jϕ Λϕ(zn))ω −i/2(y)∗Jϕ PL(Λϕ(zn)))ω ∈ L2(M )ω Λϕω (yEM ω 1 (x) − EM ω 1 (x)y) = (Λϕ(yEM1(xn) − EM1(xn)y))ω = ((y − Jϕσϕ −i/2(y)∗Jϕ) Λϕ(EM1(xn)))ω ∈ L2(M )ω. Using Claim 3 for every n ∈ N and using the ultraproduct Hilbert space structure of L2(M )ω, we obtain that Λϕω (y(x − EM ω 1 (x)y) are pairwise orthogonal in L2(M )ω. This implies that y(x − EM ω 1 (x))y and yEM ω (cid:3) 1 (x)y are pairwise ϕω-orthogonal in M ω. 1 (x))), Λϕω ((x − EM ω 1 (x)), (x − EM ω 1 (x))y) and Λϕω (yEM ω 1 (x) − EM ω 1 (x) − EM ω 4. Proofs of the main results 4.1. Proof of Theorem A and Corollaries B and C. 1 Proof of Theorem A. Let M1 ⊂ Q ⊂ M be any intermediate von Neumann subalgebra such that Q′ ∩ M ω is diffuse. Since M ϕ1 is diffuse, by [Ue11, Corollary 3.2], we have Q′ ∩ M ⊂ M′1 ∩ M ⊂ M1 and so Q′ ∩ M = Z(Q) = Q′ ∩ M1 ⊂ Z(M1). First, denote by z ∈ Q′ ∩ M the maximum projection such that M1z = Qz. We show that z = 1. Assume by contradiction that z 6= 1 and put q = z⊥ = 1 − z ∈ Q′ ∩ M = Z(Q). We have q 6= 0 and Qq ⊖ M1q 6= 0. Denote by J the nonzero σ-strongly closed two-sided ideal in Qq generated by Qq ⊖ M1q. Let e ∈ Z(Qq) = Z(Q)q be the unique nonzero central projection in Qq such that J = Qe. We necessarily have e = q. Indeed otherwise we have q − e 6= 0 and by the choice of the projection z ∈ Q′ ∩ M , we have Q(q − e) ⊖ M1(q − e) 6= 0. Now let y ∈ Q(q − e)⊖ M1(q − e) such that y 6= 0. Since y ∈ Qq ⊖ M1q, we obtain y ∈ J and so y = ye. However since y ∈ Q(q − e) ⊖ M1(q − e), we also obtain y = y(q − e) and thus y = 0. This is a contradiction. Thus, we have e = q. Next, we show that (Qq)′ ∩ (qM q)ω ⊂ (M1q)ω. Indeed, let x ∈ (Qq)′ ∩ (qM q)ω ⊂ M′1 ∩ M ω. For every y ∈ Qq ⊖ M1q ⊂ M ⊖ M1, we have 0 = yx − xy = y(x − EM ω 1 (x)) − (x − EM ω 1 (x))y + (yEM ω 1 (x) − EM ω 1 (x)y). By Theorem 3.1, y(x − EM ω orthogonal in M ω. By Pythagora's theorem, we obtain y(x− EM ω a ∈ Qq and every y ∈ Qq ⊖ M1q, we have a y (x− EM ω and a − EM1(a) ∈ Qq ⊖ M1q, we also have 1 (x)), (x − EM ω 1 (x))y and (yEM ω 1 (x) − EM ω 1 (x)y) are pairwise ϕω- 1 (x)) = 0. Likewise, for every 1 (x)) = 0 and since yEM1(a) ∈ Qq ⊖ M1q y a (x − EM ω 1 (x)) = y EM1(a) (x − EM ω 1 (x)) + y (a − EM1(a)) (x − EM ω 1 (x)) = 0. This implies that for every y ∈ J , we have y(x − EM ω Therefore, x = E(M1q)ω (x) ∈ (M1q)ω. 1 (x)) = 0 hence q(x − EM ω 1 (x)) = 0. 14 CYRIL HOUDAYER Now we have that (Qq)′ ∩ (qM q)ω = (Qq)′ ∩ (M1q)ω. Since Q′ ∩ M ω is diffuse and since (Qq)′∩ (qM q)ω = q(Q′∩ M ω)q, we have that (Qq)′∩ (M1q)ω is diffuse as well. This implies that there exists a net of unitaries Uα ∈ U ((Qq)′∩ (M1q)ω) such that Uα → 0 weakly as α → ∞. We n)n ∈ Mω(M1q) such may represent every Uα ∈ U ((Qq)′∩ (M1q)ω) by a sequence of elements (uα that uα ny → 0 ∗-strongly as n → ω for every α and every y ∈ Qq. Define the directed set n ∈ Ball(M1q) for every α and every n ∈ N, Uα = (uα n)ω for every α and yuα n − uα I = {i = (ε,F,G) : ε > 0, F ⊂ M1q and G ⊂ Qq are finite subsets} with order relation given by (ε1,F1,G1) ≤ (ε2,F2,G2) if and only if ε2 ≤ ε1, F1 ⊂ F2 and G1 ⊂ G2. Let i = (ε,F,G) ∈ I. Since Uα → 0 weakly as α → ∞, there exists α such that ϕω(b∗Uαa) ≤ n)ω ∈ U ((Qq)′ ∩ (M1q)ω), for all a, b ∈ F and all c ∈ G, we ε/2 for all a, b ∈ F. Since Uα = (uα have ε 2 ≥ ϕω(b∗Uαa) = lim kakϕ = kUαakϕω = lim 0 = kcUα − Uαckϕω = lim n→ω ϕ(b∗uα n→ω kuα nakϕ n − uα n→ω kcuα na) nckϕ. Since F ⊂ M1q and G ⊂ Qq are finite subsets, there exists n = n(α) such that maxnkakϕ − kuα n(α)akϕ,kcuα n(α)a) : a, b ∈ F, c ∈ Go ≤ ε. n(α) ∈ Ball(M1q). Thus, (wi)i∈I is a net of elements in Ball(M1q) such that n(α)ckϕ,ϕ(b∗uα n(α) − uα Put wi = uα (P1) limi∈I kwiakϕ = kakϕ for all a ∈ M1q. (P2) limi∈I kcwi − wickϕ = 0 for all c ∈ Qq. (P3) limi∈I ϕ(b∗wia) = 0 for all a, b ∈ M1q. Put E = span({q(Mi1 ⊖ C)··· (Min ⊖ C)q : n ≥ 1, 2 ∈ {i1, . . . , in} and i1 6= ··· 6= in}). Observe that E is σ-strongly dense in qM q ⊖ M1q. Claim. The following hold true. (1) For all a, b ∈ E, we have lim i∈I kEM1q(b∗wia)kϕ = 0. (2) For all b ∈ E and all y ∈ qM q ⊖ M1q, we have lim i∈I kEM1q(b∗wiy)kϕ = 0. Proof of the Claim. (1) By linearity, it suffices to prove the result for all the elements a, b ∈ E of the form a = a1 ··· a2m+1 and b = b1 ··· b2n+1 with m, n ≥ 1, a1 = q or a1 ∈ M1q ⊖ Cq, a2m+1 = q or a2m+1 ∈ M1q ⊖ Cq, b1 = q or b1 ∈ M1q ⊖ Cq, b2n+1 = q or b2n+1 ∈ M1q ⊖ Cq, a2, . . . , a2m, b2, . . . , b2n ∈ M2 ⊖ C and a3, . . . a2m−1, b3, . . . , b2n−1 ∈ M1 ⊖ C. We have b∗wia = b∗2n+1 ··· b∗2 (b∗1wia1) a2 ··· a2m+1. By the freeness property, we have EM1(b∗wia) = ϕ(b∗1wia1) EM1(b∗2n+1 ··· b∗2 a2 ··· a2m+1). Using property (P3) of the net (wi)i∈I , we obtain limi∈I kEM1q(b∗wia)kϕ = 0. GAMMA STABILITY IN FREE PRODUCT VON NEUMANN ALGEBRAS 15 (2) Let y ∈ qM q ⊖ M1q and b ∈ E. We may assume that kbk∞ ≤ 1. Since E is σ-strongly dense in qM q ⊖ M1q, for every ε > 0, there exists a ∈ E such that ky − akϕ ≤ ε/2. Thus, for every i ∈ I, we have Using the first part of the proof, this implies that lim supi∈I kEM1q(b∗wiy)kϕ ≤ ε. Since ε > 0 is arbitrary, we obtain limi∈I kEM1q(b∗wiy)kϕ = 0. This finishes the proof of the Claim. Let b ∈ E and y ∈ Qq ⊖ M1q. Using the properties (P1) and (P2) of the net (wi)i∈I , we obtain kEM1(b∗wi(y − a))kϕ ≤ kb∗wi(y − a)kϕ ≤ ky − akϕ ≤ ε. (cid:3) kEM1q(b∗y)kϕ = lim = lim i∈I kwi EM1q(b∗y)kϕ using (P1) for a = EM1q(b∗y) i∈I kEM1q(b∗y) wikϕ using (P2) for c = EM1q(b∗y) i∈I kEM1q(b∗y wi)kϕ since wi ∈ M1q i∈I kEM1q(b∗wi y)kϕ using (P2) for c = y = lim = lim using item (2) of the Claim. = 0 Since E is σ-strongly dense in qM q ⊖ M1q, we may choose a net (bj)j∈J in E such that b∗j → y∗ σ-strongly as j → ∞. Since EM1q : qM q → M1q is σ-strongly continuous, we obtain that EM1q(b∗j y) → EM1q(y∗y) σ-strongly as j → ∞ and hence EM1q(y∗y) = 0. This implies that y∗y = 0 and hence y = 0. Since y ∈ Qq ⊖ M1q is arbitrary, we derive that M1q = Qq. This contradicts the maximality of the projection z ∈ Q′ ∩ M and finishes the proof of Theorem A. (cid:3) Proof of Corollary B. Let M1 ⊂ Q ⊂ M be any intermediate von Neumann subalgebra with faithful normal conditional expectation EQ : M → Q. Denote by EM1 : M → M1 the unique ϕ-preserving normal conditional expectation. Since M ϕ1 is diffuse, we have M′1 ∩ M ⊂ M1 1 by [Ue11, Corollary 3.2] and hence EM1 is the unique faithful normal conditional expectation from M to M1 by [Co72, Th´eor`eme 1.5.5]. Since EM1 ◦ EQ is a faithful normal conditional expectation from M to M1, we have EM1 ◦ EQ = EM1. This implies that for every x ∈ M , we have ϕ(EQ(x)) = ϕ(EM1(EQ(x))) = ϕ((EM1 ◦ EQ)(x)) = ϕ(EM1(x)) = ϕ(x). t ). By [Ta03, Theorem IX.4.2], we obtain that Q is globally invariant under the modular automor- phism group (σϕ Since Q′ ∩ M = Z(Q) is abelian, there exists a sequence of pairwise orthogonal projections qn ∈ Q′ ∩ M ⊂ Z(M1) such that Pn qn = 1, (Qq0)′ ∩ q0M q0 = (Q′ ∩ M )q0 is a diffuse abelian von Neumann algebra and Qqn is a diffuse factor such that (Qqn)′∩qnM qn = (Q′∩M )qn = Cqn for every n ≥ 1. Define I = {0} ∪(cid:8)n ≥ 1 : (Qqn)′ ∩ (qnM qn)ω is diffuse(cid:9) . Put z0 = Pn∈I qn and N = (Cz0 ⊕ M1z⊥0 ) ∨ M2. If z0 = 0, then M1z0 = Qz0. Otherwise, by [Ue11, Lemma 2.2], we have that M1z0 and z0N z0 generate z0M z0 and are free in z0M z0 with respect to the state ϕz0 = ϕ(z0·z0) ϕ(z0) . Thus, we have (z0M z0, ϕz0) = (M1z0, ϕz0) ∗ (z0N z0, ϕz0). Moreover, the intermediate subalgebra M1z0 ⊂ Qz0 ⊂ z0M z0 is globally invariant under the modular automorphism group (σ ) and we have ϕz0 t (5) (Qqn)′ ∩ (qnM qn)ω ⊂ (Qz0)′ ∩ (z0M z0)ω. Mn∈I Since Q ⊂ M is globally invariant under the modular automorphism group (σϕ qn ∈ M ϕ for all n ∈ N, we have that both Ln∈I (Qqn)′ ∩ (qnM qn)ω and (Qz0)′ ∩ (z0M z0)ω t ) and since 16 CYRIL HOUDAYER are globally invariant under the modular automorphism group (σ ϕω z0 t (5) is with expectation. Since Ln∈I (Qqn)′ ∩ (qnM qn)ω is diffuse, so is (Qz0)′ ∩ (z0M z0)ω by Proposition 2.2. Applying Theorem A to the intermediate von Neumann subalgebra M1z0 ⊂ Qz0 ⊂ z0M z0, we obtain M1z0 = Qz0. For every n /∈ I, (Qqn)′ ∩ (qnM qn)ω is not diffuse. By Proposition 2.5, we obtain that (Qqn)′ ∩ (qnM qn)ω = Cqn. In particular, since Q ⊂ M is with expectation, we have (Qqn)′ ∩ (Qqn)ω ⊂ (Qqn)′ ∩ (qnM qn)ω. Thus, we have (Qqn)′ ∩ (Qqn)ω = Cqn and so Qqn is a full nonamenable factor by Proposition 2.6. This finishes the proof of Corollary B. (cid:3) ). Therefore, the inclusion Proof of Corollary C. Let M1 be any diffuse amenable von Neumann algebra with separable predual. Choose a faithful normal state ϕ1 on M1 such that the centralizer M ϕ1 is diffuse (see 1 Proposition 2.2). Define M2 = R∞ to be the unique hyperfinite type III1 factor endowed with any faithful normal state ϕ2. Then by [Ue11, Theorem 3.4], the free product (M, ϕ) = (M1, ϕ1)∗ (M2, ϕ2) is a full nonamenable type III1 factor. Moreover M1 ⊂ M is with expectation. Let M1 ⊂ Q ⊂ M be any intermediate amenable von Neumann algebra with expectation. By Corollary B, we obtain that M1 = Q. (cid:3) 4.2. Proof of Theorem D. We recall Popa's intertwining-by-bimodules theory that will play a crucial role in the proof of Theorem D. Let M be a tracial von Neumann algebra together with A ⊂ 1AM 1A and B ⊂ 1BM 1B von Neumann subalgebras. Following [Po01, Po03], we say that A embeds into B inside M and denote by A (cid:22)M B if one of the following equivalent conditions is satisfied: • There exist projections p ∈ A and q ∈ B, a nonzero partial isometry v ∈ pM q and a unital normal ∗-homomorphism ϕ : pAp → qBq such that av = vϕ(a) for all a ∈ pAp. • There exist ℓ ≥ 1, a projection q ∈ Mℓ(B), a nonzero partial isometry v ∈ M1,ℓ(1AM )q and a unital normal ∗-homomorphism ϕ : A → qMℓ(B)q such that av = vϕ(a) for all a ∈ A. • There is no net of unitaries (wi)i∈I in U (A) such that EB(x∗wiy) → 0 ∗-strongly as i → ∞ for all x, y ∈ pM q. We first prove the following intermediate result which can be regarded as a generalization of Theorem A in the case of tracial free product von Neumann algebras. Theorem 4.1. Let (M1, τ1) and (M2, τ2) be von Neumann algebras with separable predual endowed with faithful normal tracial states. Assume that M1 is diffuse. Denote by (M, τ ) = (M1, τ1) ∗ (M2, τ2) the tracial free product von Neumann algebra. For every von Neumann subalgebra Q ⊂ M such that Q∩ M1 and Q′ ∩ M ω are diffuse, we have Q ⊂ M1. Proof. Let Q ⊂ M be any von Neumann subalgebra such that Q∩ M1 and Q′ ∩ M ω are diffuse. By [IPP05, Theorem 1.1], we have Q′ ∩ M ⊂ M1. Denote by z ∈ Z(Q′ ∩ M ) the maximum projection such that Qz ⊂ zM1z. We prove that z = 1. Assume by contradiction that this not the case and put q = z⊥ = 1 − z ∈ Z(Q′ ∩ M ) ⊂ M1. We have q 6= 0. First, assume that Qq is amenable. Choose a diffuse abelian subalgebra A ⊂ q⊥M1q⊥ and put Q = Qq⊕A. Then Q is amenable and Q∩M1 is diffuse. Theorem 3.1 implies that the inclusion M1 ⊂ M has the asymptotic orthogonality property relative to the diffuse subalgebra Q ∩ M1 in the sense of [Ho12b, Definition 5.1]. Since the inclusion M1 ⊂ M is mixing (see e.g. [Ho12b, Proposition 4.7]) in the sense of [Ho12b, Definition 4.4], we have that the inclusion M1 ⊂ M is weakly mixing through the diffuse subalgebra Q ∩ M1 in the sense of [Ho12b, Definition 4.1]. Therefore [Ho12b, Theorem 8.1] implies that Q ⊂ M1 and so Qq ⊂ qM1q. This contradicts the fact that z is the maximum projection in Z(Q′ ∩ M ) such that Qz ⊂ zM1z. GAMMA STABILITY IN FREE PRODUCT VON NEUMANN ALGEBRAS 17 Second, assume that Qq is not amenable. Let q0 ∈ Z(Q′ ∩ M )q be a nonzero central projection such that Qqq0 has no amenable direct summand. Since (Qqq0)′ ∩ (qq0M qq0)ω = qq0(Q′ ∩ M ω)qq0 is diffuse and since the inclusion M1 ⊂ M is mixing, by [Pe06, Theorems 4.3, 4.5 and Lemma 5.1] and [IPP05, Theorem 4.3] (see also [Ho07, Theorem 5.6] and [Io12, Theorem 6.3]), we obtain that Qqq0 (cid:22)M Mi for some i ∈ {1, 2}. This implies that Qq (cid:22)M Mi for some i ∈ {1, 2}. There exist ℓ ≥ 1, a projection p ∈ Mℓ(Mi), a nonzero partial isometry v ∈ M1,ℓ(qM )p and a unital normal ∗-homomorphism ϕ : Qq → pMℓ(Mi)p such that av = vϕ(a) for all a ∈ Qq. Write v = [v1 ··· vℓ] ∈ M1,ℓ(qM )p. In particular, we have Qvj ⊂ Pℓ k=1 vkMi for all 1 ≤ j ≤ ℓ and so (Q∩ M1)vj ⊂ Pℓ k=1 vkMi for all 1 ≤ j ≤ ℓ. Since Q∩ M1 is diffuse, by [IPP05, Theorem 1.1], we obtain that i = 1 and that vj ∈ M1 for all 1 ≤ j ≤ ℓ. Therefore vv∗ ∈ (Qq)′ ∩ qM1q is a nonzero projection such that Qvv∗ ⊂ vv∗M1vv∗. If we denote by z0 the central support of vv∗ in (Qq)′ ∩ qM1q, we have that z0 ∈ Z(Q′ ∩ M )q, z0 6= 0 and Qz0 ⊂ z0M1z0. This contradicts again the fact that z is the maximum projection in Z(Q′ ∩ M ) such that Qz ⊂ zM1z. Consequently, we obtain that z = 1 and so Q ⊂ M1. This finishes the proof of Theorem 4.1. (cid:3) Proof of Theorem D. The proof is similar to the one of Corollary B. Let Q ⊂ M be any von Neumann subalgebra such that Q ∩ M1 is diffuse. By [Io12, Lemma 2.7], there exists a central projection z ∈ Z(Q′ ∩ M ) ∩ Z(Q′ ∩ M ω) ⊂ M1 such that (Q′ ∩ M ω)z is diffuse and (Q′ ∩ M ω)z = (Q′ ∩ M )z is discrete. Choose a diffuse abelian subalgebra A ⊂ z⊥M1z⊥ and put Q = Qz ⊕ A. We have that Q ∩ M1 and Q′ ∩ M ω are diffuse. By Theorem 4.1, we obtain Q ⊂ M1 and hence Qz ⊂ zM1z. This finishes the proof of Theorem D. (cid:3) References [AH12] [Ba93] [BC13] [Br12] H. Ando, U. Haagerup, Ultraproducts of von Neumann algebras. J. Funct. Anal. 266 (2014), 6842 -- 6913. L. Barnett, Free product von Neumann algebras of type III. Proc. Amer. Math. Soc. 123 (1995), 543 -- 553. R. Boutonnet, A. Carderi, Maximal amenable subalgebras of von Neumann algebras associated with hyperbolic groups. arXiv:1310.5864 A. Brothier, The cup subalgebra of a II1 factor given by a subfactor planar algebra is maximal amenable. Pacific J. Math., to appear. arXiv:1210.8091 [CFRW08] J. Cameron, J. Fang, M. Ravichandran, S. White, The radial masa in a free group factor is [Co72] [Co74] [Co75a] [Co75b] [Co85] [Dy92] [Fa06] [Ga09] [Ge95] [Ha84] [Ho07] [Ho12a] [Ho12b] λ and closure of inner automorphisms. J. Operator maximal injective. J. Lond. Math. Soc. 82 (2010), 787 -- 809. A. Connes, Une classification des facteurs de type III. Ann. Sci. ´Ecole Norm. Sup. 6 (1973), 133 -- 252. A. Connes, Almost periodic states and factors of type III1. J. Funct. Anal. 16 (1974), 415 -- 445. A. Connes, Outer conjugacy classes of automorphisms of factors. Ann. Sci. ´Ecole Norm. Sup. 8 (1975), 383 -- 419. A. Connes, Classification of injective factors. Ann. of Math. 104 (1976), 73 -- 115. A. Connes, Factors of type III1, property L′ Theory 14 (1985), 189 -- 211. K. Dykema, Factoriality and Connes' invariant T (M) for free products of von Neumann algebras. J. reine angew. Math. 450 (1994), 159 -- 180. J. Fang, On maximal injective subalgebras of tensor products of von Neumann algebras. J. Funct. Anal. 244 (2007), 277 -- 288. M. Gao, On maximal injective subalgebras. Proc. Amer. Math. Soc. 138 (2010), 2065 -- 2070. L. Ge, On maximal injective subalgebras of factors. Adv. Math. 118 (1996), 34 -- 70. U. Haagerup, Connes' bicentralizer problem and uniqueness of the injective factor of type III1. Acta Math. 158 (1987), 95 -- 148. C. Houdayer, Construction of type II1 factors with prescribed countable fundamental group. J. Reine Angew. Math. 634 (2009), 169 -- 207. C. Houdayer, A class of II1 factors with an exotic abelian maximal amenable subalgebra. Trans. Amer. Math. Soc. 366 (2014), 3693 -- 3707. C. Houdayer, Structure of II1 factors arising from free Bogoljubov actions of arbitrary groups. Adv. Math. 260 (2014), 414 -- 457. CYRIL HOUDAYER 18 [HS90] [Io12] [IPP05] [Pe06] [Po83] [Po01] [Po03] [Po13] [Sh05] [Ta03] [Ue98] [Ue11] [Vo85] [Vo92] U. Haagerup, E. Størmer, Equivalence of normal states on von Neumann algebras and the flow of weights. Adv. Math. 83 (1990), 180 -- 262. A. Ioana, Cartan subalgebras of amalgamated free product II1 factors. Ann. Sci. ´Ecole Norm. Sup., to appear. arXiv:1207.0054 A. Ioana, J. Peterson, S. Popa, Amalgamated free products of w-rigid factors and calculation of their symmetry groups. Acta Math. 200 (2008), 85 -- 153. P. Jolissaint, Maximal injective and mixing masas in group factors. arXiv:1004.0128 [Jo10] [MvN43] F. Murray, J. von Neumann, Rings of operators. IV. Ann. of Math. 44 (1943), 716 -- 808. [Oc85] A. Ocneanu, Actions of discrete amenable groups on von Neumann algebras. Lecture Notes in Mathematics, 1138. Springer-Verlag, Berlin, 1985. iv+115 pp. J. Peterson, L2-rigidity in von Neumann algebras. Invent. Math. 175 (2009), 417 -- 433. S. Popa, Maximal injective subalgebras in factors associated with free groups. Adv. Math. 50 (1983), 27 -- 48. S. Popa, On a class of type II1 factors with Betti numbers invariants. Ann. of Math. 163 (2006), 809 -- 899. S. Popa, Strong rigidity of II1 factors arising from malleable actions of w-rigid groups I and II. Invent. Math. 165 (2006), 369 -- 408, 409 -- 451. S. Popa, A II1 factor approach to the Kadison-Singer problem. Comm. Math. Phys. 332 (2014), 379 -- 414. J. Shen, Maximal injective subalgebras of tensor products of free group factors. J. Funct. Anal. 240 (2006), 334 -- 348. M. Takesaki, Theory of operator algebras. II. Encyclopaedia of Mathematical Sciences, 125. Oper- ator Algebras and Non-commutative Geometry, 6. Springer-Verlag, Berlin, 2003. xxii+518 pp. Y. Ueda, Amalgamated free products over Cartan subalgebra. Pacific J. Math. 191 (1999), 359 -- 392. Y. Ueda, Factoriality, type classification and fullness for free product von Neumann algebras. Adv. Math. 228 (2011), 2647 -- 2671. D.-V. Voiculescu, Symmetries of some reduced free product C∗-algebras. Operator algebras and Their Connections with Topology and Ergodic Theory, Lecture Notes in Mathematics 1132. Springer- Verlag, (1985), 556 -- 588. D.-V. Voiculescu, K.J. Dykema, A. Nica, Free random variables. CRM Monograph Series 1. American Mathematical Society, Providence, RI, 1992. CNRS - Universit´e Paris-Est - Marne-la-Vall´ee, LAMA UMR 8050, 77454 Marne-la-Vall´ee cedex 2, France E-mail address: [email protected]
1905.05589
1
1905
2019-05-14T13:29:39
Stochastic aspects of the unitary dual group
[ "math.OA" ]
In this note we study asymptotic properties of the *-distribution of traces of some matrices, with respect to the free Haar trace on the unitary dual group. The considered matrices are powers of the unitary matrix generating the Brown algebra. We proceed in two steps, first computing the free cumulants of any R-cyclic family, then characterizing the asymptotic *-distributions of the traces of powers of the generating matrix, thanks to these free cumulants. In particular, we obtain that these traces are asymptotic *-free circular variables.
math.OA
math
STOCHASTIC ASPECTS OF THE UNITARY DUAL GROUP ISABELLE BARAQUIN Abstract. In this note we study asymptotic properties of the ∗-distribution of traces of some matrices, with respect to the free Haar trace on the uni- tary dual group. The considered matrices are powers of the unitary matrix generating the Brown algebra. We proceed in two steps, first computing the free cumulants of any R-cyclic family, then characterizing the asymptotic ∗- distributions of the traces of powers of the generating matrix, thanks to these free cumulants. In particular, we obtain that these traces are asymptotic ∗-free circular variables. Keywords: unitary dual group, Haar trace, free probability, cumulants, R- cyclicity, circular variables 2010 Mathematics Subject Classification: 46L54 Résumé Aspects stochastiques du groupe dual unitaire Dans cette note, nous étudions la loi asymptotique de la trace de certaines matrices, par rapport à la trace de Haar libre sur le groupe dual unitaire. Ces matrices sont les puissances de la matrice unitaire qui engendre l'algèbre de Brown. Nous procédons en deux étapes. Tout d'abord, nous calculons les cu- mulants joints d'une famille de matrices R-cyclique. Nous caractérisons ensuite la ∗-distributions asymptotique des traces considérées, à l'aide des cumulants libres. En particulier, nous obtenons que ces traces sont des variables asymp- totiquement circulaires et ∗-libres. Mots-clés : groupe dual unitaire, trace de Haar, probabilités libres, cumulants, R-cyclicité, variables circulaires Version française abrégée Dans cette note, nous étudions le comportement asymptotique des χ(up), traces des puissances de la matrice unitaire u définissant l'algèbre de Brown U nc n , par rapport à la trace de Haar libre h déterminée par Cébron et Ulrich [2], et par McClanahan [5]. Pour cela, nous remarquons tout d'abord que la propriété de h rappelée en Proposition 2.4 entraîne que {u, u∗} est une famille de matrices R-cyclique, c'est- à-dire que les cumulants libres κr de coefficients (ue)ij s'annulent dès lors que les indices ne sont pas cycliques : Définition 0.1 ([9]). Soit A une algèbre. Une famille de matrices de Mn(A), notée ij )1≤i,j≤n}1≤l≤s, est appelée R-cyclique si κr(a(l1) {Al = (a(l) ) = 0 lorsqu'il i1j1 n'est pas vrai que j1 = i2, j2 = i3, ..., jr−1 = ir et jr = i1. , . . . , a(lr) irjr Nous pouvons donc nous ramener au calcul du cumlant libre des traces des puissances d'une famille de matrices R-cyclique, dont les cumulants considérés dans la définition ci-dessus ne dépendent pas des indices i1, . . ., ir. L'application de ce résultat dans notre cas, nous permet de montrer que : 1 2 ISABELLE BARAQUIN Théorème 0.2. La famille de variables aléatoires (χ(up))p≥1 est asymptotiquement une famille ∗-libre de variables circulaires, d'espérance 0 et de covariance 1. 1. Introduction Diaconis, Shahshahani and Evans [4, 3] show that the traces of powers of a matrix chosen at random in the unitary (respectively orthogonal) group behave asymptotically like independent complex (resp. real) Gaussian random variables. Later, Banica, Curran and Speicher investigate the case of easy quantum groups in [1], and obtain similar results in the context of free probability for free orthogonal groups. In [2], Cébron and Ulrich study the Haar states according to the five notions of convolution (free, tensor, boolean, monotone and anti-monotone) of the unitary dual group U hni. They prove in particular that there is no Haar state for each of the five notions of convolution, and even no Haar trace for the boolean, monotone or anti-monotone convolution, for n ≥ 2. They also define a faithful Haar trace on U hni for the free convolution, denoted h, which is in fact equal to the state given by McClanahan in [5]. The aim of this note is to extend the study of Diaconis, Shashahani and Evans to the framework of the unitary dual group and its free Haar trace. The paper is organized as follows. We first introduce the tools of our study. Then, in the last section, we discuss the computation of the joint cumulants of traces of powers of an R-cyclic family and determine the asymptotic ∗-distribution of the traces of powers of the generating matrix, with respect to the free Haar trace. 2. Preliminaries We recall here some facts about the unitary dual group, free cumulants and R-cyclicity. 2.1. The unitary dual group. Let n ≥ 1, and U nc n , sometimes called the Brown algebra, be the noncommutative ∗-algebra generated by n2 elements {uij}1≤i,j≤n such that the matrix u = (uij)1≤i,j≤n is unitary. It is possible to endow this algebra with a structure of dual group in the sense of Voiculescu [10], U hni = (U nc n , ∆, δ, Σ), called the unitary dual group. This is a generalization of the notion of groups, like Hopf algebras, but using the free product instead of the tensor product. Definition 2.1. Let A and B be unital ∗-algebras. The free product of A and B is the unique unital ∗-algebra A ⊔ B with two ∗-homomorphisms iA : A → A ⊔ B and iB : B → A ⊔ B, such that, for all ∗-homomorphisms f : A → C and g : B → C, there exists a unique ∗-homomorphism f ⊔ g : A ⊔ B → C satisfying f = (f ⊔ g) ◦ iA and g = (f ⊔ g) ◦ iB. We sometimes refer to A and B as the left and right legs of the free product A ⊔ B. Therefore, for each a ∈ A and b ∈ B, we denote iA(a) and iB(b) by a(1) and b(2), respectively. Let f : A1 → A2 and g : B1 → B2 be unital ∗-homomorphisms between the four g : A1 ⊔ B1 → A2 ⊔ B2 unital ∗-algebras A1, A2, B1 and B2. Then we denote by f ⊔ ¯ the unital ∗-homomorphism given by the free product (iA2 ◦ f ) ⊔ (iB2 ◦ g). STOCHASTIC ASPECTS OF THE UNITARY DUAL GROUP 3 Definition 2.2. Let n ≥ 1. The unitary dual group U hni is defined by the unital ∗-algebra U nc n → C and Σ : U nc n and three unital ∗-homomorphisms ∆ : U nc n , δ : U nc n → U nc n ⊔U nc n , such that n → U nc -- U nc n is the Brown algebra, generated by the uij's satisfying n n ∀1 ≤ i, j ≤ n, X u∗ kiukj = δij = X uiku∗ jk , k=1 -- the map ∆ is a coassociative coproduct, i.e. (id⊔ ¯ k=1 n on the generators by ∆(uij) = ∆) ◦ ∆ = (∆⊔ ¯ id) ◦ ∆, given ik u(2) u(1) kj , Pk=1 id) ◦ ∆ = id = (id⊔ ¯ -- the map δ is a counit, i.e. (δ⊔ ¯ -- the map Σ is a coinverse, i.e. (Σ ⊔ id) ◦ ∆ = δ(·)1U nc δ) ◦ ∆, given by δ(uij ) = δij, n = (id ⊔ Σ) ◦ ∆, given by Σ(uij) = u∗ ji. 2.2. Free cumulants. Note that the free cumulants characterize random variables and free independence. We will use them to study the asymptotic law of χ(up), as Banica, Curran and Speicher in [1]. Let us introduce some notations: -- [s] denotes the set {1, 2, . . . , s} for each s ≥ 1, -- the set of noncrossing partitions of [s] is denoted N C(s), -- N C2(s) corresponds to the noncrossing pairings. The family of free cumulants (κr)r∈N is uniquely characterized by the correspond- ing multiplicative family of functionals satisfying the moment-cumulant formula [8, equation (11.7)]: ∀s ∈ N, ∀{ai}1≤i≤s ⊂ U nc n , h(a1 . . . as) = X π∈NC(s) κπ[a1, . . . , as] where κπ[a1, . . . , as] = QV ∈π QV ∈π In particular, they satisfy the following property: κV [a1, . . . , as] := V ={v1<...<vl} κl(av1 , . . . , avl). Proposition 2.3 ([8, equation (11.11)]). Let s, r ∈ N and p ∈ [n]r be given such that the sum of the pi's is s. Then, for any {ai}1≤i≤s ⊂ U nc n , κr (a1 . . . ap1 , ap1+1 . . . ap1+p2 , . . . , as−pr+1 . . . as) = X κπ[a1, . . . , as] π∈N C(s) π∨γp=1s where γp denotes the noncrossing partition associated to the multi-index p, i.e. γp = {{1, . . . , p1}, . . . , {s − pr + 1, . . . , s}}. Cébron and Ulrich [2] compute the free cumulants associated to the free Haar trace h of the generators of the Brown algebra uij and their adjoints (u∗)ij = u∗ ji. Proposition 2.4 ([2]). The free cumulants of (uij )1≤i,j≤n and ((u∗)ij)1≤i,j≤n in the noncommutative probability space (U nc n , h) are given as follows. Let 1 ≤ i1, j1, . . . , ir, jr ≤ n and ǫ1, . . . , ǫr be either ∅ or ∗. If the indices are cyclic (i.e. if jl−1 = il for 2 ≤ l ≤ r and i1 = jr), r is even and the ǫi are alternating we have κr ((uǫ1)i1j1 , . . . , (uǫr )ir jr ) = n1−r(−1) where Ci = (2i)! (i+1)!i! designate the Catalan numbers. Otherwise, the left-hand side term is equal to zero. r 2 −1C r 2 −1 4 ISABELLE BARAQUIN 2.3. R-cyclicity. Definition 2.5. For an algebra A, a family of matrices {Al = (a(l) in Mn(A) is called R-cyclic if κr(a(l1) i1j1 j1 = i2, j2 = i3, ..., jr−1 = ir and jr = i1. ij )1≤i,j≤n}1≤l≤s ) = 0 whenever it is not true that , . . . , a(lr ) ir jr Thus Proposition 2.4 ensures that {u, u∗} is an R-cyclic family. Moreover, the cyclic free cumulants do not depend on the indices. Since we want to look at powers of the generating matrix u, we need th following property: Proposition 2.6 ([9, Theorem 4.3]). Let (A, φ) be a noncommutative probability space. Let d be a positive integer, and let A1, . . . , As be an R-cyclic family of matri- ces in Md(A). We denote by D the algebra of scalar diagonal matrices in Md(A), and by C the subalgebra of Md(A) which is generated by {A1, . . . , As} ∪ D. Then every finite family of matrices from C is R-cyclic. In particular every finite subset of {uk}k≥1 ∪{(u∗)k}k≥1 is also an R-cyclic family of matrices. Theorem 3.1. The family of variables (χ(up))p≥1 is asymptotically a family of ∗-free circular variables of mean 0 and covariance 1. 3. Main result To prove this, let us use the R-cyclicity of u and u∗. First, let us note that χ(up)e = χ((ue)p) for any p ≥ 1 and e ∈ {∅, ∗}. This means that we can see the calculation of κs(χ(up1)e1 , . . . , χ(ups)es ) in a more general framework and compute κs(χ(Ap1 l1 )) for {Al}l∈I an R-cyclic family of matrices. ), . . . , χ(Aps ls , . . . , a(ls) Lemma 3.2. Let {Al}l∈I be an R-cyclic family of matrices such that the cumulants κπ ha(l1) isjsi depend only on the cyclicity of the indices. Let us denote by κπ (cid:2)a(l1), . . . , a(ls)(cid:3) the common value of the cumulants with cyclic indices, i.e. such that j1 = i2, . . ., js = i1. Then i1j2 κs(χ(Ap1 l1 ), . . . , χ(Aps ls )) = np+2−s X π∈N C(p) π∨γp=1p n−πκπ ha(l1), . . . , a(l1), a(l2), . . . , a(ls)i . Note that this calculation is similar to the one in the last section of [6], and } is also an R-cyclic family, by ij denotes the coefficient (i, j) we will use similar arguments. Since {Ap1 l1 Proposition 2.6, if p is the sum of all the pi's and a(l) of the matrix Al, , . . . , Aps ls κs(χ(Ap1 l1 ), . . . , χ(Aps ls )) = X κs (cid:16)a(l1) i1i2 . . . a(l1) ip1 ip1 +1 , . . . , a(ls) ip−ps+1ip−ps+2 . . . a(ls) ipi1(cid:17) . 1≤i1,...,ip≤n i1=ip1 +1=...=ip−ps+1 By Proposition 2.3, this is equal to X X 1≤i1,...,ip≤n i1=ip1 +1=...=ip−ps+1 π∈N C(p) π∨γp=1p κπ ha(l1) i1i2 , . . . , a(l1) ip1 ip1 +1 , . . . , a(ls) ipi1i . STOCHASTIC ASPECTS OF THE UNITARY DUAL GROUP 5 By definition of κπ, we can restrict ourselves to the study of a block V of π and ) with λ = (l1, . . . , l1, l2, . . . , ls) ∈ I p and where V look at κr(a is the block {v1 < . . . < vr}. In order to have a non zero contribution, the indices have to be cyclic, i.e. to satisfy (λv1 ) iv1 iv1 +1 (λvr ) ivr ivr +1 , . . . , a ∀1 ≤ j ≤ r − 1, ivj +1 = ivj+1 and ivr+1 = iv1 . Let us denote by σπ the permutation associated to the partition π, by considering the elements of a block of π in increasing order as a cycle of σπ. Hence the conditions above can be written as iγ(vi) = iσπ(vi) where γ = (1, 2, . . . , p). Since this should be π (j) for all 1 ≤ j ≤ p. Thus, we get true for each block of π, it means that ij = iγ◦σ−1 κs(χ(Ap1 l1 ), . . . , χ(Aps ls )) = X π∈N C(p) π∨γp=1p X κπ ha(l1) i1i2 , . . . , a(l1) ip1 ip1 +1 , . . . , a(ls) ipi1i . 1≤i1,...,ip≤n i1=ip1+1=...=ip−ps+1 ij =i γ◦σ −1 π (j) Since, moreover κπ ha(l1) i1i2 , . . . , a(ls) ipi1i depends only on the cyclicity of the indices, and is denoted by κπ (cid:2)a(l1), . . . , a(l1), . . . , a(ls)(cid:3), we obtain κs(χ(Ap1 l1 ), . . . , χ(Aps ls )) = X π∈N C(p) π∨γp=1p κπ ha(l1), . . . , a(l1), a(l2), . . . , a(ls)i cπ where cπ denotes the quantity #ni ∈ {1, . . . , n}p, i1 = ip1+1 = . . . = ip−ps+1, ∀1 ≤ j ≤ p, ij = iγ◦σ−1 Thanks to [7, Lemma 14], π ∨ γp = 1p if and only if σ−1 π ◦ γ separates p1, p1 + p2, . . . and p, which is equivalent to the fact that 1, p1 + 1, . . . and p − ps + 1 are all π )−s+1, where #σ denotes the in different blocks of γ ◦ σ−1 number of cycles in the cycle decomposition of a permutation σ ∈ Sp. Notes that γ ◦σ−1 π corresponds to the Kreweras complement [8] of π, denoted K(π), conjugated by γ. Hence π . Thus, cπ = n#(γ◦σ−1 π (j)o . #(γ ◦ σ−1 π ) = K(π) = p + 1 − π where π is the number of blocks of π, and then cπ = np+2−s−π, which proves the lemma. In particular, in the dual unitary group endowed with the free Haar trace, we have I = {∅, ∗} and Proposition 2.4 ensures that, if the partition π is λ-adapted with λ = (l1, . . . , l1, l2, . . . , ls), κπ ha(λ1), . . . , a(λp)i = nπ−p(−1) p 2 −π Y V ∈π C #V 2 −1 otherwise the cumulant vanishes. Here, π ∈ NC(p) is said to be λ-adapted when the following conditions are true for each block V = {v1 < . . . < vl} of the noncrossing partition π: -- #V := l ∈ 2N, -- ∀1 ≤ i ≤ l − 1, λvi 6= λvi+1 . 6 ISABELLE BARAQUIN Finally, we get, with ǫ = (e1, . . . , e1, e2, . . . , es), κs (χ(up1 )e1 , . . . , χ(ups)es ) = n2−s(−1) (−1)π Y V ∈π p 2 X π∈N C(p) π∨γp=1p ǫ−adapted C #V 2 −1 . If s > 2, it is clear that the cumulants vanish asymptotically. If s = 2, the cumulant is non zero if and only if p1 = p2 and e1 6= e2, in this case the cumulant equals 1. Moreover, if s = 1, there is no (e)-adapted partition, and the cumulant is zero. This is the description of the cumulants of a ∗-free family of circular random variables. Acknowledgements This work was supported by the French "Investissements d'Avenir" program, project ISITE-BFC (contract ANR-15-IDEX-03). References [1] T. Banica, S. Curran and R. Speicher, Stochastic aspects of easy quantum groups, Probab. Theory Related Fields 149 (2011) 435 -- 462. [2] G. Cébron and M. Ulrich, Haar states and Lévy processes on the unitary dual group, J. Funct. Anal. 270 (2016) 2769 -- 2811. [3] P. Diaconis and S. N. Evans, Linear functionals of eigenvalues of random matrices, Trans. Amer. Math. Soc. 353 (2001) 2615 -- 2633. [4] P. Diaconis and M. Shahshahani, On the eigenvalues of random matrices, J. Appl. Probab. 31 (1994) 49 -- 62. [5] K. McClanahan, C ∗-algebras generated by elements of a unitary matrix, J. Funct. Anal. 107 (1992) 439 -- 457 [6] J. A. Mingo and R. Speicher, Schwinger-Dyson equations: classical and quantum, Probab. Math. Statist., 33 (2013) 275 -- 285. [7] J. A. Mingo, R. Speicher, and E. Tan, Second order cumulants of products, Trans. Amer. Math. Soc., 361 (2009) 4751 -- 4781. [8] A. Nica and R. Speicher, Lectures on the combinatorics of free probability, Lecture Note Ser., vol. 335, Cambridge University Press, 2006. [9] A. Nica, D. Shlyakhtenko, and R. Speicher, R-cyclic families of matrices in free probability, J. Funct. Anal., 188 (2002) 227 -- 271. [10] D. Voiculescu, Dual algebraic structures on operator algebras related to free products, J. Operator Theory 17 (1987) 85 -- 98. E-mail address: [email protected] Laboratoire de mathémathiques de Besançon, UMR CNRS 6623, Université Bour- gogne Franche-Comté, 16 route de Gray, 25030 Besançon cedex, France
1112.1455
2
1112
2016-09-28T01:38:11
A Murray-von Neumann type classification of $C^*$-algebras
[ "math.OA" ]
We define type $\mathfrak{A}$, type $\mathfrak{B}$, type $\mathfrak{C}$ as well as C*-semi-finite C*-algebras. It is shown that a von Neumann algebra is a type $\mathfrak{A}$, type $\mathfrak{B}$, type $\mathfrak{C}$ or C*-semi-finite C*-algebra if and only if it is, respectively, a type I, type II, type III or semi-finite von Neumann algebra. Any type I C*-algebra is of type $\mathfrak{A}$ (actually, type $\mathfrak{A}$ coincides with the discreteness as defined by Peligrad and Zsido), and any type II C*-algebra (as defined by Cuntz and Pedersen) is of type $\mathfrak{B}$. Moreover, any type $\mathfrak{C}$ C*-algebra is of type III (in the sense of Cuntz and Pedersen). Furthermore, any purely infinite C*-algebra (in the sense of Kirchberg and Rordam) with real rank zero is of type $\mathfrak{C}$, and any separable purely infinite C*-algebra with stable rank one is also of type $\mathfrak{C}$. We also prove that type $\mathfrak{A}$, type $\mathfrak{B}$, type $\mathfrak{C}$ and C*-semi-finiteness are stable under taking hereditary C*-subalgebras, multiplier algebras and strong Morita equivalence. Furthermore, any C*-algebra $A$ contains a largest type $\mathfrak{A}$ closed ideal $J_\mathfrak{A}$, a largest type $\mathfrak{B}$ closed ideal $J_\mathfrak{B}$, a largest type $\mathfrak{C}$ closed ideal $J_\mathfrak{C}$ as well as a largest C*-semi-finite closed ideal $J_\mathfrak{sf}$. Among them, we have $J_\mathfrak{A} + J_\mathfrak{B}$ being an essential ideal of $J_\mathfrak{sf}$, and $J_\mathfrak{A} + J_\mathfrak{B} + J_\mathfrak{C}$ being an essential ideal of $A$. On the other hand, $A/J_\mathfrak{C}$ is always C*-semi-finite, and if $A$ is C*-semi-finite, then $A/J_\mathfrak{B}$ is of type $\mathfrak{A}$.
math.OA
math
A MURRAY-VON NEUMANN TYPE CLASSIFICATION OF C ∗-ALGEBRAS CHI-KEUNG NG AND NGAI-CHING WONG Abstract. We define type A, type B, type C as well as C ∗-semi-finite C ∗- algebras. It is shown that a von Neumann algebra is a type A, type B, type C or C ∗-semi-finite C ∗-algebra if and only if it is, respectively, a type I, type II, type III or semi-finite von Neumann algebra. Any type I C ∗-algebra is of type A (actually, type A coincides with the discreteness as defined by Peligrad and Zsid´o), and any type II C ∗-algebra (as defined by Cuntz and Pedersen) is of type B. Moreover, any type C C ∗-algebra is of type III (in the sense of Cuntz and Pedersen). Conversely, any separable purely infinite C ∗-algebra (in the sense of Kirchberg and Rørdam) with either real rank zero or stable rank one is of type C. We also prove that type A, type B, type C and C ∗-semi-finiteness are stable under taking hereditary C ∗-subalgebras, multiplier algebras and strong Morita equivalence. Furthermore, any C ∗-algebra A contains a largest type A closed ideal JA, a largest type B closed ideal JB, a largest type C closed ideal JC as well as a largest C ∗-semi-finite closed ideal Jsf. Among them, we have JA + JB being an essential ideal of Jsf, and JA + JB + JC being an essential ideal of A. On the other hand, A/JC is always C ∗-semi-finite, and if A is C ∗-semi-finite, then A/JB is of type A. 6 1 0 2 p e S 8 2 ] . A O h t a m [ 2 v 5 5 4 1 . 2 1 1 1 : v i X r a This paper is dedicated to Charles Batty on the occasion of his 60th birthday. 1. Introduction In their seminal works ([27], see also [26]), Murray and von Neumann defined three types of von Neumann algebras (namely, type I, type II and type III) accord- ing to the properties of their projections. They showed that any von Neumann algebra is a sum of a type I, a type II, and a type III von Neumann subalgebras. This classification was shown to be very important and becomes the basic theory for the study of von Neumann algebras (see, e.g., [20]). Since a C ∗-algebra needs not have any projection, a similar classification for C ∗-algebras seems impossible. There is, however, an interesting classification scheme for C ∗-algebras proposed by Date: October 26, 2018. 2000 Mathematics Subject Classification. 46L05, 46L35. Key words and phrases. C ∗-algebra; open projections; Murray-von Neumann type classification. The authors are supported by National Natural Science Foundation of China (11071126), and Taiwan NSC grant (99-2115-M-110-007-MY3). 1 2 CHI-KEUNG NG AND NGAI-CHING WONG Cuntz and Pedersen in [14], which captures some features of the classification of Murray and von Neumann. The classification theme of C ∗-algebras took a drastic turn after an exciting work of Elliott on the classification of AF -algebras through the ordered K-theory, in the sense that two AF -algebras are isomorphic if and only if they have the same ordered K-theory ([16]). Elliott then proposed an invariant consisting of the tracial state space and some K-theory datum of the underlying C ∗-algebra (called the Elliott invariant) which could be a suitable candidate for a complete invariant for simple separable nuclear C ∗-algebras. Although it is known recently that it is not the case (see [38]), this Elliott invariant still works for a very large class of such C ∗-algebras (namely, those satisfying certain regularity conditions as described in [18]). Many people are still making progress in this direction in trying to find the biggest class of C ∗-algebras that can be classified through the Elliott invariant (see, e.g., [17, 36]). Notice that this classification is very different from the classification in the sense of Murray and von Neumann. In this article, we reconsider the classification of C ∗-algebras through the idea of Murray and von Neumann. Instead of considering projections in a C ∗-algebra A, we consider open projections and we twist the definition of the finiteness of projections slightly to obtain our classification scheme. The notion of open projections was introduced by Akemann (in [1]). A pro- jection p in the universal enveloping von Neumann algebra (i.e. the biduals) A∗∗ of a C ∗-algebra A (see, e.g., [37, §III.2]) is an open projection of A if there is an increasing net {ai}i∈I of positive elements in A+ with limi ai = p in the σ(A∗∗, A∗)- topology. In the case when A is commutative, open projections of A are exactly characteristic functions of open subsets of the spectrum of A. In general, there is a bijective correspondence between open projections of A and hereditary C ∗- subalgebras of A (where a hereditary C ∗-subalgebra B corresponds to an open pro- jection p such that B = pA∗∗p ∩ A; see, e.g., [31, 3.11.10]). Characterisations and further developments of open projections can be found in, e.g., [2, 3, 4, 9, 15, 30, 33]. Since every element in a C ∗-algebra is in the closed linear span of its open pro- jections, it is reasonable to believe that the study of open projections will provide fruitful information about the underlying C ∗-algebra. Moreover, because of the correspondence between open projections (respectively, central open projections) and hereditary C ∗-subalgebras (respectively, closed ideals), the notion of strong Morita equivalence as defined by Rieffel (see [34] and also [11, 35]) is found to be very useful in this scheme. One might wonder why we do not consider the classification of the universal enveloping von Neumann algebras of C ∗-algebras to obtain a classification of C ∗- algebras. A reason is that for a C ∗-algebra A, its bidual A∗∗ always contains many minimum projections (see, e.g., [1, II.17]), and hence a reasonable theory of type classification cannot be obtained without serious modifications. Furthermore, A∗∗ are usually very far away from A, and information of A might not always be A MURRAY-VON NEUMANN TYPE CLASSIFICATION OF C ∗-ALGEBRAS 3 respected very well in A∗∗; for example, c and c0 have isomorphic biduals, but the structure of their open projections can be used to distinguish them (see, e.g., Example 2.1 and also Proposition 2.3(b)). As in the case of von Neumann algebras, in order to give a classification of C ∗-algebras, one needs, first of all, to consider a good equivalence relation among open projections. After some thoughts and considerations, we end up with the "spatial equivalence" as defined in Section 2, which is weaker than the one defined by Peligrad and Zsid´o in [32] and stronger than the ordinary Murray-von Neumann equivalence. One reason for making this choice is that it is precisely the "hereditarily stable version of Murray-von Neumann equivalence" that one might want (see Proposition 2.7(a)(5)), and it also coincides with the "spatial isomorphism" of the hereditary C ∗-subalgebras (see Proposition 2.7(a)(2)). Using the spatial equivalence relation, we introduce in Section 3, the notion of C ∗-finite C ∗-algebras. It is shown that the sum of all C ∗-finite hereditary C ∗- subalgebra is a (not necessarily closed) ideal of the given C ∗-algebra. In the case when the C ∗-algebra is B(H) or K(H), this ideal is the ideal of all finite rank operators on H. Moreover, through C ∗-finiteness, we define type A, type B, type C as well as C ∗-semi-finite C ∗-algebras, and we study some properties of them. In particular, we will show that these properties are stable under taking hereditary C ∗-subalgebras, multiplier algebras, unitalization (if the algebra is not unital) as well as strong Morita equivalence. We will also show that the notion of type A coincides precisely with the discreteness as defined in [32]. In Section 4, we will compare these notions with some results in the literature and give some examples. In particular, we show that any type I C ∗-algebra (see, e.g., [31]) is of type A; any type II C ∗-algebra (as defined by Cuntz and Pedersen) is of type B; any semi-finite C ∗-algebras (in the sense of Cuntz and Pedersen) is C ∗- semi-finite; any purely infinite C ∗-algebra (in the sense of Kirchberg and Rørdam) with real rank zero and any separable purely infinite C ∗-algebra with stable rank one are of type C; and any type C C ∗-algebra is of type III (as introduced by Cuntz and Pedersen). Using our arguments for these results, we also show that any purely infinite C ∗-algebra is of type III. Moreover, a von Neumann algebra M is a type A, a type B, a type C or a C ∗-semi-finite C ∗-algebra if and only if M is, respectively, a type I, a type II, a type III, or a semi-finite von Neumann algebra. In Section 5, we show that any C ∗-algebra A contains a largest type A closed B, a largest type C closed ideal J A C as well sf . It is further shown that J A B is an C is an essential ideal of A. On the other B is always of type with ideal J A A , a largest type B closed ideal J A as a largest C ∗-semi-finite closed ideal J A essential ideal of J A B + J A hand, A/J A A if one sets B := A/J A J A A , J A sf , respectively. sf , and J A C is always a C ∗-semi-finite C ∗-algebra, while B/J B , J M (A) C . We also compare J M (A) , J M (A) B, J A C and J A and J M (A) sf A + J A A + J A A B C 4 CHI-KEUNG NG AND NGAI-CHING WONG Notation 1.1. Throughout this paper, A is a non-zero C ∗-algebra, M(A) is the multiplier algebra of A, Z(A) is the center of A, and A∗∗ is the bidual of A. Furthermore, Proj(A) is the set of all projections in A, while OP(A) ⊆ Proj(A∗∗) is the set of all open projections of A. All ideals in this paper are two-sided ideals (not assumed to be closed unless specified). If x, y ∈ A∗∗ and E is a subspace of A∗∗, we set xEy := {xzy : z ∈ E}, and denote by E the norm closure of E. For any x ∈ A∗∗, we set herA(x) to be the hereditary C ∗-subalgebra x∗A∗∗x ∩ A of A (note that if u ∈ A∗∗ is a partial isometry, then herA(u) = u∗A∗∗u ∩ A = {x ∈ A : x = u∗uxu∗u} = herA(u∗u)). When A is understood, we will use the notation her(x) instead. Moreover, px is the right support projection of a norm one element x ∈ A, i.e. px is the σ(A∗∗, A∗)-limit of {(x∗x)1/n}n∈N and is the smallest open projection in A∗∗ with xpx = x. Acknowledgement: The authors would like to thank L. Brown, E. Effros and G. Elliott for giving some comments. 2. Spatial equivalence of open projections In this section, we will consider a suitable equivalence relation on the set of open projections of a C ∗-algebra. Let us start with the following example, which shows that the structure of open projections is rich enough to distinguish c and c0, while they have isomorphic biduals (see Proposition 2.3(b) below for a more general result). Example 2.1. The sets of open projections of c0 and c can be regarded as the collections X and Y, of open subsets of N and of open subsets of the one point compactification of N, respectively. As ordered sets, X and Y are not isomorphic. In fact, suppose on the contrary that there is an order isomorphism Ψ : Y → X. Then Ψ(N) is a proper open subset of N. Let k /∈ Ψ(N) and U ∈ Y with Ψ(U) = {k}. As U is a minimal element, it is a singleton set. Thus, U ⊆ N, which gives the contradiction that {k} ⊆ Ψ(N). Secondly, we give the following well-known remarks which says that open projections and the hereditary C ∗-subalgebras they define, are "hereditarily in- variant". These will clarify some discussions later on. Remark 2.2. Let B ⊆ A be a hereditary C ∗-subalgebra and e ∈ OP(A) be the open projection with herA(e) = B. (a) For any p ∈ Proj(B∗∗), one has herB(p) = herA(p). (b) OP(B) = OP(A) ∩ B∗∗. In fact, if p ∈ OP(A) ∩ B∗∗ and {ai}i∈I is an approx- imate unit in herA(p) = herB(p), then {ai}i∈I will σ(B∗∗, B∗)-converge to p and p ∈ OP(B). A MURRAY-VON NEUMANN TYPE CLASSIFICATION OF C ∗-ALGEBRAS 5 (c) If z ∈ A satisfying zz∗, z∗z ∈ B, then z ∈ B. In fact, as z∗z ∈ herA(e) = eA∗∗e ∩ A, by considering the polar decomposition of z, we see that ze = z. Similarly, we have ez = z. (d) If f ∈ OP(A), the open projections corresponding to her(e) ∩ her(f ) and the hereditary C ∗-subalgebra generated by her(e) + her(f ) are e ∧ f and e ∨ f respectively. Let jA : M(A) → A∗∗ be the canonical ∗-monomorphism, i.e. jA(x)(f ) = f (x) (x ∈ M(A), f ∈ A∗), where f ∈ M(A)∗ is the unique strictly continuous extension of f . The proposition below can be regarded as a motivation behind the study of C ∗-algebras through their open projections. It could be a known result (especially, part (a)). However, since we need it for the equivalence of (1) and (5) in Proposition 2.7(a), we give a proof here for completeness. Proposition 2.3. Suppose that A and B are C ∗-algebras, and Φ : A∗∗ → B∗∗ is a ∗-isomorphism. (a) If Φ(cid:0)jA(M(A))(cid:1) = jB(M(B)), then Φ(A) = B. (b) If Φ(OP(A)) = OP(B), then Φ(A) = B. Proof: (a) Let pA ∈ OP(M(A)) such that herM (A)(pA) = A. It is not hard to verify that pA is the support of jA, where jA : M(A)∗∗ → A∗∗ is the ∗-epimorphism induced by jA. Consider Ψ := j−1 B ◦ ΦjA(M (A)) ◦ jA : M(A) → M(B) (which is well- defined by the hypothesis). Since jB ◦ Ψ = ΦjA(M (A)) ◦ jA, we see that jB ◦ Ψ∗∗ = Φ ◦ jA (as Φ is automatically weak-*-continuous). Thus, jB(Ψ∗∗(pA)) = 1B∗∗ which implies Ψ∗∗(pA) ≥ pB. Similarly, (Ψ∗∗)−1(pB) = (j−1 A ◦ Φ−1 jB(M (B)) ◦ jB)∗∗(pB) ≥ pA and we have Ψ∗∗(pA) = pB. Consequently, Ψ(herM (A)(pA)) = herM (B)(pB) as required. (b) If a ∈ M(A)sa and U is an open subset of σ(a) = σ(Φ(jA(a))), then χU (Φ(jA(a))) = Φ(χU (jA(a))) is an element of OP(B) (by [5, Theorem 2.2] and the hypothesis). Thus, by [5, Theorem 2.2] again, we have Φ(jA(a)) ∈ jB(M(B)). A similar ar- gument shows that Φ−1(jB(M(B))) ⊆ jA(M(A)). Now, we can apply part (a) to obtain the required conclusion. (cid:3) Remark 2.4. Note that if A and B are separable and Ψ : M(A) → M(B) is a ∗-isomorphism, then Ψ(A) = B, by a result of Brown in [10]. However, the same result is not true if one of them is not separable (e.g. take A = M(B) and Ψ = id, where B is non-unital). Proposition 2.3(a) shows that one has Ψ(A) = B if (and only if) Ψ extends to a ∗-isomorphism from A∗∗ to B∗∗. We now consider a suitable equivalence relation on OP(A). A naive choice is to use the original "Murray-von Neumann equivalence" ∼Mv. However, this choice 6 CHI-KEUNG NG AND NGAI-CHING WONG is not good because [23] tells us that two open projections that are Murray-von Neumann equivalent might define non-isomorphic hereditary C ∗-subalgebras. On the other hand, one might define p ∼her q (p, q ∈ OP(A)) whenever her(p) ∼= her(q) as C ∗-algebras. The problem of this choice is that two distinct open projections of C([0, 1]) can be equivalent (if they correspond to homeomorphic open subsets of [0, 1]), which means that the resulting classification, even if possible, will be very different from the Murray-von Neumann classification. After some thoughts, we end up with an equivalence relation ∼sp on OP(A): p ∼sp q if there is a partial isometry v ∈ A∗∗ satisfying v∗ herA(p)v = herA(q) and v herA(q)v∗ = herA(p). Note that this relation is precisely the "hereditarily stable version" of the Murray- von Neumann equivalence (see Proposition 2.7(a)(5) below and the discussion fol- lowing it). In [32, Definition 1.1], Peligrad and Zsid´o introduced another equivalence relation on Proj(A∗∗): p ∼PZ q if there is a partial isometry v ∈ A∗∗ such that (2.1) v∗ herA(p) ⊆ A and v herA(q) ⊆ A. p = vv∗, q = v∗v, It is not difficult to see that ∼PZ is stronger than ∼sp, and a natural description of ∼PZ on the set of range projections of positive elements of A is given in [29, Proposition 4.3]. Moreover, we also gave in [28, Proposition 3.1] an equivalent description of ∼PZ that is similar to ∼sp but use right ideals instead of hereditary C ∗-subalgebras. However, it is now known that ∼PZ and ∼sp are actually different even for very simple kind of C ∗-algebras (see [28, Theorem 5.3]). We decide to use ∼sp as it seems to be more natural in the way of using open projections (see Proposition 2.7(a) below). Let us start with an extension of ∼sp to the whole of Proj(A∗∗). Definition 2.5. We say that p, q ∈ Proj(A∗∗) are spatially equivalent with respect to A, denoted by p ∼sp q, if there exists a partial isometry v ∈ A∗∗ satisfying (2.2) p = vv∗, In this case, we also say that the hereditary C ∗-subalgebras herA(p) and herA(q) are spatially isomorphic. v∗ herA(p)v = herA(q) and v herA(q)v∗ = herA(p). q = v∗v, It might happen that her(p) = 0 but p 6= 0 and this is why we need to consider the first two conditions in (2.2). We will see in Proposition 2.7(a) that the first two conditions are redundant if p and q are both open projections. Obviously, ∼sp is stronger than ∼Mv (for elements in Proj(A∗∗)). Moreover, if p ∼sp q, then x 7→ v∗xv is a ∗-isomorphism from her(p) to her(q), which means that ∼sp is stronger than ∼her in the context of open projections. A good point of the spatial equivalence is that open projections are stable under ∼sp, as can be seen in part (b) of the following lemma. A MURRAY-VON NEUMANN TYPE CLASSIFICATION OF C ∗-ALGEBRAS 7 Lemma 2.6. (a) ∼sp is an equivalence relation in Proj(A∗∗). (b) Let p, q ∈ Proj(A∗∗) and u ∈ A∗∗ be a partial isometry. If p is open, u∗pu = q, herA(p) ⊆ u herA(q)u∗ and herA(q) ⊆ u∗ herA(p)u, then q is open and p ∼sp q. Consequently, if p ∼sp q and p is open, then q is open. (c) If B ⊆ A is a hereditary C ∗-subalgebra and p, q ∈ Proj(B∗∗), then p and q are spatially equivalent with respect to B if and only if they are spatially equivalent with respect to A. Proof: Definition 2.5. If w ∈ A∗∗ and r ∈ Proj(A∗∗) satisfy that (a) It suffices to verify the transitivity. Suppose that p, q and v are as in p = w∗w, r = ww∗, w herA(p)w∗ = herA(r) and w∗ herA(r)w = herA(p), then the partial isometry wv gives the equivalence r ∼sp q. (b) As p is open and herA(p) is contained in the weak-*-closed subspace uA∗∗u∗, one has p ≤ uu∗. Let v := pu. Then vv∗ = p and v∗v = u∗pu = q. Moreover, it is clear that herA(p) ⊆ v herA(q)v∗ and herA(q) ⊆ v∗ herA(p)v. Now, it is easy to see that the relations in (2.2) are satisfied. Furthermore, if {ai}i∈I is an approximate unit in herA(p), then {v∗aiv} is an increasing net in herA(q) that weak-*-converges to v∗pv = q, and so q is open. The second statement follows directly from the first one. (c) Suppose that p and q are spatially equivalent with respect to A and v ∈ A∗∗ satisfies the relations in (2.2). As vv∗, v∗v ∈ B∗∗, Remark 2.2(c) tells us that v ∈ B∗∗. Now the equivalence follows from Remark 2.2(a). (cid:3) Proposition 2.7. (a) If p, q ∈ OP(A), the following statements are equivalent. (1) p ∼sp q. (2) her(q) = u∗ her(p)u and her(p) = u her(q)u∗ for a partial isometry u ∈ A∗∗. (3) her(q) ⊆ u∗ her(p)u and her(p) ⊆ u her(q)u∗ for a partial isometry u ∈ A∗∗. (4) q ≤ v∗v and v her(q)v∗ = her(p) for a partial isometry v ∈ A∗∗. (5) There is a partial isometry w ∈ A∗∗ such that p = ww∗ and {w∗rw : r ∈ OP(A); r ≤ p} = {s ∈ OP(A) : s ≤ q}. (b) If M is a von Neumann algebra and p, q ∈ Proj(M), then p ∼sp q if and only if p ∼Mv q as elements in Proj(M). (a) The implications (1) ⇒ (2) ⇒ (3) and (1) ⇒ (4) are clear. Proof: (3) ⇒ (1). Since q is open, one has q ≤ u∗u. Thus, (uq)∗uq = q and Statement (3) also holds when u is replaced by uq. As p is also open, a similar argument shows that p ≤ uqu∗ and Statement (3) holds if we replace u by v := puq and that p = vv∗. Furthermore, since vqv∗ = vv∗ = p, Lemma 2.6(b) tells us that p ∼sp q. (4) ⇒ (2). This follows from v∗ her(p)v = v∗v her(q)v∗v = her(q). (1) ⇒ (5). Notice that OP(her(p)) = {r ∈ OP(A) : r ≤ p} (see Remark 2.2(b)). Suppose that v ∈ A∗∗ satisfies (2.2) and r ∈ OP(her(p)). If {ai}i∈I is an increasing 8 CHI-KEUNG NG AND NGAI-CHING WONG net in her(p) that σ(A∗∗, A∗)-converge to r, then {v∗aiv}i∈I is an increasing net in her(q) that σ(A∗∗, A∗)-converge to v∗rv and hence v∗rv ∈ OP(her(q)). The argument for the other inclusion is similar. (5) ⇒ (1). By Statement (5), we have q = w∗pw, and the map Φ : x 7→ w∗xw is a ∗-isomorphism from her(p)∗∗ to her(q)∗∗. By Proposition 2.3(b), we see that Φ(her(p)) = her(q) and Statement (4) holds. (b) If p ∼sp q, then p ∼Mv q as elements in Proj(M ∗∗), which implies that p ∼Mv q as elements in Proj(M) (by considering the canonical ∗-homomorphism ΛM : M ∗∗ → M). Conversely, if v ∈ M satisfying p = vv∗ and q = v∗v, then clearly v∗ her(p)v = her(q). (cid:3) One can reformulate Statement (5) of Proposition 2.7(a) in the following way. There is a partial isometry w ∈ A∗∗ that induces Murray-von Neu- mann equivalences between open subprojections of p (including p) and open subprojections of q (including q). Therefore, one may regard ∼sp as the "hereditarily stable version" of the Murray- von Neumann equivalence. Moreover, if v ∈ A∗∗ satisfies the relations in (2.2), then by Lemma 2.6(b), r ∼sp v∗rv for all r ∈ OP(her(p)), which means that spatial equivalence is automatically "hereditarily stable". Remark 2.8. (a) Let p, q ∈ Proj(A∗∗). We call the unique pint ∈ OP(A) with her(p) = her(pint) the interior of p. By the bijective correspondence between hereditary C ∗-subalgebras and open projections, pint is the largest open projection dominated by p. As a direct consequence of Proposition 2.7(a), we know that pint ∼sp qint if and only if her(q) ⊆ u∗ her(p)u and her(p) ⊆ u her(q)u∗ for a partial isometry u ∈ A∗∗. (b) Suppose that p, q ∈ OP(A). One might attempt to define p . sp q if there is q1 ∈ OP(A) with p ∼sp q1 ≤ q. However, unlike the Murray-von Neumann equivalence situation, p . sp q and q . sp p does not imply that p ∼sp q. This can be shown by using a result of Lin. More precisely, it was shown in [23, Theorem 9] that there exist a separable unital simple C ∗-algebra A as well as p ∈ Proj(A) and u ∈ A such that uu∗ = 1 and p1 = u∗u ≤ p, but her(p) and A are not ∗-isomorphic. In particular, p ≁sp 1. Now, we clearly have p . sp 1. On the other hand, as u ∈ A, we have u∗Au = her(p1) and u her(p1)u∗ = A, which implies that 1 . sp p. This example also shows that the same problematic situation appears even if we replace ∼sp with the stronger equivalence relation ∼PZ as defined in (2.1) (because u ∈ A). Nevertheless, it was shown in [32, Theorem 1.13] that a weaker conclusion holds if one adds an extra assumption on either p or q, but we will not recall the details here. A MURRAY-VON NEUMANN TYPE CLASSIFICATION OF C ∗-ALGEBRAS 9 Let us end this section with the following well-known example. We give an explicit argument here for future reference. Note that parts (a) and (b) of it mean that if a, b ∈ A+ are equivalent in the sense of Blackadar (i.e., there exists x ∈ A with a = x∗x and b = xx∗; see, e.g., [29, Definition 2.1]), then their support projections are spatially equivalence (which is also a corollary of [29, Proposition 4.3], since ∼PZ is stronger than ∼sp). Example 2.9. Suppose that x ∈ A with kxk = 1. Set a = x∗x and b = xx∗. Let x = ua1/2 be the polar decomposition. (a) It is easy to see that aAa = u∗(xAx∗)u and xAx∗ = u(aAa)u∗, i.e., xAx∗ is spatially isomorphic to aAa (by Proposition 2.7(a)). (b) Notice that u(aAa)u∗ = xAx∗ ⊇ xx∗Axx∗ ⊇ xx∗xAx∗xx∗ ⊇ ua3/2Aa3/2u∗ = u(aAa)u∗, and we have xAx∗ = bAb. Similarly, x∗Ax = aAa and x∗A∗∗x = aA∗∗a, which implies that her(x) = her(a). On the other hand, as aAa is a hereditary C ∗- subalgebra of her(a) and {a1/k}k∈N is a sequence in aAa which is an approximate unit for her(a), one has aAa = her(a). Consequently, her(x) = x∗Ax. (c) Suppose that B ⊆ A is a hereditary C ∗-subalgebra and x ∈ B. Since aAa = a2Aa2, we see that aBa = aAa. Therefore, herB(x) = herA(x) by part (b). 3. C ∗-semi-finiteness and three types of C ∗-algebras As in the case of von Neumann algebras ([27]), in order to define different "types" of C ∗-algebras, we need to define "abelian" and "finite" open projections. "Abelian" open projections are defined in the same way as that of von Neumann algebras. However, in order to define "finite" open projections, we need to use our "hereditarily stable version" of Murray-von Neumann equivalence in Section 2. Note that one cannot go very far with the original Murray-von Neumann equiv- alence, because there exist p, q ∈ OP(A) with p ∼Mv q but her(p) and her(q) are not isomorphic (see [23]). Moreover, one cannot use a direct verbatim translation of the Murray-von Neumann finiteness. Definition 3.1. (a) Let q ∈ OP(A) and p ∈ Proj(qA∗∗q). The closure of p in q, denoted by ¯pq, is the smallest closed projection of her(q) that dominates p. (b) Let p, q ∈ OP(A) with p ≤ q. The projection p is said to be i. dense in q if ¯pq = q; ii. abelian if her(p) is a commutative C ∗-algebra; iii. C ∗-finite if for any r, s ∈ OP(her(p)) with r ≤ s and r ∼sp s, one has ¯rs = s. If p is dense in q, we say that her(p) is essential in her(q). We denote by OPC(A) and OPF(A) the set of all abelian open projections and the set of all C ∗-finite open projections of A, respectively. 10 CHI-KEUNG NG AND NGAI-CHING WONG The terminology "p is dense in q" is used in many places (e.g. [32]), while the terminology "essential" comes from [39]. Some people might wonder why we do not use the finiteness as defined in [14]. The reason is that we want to give a classification scheme for C ∗-algebras using open projections (and the definition of finiteness in [14] seems not related to open projections). Remark 3.2. Let p ∈ OP(A). (a) Suppose that p is abelian. If r, s ∈ OP(her(p)) satisfying r ≤ s and r ∼sp s, then r = s. Thus, p is C ∗-finite. (b) If her(p) is finite dimensional, then p is C ∗-finite. (c) One might ask why we do not define C ∗-finiteness of p in the following way: for any r ∈ OP(her(p)) with r ∼sp p, one has ¯rp = p. The reason is that the stronger condition in Definition 3.1(b) can ensure every open subprojection of a C ∗-finite projection being C ∗-finite. Such a phenomena is automatic for von Neumann algebras. (d) A hereditary C ∗-subalgebra B ⊆ A is essential in A if and only if for any non-zero hereditary C ∗-subalgebra C ⊆ A, one has B · C 6= {0}. Thus, a closed ideal I ⊆ A is essential in the sense of Definition 3.1 if and only it is essential in the usual sense (i.e., any non-zero closed ideal of A intersects I non-trivially). Definition 3.3. A C ∗-algebra A is said to be: i. C ∗-finite if 1 ∈ OPF(A); ii. C ∗-semi-finite if every element in OP(A)\{0} dominates an element in OPF(A)\ {0}; iii. of Type A if every element in OP(A) ∩ Z(A∗∗) \ {0} dominates an element in OPC(A) \ {0}; iv. of Type B if OPC(A) = {0} but each element in OP(A)∩Z(A∗∗)\{0} dominates an element in OPF(A) \ {0}; v. of Type C if OPF(A) = {0}. Let us give an equivalent form of the above abstract definition through the re- lation between open projections (respectively, central open projections) and hered- itary C ∗-subalgebras (respectively, ideals). A C ∗-algebra A is • C ∗-finite if and only if for each hereditary C ∗-subalgebra B ⊆ A, every hereditary C ∗-subalgebra of B that is spatially isomorphic to B is essential in B; • C ∗-semi-finite if and only if every non-zero hereditary C ∗-subalgebra of A contains a non-zero C ∗-finite hereditary C ∗-subalgebra; • of type A if and only if every non-zero closed ideal of A contains a non-zero abelian hereditary C ∗-subalgebra; A MURRAY-VON NEUMANN TYPE CLASSIFICATION OF C ∗-ALGEBRAS 11 • of type B if and only if A does not contain any non-zero abelian hereditary C ∗-subalgebra and every non-zero closed ideal of A contains a non-zero C ∗-finite hereditary C ∗-subalgebra; • of type C if and only if A does not contain any non-zero C ∗-finite hereditary C ∗-subalgebra. Remark 3.4. Suppose that A is simple. (a) A is either of type A, type B or type C. (b) We will see in Corollary 4.5 that A is of type A if and only if A is of type I (see, e.g., [31, 6.1.1] for its definition). Moreover, if A is of type II (in the sense of [14]), then A is of type B (by Proposition 4.7 below), while if A is purely infinite (in the sense of [13]), then A is of type C (by Proposition 4.11(a) below and [40, Theorem 1.2(ii)]). However, we do not know if the converse of the last two statements hold. A positive element a ∈ A+ is said to be C ∗-finite if her(a) (i.e., aAa) is C ∗-finite. Proposition 3.5. (a) The sum, C(A), of all abelian hereditary C ∗-subalgebras of A is a (not necessarily closed) ideal of A. If C(A)+ := C(A) ∩ A+, then C(A) coincides with the vector space span C(A)+ generated by C(A)+. (b) The sum, F(A), of all C ∗-finite hereditary C ∗-subalgebras of A is a (not nec- essarily closed) ideal of A. If F(A)+ := F(A) ∩ A+, then F(A) = span F(A)+. (c) If B ⊆ A is a hereditary C ∗-subalgebra, then C(B)+ = C(A) ∩ B+ and F(B)+ = F(A) ∩ B+. Proof: Since parts (a) and (b) follow from the arguments of [31, Proposition 6.1.7], we will only give the proof for part (c). Moreover, we will only establish the second equality as the argument for the first one is similar. As KA is a hereditary cone, the argument of part (b) tells us that F(A)+ = KA. It is clear that F(B) ⊆ F(A) ∩ B. i=1 wi, then wi ≤ w ∈ B+, which implies that wi ∈ FA ∩ B = FB (see Example 2.9(c)). Consequently, w ∈ KB as required. (cid:3) Conversely, if w ∈ KA ∩ B and w1, ..., wn ∈ FA such that w = Pn Clearly, C(A) ⊆ F(A). We will see in Theorem 5.2(d) below that the closed ideal C(A) is of type A, while F(A) is C ∗-semi-finite. Example 3.6. (a) If A is commutative, then A is of type A and is C ∗-finite. More- over, C(A) = F(A) = A. (b) Let p ∈ OP(B(ℓ2)) ⊆ B(ℓ2)∗∗ such that her(p) = K(ℓ2) (the C ∗-algebra of all compact operators). Then p 6= 1 but her(1 −p) = (0). In fact, if T ∈ her(1 −p), we have pT = 0 and ST = SpT = 0 for any S ∈ K(ℓ2), which gives T = 0. Moreover, p is dense in 1 because K(ℓ2) is an essential closed ideal of B(ℓ2) (see Remark 3.2(d)). 12 CHI-KEUNG NG AND NGAI-CHING WONG (c) If H is an infinite dimensional Hilbert space, then K(H) is a C ∗-algebra of type A, which is not C ∗-finite but is C ∗-semi-finite. In fact, as K(H) is simple and contains many rank-one projections, it is of type A. On the other hand, suppose that e ∈ Proj(K(H)) is a rank-one projection. Then 1 − e ∈ OP(K(H)) ⊆ B(H) and there is an isometry v ∈ B(H) with vv∗ = 1 − e. Thus, v∗ her(1 − e)v = K(H) and 1 − e ∼sp 1. Moreover, as e ∈ Proj(K(H)), we see that 1−e is also a closed projection and hence it is not dense in 1. Finally, as all hereditary C ∗-subalgebras of K(H) are given by projections in B(H), they are of the form K(K) for some subspaces K ⊆ H. Hence, K(H) is C ∗-semi-finite (see Remark 3.2(b)). (d) Let H be a Hilbert space. Clearly, Proj(K(H)) ⊆ OPF(B(H)). Hence, if F(H) is the set of all finite rank operators, then F(H) ⊆ F(B(H)). Suppose that B ⊆ B(H) is a C ∗-finite hereditary C ∗-subalgebra and p ∈ Proj(B). As p is C ∗- finite and pBp = pB(H)p ∼= B(K) for a subspace K ⊆ H, we see that K is finite dimensional (see part (c)) and so p ∈ K(H). Since B ⊆ B(H) is a hereditary C ∗-subalgebra, B is generated by its projections. Thus, B is a hereditary C ∗- subalgebra of K(H), and B ∼= K(H ′) for a subspace H ′ ⊆ H. The C ∗-finiteness of B again implies that dim H ′ < ∞, and B ⊆ F(H). Consequently, F(B(H)) = F(H). On the other hand, since any finite rank projection is a sum of rank-one projections and any rank-one projection belongs to C(B(H)), we see that F(H) = C(B(H)) = F(B(H)). Furthermore, by Proposition 3.5(c), we also have F(K(H)) = C(K(H)) = F(H). Remark 3.7. Let e ∈ OP(A) and z(e) be the central support of e in A∗∗. ueu∗ (see, e.g., [31, Lemma 2.6.3]), and z(e) is an open projec- (a) z(e) = supu∈UM (A) tion (see Remark 2.2(d)) with her(z(e)) being the smallest closed ideal containing her(e). (b) Recall that B := her(e) ⊆ A is said to be full if her(z(e)) = A. In this case, B is strongly Morita equivalent to A (see, e.g., [35]). Consequently, her(e) is always strongly Morita equivalent to her(z(e)). The following provides an important tool to us in this paper. An essential ingredient of its proof (in particular, part (b)) is a result of Peligrad and Zsid´o in [32]. Proposition 3.8. Let A and B be two strongly Morita equivalent C ∗-algebras. (a) A contains a non-zero abelian hereditary C ∗-subalgebra if and only if B does. (b) A contains a non-zero C ∗-finite hereditary C ∗-subalgebra if and only if B does. A MURRAY-VON NEUMANN TYPE CLASSIFICATION OF C ∗-ALGEBRAS 13 Proof: There exist a C ∗-algebra D and e ∈ Proj(M(D)) such that both A and B are full hereditary C ∗-subalgebras of D and we have A ∼= eDe and B ∼= (1 − e)D(1 − e) (see, e.g., [8, Theorem II.7.6.9]). Thus, z(e) = 1 = z(1 − e). (a) It suffices to show that A contains a non-zero abelian hereditary C ∗-subalgebra whenever D does. Let p ∈ OPC(D) \ {0}. As pz(e) = p 6= 0, we see that pueu∗ 6= 0 for some u ∈ UM (D). By replacing p with u∗pu, we may assume that pe 6= 0, and hence e herD(p)e 6= (0). If x, y ∈ herD(p) and {bj}j∈I is an approximate unit of herD(p), then biebj ∈ herD(p) which implies that xey = lim xbiebjy = lim ybiebjx = yex. Consequently, e herD(p)e is an abelian hereditary C ∗-subalgebra of A. (b) It suffices to show that if D contains a non-zero C ∗-finite hereditary C ∗- subalgebra, then so does A. Suppose that p ∈ OPF(D) \ {0}. By [32, Theorem 1.9], there exist e0, e1 ∈ OP(herD(e)) and p0, p1 ∈ OP(herD(p)) satisfying e0 + e1 e = e, p0 + p1 p = p, z(e0)z(p0) = 0 and e1 ∼PZ p1. Suppose that p1 = 0. Then e1 = 0 and z(e0) is dense in z(e) = 1 (by [32, Lemma 1.8]). This implies that z(p0) = 0, and we have a contradiction that p0 = 0 is dense in the non-zero open projection p. Therefore, p1 6= 0 and is C ∗-finite. Since herD(e1) ∼= herD(p1) (note that ∼PZ is stronger than ∼sp), we see that herD(e1) is a non-zero C ∗-finite hereditary C ∗-subalgebra of A = herD(e). (cid:3) One may also use the argument of part (b) to obtain part (a), but we keep the alternative argument since it is also interesting. Suppose that E is a full Hilbert A-module implementing the strong Morita equivalence between A and B, i.e., B ∼= KA(E) (see, e.g., [22]). If I is a closed ideal of A, then EI is a full Hilbert I-module and KI(EI) is a closed ideal of B. We recall from [32, Definition 2.1] that A is said to be discrete if any non-zero open projection of A dominates a non-zero abelian open projection. Theorem 3.9. (a) Let A and B be two strongly Morita equivalent C ∗-algebras. Then A is of type A (respectively, type B or type C) if and only if B is of the same type. (b) A C ∗-algebra A is of type A if and only if it is discrete. Proof: (a) Suppose that A is of type B. If OPC(B) 6= {0}, then OPC(A) 6= {0} (because of Proposition 3.8(a)), which is a contradiction. Let J be a non-zero closed ideal of B. As in the paragraph above, the strong Morita equivalence of A and B gives a closed ideal J0 of A that is strongly Morita equivalent to J. As J0 contains a non-zero C ∗-finite hereditary C ∗-subalgebra, so is J (by Proposition 3.8(b)). This shows that B is of type B. The argument for the other two types are similar and easier. 14 CHI-KEUNG NG AND NGAI-CHING WONG (b) It suffices to show that if A is of type A, then it is discrete. Let B ⊆ A be a non-zero hereditary C ∗-subalgebra and J ⊆ A be the closed ideal generated by B (which is strongly Morita equivalent to B; see Remark 3.7(b)). As J contains a non-zero abelian hereditary C ∗-subalgebra, so does B (by Proposition 3.8(a)). (cid:3) The following result follows from Proposition 3.8(b) and the argument of Theorem 3.9. Corollary 3.10. (a) A is C ∗-semi-finite if and only if any non-zero closed ideal of A contains a non-zero C ∗-finite hereditary C ∗-subalgebra. (b) If A is strongly Morita equivalent to a C ∗-semi-finite C ∗-algebra, then A is also C ∗-semi-finite. (c) A is of type B if and only if it is C ∗-semi-finite and anti-liminary (i.e., it does not contain any non-zero commutative hereditary C ∗-subalgebra). Remark 3.11. (a) As in the case of von Neumann algebra, strong Morita equivalence does not preserve C ∗-finiteness. In fact, for any C ∗-algebra A, the algebra A⊗K(ℓ2) is not C ∗-finite (using the same argument as Example 3.6(c); note that 1 ⊗ (1 − e) is both an open and a closed projection of A ⊗ K(ℓ2)). Consequently, any stable C ∗-algebra is not C ∗-finite. (b) By Remark 3.7(b), Theorem 3.9(a) and Corollary 3.10(b), any type A, type B, type C or C ∗-semi-finite hereditary C ∗-subalgebra is contained in a closed ideal of the same type. Recall that a C ∗-algebra A has real rank zero in the sense of Brown and Pedersen if the set of elements in Asa with finite spectrum is norm dense in Asa (see, e.g., [12, Corollary 2.6]). The following result follows from Theorem 3.9(b), Corollary 3.10(c) as well as the fact that any hereditary C ∗-subalgebra of a real rank zero C ∗-algebra is again of real rank zero (see, e.g., [12, Corollary 2.8]). Corollary 3.12. Let A be a C ∗-algebra with real rank zero. (a) A is of type A if and only if every projection in Proj(A) \ {0} dominates an abelian projection in Proj(A) \ {0}. (b) A is of type B if and only if every projection in Proj(A) \ {0} is non-abelian but dominates a C ∗-finite projection in Proj(A) \ {0}. (c) A is of type C if and only if A does not contain any non-zero C ∗-finite projec- tion. (d) A is C ∗-semi-finite if and only if every projection in Proj(A) \ {0} dominates a C ∗-finite projection in Proj(A) \ {0}. A MURRAY-VON NEUMANN TYPE CLASSIFICATION OF C ∗-ALGEBRAS 15 Remark 3.13. Suppose that A is a C ∗-finite C ∗-algebra with real rank zero. If r, p ∈ Proj(A) such that r ≤ p and there exists u ∈ A with uu∗ = r and u∗u = p, then r ∼sp p and so, r = ¯rp = p. Corollary 3.14. If A is of real rank zero, then the closures of the ideals C(A) and F(A) (see Proposition 3.5) are the closed linear spans of abelian projections and of C ∗-finite projections in Proj(A), respectively. If B ⊆ A is a C ∗-finite hereditary C ∗-subalgebra, then B is the closed Proof: linear span of Proj(B) ∩ OPF(B). Thus, F(A) lies inside the closed linear span of Proj(A) ∩ OPF(A). Conversely, it is clear that Proj(A) ∩ OPF(A) ⊆ F(A). The argument for the statement concerning C(A) is similar. (cid:3) Corollary 3.15. Let A be of type A (respectively, of type B, of type C or C ∗-semi- finite). (a) If B is a hereditary C ∗-subalgebra of A, then B is of type A (respectively, of type B, of type C or C ∗-semi-finite). (b) If A is a hereditary C ∗-subalgebra of A0 that generates an essential ideal I ⊆ A0, then A0 is of type A (respectively, of type B, of type C or C ∗-semi-finite). Proof: (a) As any hereditary C ∗-subalgebra of B is a hereditary C ∗-subalgebra of A, this result follows directly from the definitions, Theorem 3.9(b) and Corollary 3.10(c). (b) Note that A is strongly Morita equivalent to I and any hereditary C ∗-subalgebra of A0 intersects I non-trivially. Thus, this part follows from the definitions, The- orem 3.9 and Corollary 3.10. (cid:3) Consequently, we have the following result. Corollary 3.16. Suppose that A is non-unital, and A is the unitalization of A. Then A is of type A (respectively, of type B, of type C or C ∗-semi-finite) if and only if A is of type A (respectively, of type B, of type C or C ∗-semi-finite). The same is true when A is replaced by M(A). Our next lemma is probably well-known, but we give a simple argument here for completeness. Lemma 3.17. Let e, f ∈ OP(A) and p, q ∈ OP(A) ∩ Z(A∗∗). (a) ep ∈ OP(A) and her(ep) = her(e) ∩ her(p). (b) If e 6= 0 and her(e) ⊆ her(p) + her(q), then her(e) ∩ her(p) 6= (0) or her(e) ∩ her(q) 6= (0). (c) If z(e)z(f ) = 0, then her(e) + her(f ) = her(e + f ). 16 CHI-KEUNG NG AND NGAI-CHING WONG Proof: Parts (a) and (c) are obvious (see Remark 2.2(d)). To show part (b), note that as her(p)+her(q) ⊆ her(p+q −pq), we have e ≤ p+q −pq. If ep = 0 = eq, one obtains a contradiction that e = e(p + q − pq) = 0. Thus, the conclusion follows from part (a). (cid:3) Lemma 3.18. If {pi}i∈I is a family in OPF(A) with z(pi)z(pj) = 0 for i 6= j, then p :=Pi∈I pi ∈ OPF(A). It is clear that p is an open projection and z(p) = Pi∈I z(pi). Suppose Proof: that r, q ∈ OP(her(p)) with r ≤ q and r ∼sp q. Let u ∈ A∗∗ with q = u∗u and u her(q)u∗ = her(r). For any i ∈ I, we set qi := z(pi)q, ri := z(pi)r ∈ OP(A) i ui and and ui := z(pi)u. It is easy to see that q = Pi∈I qi, r = Pi∈I ri, qi = u∗ qj(cid:1)(cid:1) = her(qi). z(pi) her(q) = z(pi)(cid:0) her(qi) + her(cid:0)Xj∈I\{i} ri ≤ qi ≤ z(pi)p = pi. By Lemma 3.17(c), we see that Similarly, z(pi) her(r) = her(ri) and we have ui her(qi)u∗ i = her(ri). By Proposition 2.7(a), we know that ri ∼sp qi and the C ∗-finiteness of pi tells us that ri is dense in qi. If e ∈ OP(her(q)) with re = 0, then ei := z(pi)e ∈ OP(her(qi)) with riei = 0, which means that ei = 0 (because ri is dense in q as required. qi = qi). Consequently, e =Pi∈I ei = 0 and r (cid:3) Part (a) of the following result is the equivalence of statements (i) and (iii) in [32, Theorem 2.3], while part (b) follows from the proof of [32, Theorem 2.3], Lemma 3.18, Theorem 3.9(a) and Corollary 3.15(b). Proposition 3.19. (a) A C ∗-algebra A is of type A if and only if there is an abelian hereditary C ∗-subalgebra of A that generates an essential closed ideal of A. (b) A C ∗-algebra A is C ∗-semi-finite if and only if there is a C ∗-finite hereditary C ∗-subalgebra of A that generates an essential closed ideal of A. 4. Comparison with existing theories In this section, we compare our "Murray-von Neumann type classification" with existing results in the literature. Through these comparisons, we obtain many new examples of C ∗-algebras of different types. Moreover, we will show that a von Neumann algebra is a type A, type B, type C or C ∗-semi-finite C ∗-algebra if and only if it is, respectively, a type I, type II, type III or semi-finiteness von Neumann algebra. A MURRAY-VON NEUMANN TYPE CLASSIFICATION OF C ∗-ALGEBRAS 17 4.1. Comparison with type I algebras. Recall that a C ∗-algebra A is said to be of type I if for any irreducible repre- sentation (π, H) of A, one has K(H) ⊆ π(A). We have already seen in Theorem 3.9(b) that type A is the same as discreteness. Thus, the following result is a direct consequence of [32, Theorem 2.3]. Note that one can also obtain it using Theorem 3.9(a) and [6, Theorems 1.8 and 2.2]. Corollary 4.1. Any type I C ∗-algebra is of type A. The converse of the above is not true even for real rank zero C ∗-algebras, as can be seen in the following example. Example 4.2. Example 3.6(c) and Corollary 3.15(b) tell us that B(ℓ2) is of type A. However, B(ℓ2) is not a type I C ∗-algebra (see, e.g., [31, 6.1.2]). Proposition 4.3. (a) A is of type I if and only if every primitive quotient of A is of type A. (b) If A is of type A and contains no essential primitive ideal, then A is of type I. (a) Because of Corollary 4.1 and the fact that quotients of type I C ∗- Proof: algebras are also of type I, we only need to show the "if" part. Let π : A → B(H) be an irreducible representation and B be a non-zero abelian hereditary C ∗-subalgebra of A/ ker π. If π : A/ ker π → B(H) is the induced representation, the restriction πB : B → B(π(B)H) is non-zero and irreducible. Thus, dim π(B)H = 1 and π(b) is a rank-one operator (and hence is compact) for any b ∈ B \ {0}. This shows that π(A/ ker π) ∩ K(H) 6= (0), and π(A) ⊇ K(H). (b) Suppose that π : A → B(H) is an irreducible representation and J is a non-zero closed ideal of A with J ∩ ker π = (0). If B ⊆ J is a non-zero abelian hereditary C ∗-subalgebra, the restriction πB : B → B(π(B)H) is non-zero and irreducible. The same argument as in part (a) tells us that π(A) ⊇ K(H). (cid:3) Remark 4.4. (a) Proposition 4.3(a) actually shows that A is of type I if and only if any primitive quotient contains a non-zero abelian hereditary C ∗-subalgebra, which is likely to be a known fact. (b) If every quotient of B(ℓ2) were of type A, then Proposition 4.3(a) told us that B(ℓ2) were a type I C ∗-algebra, which contradicted [31, 6.1.2]. Consequently, not every quotient of a type A C ∗-algebra is of type A. If A is simple and of type A, then by Proposition 4.3(b), it is of type I. This, together with Example 3.6(c), gives the following. 18 CHI-KEUNG NG AND NGAI-CHING WONG Corollary 4.5. If A is a simple C ∗-algebra of type A, then A = K(H) for some Hilbert space H. If, in addition, A is C ∗-finite, then A = Mn for some positive integer n. 4.2. Comparison with type II and (semi-)finite C ∗-algebras. The following is a direct consequence of Remark 3.4(a) and Corollary 4.5. Corollary 4.6. Any infinite dimensional C ∗-finite simple C ∗-algebra is of type B. In the following, we compare type B and type C with the notions of type II and type III as introduced by Cuntz and Pedersen in [14]. Let us recall from [14, p. 140] that x ∈ A+ is said to be finite if for any sequence {zk}k∈N in A with k. We also recall that A is said to be finite (respectively, semi-finite) if every x ∈ A+ \ {0} is finite (respectively, x dominates a non-zero finite element). Furthermore, A is said to be of type II if it is anti-liminary and finite, while A is said to be of type III if it has no non-zero finite elements (see [14, p. 149]). k ≤ x will imply x = P∞ kzk, the condition P∞ k=1 zkz∗ k=1 z∗ x = P∞ k=1 zkz∗ Let Ts(A) be the set of all tracial states on A. It follows from [14, Theorem 3.4] that Ts(A) separates points of A+ if A is finite. Proposition 4.7. If Ts(A) separates points of A+, then A is C ∗-finite. Conse- quently, if A is finite, then A is C ∗-finite. Proof: Suppose on the contrary that there exist r, q ∈ OP(A) with r ≤ q, r ∼sp q but ¯rq (cid:12) q. For any τ ∈ Ts(A), if τ is the normal tracial state on A∗∗ extending τ , then τ (r) = τ (q) (because r = vv∗ and q = v∗v for some v ∈ A∗∗). Moreover, if {ai}i∈I is an approximate unit in her(r), one has τ (r) = lim τ (ai). Since ¯rq (cid:12) q, there exists s ∈ OP(her(q))\{0} with rs = 0. If x ∈ her(s)+ with kxk = 1, one can find τ0 ∈ Ts(A) with τ0(x) > 0. Thus, we have τ0(ai) + τ0(x) ≤ τ0(q) (as ai + x ≤ q because aix = 0), which gives the contradiction that τ0(r) + τ0(x) ≤ τ0(q). (cid:3) As in [14], we denote by F A the set of all finite elements in A+. If B ⊆ A is a hereditary C ∗-subalgebra, then F B = F A ∩ B. In fact, it is obvious that F A ∩ B ⊆ F B. Conversely, suppose that x ∈ F B. k=1 zkz∗ k kzk. Since B+ is a hereditary cone of A+, we have y ∈ B+ and k ∈ B+ (k ∈ N). By Remark 2.2(c), we know that zk ∈ B and so, y = x Consider y ∈ A+ and a sequence {zk}k∈N in A satisfying y ≤ x, y = P∞ and x = P∞ k=1 z∗ z∗ kzk, zkz∗ as required. A MURRAY-VON NEUMANN TYPE CLASSIFICATION OF C ∗-ALGEBRAS 19 Corollary 4.8. (a) A is semi-finite if and only if every non-zero hereditary C ∗- subalgebra of A contains a non-zero finite hereditary C ∗-subalgebra. (b) If A is semi-finite (respectively, of type II), then A is C ∗-semi-finite (respec- tively, of type B). (a) For the necessity, let B ⊆ A be a non-zero hereditary C ∗-subalgebra. Proof: If y ∈ B+ \ {0}, there is x ∈ F A \ {0} with x ≤ y. By [14, Lemma 4.1] and [14, Theorem 4.8] as well as their arguments, one can find a non-zero finite hereditary C ∗-subalgebra of her(x). More precisely, let f ∈ C(σ(x))+ such that f vanishes in a neighborhood of 0 and f (t) ≤ t ≤ f (t) + kxk (t ∈ σ(x)). There exists g ∈ C(σ(x))+ 2 and λ > 0 such that f = f g and g(t) < λt (t ∈ σ(x)). Then g(x) ∈ F A and f (x) = f (x)g(x), i.e., f (x) ∈ F0 := {a ∈ A+ : a = ay for some y ∈ F A} ⊆ F A. For any z ∈ her(f (x))+, we have zg(x) = z and z ∈ F0 ∩ her(f (x)) ⊆ F A ∩ her(f (x)) = F her(f (x)). Thus, her(f (x)) is a non-zero finite hereditary C ∗-subalgebra of her(x). For the sufficiency, let y ∈ A+ \ {0} and C be a non-zero finite hereditary C ∗-subalgebra of her(y). Observe that C+ = F C = F A ∩ C. Take any x ∈ C+ with kxk = 1. Since x1/2yx1/2 ≤ kykx ∈ F A, we know, from [14, Lemma 4.1], that y1/2xy1/2 = y1/2x1/2(y1/2x1/2)∗ ∈ F A. Moreover, as y1/2xy1/2 ≤ y, we see that A is semi-finite. (b) This follows from part (a), Proposition 4.7 and Corollary 3.10(c). (cid:3) Example 4.9. (a) If A is an infinite dimensional simple C ∗-algebra with a faithful tracial state, then A is of type B (by Corollary 4.6 and Proposition 4.7). In particular, if Γ is an infinite discrete group such that C ∗ r (Γ) is simple (see, e.g., [7] for some examples of such groups), then C ∗ (b) Every simple AF algebra which is not of the form K(H) is of type B (because of [14, Proposition 4.11] as well as Corollaries 4.5 and 4.8(b)). r (Γ) is of type B. 4.3. Comparison with type III and purely infinite C ∗-algebras. If a C ∗-algebra A contains a non-zero (positive) finite element x, the argument of the necessity of Corollary 4.8(a) tells us that there is a non-zero finite hereditary C ∗-subalgebra of A, and hence A is not of type C, because of Proposition 4.7. This gives the following corollary. Corollary 4.10. If A is of type C, then it is of type III. 20 CHI-KEUNG NG AND NGAI-CHING WONG In the following, we will also compare type C with the notion of pure infinity as defined by Cuntz (in the case of simple C ∗-algebras) and by Kirchberg and Rørdam (in the general case). Suppose that a ∈ Mn(A) and b ∈ Mm(A) (m, n ∈ N). As in [21, Definition 2.1], we say that a - b relative to Mm,n(A) if there is a sequence {xk}k∈N in Mm,n(A) such that kx∗ kbxk − ak → 0. An element a ∈ A is said to be properly infinite if a ⊕ a - a relative to M1,2(A). Moreover, A is said to be purely infinite if every element in A+ is properly infinite (see [21, Theorem 4.16]). Note that if A is simple, this notion coincides with the one in [13], namely, every hereditary C ∗-subalgebra of A contains a non-zero infinite projection (see, e.g., the work of Lin and Zhang in [24]). Proposition 4.11. (a) If A has real rank zero and is purely infinite, then it is of type C. (b) If A is a separable purely infinite C ∗-algebra with stable rank one, then A is of type C. Proof: (a) By [21, Theorem 4.16], any element p ∈ Proj(A) \ {0} is properly infinite and hence is infinite, in the sense that there exist q ∈ Proj(A) and v ∈ A such that q ≤ p, v∗v = p and q = vv∗ (see, e.g., [21, Lemma 3.1]). Thus, p ∼sp q (as v ∈ A) but q is not dense in p (because p − q ∈ Proj(A) \ {0}). Consequently, any non-zero projection in A is not C ∗-finite, and Corollary 3.12(c) shows that A is of type C. (b) Suppose on contrary that A contains a non-zero C ∗-finite hereditary C ∗- subalgebra B and we take any z ∈ B+ with kzk = 1. By [21, Theorem 4.16], one has z ⊕ z - z ⊕ 0 relative to M2(A), and so, z ⊕ z - z ⊕ 0 relative to M2(her(z)) (by [21, Lemma 2.2(iii)]). Thus, [29, Proposition 4.13] implies pz ⊕ pz = pz⊕z -Cu pz⊕0 = pz ⊕ 0 (see [29, §3] for the meaning of -Cu). Moreover, one obviously has pz⊕0 -Cu pz⊕z. Since A has stable rank one, we conclude that pz ⊕ pz ∼PZ pz ⊕ 0 (by [29, 6.2(1)'&(2)']) and hence pz ⊕ pz ∼sp pz ⊕ 0. This means that M2(her(z)) is spatially isomorphic (and hence ∗-isomorphic) to its hereditary C ∗-subalgebra her(z) ⊕ (0), which is not essential in M2(her(z)) (because (0) ⊕ her(z) is a non- zero hereditary C ∗-subalgebra and we can apply Remark 3.2(d)). As her(z) is ∗-isomorphic to her(z) ⊕ (0) and hence to M2(her(z)), we know that her(z) is also spatially isomorphic to an inessential hereditary C ∗-subalgebra. Consequently, her(z) is not C ∗-finite, which contradicts the fact that B is C ∗-finite. (cid:3) One may regard parts (a) and (b) of the above as two extremes, because any real rank zero C ∗-algebras has plenty of projections, while a purely infinite C ∗- algebra with stable rank one is stably projectionless. Let us make the following conjecture. Conjecture 4.12. Every purely infinite C ∗-algebra is of type C. A MURRAY-VON NEUMANN TYPE CLASSIFICATION OF C ∗-ALGEBRAS 21 On the other hand, by Proposition 4.11 and Corollary 4.10, we know that any separable purely infinite C ∗-algebra A having real rank zero or stable rank one is of type III. This implication actually holds without these extra assumptions, as can be seen in the following proposition, which gives another evidence for Conjecture 4.12. Note that this proposition also implies [21, Proposition 4.4]. To show this result, let us recall the following notation from [29, p. 3476]. For any ǫ > 0, let fǫ : R+ → R+ be the function fǫ(t) = (t/ǫ 1 if t ∈ [0, ǫ) if t ∈ [ǫ, ∞). If µ ∈ Ts(A) and a ∈ A+, we define dµ(a) := supǫ>0 µ(fǫ(a)) (note that the definition in [29] is for tracial weights but we only need tracial states here). Proposition 4.13. Any purely infinite C ∗-algebra A is of type III. Proof: Suppose on the contrary that F A 6= {0}. By the argument of the necessity of Corollary 4.8(a), there is z ∈ A+ with kzk = 1 and her(z) being a finite C ∗- algebra. By the argument of Proposition 4.11(b), one has z ⊕ z - z ⊕ 0 relative to M2(her(z)). By [29, Remark 2.5], we see that dµ(z ⊕ z) ≤ dµ(z ⊕ 0) for each µ ∈ Ts(M2(her(z))). Now, if τ ∈ Ts(her(z)), then τ ⊗ Tr2 ∈ Ts(M2(her(z))) (where Tr2 is the canonical tracial state on M2), and the above tells us that τ (fǫ(z)) = sup ǫ>0 sup ǫ>0 which gives dτ (z) = 0 and hence τ (z) = 0. This contradicts [14, Theorem 3.4]. (cid:3) (τ ⊗Tr2)(fǫ(z) ⊕fǫ(z)) ≤ sup ǫ>0 (τ ⊗Tr2)(fǫ(z) ⊕0) = sup ǫ>0 τ (fǫ(z)) , 2 If one can show that her(a) is not C ∗-finite, for every properly infinite positive element a in any C ∗-algebra, then the above conjecture is verified. Let us recall from [21, Proposition 3.3(iv)] that a ∈ A+ is properly infinite if and only if there are sequences {xn}n∈N and {yn}n∈N in her(a) such that x∗ n → a and x∗ nyn → 0. The following remark tells us that if a ∈ A+ satisfies a stronger condition than the above, then her(a) is indeed non-C ∗-finite. nxn → a, yny∗ Remark 4.14. Let a ∈ A+ such that there exist x, y ∈ her(a) with x∗x = a = y∗y as well as x∗y = 0. By Example 2.9(a)&(b), we see that her(a) is spatially isomorphic to its hereditary C ∗-subalgebra her(x∗). As her(x∗) her(y∗) = (0), we see that her(x∗) is not essential in her(a). Thus, her(a) is not C ∗-finite. Example 4.15. For any AF -algebra B, the C ∗-algebra O2 ⊗ B is purely infinite (by [21, Proposition 4.5]) and is of real rank zero (by [12, Theorem 3.2]), which means that O2 ⊗ B is of type C (by Proposition 4.11(a)). Note that one may replace O2 22 CHI-KEUNG NG AND NGAI-CHING WONG with any unital, simple, separable, purely infinite, nuclear C ∗-algebra (which has real rank zero because of [40, Theorem 1.2(ii)]). 4.4. The case of von Neumann algebras. In this subsection, we consider the case of von Neumann algebras. Let us start with the following lemma. Note that the necessity of part (a) of this result follows directly from Proposition 4.7, but we give an alternative proof here as this argument is also interesting (see Remark 4.17 below). Lemma 4.16. (a) Let M be a von Neumann algebra. Then p ∈ Proj(M) is finite as a projection in M if and only if it is C ∗-finite. (b) The ideal F(M) in Proposition 3.5 is a dense subalgebra of the ideal J(M) generated by finite projections (as defined in [19]). (a) Assume that p is finite. Let ΛM : M ∗∗ → M be the canonical ∗- Proof: epimorphism. If q ∈ OP(pMp), then herM (q) ⊆ herM (ΛM (q)) and ΛM (q) ≤ p, which imply that ΛM (q) = ¯qp (notice that ¯qp ∈ pMp because of [2, Theorem II.1]). Suppose that r, q ∈ OP(pMp) such that r ≤ q and r ∼sp q. Consider w ∈ M ∗∗ satisfying q = ww∗, r = w∗w, w∗ her(q)w = her(r) and w her(r)w∗ = her(q). Define v := ΛM (w). Then ΛM (q) = vv∗ and ΛM (r) = v∗v. Since ΛM (r) ≤ ΛM (q) ≤ p, the finiteness of p tells us that ¯rp = ΛM (r) = ΛM (q) = ¯qp. If ¯rq (cid:12) q, there is e ∈ OP(her(q)) \ {0} with re = 0. Since e ∈ OP(her(p)), we obtain a contradiction that ¯rp 6= ¯qp (as r ≤ p − e but q (cid:2) p − e). This shows that p is C ∗-finite. Conversely, if p is C ∗-finite, then Remark 3.13 implies that p is finite. (b) This follows from part (a) and Corollary 3.14. (cid:3) Remark 4.17. (a) Let p ∈ M be a finite projection. If r ∈ Proj(pMp) with r ∼sp p, then Lemma 4.16(a) and Remark 3.13 tell us that r = p. The same is true if we relax the assumption to r ∈ OP(pMp). In fact, we first notice that the C ∗- finiteness of p gives ¯rp = p. Moreover, suppose that w ∈ M ∗∗ and v ∈ M are as in the proof of Lemma 4.16 for the case when q = p. Then vv∗ = p = ¯rp = v∗v. This means that v is a unitary in pMp. As v her(r)v∗ = ΛM (w her(r)w∗) = pMp, we have her(r) = pMp and hence r = p. (b) If A is a C ∗-algebra and p ∈ OP(A) satisfying ¯rp = ¯qp for any r, q ∈ OP(her(p)) with r ≤ q and r ∼sp q, then by the argument of Lemma 4.16, we see that p is C ∗-finite. The following is a direct consequence of Lemma 4.16 and Corollary 3.12. A MURRAY-VON NEUMANN TYPE CLASSIFICATION OF C ∗-ALGEBRAS 23 Theorem 4.18. Let M be a von Neumann algebra. (a) M is of type A if and only if M is a type I von Neumann algebra. (b) M is of type B if and only if M is a type II von Neumann algebra. (c) M is of type C if and only if M is a type III von Neumann algebra. (d) M is C ∗-semi-finite if and only if M is a semi-finite von Neumann algebra. 5. Factorisations In this section, we give two factorization type results for general C ∗-algebras. Let us first state the following easy lemma. Notice that if A contains a non-zero abelian hereditary C ∗-subalgebra B, the closed ideal generated by B is of type A (by Corollary 3.15(b) and Remark 3.7(b)), and the same is true for C ∗-finite hereditary C ∗-subalgebra. Lemma 5.1. If A is not of type C, then A contains a non-zero closed ideal of either type A or type B. The following is our first factorization type result, which mimics the corre- sponding situation for von Neumann algebras. Theorem 5.2. Let A be a C ∗-algebra. (a) There is a largest type A (respectively, type B, type C and C ∗-semi-finite) hereditary C ∗-subalgebra JA (respectively, JB, JC and Jsf) of A, which is also an ideal of A. (b) JA, JB and JC are mutually disjoint such that JA + JB + JC is an essential closed ideal of A. If eA, eB, eC ∈ OP(A) ∩ Z(A∗∗) with JA = her(eA), JB = her(eB) and JC = her(eC), then 1 = eA + eB 1 + eC. (c) JA + JB is an essential closed ideal of Jsf. If esf ∈ OP(A) with Jsf = her(esf), then esf = eA esf + eB. (d) The closure of C(A) and F(A) (in Proposition 3.5) are essential closed ideals of JA and Jsf, respectively. Proof: (a) We first consider the situation of type B hereditary C ∗-subalgebra. Let JB be the set of all type B closed ideals of A. If JB = {(0)}, then JB := (0) is the largest type B hereditary C ∗-subalgebra of A (see Remark 3.11(b)). Suppose that there exist distinct elements J1 and J2 in JB. If J1 + J2 contains a non-zero abelian hereditary C ∗-algebra B, then by Lemma 3.17(b), one of the two abelian hereditary C ∗-subalgebras B∩J1 and B∩J2 is non-zero, which contradicts J1, J2 ∈ JB. On the other hand, consider a non-zero closed ideal I of J1 +J2. Again, by Lemma 3.17(b), we may assume that the closed ideal I ∩ J1 is non-zero. Thus, I ∩ J1 contains a 24 CHI-KEUNG NG AND NGAI-CHING WONG non-zero C ∗-finite hereditary C ∗-subalgebra B. This shows that J1 + J2 ∈ JB and JB is a directed set. For any ideal J of A, we consider eJ ∈ OP(A) ∩ Z(A∗∗) with J = her(eJ ). Set Then eJB = w∗-limJ∈JB eJ . If there is p ∈ OPC(A) \ {0} such that her(p) ⊆ JB, then JB := XJ∈JB J. p = peJB = peJBp = w∗-limJ∈JBpeJ p, and one can find J ∈ JB with the abelian algebra her(p) ∩ J being non-zero (because of Lemma 3.17(a)), which is absurd. On the other hand, suppose that I is a non-zero closed ideal of JB. The argument above tells us that I ∩ J 6= (0) for some J ∈ JB, and hence it contains a non-zero C ∗-finite hereditary C ∗-subalgebra. Consequently, JB ∈ JB. Finally, if B ⊆ A is a hereditary C ∗-subalgebra of type B, then, by Remark 3.11(b), one has B ⊆ JB. The arguments for the statements concerning JA, JC and Jsf are similar and easier. (b) The first statement follows directly from Lemma 5.1 (any non-type C ideal interests either JA or JB). For the second statement, one obviously has eA + eB ≤ 1 − eC. Suppose that p ∈ OP(A) with eA + eB ≤ 1 − p. We have p(eA + eB) = 0. If p (cid:2) eC, then her(p) will contain a hereditary C ∗-subalgebra of either type A or type B (by Lemma 5.1) and Lemma 3.17(a) will give a contradiction that either peA 6= 0 or peB 6= 0. Thus, 1 − eC is the smallest closed projection dominating eA + eB. (c) This follows from a similar (but easier) argument as part (b). (d) Clearly, F(A) ⊆ Jsf and C(A) ⊆ JA (see Remark 3.11(b)). Their closure are both essential because of Proposition 3.19. (cid:3) By Proposition 3.19, there is an abelian (respectively, a C ∗-finite) hereditary C ∗-subalgebra that generates an essential ideal of JA (respectively, of JB). More- over, by [32, Theorem 2.3(vi)], the largest type I closed ideal Apostlim of A is an essential ideal of JA. A = A⊥ 0 = J ⊥. Remark 5.3. For any closed ideal J of A, we write J ⊥ for the closed ideal {a ∈ A : aJ = (0)}. It is easy to see that if J0 is an essential ideal of J, then J ⊥ postlim is the largest anti-liminary hereditary C ∗-subalgebra of A (note (a) J ⊥ that aJAa is a hereditary C ∗-subalgebra of JA for every a ∈ A+). Furthermore, JB + JC is an essential ideal of J ⊥ (b) J ⊥ (c) J ⊥ sf = (JA + JB)⊥ = JC. A ∩ Jsf = JB (compare with Corollary 3.10(c)). A (by Lemma 5.1). A MURRAY-VON NEUMANN TYPE CLASSIFICATION OF C ∗-ALGEBRAS 25 From now on, we denote by J A sf , respectively, the largest type A, the largest type B, the largest type C and the largest C ∗-semi-finite closed ideals of a C ∗-algebra A. C and J A A , J A B, J A The following is a direct application of Theorem 4.18. Corollary 5.4. Let M be a von Neumann algebra. If MI , MII and MIII are respectively the type I summand, the type II summand and the type III summand of M, then J M B = MII and J M A = MI, J M C = MIII. Our next theorem is the second factorization type result, which seems to be more interesting for C ∗-algebra (c.f. [14, Proposition 4.13]). Theorem 5.5. Let A be a C ∗-algebra. (a) A/J A C is C ∗-semi-finite and A/(J A (b) If A is C ∗-semi-finite, then A/J A A )⊥ is of type A. B is of type A. (a) Assume, without loss of generality, that A/J A C 6= (0) and consider Proof: C to be the canonical map. Let I be a non-zero closed ideal of A/J A Q : A → A/J A C C , one knows that J contains a non-zero C ∗-finite and J := Q−1(I). Since J ) J A hereditary C ∗-subalgebra B. Since B ∩ J A C = (0), the ∗-homomorphism Q restricts to an injection on B. Thus, Q(B) ⊆ I is also a non-zero C ∗-finite hereditary C ∗- subalgebra, and A/J A C is C ∗-semi-finite (by Corollary 3.10(a)). The proof of the second statement is similar. (b) This follows from part (a) and Remark 5.3(c). (cid:3) Remark 5.6. Let S be a statement concerning C ∗-algebras that is stable under extensions of C ∗-algebras (i.e. if I is a closed ideal of a C ∗-algebra A such that S is true for both I and A/I, then S is true for A). (a) If S is true for all type A and all type B C ∗-algebras, S is true for all C ∗-semi- finite C ∗-algebras. If, in addition, S is true for all type C C ∗-algebras, it is true for all C ∗-algebras. (b) If S is true for all discrete C ∗-algebras and all anti-liminary C ∗-algebras, then S is true for all C ∗-algebras. The following results follows from Theorem 3.9(a). Corollary 5.7. If A and B are strongly Morita equivalent, then the closed ideal of B that corresponds to J A sf ) under the strong Morita equivalence (see the paragraph preceding Theorem 3.9) is precisely J B A (respectively, J B B, J B A (respectively, J A C and J A C and J B B, J A sf ). 26 CHI-KEUNG NG AND NGAI-CHING WONG Remark 5.8. It is natural to ask if the closure C(·) of C(·) (see Proposition 3.5) is also stable under strong Morita equivalence. Unfortunately, it is not the case. Suppose that A is any type I C ∗-algebra. Then by [6, Theorems 1.8 and 2.2], there is a commutative C ∗-algebra B that is strongly Morita equivalent to A. Notice that C(B) = B and C(A) is of type I0 (by [31, Proposition 6.1.7]). Thus, if C(·) is stable under strong Morita equivalence, then any type I C ∗-algebra A will coincide with C(A) and hence is liminary (see, e.g., [31, Corollary 6.1.6]), which is absurd. To end this section, we compare J A ∗ with J M (A) ∗ . C = J A B ∩ B, J B Proposition 5.9. (a) If B ⊆ A is a hereditary C ∗-subalgebra, then J B J B B = J A (b) J M (A) (c) J M (A) (d) J M (A) C ∩ B and J B = {x ∈ M(A) : xA ⊆ J A = {x ∈ M(A) : xJ A A = (0) and xA ⊆ J A sf } sf = (0)} = {x ∈ M(A) : xJ A sf = J A A }. Similar statements hold for JB, JC and Jsf. = {x ∈ M(A) : xJ A A = (0) and xJ A B = (0)}. A = J A sf ∩ B. A ∩ B, C A B A ⊆ B ∩ J A (a) Clearly, J B A . Conversely, since B ∩ J A Proof: ideal of B (by Corollary 3.15(a)), we have B ∩ J A from similar arguments. (b) We will only consider the case of JB (since the other cases follow from similar and easier arguments). Notice that J M (A) B (by part (a)) and A is a type A closed A . The other cases follow · A = J M (A) ∩ A = J A A ⊆ J B B B J M (A) B ⊆ J0 := {x ∈ M(A) : xA ⊆ J A B}. Suppose that the closed ideal J0 ⊆ M(A) contains a non-zero abelian hereditary C ∗-subalgebra B. The abelian hereditary C ∗-subalgebra B ∩ A = B · A · B is contained in J A B and so, B · A = (0), which contradicts the fact that A is essential in M(A) (see Remark 3.2(d)). Furthermore, let I be a non-zero closed ideal of J0. Then I · A = I ∩ A 6= (0) and is a closed ideal of J A B. Thus, I ∩ A contains a non-zero C ∗-finite hereditary C ∗-subalgebra. Consequently, J0 is of type B and is a subset of J M (A) . A = (0) if and only if xAJ A (c) Obviously, xJ A part (b) and Remark 5.3(c). (d) This part follows from a similar argument as part (c) as well as Remark 5.3(b). (cid:3) A = (0). Thus, this part follows from B References [1] C. A. Akemann, The General Stone-Weierstrass problem, J. Funct. Anal., 4 (1969), 277-294. [2] C. A. Akemann, Left ideal structure of C ∗-algebras, J. Funct. Anal., 6 (1970), 305-317. [3] C. A. Akemann and S. Eilers, Regularity of projections revisited, J. Oper. Theory, 48 (2002), 515-534. A MURRAY-VON NEUMANN TYPE CLASSIFICATION OF C ∗-ALGEBRAS 27 [4] C. A. Akemann and G. K. Pedersen, Complications of semicontinuity in C ∗-algebra theory, Duke Math. J., 40 (1973), 785-795. [5] C. A. Akemann, G. K. Pedersen and J. Tomiyama, Multipliers of C ∗-algebras, J. Funct. Anal., 13 (1973), 277-301. [6] W. Beer, On Morita equivalence of nuclear C ∗-algebras, J. Pure Appl. Alg., 26 (1982), 249-267. [7] M. Bekka, M. Cowling and P. de la Harpe, Some groups whose reduced C ∗-aglebra is simple, Publ. Math. I.H.E.S., 80 (1994), 117-134. [8] B. Blackadar, Operator algebras - theory of C ∗-algebras and von Neumann algebras, in Operator Algebras and Non-commutative Geometry III, Encyc. Math. Sci., 122, Springer- Verlag, Berlin (2006). [9] L. G. Brown, Semicontinuity and multipliers of C ∗-algebras, Canad. J. Math., XL, 4 (1988), 865-988. [10] L. G. Brown, Determination of A from M (A) and related matters, C. R. Math. Rep. Acad. Sci. Canada, 10 (1988), 273-278. [11] L. G. Brown and M. A. Rieffel, Stable isomorphism and strong Morita equivalence of C ∗- algebras, Pacific J. Math., 71 (1977), 349-363. [12] L. G. Brown and G. K. Pedersen, C ∗-algebras of real rank zero, J. Funct. Anal., 99 (1991), 131-149. [13] J. Cuntz, K-theory for certain C ∗-algebras, Ann. of Math., 113 (1981), 181-197. [14] J. Cuntz and G. K. Pedersen, Equivalence and traces on C ∗-algebras, J. Funct. Anal., 33 (1979), 135-164. [15] E. G. Effros, Order ideals in C ∗-algebras and its dual, Duke Math. J., 30 (1963), 391-412. [16] G. A. Elliott, On the classification of inductive limits of sequences of semisimple finite- dimensional algebras, J. Alg. 38 (1976), 29-44. [17] G. A. Elliott, On the classification of C ∗-algebras of real rank zero, J. Reine Angew. Math., 443 (1993), 179-219. [18] G. A. Elliott and A. S. Toms, Regularity properties in the classification program for separable amenable C ∗-algebras, Bull. Amer. Math. Soc., 45 (2008), 229 - 245. [19] H. Halpern, V. Kaftal, P. W. Ng and S. Zhang, Finite sums of projections in von Neumann algebras, preprint (arXiv:1007.4679v1). [20] R. V. Kadison and J. R. Ringrose, Fundamentals of the theory of operator algebras Vol. II: Advanced theory, Pure and Applied Mathematics 100, Academic Press (1986). [21] E. Kirchberg and M. Rørdam, Non-simple purely infinite C ∗-algebras, Amer. J. Math., 122 (2000), 637-666. [22] E. C. Lance, Hilbert C ∗-modules - A toolkit for operator algebraists, Lond. Math. Soc. Lect. Note Ser. 210, Camb. Univ. Press (1995). [23] H. Lin, Equivalent open projections and corresponding hereditary C ∗-subalgebras, J. Lond. Math. Soc., 41 (1990), 295-301. [24] H. Lin and S. Zhang, On infinite simple C ∗-algebras, J. Funct. Anal., 100 (1991), 221-231. [25] G.J. Murphy, C ∗-algebras and operator theory, Academic Press (1990). [26] F. J. Murray, The rings of operators papers, in The legacy of John von Neumann (Hemp- stead, NY, 1988), 57-60, Proc. Sympos. Pure Math. 50, Amer. Math. Soc., Providence, R.I. (1990). [27] F. J. Murray and J. Von Neumann, On rings of operators, Ann. of Math. (2), 37 (1936), 116-229. [28] C.K. Ng and N.C. Wong, Comparisons of equivalence relations on open projections, preprint. [29] E. Ortega, M. Rørdam and H. Thiel, The Cuntz semigroup and comparison of open projec- tions, J. Funct. Anal., 260 (2011), 3474-3493. [30] G. K. Pedersen, Applications of weak-*-semicontinuity in C ∗-algebra theory, Duke Math. J., 39 (1972), 431-450. 28 CHI-KEUNG NG AND NGAI-CHING WONG [31] G. K. Pedersen, C ∗-algebras and their automorphism groups, Academic Press (1979). [32] C. Peligrad and L. Zsid´o, Open projections of C ∗-algebras: Comparison and Regularity, Operator Theoretical Methods, 17th Int. Conf. on Operator Theory, Timisoara (Romania), June 23-26, 1998, Theta Found. Bucharest (2000), 285-300 [33] R. T. Prosser, On the ideal structure of operator algebras, Memoirs Amer. Math. Soc., 45 (1963). [34] M. A. Rieffel, Morita equivalence for C ∗-algebras and W ∗-algebras, J. Pure Appl. Alg., 5 (1974), 51-96. [35] M. A. Rieffel, Morita equivalence for operator algebras, in Operator algebras and applica- tions, Part I (Kingston, Ont., 1980), Proc. Sympos. Pure Math. 38, Amer. Math. Soc., Providence, R.I. (1982), 285 - 298. [36] M. Rørdam, Classification of nuclear, simple C ∗-algebras, in Classification of nuclear C ∗- algebras, Entropy in operator algebras, Encyclopaedia Math. Sci. 126, Springer, Berlin (2002), 1-145. [37] M. Takesaki, Theory of Operator algebras I, Springer-Verlag New York (1979). [38] A. S. Toms, On the classification problem for nuclear C ∗-algebras, Ann. of Math. (2), 167 (2008), 1029-1044. [39] S. Zhang, Stable isomorphism of hereditary C ∗-subalgebras and stable equivalence of open projections, Proc. Amer. Math. Soc., 105 (1989), 677-682. [40] S. Zhang, Certain C ∗-algebras with real rank zero and their corona and multiplier algebras part I, Pac. J. Math. 155 (1992), 169 - 197. (Chi-Keung Ng) Chern Institute of Mathematics and LPMC, Nankai University, Tianjin 300071, China. E-mail address: [email protected] (Ngai-Ching Wong) Department of Applied Mathematics, National Sun Yat-sen University, Kaohsiung, 80424, Taiwan. E-mail address: [email protected]
1101.0033
1
1101
2010-12-30T02:56:24
Quantum Symmetries and Strong Haagerup Inequalities
[ "math.OA" ]
In this paper, we consider families of operators $\{x_r\}_{r \in \Lambda}$ in a tracial C$^\ast$-probability space $(\mathcal A, \phi)$, whose joint $\ast$-distribution is invariant under free complexification and the action of the hyperoctahedral quantum groups $\{H_n^+\}_{n \in \N}$. We prove a strong form of Haagerup's inequality for the non-self-adjoint operator algebra $\mathcal B$ generated by $\{x_r\}_{r \in \Lambda}$, which generalizes the strong Haagerup inequalities for $\ast$-free R-diagonal families obtained by Kemp-Speicher \cite{KeSp}. As an application of our result, we show that $\mathcal B$ always has the metric approximation property (MAP). We also apply our techniques to study the reduced C$^\ast$-algebra of the free unitary quantum group $U_n^+$. We show that the non-self-adjoint subalgebra $\mathcal B_n$ generated by the matrix elements of the fundamental corepresentation of $U_n^+$ has the MAP. Additionally, we prove a strong Haagerup inequality for $\mathcal B_n$, which improves on the estimates given by Vergnioux's property RD \cite{Ve}.
math.OA
math
QUANTUM SYMMETRIES AND STRONG HAAGERUP INEQUALITIES MICHAEL BRANNAN Abstract. In this paper, we consider families of operators {xr}r∈Λ in a tracial C∗-probability space (A, ϕ), whose joint ∗-distribution is invariant under free complexification and the action of the hyperoctahedral quantum groups {H + n }n∈N. We prove a strong form of Haagerup's inequal- ity for the non-self-adjoint operator algebra B generated by {xr}r∈Λ, which generalizes the strong Haagerup inequalities for ∗-free R-diagonal families obtained by Kemp-Speicher [22]. As an appli- cation of our result, we show that B always has the metric approximation property (MAP). We also apply our techniques to study the reduced C∗-algebra of the free unitary quantum group U + n . We show that the non-self-adjoint subalgebra Bn generated by the matrix elements of the fundamental corepresentation of U + n has the MAP. Additionally, we prove a strong Haagerup inequality for Bn, which improves on the estimates given by Vergnioux's property RD [34]. 0 1 0 2 c e D 0 3 ] . A O h t a m [ 1 v 3 3 0 0 . 1 0 1 1 : v i X r a 1. Introduction Let Fn denote the free group on n ≤ ∞ generators g1, g2, . . . , gn, and let C∗λ(Fn) ⊆ B(ℓ2(Fn)) be the C∗-algebra generated by the left regular representation λ : Fn → U (ℓ2(Fn)). Denote by ℓ : Fn → N ∪ {0}, the natural reduced word length function associated to the generating set S = {gr, g−1 r=1 ⊂ Fn. In 1978, Haagerup published the following result, known as the Haagerup inequality: Theorem 1.1. [18, Lemma 1.4] Let d ∈ N and suppose f ∈ ℓ2(Fn) is supported on the set Wd = {g ∈ Fn : ℓ(g) = d}. Then r }n kfkℓ2(Fn) ≤ kλ(f )kC ∗ λ(Fn) ≤ (d + 1)kfkℓ2(Fn). The above inequality was used by Haagerup (together with the fact that the map λ(g) 7→ e−ℓ(g)tλ(g) defines a unital completely positive map on C∗λ(Fn), for all t ≥ 0) to show that C∗λ(Fn) has the metric approximation property [18], even though it is a non-nuclear C∗-algebra for n ≥ 2. Since the publication of this foundational result, the Haagerup inequality has continued to find numerous applications and generalizations in operator algebras, noncommutative harmonic analysis, and geometric group theory. See [10, 12, 15, 18, 19, 25, 26] for example. One of the key ingredients, implicit in Haagerup's original proof of Theorem 1.1, is the fact that r=1 of C∗λ(Fn) are algebraically free (and in fact ∗-freely independent in the the generators {λ(gr)}n sense of Voiculescu's free probability theory). Using this connection with free independence, Kemp and Speicher [22] showed, using combinatorial techniques from Voiculescu's free probability theory, that if one restricts to the non-self-adjoint operator algebra of convolution operators λ(f ) ∈ C∗λ(Fn) supported on the free semigroup F+ r=1, then the constants in the Haagerup inequality enjoy a substantial improvement: n generated by {gr}n Date: December 22, 2010. Key words and phrases. Free probability, Haagerup inequality, quantum groups, quantum symmetries, metric approximation property. 2010 Mathematics Subject Classification. Primary 46L54; Secondary 46L65. 1 Theorem 1.2. [22, Theorem 1.4] For any d ∈ N and any f ∈ ℓ2(Fn) supported on Wd ∩ F+ supported on words in Fn of length d in g1, . . . , gn but not their inverses), we have the estimate n (i.e. kfkℓ2(Fn) ≤ kλ(f )kC ∗ λ(Fn) ≤ √e√d + 1kfkℓ2(Fn). Furthermore, in [22] the authors were able to generalize the above strong Haagerup inequality to the much broader context of operator algebras generated by ∗-free, identically distributed, R- diagonal operators in a tracial C∗-probability space (A, ϕ). (Please consult Section 2 for the relevant definitions.) Their result can be stated as follows: Theorem 1.3. [22, Theorem 1.3] Let Λ be an index set, let {xr}r∈Λ be a ∗-free, identically dis- tributed family of R-diagonal operators in a tracial C∗-probability space (A, ϕ), and let x ∈ {xr}r∈Λ be some fixed reference variable. Then, for any d ∈ N and any homogeneous polynomial of degree d in the variables {xr}r∈Λ, we have (ai ∈ C), aixi(1)xi(2) . . . xi(d), T = Xi:{1,...,d}→Λ kTkL2(A,ϕ) ≤ kTkA ≤ Cx√dkTkL2(A,ϕ), L2(A,ϕ). Moreover, if x has non-negative free cumulants, then Cx ≤ A/kxk2 where Cx ≤ 515√ekxk2 √ekxkA/kxkL2(A,ϕ). The remarkable feature of Theorems 1.2 and 1.3 is the fact that the order of growth of the con- stants in these inequalities improves from O(d) to O(√d). This slower growth rate is a consequence of two things: (1) the fact that we are restricting to a non-self-adjoint subalgebra of our C∗-algebra A, and (2) the fact that the R-diagonal operators under consideration possess a great deal of ro- tational symmetry in their ∗-distributions, which can be exploited. The basic method of proof in [22] is to approximate the norm kTkA of a given homogeneous polynomial with the associated noncommutative Lp-norms kTkLp(A,ϕ), (p → ∞). For p ∈ 2N (an even integer), this amounts to the calculation of certain joint moments, which can then be expressed in terms of Speicher's free cumulants, and estimated efficiently using the freeness and R-diagonality assumptions. We remark here that de la Salle [16] has recently considered the framework of Theorem 1.3 in the category of operator spaces, and obtained strong Haagerup inequalities with operator coefficients. On a different note, several deep connections between free probability and certain classes of compact quantum groups have recently emerged, particularly in the study of quantum symmetries of families of random variables. Perhaps most illustrative of this connection is the free de Finetti theorem of Kostler-Speicher [23] (and its generalizations [6]), which says that an infinite sequence of random variables in a W∗-probability space is conditionally free if and only if its joint ∗-distribution is invariant under the action of the quantum permutation groups {S+ n }n∈N. Another example illus- trating this connection is the asymptotic freeness of the standard generators of many combinatorial quantum groups such as S+ n , H + n , O+ n , U + n [4, 6]. In this paper, we prove a generalization of the strong Haagerup inequality of Kemp-Speicher (The- orem 1.3). Continuing with the above theme of connecting free probability and compact quantum groups, we consider families of operators {xr}r∈Λ in a C∗-probability space (A, ϕ), which are not necessarily ∗-free, but instead possess certain quantum symmetries in their joint ∗-distributions. The setup for our result is as follows: for each n ∈ N, let H + n denote the hyperoctahedral quantum group of dimension n [3]. We say that a family of random variables {xr}r∈Λ in a C∗-probability space (A, ϕ) has an H +-invariant joint ∗-distribution if the joint ∗-distribution each sub-n-tuple {xr(l)}n of {xr}r∈Λ is invariant under the natural action of H + n . We say that the joint ∗-distribution of {xr}r∈Λ is invariant under free complexification if {zxr}r∈Λ has the same joint ∗-distribution as l=1 2 {xr}r∈Λ, for any Haar unitary z ∈ (A, ϕ) which is ∗-free from {xr}r∈Λ. Our main theorem is the following: Theorem 1.4. Let (A, ϕ) be a tracial C∗-probability space, and let {xr}r∈Λ ⊂ (A, ϕ) be a family of random variables. Suppose that the joint ∗-distribution of {xr}r∈Λ is H +-invariant and in- variant under free complexification. Let x ∈ {xr}r∈Λ be a fixed reference variable. Then for any homogeneous polynomial T = Xi:[d]→Λ aixi(1)xi(2) . . . xi(d), (ai ∈ C), of degree d in the variables {xr}r∈Λ and any p ∈ 2N ∪ {∞}, we have kTkL2(A,ϕ) ≤ kTkLp(A,ϕ) ≤ 45 · (3e)2√ekxk2 kxk2 Lp(A,ϕ) L2(A,ϕ) √d + 1kTkL2(A,ϕ). Using Theorem 1.4, we study the structure of the non-self-adjoint operator algebra B generated by the family {xr}r∈Λ. In particular, we show that B always has the metric approximation property. We also indicate how any ∗-free, identically distributed R-diagonal family {xr}r∈Λ ⊂ (A, ϕ) satisfies the hypotheses of Theorem 1.4. On the other hand, we show that there are many natural families of random variables (coming from certain free complexified quantum groups) satisfying the hypotheses of Theorem 1.4, which are not ∗-free. Using a modification of Theorem 1.4 for bi-invariant arrays of random variables, we obtain a strong Haagerup inequality for Wang's free unitary quantum group U + n ([35]), improving on the Haagerup inequalities for U + n obtained by Vergnioux [34]. Viewing U + n as the non-cocommutative analogue of the compact quantum group associated to C∗λ(Fn), our result is the non-cocommutative analogue of the Kemp-Speicher strong Haagerup inequality (Theorem 1.2) for Fn. The remainder of the paper is organized as follows: Section 2 contains all of the notation and basic facts we will need from free probability and compact quantum groups. In Section 3, we restate and prove Theorem 1.4. In Section 4, we generalize Theorem 1.4 to arrays {xrs}1≤r,s≤n of random variables, whose joint ∗-distribution is H + n -bi-invariant (Theorem 4.1). In Section 5, we study the unital (norm closed) non-self-adjoint operator algebras B ⊂ (A, ϕ) generated by families {xr}r∈Λ ⊂ (A, ϕ) satisfying the hypotheses of Theorem 1.4 or 4.1. We show that the natural complex Ornstein- Uhlenbeck type semigroup {Γt}t>0 acting on B by Γt(xi(1)xi(2) . . . xi(d)) = e−dtxi(1)xi(2) . . . xi(d), is completely contractive for all t ≥ 0. Using this fact, together with the norm estimates provided by our Haagerup inequalities, we prove that B has the metric approximation property. In Section 6, we consider the free unitary quantum groups U + n (n ∈ N), and show that the strong Haagerup inequalities and metric approximation properties obtained in Sections 3 - 5 apply to the non-self- adjoint operator algebra Bn ⊂ L∞(U + n ), generated by the matrix elements of the fundamental corepresentation of U + n . This yields Theorem 6.3, which is our non-cocommutative analogue of Theorem 1.2. We also discuss a general method for obtaining more non-trivial examples of families which satisfy the hypotheses of our theorems. Acknowledgements. It is a pleasure to thank my doctoral supervisors James A. Mingo and Roland Speicher for many fruitful discussions and for their continued guidance while working on this project. This research was partially supported by an NSERC Canada Graduate Scholarship. 2. Preliminaries and Notation We begin by briefly reviewing the relevant facts from free probability and C∗-algebraic compact matrix quantum groups that will be needed in this paper. Our main reference for free probability will be the monograph [28]. For the basics of compact quantum groups we refer to the textbook [32] and the foundational paper of Woronowicz [36]. 3 2.1. Noncommutative Probability Spaces and Free Independence. Definition 2.1. (1) A noncommutative probability space (NCPS) is a pair (A, ϕ), where A is a unital ∗-algebra over C, and ϕ : A → C is a state (i.e. a linear functional such that ϕ(1A) = 1 and ϕ(aa∗) ≥ 0 for all a ∈ A). We say that (A, ϕ) is tracial if ϕ is a trace on A. (2) A C∗-probability space is a NCPS (A, ϕ) where A is a unital C∗-algebra, and ϕ is a faithful state (i.e. ϕ(aa∗) = 0 ⇔ a = 0). (3) If (A, ϕ) is a NCPS, we call elements of A random variables. A random variable x ∈ (A, ϕ) is called centered if x ∈ ker ϕ. (4) Let X = {xr}r∈Λ be a family of random variables in a NCPS (A, ϕ), and let Chtr, t∗r : r ∈ Λi denote the ∗-algebra of noncommutative polynomials in the indeterminates {tr}r∈Λ. Denote by evX : Chtr, t∗r : r ∈ Λi → A, evX (tr) = xr the canonical evaluation ∗-homomorphism determined by the family X. The joint ∗- distribution of the family X = {xr}r∈Λ is the linear functional ϕX : Chtr, t∗r : r ∈ Λi → C given by ϕX (p) = ϕ(evX (p)), (p ∈ Chtr, t∗r : r ∈ Λi). (5) Let (Aj, ϕj) (j = 1, 2) be two NCPS's, and consider two families of random variables X = {xr}r∈Λ ⊂ (A1, ϕ1) and Y = {yr}r∈Λ ⊂ (A2, ϕ2). We say that X and Y are identically distributed if they have the same joint ∗-distributions: ϕX = ϕY . (6) If (A, ϕ) is a tracial C∗-probability space, and p ∈ [1,∞) we denote by Lp(A, ϕ) the asso- ciated noncommutative Lp-space, which is the completion of A with respect to the norm kxkLp(A,ϕ) = ϕ((xx∗)p/2)1/p. For p = ∞, we identify L∞(A, ϕ) with A, and note that since ϕ is faithful. kxkL∞(A,ϕ) = lim p→∞kxkLp(A,ϕ), We now recall the definition of free independence for noncommutative probability spaces. Definition 2.2. (1) Let (A, ϕ) be a NCPS and let {Ar}r∈Λ be a family of unital subalgebras of A. The family {Ar}r∈Λ is said to be freely independent with respect to ϕ (or just free for any choice of indices if the state ϕ is understood), if the following condition holds: r(1) 6= r(2), r(2) 6= r(3), . . . , r(k − 1) 6= r(k) ∈ Λ and any choice of centered random variables variables xr(j) ∈ Ar(j) ∩ ker ϕ, we have the equality ϕ(xr(1)xr(2) . . . xr(k)) = 0. If each Ar is a ∗-subalgebra, the we say that {Ar}r∈Λ are ∗-free. unital subalgebras (2) A family of noncommutative random variables {xr}r∈Λ ⊆ (A, ϕ) is free if the family of {Ar}r∈Λ, Ar = Algh1A, xri, are free in the sense of (1). We say that {xr}r∈Λ are ∗-free if we replace the subalgebras Ar = Algh1A, xri above by the ∗-subalgebras Algh1A, xr, x∗ri. 2.2. Non-Crossing Partitions and Free Cumulants. An important tool for studying the no- tion of free independence and for our calculations will be the free cumulant functionals, {κn : An → C}n∈N, associated to any NCPS (A, ϕ). These were first introduced by Speicher in [31]. Throughout this section and the rest of the paper, we will frequently be dealing with partitions of the ordered sets {1, 2, . . . , k} and multi-indices i = (i(1), . . . , i(k)) ∈ Λk where Λ is any index set and 4 k ∈ N. For ease of notation, we will denote the ordered set {1, . . . , k} by [k], and interchangeably view multi-indices as functions i = (i(1), . . . , i(k)) ∈ Λk, i : [k] → Λ, j 7→ i(j). Also, given r functions i1, . . . , ir : [k] → Λ, we will often regard the r-tuple I = (i1, . . . , ir) as the function I : [rk] → Λ in the obvious way. Definition 2.3. (1) A partition π of the set [k] is a collection of disjoint, non-empty subsets V1, . . . , Vr ⊆ [k] such that V1 ∪ . . . ∪ Vr = [k]. V1, . . . , Vr are called the blocks of π, and we set π = r - the number of blocks of π. If s, t ∈ [k] are in the same block of π, we write s ∼π t. The collection of partitions of [k] is denoted by P(k). (2) Given π, σ ∈ P(k), we say that π ≤ σ if each block of π is contained in a block of σ. There is a least element of P(k) which is larger than both π and σ, which we denote by π ∨ σ. There is also a maximal element π ∧ σ ∈ P(k) which is smaller than both π and σ. With these operations, P(k) is a finite lattice. (3) Given a function i : [k] → Λ, we denote by ker i the element of P(k) whose blocks are the equivalence classes of the relation s ∼ker i t ⇐⇒ i(s) = i(t). Note that if π ∈ P(k), then π ≤ ker i is equivalent to the condition that whenever s and t are in the same block of π, i(s) must equal i(t) (i.e the function i : [k] → Λ is constant on the blocks of π). (4) We say that π ∈ P(k) is non-crossing if whenever V, W are blocks of π and s1 < t1 < s2 < t2 are such that s1, s2 ∈ V and t1, t2 ∈ W , then V = W . We can also define non-crossing partitions recursively: a partition π ∈ P(k) is non-crossing if and only if it has a block V which is an interval, such that π\V is a non-crossing partition of [k]\V ∼= [k −V ]. The set of non-crossing partitions of [k] is denoted by N C(k). (5) Given π, σ ∈ N C(k), we define π ∨N C σ ∈ N C(k) to be the least element of N C(k) which is larger than both π and σ. With the operation ≤ and ∧ induced by the inclusion N C(k) ⊂ P(k), and the operation ∨N C, N C(k) becomes a finite lattice itself. (6) A partition π ∈ P(k) is called a pairing if every block of π contains exactly 2 elements. The set of all pairings of [k] is denoted by P2(k). We call a partition π ∈ N C(k) ∩ P2(k) a non-crossing pairing, and denote this set of pairings by N C2(k). A partition π ∈ P(k) is called even if every block of π has even cardinality. We denote the set of even partitions (resp. even non-crossing partitions) of [k] by Pe(k) (resp. N Ce(k)). Note that k must be even for Pe(k) to be non-empty. (7) Given a function ǫ : [k] → {1,∗}, we define the set P ǫ(k) of ǫ-partitions of [k] to be the set of all even partitions π ∈ Pe(k) with the property that on each block V = {j1 < j2 < . . . < j2r−1 < j2r}, of π, ǫV is an alternating function. I.e. ǫV = (1,∗, . . . , 1,∗) , or ǫV = (∗, 1, . . . ,∗, 1) , {z for all blocks V of π. We also put N C ǫ(k) := N C(k) ∩ P ǫ(k). {z 2r times 2r times } } Since N C(k) is a lattice, it has a Mobius function µk : N C(k) × N C(k) → Z, which is is well- known. The first major study of the lattice structure of N C(k) was done by Kreweras in 1972. 5 We refer to Kreweras' original paper [24] and [28, Chapters 9-10] for more details on this lattice structure. For our purposes, we will only need the fact that µk(0k, 1k) = (−1)k−1Ck−1, and that for any σ ∈ N C(k) µk(σ, 1k) ≤ µk(0k, 1k) ≤ Ck−1. where 0k, 1k are respectively the smallest and biggest elements of N C(k), and Ck denotes the kth Catalan number Ck = N C(k) = 1 k(cid:18) 2k k − 1 (cid:19) ≤ 4k. Let A be a C-vector space equipped with a sequence of multilinear functionals {ψn : An → C}n∈N. For each k ∈ N and π ∈ N C(k), there is a canonical way to define a new k-linear functional ψπ : Ak → C from the family {ψn}n as follows: if {Vj}r j=1 denote the blocks of π, where Vj = {ij(1) < ij(2) < . . . < ij(Vj)}, then ψπ is defined by the equation ψπ[x1, . . . , xk] = ψVj[xij (1), . . . , xij (Vj)], ((x1, . . . xk) ∈ Ak). rYj=1 We are now ready to define the free cumulants associated to a NCPS (A, ϕ). Definition 2.4. (1) Let (A, ϕ) be a NCPS. The free cumulant functionals of (A, ϕ) are the family of multilinear functionals {κn : An → C}n∈N defined recursively by the equations (2.1) ϕ(x1x2 . . . xk) = Xπ∈N C(k) κπ[x1, x2, . . . , xk], ((x1, . . . , xk) ∈ Ak). (2) Given a random variable x ∈ (A, ϕ), the free cumulants of x are simply the collection of all quantities κk[xǫ(1), . . . , xǫ(k)], (ǫ : [k] → {1,∗}, k ∈ N). Alternatively, by considering the family of moment functionals {ϕk : Ak → C}k∈N given by ϕk[x1, x2, . . . , xk] = ϕ(x1x2 . . . xk), and applying Mobius inversion to (2.1), we have ((x1, . . . , xn) ∈ Ak), (2.2) (2.3) More generally, we have κk[x1, . . . , xk] = Xσ∈N C(k) κπ[x1, . . . , xk] = Xσ∈N C(k) σ≤π µk(σ, 1k)ϕσ[x1, . . . , xk], ((x1, . . . , xk) ∈ Ak). µk(σ, π)ϕσ[x1, . . . , xk], (π ∈ N C(k)). The following result, originally proved by Speicher [31] (see also [28, Theorem 11.16]), charac- terizes freeness in terms of the vanishing of mixed free cumulants. Theorem 2.5. Let (A, ϕ) be a NCPS with free cumulant functionals {κn}n∈N A family of unital subalgebras {Ar}r∈Λ is free if and only if for any k ≥ 2 and any function i : [k] → Λ, we have κkAi(1)×Ai(2)×...×Ai(k) = 0, (unless i(1) = i(2) = . . . = i(k)). 6 2.3. R-Diagonal Elements and Free Complexification. R-diagonal elements are defined in terms of the structure of their free cumulants. Definition 2.6. A random variable x ∈ (A, ϕ) is called an R-diagonal element if the only non-zero free cumulants of x are the alternating even ones. I.e. κ2k+1[xǫ(1), . . . , xǫ(2k+1)] = 0, (∀k ∈ N, ǫ : [2k + 1] → {1,∗}), and only if ǫ = (1,∗, . . . , 1,∗), or ǫ = (∗, 1, . . . ,∗, 1). κ2k[xǫ(1), . . . , xǫ(2k)] 6= 0, The standard examples of R-diagonal elements are Haar unitary random variables and circular random variables. A random variable z ∈ (A, ϕ) is Haar unitary if z is unitary, and ϕ(zn) = δn,0 for all n ∈ Z. From these equations it follows that the free cumulants of a Haar unitary z satisfy κ2k[z, z∗, . . . , z, z∗] = κ2k[z∗, z, . . . , z, z∗] = (−1)k−1Ck−1, and are identically zero otherwise ([28, Proposition 15.1]). A random variable c ∈ (A, ϕ) is (stan- dard) circular if and only if its free cumulants satisfy κ2[c, c∗] = κ2[c∗, c] = 1, and are zero otherwise. The standard unitary generators λ(g1), . . . , λ(gk) of the reduced group C∗-algebra C∗λ(Fk) are Haar unitary, and ∗-free with respect to the canonical trace-state τ : C∗λ(Fk) → C. If s1, s2 ∈ (A, ϕ) are two free standard semicircular random variables (i.e. s1 and s2 are free, identically distributed, and self-adjoint with spectral measures dµsi(x) = 1 2π √4 − x2χ[−2,2]dx), then s1 + is2√2 , c = is a circular random variable. In Voiculescu's free probability theory, the semicircular variable is the "free analogue" of a real Gaussian random variable. The circular variable c therefore can be thought of as the free analogue of a complex Gaussian random variable. We now introduce the important notion of free complexification, which can be thought of as the free analogue of multiplying a classical complex random vector by a randomly chosen phase factor eiθ ∈ T. Definition 2.7. Let {xr}r∈Λ be a family random variables in a NCPS (A, ϕ), and let z ∈ (A, ϕ) be a Haar unitary which is ∗-free from {xr}r∈Λ. We call the family {wr}r∈Λ, where wr = zxr, the free complexification of {xr}r∈Λ. We say that the joint ∗-distribution of {xr}r∈Λ is invariant under free complexification if {xr}r∈Λ and {wr}r∈Λ are identically distributed. Remark 2.8. If z′ ∈ (A, ϕ) is another Haar unitary which is ∗-free from {xr}r∈Λ and w′r = z′xi, then {wr}r∈Λ and {w′r}r∈Λ are identically distributed families. Therefore, the notion of free complexifi- cation as an operation on joint ∗-distributions, is well defined. We also record here the elementary fact if {wr}r∈Λ denotes the free complexification of {xr}r∈Λ, then the joint ∗-distribution of {wr}r∈Λ is invariant under free complexification. Free complexification and R-diagonality are intimately related through the following theorem. Theorem 2.9. (i) Let z, x ∈ (A, ϕ) be random variables, and suppose that z is Haar unitary and ∗-free from x. Then x is R-diagonal if and only if x and zx are identically distributed. In particular if x is any random variable and z is Haar unitary and ∗-free from x, then zx is R-diagonal. (ii) More generally, suppose that {xr}r∈Λ ⊂ (A, ϕ) is a ∗-free family of even elements (i.e. all odd ∗-moments of each xr are zero), and z ∈ (A, ϕ) is a Haar unitary which is ∗-free from {xr}r∈Λ, then the free complexification {zxr}r∈Λ is a ∗-free family of R-diagonal elements. 7 Proof. Statement (i) is [28, Corollary 15.9 and Theorem 15.10]. Statement (ii) is a special case of [29, Theorem 1.13]. (cid:3) 2.4. Quantum Groups and Invariant Distributions. Definition 2.10. A compact matrix quantum group (CMQG) is a pair G = (A, U ), where A is a unital C∗-algebra, and U = [urs]1≤r,s≤n ∈ Mn(A) is a unitary element satisfying the following three conditions: (1) A is generated as a C∗-algebra by the set {urs : 1 ≤ r, s ≤ n}; (2) The matrix U := [u∗rs]1≤r,s≤n is invertible in Mn(A); (3) There exists a ∗-homomorphism ∆ : A → A ⊗ A such that ∆urs = urk ⊗ uks, (1 ≤ r, s ≤ n). nXk=1 kXl=1 By definition, the ∗-homomorphism ∆ (called the coproduct of A) satisfies the coassociativity law A k-dimensional unitary corepresentation of G = (A, U ) is a unitary matrix V = [vrs]1≤r,s≤k ∈ Mk(A) such that (∆ ⊗ id) ◦ ∆ = (id ⊗ ∆) ◦ ∆. ∆vrs = vrl ⊗ vls, (1 ≤ r, s ≤ k). By definition, the matrix U defining G is a corepresentation, called the fundamental corepresentation of G. The unit 1A is also a corepresentation of G, called the trivial corepresentation. Recall that for any CMQG G = (A, U ), there exists a unique state h : A → C, called the Haar state, which is bi-invariant with respect to the coproduct ∆. I.e. (2.4) (h ⊗ id) ◦ ∆a = (id ⊗ h) ◦ ∆a = h(a)1, (a ∈ A). The existence and uniqueness of the Haar state was shown by Woronowicz in [36]. When h is a trace (as it will always be for us) and p ∈ [1,∞], we denote by Lp(G) := Lp(A, h) the noncommutative Lp-space associated to the GNS representation of the Haar state h. Definition 2.11. Let G = (A, U ) be a CMQG with fundamental corepresentation U = [urs]1≤r,s≤n ∈ Mn(A), and let (A, ϕ) be a NCPS. (i). We say that an n-tuple of random variables X = {xr}n r=1 ⊂ (A, ϕ) has a (left) G-invariant joint ∗-distribution if the family Y = {yr}n r=1 ⊂ A ⊗ A defined by yr = urk ⊗ xk, (1 ≤ r ≤ n), nXk=1 has the same joint ∗-distribution (with respect to id ⊗ ϕ) as X. That is, for any ∗-polynomial p ∈ Chtr, t∗r : 1 ≤ r ≤ ni, (id ⊗ ϕ)(evY (p)) = ϕX (p)1A. (ii). We say that an n × n array of random variables X = {xrs}1≤r,s≤n ⊂ (A, ϕ) has a left-G- invariant joint ∗-distribution if the family Y = {yrs}1≤r,s≤n ⊂ A ⊗ A defined by yrs = nXk=1 urk ⊗ xks, (1 ≤ r, s ≤ n), 8 has the same joint ∗-distribution (with respect to id⊗ϕ) as X. Similarly, X has a right-G-invariant joint ∗-distribution if the family Z = {zrs}1≤r,s≤n ⊂ A ⊗ A defined by zrs = xrk ⊗ uks, (1 ≤ r, s ≤ n), nXk=1 has the same joint ∗-distribution (with respect to ϕ⊗ id) as X. We say that X has a G-bi-invariant joint ∗-distribution if the ∗-distribution is both left and right G-invariant. I.e., for any ∗-polynomial p ∈ Chtrs, t∗rs : 1 ≤ r, s ≤ ni, (id ⊗ ϕ)(evY (p)) = (ϕ ⊗ id)(evZ (p)) = ϕX (p)1A. Remark 2.12. In other words, an n-tuple X = {xi}n i=1 ⊂ (A, ϕ) has a G-invariant joint ∗- distribution if and only if for all k ∈ N, i : [k] → [n], and ǫ : [k] → {1,∗}, we have the algebraic identity Similarly, X = {xrs}1≤r,s≤n has a G-bi-invariant joint ∗-distribution if and only if for all k ∈ N, i, j : [k] → [n], and ǫ : [k] → {1,∗}, we have the algebraic identities i(1)j(1) . . . uǫ(k) uǫ(1) j(1) . . . xǫ(k) i(k)j(k)ϕ(cid:0)xǫ(1) j(k)(cid:1). i(1) . . . xǫ(k) ϕ(cid:0)xǫ(1) i(k)(cid:1)1A = Xj:[k]→[n] ϕ(cid:0)xǫ(1) = Xl:[k]→[n] = Xl:[k]→[n] i(1)j(1) . . . xǫ(k) i(k)j(k)(cid:1)1A i(1)l(1) . . . uǫ(k) uǫ(1) l(1)j(1) . . . xǫ(k) i(k)l(k)ϕ(cid:0)xǫ(1) l(k)j(k)ϕ(cid:0)xǫ(1) l(k)j(k)(cid:1) i(k)l(k)(cid:1). l(1)j(1) . . . uǫ(k) uǫ(1) i(1)l(1) . . . xǫ(k) Note also that any column of a n × n left G-invariant array of random variables is a G-invariant n-tuple. Remark 2.13. If G = (A, U ) and H = (B, V ) are two CMQG's with U ∈ Mn(A) and V ∈ Mn(B), we say that H is a quantum subgroup of G if there exists a surjective ∗-homomorphism π : A → B, such that (id ⊗ π)(U ) = V . We note here the elementary fact that a family of random variables X = {xi}n i=1 ⊂ (A, ϕ) with a G-invariant joint ∗-distribution also has an H-invariant joint ∗- distribution, for any quantum subgroup H of G. The same is statement is true for arrays X = {xrs}1≤r,s≤n ⊂ (A, ϕ), which have G-bi-invariant joint ∗-distributions. The proof of these facts is an elementary application of the above definitions, and is left to the reader. 2.5. The Hyperoctahedral Quantum Group. Consider the hypercube In = [−1, 1]n in Rn. The symmetry group of In is called the hyperoctahedral group, and is denoted by Hn. We now quantize the definition of Hn. Definition 2.14. Let A be a unital C∗-algebra, and let U = [urs]1≤r,s≤n ∈ Mn(A) be an orthogonal matrix (i.e. U is unitary, and urs = u∗rs for all r, s.) We call U a cubic unitary if, in addition to orthogonality, we have That is, on each row or column of U , distinct entries a 6= b satisfy ab = ba = 0. uijuik = ujiuki = 0, (1 ≤ i ≤ n, 1 ≤ j 6= k ≤ n). It is easily shown that C(Hn), the commutative algebra of complex functions on Hn, is isomorphic as a C∗-algebra to the universal commutative C∗-algebra A, generated by n2 generators {urs}1≤r,s≤n which satisfy the relations which make U = [urs] ∈ Mn(A) a cubic unitary. By removing the commutativity in the above statement, we arrive at the definition of the hyperoctahedral quantum group [3]. 9 vertices), the quantum group H + Definition 2.15. [3] Let AH(n) denote the universal C∗-algebra generated by n2 generators {urs}1≤r,s≤n subject to the relations which make U = [urs] ∈ Mn(AH (n)) a cubic unitary. The pair H + n = (AH (n), U ) is a CMQG, and is called the hyperoctahedral quantum group (of dimension n). By regarding the hypercube In ⊂ Rn as the graph formed by n segments (n edges and 2n n = (AH (n), U ) is indeed the quantum symmetry group of In [3]. There is a useful combinatorial formula describing the Haar state h : AH(n) → C, which we now briefly describe: consider the family of matrices {Gn,2k}k∈N, where Gn,2k is the N Ce(2k) × N Ce(2k) matrix indexed by partitions π, σ ∈ N Ce(2k), with entries (π, σ ∈ N Ce(2k)). Gn,2k(π, σ) = nπ∨σ, Theorem 2.16. ([3, 6]) For every k, the matrix Gn,2k is invertible. Let Wn,2k its matrix inverse. Then for any pair i, j : [2k] → [n], we have h(cid:0)ui(1)j(1)ui(2)j(2) . . . ui(2k)j(2k)(cid:1) = Xπ,σ∈N Ce(2k) ker i≥π, ker j≥σ Wn,2k(π, σ), and h(cid:0)ui(1)j(1)ui(2)j(2) . . . ui(2k−1)j(2k−1)(cid:1) = 0. If X ⊆ (A, ϕ) is a family (array) with an H + n (-bi)-invariant joint ∗-distribution, Theorem 2.16 allows us to get a combinatorial description of which joint ∗-moments of the family X must always vanish. Proposition 2.17. (i). Let X = {xr}n joint ∗-distribution is invariant under the hyperoctahedral quantum group H + ǫ : [k] → {1,∗}, and i : [k] → [n], r=1 be a family of random variables in a NCPS (A, ϕ), whose n . Then for any k ∈ N, only if there exists some π ∈ N Ce(k) such that ker i ≥ π. (ii). Let X = {xrs}1≤r,s≤n ⊆ (A, ϕ) be an array whose joint ∗-distribution is H + Then for any k ∈ N, ǫ : [k] → {1,∗}, and any pair i, j : [k] → [n], we have n -bi-invariant. ϕ(cid:0)xǫ(1) i(1) xǫ(2) i(2) . . . xǫ(k) i(k)(cid:1) 6= 0 ϕ(cid:0)xǫ(1) i(1)j(1)xǫ(2) i(2)j(2) . . . xǫ(k) i(k)j(k)(cid:1) 6= 0 only if there exist π, σ ∈ N Ce(k) such that ker i ≥ π and ker j ≥ σ. (iii). In both (i) and (ii), the family X is identically distributed, and orthogonal in L2(A, ϕ). Proof. We use the formulae given in Remark 2.12. (i). Since X = {xr}n ϕ(cid:0)xǫ(1) ui(1)j(1) . . . ui(k)j(k)ϕ(cid:0)xǫ(1) i(k)(cid:1)1AH (n) = Xj:[k]→[n] n -invariant, we have j(k)(cid:1). j(1) . . . xǫ(k) i(1) . . . xǫ(k) r=1 is H + Applying the Haar state to both sides of the equation and using Theorem 2.16, we get ϕ(cid:0)xǫ(1) i(1) . . . xǫ(k) i(k)(cid:1) = 0, 10 if k is odd, and i(1) . . . xǫ(k) ϕ(cid:0)xǫ(1) i(k)(cid:1) = Xj:[k]→[n] Xπ,σ∈N Ce(k) = Xπ,σ∈N Ce(k) ker i≥π, ker j≥σ Wn,k(π, σ) Xj:[k]→[n] Wn,k(π, σ)ϕ(cid:0)xǫ(1) ϕ(cid:0)xǫ(1) ker j≥σ ker i≥π j(1) . . . xǫ(k) j(k)(cid:1) j(k)(cid:1), j(1) . . . xǫ(k) n -bi-invariant, we have the two equalities i(1)j(1) . . . xǫ(k) Applying the Haar state to these equalities and using Theorem 2.16 gives, for k odd, if k is even. Clearly (i) follows from these equalities. (ii). Since X = {xrs}1≤r,s≤n is H + ϕ(cid:0)xǫ(1) = Xl:[k]→[n] = Xl:[k]→[n] i(1)j(1) . . . xǫ(k) i(k)j(k)(cid:1)1AH (n) ui(1)l(1) . . . ui(k)l(k)ϕ(cid:0)xǫ(1) ul(1)j(1) . . . ul(k)j(k)ϕ(cid:0)xǫ(1) ϕ(cid:0)xǫ(1) i(k)j(k)(cid:1) = 0, Wn,k(π, σ) Xl:[k]→[n] ϕ(cid:0)xǫ(1) Wn,k(π, σ) Xl:[k]→[n] ϕ(cid:0)xǫ(1) i(k)j(k)(cid:1) ker l≥σ ker l≥π i(1)j(1) . . . xǫ(k) ϕ(cid:0)xǫ(1) = Xπ,σ∈N Ce(k) = Xπ,σ∈N Ce(k) ker i≥π ker j≥σ and l(1)j(1) . . . xǫ(k) i(1)l(1) . . . xǫ(k) l(k)j(k)(cid:1) i(k)l(k)(cid:1). l(1)j(1) . . . xǫ(k) i(1)l(1) . . . xǫ(k) l(k)j(k)(cid:1) i(k)l(k)(cid:1), for k even, which clearly implies (ii). (iii). The fact that the families X in (i) and (ii) are both orthogonal is just a special case, when k = 2, of (i) and (ii). Indeed, when k = 2 we have Gn,k = n, so the equations in the proof of (i) reduce to ϕ(xi(1)x∗i(2)) = δi(1),i(2) n and the equations in the proof of (ii) reduce to nXl=1 ϕ(xlx∗l ), ϕ(xi(1)j(1)x∗i(2)j(2)) = δi(1),i(2) n ϕ(xlj(1)x∗lj(2)) = δj(1),j(2) n ϕ(xi(1)lx∗i(2)l). nXl=1 nXl=1 To show that these families are identically distributed, note that the classical permutation group Sn is a quantum subgroup of H + n . Therefore the family X (in either (i) or (ii)) has a Sn-(bi-)invariant joint ∗-distribution, and this implies in particular that the variables in X are identically distributed (see [23, Section 2] for instance). Remark 2.18. We say that a family X = {xr}r∈Λ ⊂ (A, ϕ) (Λ ≤ ∞), has an H +-invariant joint ∗-distribution if every finite subsequence {xr(l)}n n -invariant joint ∗-distribution in the sense of Definition 2.11. l=1 of X has an H + (cid:3) 11 In this section we prove Theorem 1.4, which we restate here for convenience. 3. Strong Haagerup Inequalities Theorem 3.1. Let (A, ϕ) be a tracial C∗-probability space, and let {xr}r∈Λ ⊂ (A, ϕ) be a family of random variables. Suppose that the joint ∗-distribution of {xr}r∈Λ is H +-invariant and in- variant under free complexification. Let x ∈ {xr}r∈Λ be a fixed reference variable. Then for any homogeneous polynomial T = Xi:[d]→Λ aixi(1)xi(2) . . . xi(d), (ai ∈ C), of degree d in the variables {xr}r∈Λ and any p ∈ 2N ∪ {∞}, we have kTkL2(A,ϕ) ≤ kTkLp(A,ϕ) ≤ 45 · (3e)2√ekxk2 kxk2 Lp(A,ϕ) L2(A,ϕ) √d + 1kTkL2(A,ϕ). Remark 3.2. As mentioned in the introduction, Theorem 3.1 applies in particular to ∗-free, identi- cally distributed R-diagonal families {xr}r∈Λ ⊆ (A, ϕ). To see this, note that such a family {xr}r∈Λ is ∗-free, identically distributed, and each xr is even (from the definition of R-diagonality). There- fore it follows from [6, Proposition 4.3] that {xr}r∈Λ has an H +-invariant joint ∗-distribution. On the other hand, if z is a Haar unitary which is ∗-free from {xr}r∈Λ, then by Theorem 2.9, {zxr}r∈Λ and {xr}r∈Λ are identically distributed. As a consequence, we recover as a special case the strong Haagerup inequality of Kemp-Speicher ([22, Theorem 1.3]), although with weaker universal con- stants than theirs. This is to be expected, as our assumptions on the joint ∗-distribution of our operators are weaker. In Section 6, we will consider some non-trivial examples coming from compact quantum groups, where the greater generality of Theorem 3.1 will be essential. 3.1. Proof of Theorem 3.1. Let {xr}r∈Λ satisfy the hypotheses of Theorem 3.1. Given a function i : [d] → Λ, we will use the notation (3.1) to denote monomials of degree d in the variables {xr}r∈Λ. Fix d ∈ N, and let Xi := xi(1)xi(2) . . . xi(d) T = Xi:[d]→Λ aiXi, be homogeneous polynomial of degree d as in Theorem 3.1. Our strategy is similar to the ones used in [22] and [16] to prove strong Haagerup inequalities: we will first express the norms kTk2m := kTkL2m(A,ϕ) (m ∈ N) in terms of the joint free cumulants of the family {xr}r∈Λ, then use the assumed properties of the distribution to obtain meaningful estimates. The inequality for kTkA = kTkL∞(A,ϕ) will follow from our estimates by letting m → ∞. Explicitly we have 2m = ϕ((T T ∗)m) aiXi(cid:17)(cid:16) Xi:[d]→Λ aiXi(cid:17)∗(cid:17)m(cid:17) kTk2m = ϕ(cid:16)(cid:16)(cid:16) Xi:[d]→Λ = Xi1,...,i2m:[d]→Λ ai1ai2 . . . ai2m−1ai2m ϕ(Xi1 (Xi2)∗ . . . Xi2m−1 (Xi2m )∗). Now, for any function i : [d] → Λ, define the function i : [d] → Λ by reversing the order of the d-tuple i. That is, (i(1), . . . ,i(d)) = (i(d), . . . , i(1)). 12 Then, by a simple change of indices in the above sum, we can write 2m kTk2m = Xi1,...,i2m:[d]→Λ where ai1ai2 . . . ai2m−1ai2m ϕ(Xi1 (X∗)i2 . . . Xi2m−1 (X∗)i2m ), (X∗)ik := x∗ik(1)x∗ik(2) . . . x∗ik(d). We will now write each moment ϕ(Xi1 (∗)i2 . . . Xi2m−1 (X∗)i2m ) in terms of free cumulants, using equation (2.1). This gives, for each function I = (i1, . . . , i2m) : [2dm] → Λ, κπ[I], ϕ(Xi1 (X∗)i2 . . . Xi2m−1 (X∗)i2m) = Xπ∈N C(2dm) where κπ[I] := κπ[xi1(1), . . . , xi1(d) , x∗i2(1), . . . , x∗i2(d) , x∗i2m(1), . . . , x∗i2m(d) ]. {z This gives the equation 2m = Xπ∈N C(2dm) kTk2m from Xi1 (3.2) } {z X I=(i1,...,i2m):[2dm]→Λ from (X ∗)i2 from (X ∗)i2m , . . .{z}... } {z } ai1ai2 . . . ai2m−1ai2m κπ[I]. We now use the fact that the joint ∗-distribution of {xr}r∈Λ is invariant under free complexifi- cation, to show that for many π ∈ N C(2dm), the function I 7→ κπ[I] is identically zero. We state this as the following lemma. Lemma 3.3. Let (A, ϕ) be a NCPS, let z ∈ (A, ϕ) be a Haar unitary which is ∗-free from a family {xr}r∈Λ. Then for any k ∈ N, i : [k] → Λ, ǫ : [k] → {1,∗}, and π ∈ N C(k), we have κπ[(zxi(1))ǫ(1), . . . , (xbi(k))ǫ(k)] 6= 0 only if π ∈ N C ǫ(k). In particular, this quantity is zero if k is odd. Proof. From the multiplicative definition of κπ (π ∈ N C(k)), it suffices to assume that π = 1k. We then need to show that κk[(zxi(1))ǫ(1), . . . , (zxi(k))ǫ(k)] 6= 0 only when ǫ is alternating and k is even. When i(1) = i(2) = . . . = i(d), this just reduces to the proof of the fact that zxi(1) ∈ (A, ϕ) is R-diagonal (Theorem 2.9 (i)). We refer to the proof of [28, Proposition 15.8] for an explicit proof of this fact. For the general case (where i : [k] → Λ is not constant), one just has to notice that the argument in the proof of [28, Proposition 15.8] still applies when i is non-constant - we just have to carry around the extra indices. (cid:3) Since {xr}r∈Λ and {zxr}r∈Λ are identically distributed for any Haar unitary z ∈ (A, ϕ) which is ∗-free from {xr}r∈Λ, Lemma 3.3 implies that each function I 7→ κπ[I] appearing in equation (3.2) is identically zero for any π ∈ N C(2dm)\N C ǫd(2dm), where ǫd : [2dm] → {1,∗} is the pattern given by (The symbol ǫd will denote the above pattern for the rest of the paper.) We now get z ǫd = 1, 1, . . . , 1 d times {z 2m = Xπ∈N C ǫd (2dm) kTk2m } ,∗,∗, . . . ∗ d times {z } X I=(i1,...,i2m):[2dm]→Λ 13 2m groups , . . . , 1, 1, . . . , 1 } {z d times d times ,∗,∗, . . . ∗ . } {z { } ai1ai2 . . . ai2m−1 ai2mκπ[I]. (3.3) When m = 1, we have ǫd = (1, 1, . . . , 1 ), and it is clear in this case that N C ǫd(2d) contains exactly one partition: the fully nested pairing ω = {{1, 2d},{2, 2d−1}, . . . ,{d−1, d+2},{d, d+1}}. Therefore kTk2 d times {z 2 = d times {z ,∗,∗, . . . ,∗ } } XI=(i1,i2):[d]→Λ = Xi1,i2:[d]→Λ = Xi1,i2:[d]→Λ 2 Xi:[d]→Λ = kxk2d ai1ai2κω[I] ai1ai2 dYl=1 ai1ai2kxk2d 2 δi1,i2 ai2, ϕ(xi1(l)x∗i2(d−l+1)) where in the second last line we have used Proposition 2.17 (iii). Remark 3.4. The above calculation tells us that {Xi}i:[d]→Λ is an orthogonal system for each d ∈ N. Furthermore, Lemma 3.3 together with the moment-cumulant formula (2.1) tell us that ϕ(XiX∗j ) = 0 whenever i : [d] → Λ, j : [d′] → Λ are such that d 6= d′. I.e. {1A} ∪S∞d=1{Xi}i:[d]→Λ is an orthogonal system in L2(A, ϕ). We now proceed to the general case m ≥ 2. Applying Holder's inequality to equation (3.3), we have ai1ai2 . . . ai2m−1ai2m · κπ[I] ai1ai2 . . . ai2m−1 ai2m · κπ[I]o 2m × max X X I=(i1,...,i2m):[2dm]→Λ I=(i1,...,i2m):[2dm]→Λ kTk2m ≤ Xπ∈N C ǫd (2dm) ≤ N C ǫd(2dm) π∈N C ǫd (2dm)n π∈N C ǫd (2dm), I:[2dm]→Λnκπ[I]o ≤ N C ǫd(2dm) {z } } {z ai1ai2 . . . ai2m−1ai2m) π∈N C ǫd (2dm)( XI:[2dm]→Λ } {z κπ[I]6=0 max max × (A) (C) (B) . So we need to estimate the quantities (A), (B), and (C). To do this, we first collect some useful facts about the set N C ǫd(2dm). 3.2. Properties of N C ǫd(2dm) and Estimates of (A), (B), and (C). Denote by N C ǫd 2 (2dm) ⊆ N C ǫd(2dm) the subset of all pairings in N C ǫd(2dm). The sets N C ǫd 2 (2dm) and N C ǫd(2dm) have been studied extensively in [16, 22, 27, 30]. We will only record the results from these articles which are relevant to us. Denote by N C(m)(d) the set of d-chains in N C(m): N C(m)(d) :=(cid:8)(σ1, . . . , σd) ∈ N C(m)d : σ1 ≥ σ2 ≥ . . . ≥ σd(cid:9). 14 1 m − 1 (cid:19). What is more, it turns out that the sets N C (d)(m) and N C ǫd In [17], Edelman studied the combinatorics of the set N C(m)(d), and proved that N C(m)(d) = m(cid:18) m(d + 1) 2 (2dm) are naturally in bijection. Lemma 3.5. ([22, Section 3.1 and Corollary 3.2], [30], [27]) For any d, m ∈ N, the sets N C ǫd and N C(m)(d) are in bijection. Therefore 2 (2dm) N C ǫd 2 (2dm) = N C(m)(d) = 1 m(cid:18) m(d + 1) m − 1 (cid:19) ≤ (e(d + 1))m. The last inequality above follows from an application of Sterling's formula. To get a handle on the set N C ǫd(2dm), the idea is to compare N C ǫd(2dm) to the subset 2 (2dm). It is a remarkable fact, proved in [16], that these two sets are quite "close" to being N C ǫd equal. Proposition 3.6. ([16, Theorem 1.5]) For any d, m ∈ N, we have 2 (2dm). N C ǫd(2dm) ≤ 42mN C ǫd Moreover, for any π ∈ N C ǫd(2dm), π has at least dm − 2m blocks of size two, and the size of any block of π is at most 2m. In particular, Proposition 3.6 and Lemma 3.5 give the estimate (3.4) (A) = N C ǫd(2dm) ≤ 42m(e(d + 1))m. We now turn to the estimation of (B). To do this, we use the following elementary free cumulant bound, variants of which can also be found in [16, Lemma 3.1] and [22, Lemma 4.3]. Lemma 3.7. Let σ ∈ N C(n), and suppose σ has at least K blocks of size 2 and all blocks σ have size at most N . Then for any centered family of random variables {b1, . . . , bn} in a tracial C∗-probability space (A, ϕ), we have κσ[b1, . . . , bn] ≤(cid:16) max i kbikL2(A,ϕ)(cid:17)2K(cid:16)16 max i kbikLN (A,ϕ)(cid:17)n−2K Proof. Since both sides of the claimed inequality are multiplicative with respect to the block struc- ture of σ, it suffices to prove this result for the case σ = 1n. The case n = 1 is trivial since κ1[b1] = ϕ[b1] = 0. . When n = 2, we have necessarily K = 1, and since the bi's are centered, κ2[b1, b2] = ϕ(b1b2) − ϕ(b1)ϕ(b2) = ϕ(b1b2). Therefore by the Cauchy-Schwarz inequality for ϕ, we have κ2[b1, b2] ≤ kb∗1kL2(A,ϕ)kb2kL2(A,ϕ) = kb1kL2(A,ϕ)kb2kL2(A,ϕ) ≤(cid:16) max i kbikL2(A,ϕ)(cid:17)2 . When n ≥ 3, we have K = 0. Using equation (2.2), we have κn[b1, . . . , bn] =Pπ∈N C(n) µ(π, 1n)ϕπ[b1, . . . , bn]. By taking absolute values and using the fact that µ(π, 1n) ≤ Cn−1 for all π ∈ N C(n), we get ϕπ[b1, . . . , bn] κn[b1, . . . , bn] ≤ Cn−1 Xπ∈N C(n) ≤ CnCn−1(cid:16) max ≤ CnCn−1 max 15 π∈N C(n) ϕπ[b1, . . . , bn] i kbikLN (A,ϕ)(cid:17)n , where in the last line we have applied Holders inequality (for the trace ϕ) to the quantities ϕπ[b1, . . . , bn], and used the fact that kxkLr (A,ϕ) ≤ kxkLN (A,ϕ), for all x ∈ A and r ≤ N . The proof is now completed by noting that Ck ≤ 4k for any k. (cid:3) Using Lemma 3.7, we obtain the following estimate for (B). Corollary 3.8. For any π ∈ N C ǫd(2dm) and any function I : [2dm] → Λ, (3.5) . κπ[I] ≤ kxk2dm 2 (cid:16) 16kxk2m kxk2 (cid:17)4m Proof. From Lemma 3.7 and the fact that the variables {xr}r∈Λ are identically distributed according to our reference variable x, we have κπ[I] ≤ kxk2K 2 (16kxk2m)2dm−2K , where K denotes the number of blocks of π which are pairings. Now, since π ∈ N C ǫd(2dm), Proposition 3.6 implies that K ≥ dm − 2m. Using this estimate in the above inequality we have kxk2K 2 (16kxk2m)2dm−2K = kxk2dm ≤ kxk2dm kxk2 (cid:17)2dm−2K 2 (cid:16) 16kxk2m kxk2 (cid:17)4m 2 (cid:16) 16kxk2m . (cid:3) We will now prove the inequality (3.6) (C) = π∈N C ǫd (2dm)( XI:[2dm]→Λ max ai1ai2 . . . ai2m−1 ai2m) ≤ (3e)4m(cid:16) Xi:[d]→Λ ai2(cid:17)m . κπ[I]6=0 To do this, we will need the following two lemmas. Lemma 3.9. Let π ∈ N C ǫd(2dm), and fix σ ≤ π. Then σ ∈ N C ǫd(2dm) if and only if σ ∈ N Ce(2dm). Furthermore, {σ ∈ N C ǫd(2dm) : σ ≤ π} ≤ N C(2m)(2) ≤ (3e)2m. Proof. Fix π ∈ N C ǫd(2dm), and let σ ≤ π. The "only if" part of the first statement follows from the definition of N C ǫd(2dm). To prove the "if" part, suppose σ ∈ N Ce(2dm). Let V = {j1 < j2 < . . . < j2r} be a block of σ. Since σ is even and non-crossing, a simple inductive argument shows that the parity of the sequence j1, j2, . . . , j2r must be alternating. Now let W be the block of π containing V . Then by definition ǫdW is alternating, and since the parity of the elements of V ⊆ W alternates as well, ǫdV = (ǫd(j1), . . . , ǫd(j2r)) is also alternating. As V was arbitrary, we get σ ∈ N C ǫd(2dm). We now prove the second statement of the lemma. By Proposition 3.6, π contains at least dm − 2m blocks of size 2. Let A ⊆ [2dm] denote the union of these blocks of size 2. For any σ ∈ N C ǫd(2dm) ⊆ N Ce(2dm) with σ ≤ π, we must then have σA = πA, and σ[2dm]\A ≤ π[2dm]\A ∈ N Ce([2dm]\A). Therefore we clearly have an injection {σ ∈ N C ǫd(2dm) : σ ≤ π} ֒→ N Ce([2dm]\A), σ 7→ σ[2dm]\A. This gives {σ ∈ N C ǫd(2dm) : σ ≤ π} ≤ N Ce([2dm]\A) ≤ N Ce(4m), 16 since [2dm]\A ≤ 4m. To complete the proof, we use the fact that there is a bijection between N Ce(4m) and the set of 2-chains N C(2m)(2) (cf. [28, Exercise 9.42]), together with the bound on N C(2m)(2) given by Lemma 3.5. (cid:3) Lemma 3.10. Let {ai}i:[d]→Λ ⊂ C be a finitely supported family of complex numbers, and let π ∈ P ǫd(2dm). Then ai1 ai2 . . . ai2m−1ai2m ≤(cid:16) Xi:[d]→Λ Proof. Given π ∈ P ǫd(2dm), associate to it a pairing πr ∈ P ǫd for each block V = {k1 < k2 < . . . < k2s} of π, declare {k1, k2}, . . . ,{k2r−1, k2s} to be blocks of πr. Observe that πr ≤ π by construction. In particular, for all I : [2dm] → Λ, XI:[2dm]→Λ ai2(cid:17)m 2 (2dm), as follows: ker I≥π . Therefore XI:[2dm]→Λ ker I≥π ker I ≥ π =⇒ ker I ≥ πr. ai1ai2 . . . ai2m−1 ai2m ≤ XI:[2dm]→Λ = XI:[2dm]→Λ ker I≥πr ker I≥πr ai1ai2 . . . ai2m−1 ai2m mYk=1 ai2k−1 mYk=1 ai2k. Let B1 := ǫ−1 d {1} and B∗ := ǫ−1 d {∗}, so that [2dm] = B1 ∪B∗. Since πr ∈ P ǫd(2dm), we can identify πr with the bijection πr : B1 → B∗, πr(t) = s ⇐⇒ {t, s} is a block of πr. Now fix a function I = (i1, . . . , i2m) : [2dm] → Λ. Let I1 = (i1, i3, . . . , i2m−3, i2m−1) = IB1, and let I∗ = (i2, i4, . . . , i2m−2, i2m) = IB∗ . Then we can rewrite the condition ker I ≥ πr as I∗ = I1 ◦ πr. k=1 ai2k our previous inequality now becomes Writing AI1 =Qm k=1 ai2k−1 and AI∗ =Qm (by the Cauchy-Schwarz inequality) A2 I1 (since πr is a bijection) ai12ai32 . . . ai2m−12 =(cid:16) Xi:[d]→Λ ai2(cid:17)m . 17 (cid:3) ai1ai2 . . . ai2m−1ai2m ≤ XI:[2dm]→Λ ker I≥πr AI1AI1◦πr A2 I1(cid:17)1/2(cid:16) XI1=(i1,i3,...,i2m−1) AI1AI∗ A2 I1◦πr(cid:17)1/2 = ker I≥π XI:[2dm]→Λ XI1=(i1,i3,...,i2m−1) ≤ (cid:16) XI1=(i1,i3,...,i2m−1) XI1=(i1,i3,...,i2m−1) Xi1,i3,...,i2m−1:[d]→Λ = = Using Lemmas 3.9 and 3.10, we are now ready to obtain the bound (3.6). Fix π ∈ N C ǫd(2dm) and consider the sum XI:[2dm]→Λ κπ[I]6=0 ai1ai2 . . . ai2m−1 ai2m. For any I(i1, i2, . . . , i2m) : [2dm] → Λ, we can use equation (2.3) to write κπ[I] = Xσ∈N C(2dm) σ≤π µ(σ, π)ϕσ[I], (3.7) ϕσ[I] := ϕσ[xi1(1), . . . , xi1(d), x∗i2(1), . . . , x∗i2(d), . . . , x∗i2m(1), . . . , x∗i2m(d)]. In particular, κπ[I] 6= 0 only if there exists some σ ∈ N C(2dm) with σ ≤ π such that ϕσ[I] 6= 0. Fix such a σ ≤ π. Since {xr}r∈Λ has an even joint ∗-distribution (Proposition 2.17), it follows that ϕσ[I] 6= 0 only if each block of σ has even cardinality. I.e. σ ∈ N Ce(2dm), which is equivalent to σ ∈ N C ǫd(2dm) by Lemma 3.9. Therefore, by simply over-counting, we have the estimate κπ[I]6=0 ai1ai2 . . . ai2m−1 ai2m XI:[2dm]→Λ ≤ Xσ∈N C ǫd (2dm) ≤ {σ ∈ N C ǫd(2dm) : σ ≤ π} ϕσ[I]6=0 σ≤π XI:[2dm]→Λ ( XI:[2dm]→Λ ( XI:[2dm]→Λ ϕσ[I]6=0 ϕσ[I]6=0 × σ∈N C ǫd (2dm): max σ≤π ≤ (3e)2m σ∈N C ǫd (2dm): max σ≤π ai1ai2 . . . ai2m−1ai2m ai1ai2 . . . ai2m−1 ai2m) ai1ai2 . . . ai2m−1 ai2m), where in the last line we have used Lemma 3.9. Now fix σ ∈ N C ǫd(2dm) with σ ≤ π. By applying Proposition 2.17 (i) to each block V of σ, we get ϕσ[I] 6= 0 only if there exists some ρ ∈ N Ce(2dm) such that ker I ≥ ρ and ρ ≤ σ. (In particular, ρ ∈ N C ǫd(2dm), by Lemma 3.9). Therefore, by possibly over-counting again, we have 18 the uniform estimate ϕσ[I]6=0 ai1ai2 . . . ai2m−1 ai2m XI:[2dm]→Λ ≤ Xρ∈N C ǫd (2dm): ≤ {ρ ∈ N C ǫd(2dm) : ρ ≤ σ} ρ≤σ ker I≥ρ XI:[2dm]→Λ ( XI:[2dm]→Λ ai2(cid:17)m ker I≥ρ × ρ∈N C ǫd (2dm): max ρ≤σ ≤ (3e)2m(cid:16) Xi:[d]→Λ ai1ai2 . . . ai2m−1 ai2m ai1ai2 . . . ai2m−1ai2m) where in the last line we have used both Lemmas 3.9 and 3.10. This completes the proof of inequality (3.6). Note that the presence of the "checks" i2k, 1 ≤ k ≤ m, in the previous sums do not effect the applicability of Lemma 3.10, since for any ρ ∈ N C ǫd(2dm) we can write XI:[2dm]→Λ ker I≥ρ ai1ai2 . . . ai2m−1 ai2m = XI:[2dm]→Λ ker I≥ρ′ ai1ai2 . . . ai2m−1ai2m, where ρ′ ∈ P ǫd(2dm) is the unique partition with the property that ker(i1, . . . , i2m) ≥ ρ ⇐⇒ ker(i1, i2, . . . , i2m−1,i2m) ≥ ρ′, for all I = (i1, . . . , i2m) : [2dm] → Λ. Putting inequalities (3.4), (3.5), and (3.6) together we get kTk2m 2m ≤ (A) · (B) · (C) 2 (cid:16) 16kxk2m kxk2 (cid:17)4m ≤ 42m(e(d + 1))m · kxk2dm = 410m34me5m(d + 1)m(cid:16)kxk2m kxk2 (cid:17)4m kTk2m 2 . 2 p(d + 1)kTk2, kTk2m ≤ 45 · (3e)2√ekxk2 kxk2 2m After taking 2m'th roots we get · (3e)4m(cid:16) Xi:[d]→Λ ai2(cid:17)m (m ∈ N). This completes the proof of Theorem 3.1, since the lower bound kTk2 ≤ kTk2m is automatic. 4. Strong Haagerup Inequalities for Bi-Invariant Arrays. In this section, we extend Theorem 3.1 to the case where we have an n × n array {xrs}1≤r,s≤n n -bi- n in Section of random variables in a tracial C∗-probability space (A, ϕ), whose joint ∗-distribution is H + invariant. We will need this extension to prove our strong Haagerup inequalities for U + 6. Here is the main result. Theorem 4.1. Let (A, ϕ) be a tracial C∗-probability space, and suppose {xrs}1≤r,s≤n ⊂ (A, ϕ) is an n× n array whose joint ∗-distribution is H + n -bi-invariant and invariant under free complexification. 19 Let x ∈ {xr,s}1≤r,s≤n be a fixed reference variable. Then for any homogeneous polynomial T = Xk,l:[d]→[n] ak,lxk(1)l(1)xk(2)l(2) . . . xk(d)l(d), (ak,l ∈ C), of degree d in the variables {xrs}1≤r,s≤n and any p ∈ 2N ∪ {∞}, we have kTkL2(A,ϕ) ≤ kTkLp(A,ϕ) ≤ 45 · (3e)3√ekxk2 kxk2 Lp(A,ϕ) L2(A,ϕ) √d + 1kTkL2(A,ϕ). Remark 4.2. Note that there is an additional factor of 3e in the constant of Theorem 4.1 that does not appear in Theorem 3.1. Proof. The proof of this result is almost identical to that of Theorem 3.1, so we only sketch it. Fix d ∈ N, and T as in the statement of the theorem. Let Λ = [n] × [n], and identify any pair of functions k, l : [d] → [n] with the function i : [d] → Λ given by With this change of indices, T can be written as i(j) = (k(j), l(j)), (j ∈ [d]). T = Xi:[d]→Λ aixi(1)xi(2) . . . xi(d), (ai ∈ C). Repeating the arguments in the proof of Theorem 3.1 (and using the same notation therein), we get the same (in)equalities we had there: kTk2 2 = kxk2d 2 Xi:[d]→Λ ai2 = kxk2d 2 Xk,l:[d]→[n] ak,l2, and for general m ∈ N, kTk2m 2m ≤ N C ǫd(2dm) } (A′) (B′) max π∈N C ǫd (2dm), I:[2dm]→Λnκπ[I]o {z } {z ai1ai2 . . . ai2m−1 ai2m) π∈N C ǫd (2dm)( XI:[2dm]→Λ } {z κπ[I]6=0 max (C′) . × These (in)equalities remain valid in our new setting because, up to this point in the proof, they only rely on the fact that our random variables are orthogonal in L2(A, ϕ), identically distributed, and that their joint ∗-distribution is invariant under free complexification. Consider the quantities (A), (B), and (C) defined in Section 3.1. Obviously we have (A′) = (A). Also note that the bound (3.5) obtained for (B) only relied on the structure of collection of partitions N C ǫd(2dm), so the same bound applies to (B′): (4.1) (4.2) For (C′) we will prove the inequality . (B′) ≤ kxk2dm kxk2 (cid:17)4m 2 (cid:16) 16kxk2m ak,l2(cid:17)m (C′) ≤ (3e)6m(cid:16) Xk,l:[d]→[n] . 20 Theorem 4.1 now follows from the inequalities (3.4), (4.1), and (4.2). The proof of (4.2) is only a minor modification of the proof of inequality (3.6) for (C). Indeed, for any π ∈ N C ǫd(2dm), equation (2.3) and Lemma 3.9 give (just as before) κπ[I] = Xσ∈N C ǫd (2dm) σ≤π µ(σ, π)ϕσ[I], and therefore κπ[I]6=0 XI:[2dm]→Λ ≤ Xσ∈N C ǫd (2dm) σ≤π ≤ (3e)2m ai1ai2 . . . ai2m−1 ai2m XI:[2dm]→Λ ϕσ[I]6=0 ai1ai2 . . . ai2m−1ai2m σ∈N C ǫd (2dm): max σ≤π ( XI:[2dm]→Λ ϕσ[I]6=0 ai1ai2 . . . ai2m−1 ai2m), where in the last line we have applied Lemma 3.9. Fix σ ∈ N C ǫd(2dm) with σ ≤ π, and write ai1ai2 . . . ai2m−1ai2m XI:[2dm]→Λ = XK,L:[2dm]→[n] ϕσ[I]6=0 ϕσ[K,L]6=0 ak1,l1ak2,l2 . . . ak2m−1,l2m−1ak2m,l2m, where ϕσ[K, L] = ϕσ[xk1(1)l1(1), . . . , xk1(d)l1(d), . . . , x∗k2m(1)l2m(1), . . . , x∗k2m(d)l2m(d)] = ϕσ[I]. Next, observe that H + n -bi-invariance (i.e. Proposition 2.17 (ii) applied to each block of σ) implies that ϕσ[K, L] 6= 0 only if there exist ρ, δ ∈ N Ce(2dm) such that ρ, δ ≤ σ (so ρ, δ ∈ N C ǫd(2dm) by Lemma 3.9) with the property that ker K ≥ ρ, and ker L ≥ σ. Using this fact, we can over count 21 the possible non-zero contributions to the above sum and get ≤ ρ,δ≤σ ϕσ[K,L]6=0 ak1,l1ak2,l2 XK,L:[2dm]→[n] XK,L:[2dm]→[n] Xρ,δ∈N C ǫd (2dm) ker I≥ρ, ker J≥δ ≤ {ρ ∈ N C ǫd(2dm) : ρ ≤ σ}2 ρ,δ∈N C ǫd (2dm)( XK,L:[2dm]→[n] ρ,δ∈N C ǫd (2dm)( XK,L:[2dm]→[n] ≤ (3e)4m × × max max ker K≥ρ, ker L≥δ ker K≥ρ, ker L≥δ . . . ak2m−1,l2m−1ak2m,l2m ak1,l1ak2,l2 . . . ak2m−1,l2m−1 ak2m,l2m ak1,l1ak2,l2 . . . ak2m−1,l2m−1 ak2m,l2m) ak1,l1ak2,l2 . . . ak2m−1,l2m−1 ak2m,l2m), where in the last inequality we have used Lemma 3.9. Finally, we rewrite the sums XK,L:[2dm]→[n] ker K≥ρ, ker L≥δ ak1,l1ak2,l2 . . . ak2m−1,l2m−1 ak2m,l2m, (ρ, δ ∈ N C ǫd(2dm)), appearing above, in the equivalent form J=(k1,l1,k2,l2,...,k2m,l2m):[2(2d)m]→[n] X ker J≥ρ⊔δ ak1,l1ak2,l2 . . . ak2m−1,l2m−1 ak2m,l2m, where ρ ⊔ δ ∈ P([2dm] ⊔ [2dm]) ∼= P(4dm) is the disjoint union of ρ and δ. Since it is obvious that ρ⊔ δ ∈ P ǫ2d(4dm), we may apply Lemma 3.10 (the same way we did in the proof of inequality (3.6)) to the above sums obtain XK,L:[2dm]→[n] ker K≥ρ, ker L≥δ ak1,l1ak2,l2 . . . ak2m−1,l2m−1ak2m,l2m ≤ (cid:16) Xk,l:[d]→[n] ak,l2(cid:17)m , for all ρ, δ ∈ N C ǫd(2dm). Putting all these uniform estimates together gives the bound (4.2). (cid:3) 5. Application to the Metric Approximation Property In this section we study the norm-closed, non-self-adjoint (unital) operator algebra B generated by a family of random variables satisfying the hypotheses of Theorems 3.1 or 4.1. We show that B always has the metric approximation property, and derive some other intermediate results along the way, which may be of independent interest. We begin by recalling the definition of the metric approximation property. Definition 5.1. Let Y be a Banach space. We say that Y has the metric approximation property (MAP) if there exists a net {Sα}α∈A ⊂ B(Y ) of finite rank contractions converging to the identity map in the strong operator topology (s.o.t.) on B(Y ). That is, for all y ∈ Y lim α∈A kSαy − yk = 0. 22 Let (A, ϕ) be a tracial C∗-probability space, and let X = {xr}r∈Λ be a family of operators satisfying the hypotheses of Theorem 3.1 or Theorem 4.1. Denote by B = BX ⊆ A the norm- closed, non-self-adjoint, unital operator algebra generated by {xr}r∈Λ. We prove: Theorem 5.2. The operator algebra B has the metric approximation property. To prove Theorem 5.2, we use a fairly standard truncation argument, originating from Haagerup in [18]. We start with some notation and two preliminary results. Let (A, ϕ) be a C∗-probability space (not necessarily tracial), let X = {xr}r∈Λ ⊂ (A, ϕ) be a family of random variables, and assume that the joint ∗-distribution of X is invariant under free complexification. Let B = BX ⊆ A be the norm-closed, non-self-adjoint, unital operator algebra generated by X. Denote by L2(B) := Bk·k2 ⊆ L2(A, ϕ), the Hilbert space generated by B, and for d ∈ N ∪ {0}, let Pd : L2(B) → L2 projection onto the degree d subspace d(B) be the orthogonal d(B) := span{Xi = xi(1) . . . xi(d) : i : [d] → Λ}k·k2. L2 (Here we use the convention L2 0(B) := C1A). Since the joint ∗-distribution of X is invariant under free complexification, it follows from Lemma 3.3 and the moment cumulant formula (2.1), that the subspaces {L2 Proposition 5.3. Consider the Ornstein-Uhlenbeck type semigroup {Γt}t≥0 ⊂ B(L2(B)) given by d(B)}d∈N∪{0} are orthogonal. Γt = ∞Xd=0 e−dtPd. Then {Γt}t≥0 is a contraction semigroup on L2(B), furthermore for each t ≥ 0, Γt restricts to a unital complete contraction Γt : B → B. Proof. The first statement is immediate since {Pd}d≥0 is an orthogonal family of projections on L2(B) and e−dt ≤ 1 for all t, d ≥ 0. When t = 0 the second statement is also immediate since Γ0B = idB, so assume for the remainder that t > 0. Let T denote the unit circle in the complex plane and let z = idT denote the canonical unitary generator of the C∗-algebra C(T). Then z is a Haar unitary in the C∗-probability space (C(T), ψ), where ψ denotes integration with respect to normalized Haar measure on T. Since the joint ∗- distribution of X is invariant under free complexification, there exists a state-preserving injective ∗-homomorphism of C∗-probability spaces π :(cid:0)C∗h1A, Xi, ϕ(cid:1) → (cid:0)C(T) ∗red C∗h1A, Xi, ψ ∗ ϕ(cid:1), π(xr) = zxr, where C∗h1A, Xi ⊂ A is the unital C∗-algebra generated by X. Let {ρt}t>0 denote the Poisson convolution semigroup on T, with convolution kernel given by the probability density Pt(eiθ) = 1−2e−tcos(θ)+e−2t for θ ∈ [0, 2π). It is well known that ρt acts on C(T) by the formula 1−e−2t ρt(zn) = Pt ∗ (zn) = e−ntzn, 23 (n ∈ Z). Furthermore, since Pt is a positive kernel for all t > 0, ρt is completely positive on C(T). On the other hand, for any d ∈ N and i : [d] → Λ, we have the identity π(Γt(xi(1)xi(2) . . . xi(d))) = π(e−dtxi(1)xi(2) . . . xi(d)) = e−dtzxi(1)zxi(2) . . . zxi(d) = (ρtz)xi(1)(ρtz)xi(2) . . . (ρtz)xi(d) = (ρt ∗ id)(zxi(1)zxi(2) . . . zxi(d)) = (ρt ∗ id)π(xi(1)xi(2) . . . xi(d)), where ρt ∗ id : C(T) ∗red C∗h1A, Xi → C(T) ∗red C∗h1A, Xi is the reduced free product of the completely positive, state-preserving maps ρt and idC ∗h1A,Xi (which, by [11], is again completely positive and state preserving). Therefore, by linearity and continuity, we have π ◦ Γt = (ρt ∗ id) ◦ πB, (t > 0). Finally, since π is a complete isometry, the completely bounded norm kΓtkcb = kΓtkCB(B) satisfies kΓtkcb = kπ ◦ Γtkcb = k(ρt ∗ id) ◦ πBkcb ≤ kρt ∗ idkcb = 1, (t > 0). (cid:3) Remark 5.4. It is easy to see that (ρt ∗ id)π(C∗h1A, Xi) ⊂ π(C∗h1A, Xi). Therefore {π−1 ◦ (ρt ∗ id) ◦ π}t>0 ⊂ CB(C∗h1A, Xi) is a ϕ-preserving completely positive extension of the semigroup {Γt}t>0 defined on B. This extension problem has been previously considered in the context of free R-diagonal families in [21]. Remark 5.5. We use the name "Ornstein-Uhlenbeck" in the previous proposition because of the fact that when X is a free circular system, Γt actually is the free Ornstein-Uhlenbeck semigroup [22]. Now suppose that X = {xr}r∈Λ ⊂ (A, ϕ) satisfies the hypotheses of Theorem 3.1 or 4.1. Let Λ denote the free semigroup with Λ generators {gr}r∈Λ, and denote by Wd ⊂ F+ F+ Λ the set of words of length d in F+ Λ . Given i : [d] → Λ, recall that Xi := xi(1)xi(2) . . . xi(d) and similarly put Λ . From Remark 3.4, it follows that the map gi 7→ Xi, identifies ℓ2(F+ gi := gi(1)gi(2) . . . gi(d) ∈ F+ Λ ) with L2(B), and consequently any ψ ∈ ℓ∞(F+ Λ ) defines a multiplication operator Mψ ∈ B(L2(B)) given by Mψ(Xi) = ψ(gi)Xi, (i : [d] → Λ). Λ ). The next lemma says that if ψ ∈ ℓ∞(F+ Λ ) decays sufficiently Note that kMψkB(L2(B)) = kψkℓ∞(F+ rapidly, then Mψ(L2(B)) ⊆ B. Lemma 5.6. Let ψ ∈ ℓ∞(F+ Λ ) be such that K(ψ) := sup d≥0 (d + 1)3/2kψWdk∞ < ∞. Then Mψ(L2(B)) ⊆ B and kMψkB(L2(B),B) ≤ CBK(ψ), where CB is a constant only depending on B. Proof. Let ψ ∈ ℓ∞(F+ for B (Theorem 3.1 or 4.1), there is a constant cB > 0 such that (T ∈ L2 Λ ) satisfy the above hypothesis. Then, from our strong Haagerup inequality kTkB ≤ cB √d + 1kTkL2(B), 24 d(B)). Put CB = cB(cid:16)P∞d=0 all d ≥ 0) 1 (d+1)2(cid:17)1/2 kMψTkB = (cid:13)(cid:13)(cid:13) . Then for any T ∈ L2(B) we have (noting that PdMψ = MψPd for ∞Xd=0 kMψPdTkB PdMψT(cid:13)(cid:13)(cid:13)B ≤ ∞Xd=0 ∞Xd=0 √d + 1kMψPdTkL2(B) ∞Xd=0 √d + 1kψWdk∞kPdTkL2(B) ∞Xd=0 √d + 1 (d + 1)3/2 kPdTkL2(B) K(ψ) ≤ cB ≤ cB ≤ cB ≤ cBK(ψ)(cid:16) ∞Xd=0 1 (d + 1)2(cid:17)1/2(cid:16) ∞Xd=0 kPdTk2 L2(B)(cid:17)1/2 = CBK(ψ)kTkL2(B). Hence MψL2(B) ⊆ B and kMψkB(L2(B),B) ≤ CBK(ψ). We are now ready to prove the main theorem of this section. (cid:3) 5.1. Proof of Theorem 5.2. Denote by χWd the characteristic function of the set Wd. For each N ∈ N and t > 0, define Γt,N : B → B, Γt,N = e−dtPd = e−dtMχWd . NXd=0 NXd=0 (5.1) Let ψt,N =Pd≥N +1 e−dtχWd ∈ ℓ∞(F+ Γt − Γt,N = Xd≥N +1 e−dtPd = Xd≥N +1 Since the inclusion B ֒→ L2(B) is a contraction, we have Λ ), so that e−dtMχWd = Mψt,N . kΓt − Γt,NkB(B) ≤ kΓt − Γt,NkB(L2(B),B) ≤ CB sup d≥N +1 (d + 1)3/2e−dt, where the last inequality follows from (5.1) and Lemma 5.6. So for each t > 0, lim (5.2) N→∞kΓt,N − ΓtkB(B) = 0, and Let Qt,N = kΓt,Nk−1Γt,N . Then (5.2) also gives (5.3) lim N→∞kQt,N − ΓtkB(B) = 0, (t > 0). lim N→∞kΓt,NkB(B) = kΓtkB(B) = 1. We now claim that the identity map idB : B → B is contained in the strong closure of the set of contractions {Qt,N}t>0,N∈N ⊂ B(B). To prove this, first note that by (5.3), {Γt}t>0 is contained in the strong closure of {Qt,N}t>0,N∈N. Next, note that since limt→0 e−dt = 1 for all d > 0, we have limt→0 kΓtT − TkA = 0 for any polynomial T ∈ Algh1A, Xi. Since {Γt}t>0 is uniformly norm- bounded (by Proposition 5.3), this limit is valid for all T ∈ B = Algh1A, Xik·kA. Therefore idB is contained in {Γt}t>0 If Λ < ∞, then each of the maps Qt,N is finite rank, so the MAP for B follows from the fact that {Qt,N}t>0,N∈N contains idB in its strong closure. If Λ = ∞, then the contractions {Qt,N}t>0,N∈N ⊂ {Qt,N}t>0,n∈N , proving the claim. s.o.t. s.o.t. 25 constructed above are no longer finite rank. Let F denote the collection of finite subsets of Λ. For each F ∈ F, let XF = {xr}r∈F , and let EF be the unique ϕ-preserving conditional expectation from the von Neumann algebra C∗h1A, Xi′′ ⊆ B(L2(A, ϕ)) onto C∗h1A, XFi′′ ⊆ B(L2(A, ϕ)). (Recall that on the L2-level, EB is just the orthogonal projection from L2(B) onto the Hilbert subspace generated by {1A, XF}). Put Qt,N,F = Qt,N ◦ EFB. Then {Qt,N,F}t>0,N∈N is a family of finite rank contractions on B, which, by the argument in the previous paragraph, contains EFB in its strong closure. For any T ∈ B we have limF∈F kEF T − Tk = 0, and therefore {Qt,N,F}t>0,N∈N,F∈F contains idB in its strong closure, so B has the MAP when Λ = ∞ as well. 6. Applications to Free Unitary Quantum Groups In this final section, we consider applications of our results to the reduced C∗-algebra associated to the free unitary quantum group U + n (of dimension n), introduced by Wang [35]. Let us briefly recall the definition of this quantum group: U + n is the CMQG given by the pair (Au(n), U ), where Au(n) is the universal C∗-algebra with generators {urs : 1 ≤ r, s ≤ n} subject to the relations which make the matrices U = [urs]1≤r,s≤n and U = [u∗rs]1≤r,s≤n unitary in Mn(Au(n)). It is clear that for each n ∈ N, U + n = (Au(n), U ) satisfies Definition 2.10. The free orthogonal quantum group O+ n (of dimension n), also introduced in [35], will also be of use to us here. O+ n is the CMQG given by the pair (Ao(n), V ), where Ao(n) = Au(n)/hurs = u∗rs : 1 ≤ r, s ≤ ni, and V = [vrs]1≤r,s≤n is the image of U under the canonical quotient map Au(n) → Ao(n). For the rest of this section (and with a slight abuse of notation), we will also use the symbols urs and vrs to denote the canonical generators of L∞(U + n ), respectively. We also note that these quantum groups are always unimodular, i.e. their Haar states are tracial [4]. n ) and L∞(O+ In recent years, the series {U + n }n∈N have been intensively studied from both operator algebraic and probabilistic perspectives [1, 4, 6, 7, 8, 33, 34]. In particular, Vergnioux has proved a version of Haagerup's inequality for O+ n [34, Section 4.3]. Vergnioux's result can be formulated as follows (refer to [32] for the unexplained terminologies below). Let G = (A, U ) be a n }n∈N and {O+ n and U + {U α = [uα ij]1≤i,j≤dα}α∈ bG, classes of irreducible finite dimensional unitary corepresentations of G, and let CMQG with fundamental corepresentation U , let bG denote the collection of all unitary equivalence be a complete family of representatives for bG. Then by the Peter-Weyl Theorem ([36]) forms an orthogonal basis for L2(G). Denote by C the category of equivalence classes of finite dimen- sional unitary corepresentations of G, and let S = {α1, . . . αs} ⊆ bG be a generating set for C which is closed under conjugation of representations and does not contain the trivial corepresentation 1A. Let ℓ = ℓS : bG → N be the "word length" function given by ij : α ∈ bG, 1 ≤ i, j ≤ dα} ⊂ L∞(G) {uα l(α) = min{k ∈ N : α ⊆ αi(1) ⊠ . . . ⊠ αi(k), αi(j) ∈ S}, and define Then we have the following definition. L2 d(G) := span{uα ij : ℓ(α) = d}k·k2 ⊂ L2(G). Definition 6.1. ([34]) G has the property of rapid decay (property RD) (with respect to ℓ = ℓS : bG → N) if there exists a polynomial P ∈ R+[x] such that for all T ∈ L2 kTkL∞(G) ≤ P (d)kTkL2(G) d(G). 26 For O+ U + g1, g2 are the generators of F+ n , there is a natural labeling cO+ n = {V (k)}k∈N∪{0} such that 1 = V (0) and V = V (1). For such that 1 = U e, U = U g1, and U = U g2, where n , and SU = {U, U} as n , the corresponding length functions ℓSO , ℓSU are identified with the natural 2 , respectively [34, Section 4.3]. We collect here the main result n there is a natural labeling cU + generating set for cU + length functions on N ∪ {0} and F+ from [34, Section 4.3]. Theorem 6.2. For each n ∈ N, there exist positive constants Cn, Dn > 0 such that O+ RD with P (x) = Cn(x + 1), and U + n = {U g}g∈F+ 2 . Taking SO = {V } as a generating set for cO+ n has property RD with P (x) = Dn(x + 1). n has property 2 Viewing L∞(U + n ) as a non-cocommutative analogue of L(Fn) = C∗λ(Fn)′′, Theorem 6.2 can be regarded as the non-cocommutative analogue of Haagerup's classical inequality (Theorem 1.1) for L(Fn). It is interesting to note that the order of growth (with respect to d) in Theorems 1.1 and 6.2 is the same. Using our results from the previous sections we can also obtain the following non- cocommutative analogue of Theorem 1.2, which improves on the growth of the bounds in Theorem 6.2. Theorem 6.3. Let Bn ⊂ L∞(U + by the coefficients {urs}1≤r,s≤n of the fundamental corepresentation of U + Then for any d ∈ N, p ∈ 2N ∪ {∞}, and any T ∈ L2 n ) ∩ Bn, we have n ) ≤ 46 · (3e)3√e√d + 1kTkL2(U + n ). n ) be the norm-closed, non-self-adjoint unital subalgebra generated n (and not their adjoints). Proof. Let h : L∞(U + (equation (2.4)) implies that the joint ∗-distribution of the array {urs}1≤r,s≤n ⊂ (L∞(U + bi-invariant. By comparing the defining relations for H + subgroup of U + kTkL2(U + n ) → C denote the Haar state. The bi-invariance of h with respect to ∆ n ), h) is U + n - n is a quantum n , it is clear that H + n ) ≤ kTkLp(U + n . So by Remark 2.13, {urs}1≤r,s≤n has an H + n -bi-invariant joint ∗-distribution. In [1] (see also [4, Theorem 9.2]), it was shown that {urs}1≤r,s≤n is the free complexification of {vrs}1≤r,s≤n. In particular, the joint ∗-distribution of {urs}1≤r,s≤n is invariant under free complex- ification by Remark 2.8. Therefore the array {urs}1≤r,s≤n ⊂ (L∞(U + n ), h) satisfies the hypotheses of Theorem 4.1. Since the linear span of the homogeneous polynomials of degree d in the variables {urs}1≤r,s≤n ⊂ L∞(U + n ) is precisely L2 n and U + d(U + d(U + n ) ∩ Bn [1, Theorem 1], Theorem 4.1 gives n ) ≤ 45 · (3e)3√eku11k2 ku11k2 √d + 1kTkL2(U + n ), p n ) ≤ kTkLp(U + kTkL2(U + n )∩Bn. To complete the proof, we use a result of Banica, Collins, and Zinn-Justin n ) with respect for any T ∈ L2 [8, Theorem 5.3], which says that the spectral measure of any generator vrs ∈ L∞(O+ to the Haar state has support equal toh −2√n+2 2 = n−1, we get 2√n+2i. Since also kv11k2 d(U + 2 , p ku11k2 ku11k2 2 ≤ ku11k2 ∞ ku11k2 2 = kzv11k2 ∞ kzv11k2 2 = kv11k2 ∞ kv11k2 2 = 4n n + 2 ≤ 4. (cid:3) n ) and χU = (T r⊗ id)U ∈ L∞(U + n ), h), and χU is a standard circular random variable in (L∞(U + Remark 6.4. We can show that the order of the growth of the bounds in both Theorems 6.2 and 6.3 is optimal. Let χO = (T r⊗ id)V ∈ L∞(O+ n ) denote the fundamental characters of O+ n , respectively. Then χO is a standard semicircular random variable in (L∞(O+ n ), h) [1, 4]. Let {Td}d∈N∪{0} denote the Chebyshev II polynomials determined by the initial conditions T0(x) = 1, T1(x) = x and the recursion n and U + (6.1) xTd(x) = Td+1(x) + Td−1(x), 27 (d ≥ 1). It is well known that these polynomials are orthonormal for the standard semicircular law. Since both χo and √2ReχU are standard semicircular, functional calculus gives kTd(χO)kL∞(O+ n ) = kTd(√2ReχU )kL∞(U + n ) = sup t∈[−2,2]Td(t) = d + 1. n ). So the growth rate of O(d + 1) given by Theorem 6.2 is actually obtained. On the other hand, a simple inductive argument on d ∈ N using the orthogonality of the fami- n ) and Td(√2ReχU ) ∈ lies {Td(χO)}d and {Td(√2ReχU )}d and (6.1) shows that Td(χO) ∈ L2 d(U + L2 n )∩Bn, kχd Similarly, since χU is standard circular, we have χd n ) = 1, and by [22, UkL2(U + n ) = (1 + 1/d)d/2√d + 1 ∼d→∞pe(d + 1). So the growth rate of O(√d + 1) given by Theorem 6.3 is actually obtained. Of course, the universal constant 46(3e)3√e given by Theorem 6.3 can probably be greatly improved. UkL∞(U + Corollary 3.2] U ∈ L2 d(U + d(O+ kχd Denote by C(U + n ) and L∞(U + n ) is non-nuclear and simple, L∞(U + n ) share with the free group algebras. For example, does C(U + 2 ) ∼= L(F2) = C∗λ(F2)′′. It is an interesting open question whether L∞(U + n ) ⊂ L∞(U + n ) the GNS representation of Au(n) with respect to the Haar state. In [1], it is shown that C(U + n ) is a non-injective II1-factor for all n ∈ N, and L∞(U + n ) is a free group factor for all n ∈ N? It would also be interesting to know what other properties the algebras C(U + n ) always have the MAP? A related question on the von Neumann level is: does L∞(U + n ) always have the Haagerup approximation property? See [20] for the definition of this property. We note that the answer to these last two questions would be "yes" if one could show that the exponentiated length n )) restricted to a unital completely positive map on L∞(U + n ), for each t > 0. Unfortunately this is not true [34, Section 1], and these questions remain open. On the positive side, a direct application of Theorem 5.2 gives the following partial result concerning the MAP. Theorem 6.5. Let Bn ⊂ C(U + ated by the coefficients of the fundamental corepresentation of U + n ∈ N. n ) be the unital non-self-adjoint operator algebra gener- n . Then Bn has the MAP for all function e−ℓt :=Ld≥0 e−dtidL2 n ) ∈ B(L2(U + n ) ⊂ L∞(U + d(U + In [2] the notion of free complexification of a compact matrix quantum group was introduced. Let G = (A, U = [urs]1≤r,s≤n) be a CMQG, and let z = idT be the canonical generator of C(T). The free complexification of G is the CMQG G = ( A, U ), given by U = [zurs]1≤r,s≤n ∈ Mn(C(T) ∗ A), and A = C∗hzurs : 1 ≤ r, s ≤ ni ⊆ C(T)∗ A. The Haar state on A is the restriction to A of the free product of the Haar states on C(T) and A. Using this notion, we can construct more families of random variables which satisfy the hypotheses of Theorems 3.1 and 4.1. Indeed, if G = (A, U ) is any CMQG containing H + n as a quantum subgroup [2], so the array of variables given by U is H + n -bi-invariant, and invariant under free complexification. The special case U + n was considered above. The fact that these variables are not ∗-free follows from an argument similar to [13, Section 4.9]. n as a quantum subgroup, then G = ( A, U ) also contains H + n = fO+ References [1] Banica, T.: Le groupe quantique compact libre U (n). Comm. Math. Phys. 190, 143 -- 172 (1997). [2] Banica, T.: A note on free quantum groups. Ann. Math. Blaise Pascal. 15, 135 -- 146 (2008). [3] Banica, T., Bichon, J., Collins, B.: The hyperoctahedral quantum group. J. Ramanujan Math. Soc. 22, 345 -- 384 (2007). [4] Banica, T., Collins, B.: Integration over compact quantum groups. Publ. Res. Inst. Math. Sci. 43, 277 -- 302 (2007). [5] Banica, T., Curran, S., Speicher, R.: Classification results for easy quantum groups. Pacific J. Math. 247, 1-26 (2010). 28 [6] Banica, T., Curran S., Speicher, R.: De Finetti theorems for easy quantum groups. Ann. Prob. To appear. [7] Banica, T., Curran, S., Speicher, R.: Stochastic aspects of easy quantum groups. Probab. Theory Related Fields. To appear. [8] Banica, T., Collins, B., Zinn-Justin, P.: Spectral analysis of the free orthogonal matrix. Int. Math. Res. Notices (2009). DOI: 10.1093/imrn/rnp054 [9] Banica , T., Speicher, R.: Liberation of orthogonal Lie groups. Adv. Math. 222, 1461 -- 1501 (2009). [10] Bozejko, M.: Remark on Herz-Schur multipliers on free groups. Math. Ann. 258, 11-15 (1981). [11] Choda, M.: Reduced free products of completely positive maps and entropy for free product of automorphisms. Publ. Res. Inst. Math. Sci. 32, 371 -- 382 (1996). [12] Cowling, M., Haagerup, U.: Completely bounded multipliers of the Fourier algebra of a simple Lie group of real rank one. Invent. Math. 96, 507-549 (1989). [13] Curran, S.: Quantum Rotatability. Trans. Amer. Math. Soc. To appear. [14] Curran, S., Speicher, R.: Asymptotic infinitesimal freeness with amalgamation for Haar quantum unitary random matrices. Comm. Math. Phys. To appear. [15] de Canniere, J., Haagerup, U.: Multipliers of the Fourier algebras of some simple Lie Groups and their discrete subgroups. Amer. Journ. Math. 107, 455 -- 500 (1985). [16] de la Salle, M.: Strong Haagerup inequalities with operator coefficients. J. Funct. Anal. 257, 3968-4002 (2009). [17] Edelman, P.: Chain enumeration and non-crossing partitions. Discrete Math. 31, 171 -- 180 (1980). [18] Haagerup, U.: An example of a nonnuclear C∗-algebra, which has the metric approximation property. Invent. Math. 50, 279 -- 293 (1978/79). [19] Jolissaint, P.: K-theory of reduced C∗-algebras and rapidly decreasing functions on groups. K-Theory. 2, 723 -- 735 (1989). [20] Jolissaint, P.: The Haagerup approximation property for finite von Neumann algebras. J. Operator Theory. 48, 549 -- 571 (2002). [21] Kemp, T.: R-diagonal dilation semigroups. Math. Z. 264, 111 -- 136 (2010). [22] Kemp, T. and R. Speicher, Strong Haagerup inequalities for free R-diagonal elements. J. Funct. Anal. 251, 141 -- 173 (2007). [23] Kostler, C., Speicher, R.: A noncommutative de Finetti theorem: invariance under quantum permu- tations is equivalent to freeness with amalgamation. Comm. Math. Phys. 291, 473 -- 490 (2009). [24] Kreweras, G.: Sur les partitions non-croisses d'un cycle. Discrete Math. 1, 333 -- 350 (1972). [25] Lafforgue, V.: A proof of property (RD) for cocompact lattices of SL(3, R) and SL(3, C). J. Lie Theory. 10, 255 -- 267 (2000). [26] Lafforgue, V.: K-th´eorie bivariante pour les alg`ebres de Banach et conjecture de Baum-Connes. Invent. Math. 149, 1 -- 95 (2002). [27] Larsen, F.: Powers of R-diagonal elements. J. Operator Theory. 47, 197 -- 212 (2002). [28] Nica, A., Speicher, R.: Lectures on the combinatorics of free probability. LMS Lecture Notes Series 335, Cam- bridge Univ. Press, (2006). [29] Nica, A., Speicher, R.: R-diagonal pairs - a common approach to Haar unitaries and circular elements. Fields Inst. Commun. 12, 149 -- 188 (1997). [30] Oravecz, F.: On the powers of Voiculescus circular element. Studia Math. 145, 85 -- 95 (2001). [31] Speicher, R.: Multiplicative functions on the lattice of non-crossing partitions and free convolution. Math. Annalen. 298, 611 -- 628 (1994). [32] Timmerman, T.: An invitation to quantum groups and duality. EMS Textbooks in Mathematics, Zurich (2008). [33] Vaes, S., Vergnioux, R.: The boundary of universal discrete quantum groups, exactness, and factoriality. Duke Math. J. 140, 35 -- 84 (2007). [34] Vergnioux, R.: The property of rapid decay for discrete quantum groups. J. Operator Theory. 57, 303 -- 324 (2007). [35] Wang, S.: Free products of compact quantum groups. Comm. Math. Phys. 167, 671 -- 692 (1995). [36] Woronowicz, S.: Compact matrix pseudogroups. Comm. Math. Phys. 111, 613 -- 665 (1987). Michael Brannan: Department of Mathematics and Statistics, Queen's University, 99 University Avenue, Kingston, ON CANADA, K7L 3N6. Email: [email protected] 29
1308.3429
1
1308
2013-08-15T15:04:28
On the Moore-Penrose inverse in $C^*$-algebras
[ "math.OA" ]
In this article, two results regarding the Moore-Penrose inverse in the frame of $C^*$-algebras are considered. In first place, a characterization of the so-called reverse order law is given, which provides a solution of a problem posed by M. Mbekhta. On the other hand, Moore-Penrose hermitian elements, that is $C^*$-algebra elements which coincide with their Moore-Penrose inverse, are introduced and studied. In fact,these elements will be fully characterized both in the Hilbert space and in the $C^*$-algebra setting. Furthermore, it will be proved that an element is normal and Moore-Penrose hermitian if and only if it is a hermitian partial isometry.
math.OA
math
ON THE MOORE-PENROSE INVERSE IN C ∗-ALGEBRAS ENRICO BOASSO Abstract. In this article, two results regarding the Moore-Penrose inverse in the frame of C ∗-algebras are considered. In first place, a characterization of the so-called reverse order law is given, which provides a solution of a problem posed by M. Mbekhta. On the other hand, Moore-Penrose hermitian elements, that is C ∗-algebra elements which coincide with their Moore-Penrose inverse, are intro- duced and studied. In fact, these elements will be fully character- ized both in the Hilbert space and in the C ∗-algebra setting. Fur- thermore, it will be proved that an element is normal and Moore- Penrose hermitian if and only if it is a hermitian partial isometry. 1. Introduction Given an unitary ring A, an element a ∈ A will be called regular, if it has a generalized inverse, also called pseudo-inverse, in A, that is if there exists a′ ∈ A for which a = aa′a. It is clear that in this case aa′ and a′a are idempotents of A. In addition, a generalized inverse a′ of a regular element a ∈ A will be called normalized, if a′ is regular and a is a pseudo-inverse of a′, that is if a = aa′a, a′ = a′aa′. In the presence of an involution ∗ : A → A, it is also possible to enquire if the idempotents aa′ and a′a are self-adjoint, equivalently whether or not (aa′)∗ = aa′, (a′a)∗ = a′a. In this case a′ is called the Moore-Penrose inverse of a, and it is denoted by a†, see [16], where this concept was introduced for matrices, and the related works [10], [11], and [14]. In [10] it was proved that each regular element a in a C ∗-algebra A has a Moore-Penrose inverse, which in addition is unique. Conse- quently, the Moore-Penrose inverse of a regular element a ∈ A is the 1 2 ENRICO BOASSO unique solution x ∈ A to the following set of equations: a = axa, x = xax, (ax)∗ = ax, (xa)∗ = xa. According to the uniqueness of the Moore-Penrose inverse of a regu- lar element a, a∗ also has a Moore-Penrose inverse and (a∗)† = (a†)∗. Furthermore, according to the above equations, if a is a regular ele- ment, then a† also is and (a†)† = a. The so-called reverse order law is one of the most important proper- ties of the Moore-Penrose inverse that have been deeply studied, that is under what condition the equation (ab)† = b†a† holds. In the well-known article [7], T. N. E. Greville proved that the fol- lowing facts are equivalent: i- (ab)† = b†a†, ii- a†abb∗a∗ = bb∗a∗ and bb†a∗ab = a∗ab, iii- a†a commutes with bb∗ and a∗a with bb†, iv- a†abb∗a∗abb† = bb∗a∗a, v- a†ab = b(ab)†ab and bb†a∗ = a∗ab(ab)†, where a and b are two matrices. However, it is worth noticing that the proofs in [7] are also valid in the more general context of C ∗-algebras. The key results of [7] were extended in some works devoted to gener- alized inverses of matrices, see for example [2], [3], and [18]. As regard Hilbert space operators, in [4] R. Bouldin gave a characterization in terms of invariant subspaces, which was refined in [12] and [17]. Ob- serve that the main result in [4], Theorem 3.1, is equivalent to the generalization of Theorem 2 of [7], the above mentioned condition iii, to Hilbert space operators, see Remark 3.2 of [4]. On the other hand, in the work [14] M. Mbekhta studied the reverse order law for generalized inverses in the frame of C ∗-algebras. In fact, given two regular elements a and b in a C ∗-algebra A, it was proved that the following statements are equivalent: i- b′a′ is a generalized inverse of ab, ii- a(pq − qp)b = 0, iii- qp is an idempotent, MOORE-PENROSE INVERSE 3 where a′ and b′ are generalized inverses of a and b respectively, p = bb′ and q = a′a, see Theorem 3.1 of [14]. Naturally, this characteriza- tion remains true in Banach algebras, in fact in a ring. Furthermore, in [5] R. Bouldin proved the same characterization for Banach space operators. In addition, in [14] M. Mbekhta posed the problem of finding neces- sary and sufficient conditions, analogues to the ones of Theorem 3.1 of [14], which ensures that (ab)† = b†a†, for a and b in a C ∗-algebra A. In the work [13] it was claimed that the question of M. Mbekhta in [14] was solved. However, the answer to this problem, Theorem 5 of [13], not only does not provide conditions analogues to the one of Theorem 3.1 in [14], but also it consists in the formulation of the well-known Theorems 1 and 2 of [7] in C ∗-algebras, the above reviewed conditions i, ii, and iii, whose proofs are also valid in C ∗-algebras. The first and main objective of the present work consists in solving the problem posed by M. Mbekhta, that is to give a characterization of the reverse order law for the Moore-Penrose inverse in C ∗-algebras which is analogue to the one of Theorem 3.1 of [14]. Due to the fact that the Moore-Penrose inverse is determined by four equations instead of one, and that it involves not only the product but also the involution, several modifications must be made, however the form of M. Mbekhta's characterization is preserved. What is more, in section 3 four equivalent characterizations with this characteristic will be proved. To this end it will be necessary to reformulate the equations that define the Moore- Penrose inverse of a regular element, which will be done in section 2 following an argument in [16]. On the other hand, given a regular element a in a C ∗-algebra A, according to a general argument, or even as an application of the results of section 3, it is easy to prove that (aa†)† = aa† and (a†a)† = a†a. Now well, since the Moore-Penrose inverse is a particular generalized inverse, it can be thought of a sort of inverse, however, these two identities also suggest that the Moore-Penrose inverse has properties that are similar to the ones of the involution of the algebra. This observation has led to the second objective of this work, namely, the study of the regular elements a ∈ A for which a† = a. These elements will be called Moore- Penrose hermitian, and its basic properties will be studied in section 4. Furthermore, in section 5 Moore-Penrose hermitian elements will be fully characterize both in the Hilbert space and in the C ∗-algebra In addition, it will be also proved that a ∈ A is a normal setting. 4 ENRICO BOASSO Moore-Penrose hermitian element if and only if it is a hermitian partial isometry. This article was written during a research visit to the Abdus Salam International Centre for Theoretical Physics, thanks to the Research Fellowships 2005 Programm. The author wishes to express his indebt- edness to the authorities of the Mathematics Section of the ICTP. In fact, the stimulating atmosphere and the warm hospitality of the men- tioned centre were two extraordinary helps to the research work of the author. This work was also supported by UBACyT and CONICET. 2. Equivalent formulations of the Moore-Penrose inverse Consider a C ∗-algebra A, and a ∈ A a regular element. In this section several equivalent formulations of the equations defining the Moore-Penrose inverse of a will be considered. These formulations will be central in the proof of the characterizations of the next section. In addition, the argument in Proposition 2.1 will follow ideas of Theorem 1 of [16]. Proposition 2.1. Consider a C ∗-algebra A, and two elements in A, a and x. Then, i- the equations a = axa and (ax)∗ = ax are equivalent to a = x∗a∗a. ii- the equations a = axa and (xa)∗ = xa are equivalent to a = aa∗x∗. iii- the equations x = xax and (ax)∗ = ax are equivalent to x = xx∗a∗. iv- the equations x = xax and (xa)∗ = xa are equivalent to x = a∗x∗x. Proof. The third equivalence was proved in Theorem 1 of [16]. The other three statements can be proved in a similar way. As a consequence, the following equivalent conditions are obtained. Proposition 2.2. Consider a C ∗-algebra A and a ∈ A. Then the following statements are equivalent: i- x ∈ A is the Moore-Penrose inverse of a, ii- a = x∗a∗a and x = a∗x∗x, iii- a = aa∗x∗ and x = xx∗a∗. Proof. Is is a consequence of Proposition 2.1 and the equations defining the Moore-Penrose inverse. MOORE-PENROSE INVERSE 5 Remark 2.3. Consider a C ∗-algebra A, a ∈ A a regular element of A, and x = a†. Then, according to Proposition 2.2 and to the fact that a∗ is also regular and (a∗)† = (a†)∗, the following statements are equivalent: i- x ∈ A is the Moore-Penrose inverse of a, ii- a∗ = a∗ax and x∗ = x∗xa, iii- a∗ = xaa∗ and x∗ = axx∗. Next follows the equivalent formulations of the Moore-Penrose in- verse that will be central in the next section. Proposition 2.4. Consider a C ∗-algebra A and a ∈ A. Then the following statements are equivalent: i- x ∈ A is the Moore-Penrose inverse of a, ii- a∗ = xaa∗ and x = xx∗a∗, iii- a = aa∗x∗ and x∗ = axx∗, iv- a∗ = a∗ax and x = a∗x∗x, v- a = x∗a∗a and x∗ = x∗xa. Proof. It is a consequence of Proposition 2.2 and Remark 2.3. 3. The reverse order law for the Moore-Penrose inverse In this section the relationship between the product and the Moore- Penrose inverse will be studied. In fact, four equivalent characteriza- tions of the so-called reverse order law for the Moore-Penrose inverse will be proved. These characterizations are analogue to the one given in Theorem 3.1 of [14] for the generalized inverse of the product of two C ∗-algebra elements. The results of this section provide an answer to a question posed by M. Mbekhta in [14]. Theorem 3.1. Consider a C ∗-algebra A, and two regular elements of A, a and b, such that ab is also regular. Define p = bb†, q = a†a†∗ , r = bb∗ and s = a†a. Then, the following statements are equivalent: (ab)† = b†a†, i- ii- a(pq − qp)b†∗ = 0, and a(rs − sr)b†∗ = 0, iii- spqp = qp, and srsp = sr. Proof. First of all, observe that p, q, r and s are hermitian elements of A. Consider a†, b† and (ab)†, the Moore-Penrose inverses of a, b and ab respectively. According to the third statement of Proposition 2.4, the 6 ENRICO BOASSO following equations hold: a = aa∗a†∗ a†∗ , = aa†a†∗ , b = bb∗b†∗ b†∗ , = bb†b†∗ ab = ab(ab)∗(ab)†∗ , , (ab)†∗ = ab(ab)†(ab)†∗ . Furthermore, note that according again to the third statement of Proposition 2.4, a = as, a†∗ = aq, b = rb†∗ , b†∗ = pb†∗ . Now suppose that (ab)† = b†a†. Then, since (ab)∗ = b∗a∗ and (ab)†∗ = (b†a†)∗ = a†∗ b†∗, it is clear that ab = abb∗a∗a†∗ b†∗ , a†∗ b†∗ = abb†a†a†∗ b†∗ , which is equivalent to asrb†∗ = arsb†∗ , aqpb†∗ = apqb†∗ , which in turn is equivalent to the following identities: a(rs − sr)b†∗ a(pq − qp)b†∗ = 0, = 0. Next suppose that the second statement of the theorem holds. Then, it is clear that a†apqb†∗ b∗ = a†aqpb†∗ b∗, a†arsb†∗ b∗ = a†asrb†∗ b∗. However, according again to the third statement of Proposition 2.4, and to the fact that s = s∗ and p = p∗, these equations can be rewritten as spqp = a†(aa†a†∗ srsp = (a†aa∗)a†∗ )b(b†b†∗ (bb∗b†∗ b∗) = a†a†∗ )b∗ = a∗a†∗ bb† = qp, bb∗ = sr. Finally suppose that the third statement of the theorem holds. Then, since p = p∗ and s = s∗, it is clear that a†abb†a†a†∗ b†∗ a†abb∗a∗a†∗ b†∗ b∗ = a†a†∗ b∗ = a∗a†∗ bb†, bb∗, which implies that (aa†a)bb†a†a†∗ (b†∗ b∗b†∗ ) = (aa†a†∗ )(bb†b†∗ ), (aa†a)bb∗a∗a†∗ (b†∗ b∗b†∗ ) = (aa∗a†∗ )(bb∗b†∗ ). However, according to the third statement of Proposition 2.4 and to the fact that a† and b† are the Moore-Penrose inverse of a and b respectively, the previous equations are equivalent to ab(b†a†)(b†a†)∗ = (b†a†)∗, ab(ab)∗(b†a†)∗ = ab, MOORE-PENROSE INVERSE 7 which, according again to the third statement of Proposition 2.4 implies that (ab)† = b†a†. Note that in a C ∗-algebra, under the same conditions of Theorem 3.1, when instead of generalized inverses Moore-Penrose inverses are con- sidered, the characterization of Theorem 3.1 in [14] determines if b†a† is a normalized generalized inverse of ab. However, in order to char- acterize the reverse order law for the Moore-Penrose inverse, another equation is necessary as well as new elements must be introduced. The next three theorems provide characterizations which are equiv- alent to the one in Theorem 3.1. However, for sake of completeness they are included. Theorem 3.2. Under the same conditions and notations of Theorem 3.1, the following statements are equivalent: (ab)† = b†a†, i- ii- b†(qp − pq)a∗ = 0, and b†(sr − rs)a∗ = 0, iii- pqps = pq, and psrs = rs. Proof. The proof is similar to the one of Theorem 3.1. However, instead of the third statement of Proposition 2.4, the second statement of the mentioned proposition must be used. Theorem 3.3. Under the same conditions and notations of Theorem 3.1, the following statements are equivalent: (ab)† = b†a†, i- ii- b∗(q†p − pq†)a† = 0, and b∗(sr† − r†s)a† = 0, iii- pq†ps = pq†, and psr†s = r†s. Proof. The proof is similar to the one of Theorem 3.1. However, instead of the third statement of Proposition 2.4, the forth statement of the mentioned proposition must be used. In addition, in order to compute q† and r†, Theorem 7 of [10] must be considered. Theorem 3.4. Under the same conditions and notations of Theorem 3.1, the following statements are equivalent: (ab)† = b†a†, i- ii- a†∗(pq† − q†p)b = 0, and a†∗(r†s − sr†)b = 0, iii- spq†p = q†p, and sr†sp = sr†. 8 ENRICO BOASSO Proof. The proof is similar to the one of Theorem 3.1. However, instead of the third statement of Proposition 2.4, the fifth statement of the mentioned proposition must be used. In addition, in order to compute q† and r†, Theorem 7 of [10] must be considered. Remark 3.5. Consider a C ∗-algebra A, and two regular elements of A, a and b, such that ab is also regular. It is well-known that the reverse order law for the product ab is equivalent to the conditions a†abb∗ = bb∗a†a, bb†a∗a = a∗abb†, see for example Theorem 2 of [7], which was proved for matrices but whose proof remains valid in a C ∗-algebra, Proposition 4.4 of [18], Remark 3.2 of [4], Corollary 3.11 of [12], and also Theorem 5 of [13]. However, the above conditions are equivalent to the equalities pq = qp, rs = sr. In fact, the first condition is exactly rs = sr. As regard the second condition, since bb† commutes with a∗a, accord- ing to Theorem 5 of [10], bb† commutes with (a∗a)†, which, according to Theorem 7 of [10], proves that pq = qp. On the other hand, if bb† commutes with a†a†∗, according again to Theorem 5 of [10], bb† commutes with the Moore-Penrose inverse of a†a†∗. In particular, according to Theorem 7 of [10], bb† commutes with a∗a. Furthermore, note that, according to Theorems 5 of [10], and to the fact that p, q, r and s are hermitian elements of A, the above conditions and equalities are equivalent to q†p = pq†, r†s = sr†. Consequently, the second condition of Theorems 3.1 and 3.2 (resp. Theorems 3.3 and 3.4) could have been replaced by the commutativity of p and q, and of r and s (resp. the commutativity of q† and p and of r† and s), however, this has not been done for two reasons. In first place, the conditions ii in the aforesaid Theorems are weakers, but above all, Theorems 3.1 - 3.4 have been presented in a way that they provide a characterization of the reverse order law for the Moore-Penrose inverse analogue to the one of Theorem 3.1 of [14] for generalized inverses. MOORE-PENROSE INVERSE 9 4. Moore-Penrose hermitian elements In first place, the main notion of this and the following section is introduced. Definition 4.1. Consider a C ∗-algebra A. A regular element a ∈ A will be called Moore-Penrose hermitian, if a† = a. Next follow the basic facts regarding the concept just introduced. In the next section Moore-Penrose hermitian elements will be fully characterize. Proposition 4.2. Consider a C ∗-algebra A and an element a ∈ A. Then, the following statements hold: i- Necessary and sufficient for a to be a Moore-Penrose hermitian element is a = a3 and (a2)∗ = a2. ii- If a is a Moore-Penrose hermitian element, then an also is, n ∈ N. iii- The element a is Moore-Penrose hermitian if and only if a∗is. iv- If a is a Moore-Penrose hermitian element, then σ(a) ⊆ {0, −1, 1}, where σ(a) denotes the spectrum of a. Proof. Definition 4.1 and the equations defining the Moore-Penrose in- verse prove the first point, which in turn proves the second. The third point is clear, and the fouth is a consequence of the fact that p = a2 is a hermitian idempotent. 5. Characterizations of Moore-Penrose hermitian elements This section begins with the characterization of Moore-Penrose her- mitian C ∗-algebra elements. In first place, some notation is given. Recall that if A is a C ∗-algebra and a ∈ A, then La : A → A is the map defined by left multiplication by a, that is La(x) = ax, (x ∈ A). In addition, the range and the null space of La will be denoted by R(La) = aA and N(La) = a−1(0) respectively. Theorem 5.1. Consider a C ∗-algebra A. Then, the following state- ments are equivalent: i- a ∈ A is a Moore-Penrose hermitian element, ii- aA = a∗A, a−1(0) = a∗−1(0), A = aA ⊕ a−1(0), and if L = La aA : aA → aA and L = La∗ aA : aA → aA, then L2 = L2 = I, where I denotes the identity map of aA. 10 ENRICO BOASSO Proof. Suppose that a is a Moore-Penrose hermitian element of A, and consider the map La2 : A → A. Since a2 is an idempotent, La2 is a projection defined in A. Consequently, A = R(La2)⊕N(La2 ). However, since a is a Moore-Penrose hermitian element of A, an easy calculation proves that R(La2) = aA and N(La2) = a−1(0). Moreover, since La2 is a projection, it is clear that L2 = I. In addition, according to the third statement of Proposition 4.2, a∗ is a Moore-Penrose hermitian element. Moreover, according to the fifth statement of Proposition 2.4, a = a∗a∗a and a∗ = aaa∗, which implies that aA = a∗A. Furthermore, since according to the third statement of Proposition 2.4, a = aa∗a∗ and a∗ = a∗aa, it follows that a∗−1(0) = a−1(0). However, since a2 is hermitian, L2 = L2 = I. Conversely, if the second statement holds, a straightforward calcu- lation porves that La = L3 a, which clearly implies that a = a3 and (a2)∗ = a2, that is a is a Moore-Penrose hermitian element. a and L2 a∗ = L2 Next follows the characterization of Moore-Penrose hermitian ele- ments in the frame of Hilbert spaces. However, firstly several notions and results need to be reviewed As in the case of a C ∗-algebra element, a Hilbert space operator will be said Moore-Penrose hermitian, if it has a Moore-Penrose inverse T † and T † = T. Recall that if T : H → H is a bounded linear operator defined on the Hilbert space H, then the Moore-Penrose inverse of T is the unique linear and continuous map T † for which the following equations hold: T = T T †T, T † = T †T T †, (T T †)∗ = T T †, (T †T )∗ = T †T. Note that the operator T admits a generalized inverse in A = L(H) if and only if R(T ) is closed, see Theorem 3.8.2 of [9]. However, when the operator T admits a generalized inverse, it can be chosen to be the Moore-Penrose inverse of T in A = L(H), see Theorem 5 of [10]. Moreover, in this case it is unique, and it coincides with the Moore- Penrose inverse of T viewed as an operator defined on L(H), see [10], [11], [14] and [15]. In addition, a bounded linear map which has a generalized inverse will be called a regular operator. Theorem 5.2. Consider a Hilbert space H, and T a regular operator defined on H. Then the following statements are equivalent: i- T is a Moore-Penrose hermitian operator, MOORE-PENROSE INVERSE 11 ii- there exist two orthogonal Hilbert subspaces H1 and H2 such that H = H1 ⊕ H2, T H1 = 0, T (H2) ⊆ H2, and if T2 denotes the restriction of T to H2, then T 2 2 = I2, where I2 denotes the identity map of H2. Proof. Suppose that T † = T and consider the self-adjoint projection P = T †T = T T † = T 2. In particular, the Hilbert space can be pre- sented as the orthogonal direct sum H = N(T 2) ⊕ R(T 2). However, as in the case of a C ∗-algebra, since T is a Moore-Penrose hermitian operator, a straightforward calculation proves that N(T ) = N(T 2) and R(T ) = R(T 2). Define H1 = N(T ), H2 = N(T )⊥ = R(T ), where N(T )⊥ denotes the orthogonal subspace of N(T ). Then, it is clear that T (H2) ⊆ H2, and T 2 2 = I2. Conversely, it the second statement of the theorem holds, it is clear that T 3 = T and T 2 is the orthogonal projection onto H2, in particular T 2 is an hermitian projection. Next normal Moore-Penrose hermitian elements will be considered. However, first of all some preparation is necessary. Given a C ∗-algebra A, the conorm of an element a ∈ A is defined by c(a) = inf {k ax k : dist(x, a−1(0)) = 1, x ∈ A}, see [11] and [14]. It is worth noticing that if a is a regular element, then c(a) = 1 k a† k , see Proposition 1.3 of [14] and Theorem 2 of [11]. Next consider a bounded linear operator T : H → H, where H is a Hilbert space. Then, T is said a partial isometry, if T admits a Moore-Penrose inverse and T † = T ∗, see [15] and Chapter 15 of [8]. In order to keep an analogy with the Hilbert space case, an element a of a C ∗-algebra A will be called a partial isometry, if a is regular and a† = a∗. It is clear that if a ∈ A is a partial isometry, then a∗a is a hermitian idempotent. Conversely, consider a ∈ A such that a∗a is a hermitian idempotent. Then, since each C ∗-algebra has a faithful representation in a Hilbert space, see for example Theorem 7.10 of [6], according to problem 127, Chapter 15, of [8], a straightforward calculation shows that a is a partial isometry. Furthermore, since a is a partial isometry 12 ENRICO BOASSO if and only if a∗ is, then necessary and sufficient for a to be a par- tial isometry is that aa∗ is a hermitian idempotent. See [1] where an equivalent definition of the notion under consideration was considered. In the following proposition a generalization of Corollary 3.2 of [15] will be proved. This result will be central for the characterization of normal Moore-Penrose hermitian elements. Proposition 5.3. Consider a C ∗-algebra A, and a non-zero regular element a ∈ A. Then, necessary and sufficient for a to be a partial isometry is c(a) =k a k= 1. Proof. Let a ∈ A be a non-zero regular element, and consider, ac- cording to Theorem 7.10 of [6], a Hilbert space H and π : A → L(H) a faithful representation of A. It is worth noticing that in this case π(a) ∈ L(H) is regular and π(a)† = π(a†). Suppose that a is a partial isometry. Then π(a) ∈ L(H) also is a partial isometry. Then, according to Corollary 3.2 of [15], k π(a) k = 1. In particular, k a k= 1. Moreover, according to Proposition 1.3 of [14], c(a) = 1 k a∗ k = 1 k a k = 1. Conversely, suppose that a regular element a ∈ A is such that c(a) = k a k= 1. Then, k π(a) k= 1, and according again to Proposition 1.3 of [14], k a† k= 1. On the other hand, since π(a†) = π(a)†, according to Corollaries 2.3 and 3.2 of [15], π(a) is a partial isometry. However, since π : A → L(H) is a faithful representation, a is a partial isometry. Theorem 5.4. Consider a C ∗-algebra A. Then, an element a ∈ A is a normal Moore-Penrose hermitian element if and only if a is a hermitian partial isometry. Proof. Suppose that a ∈ A is a normal Moore-Penrose hermitian ele- ment. Then, according to Theorem 2.9 of [6], to the fourth statement of Proposition 4.2, and to Corollary 1.6 of [14], c(a) = 1. Moreover, since a is a Moore-Penrose hermitian element, according to Proposition 1.3 of [14] or to Theorem 2 of [11], k a k= 1. Consequently, according to Proposition 5.3, a is a partial isometry. However, a = a† = a∗, that is a is hermitian. The converse is clear. MOORE-PENROSE INVERSE 13 Note that if T : H → H is a linear and continuous Hilbert space map, then T is a normal Moore-Penrose hermitian operator if and only if the map T2 in Theorem 5.2 is a hermitian unitary operator. Acknowledgements. The author wishes to express his indebtedness to the referees of this article, for their remarks and suggestions led to an improvement of the present work. References [1] E. Andruchow and G. Corach, Differential geometry of partial isometries and partial unitaries, Illinois J. Math. 48 no. 1 (2004), 97-120. [2] D. T. Barwick and J. D. Gilbert, On generalizations of the reverse order law, SIAM J. Appl. Math. 27 no. 2 (1974), 326-330. [3] D. T. Barwick and J. D. Gilbert, Generalizations of the reverse order law with related results, Linear Algebra Appl. 8 (1974), 345-349. [4] R. Bouldin, The pseudo-inverse of a product, SIAM J. Appl. Math. 24 no. 4 (1973), 489-455. [5] R. Bouldin, Closed range and relative regularity for products, J. Math. Anal. Appl. 61 (1977), 397-403. [6] J. B. Conway, A course in operator theory, Graduate Studies in Mathematics, 21, American Mathematical Society, Providence, RI, 2000. [7] T. N. E. Greville, Note on the generalized inverse of a matrix product, SIAM Rev. 8 no. 4 (1966), 518-521. Erratum, ibid 9 no. 2 (1967), 249. [8] P. R. Halmos, A Hilbert space problem book, Van Nostrand, New York, 1966. [9] R. Harte, Invertibility and singularity for bounded linear operators, Pure and Applied Mahematics, Marcel Dekker Inc., 1988. [10] R. Harte and M. Mbekhta, On generalized inverses in C ∗-algebras, Studia Math. 103 (1992), 71-77. [11] R. Harte and M. Mbekhta, On generalized inverses in C ∗-algebras II, Studia Math. 106 (1993), 129-138. [12] S. Izumino, The product of operators with closed range and an extension of the reverse order law, Tohoku Math. J. 34 (1982), 43-52. [13] A. Khosravi and M. H. Alizadeh, Generalized inverses of products, Int. J. Appl. Math. 10 no. 2 (2002), 141-148. [14] M. Mbekhta, Conorme et inverse g´en´eralis´e dans les C ∗-alg`ebres, Canadian Math. Bull. 35 no. 4 (1992), 515-522. [15] M. Mbekhta, Partial isometries and generalized inverses, Acta Sci. Math. (Szeged) 70 no. 3-4 (2004), 767-781. [16] R. Penrose, A generalized inverse for matrices, Proc. Cambridge Philos. Soc. 51 (1955), 406-413. [17] L. Shijie, The range and pseudo-inverse of a product, Tohoku Math. J. 89 (1987), 89-94. [18] N. Shinozaki and M. Sibuya, The reverse order law (AB)− = B−A−, Linear Algebra Appl. 9 (1974), 29-40. 14 ENRICO BOASSO Enrico Boasso E-mail address: enrico [email protected]
1501.06495
3
1501
2018-11-21T08:15:27
On operator algebras associated with monomial ideals in noncommuting variables
[ "math.OA" ]
We study operator algebras arising from monomial ideals in the ring of polynomials in noncommuting variables, through the apparatus of subproduct systems and C*-correspondences. We provide a full comparison amongst the related operator algebras. For our analysis we isolate a partially defined dynamical system, to which we refer as the {\em quantised dynamics} of the monomial ideal. In addition we revisit several previously considered constructions. These include Matsumoto's subshift C*-algebras, as well as the tensor and the Pimsner algebras associated with dynamical systems or graphs. We sort out the various relations by giving concrete conditions and counterexamples that orientate the operator algebras of our context. It appears that the boundary C*-algebras do not arise as the quotient with the compact operators unconditionally. We establish a dichotomy to this effect by examining the resulting tensor algebras. We identify their boundary representations, we analyse their C*-envelopes, and we give criteria for hyperrigidity. Moreover we completely classify them in terms of the data provided by the monomial ideals. For tensor algebras of C*-correspondences and bounded isomorphisms this is achieved up to the level of local conjugacy (in the sense of Davidson and Roydor) for the quantised dynamics. For tensor algebras of subproduct systems and algebraic isomorphisms this is achieved up to the level of equality of monomial ideals modulo permutations of the variables. In the process we accomplish more in different directions. Most notably we show that tensor algebras form a complete invariant for isomorphic (resp. similar) subproduct systems of homogeneous ideals up to isometric (resp. bounded) isomorphisms. The results on local conjugacy are obtained via an alternative proof of the breakthrough result of Davidson and Katsoulis on piecewise conjugate systems. For our purposes we use appropriate compressions of the Fock representation. We then apply this alternative proof locally for the partially defined quantised dynamics. In this way we avoid the topological graphs machinery and pave the way for further applications. These include operator algebras of dynamical systems over commuting contractions or over row commuting contractions.
math.OA
math
OPERATOR ALGEBRAS OF MONOMIAL IDEALS IN NONCOMMUTING VARIABLES EVGENIOS T.A. KAKARIADIS AND ORR MOSHE SHALIT Abstract. We study operator algebras arising from monomial ideals in the ring of poly- nomials in noncommuting variables, through the apparatus of subproduct systems and C*- correspondences. We provide a full comparison amongst the related operator algebras. For our analysis we isolate a partially defined dynamical system, to which we refer as the quan- tised dynamics of the monomial ideal. In addition we revisit several previously considered constructions. These include Mat- sumoto's subshift C*-algebras, as well as the tensor and the Pimsner algebras associated with dynamical systems or graphs. We sort out the various relations by giving concrete conditions and counterexamples that orientate the operator algebras of our context. It appears that the boundary C*-algebras do not arise as the quotient with the compact operators unconditionally. We establish a dichotomy to this effect by examining the resulting tensor algebras. We identify their boundary representations, we analyse their C*-envelopes, and we give criteria for hyperrigidity. Moreover we completely classify them in terms of the data provided by the monomial ideals. For tensor algebras of C*-correspondences and bounded isomorphisms this is achieved up to the level of local conjugacy (in the sense of Davidson and Roydor) for the quantised dynamics. For tensor algebras of subproduct systems and algebraic isomorphisms this is achieved up to the level of equality of monomial ideals modulo permutations of the variables. In the process we accomplish more in different directions. Most notably we show that tensor algebras form a complete invariant for isomorphic (resp. similar) subproduct systems of homogeneous ideals up to isometric (resp. bounded) isomorphisms. The results on local conjugacy are obtained via an alternative proof of the breakthrough result of Davidson and Katsoulis on piecewise conjugate systems. For our purposes we use appropriate compressions of the Fock representation. We then apply this alternative proof locally for the partially defined quantised dynamics. In this way we avoid the topological graphs machinery and pave the way for further applications. These include operator algebras of dynamical systems over commuting contractions or over row commuting contractions. Contents Introduction 1. 2. Preliminaries 3. Subproduct systems of homogeneous ideals 4. Monomial ideals 5. The C*-correspondence of a monomial ideal 6. C*-algebras associated with a monomial ideal 7. Tensor algebras associated with a monomial ideal 2 7 12 15 24 28 32 2010 Mathematics Subject Classification. 47L65, 47L75, 46L08, 46L55, 46L40, 46L89. Key words and phrases: C*-correspondences, C*-envelope, hyperrigidity, monomial ideals, nonselfadjoint operator algebras, operator algebras on Fock spaces, subproduct systems, tensor algebras, subshifts. 1 8. Encoding via C*-correspondences 9. Encoding via subproduct systems 10. Comparison with other constructions 11. Appendix References 38 55 59 66 74 1. Introduction A contemporary trend in noncommutative geometry is to encode geometrical and topo- logical objects in terms of operator algebras. This program is developed around two main objectives: (a) Use C*-algebras and nonselfadjoint operator algebras as an invariant (resp. a com- plete invariant) for classifying (resp. encoding) the objects. (b) Explore the passage from intrinsic properties of the object into properties of the associated operator algebras. In turn the invariants of the related operator algebras may be used to classify the objects. In this course of study one has to answer two interrelated problems that specify the context: Problem 1. Which are the (desirable) features of the object that determine the asso- ciated operator algebra? Problem 2. What is the (desirable) level of equivalence for classifying objects in the "same" family? The enumeration is in fact redundant; an answer to any of the above problems initiates a new program in finding an answer to the other. There is an extensive literature on the subject and we cannot list all papers. We will only elaborate on the direct connections with results concerning: graphs [46, 82], dynamical systems [21, 22, 23, 44], topological graphs [27], homogeneous ideals [25], complex analytic varieties [26], stochastic matrices [31, 32], and C*-correspondences [1, 43, 62, 64]. The structure under consideration in the current paper is monomial ideals in the ring of polynomials in noncommuting variables. This links to an ongoing effort to develop operator- algebraic geometry in the unit ball [25, 75]. The goal is to study a number of C*- and nonselfadjoint algebras related to one monomial ideal, and study those in conjunction with Questions (a)-(b) and Problems (1)-(2) described above. Motivation. Our approach relies on the interplay between C*-algebras and nonselfadjoint operator algebras related to three classes: C*-correspondences, subshifts, and subproduct systems. All three form active research areas with a rich literature in their own merits. We provide a short presentation on key elements that we use in the current paper. The theory of operator algebras of C*-correspondences has been under considerable de- velopment since their initiation by Pimsner [72]. They have become a central part in the theory of C*-algebras covering a broad number of constructions that generalise the Toeplitz algebra and the Cuntz algebra. Notable examples of Pimsner algebras are C*-algebras as- sociated with topological graphs (hence Cuntz-Krieger algebras, graph algebras, dynamical systems etc.), Hilbert bimodules, and bimodules of unital completely positive maps (see [48] 2 and [15, Examples 4.6.10 -- 4.6.12]). In addition the class of tensor algebras arising from C*- correspondences includes Peters' semicrossed product [71] and Popescu's noncommutative disc algebras [74]. A subshift is the dynamical system obtained by restricting the left shift σ on {1, . . . , d}Z to a closed invariant subspace Λ ⊆ {1, . . . , d}Z . The study of subshifts is a significant branch of topological dynamics, e.g. [55]. Motivated by the Cuntz-Krieger algebras [19], Matsumoto [58] introduced a way to construct a C*-algebra out of a subshift and initiated an in-depth study of these so-called subshift C*-algebras. In [58] Matsumoto constructs C∗(T ) and then concentrates on its quotient by the compacts. In a series of papers (e.g. [59, 60]) he shows how the C*-algebraic structure of this quotient is determined by the topological and combinatorial properties of Λ. Later Carlsen and Matsumoto [18] proposed another construction that differs in general from the one in [58]. The reason for doing so was the discovery of a flaw in one of the earlier papers. They found that many of the theorems in the earlier papers which relied on the flawed one are true if one replaces the original algebra by their new construction, see [18, Introduction]. In turn the construction of [18] did not satisfy a desirable universal property, and around the same time Carlsen [17] proposed a third subshift C*-algebra by using an associated C*-correspondence. In addition the object in [17] is shown to be an invariant for conjugacy of subshifts. Matsumoto [61] introduced later C*-algebras associated with symbolic matrix systems and λ-graph systems that generalise the previous constructions. Subshift algebras form an example of operator algebras related to subproduct systems [80, Section 12]. A subproduct system is a collection of Hilbert (C*- or W*-) correspondences X = {X(s)}s∈S indexed by a semigroup S, which carries an associative family of multiplication operators Us,t : X(s) ⊗ X(t) → X(s + t). They were formally introduced in [80] and [11] as a generalisation of product systems [6], and now they form a basic technical tool in the analysis of noncommutative dynamics, e.g. semigroups of CP-maps and ∗-endomorphisms. In fact they were present in the literature in various guises prior to their formal introduction, see [6, 12, 57, 65]. Simple cases like S = N are still exploitable with numerous successful applications [11, 67, 78, 79, 80, 85]. The operator algebras of a subproduct system are defined concretely by an appropriate compression on the Fock representation of X(1) and quotients by the compacts. The mo- tivation was to approach the unstudied subproduct systems from the well trodden path of C*-correspondences. Soon it was realised that there are substantial differences, yet there is enough structure to be tractable to analysis. Several subsequent works have thus consid- ered operator algebras of subproduct systems per se [25, 31, 38, 52, 86, 87]. There are also works without regard to quantum dynamics or operator algebras, e.g. [37, 83, 84]. Examples include the continuous functions on the unit sphere C(∂Bd), operator algebras of polynomial relations, algebras of analytic functions on homogeneous varieties of the unit ball, as well as algebras of noncommutative analytic functions on homogeneous subvarieties of the noncommutative unit ball. A main obstacle so far in the category of subproduct systems is the lack of a universal property for the concretely defined Toeplitz and Cuntz algebras, that could act in analogy to the theory of C*-correspondences. For example Katsoulis and Kribs [47] use such results of Katsura [50] to obtain that the C*-envelope (in the sense of Arveson [3]) of the tensor algebra of a C*-correspondence is its Cuntz-Pimsner algebra. Such a result is known only in some 3 cases for product systems, e.g. [20], and does not hold for subproduct systems. In fact, until recently all examples indicated that the C*-envelope of the tensor algebra of a subproduct system could be either its Toeplitz algebra or its Cuntz algebra. In the course of writing this paper we were informed by Dor-On and Markiewicz that they discovered a subclass for which there may be several possibilities between the Toeplitz and the Cuntz algebras [32]. The on-going program asks for properties that control such behaviour and includes a number of further questions about: (a) nuclearity, exactness, ideal structure, KMS structure for the C*-algebras, and (b) boundary representations, hyperrigidity, (hyper)reflexivity for the tensor algebras. A unified treatment for all the aforementioned objects cannot be proposed at this moment. Our study on monomial ideals should be seen as one more step in this larger program. Main results. We fix an orthonormal basis {e1, . . . , ed} for Cd and we write eµ = eµ1 ⊗ +. Given a monomial ideal I in C(cid:104)x1, . . . , xd(cid:105) let ··· ⊗ eµn for every word µ = µ1 . . . µn ∈ Fd X = (X(n)) such that X(n) = (Cd)⊗n (cid:9) {eµ xµ ∈ I}, and let the multiplication Un,m be concatenation of words followed by the projection on + xµ /∈ I} of allowable words. Let us write X(n + m). We fix the set Λ∗ = {µ ∈ Fd FX = ⊕n≥0X(n) and let the shift operators {Ti}d i=1 defined by if iν ∈ Λ∗, otherwise. (cid:40) eiν 0 Tieν = The C*-algebras C∗(T ) := C∗(I, Ti i = 1, . . . , d) and C∗(T )/K(FX) are the Toeplitz and the Cuntz algebra of X. The nonselfadjoint subalgebra AX := alg{I, Ti i = 1, . . . , d} of C∗(T ) is the tensor algebra of X. It may happen that C∗(T )/K(FX) remembers remarkably less than the original data. For example, if I = (cid:104)xx, xy(cid:105) (cid:67) C(cid:104)x, y(cid:105) then C∗(T ) = C∗(Tx, Ty) is highly not commutative whereas C∗(T )/K(FX) (cid:39) C(T) (Example 6.9). The family {Ti}d T ∗ µ TµTi = TiT ∗ i=1 satisfies a number of properties; for example it is orthogonal and µiTµi. Hence the linear space E = span{Tia a ∈ A, i = 1, . . . , d}, becomes a C*-correspondence over the commutative unital C*-algebra A = C∗(T ∗ µ Tµ µ ∈ Λ∗). Consequently we obtain the Toeplitz-Pimsner algebra TE, the Cuntz-Pimsner algebra OE, and the tensor algebra T + E in the sense of Muhly and Solel [63]. We give a list of equivalent conditions for the left action to be injective (Proposition 5.8). For the discussion let us mention that the left action is injective if and only if there exists an i0 ∈ {1, . . . , d} such that Λ∗i0 ⊆ Λ∗. Notice that the left action is not injective for the example of I = (cid:104)xx, xy(cid:105) (cid:67) C(cid:104)x, y(cid:105). 4 In the first part of the paper we settle the relation between the appearing operator algebras (Theorem 6.1). We show that C∗(T ) is the J-relative Cuntz-Pimsner algebra of E for the ideal J of A generated by {I − T ∗ µ Tµ µ ∈ Λ∗}, whereas C∗(T )/K(FX) is the A-relative Cuntz-Pimsner algebra of E. In particular there are canonical ∗-epimorphisms TE φ3−→ C∗(T )/K(FX) φ1−→ C∗(T ) φ2−→ OE where: (a) φ1 is injective if and only if I = (0); (b) exactly one of the φ2 and φ3 is injective; and (c) φ3 is injective if and only if the left action is injective. Thus we get that AX ⊆ T + (even as much as one generator) when passing to the quotient. E (Corollary 7.1). Item (c) above explains why we may lose data There are two immediate remarks following these first results. First we get an example of a class of subproduct systems whose operator algebras coincide in a non-trivial way with those of a class of C*-correspondences. This is quite surprising and contradicts to what was thought in the past1, e.g. [87]. Secondly we get that an appropriate Cuntz-type C*-algebra for a subshift is either C∗(T ) or C∗(T )/K(FX) accordingly to the form of the subshift. This is again surprising and should be compared to Matsumoto's work where C∗(T )/K(FX) is considered a priori, e.g. [58]. Our analysis works in parallel with what is known for graphs with sources which were not tractable before Katsura's work [50]. In analogy the new treatment of subshift C*-algebras and the identification of OE that we offer here can work effectively to treat general subshifts, that may not satisfy the property (I) of Matsumoto and Carlsen [17, 18, 58, 59, 60]. In Section 10 we include a review of a number of constructions, including the previous approaches on subshifts, with which we compare our findings. On the one hand, we find that our construction differs from ones that have been considered. On the other hand, in Section 10.4, we show that when E arises from a sofic subshift, OE arises via two familiar constructions: it the graph C*-algebra of the follower set graph of the subshift. In this setting, any structure of I is tracked via the related E. We define an intrinsic dynamical system that gives the encoding by αi : A → A : a (cid:55)→ T ∗ i aTi, for all i = 1, . . . , d. We coin (A, α) ≡ (A, α1, . . . , αd) as the quantised dynamics of I (Section 4.4). Even though the quantised dynamics form a complete conjugacy invariant for the ideal I (Theorem 4.16), some elements are not recognised by E. Following Davidson and Katsoulis [23] and Davidson and Roydor [27] we show that unitary equivalence of the C*-correspondences is equivalent to local conjugacy (which is a weaker condition than conjugacy) of the quantised dynamics. By universality, local conjugate systems have ∗-isomorphic Toeplitz-Pimsner algebras and completely isometric tensor algebras. By the co-universal property of the C*-envelope and 1After this paper was complete, we were made aware of the paper [2], which considers interacting Fock spaces. Interacting Fock spaces provide a framework that encompasses subproduct systems. A major theme in [2] was the problem of representing the creation operators on an interacting Fock space by operators on the full Fock module of a C*-correspondence, similar to the way we represent T in E and in TE. 5 by using [47] this passes to the Cuntz-Pimsner algebras. In addition we show that this holds for the C*-algebras C∗(T ) and C∗(T )/K(FX) as well (Corollary 8.19). E and AX. By using a universal property for Next we turn our focus to the structure of T + E is hyperrigid (Theorem 7.4). If I is finitely OE (Theorem 6.8) and [47] we prove that T + generated then the same holds for q(AX) where q : C∗(T ) → C∗(T )/K(FX) is the quotient map (Proposition 7.5). Thus C∗(T )/K(FX) is the C*-envelope of AX if qAX is completely isometric, and we provide a criterion for this to happen (Theorem 7.6). This criterion becomes a necessary condition for a big class of monomial ideals, which contains subshifts of finite type. Hyperrigidity is further used to obtain a universal property for C∗(T )/K(FX) when I is of finite type (Theorem 7.11). This is quite pleasing as it establishes a strong interaction between C*-algebras and nonselfadjoint operator algebras. We already commented on that T + E is an isometric isomorphic invariant for local conju- gate quantised dynamics. Remarkably the converse also holds, and moreover it holds for continuous isomorphisms (Corollary 8.12). This can be derived by the results of Davidson and Roydor [27] (see Remark 8.5 as well). Nevertheless, we do so by applying the tools we exhibit in the appendix instead of the topological graph language. In particular we provide an alternative proof of [23, Theorem 3.22]. Our remark here is that one can work directly on an appropriate compression of the Fock representation. We then obtain the proof of Corollary 8.12 by applying locally these ideas. On the other hand AX appears to provide a sharper encoding of the monomial ideals. This is not a coincidence, as AX is in a sense the universal operator algebra generated by a row contraction satisfying the relations in I. Therefore its space of (completely contractive) representations is parameterized by the noncommutative variety V (I) = {S = [S1, . . . , Sd] ∈ (B(H)d)1 p(S) = 0 for all p ∈ I}. However, our results are stronger than expected. We show that AX is a complete (algebraic isomorphic) invariant for monomial ideals that coincide up to a permutation of the symbols (Theorem 9.2). In the process we show that this is true for C(cid:104)x1, . . . , xd(cid:105) /I up to graded algebraic isomorphisms. In contrast to what one would expect this behaviour is not met in other settings and it is quite unique to monomial ideals (Remark 9.7). Key to the proof of Theorem 9.2 is Lemma 3.3 that applies in a more general setting. We prove that algebraic isomorphisms between tensor algebras of subproduct systems related to homogeneous ideals can be substituted by graded ones. Then by using results of the second author with Davidson and Ramsey [25], and of Dor-On and Markiewicz [31] we settle the following classification problem (Theorem 3.4): tensor algebras form a complete invariant up to isometric isomorphisms (resp. bounded isomorphisms) for isomorphic (resp. similar) subproduct systems of homogeneous ideals. Open Questions. Our alternative proof of [23, Theorem 3.22] is flexible enough to treat other classes of operator algebras related to classical systems. The key requirement for apply- ing our arguments is for a specific representation on a finite Hilbert space to be completely contractive for the class of the operator algebras under study. We briefly describe some more examples of this phenomenon in Section 11.1. We reserve the full discussion on general classes for a forthcoming project. 6 Some of our intermediate results are not obtained in the full generality of monomial ideals. In Questions 7.10, 8.13, and 8.18 we gather some of the possible considerations. For the conclusion let us add some more that appear to be programs on their own. There are several notions of equivalence for subshifts such as conjugacy and flow equiva- lence. It will be interesting to research similarities and differences with local conjugacy of the quantised dynamics. Since local conjugacy (of the quantised dynamics) requires the same number of symbols we expect it to be stronger than conjugacy (of the subshifts). However it is reasonable to ask whether local conjugate systems imply conjugacy on the corresponding edge shifts. On the other hand flow equivalence is translated into strong shift equivalence of the associated 0-1 matrices [68, 88] (see also [55]). There is also a fourth equivalent relation, that of shift equivalence [88]. It is reasonable to ask for the connections with local conjugacy. The related C*-correspondences that we provide herein suggest a second direction. The first author and Katsoulis [43] have formulated shift equivalence relations for C*-correspon- dences, following the work of Abadie, Eilers and Exel [1], of Muhly and Solel [64], and of Muhly, Pask and Tomforde [62]. A combination of these works shows that these relations provide strong Morita equivalence of the Cuntz-Pimsner algebras. A further project is to study this question for the relative Cuntz-Pimsner algebras C∗(T ). Moreover another direction is to analyse the case of subshift matrix systems and λ-graph systems under the new prism we offer by Theorem 6.1. A similar phenomenon is expected to appear, as [61, Theorem A, equation 1.3] suggests. Again we anticipate that the appropriate C*-algebra depends on the injectivity of a particular C*-correspondence. We remark that the construction in [61, Section 6] does not suffice for this purpose: in the case of subshifts it coincides with q(E) instead of E (see also Remark 5.10). Theorem 9.2 implies that the tensor algebras of the subproduct systems completely encode the subshifts, hence offering a complete invariant for languages. All results include (in fact they are stronger in the case of) two-sided subshifts, and in particular for deterministic automata. It now looks reasonable to move to other directions, for example towards non- deterministic automata or multivariable subshifts. The language of subproduct systems is flexible enough to encode such structures and accommodate such results. Acknowledgements. The authors acknowledge support from London Mathematical Soci- ety (Scheme 4, Grant Ref: 41411). The second author was partially supported by Israel Science Foundation Grant no. 474/12, by EU FP7/2007-2013 Grant no. 321749, and by GIF Grant no. 2297-2282.6/20.1. The authors would like to thank Guy Salomon for the helpful remarks and comments. The first author would like to dedicate this paper to Doukissa Markopoulou. Thank you for your courage; life became darker but the light stays on. 2. Preliminaries 2.1. Dilation theory. The reader should be well acquainted with the representation the- ory of nonselfadjoint operator algebras [13, 70]. For this paper a nonselfadjoint operator algebra will be a norm-closed non-involutive subalgebra of B(H) for a Hilbert space H. The morphisms in the category of operator algebras consist of completely contractive homomor- phisms, which we will refer to as representations. 7 Arveson [3] noticed that a nonselfadjoint operator algebra A may admit a number of com- pletely isometric homomorphisms ιk : A → B(Hk) such that C∗(ιk(A)) is not ∗-isomorphic to C∗(ιk(cid:48)(A)) for k (cid:54)= k(cid:48). Every C∗(ιk(A)) is called a C*-cover of A. This striking event can be realised in the sense that different algebraic relations (even ∗-algebraic relations) on the generators may define the same object. Thus (and in contrast to C*-algebras) ∗-algebraic relations may not define uniquely universal objects. It is dilation theory that facilitates the comparison and the identification of such objects. The interested reader is addressed to [20] for a discussion on nonselfadjoint operator algebras relative to families of homomorphisms. A second question that Arveson [3] posed was whether there is a co-universal C*-cover, with a role similar to the one that the injective envelope plays in ring theory. Hamana [39] showed that this is indeed true: given a nonselfadjoint operator algebra A there is env(A), such that for any other C*-cover C∗(ι(A)) there exists a a C*-cover, denoted by C∗ necessarily unique ∗-epimorphism Φ : C∗(ι(A)) → C∗ env(A) such that Φι(a) = a for all a ∈ A. The kernel of Φ is called the Silov ideal of A in analogy to the Silov boundary of commutative Banach algebras. This result holds for unital operator spaces as well. The first author [40] has provided an elementary proof of Hamana's Theorem. Arveson's motivation was to find an interplay between the C*-envelope and dilation theory. A representation ρ : A → B(K) is a dilation of a representation φ : A → B(H) if H ⊆ K and PHρ(a)H = φ(a) for all a ∈ A. Dritschel and McCullough [33] showed that every completely isometric representation φ of A admits a maximal dilation ρ (in the sense that env(A) (cid:39) C∗(ρ(A)). The all dilations of ρ are trivial). Remarkably they then obtain that C∗ interested reader is addressed to [41] for an overview on the subject. It follows that the maximal representation ρ : A → B(K) (that is, ρ admits only trivial env(A) → B(K) [8]. It remains an open problem whether the converse of this scheme holds. To capture this property, Arveson [9] introduced the notion of hyperrigidity. An operator algebra A ⊆ C∗(A) is hyperrigid if for every faithful ∗-representation π : C∗(A) → B(K) the restriction πA is maximal. Then the same is true for non-injective ∗-representations π of C∗(A), and C∗(A) is the C*-envelope of A. dilations) extends to a unique ∗-representation (cid:101)ρ : C∗ We mention that there is also the notion of the Choquet boundary. In fact Arveson's initial vision was to derive the existence of the C*-envelope through it. He was able to achieve this in the separable case [8], and eventually Davidson and Kennedy [24] solved this long standing open problem. 2.2. C*-correspondences. The reader should be well acquainted with the general theory of Hilbert C*-modules [54, 56]. Throughout the last 15 years there have been many influential authors working on C*-correspondences (e.g. [1, 30, 34, 35, 36, 47, 51, 53, 62, 63, 64, 66, 69] to mention but a few) starting with the seminal work of Pimsner [72]. The notation and definitions have been under several changes and in what follows we will try to highlight the main features by fixing notation for this paper. Mainly we follow the breakthrough work of Katsura [50]. Our suggestions to the reader include [15] for an introduction to the language of C*-correspondences, and [42] for a brief history on the gauge invariant uniqueness theorem. A C*-correspondence AEA is a right Hilbert C*-module E over a C*-algebra A that admits a left action by a ∗-homomorphism φE : A → L(E). We will often write E instead of AEA for simplicity. A C*-correspondence is called regular if it is injective (i.e. φE is injective) 8 and φE(A) sits inside the compacts K(E). It is called non-degenerate if span{φE(a)ξ a ∈ A, ξ ∈ E} = E. It is called full if span{(cid:104)ξ, η(cid:105) ξ, η ∈ E} = A. A pair (π, t) is said to be a representation of E if π : A → B(H) is a ∗-representation and t : E → B(H) is a linear mapping such that t(ξ)∗t(η) = π((cid:104)ξ, η(cid:105)) and π(a)t(ξ) = t(φE(a)ξ) for all ξ, η ∈ E and a ∈ A; then we get t(ξ)π(a) = t(ξa) for free. If π is injective then t is isometric. Every such pair defines a ∗-representation ψt : K(E) → B(H) such that ξ,η) = t(ξ)t(η)∗ for all ξ, η ∈ E [51]. A pair (π, t) is said to admit a gauge action if ψt(ΘE there is a point-norm continuous family {βz}z∈T of ∗-automorphisms on the C*-algebra C∗(π, t) := C∗(π(a), t(ξ) a ∈ A, ξ ∈ E) such that for all a ∈ A, ξ ∈ E, and z ∈ T. Let J ⊆ φ−1 (π, t) of E is called J-covariant if E (K(E)) be an ideal in A. A representation βz(π(a)) = π(a) and βz(t(ξ)) = zt(ξ) π(a) = ψt(φE(a)) for all a ∈ J. If J is Katsura's ideal JE := ker φ⊥ E ∩ φ−1 E (K(E)) then (π, t) is called simply covariant [50]. The covariance of a pair (π, t) is quantified by the ideal (π,t) := {a ∈ A π(a) ∈ ψt(K(E))}. I(cid:48) This follows by the insightful [50, Proposition 3.3] where it is shown that I(cid:48) π is faithful. There are special cases where the ideal I(cid:48) Lemma 2.1. Suppose that K(E) admits a unit e and let (π, t) be a representation of AEA. Then π(a) ∈ ψt(K(E)) if and only if π(a)(I − ψt(e)) = 0 if and only if (I − ψt(e))π(a) = 0. Proof. It follows by the observation that π(a)ψt(e) = ψt(φE(a)e) and that ψt(e)π(a) = ψt(eφE(a)). (π,t) can be described fairly easily. (π,t) ⊆ JE when Hilbert module FE :=(cid:80) Specific representations are given on the full Fock space on E denoted by FE. Let E⊗n+1 = E⊗n ⊗A E (the interior tensor product) for n ≥ 1, and set E⊗0 = A. On the direct sum n≥0 E⊗n let the operators (the sums are taken in the strong operator (cid:88) (cid:88) n topology) n≥0 t∞(ξ) = (ξ) and π∞(a) = τ n+1 φn(a) n (ξ) : E⊗n → E⊗n+1 is such that τ n+1 n≥0 for all ξ ∈ E and a ∈ A, where τ n+1 (ξ)η = ξ⊗η, φ0 = id, and φn = φE ⊗id for n ≥ 1. It is not immediate but if we let q : L(FE) → L(FE)/K(FEJ) be E (K(E)) then (qπ∞, qt∞) is a J-covariant representation [42]. the quotient map for J ⊆ φ−1 The universal C*-algebra TE generated by A and E with respect to the pairs (π, t) is called the Toeplitz-Pimsner algebra of E. The norm closed (non-involutive) subalgebra of TE generated by A and E is called the tensor algebra of E, and will be denoted by T + E . The universal C*-algebra O(J, E) generated by A and E with respect to the J-covariant representations (π, t) is called the J-relative Cuntz-Pimsner algebra of E [63]. In particular n 9 OE := O(JE, E) is called the Cuntz-Pimsner algebra [50]. Katsoulis and Kribs [47] show that C∗ E ) (cid:39) OE for any C*-correspondence E. By universality O(J, E) is the quotient of TE by the ideal generated by the differences env(T + π(a) − ψt(φE(a)) for all a ∈ J, where (π, t) defines a faithful representation of TE. In general A may not embed faithfully in O(J, E). In fact we have that A ⊆ O(J, E) if and only if J ⊆ JE. In particular when J ⊆ JE we obtain a commutative diagram TE OE O(J, E) where all arrows are ∗-epimorphisms that map elements of some index to elements of the same index. These are immediate consequences of the gauge invariant uniqueness theorem as presented in [42] (see also [45]): if J ⊆ JE then (π, t) defines a faithful representation of O(J, E) if and only if (π, t) admits a gauge action, π is injective and I(cid:48) (π,t) = J. In particular if (π, t) admits a gauge action and π is faithful then: (i) C∗(π, t) is ∗-isomorphic to O(J, E) (π,t); (ii) J ⊆ JE by [50, Proposition 3.3]; and (iii) O(J, E) is a C*-cover of T + for J = I(cid:48) E . Let EA and FB be right Hilbert C*-modules and let γ : A → B be a ∗-homomorphism. A right A-module map U ≡ (γ, U ) : EA → FB is called an inner-product map if there exists a linear mapping U∗ : F → E such that (cid:104)U ξ, η(cid:105)F = γ((cid:104)ξ, U∗η(cid:105)E) for all ξ, η ∈ E. When γ : A → B is a ∗-isomorphism then an inner-product map is in L(E, F ) with U∗ ≡ (γ−1, U∗). An adjointable map is called unitary if U∗U = IE and U U∗ = IF , in which case EA is said to be unitarily equivalent to FB. An adjointable map is called invertible if there exists an adjointable (γ−1, V ) : F → E such that V U = IE and U V = IF ; in this case we write V = U−1. If U is an invertible adjointable map then U∗U ∈ L(E) is invertible as well and the map UU−1 defines a unitary adjointable map in L(E, F ). An adjointable mapping (γ, U ) : AEA → BFB is a C*-correspondences mapping if it is a left module map as well. If (γ, U ) is a unitary mapping then AEA and BFB are unitarily equivalent. Inverses of left module maps are again left module maps. Therefore similar C*-correspondences are automatically unitarily equivalent. If AEA and BFB are unitarily equivalent by a (γ, U ) then (π, t) is a representation for F if and only if (πγ, tU ) is a representation for E. Due to the universal property we then obtain In particular the ∗-isomorphism is that the Toeplitz-Pimsner algebras are ∗-isomorphic. completely isometric and restricts to an isomorphism of the tensor algebras. By [47] and the C*-envelope machinery (or by algebraic manipulations and Proposition 2.2 below) we then get that the Cuntz-Pimsner algebras are also ∗-isomorphic. For J-relative Cuntz-Pimsner algebras we have the following. Proposition 2.2. Let (γ, U ) : AEA → BFB be a C*-correspondences unitary mapping. Let J1 ⊆ φ−1 If γ(J1) = J2 then the relative Cuntz-Pimsner algebras O(J1, E) and O(J2, F ) are ∗-isomorphic. E (K(E)) and J2 ⊆ φ−1 F (K(F )). 10 / / # # : : Proof. First notice that (γ, U ) induces an isomorphism ψU : K(E) → K(F ). The proof is identical to [51, Lemma 2.2] (in fact one may use concrete faithful representations of K(E) and K(F ) to achieve this). Let (π, t) be a J2-covariant pair for F . It suffices to show that ψt(φF γ(a)) = ψtU (φE(a)) for all a ∈ J1. Indeed, in this case γ(a) ∈ J2 and the computation πγ(a) = ψt(φF γ(a)) = ψtU (φE(a)) gives that (πγ, tU ) is a J1-covariant pair for E. Universality of the relative Cuntz-Pimsner algebras will then complete the proof. For convenience let γ(a) = b and observe that ψt(φF (b))t(U (ξ)) = t(φF (b)U (ξ)) = tU (φE(a)ξ), ψt(φF (b)ψU (k)) = ψt(φF (b))ψt(ψU (k)) = ψtU (φE(a)k), for all ξ ∈ E. Then by applying t(U (η))∗ for all η ∈ E and by taking limits of finite sums we get that for all k ∈ K(E). However φF (b) ∈ K(F ) = ψU (K(E)) and by considering an approximate identity for K(E) we then have the required equation. 2.3. Subproduct systems. We will require knowledge of subproduct systems as initiated by the second author and Solel [80] for W*-correspondences and by Viselter [86] for C*- correspondences. Here we will restrict our attention to subproduct systems over Z+. We strongly recommend the work of Dor-On and Markiewicz [31] for a series of elegant results. Let E be a non-degenerate C*-correspondence over A. A standard subproduct system X consists of a sequence (X(n), pn) such that: (i) X(0) = A, X(1) = E; (ii) X(n) is an orthocomplemented subcorrespondence of E⊗n, for every n; (iii) X(m + n) ⊆ X(m) ⊗ X(n) for every m, n; and (iv) if pn is the orthogonal projection of E⊗n onto X(n) then the following associativity condition holds pk+m+n(IE⊗k ⊗ pm+n) = pk+m+n(pk+m ⊗ IE⊗n) = pk+m+n. Non-degeneracy implies that X(0) ⊗ X(n) = X(n) for all n ∈ Z+. In the case of W*- correspondences it suffices to assume that the X(n) are closed subspaces (since then they are automatically orthocomplemented [69]). n pn in the full Fock space FE define FX = pFE and the compression operators For p =(cid:80) T (a) = pπ∞(a)FX and T (ξ) = pt∞(ξ)FX for all a ∈ A and ξ ∈ X(n) with n ≥ 1. The associativity condition on the pn shows that FX is co-invariant for t∞ and reducing for π∞. The nonselfadjoint operator algebra AX generated by the T (a) and T (ξ) will be called the tensor algebra of X. Notice here that AX = alg{pπ∞(a), pt∞(ξ) a ∈ A, ξ ∈ FX} = p · T + E . C∗(T ) := C∗(AX) ⊆ B(FX). We will write The C*-algebras C∗(T ) and C∗(T )/K(FX) are the so-called Toeplitz algebra and Cuntz al- gebra, respectively, of the subproduct system X. In parts of the literature they are denoted by TX and OX [25], or T (X) and O(X) [31, 87] (in fact, the Cuntz algebra O(X) is de- fined slightly differently in [87], but for all subproduct systems encountered in this paper, Viselter's definition coincides with C∗(T )/K(FX)). We avoid using this abbreviated notation 11 here so as not to create confusion with TE or OE, with a couple of exceptions required so as to make a connection to the literature. Likewise, the tensor algebra AX is also denoted by T +(X) or T + The subproduct systems X = (X(n), pn) and Y = (Y (n), qn) are called similar if there exists a sequence V = (Vn) of invertible C*-correspondences maps Vn : X(n) → Y (n) such that supn (cid:107)Vn(cid:107) < ∞, supn (cid:107)V −1 X , notation that we will also avoid. n (cid:107) < ∞ and (†) Vm+n(pm+n(ξ ⊗ η)) = qm+n(Vmξ ⊗ Vnη) for all m, n and all ξ ∈ X(m), η ∈ X(n). Then V = (Vn) is said to be a similarity. It follows that V satisfies (†) if and only if Vnpn = qn · V1 ⊗ ··· ⊗ V1 , for all n ∈ Z+. (cid:123)(cid:122) n−times (cid:125) By assumption E⊗n = X(n) ⊕ X(n)⊥, hence V = (Vn) satisfies (†) if and only if ∈ L(X(n) ⊕ X(n)⊥, Y (n) ⊕ Y (n)⊥). Reflexivity of similarity is induced by (V1 ⊗ ··· ⊗ V1)−1 = V −1 The subproduct systems X = (X(n), pn) and Y = (Y (n), qn) are called isomorphic if they are similar by a sequence U = (Un) of unitary C*-correspondences maps Un : X(n) → Y (n). However in this case every Un is a unitary and a compression of the unitary U⊗n , hence we obtain 1 ⊗ ··· ⊗ V −1 . 1 1 ∈ L(X(n) ⊕ X(n)⊥, Y (n) ⊕ Y (n)⊥). (cid:124) (cid:20)Vn 0 (cid:21) ∗ ∗ (cid:20)Un 0 (cid:21) ∗ 0 (cid:125) (cid:125) (cid:124) (cid:124) V1 ⊗ ··· ⊗ V1 = (cid:123)(cid:122) n−times U1 ⊗ ··· ⊗ U1 = (cid:123)(cid:122) n−times for every n. Therefore there is a significant difference between similar and isomorphic sub- product systems. The reader is addressed to [25] for examples that depict this phenomenon. 3. Subproduct systems of homogeneous ideals When E = Cd (over C) then the resulting subproduct systems are characterised by ho- mogeneous ideals of the polynomial ring C(cid:104)x1, . . . , xd(cid:105) in d noncommuting variables [80]. Let {e1, . . . , ed} be an o.n. basis of Cd. For any word w ∈ Fd + with w = w1 . . . wn we write w := n for the length of w. For an X = (X(n), pn) with X(0) = C and X(1) = E = Cd let IX := span{f ∈ C(cid:104)x1, . . . , xd(cid:105) ∃n > 0 such that f (e) ∈ E⊗n (cid:9) X(n)}, with the understanding that f (e) = λwew1 ⊗ ··· ⊗ eww when f (x) = λwxw. (cid:88) w∈Fd + (cid:88) w∈Fd + Then IX is a homogeneous ideal of C(cid:104)x1, . . . , xd(cid:105). On the other hand if I is a homogeneous ideal in C(cid:104)x1, . . . , xd(cid:105) let XI(n) := E⊗n (cid:9) {f (e) f ∈ I}, and pn be the projection of E⊗n onto XI(n). Then XI = (XI(n), pn) is a subproduct system. By [80, Proposition 7.2] we get the connection XIX = X and IXI = I. 12 When I = (0) then we write Ad for the multivariable noncommutative disc algebra. It is accustomed to denote by L = [L1, . . . , Ld] the generators of Ad. When I is the commutator ideal Id, i.e. the ideal generated by LiLj − LjLi, then we will write Ad for the multivariable commutative disc algebra, that is, Ad is the compression of Ad by the closure commutator ideal Id. Given a homogeneous ideal I in C(cid:104)x1, . . . , xd(cid:105) the operators Ti = T (ei) that generate AXI satisfy f (T ) = 0 for all f ∈ I. If S = [S1, . . . , Sd] ∈ B(H d, H) is a row contraction such that f (S) = 0 for all f ∈ I then the mapping Ti (cid:55)→ Si extends to a unital completely contractive representation of AXI in B(H) (for example, combine [80, Theorem 7.5] with [80, Theorem 8.2]). There is an "algebraic" identification which we record for further use. Proposition 3.1. Let XI = (X(n), pn) be the subproduct system associated with a homoge- neous ideal I in C(cid:104)x1, . . . , xd(cid:105). If I denotes the closure of I ≡ {f (L) f ∈ I} in Ad, then AXI is completely isometrically isomorphic to Ad/I. In particular, the isomorphism is given Proof. Let us write X for XI. Let T = [T1, . . . , Td] be the generators of AX. By construction there is a unital completely contractive homomorphism Ad → AX : Li (cid:55)→ Ti for all i = 1, . . . , d. Its kernel contains I and consequently it contains I. Thus we obtain a completely contractive homomorphism by ψ(x + I) = px, where p =(cid:80) n pn, and AXI = p · Ad. ψ : Ad/I → AX : Li + I (cid:55)→ Ti. On the other hand the row contraction (cid:98)L = [L1 + I, . . . , Ld + I] satisfies f ((cid:98)L) = 0 for all f ∈ I. By the remarks preceding the statement there is a unital completely contractive homomorphism σ : AX → Ad/I such that σ(Ti) = Li + I for all i = 1, . . . , d. Thus ψσ is a completely contractive homomorphism sending Ti to itself. Hence it is the identity, and likewise for σψ. Therefore ψ : Ad/I → AX is a completely isometric isomorphism that maps every Li + I to Ti = pLi. Since pLi = pLip, this extends to all polynomials f ∈ Ad. The proof is completed by recalling that the quotient map Ad → Ad/I has closed range. 3.1. The character space. A character of an operator algebra, i.e. an algebraic homo- morphism in C, is automatically completely contractive. When X is associated with a homogeneous ideal I in C(cid:104)x1, . . . , xd(cid:105) then the character space MAX of AX is identified with the set Z(I) = {z ∈ Bd f (z) = 0 for all f ∈ I}, MAX (cid:51) ρ ←→ (ρ(T1), . . . , ρ(Td)) ∈ Z(I). by the bijection Indeed, by Proposition 3.1 we have that a character ρ of AX must vanish on I and hence on I. Since ρ is an algebraic homomorphism, we obtain that f (ρ(T )) = ρ(f (T )) = 0 for all f ∈ I, which shows that ρ(T ) = [ρ(T1), . . . , ρ(Td)] ∈ Z(I). On the other hand if z ∈ Z(I) let ρ(Ti) = zi. Hence we have that ρ(f (T )) := f (ρ(T )) = 0 and by the remarks preceding Proposition 3.1 we have that ρ extends to a contractive homomorphism from AX into C. Given λ ∈ Z(I) we write ρλ ∈ MAX obtained by the above identification. Then for any element t ∈ AX such that t = limn fn(T ) for some polynomials fn ∈ C(cid:104)x1, . . . , xd(cid:105) we get that ρλ(t) = lim n ρλ(fn(T )) = lim n 13 fn(ρλ(T )) = lim n fn(λ). In particular, for every x ∈ Ad the map (cid:98)x : Bd → C given by λ (cid:55)→ ρλ(x) is continuous on Since AX is the norm closure of polynomials in T , the above formula defines ρλ. Furthermore every λ ∈ Z(I) is in Bd, and the latter is the character space of Ad. Even more Bd is the character space of the commutative algebra Ad as well. Since it will be clear by the domain, we will write ρλ for the characters of Ad, Ad, and AX defined by the same λ ∈ Z(I) ⊆ Bd. Lemma 3.2. Let t ∈ AX and x ∈ Ad such that t = x + I given by Proposition 3.1. If λ ∈ Z(I) then we obtain that ρλ(t) = ρλ(x + I) = ρλ(x). Bd and holomorphic in Bd. Proof. The first equality holds because the isomorphism between AX and Ad/I is isometric and preserves polynomials. For the second equality recall that f (λ) = 0 for all f ∈ I. The von Neumann inequality f (λ) ≤ (cid:107)f (L)(cid:107) on Ad [73] implies that ρλ vanishes on I, thus it factors through the quotient. Now we may apply in particular for the commutator ideal Id to obtain that the function λ (cid:55)→ ρλ(x) = ρλ(x + I d) d ) (see [5]) is in Ad. The latter is identified with the norm closure of polynomials in Mult(H 2 and the proof is complete. 3.2. Isomorphisms. For the rest of the section fix the homogeneous ideals I(cid:67)C(cid:104)x1, . . . , xd(cid:105) and J (cid:67) C(cid:104)y1, . . . , yd(cid:48)(cid:105). If φ : AX → AY is an isomorphism then it defines a continuous map φ∗ : MAY → MAX . We say that φ is a vacuum preserving isomorphism if φ∗ρ0 = ρ0 where we write 0 for both zeroes in Bd and in Bd(cid:48). Lemma 3.3. Let X and Y be subproduct systems associated with the homogeneous ideals I (cid:67) C(cid:104)x1, . . . , xd(cid:105) and J (cid:67) C(cid:104)y1, . . . , yd(cid:48)(cid:105). If φ : AX → AY is an algebraic (resp. bounded, isometric) isomorphism then there exists a vacuum preserving algebraic (resp. bounded, isometric) isomorphism φ(cid:48) : AX → AY . that if φ : AX → AY is an isomorphism, then there exists a continuous map (cid:101)F : Bd(cid:48) → Cd Proof. The proof follows verbatim by [25, Proposition 4.7] once we prove an alternative of [25, Lemma 4.4] for algebraic and for bounded isomorphisms. To this end we have to show that is holomorphic on Bd(cid:48) and extends φ∗ : MAY → MAX . Let F = φ∗ be the (continuous) map after the identification MAX (cid:39) Z(I) and MAY (cid:39) Z(J ). For every i = 1, . . . , d, let wi ∈ Ad(cid:48) so that where ψ : Ad(cid:48)/J → AY is the isomorphism obtained by Proposition 3.1, and q is the defining projection of Y . Then the map λ (cid:55)→ φ(Ti)(λ) on Z(J ) extends to the continuous map (cid:98)wi : Bd(cid:48) (cid:55)→ C which is holomorphic on Bd(cid:48) by Lemma 3.2. The required map (cid:101)F : Bd(cid:48) → Cd is then defined by (cid:101)F (λ) = ((cid:99)w1(λ), . . . ,(cid:99)wd(λ)). F : Z(J ) → Z(I) : λ (cid:55)→ (φ(T1)(λ), . . . , φ(Td)(λ)), φ(Ti) = ψ(wi + J ) = qwi, Theorem 3.4. Let X and Y be subproduct systems associated with the homogeneous ideals I (cid:67) C(cid:104)x1, . . . , xd(cid:105) and J (cid:67) C(cid:104)y1, . . . , yd(cid:48)(cid:105). Then: (i) AX and AY are completely isometrically isomorphic if and only if AX and AY are isometrically isomorphic if and only if X and Y are isomorphic; 14 (ii) AX and AY are isomorphic by a completely bounded map if and only if AX and AY are isomorphic as topological algebras if and only if X and Y are similar. Proof. It suffices to show that isometric isomorphism and bounded isomorphism imply iso- morphism and similarity respectively. By Lemma 3.3 we may assume that the isomorphisms are vacuum preserving. Then the isometric case follows by [80, Theorem 9.7]. On the other hand if φ : AX → AY is a vacuum preserving bounded isomorphism, then it is semi-graded by [31, Proposition 6.16]. As a consequence it implements a similarity by [31, Proposition 6.17] and [31, Proposition 6.12] finishes the proof. The converses follow from [31, Proposition 6.12, Corollary 6.13]. 4. Monomial ideals An ideal I of C(cid:104)x1, . . . , xd(cid:105) is called monomial if it is generated by monomials. Monomial ideals in noncommuting variables are homogeneous and may be generated by infinite sets. For example take I = (cid:104)xynx n ≥ 2(cid:105). A set S of monomials is called generating for I if I = (cid:104)S(cid:105). The degree deg S of S is the maximum of the degrees of the monomials in S. We say that I is of type k if it is generated by a set S with deg S = k + 1 and deg S ≤ deg S(cid:48) for any generating set S(cid:48) of I. Then k is finite if and only if I is generated by a finite set S. Every monomial in C(cid:104)x1, . . . , xd(cid:105) corresponds to a finite sequence of the xi. Therefore we identify monomials in C(cid:104)x1, . . . , xd(cid:105) with finite words in Fd + xw ∈ I} the set of forbidden words, and its complement Λ∗ in Fd + the set of allowable words. We write +. We call F := {w ∈ Fd Fl := {µ ∈ F µ ≤ l} and Λ∗ l := {µ ∈ Λ∗ µ ≤ l}. We see that a word is forbidden if and only if it contains a forbidden word. Hence if νµ ∈ Λ∗ then ν ∈ Λ∗ and µ ∈ Λ∗ as well. In general an ideal I of C(cid:104)x1, . . . , xd(cid:105) may not have a basis, that is a smallest generating set. However when I is monomial we can find a basis in the following way. Let I (n) = {f ∈ I deg f = n} = span{xw ∈ I w ∈ Fd +,w = n}. and begin with I (n) such that I (k) = (0) for all k < n and set S(n) = I (n) and S(k) = ∅ for k < n. In I (n+1) let S(n+1) be the set that consists of monomials that cannot be generated by S(n) and continue inductively. Then S = ∪∞ n=0S(n) is a generating set we can get for I. In particular if I is of finite type k then S = ∪k n=0S(n). Without loss of generality, we will only consider monomial ideals which are generated by monomials of degree 2 or more, i.e of type greater than 1. Equivalently, we will consider all the letters of the alphabet as allowed. Indeed, if xi is in the ideal then we may simply throw out of the alphabet the letter i, and the following constructions produce the same outcome. Definition 4.1. We say that a monomial xµ ∈ C(cid:104)x1, . . . , xd(cid:105) \ I is a sink on the left for I (resp. a sink on the right for I) if xixµ ∈ I (resp. xµxi ∈ I) for all i = 1, . . . , d. 4.1. Fock representation. We will write X instead of XI for the subproduct system as- sociated with a monomial ideal I. Let the operators Ti = T (ei) ∈ B(FX). We fix once and for all C∗(T ) := C∗(I, Ti i = 1, . . . , d). 15 By definition the projections pn are diagonal, meaning that (cid:40) eν 0 pneν = ν = n and ν ∈ Λ∗, if otherwise. Therefore X(n) = span{eν ∈ (Cd)⊗n ν ∈ Λ∗,µ = n} and for every µ ∈ Λ∗ we have that Tµeν = eµν 0 if µν ∈ Λ∗, otherwise, with the understanding that Tµe∅ = eµ and T∅ = I. Moreover if µν ∈ Λ∗, otherwise. TµTν = Tµν 0 (cid:40) (cid:40) However it is convenient to write TµTν = Tµν even when Tµν = 0, i.e. Tµν = 0 if and only if µν /∈ Λ∗. We will also write T∅ = I. The operators Tµ satisfy a list of properties which we gather in the following lemma. Lemma 4.2. Let Tµ, Tν for µ, ν ∈ Λ∗ as above. Then the following hold: (i) T ∗ (ii) TνT ∗ (iii) (iv) T ∗ (v) T ∗ (vi) (cid:80)d µ Tµ is an orthogonal projection on span{eν µν ∈ Λ∗}; ν is an orthogonal projection on span{eνµ µ ∈ Λ∗}; if µ = ν, then T ∗ µ Tµ commutes with T ∗ µ Tµ · Ti = Ti · T ∗ i=1 TiT ∗ the rank one operator eν (cid:55)→ eµ equals TµP∅T ∗ ν ; i + P∅ = I where P∅ is the projection on Ce∅; ν Tν, and with TνT ∗ ν ; µiTµi for all i = 1, . . . , d; µ Tν = 0 if and only if µ (cid:54)= ν; (vii) (viii) K(FX) ⊆ C∗(T ). On the other hand such a structure identifies the monomial ideals. Proposition 4.3. Let I be a homogeneous ideal in C(cid:104)x1, . . . , xd(cid:105). Then I is a monomial ideal if and only if the Ti := T (ei) are partial isometries on some eν, and have pairwise orthogonal ranges. j Tieν = (cid:80) true for pn. Fix a word ν ∈ Λ∗ such that ν = n and suppose that pn+1eiν =(cid:80)µ=n+1 λµeµ. Proof. Lemma 4.2 contains the forward implication. For the converse we have to show that pneν is 0 or eν, for all eν ∈ (Cd)⊗n. It is evident that this holds for n = 1 and suppose that it is µ=jµ(cid:48) λµeµ(cid:48) for all j = 1, . . . , d. For j (cid:54)= i we get that λµ = 0 for all Then T ∗ i Tieν is either 0 or eν. Hence we get that λµ = 0 when µ = iµ(cid:48) and µ(cid:48) (cid:54)= ν, µ = jµ(cid:48). Now T ∗ and that λiν = 0 or 1. Therefore pn+1eiν is either 0 or eiν for all i, which shows that pn+1 is diagonal. 4.2. Q-Projections. We will be using the projections generated by the T ∗ i Ti. To this end we introduce the following enumeration. Write all numbers from 0 to 2d−1 by using 2 as a base, but in reverse order. Hence we write [m]2 ≡ [m] = [m1m2 . . . md] so that 2 = [0100 . . . 0], and we define supp[m] := {i = 1, . . . d mi = 1}. 16 Let the (not necessarily one-to-one) assignment [m] (cid:55)→ Q[m] given by (I − T ∗ T ∗ i Ti), (cid:89) mi∈supp[m] i Ti · (cid:89) mi=0 Q[m] ≡ Q[m1...md] := where I ∈ B(FX). For example we write (I − T ∗ d(cid:89) Consequently we obtain(cid:80)2d−1 Q0 = Q[0...0] = i=1 [m]=0 Q[m] = I. The Q[m] are the minimal projections in the C*-subalgebra C∗(T ∗ i Ti) and Q2d−1 = Q[1...1] = d(cid:89) i Ti i = 1, . . . , d) of C∗(T ). T ∗ i Ti. i=1 (I − T ∗ i Ti)) = T ∗ µ Tµ. Lemma 4.4. Let I be a monomial ideal in C(cid:104)x1, . . . , xd(cid:105). If Q[m] are the projections defined above, then µ Tµ · 2d−1(cid:88) 2d−1(cid:88) i Ti i = 1, . . . , d) coincides with I ∈ B(FX) Q[m] · T ∗ µ Tµ [m]=1 ν TνTj ≤ T ∗ j Tj and therefore T ∗ µ Tµ · T ∗ µ Tµ = T ∗ [m]=1 Q[m] = for all ∅ (cid:54)= µ ∈ Λ∗. In particular the unit of C∗(T ∗ if and only if there are no sinks on the left for I. (cid:81)d Proof. For µ = νj ∈ Λ∗ we have that T ∗ i=1(I − T ∗ T ∗ µ Tµ(I − d(cid:89) For the second part of the proof we have that I = (cid:80)2d−1 µ Tµ · 2d−1(cid:88) µ Tµ(I − Q0) = T ∗ i Ti) = 0. Hence we obtain µ Tµ = T ∗ Q[m] = T ∗ j T ∗ [m]=1 i=1 [m]=1 Q[m] if and only if Q0 = 0. Suppose that Q0 = 0 and that there is a word µ ∈ Λ∗ such that iµ /∈ Λ∗ for all i = 1, . . . , d. For this µ we obtain d(cid:89) 0 = Q0eµ = (I − T ∗ i Ti)eµ = eµ which is a contradiction. Conversely, if there are no sinks on the left for I then for every µ ∈ Λ∗ there is an i0 such that i0µ ∈ Λ∗, hence (I − T ∗ i0Ti0)eµ = 0. Commutativity in C∗(T ∗ i Ti i = 1, . . . , d) shows that Q0eµ = 0 for all µ ∈ Λ∗, and the proof is complete. i=1 4.3. Subshifts. Special examples of monomial ideals relate to symbolic dynamics. Let us give a brief description. For the fixed symbol set Σ = {1, . . . , d} let Σ Z let σ : Σ a left subshift if Λ− is a closed subset of Σ called a right subshift if Λ+ is a closed subset of Σ Z a two-sided subshift if Λ is a closed subset of Σ be endowed with the product topology and be the backward shift with σ((xi))k = xk+1. The pair (Λ−, σ−1) is called Z− with σ−1(Λ−) ⊆ Λ−. Respectively (Λ+, σ) is Z+ with σ(Λ+) ⊆ Λ+, and (Λ, σ) is called with σ(Λ) = Λ. Z → Σ Z A word µ = i1 . . . in is said to occur in some (one-sided or two-sided) sequence (xi) if there is an m such that xm = i1, . . . , xm+n−1 = in. Let F be a set of words and let ΛF := {(xi) ∈ Σ Z− no µ ∈ F occurs in (xi) }. 17 All left subshifts arise this way, and a similar construction gives rise to all right or two-sided subshifts. Indeed, if Λ− is a left subshift, then by setting Fk = {µ ∈ F µ ≤ k, µ does not occur in (xi) ∈ Λ−} we see that Fk ⊆ Fk+1 and F =(cid:83) k Fk; then Λ− = ∩kΛFk = ΛF. As in the case of monomial ideals the set of forbidden words admits a basis. We say that Λ− is a left subshift of finite type k + 1 if the longest word in the basis of F has length k. Similar comments hold for the right or two-sided subshifts. The forbidden words form a monomial ideal in C(cid:104)x1, . . . , xd(cid:105) and hence define a subproduct system. The second author and Solel [80, Section 12] give a characterisation of two-sided subshifts by monomial ideals. This generalises to the one-sided shifts as well. Proposition 4.5. Let I = (cid:104)S(cid:105) be a monomial ideal of C(cid:104)x1, . . . , xd(cid:105) with basis S. Then I gives rise to a left subshift (Λ−, σ−1) (resp. right subshift (Λ+, σ), two-sided subshift (Λ, σ)) on the forbidden words S if and only if there are no sinks on the left (resp. on the right, or on either side) for I. In addition I is of finite type k if and only if the subshift is of finite type k. Proof. There are no sinks on the left for I if and only if for all n, k ∈ Z+ we have that for every µ ∈ Λ∗ of length n there exists a ν ∈ Λ∗ of length k such that νµ ∈ Λ∗. The proof then follows in the same way as in [80, Proposition 12.3] with the modification that the required sequences are taken in Σ Z−. An analogous proof follows for the right subshifts. The main point is that the allowable words in a left (resp. right) subshift come from a left-infinite (resp. right-infinite) sequence of symbols. Hence for every length there must be at least one word that we can concatenate on the left (resp. on the right). It is evident that not all monomial ideals come from subshifts. In particular, the allowable words in a subshift related to a set S of forbidden words may be different from the allowable words given directly by the monomial ideal generated by S. We underline this by the following examples. Examples 4.6. Let the symbol space {1, 2}. Then the subproduct system of the ideal (cid:104)x2 1, x1x2(cid:105)(cid:67)C(cid:104)x1, x2(cid:105) differs from the subproduct system related to the right subshift on the forbidden words with basis {11, 12}. Indeed in the latter case we have only one allowable right-infinite word (22 . . . ). 1, x1x2(cid:105)(cid:67) C(cid:104)x1, x2(cid:105) coincides with the subproduct system of the left subshift on the forbidden words with basis {11, 12}. In this case the subshift contains the left-infinite words (. . . 22) and (. . . 21). As a second example, we have that the subproduct system associated with the ideal (cid:104)x1x2, x2x1(cid:105) coincides with the subproduct system of the left subshift, the right subshift and the two-sided subshift on the forbidden words with basis {12, 21}. All subshift spaces consist of two points (. . . 11.11 . . . ) and (. . . 22.22 . . . ). On the other hand the subproduct system associated with the monomial ideal (cid:104)x2 4.4. The quantised dynamics on the allowable words. Given a monomial ideal we isolate a dynamical system originating from the C*-algebra generated by the T ∗ µ Tµ. As we will see later there is a strong connection between these dynamics and a class of nonselfadjoint operator algebras (a further description and study of the quantised dynamics can be found in the preprint of Christopher Barrett with the first author [10], that followed the current work). 18 We fix once and for all the unital C*-subalgebra A := C∗(T ∗ µ Tµ µ ∈ Λ∗) µ Tµ) = T ∗ i and αi(T ∗ of C∗(T ). We define the positive maps αi : A → A such that αi(a) = T ∗ µ Tµ com- µiTµi ∈ A for all µ ∈ Λ∗ then every αi is a ∗-endomorphism mutes with TiT ∗ of A. Our objective is to describe the dynamical system (A, α) ≡ (A, α1, . . . , αd) in terms of the original data. Proposition 4.7. The C*-algebra A = C∗(T ∗ algebra. Proof. We see that A = ∪lAl for the commutative C*-subalgebras µ Tµ µ ≤ l, µ ∈ Λ∗) with l ≥ 1, µ Tµ µ ∈ Λ∗) is a unital commutative AF i aTi. Since T ∗ Al = C∗(T ∗ of C∗(T ). Every Al is finite dimensional since it is generated by a finite family of projections and by definition Al ⊆ Al+1. The C*-algebra A can be characterised by using the allowable words in Λ∗. For l ≥ 0 we define the equivalence relation ∼l on Λ∗ by the rule µ ∼l ν ⇔ {w ∈ Λ∗ l wµ ∈ Λ∗} = {w ∈ Λ∗ l wν ∈ Λ∗}. We stress that this equivalence relation is different than the ones considered by Matsumoto [58] and by Carlsen-Matsumoto [18]. We define the topological space Ωl = Λ∗/ ∼l and write [µ]l for the points in Ωl. Every µ ∈ Λ∗ partitions Λ∗ into the set of the wi ∈ Λ∗ l for which wiµ ∈ Λ∗ and its complement. There is a finite number of such partitions since Λ∗ l is finite. These partitions completely identify single points in Ωl, hence Ωl is a discrete finite space. Furthermore the mapping l ϑ : Ωl+1 → Ωl : [µ]l+1 (cid:55)→ [µ]l is well defined (continuous) and onto. Indeed if µ (cid:54)∼l ν then we may suppose that there exists a word w with w ≤ l such that wµ ∈ Λ∗ and wν /∈ Λ∗; hence [µ]l+1 (cid:54)= [ν]l+1, so ϑ is well defined. Continuity and surjectivity are evident. We can then form the projective limit Ω by the directed sequence ϑo Ω0 ϑo Ω1 ϑo Ω2 . . . Ω for which we obtain the following identification. We write ΩI for Ω when we want to highlight the ideal I which Ω is related to. Proposition 4.8. Let I (cid:67) C(cid:104)x1, . . . , xd(cid:105) be a monomial ideal. With the aforementioned notation, we have that Al (cid:39) C(Ωl) and consequently A (cid:39) C(Ω), where A = ∪lAl for Al = C∗(T ∗ Proof. Recall that Al = C∗(T ∗ l = {µ1, . . . , µn} then the minimal projections arise as products of some T ∗ Λ∗ rest of the I − T ∗ If Tµi and the Tµj . For the sake of clarity suppose that we have the minimal projection l ) is generated by its minimal projections. µ Tµ µ ∈ Λ∗ l ). µi µj µ Tµ µ ∈ Λ∗ n(cid:89) k(cid:89) T ∗ µj Tµj a = (I − T ∗ µj Tµj ) j=1 j=k+1 19 o o o o o and let a point [ν]l ∈ Ωl such that µ1ν, . . . , µkν ∈ Λ∗ and µk+1ν, . . . , µnν /∈ Λ∗. At least one such ν exists if and only if a (cid:54)= 0. Furthermore [ν]l = [ν(cid:48)]l for any ν(cid:48) ∈ Λ∗ with µ1ν(cid:48), . . . , µkν(cid:48) ∈ Λ∗ and µk+1ν(cid:48), . . . , µnν(cid:48) /∈ Λ∗. Therefore the mapping that associates a to χ[ν]l for the ν as above is well defined and one-to-one. Thus it extends to an injective ∗- homomorphism Al → C(Ωl). Now every [ν]l ∈ Ωl is completely characterised by the µj ∈ Λ∗ for which µjν ∈ Λ∗. Hence the ∗-homomorphism is surjective. Finally in order to show that A (cid:39) C(Ω) it suffices to show that the diagram l C(Ωl) Al αϑ id C(Ωl+1) / Al+1 l then µ ∼l ν. On the other hand if wµ /∈ Λ∗ for all w ∈ Λ∗ commutes, where αϑ(f ) = f ϑ. This follows by the definitions since αϑ(χ[ν]l) is the charac- teristic function on the set {[µ]l+1 ∈ Ωl+1 [µ]l = [ν]l}. Proposition 4.9. Let I be a monomial ideal of C(cid:104)x1, . . . , xd(cid:105). If I is of finite type k then Al = Ak for all l ≥ k + 1. Consequently A is generated by a finite number of pairwise orthogonal projections, hence Ω is a finite (discrete) set. Proof. Fix l ≥ k + 1. We have to show that if µ ∼k ν then µ ∼l ν. If wµ /∈ Λ∗ and wν /∈ Λ∗ for all w ∈ Λ∗ l but there is one such that wν ∈ Λ∗ then we exchange the roles of µ and ν. Hence without loss of generality let w ∈ Λ∗ Necessarily we have that ν (cid:54)= ∅, hence wν is a forbidden word of length l+ν ≥ l+1 > k+1. Therefore there are words y, z ∈ Λ∗ and q /∈ Λ∗ with q ≤ k + 1 such that wν = yqz. The word z cannot contain ν, for if z = ν(cid:48)ν then w = yqν(cid:48) /∈ Λ∗ which is a contradiction. Also the word y cannot contain w, for if y = ww(cid:48) then ν = w(cid:48)qz /∈ Λ∗ which is a contradiction. In the above cases we have included the options z = ν or y = w; consequently ν = ν(cid:48)z and w = yw(cid:48) with ν(cid:48), w(cid:48) (cid:54)= ∅. Thus we obtain w(cid:48)ν = qz, hence w(cid:48) +ν = q +z. Since z < ν we get w(cid:48) < q ≤ k + 1 so that w(cid:48) ≤ k. The equation w(cid:48)ν = qz shows that w(cid:48)ν /∈ Λ∗ and since ν ∼k µ we have that w(cid:48)µ /∈ Λ∗. As a consequence wµ = yw(cid:48)µ /∈ Λ∗ which is a contradiction. l such that wµ ∈ Λ∗ and suppose that wν /∈ Λ∗ in order to reach contradiction. The converse of the above proposition does not hold. 2 , xm 2 xm 1 xk 1 , and xn Example 4.10. Consider the ideal I = (cid:104)x1xn 2 x1 n ≥ 1(cid:105), which is not of finite type. The sets Ωl stabilise at l = 2 at the equivalence classes [∅]2, [x1]2 and [x2x1]2, and therefore Ω = {[∅]2, [x1]2, [x2x1]2}. Indeed, the only allowable words are of the form ∅, xm 1 xn 2 , xn 2 xm 2 for k, m, n > 0, and it is straightforward to check that all allowable words fall into one of the three specified ∼l equivalence classes for l ≥ 2. Remark 4.11. In light of the above example, we wish to describe the class of monomial ideals for which A is finite dimensional. In the case where I comes from a two-sided subshift (Λ, σ) then A is finite dimensional if and only if the subshift is sofic. Indeed, recall that (Λ, σ) is called sofic if it is a factor of a subshift of finite type. This is equivalent to having a finite number of equivalence classes with respect to the equivalence relation 1 , xn µ ∼ ν ⇔ {w ∈ Λ∗ µw ∈ Λ∗} = {w ∈ Λ∗ νw ∈ Λ∗}, 20 / /     / e.g. [55, Theorem 3.2.10]. By using the forward shift (Λ, σ−1) and similar ideas one may show that the same holds for the equivalence relation µ ∼ ν ⇔ {w ∈ Λ∗ wµ ∈ Λ∗} = {w ∈ Λ∗ wν ∈ Λ∗}, and eventually obtain that A is finite dimensional if and only if the subshift (Λ, σ) is sofic. In short if A is not finite dimensional then the sets Ωl never stabilize. But the number of equivalence classes with respect to ∼ is at least Ωl, thus the subshift is not sofic. Conversely if A is finite dimensional then the Al stabilise eventually, say at k, thus µ ∼k ν implies that µ ∼l ν for all l ≥ k, and hence that µ ∼ ν. More generally, a monomial ideal I may not come from a subshift. In this case we pass to an augmented system. That is if I (cid:67) C(cid:104)x1, . . . , xd(cid:105) then let (Λ, σ) be the two-sided subshift generated by I(cid:48) = (cid:104)I(cid:105) as an ideal in C(cid:104)x0, x1, . . . , xd(cid:105). Indeed I(cid:48) has no sinks on either side and the allowable words of the augmented shift (Λ, σ) are then of the form 0n1µ10n2µ2 . . . µk0nk for µi ∈ Λ∗ and some ni ∈ Z+. Then the quantised space A related to Λ∗ coincides with the quantised space A of Λ∗. For example the equivalence class of µ10n2 . . . µk0nk coincides with the equivalence class of the last word µ1 appearing on the left, since 0 does not interfere within the forbidden words from I. Therefore A is finite dimensional if and only if the augmented shift (Λ, σ) is sofic. Note that it does not make any difference if we introduce more than one new variables for defining the augmented two-sided subshift. Next we use the identification of A with C(Ω) to get the following translation of each i TiA in A i Ti, then αi defines a unital ∗-homomorphism αi : A → Ai. The ideal Ai is the l = Al ∩ Ai, and the corresponding projective limit Ωi is determined by αi : A → A in a continuous partially defined map ϕi on Ω. If Ai is the ideal T ∗ with unit T ∗ direct limit of Ai ϑ : Ωl+1 → Ωl and the spaces l = {[µ]l ∈ Ωl iµ ∈ Λ∗}. Ωi Hence αi : A → Ai is a unit preserving map from A = C(Ω) into Ai := T ∗ therefore induces a continuous map ϕi : Ωi → Ω. If Λ∗ Ωi l+1 in the following way: if [ν]l+1 ∈ Ωi ϕi([ν]l+1) = [ν(cid:48)]l is such that µ1ν(cid:48), . . . , µkν(cid:48) ∈ Λ∗ and µk+1ν(cid:48), . . . , µnν(cid:48) /∈ Λ∗. l+1 is such that µ1iν, . . . , µkiν ∈ Λ∗ and µk+1iν, . . . , µniν /∈ Λ∗, then i TiA = C(Ωi), and l = {µ1, . . . , µn} then ϕi is defined on Recall here that the universal property of the projective limit implies that this is enough to l = {µ1, . . . , µn}. Then describe ϕi. Indeed let the identification π : A → C(Ω) and fix µ ∈ Λ∗ we get that π(T ∗ µ Tµ) = χ{[w]l µw∈Λ∗} and παi(T ∗ µ Tµ) = χ{[w]l µiw∈Λ∗}. Now for every [ν]l+1 ∈ Ωi µk+1iν, . . . , µniν /∈ Λ∗. Then we obtain l+1 we get a split into some µ1iν, . . . , µkiν ∈ Λ∗ and the rest παi(T ∗ µ Tµ)([ν]l+1) = if µiν ∈ Λ∗, otherwise. (cid:40) 1 0 21 Hence παi(T ∗ [ν(cid:48)]l with µ1ν(cid:48), . . . , µkν(cid:48) ∈ Λ∗ and µk+1ν(cid:48), . . . , µnν(cid:48) /∈ Λ∗, thus µ Tµ)([ν]l+1) = 1 if and only if µ ∈ {µ1, . . . , µk}. On the other hand ϕi([ν]l+1) = π(T ∗ µ Tµ)ϕi([ν]l+1) = if µν(cid:48) ∈ Λ∗, otherwise. (cid:40) 1 0 µ Tµ)ϕi([ν]l+1) = 1 if and only if µ ∈ {µ1, . . . , µk}. Thus π(T ∗ Definition 4.12. Let I be a monomial ideal of C(cid:104)x1, . . . , xd(cid:105). We call the (A, α) ≡ (A, α1, . . . , αd), or alternatively the (Ω, ϕ) ≡ (Ω, ϕ1, . . . , ϕd), the quantised dynamics on the allowable words Λ∗ of I. We chose the above terminology to prevent confusion between the above dynamical system and the dynamical system determined by the shift, when I gives rise to a subshift. Remark 4.13. When the ideal I is of finite type 1 then there is a simple description of the quantised dynamical system. In this case Ω is a finite space by Proposition 4.9. In particular we have that Ω = {[∅]1, [i1]1, . . . , [ik]1}, where it may happen or not that [∅]1 contains some of the variables. Every set Ωi is the set of equivalence classes of variables (and the empty word) to which one may append i on the left to obtain a legal word. On Ωi the map ϕi is defined to be the constant function with value [i]1. To see this, recall : C(Ω) → C(Ωi). But A is generated that ϕi : Ωi → Ω is the dual of the map αi = adT ∗ by T ∗ j Tj for j = 1, . . . , d. Hence by identifying A with C(Ω) we see that it is generated by the functions χΩj for j = 1, . . . , d. Thus we need only check that if we define ϕi to be constant [i]1 on Ωi then αi(χΩj ) = χΩj ϕi. Since we are in a subshift of type 1, we have the identifications i αi(χΩj ) = T ∗ i T ∗ j TjTi = T ∗ i Ti = 1Ωi 0 if ji ∈ Λ∗, if ji /∈ Λ∗. On the other hand we have ji ∈ Λ∗ if and only if [i]1 ∈ Ωj, so (cid:40) (cid:40) χΩj ϕi = 1Ωi 0 if ji ∈ Λ∗, if ji /∈ Λ∗. This can be pictured as a graph on the points [∅]1, [i1]1, . . . , [ik]1, with an edge labeled i from every p ∈ Ωi to [i]1 (including, maybe, loops). The language can be completely read from this graph: there is an edge labeled i from [j]1 to [i]1 if and only if the word ij is a legal word. Remark 4.14. Similarly to the previous remark, whenever the augmented system of an ideal is sofic, then one may picture the quantised dynamics as a finite labeled graph. Two dynamical systems (A, α1, . . . , αd) and (B, β1, . . . , βd) are said to be conjugate if (after perhaps reordering the maps) there is a ∗-isomorphism γ : A → B such that γαi = βiγ for all i. Equivalently, (after perhaps reordering the maps) there is a homeomorphism γs : ΩJ → ΩI mapping ΩiJ into ΩiI such that γsϕJ ,iΩiJ = ϕI,iγsΩiJ for all i. 2}. This ideal corre- Example 4.15. Let J be the ideal in C(cid:104)x1, x2(cid:105) generated by {x2 sponds to the two-sided subshift ΛJ on two symbols {1, 2} with illegal words F = {11, 22}. This subshift consists of two points 1, x2 ΛJ = {(. . . 121.2121 . . . ), (. . . 212.1212 . . . )}, 22 and the shift just permutes these two points. In this case, the quantised dynamics attain the following description. The space ΩJ is a three point set {0, 1, 2} where 1 corresponds to the equivalence class of all words beginning with 1, 2 likewise, and 0 corresponds to the equivalence class of the empty word. Then Ω1J = {0, 2} and Ω2J = {0, 1}. The map ϕJ ,1 : Ω1J → ΩJ is the constant function 1, and ϕJ ,2 : Ω2J → ΩJ is the constant function 2. Now let I be the ideal generated by {x1x2, x2x1}. We find that ΩI = {0, 1, 2} as above. We also have that Ω1I = {0, 1} and Ω2I = {0, 2}, that ϕI,1 : Ω1I → ΩI is the constant function 1, and the map ϕI,2 : Ω2I → ΩI is the constant function 2. The form of the partial dynamical systems is depicted in the following graphs 1 1 0 1 2 graph for J 2 1 0 2 5 2 1 1 2 2 graph for I where 0 = [∅]1, 1 = [1]1, 2 = [2]1, and the solid arrows represent ϕ1 and the broken arrows represent ϕ2. We have here two very simple dynamical systems: each one is a three point set with a pair of distinct constant valued maps acting on it. However, they cannot be conjugate, since these maps are partially defined constant maps. Theorem 4.16. The quantised dynamical system is a complete invariant of the monomial ideal: if the quantised dynamical systems of two monomial ideals I and J are conjugate, then I and J are the same modulo a permutation of the variables. In particular, if the quantised dynamics of two subshifts on the same set of symbols are conjugate, then the subshifts are conjugate. Proof. Note that conjugacy determines the number of variables. Letting αµ = adT ∗ have that µ = µ1 . . . µk is a forbidden word if and only if T ∗ Since this determines the forbidden words, this determines the ideal, as well. µ , we µ Tµ = αµ(I) = αµk ··· αµ1(I) = 0. Note that the converse of the second assertion in the proposition fails, since the number of symbols is invariant under conjugacy of the quantised dynamics but not under conjugacy of the shifts. Example 4.17. Let J be as in Example 4.15. Let K be the ideal in the polynomial algebra C(cid:104)x1, . . . , x4(cid:105) generated by This corresponds to the two-sided subshift ΛK on four symbols {1, 2, 3, 4} where a word is legal if and only if it contains no two consecutive even symbols and no two consecutive odd symbols. The subshift ΛK differs from ΛJ , since it has uncountably many points. On the other hand, the space ΩK is also a three point set {0, 1, 2}, where 0 is the equivalence class of ∅, 1 is the equivalence class of all words beginning with an odd symbol, and similarly 2. Here 23  x2 1 x3x1  . x1x3 x2 3 x2 2 x4x2 x2x4 x2 4         5 u u Y Y E E too we have ΩiK = {0, 2} for i odd and ΩiK = {0, 1} for i even; the map ϕK,i is identically 1 (resp. 2), if i is odd (resp. even). We see that the quantized dynamics (ΩJ , ϕJ ,1, ϕJ ,2) and (ΩK, ϕK,1, . . . , ϕK,4) are given by the same maps on the same space. This may seem strange in light of Theorem 4.16 but the difference lies on that in the second dynamical system each map is repeated twice. Remark 4.18. In the previous discussion we enforced I to be included in the C*-algebra A. We could as well restrict our attention to the C*-algebra A0 := C∗(T ∗ µ Tµ ∅ (cid:54)= µ ∈ Λ∗) ⊆ C∗(T ). Notice here that αi(A0) ⊆ A0 and that A0 is also unital by Lemma 4.4. Therefore similar conclusions can be derived for A0 which can be identified with a compact Hausdorff space Ω0. The reason of considering this setting is explained in Remark 5.2 that will follow. 5. The C*-correspondence of a monomial ideal The C*-algebra C∗(T ) is trivially a C*-correspondence over itself. We want to isolate a C*-correspondence AEA generated by the Ti inside C∗(T ) with the inherited inner prod- uct. There are (at least) three equivalent identifications. The use of the symbol A is not coincidental. 5.1. Concrete construction. First we remark that all (cid:104)Ti, Ti(cid:105) = T ∗ i Ti must be in the A we are looking for. Now equation T ∗ jiTji must be in the module E we are looking for (with the understanding that Tji = 0 if ji /∈ Λ∗). Consequently the inner product (Ti · T ∗ jiTji implies that all Ti · T ∗ j Tj · Ti = Ti · T ∗ must be in A. Inductively E contains all "vectors" Ti · T ∗ µ Tµ, for µ ∈ Λ∗ and i = 1, . . . , d. T ∗ Definition 5.1. Let I be a monomial ideal of C(cid:104)x1, . . . , xd(cid:105) and let Λ∗ be the set of allowable words of I. Then the linear space µ Tµ and A contains all projections jiTj · TiT ∗ i Ti · T ∗ jiTji)∗Ti · T ∗ jiTji = T ∗ jiTjTiT ∗ jiTji = T ∗ jiTji = T ∗ jiTji E := span{Tia a ∈ A, i = 1, . . . , d} becomes a C*-correspondence over the C*-algebra A = C∗(T ∗ µ Tµ µ ∈ Λ∗) by defining the operations a · ξ · b := aξb and (cid:104)ξ, η(cid:105) = ξ∗η, for all a, b ∈ A and ξ, η ∈ E, inside C∗(T ). We refer to AEA as the C*-correspondence associated with the monomial ideal I. We write 1 ≡ 1A = IFX ≡ I for the unit of A. Remark 5.2. Alternatively we consider the minimal C*-correspondence generated by the Ti. In this case we get the linear space E0 := span{Ti, Tia a ∈ A0, i = 1, . . . , d} which becomes a C*-correspondence over the C*-algebra A0 := C∗(T ∗ µ Tµ ∅ (cid:54)= µ ∈ Λ∗) 24 admits the unit(cid:80)2d−1 in the same way. We include the subscript 0 in the notation of E0 to make a distinction with E as Hilbert modules, even though E0 and E coincide as linear spaces. The C*-algebra A0 [m]=1 Q[m] by Lemma 4.4, but it may not contain I ∈ B(FX). Consequently the left action on E0 need not be unital, even though the right action is, since we have that A number of the arguments we will present henceforth hold by substituting 1 with(cid:80)2d−1 [m]=1 [m]=1 i Ti = Ti. Q[m] = TiT ∗ i Ti Q[m] = TiT ∗ [m]=1 Q[m]. Ti · 2d−1(cid:88) 2d−1(cid:88) We will keep track of when this happens. 5.2. Direct sum construction. For every i = 1, . . . , d, let Ei = Ai := T ∗ space and define (cid:104)ξi, ηi(cid:105) = ξ∗ i ηi and ξi · a = ξia, for all a ∈ A and ξi, ηi ∈ Ei, with the operations taking place inside C∗(T ). Then Ei = (cid:104)δi(cid:105) for δi = T ∗ A-module and i Ti ∈ A as a Hilbert i TiA as a vector Recall the ∗-homomorphism αi : A → A such that αi(a) = T ∗ C*-correspondence over the commutative A by defining (cid:104)δi · a, δi · b(cid:105) = a∗T ∗ i Tib = T ∗ i Tia∗b. φi(a)(δi · b) = δi · αi(a)b. Since αi(a)T ∗ i Ti = αi(a) the computation ΘEi δi·αi(a),δi (δi · b) = δi · αi(a)T ∗ i Tib i aTi. Then Ei becomes a gives that φi(a) = ΘEi δi·αi(a),δi . Furthermore we get that i TiT ∗ (cid:104)δi · T ∗ i Tib = (cid:104)δi, δi · b(cid:105) i Ti, thus every φi is unital. i Ti, δi · b(cid:105) = T ∗ for all b ∈ A. Therefore we have that δi = δi · T ∗ Proposition 5.3. Let AEA be the C*-correspondence of a monomial ideal I (cid:67) C(cid:104)x1, . . . , xd(cid:105) and let A(Ei)A be as above. Then E is unitarily equivalent to the sum C*-correspondence of the Ei over A. Proof. The map U (1· a1, . . . , 1· ad) := T1a1 +··· + Tdad defines the required equivalence. 5.3. Topological graph construction. Katsura introduced a construction that generalises both graph algebras and dynamical systems [49]. A topological graph is a tuple (Υ0, Υ1, r, s) such that Υ0 and Υ1 are locally compact Hausdorff spaces, r : Υ1 → Υ0 is a continuous proper map and s : Υ1 → Υ0 is a local homeomorphism. For ξ ∈ C0(Υ1) let the map (cid:104)ξ, ξ(cid:105) : Υ0 → [0,∞] given by (cid:104)ξ, ξ(cid:105) (v) =(cid:80) becomes a C*-correspondence over C0(Υ0) by defining the inner product Cd(Υ1) := {ξ ∈ C0(Υ1) (cid:104)ξ, ξ(cid:105) ∈ C0(Υ0)} y∈s−1(v) ξ(y)2. The linear space (cid:88) ξ(y)η(y) (cid:104)ξ, η(cid:105) (v) = y∈s−1(v) 25 and the module actions for all a ∈ C0(Υ0) and ξ, η ∈ Cd(Υ1). (ξ · a)(y) = ξ(y)a(s(y)) and (a · ξ)(y) = a(r(y))ξ(y) Suppose that AEA is the C*-correspondence associated with a monomial ideal. Then A is identified with C(Ω) for some compact Hausdorff space Ω and every αi gives rise to a continuous mapping ϕi : Ωi → Ω, where Ωi is the clopen set induced by the projection T ∗ i Ti under the identification A (cid:39) C(Ω). Let χi be the characteristic function on Ωi. Let Υ1 i = Ωi i for all i = 1, . . . , d. Thus a y ∈ Υ1 is determined and define Υ1 be the disjoint union of the Υ1 i . Let Υ0 = Ω and define the continuous maps s, r : Υ1 → Υ0 by a tuple (i, yi) with yi ∈ Υ1 by s(i, yi) = yi and r(i, yi) = ϕi(yi). Then s is a local homeomorphism. Proposition 5.4. Let AEA be the C*-correspondence of a monomial ideal I (cid:67) C(cid:104)x1, . . . , xd(cid:105). Then E is unitarily equivalent to the C*-correspondence of the topological graph (Υ0, Υ1, r, s) as constructed above. Proof. Every ξ ∈ Cd(Υ1) can be decomposed into the sum of the ξi = χiξ for i = 1, . . . , d. Each ξi is a continuous function in Ai = C(Ωi) ⊆ C(Υ0). Hence the map U (ξ) := T1ξ1 + ··· + Tdξd defines the required equivalence. For surjectivity recall that Tia = Ti(T ∗ i Ti)a with i Tia ∈ C(Υ1 T ∗ i ). 5.4. Analysis of the C*-correspondence. We proceed to the properties of the C*-cor- respondence AEA. Proposition 5.5. Let AEA be the C*-correspondence of a monomial ideal I (cid:67) C(cid:104)x1, . . . , xd(cid:105). Then the left action is non-degenerate and by the compacts. If, in addition, there are no sinks on the left for I then AEA is full. Proof. By definition the left action φE is unital, hence E is non-degenerate. The computa- tion d(cid:88) µ Tµ) =(cid:80)d Tj T ∗ µj Tµj ,Tj ΘE j=1 d(cid:88) j=1 (Tia) = TjT ∗ µjTµjT ∗ j Tia = TiT ∗ µiTµia = T ∗ µ TµTia, shows that φE(T ∗ j=1 ΘE Tj T ∗ µj Tµj ,Tj , hence φE(A) ⊆ K(E). a =(cid:80)d Lemma 4.4. By using the minimal projections Q[m] in C∗(T ∗ If I has no sinks on the left, then the joint projection of the T ∗ i Ti equals I ∈ B(FX) by i Ti i = 1, . . . , d) we can write i Tiai for suitable ai ∈ A. Since ai = (cid:104)δi, δiai(cid:105) we get that ai ∈ (cid:104)E, E(cid:105) i=1 ai with ai = T ∗ for all i = 1, . . . , d which completes the proof. Remark 5.6. In particular we have that E0 of Remark 5.2 is full and the action is by the compacts. It will be essential to identify the kernel of the left action on E. By definition we have that ker φE = {a ∈ A aTi = 0 for all i = 1, . . . , d}. Lemma 5.7. If for each i = 1, . . . , d there is a µi ∈ Λ∗ such that µii /∈ Λ∗, then P∅ = µ1Tµ1 ··· T ∗ T ∗ Tµd. In particular we get that P∅ ∈ A. µd 26 (i) P∅ ∈ A; (ii) ker φE = (cid:104)P∅(cid:105); (iii) ker φE = C · P∅; (iv) ker φE (cid:54)= (0); (v) (vi) (vii) JE := ker φ⊥ (viii) 1 /∈ JE. for every i = 1, . . . , d there is a µi ∈ Λ∗ such that µii /∈ Λ∗; for every i = 1, . . . , d there is a µi ∈ Fd + such that xµi /∈ I and xµixi ∈ I; E ∩ φ−1(K(E)) = (cid:104)1 − P∅(cid:105) = A(1 − P∅); (cid:11) for any tuple of words (µ1, . . . , µd) µ1Tµ1 ··· T ∗ µd Tµd If these conditions hold then ker φE =(cid:10)T ∗ Proof. We will be using that I − P∅ =(cid:80)d such that µii /∈ Λ∗ for all i = 1, . . . , d. i . Consequently if a ∈ ker φE then we get µ Tµ) = 0 for all µ ∈ Λ∗; hence P∅T ∗ µ Tµ = P∅ It is immediate that P∅Ti = 0 for all i = 1, . . . , d. Thus if P∅ ∈ A then µ Tµ = P∅ for all µ ∈ Λ∗, and that the i=1 TiT ∗ that a(I − P∅) = 0. Also observe that P∅(I − T ∗ for all µ ∈ Λ∗. [(i) ⇒ (ii)]: P∅ ∈ ker φE. Now if a ∈ ker φE, then a(I − P∅) = 0, which shows that a = aP∅ ∈ (cid:104)P∅(cid:105). [(ii) ⇒ (iii)]: This follows from the facts that P∅T ∗ projections T ∗ µ Tµ generate the unital C*-algebra A. [(iii) ⇒ (iv)]: Immediate, since P∅ (cid:54)= 0. [(iv) ⇒ (v)]: Recall that A is an AF algebra by Proposition 4.7. Hence ker φE = ∪l ker φE ∩ Al. Suppose that Λ∗ l = n and let the orthogonal minimal projections {Qj} that generate Al µsTµs for some enumeration {µ1, . . . , µn} of Λ∗ l . Fix a non-zero element a ∈ Al ∩ ker φE; then a = aP∅. Since P∅(I − T ∗ (cid:88) µ Tµ) = 0 we obtain that P∅Qj = δ0,jP∅ and therefore with the understanding that Q0 = (cid:81)n s=1 T ∗ As a (cid:54)= 0, it follows that P∅ ∈ Ak. Thus we can write P∅ =(cid:80) λjQjP∅ = λ0P∅. a = aP∅ = Qj are orthogonal and thus for i (cid:54)= 0 we have j j λ(cid:48) µ1Tµ1 ··· T ∗ µnTµn ≥ P∅ for any µ1, . . . , µn ∈ Λ∗. Suppose that the Proof. First note that T ∗ µi are as in the statement and let eν ∈ Λ∗ with ν (cid:54)= ∅. We can write ν = iν(cid:48) for some i ∈ {1, . . . , d} and ν(cid:48) ∈ Λ∗. Then we get that T ∗ Tµieν = 0 since µii /∈ Λ∗ by assumption. Tµdeν = 0. Since ν (cid:54)= ∅ was arbitrary, the proof is Therefore we obtain that T ∗ complete. Proposition 5.8. Let AEA be the C*-correspondence of a monomial ideal I (cid:67) C(cid:104)x1, . . . , xd(cid:105). The following are equivalent: µ1Tµ1 ··· T ∗ µd µi 0 = P∅Qi = λ(cid:48) iQi. jQj. However the projections 0 = 1 and so P∅ =(cid:81)n 0Q0 which implies that λ(cid:48) Consequently we have that P∅ = λ(cid:48) As P∅ei = 0 for i ∈ {1, . . . , d} there is a µs ∈ Λ∗ [(v) ⇒ (i)]: This follows by Lemma 5.7. [(v) ⇔ (vi), (ii) ⇔ (vii), (iv) ⇔ (viii)]: These are immediate (recall that the left action is by the compacts, hence JE = ker φ⊥ E). Remark 5.9. Proposition 5.8 reads the same for the E0 of Remark 5.2. Indeed if 1 is the identity of A0 then a(1 − T ∗ µ Tµ) and aP∅ = P∅a for all a ∈ A0 and µ ∈ Λ∗. k such that µsi /∈ Λ∗. µ Tµ) = a(I − T ∗ s=1 T ∗ µsTµs. 27 First of all (cid:80)d Remark 5.10. There is another C*-correspondence we could relate to a monomial ideal. Let q : C∗(T ) → C∗(T )/K(FX) be the quotient map, and form the C*-correspondence q(E) over the C*-algebra q(A) by using the same ∗-algebraic relations with E. The left action on q(E) is again by the compacts since this is verified by ∗-algebraic relations. However there are some substantial differences between q(E) and E. i=1 q(Ti)q(Ti)∗ = I and therefore q(E) is always injective. Furthermore C∗(T )/K(FX) is the Cuntz-Pimsner algebra of q(E) given by the covariant representation π : q(a) (cid:55)→ q(a) and t : q(Ti) (cid:55)→ q(Ti). This follows from almost tautological algebraic equa- tions, and by that (π, t) is injective and admits a gauge action. Similar remarks hold also for the C*-correspondence q(E0) over q(A0) for the C*-correspon- dence E0 of Remark 5.2. 6. C*-algebras associated with a monomial ideal Given a monomial ideal I (cid:67) C(cid:104)x1, . . . , xd(cid:105) we can form the C*-algebras C∗(T ) and C∗(T )/K(FX) related to the subproduct system XI. On the other hand the C*-corresponden- ce AEA associated with I initiates automatically two more C*-algebras, namely the Pimsner algebras TE and OE. In this section we show the connection between all four. Theorem 6.1. Let AEA be the C*-correspondence of a monomial ideal I (cid:67) C(cid:104)x1, . . . , xd(cid:105). Then the following diagram holds: C∗(T ) (cid:54)(cid:39) TE ⇔ I (cid:54)= (0) ⇔ E (cid:54)(cid:39) Cd C∗(T ) (cid:39) TE ⇔ I = (0) ⇔ E (cid:39) Cd ker φE (cid:54)= (0) ker φE = (0) OE (cid:39) C∗(T )/K(FX ) (cid:39) Od , ker φE = (0) OE (cid:39) C∗(T ) OE (cid:39) C∗(T )/K(FX ) with the understanding that all ∗-isomorphisms are canonical. The proof is induced by Proposition 6.3, Proposition 6.4, and Corollary 6.5 that will follow. Remark 6.2. Theorem 6.1 reads the same for the C*-correspondence E0 of Remark 5.2 by substituting C∗(T ) with C∗(Tµ ∅ (cid:54)= µ ∈ Λ∗). We will deliberately use both I and the unit 1 of the C*-algebras throughout the proofs even when these operators coincide. By doing so, the proofs will read the same for both E and E0. Proposition 6.3. Let AEA be the C*-correspondence of a monomial ideal I in C(cid:104)x1, . . . , xd(cid:105). Then C∗(T ) is the relative Cuntz-Pimsner algebra O(J, E) for the ideal J generated by {1 − µ Tµ µ ∈ Λ∗}. Moreover C∗(T )/K(FX) is the relative Cuntz-Pimsner algebra O(A, E). T ∗ In particular J ⊆ JE ⊆ A, and there are canonical ∗-epimorphisms TE → C∗(T ) → OE → C∗(T )/K(FX). Proof. The mappings t : Ti (cid:55)→ Ti and π : a (cid:55)→ a define a representation (π, t) for E such that C∗(T ) = C∗(π, t). If we let uzeν = zνeν for z ∈ T and ν ∈ Λ∗ then the family {aduz}z∈T 28  K S  K S  defines a gauge action of C∗(T ). Since π is faithful on A we obtain that C∗(π, t) (cid:39) O(J, E) for the ideal J = {a ∈ A π(a) ∈ ψt(K(X))} by [42], and that J ⊆ JE by [50]. is a unit of K(E) we have that a ∈ J if and only if π(a)(I − ψt(e)) = 0 by Lemma 2.1. That is a ∈ J if and only if µ Tµ for all µ ∈ Λ∗. Since e =(cid:80)d We now show that J is generated by 1 − T ∗ i=1 ΘE δi,δi aP∅ = a(I − d(cid:88) TiT ∗ i ) = 0. µ Tµ ∈ J for all µ ∈ Λ∗. On the other hand recall that A Since P∅ ≤ T ∗ is an AF algebra. If a ∈ J ∩ Al (cid:54)= (0) then without loss of generality we may assume that a =(cid:80) λjQj where each minimal projection Qj of Al satisfies λjQj = Qja ∈ J and has the i=1 µ Tµ we get that 1 − T ∗ k(cid:89) T ∗ Tµj for some enumeration {µ1, . . . , µn} = Λ∗ appear in the sum for a above must satisfy k < n. Hence a is generated by some 1 − T ∗ and the proof of the first part is complete. l . Since aP∅ = 0 then the products giving the Qj that µ Tµ Tµj ), (I − T ∗ j=k+1 j=1 µj µj form n(cid:89) (cid:32) d(cid:88) i=1 For the second part notice that π(a) − ψt(φE(a)) = π(a) − ψt (cid:33) = π(a) − d(cid:88) ΘE ξiαi(a),ξi Tiπαi(a)T ∗ i . However π(a)Ti = Tiπαi(a) and therefore π(a) − ψt(φE(a)) = π(a)(I − d(cid:88) i=1 TiT ∗ i ) = π(a)P∅. i=1 Since C∗(T )/K(FX) is the quotient of C∗(T ) by the ideal generated by P∅ we have that C∗(T )/K(FX) is the relative Cuntz-Pimsner algebra for the ideal generated by A; hence C∗(T )/K(FX) is O(A, E). The canonical ∗-epimorphism TE → OE is not faithful because JE (cid:54)= (0). Indeed if φE is injective then JE = A, whereas if φE is not injective then JE (cid:54)= (0) by Proposition 5.8. Now we proceed to the examination of the relation between the remaining C∗(T ), OE, and C∗(T )/K(FX). Proposition 6.4. Let AEA be the C*-correspondence of a monomial ideal I (cid:67) C(cid:104)x1, . . . , xd(cid:105). Then we get the following dichotomy: (i) φE is injective if and only if OE (cid:39) C∗(T )/K(FX) by the canonical ∗-homomorphism; (ii) φE is not injective if and only if OE (cid:39) C∗(T ) by the canonical ∗-homomorphism. Hence there are canonical ∗-epimorphisms C∗(T ) → OE → C∗(T )/K(FX) where in any case exactly one of them is faithful. Proof. Recall that JE = ker φ⊥ E since the left action is by the compacts. Proposition 6.3 implies that OE (cid:39) C∗(T )/K(FX) if and only if JE = A which settles item (i). Hence if OE (cid:39) C∗(T ) then φE is not injective since q : C∗(T ) → C∗(T )/K(FX) cannot be injective. On the other hand if ker φE (cid:54)= (0) then P∅ ∈ A which implies that JE = (cid:104)1 − P∅(cid:105). Since 29 (1 − P∅)P∅ = 0 we have that 1 − P∅ ∈ J as in the proof of Proposition 6.3. Hence we get that J = JE which implies that OE (cid:39) O(J, E) = C∗(T ). For the next result recall that Td denotes the Toeplitz-Cuntz algebra associated with the row isometries of multiplicity d. Corollary 6.5. If AEA is the C*-correspondence of a monomial ideal I (cid:67) C(cid:104)x1, . . . , xd(cid:105), then C∗(T ) (cid:39) TE by the canonical ∗-homomorphism if and only if I = (0), if and only if C∗(T ) (cid:39) Td, if and only if E = Cd. Proof. By definition TE (cid:39) C∗(T ) = O(J, E) if and only if J = (0). By Proposition 6.3 this µ Tµ = I for all µ ∈ Λ∗. In particular iν ∈ Λ∗ for all i = 1, . . . , d is equivalent to having that T ∗ and ν ∈ Λ∗, which implies that I = (0). Conversely if I = (0) then obviously C∗(T ) (cid:39) Td and E = Cd. It is an open problem to determine when the Toeplitz algebra T (X) or when Cuntz- Pimsner algebra O(X) (as defined in [87]) of a subproduct system X is nuclear or exact. For the Toeplitz and the Cuntz-Pimsner algebras of a C*-correspondence these problems were solved in [34, 50]. Using the above results this problem is now resolved in the case where X is the subproduct system coming from a monomial ideal. The key is the identification of C∗(T ) as a relative Cuntz-Pimsner algebra. Corollary 6.6. Let X = XI be a subproduct system associated with a monomial ideal I (cid:67) C(cid:104)x1, . . . , xd(cid:105). Then the Toeplitz algebra T (X) := C∗(T ) = C∗(AX) and the Cuntz- Pimsner algebra O(X) := C∗(T )/K(FX) associated with X are both nuclear. Proof. Since A is commutative then TE is nuclear by [50, Theorem 7.2]. Therefore so are its quotients C∗(T ) and C∗(T )/K(FX). Remark 6.7. Nuclearity of C∗(T )/K(FX) has been observed by Matsumoto [59, Lemma 4.10] in the case of the two-sided subshifts and by using a different line of reasoning. When I = (0) then OE is the Cuntz algebra Od. We provide a universal property for OE when I (cid:54)= (0) as well. A universal property for C∗(T )/K(FX) when I is of finite type will be given in Theorem 7.11. Both Theorem 6.8 and Theorem 7.11 should be compared with [58, Theorem 4.9]. In particular [58, Theorem 4.9] follows from Theorem 6.8 below when φE is injective (see also in conjunction with Remark 5.10). Theorem 6.8. Let AEA be the C*-correspondence of a monomial ideal I (cid:67) C(cid:104)x1, . . . , xd(cid:105) with I (cid:54)= {0}. Then OE is the universal C*-algebra generated by d partial isometries si such that µsµ extends to a ∗-representation of A; µ Tµ (cid:55)→ s∗ the mapping T ∗ (i) j si = δi,js∗ (ii) s∗ i si; µsµ for all µ ∈ Λ∗ and i = 1, . . . , d; µsµsis∗ i = sis∗ (iii) s∗ i s∗ (iv) I − s∗ µ1sµ1 . . . s∗ µi ∈ Λ∗ such that µii /∈ Λ∗, or I =(cid:80)d sµd = (cid:80)d i=1 sis∗ i µd for any (and hence for every) d-tuple of words i=1 sis∗ i if no such d-tuple of words exist. Proof. If si are as above then they define a representation of E, which we will denote by µsµ and t : δi (cid:55)→ si. Indeed items (i) and (ii) show the compatibility with the π : T ∗ inner product and item (iii) shows the compatibility with the left action. µ Tµ (cid:55)→ s∗ 30 We show that item (iv) implies that (π, t) is covariant. Since 1 ∈ A acts as an identity on E we may suppose that π is non-degenerate, i.e. π(1) = I. If there are no words µii /∈ Λ∗ for therefore all i = 1, . . . , d, then 1 ∈ JE by Proposition 5.8. Recall that φE(1) =(cid:80)d i=1 ΘE δi,δi we obtain π(1) = I = sis∗ i = ψt(ΘE δi,δi ). d(cid:88) i=1 Hence (π, t) is covariant in this case. Now if µii /∈ Λ∗ for some word µi ∈ Λ∗ for all i = 1, . . . , d, then T ∗ Tµd ∈ ker φE and µ1Tµ1 . . . T ∗ µd E =(cid:10)1 − T ∗ ker φ⊥ µ1Tµ1 . . . T ∗ µd Tµd by Proposition 5.8. Thus we get that φE(1 − T ∗ µ1Tµ1 . . . T ∗ µd Tµd) = φE(1) = ΘE δi,δi . (cid:11) (cid:67) A, d(cid:88) i=1 Covariance of (π, t) is then given by the computation Tµd) = I − s∗ d(cid:88) d(cid:88) µ1sµ1 . . . s∗ sis∗ µ1Tµ1 . . . T ∗ π(1 − T ∗ i = ψt( = µd µd sµd ΘE δi,δi ). i=1 i=1 Finally to see that OE is the universal C*-algebra of the statement it suffices to find a faithful representation of OE that satisfies these properties. If there are no words µi ∈ Λ∗ such that µii /∈ Λ∗ then OE (cid:39) C∗(T )/K(FX) and the q(Ti) ∈ C∗(T )/K(FX) satisfy the enlisted properties and provide a faithful representation. If there are such words then OE (cid:39) C∗(T ) and the Ti satisfy the properties and provide a faithful representation of OE. In the literature there are several C*-algebras associated with subshifts and consequently with monomial ideals. These have been introduced as generalisations of the Cuntz-Krieger algebras. In Section 10.3 we show that they are distinct from the algebras OE and TE that we introduced above. The main difference is that in our case we pass to the quotient by the compacts only when the left action is injective. As we show with the next example, cutting C∗(T ) by the compacts unconditionally may erase important information from the original data. Example 6.9. Let I be the monomial ideal generated by {x2 definition 1, x1x2} inside C(cid:104)x1, x2(cid:105). By (cid:40) e1 0 T1eµ = if µ = ∅, otherwise, and T2eµ = e2µ. Therefore T1 is a compact operator, and T2 is unitarily equivalent to the direct sum of the forward shift with itself (one on the e2n and one on the e2n1 for n ∈ Z+). As a consequence we have that q(T1) = 0 and q(T2) = u ⊕ u, where u is the bilateral shift. Therefore we obtain that C∗(T )/K(FX) = C(T). 31 7. Tensor algebras associated with a monomial ideal We focus on two kinds of nonselfadjoint operator algebras associated with a monomial ideal I. These are: (i) the tensor algebra T + E of the C*-correspondence AEA in the sense of Muhly and (ii) the tensor algebra AX of the subproduct system XI in the sense of Shalit and Solel Solel [63]; and [80]. Let us record here the following corollary of our previous analysis. Corollary 7.1. Let AEA and X be the C*-correspondence and the subproduct system re- spectively of a monomial ideal I (cid:67) C(cid:104)x1, . . . , xd(cid:105). Then there is a completely isometrical embedding AX (cid:44)→ T + E . Proof. It is immediate since C∗(T ) = C∗(AX) is a C*-cover of T + 7.1. The tensor algebra T + allowable words determine the representations of the C*-correspondence AEA. Proposition 7.2. Let AEA be the C*-correspondence of a monomial ideal I (cid:67) C(cid:104)x1, . . . , xd(cid:105). Then a family (π,{Vi}d E . First we show that the quantised dynamics (A, α) of the i=1) defines a representation (π, t) of AEA if and only if E by Proposition 6.3. (i) π : A → B(H) is a ∗-representation; (ii) π(a)Vi = Viπαi(a) for all a ∈ A and i = 1, . . . d; (iii) V ∗ j Ti) for all j, i = 1, . . . d. j Vi = π(T ∗ Proof. Write E as the direct sum of Ei = (cid:104)δi(cid:105) by Proposition 5.3. If (π, t) is a representation of E then let Vi = t(δi). Then we obtain that π(a)Vi = π(a)t(δi) = t(φE(a)δi) = t(δiαi(a)) = t(δi)παi(a) = Viπαi(a), and that j Vi = t(δj)∗t(δi) = π((cid:104)δj, δi(cid:105)) = π(T ∗ V ∗ for all a ∈ A and j, i = 1, . . . , d. Conversely, if a family (π,{Vi}d the statement then (π, t) defines a representation for E with t((cid:80)d j Ti), i=1) satisfies the properties of i=1 δiai) :=(cid:80)d i=1 Viπ(ai). The quantised dynamics determine the completely contractive representations of T + E as well. Since E is a non-degenerate C*-correspondence the following result can be obtained by using the completely contractive covariant representations of E in the sense of Muhly and Solel [63]. We sketch an alternative proof that settles also the case of E0. Proposition 7.3. Let AEA be the C*-correspondence of a monomial ideal I (cid:67) C(cid:104)x1, . . . , xd(cid:105). Then a family (π,{Vi}d E if and only if i=1) defines a completely contractive representation of T + (i) π : A → B(H) is a ∗-representation; (ii) π(a)Vi = Viπαi(a) for all a ∈ A and i = 1, . . . d; (iii) j Vi] ≤ [π(T ∗ [V ∗ j Ti)]. A similar conclusion holds for T + and representations π : A0 → B(H). E0 32 Proof. If (π, t) defines a completely contractive representation π × t of T + E then let T1 0 ... 0  ··· Td ··· 0 ··· ... ··· 0 G = and compute j Vi] = t ⊗ idn(G)∗ · t ⊗ idn(G) [V ∗ = (π × t) ⊗ idn(G)∗ · (π × t) ⊗ idn(G) ≤ ((π × t) ⊗ idn) (G∗G) = [π(T ∗ j Ti)], for Vi = t(δi) with i = 1, . . . , d. For the converse we write Hence [V1, . . . , Vd] is a row contraction since π(T ∗ j Ti) ≤ δi,jI. By using the defect operator [V ∗ j Vi] = [V1, . . . , Vd]∗[V1, . . . , Vd]. D =(cid:0)[π(T ∗ j Vi](cid:1)1/2 j Ti)] − [V ∗ the proof then follows as in [23, Theorem 2.1]. The case of E0 follows in the same way. Suppose that A acts faithfully on a Hilbert space H and let the Hilbert space K = ⊕µ∈Λ∗H. On K we define the ∗-representation π0(a) = diag{αν(a) ν ∈ Λ∗}, ν . We remark here that α : Fd where αν = adT ∗ αναµ = αµν. Moreover let Li be the regular shifts on F+ and define + → End(A) is an anti-homomorphism because µ Tµ d , i.e. Lieν = eiν. Let P = ⊕µ∈Λ∗T ∗ π(·) = P π0(·)P and Vi = P LiP for all i = 1, . . . , d. Since A is commutative we get that P commutes with π0. Moreover we have that P LiP = P Li, hence Vµ = P Lµ for all µ ∈ Λ∗. The family (π,{Vi}d i=1) satisfies the conditions of Proposition 7.2, hence it defines a representation (π, t) of E. In addition (π, t) admits the gauge action given by βz := aduz with uz(ξ ⊗ eµ) = zµξ ⊗ eµ. In particular we get d(cid:88) (cid:40) i = ψt(e) for the identity e =(cid:80)d i (ξ ⊗ eµ) = ViV ∗ i=1 ξ ⊗ eµ 0 i=1 ΘE δi,δi and(cid:80)d i=1 ViV ∗ have that π(a) ∈ ψt(K(E)) if and only if if µ (cid:54)= ∅, otherwise, of K(E). In view of Lemma 2.1 we 0 = π(a)(I − ψt(e)) = π(a)P∅ = a. Hence by the gauge invariant uniqueness theorem [50] the pair (π, t) defines a faithful rep- resentation of TE. Therefore the restriction of π × t to the tensor algebra T + E is a completely isometric homomorphism. 33 We will use that V ∗ µ Vµ = π(T ∗ µ Tµ) for all µ ∈ Λ∗. This follows by induction and the computation V ∗ i V ∗ µ VµVi = V ∗ = π(T ∗ µ Tµ)Vi = V ∗ i T ∗ µ Tµ) is a projection we obtain i π(T ∗ i TiT ∗ i Viπαi(T ∗ i T ∗ µ TµTi) = π(T ∗ µ Tµ) µ TµTi). Vµπ(a) = VµV ∗ µ Vµπ(a) = Vµπ(T ∗ µ Tµa). E → A be the compression to the (∅, µ)-entry. Then Eµ is a contractive Since V ∗ µ Vµ = π(T ∗ For µ ∈ Λ∗, let Eµ : T + map such that Therefore every element f ∈ T + (cid:88) ν∈Λ∗ l Eµ( Vνπ(aν)) = T ∗ µ Tµaµ. (cid:88) E can be written as f = f(cid:48) + Vµπ(aµ) l . Moreover if f = g(cid:48) +(cid:80) µ∈Λ∗ l with Eµ(f(cid:48)) = 0 for every µ ∈ Λ∗ Eµ(g(cid:48)) = 0 for every µ ∈ Λ∗ l , then µ Tµ(bµ − aµ) = 0 for all µ ∈ Λ∗ T ∗ l , µ∈Λ∗ l i.e. the coefficients aµ are unique modulo T ∗ that µ Tµ, for every µ ∈ Λ∗ l . Consequently we obtain Vµπ(bµ) for some bµ ∈ A with therefore Vµπ(bµ) = Vµπ(T ∗ g(cid:48) = f − (cid:88) µ Tµbµ) = Vµπ(T ∗ Vµπ(bµ) = f − (cid:88) µ∈Λ∗ l µ∈Λ∗ l µ Tµaµ) = Vµπ(aµ), Vµπ(aµ) = f(cid:48), The universal property of OE obtained in Theorem 6.8 enables us to provide the following which shows in what sense the decomposition is unique. result for the tensor algebra T + E . Theorem 7.4. Let AEA be the C*-correspondence of a monomial ideal I (cid:67) C(cid:104)x1, . . . , xd(cid:105). Then the tensor algebra T + E is hyperrigid. Proof. By [47] the C*-envelope of T + {si}d Si. Let ρ : T + E is OE. Suppose that OE is generated by a family i=1 as in Theorem 6.8 and let Φ : OE → B(H) be a unital ∗-representation with Φ(si) = (cid:21) E → B(K) be a maximal dilation of ΦT + and write E (cid:20)Si Xi Yi Zi ρ(si) = with respect to the decomposition of K = H ⊕ H⊥. We will denote by the same symbol the unique extension of ρ to a representation of OE. Then the ρ(si) satisfy the relations of Theorem 6.8. We aim to show that ρ is a trivial dilation. It suffices to show this for the generators a ∈ A and si of T + E . Φ(si)∗Φ(si) = S∗ First ΦA is a ∗-representation, hence it is a direct summand of ρA. Since Φ(T ∗ i Si then by equating the (1, 1)-entries of the equation ρ(T ∗ i Ti) = i Ti) = ρ(si)∗ρ(si) 34 I − (cid:20)Φ(s∗ If ker φE = (0) then I =(cid:80)d E we get that Yi = 0 for all i = 1, . . . , d. If ker φE (cid:54)= (0) then there are words µi ∈ Λ∗ such that µii /∈ Λ∗ and I − s∗ i . By applying ρ we obtain that µ1sµ1 . . . s∗ i=1 sis∗ µd sµd =(cid:80)d µ1sµ1 . . . s∗ µd 0 sµd) 0 ∗ = (cid:21) n(cid:88) (cid:20)SiS∗ i=1 i + XiX∗ i ∗ (cid:21) ∗ ∗ , and by equating the (1, 1)-entries we obtain that Xi = 0 for all i = 1, . . . , d. Hence ρ is a trivial dilation of ΦT + . i=1 sis∗ i and a same computation completes the proof. 7.2. The tensor algebra AX. Let us continue with the analysis of the nonselfadjoint op- erator algebra AX generated by the Tµ for µ ∈ Λ∗. Proposition 7.5. Let X be the subproduct system of a monomial ideal I (cid:67) C(cid:104)x1, . . . , xd(cid:105) of finite type k and let q : C∗(T ) → C∗(T )/K(FX). Then q(AX) is hyperrigid in C∗(T )/K(FX), hence C∗ Proof. Let Φ : C∗(T )/K(FX) → B(H) be a unital ∗-representation and let us write Φ(q(Ti)) = Wi for all i = 1, . . . , d. Let ρ be a maximal dilation of Φq(AX ) and let us denote by the same symbol the extension of ρ to C∗(T )/K(FX). Then we get env(q(AX)) = C∗(T )/K(FX). (cid:20)Wi Xi (cid:21) Yi Zi . Applying Φ and ρ to the equation I =(cid:80)d ρ(q(Ti)) = we get that Xi = 0 for all i = 1, . . . , d. We will require one more equation from [80, Section 12]; that is q(Ti)∗q(Ti) = q(Tµ)q(Tµ)∗ for all i = 1, . . . , d, (cid:88) µ∈Ek i i=1 q(Ti)q(Ti)∗ and by equating the (1, 1)-entries i := {µ ∈ Λ∗ k iµ ∈ Λ∗}. For a short proof, recall that the T ∗ i Ti is the projection on where Ek the space G := span{eµ µ ∈ Λ∗ such that iµ ∈ Λ∗}. Define the subspace G(cid:48) := span{eµ µ ∈ Λ∗ such that iµ ∈ Λ∗ and µ ≥ k} of finite co-dimension in G; then G(cid:48) = span{eµν µν ∈ Λ∗ and µ ∈ Ek i }. Indeed, if w = µν with iµ ∈ Λ∗ then iw ∈ Λ∗, otherwise we would have a forbidden word of length greater than k + 1 that does not contain a forbidden word. This expression of G(cid:48) shows that it is the range of the projection of(cid:80) TµT ∗ µ . µ∈Ek i Applying Φ and ρ to the above equation, and by restricting to the (1, 1)-entries we get that Yi = 0 for all i = 1, . . . , d. Hence ρ(q(Ti)) is a trivial dilation of Wi for all the generators of AX. 35 Theorem 7.6. Let X be the subproduct system of a monomial ideal I (cid:67) C(cid:104)x1, . . . , xd(cid:105) of finite type and let q : C∗(T ) → C∗(T )/K(FX). Then items (1), (2), and (3) of the following diagram hold: qAX is not completely isometric. qAX is completely isometric. (1) env(AX ) (cid:39) C∗(T ) C∗ (2) env(AX ) (cid:39) C∗(T )/K(FX ) C∗ (3) (4) (4) (3) ∀i = 1, . . . , d,∃µi ∈ Λ∗.µii /∈ Λ∗. ∃i ∈ {1, . . . , d},∀µ ∈ Λ∗.µi ∈ Λ∗. If in addition the µi can be chosen to have the same length then item (4) also holds. In particular item (4) holds when the ideal I has no sinks on the left. env(AX) (cid:39) Proof. Referring to item (1), if qAX is not completely isometric then we obtain C∗ env(AX) (cid:39) C∗(T ) then qAX cannot be com- C∗(T ) by [4, Theorem 2.1.1]. Conversely if C∗ pletely isometric since K(FX) (cid:54)= (0). Referring to item (2), if qAX is completely isometric then Proposition 7.5 implies that env(AX) (cid:39) C∗(T )/K(FX). The converse is trivial. C∗ Referring to items (3) and (4), if there is an i ∈ {1, . . . , d} such that µi ∈ Λ∗ for all µ ∈ Λ∗ E ) = OE = C∗(T )/K(FX) by Theorem 6.1. This shows that then ker φE = (0) hence C∗ qT + E by Corollary 7.1, then qAX is completely isometric, as well. What remains to show is that if there are words µ1, . . . , µd ∈ Λ∗ of the same length such that µii /∈ Λ∗ for all i = 1, . . . , d, then qAX is not completely isometric. Initially the words µi may not be distinct. However we can restrict to a subset and assume that we have n distinct words µk such that for any i = 1, . . . , d there exists a ki ∈ {1, . . . , n} with µkii /∈ Λ∗. Let the element T = Tµ1 + ··· + Tµn in AX. Then T ∗ Tµj = 0 since the distinct words have the same length so that is completely isometric. Since AX ⊆ T + env(T + µi E (cid:107)T(cid:107)2 = (cid:107)T ∗T(cid:107) =(cid:13)(cid:13)T ∗ µ1Tµ1 + ··· + T ∗ µnTµn (cid:13)(cid:13) . µ1Tµ1 ··· T ∗ Recall that by Lemma 5.7 we have T ∗ µnTµn = P∅. By splitting every projection into a sum of minimal subprojections with respect to the family {T ∗ Tµi i = 1, . . . , n} µ1Tµ1 . . . T ∗ we find that the common subprojection P∅ = T ∗ µnTµn appears n times and con- sequently (cid:107)T(cid:107)2 = n. The same decomposition in the quotient shows that every subpro- Tµk) appears at most n − 1 times since their only common subprojection jection of q(T ∗ q(T ∗ µ1Tµ1 . . . T ∗ µk µnTµn) is equal to q(P∅) = 0. Thus µi (cid:107)q(T )(cid:107)2 =(cid:13)(cid:13)q(T ∗ µ1Tµ1) + ··· + q(T ∗ Therefore qAX is not completely isometric in this case. µnTµn)(cid:13)(cid:13) ≤ n − 1. Finally we mention that if there are no sinks on the left for I then we can add more letters to the left of each µi and arrange them to have the same length. Note that the new elements will still satisfy the same condition. 36 K S  K S   K S  K S Remark 7.7. Let us compare the C*-correspondences E and q(E). In Remark 5.10 we noted that C∗(T )/K(FX) is the Cuntz-Pimsner algebra of q(E). However C∗(T ) may not be in general a relative Cuntz-Pimsner algebra of q(E) in the canonical way. Indeed, if it were then T + E ), which we showed that does not hold in general. E should be completely isometric to T + q(E) = q(T + E E and qT + Remark 7.8. We wish to record here a different proof of item (3) of Theorem 7.6 that does not use the facts that AX ⊆ T + is completely isometric. It is an adaptation of [47, Lemma 3.2] and it can be generalised to other subproduct systems where a convenient relative Cuntz-Pimsner algebra may not be present. Let us content ourselves in showing that the quotient map is isometric on the normed algebra alg{1, T1, . . . , Td}. Let p(T ) be a polynomial in T1, . . . , Td, and let x be a unit vector in the linear span of X(m) for m ≥ 0, such that (cid:107)p(T )x(cid:107) > (cid:107)p(T )(cid:107) − ε. We have that µin is an allowed word for all n and all µ ∈ Λ∗, thus pm+n(y ⊗ ein) = pm(y) ⊗ ein for all y ∈ X. Let the contractive operator Ri ∈ B(FX) defined by Riy = pm+1(y ⊗ ei) = pm(y) ⊗ ei for y ∈ X(m). Therefore Ri commutes with p(T ) and (cid:107)Rn then have i p(T )x(cid:107) ≥ (cid:107)p(T )(cid:107) − ε. For every K ∈ K(FX) we i x(cid:107) . (cid:107)p(T ) + K(cid:107) ≥ (cid:107)(p(T ) + K)Rn i x(cid:107) > (cid:107)p(T )(cid:107) − ε − (cid:107)KRn Letting n → ∞ we obtain (cid:107)q(p(T ))(cid:107) ≥ (cid:107)p(T )(cid:107) − ε for all ε, as required. Remark 7.9. There is a similarity of Theorem 7.6 to what is known for commutative sub- env(q(AX)) = OX product systems. By [52] if X is a commutative subproduct system then C∗ if and only if the shift [T1, . . . , Td] is essentially normal. By [7] monomial ideals in commut- env(q(AX)) = OX ing variables give rise to essentially normal shifts. Thus we obtain that C∗ for monomial ideals in the commutative case as well. Recall that in the commutative case all ideals are finitely generated. Question 7.10. The only part in Theorem 7.6 where we use that I is of finite type is when env(AX) = C∗(T )/K(FX). Finiteness is needed to showing that if qAX is isometric then C∗ apply Proposition 7.5. We ask if Proposition 7.5 holds in general, or even more if equivalence (2) of Theorem 7.6 holds in general. We now give a universal property of C∗(T )/K(FX) for the case where I is of finite type (compare the following with Theorem 6.8). This is an illustration of the utility of the dilation techniques used for hyperrigidity. the equation C∗(T ) = spanAXA∗ therein. We do not know whether this claim is always true. Theorem 7.11. Let I be a monomial ideal of finite type k. Then the algebra C∗(T )/K(FX) is the universal C*-algebra generated by a row contraction s = [s1, . . . , sd] such that Also we take this opportunity to fill a gap in the proof of [80, Theorem 12.7]. In particular, X used in the proof of [80, Theorem 12.7] was not justified (i) I =(cid:80)d i si =(cid:80) (ii) p(s) = 0 for all p ∈ I; (iii) s∗ i=1 sis∗ i ; sµs∗ µ∈Ek i µ where Ek i = {µ ∈ Λ∗ k iµ ∈ Λ∗}, for all i = 1, . . . , d. 37 Proof. Denote w = [q(T1), . . . , q(Td)]. The identities (i) and (ii) then hold for w. Identity (iii) for w is provided in the course of the proof of Proposition 7.5. It remains to show that whenever s is a d-tuple as in the statement of the theorem, then there exists a ∗- homomorphism π : C∗(T )/K(FX) → C∗(s) with π(wi) = si. To this end, denote EX := spanAXA∗ positive map Ψ : EX → B(H) satisfying ν ) = sµs∗ X. By [80, Theorem 8.2], there is a unital completely and Ψ(ab) = Ψ(a)Ψ(b) Ψ(TµT ∗ for all µ, ν ∈ Fd + and for all a ∈ AX, b ∈ EX. Let Ψ denote also a unital completely positive extension of Ψ to C∗(T ). Using property (i) for s and Lemma 4.2 items (vi) and (vii) for T , we find that K(FX) ⊆ ker Ψ. Indeed we directly compute i T ∗ Ψ(TµTiT ∗ ν ) ν ) = Ψ(TµT ∗ Ψ(TµP∅T ∗ ν ν ) − d(cid:88) ν − d(cid:88) i=1 νi = sµ(I − d(cid:88) sµis∗ = sµs∗ ν in B(FX). Thus Ψ induces a unital completely positive for all rank one operators TµP∅T ∗ map Φ : C∗(T )/K(FX) → B(H) for which Φ(q(a)) = Ψ(a), for all a ∈ C∗(T ). In particular, d(cid:88) we have Φ(wi) = si, as well as ν = 0, i )s∗ sis∗ i=1 i=1 Φ(wiw∗ i ) = I = sis∗ i , and Φ(w∗ i wi) = Φ(wµw∗ µ) = sµs∗ µ = s∗ i si, for all i = 1, . . . , d. Let ρ : C∗(T )/K(FX) → B(K) be the Stinespring dilation of Φ. We can now use the items (i), (ii), and (iii) as in the proof of Proposition 7.5 to obtain that ρq(AX ) is a trivial dilation of Φq(AX ). Since q(AX) generates C∗(T )/K(FX) we get that Φ is a ∗-representation. Corollary 7.12. Let I be a monomial ideal of finite type k, and let s = [s1, . . . , sd] be a row contraction satisfying items (i) -- (iii) of Theorem 7.11. The each si is a partial isometry. i 8. Encoding via C*-correspondences We turn our attention to the classification of our data by using C*-correspondences. Our aim is to show that C*-correspondences and their tensor algebras form a complete invariant for monomial ideals up to local conjugacy of the induced quantised dynamics. To allow comparisons we fix the following notation. Let J be a monomial ideal of C(cid:104)y1, . . . , yd(cid:48)(cid:105), and let BFB be the C*-correspondence associated with J . We write M∗ for the allowable words related to J , and R∗ wRw with w ∈ M∗ for the generators of B. Moreover suppose that ζi = R∗ i Ri ∈ B generate each direct summand Fi of F as in Section 5.2. For [n]2 ≡ [n] = [n1n2 . . . nd(cid:48)] ∈ {0, . . . , 2d(cid:48) − 1} we write (cid:89) i Ri · (cid:89) R∗ ni=0 38 P[n] := ni∈supp[n] (I − R∗ i Ri), d(cid:88) (cid:88) i=1 µ∈Ek i=1 (cid:88) µ∈Ek i i Ri i = 1, . . . , d(cid:48)). The quantised dynamics associated i bRi for i = 1, . . . , d(cid:48). In analogy to E0 of for the minimal projections in C∗(I, R∗ with J are denoted by (B, β) where βi(b) = R∗ Remark 5.6 we write B0 and F0 for J . 8.1. Unitary equivalence. We collect a few facts about module maps and their matrix representations. Since both EA and FB are finitely generated, an adjointable map (γ, U ) is i=1 ζibij with bij ∈ B, then characterised by the images U (δj) for j = 1, . . . , d. If U (δj) =(cid:80)d(cid:48)  : δj (cid:55)→ d(cid:48)(cid:88) we can represent U as the matrix  →  F1 E1 [Θζibij ,δj ] : ζibij, ... Ed ... Fd(cid:48) i=1 which shows that L(E, F ) = K(E, F ). Since Θζibij ,δj = Θζi,δj γ−1(b∗ j Tj, we obtain that the elements bij ∈ A are unique up to the equations and δj = δj · T ∗ R∗ i Ribij = bij = bijγ(T ∗ j Tj) for all i = 1, . . . , d(cid:48) and j = 1, . . . , d. ij ) as well as ζi = ζi · R∗ Conversely if there is a ∗-isomorphism γ : A → B then a matrix [bij] ∈ Md(cid:48)×d(B) defines a (γ, U ) ∈ L(E, F ) by i Ri Indeed we can write U =(cid:80)d(cid:48) so that the adjoint of U is given by U (δj) = ζibij for all j = 1, . . . , d. j=1 Θζibij ,δi. It is immediate that (Θζibij ,δi)∗ = Θδj γ−1(bij )∗,ζi i=1 U∗(ζj) = δiγ−1(b∗ ij) for all j = 1, . . . , d. i Ribijγ(T ∗ Both [bij] and [R∗ j Tj)] produce the same U , thus the correspondence [bij] (cid:55)→ U is not one-to-one. We will often be replacing the bij by the R∗ j Tj). We will call this procedure the calibration of [bij]. Hence the identity maps on E and on F are represented respectively by i Ribijγ(T ∗ [tij] = diag{T ∗ [rij] = diag{R∗ i Ti i = 1, . . . , d} i Ri i = 1, . . . , d(cid:48)}. and If ξ =(cid:80)d j=1 δjaj ∈ E is identified with the column vector [a1, . . . , ad]t ∈ ⊕d i TiA, then U (ξ) may be identified with the vector [b1, . . . , bd(cid:48)]t ∈ ⊕d(cid:48) j=1 bijγ(aj). Hence the composition (γ1, U1) ∈ L(E, F ) related to [bij] with (γ2, U2) ∈ L(D, E) related to [aij] is represented by the matrix [bij]· [γ1(aij)]. In particular a W ∈ L(E) that is represented by [wij] is injective (resp. surjective) if and only if there is a matrix [vij] ∈ Md(A) such that [vij] · [wij] = [tij] (resp. [wij] · [vij] = [tij]). If in addition (γ, U ) : AEA → BFB is a C*-correspondences map, then the left module j=1R∗ i=1T ∗ i RiB where bi =(cid:80)d properties aδj = δjαj(a) and bζi = ζiβi(b) give that d(cid:48)(cid:88) i=1 (cid:80)d d(cid:88) i=1 d(cid:48)(cid:88) d(cid:48)(cid:88) ζiβi(γ(a))bij = γ(a)U (δj) = U (δj)γ(αj(a)) = ζibijγ(αj(a)), i=1 j=1 39 for all a ∈ A. Therefore we get that βiγ(·) bij = bij γαj(·) for all i = 1, . . . , d(cid:48) and j = 1, . . . , d. Conversely if there is a ∗-isomorphism γ : A → B, then a matrix [bij] ∈ Md(cid:48)×d(B) satisfying βiγ(·) bij = bij γαj(·) defines a C*-correspondences map (γ, U ) : AEA → BFB. In particular, we can replace the coefficients bij with the calibrated elements R∗ j Tj). This follows by the fact that A and B are commutative, and because βi(·)R∗ j Tjαj(·) = αj(·). Proposition 8.1. Let I (cid:67) C(cid:104)x1, . . . , xd(cid:105) and J (cid:67) C(cid:104)y1, . . . , yd(cid:48)(cid:105) be monomial ideals. If there is an invertible C*-correspondence map between the associated C*-correspondences AEA and BFB, then d = d(cid:48), and i Ribijγ(T ∗ i Ri = βi(·) and T ∗ (cid:12)(cid:12)(cid:8)j ∈ {1, . . . , d} P[n]γ(T ∗ j Tj) (cid:54)= 0(cid:9)(cid:12)(cid:12) = supp[n], for all [n] ∈ {1, . . . , 2d(cid:48) − 1}. Proof. Let (γ, V ) : E → F be an invertible C*-correspondence map. Then there are matrices [bij] ∈ Md(cid:48)×d(B) and [aij] ∈ Md×d(cid:48)(A) such that [bij] · [γ(aij)] = [rij]. By symmetry there are also matrices [fij] ∈ Md×d(cid:48)(A) and [gij] ∈ Md(cid:48)×d(B) so that [fij][γ−1(gij)] = [tij]. Without loss of generality we may assume that all bij, aij, fij, and gij are calibrated. First we show that d = d(cid:48). To this end for P := P[1...1] = R∗ 1R1 . . . R∗ d(cid:48)Rd(cid:48) we obtain that P ⊗ Id(cid:48) · [bij] · [γ(aij)] · P ⊗ Id(cid:48) = P ⊗ Id(cid:48) · [rij] · P ⊗ Id(cid:48). Since P is a non-zero subprojection of every R∗ that [P bijP ] · [P γ(aij)P ] = P ⊗ Id(cid:48). i Ri and by using commutativity of B we get This shows that the matrix [P bijP ] with entries from the commutative and unital C*- algebra P BP is right invertible. As a consequence we obtain that d(cid:48) ≤ d. By symmetry on [fij][γ−1(gij)] = [tij] we obtain that d ≤ d(cid:48) as well. i Ri i = 1, . . . , d). Without loss of gener- We repeat for the minimal projections of C∗(R∗ k+1Rk+1) . . . (I − R∗ ality fix [n] = [1 . . . 10 . . . 0] and the associated projection P := P[n] = R∗ kRk(I − R∗ 1R1 . . . R∗ dRd). We may assume so by applying a permutation on the variables i = 1, . . . , d, which produces a unitary equivalence on BFB. For convenience let us set d − l = {j = 1, . . . , d P γ(T ∗ j Tj) = 0}. After permuting the variables (which produces a unitary equivalence on AEA), we may assume that 0 40 We aim to show that l = k. By commutativity in B we have that the equation P ⊗ Id(cid:48) · [bij] · [γ(aij)] · P ⊗ Id(cid:48) = P ⊗ Id(cid:48) · [rij] · P ⊗ Id(cid:48), {j = 1, . . . , d P γ(T ∗ j Tj) = 0} = {l + 1, . . . , d}. (cid:20)[P ]k×k 0 (cid:21) (cid:21)(cid:20)[P γ(aij)P ]l×k 0 (cid:21) = 0 0 . 0 (cid:20)[P bijP ]k×l 0 0 0 implies Reasoning as above for the commutative and unital P BP we get that k ≤ l. To reach contradiction suppose that k < l. Now P is written as the sum of P γ(Q[m]) for [m] = 0, . . . , 2d − 1 and k ≥ 1. Thus there exists an [m] with supp[m] = l > 1 such that P γ(Q[m]) (cid:54)= 0. Since we have permuted the variables to have {j = 1, . . . , d P γ(T ∗ j Tj) = 0} = {l + 1, . . . , d}, l+1Tl+1) . . . (I − T ∗ l Tl(I − T ∗ we can take Q ≡ Q[m] = T ∗ Since Q is minimal and P γ(Q) (cid:54)= 0, then we obtain that W := γ−1(P )Q = Q. Applying on [fij] · [γ−1(gij)] = [tij] we obtain 1 T1 . . . T ∗ d Td). [W fijW ]l×k · [W γ−1(gij)W ]l×k = [W tijW ]l×l = W ⊗ Il. Reasoning as above for the commutative unital C*-algebra QAQ we get that l ≤ k which is a contradiction. Therefore k must be equal to l. Proposition 8.2. Let I (cid:67) C(cid:104)x1, . . . , xd(cid:105) and J (cid:67) C(cid:104)y1, . . . , yd(cid:48)(cid:105) be monomial ideals, and let AEA and BFB be the associated C*-correspondences. Then AEA is unitarily equivalent to BFB if and only if: (a) d = d(cid:48); (b) there exists a ∗-isomorphism γ : A → B; and (c) there exist [bij] ∈ Md(B) and [aij] ∈ Md(A) such that βiγ(·)bij = bijγαj(·) and (i) (cid:12)(cid:12)(cid:8)j ∈ {1, . . . , d} P[n]γ(T ∗ j Tj) (cid:54)= 0(cid:9)(cid:12)(cid:12) = supp[n]; (ii) [P[n]bijP[n]] · [P[n]γ(aij)P[n]] = [P[n]rijP[n]]; for all the minimal projections P[n] ∈ C∗(R∗ i Ri i = 1, . . . , d). Proof. If (γ, V ) : E → F is an invertible C*-correspondence map then that d = d(cid:48) and item (i) follow from Proposition 8.1. In this case if [bij] is the matrix associated with V and [aij] is the matrix associated with V −1 then the identity V V −1 = IF implies that [bij]·[γ(aij)] = [rij]. Multiplying the latter equation by P[n] ⊗ Id and using commutativity of B gives item (ii). Conversely, the assumption βiγ(·)bij = bijγαj(·) implies that the matrix [bij] defines a C*-correspondence map between E and F . We may assume that the [bij] and (hence) the i Ri i = 1, . . . , d). Without loss of [aij] are calibrated. Fix a minimal projection P ∈ C∗(R∗ generality we may suppose that P = R∗ 1R1 . . . R∗ kRk(I − R∗ k+1Rk+1) . . . (I − R∗ After permuting the columns we have that item (ii) implies 0 where l = {j ∈ {1, . . . , d} P γ(T ∗ P BP is stably finite with identity P we derive that (cid:20)[P bijP ]k×l 0 (cid:20)[P γ(aij)P ]k×k 0 (cid:21)(cid:20)[P γ(aij)P ]l×k 0 (cid:21) (cid:21)(cid:20)[P bijP ]k×k 0 (cid:21) dRd). (cid:20)[P ]k×k 0 (cid:21) (cid:20)[P ]k×k 0 (cid:21) = 0 0 0 0 0 , 0 0 0 = 0 0 . 0 j Tj) = 0}. However item (i) implies that k = l. Since Equivalently [P γ(aij)P ] · [P bijP ] = [P rijP ]. By using commutativity once more we derive that and the symmetrical P ⊗ Id · [γ(aij)] · [bij] = P ⊗ Id · [rij], P ⊗ Id · [bij] · [γ(aij)] = P ⊗ Id · [rij], 41 from item (ii). Since the projections P add up to an identity for the calibrated elements we get that [bij] · [γ(aij)] = [rij] and that [γ(aij)] · [bij] = [rij]. The first equation shows that the C*-correspondence map U : E → F associated with [bij] is surjective and the second one that U is injective. Therefore the unitary UU−1 provides the unitary equivalence. 8.2. Isometric isomorphisms. The following result will be refined later by Corollary 8.12. We include an independent proof as it provides evidence of some additional structure (see also Remark 8.7). Theorem 8.3. Let I (cid:67) C(cid:104)x1, . . . , xd(cid:105) and J (cid:67) C(cid:104)y1, . . . , yd(cid:48)(cid:105) be monomial ideals, and let AEA and BFB be the associated C*-correspondences. The following are equivalent: E and T + E and T + (i) T + (ii) T + (iii) AEA and BFB are unitarily equivalent. F are completely isometrically isomorphic; F are isometrically isomorphic; Proof. The implications [(iii) ⇒ (i) ⇒ (ii)] are immediate by the general theory of C*- correspondences. For the implication [(ii) ⇒ (iii)], fix an isometric isomorphism Φ : T + E → T + F and denote γ = ΦA. Recall that A and B are maximal C*-algebras inside T + E and T + F , respectively. Since γ is isometric it is a ∗-homomorphism and thus maps A into B. The symmetrical argument for γ−1 = Φ−1B gives that γ : A → B is a ∗-isomorphism. Let us fix notation. Suppose that T + T + F is generated by B and wi for i = 1, . . . d(cid:48). Claim 1. There exists a C*-correspondence map (γ, U ) : E → F . Proof of Claim 1. By the Fourier analysis of Section 7.1 we can write E is generated by A and vj for j = 1, . . . , d, and that (cid:33) (cid:32) d(cid:48)(cid:88) i=1 Φ(vj) = b0 + wibij + f, with f having zero Fourier coefficients on ∅, 1, . . . , d(cid:48). Even more we can choose the bij so that R∗ hence wibij = wibijγ(T ∗ i Ribij = bij. Furthermore we have that Φ(vj) = Φ(vjT ∗ j Tj) = Φ(vj)γ(T ∗ j Tj). Thus we obtain the calibrated bij = R∗ i Ribij = R∗ i Ribijγ(T ∗ j Tj) = bijγ(T ∗ j Tj). j Tj) Recall the algebraic relations a · vj = vj · αj(a) and b · wi = wi · βi(b). Applying then Φ on the first one and using the second we derive that (cid:32) d(cid:48)(cid:88) i=1 (cid:33) (cid:32) d(cid:48)(cid:88) i=1 (cid:33) + f(cid:48)(cid:48). γ(a)b0 + wiβiγ(a)bij + f(cid:48) = b0γαj(a) + wibijγαj(a) Then the uniqueness of the coefficients gives that βiγ(·) bij = bij γαj(·). Letting U : E → F be determined by [bij] completes the proof of Claim 1. Our goal is to show that U is invertible, since then UU−1 will produce the unitary equivalence between E and F . We proceed in two steps. Claim 2. The C*-correspondence map U : E → F is surjective. 42 Proof of Claim 2. The arguments of [44, Lemma 4.2] apply in our case. Indeed the argu- ments of [44, Lemma 4.2] depend only on surjectivity, continuity, and the covariant relations βiγ(·) bij = bij γαj(·). Hence for any tuple [y1, y2, . . . , yd(cid:48)] ∈ ⊕d(cid:48) i=1B there exist a sequence 1, xk 2, . . . , xk j=1B such that wi(bi1xk 1 + bi2xk 2 + ··· + bidxk d), for all i = 1, 2, . . . , d(cid:48). (cid:0)[xk d](cid:1) k in ⊕d wiyi = lim k Since R∗ i Ribij = bij we get that R∗ i Riyi = lim k bi1xk Applying this for the tuples [R∗ Proposition 4.3] we can find [xij] such that 1 + bi2xk 1R1, 0, . . . , 0], . . . , [0, . . . , 0, R∗ 2 + ··· + bidxk d, for all i = 1, 2, . . . , d(cid:48). dRd] and proceeding as in [44, (cid:107)[rij] − [bij][xij](cid:107) < 1, where (cid:107)·(cid:107) denotes the norm in Md(cid:48)(B). We may as well substitute xij by the elements γ(T ∗ j Rj and still reach the same conclusion. Indeed we have that i Ti)][xijR∗ j Rj](cid:13)(cid:13) =(cid:13)(cid:13)[rij] − [bijγ(T ∗ (cid:13)(cid:13)[rij] − [bij][γ(T ∗ j Rj](cid:13)(cid:13) i Ti)xijR∗ i Ti)xijR∗ = (cid:107)([rij] − [bij][xij])[rij](cid:107) ≤ (cid:107)[rij] − [bij][xij](cid:107) < 1. Therefore we obtain that Set for convenience g = [rij]− [bij][xij] which is a matrix in Md(cid:48)(B). Then the series(cid:80)N [rij][bij][xij] = [bij][xij][rij] = [bij][xij]. n=1 gn converge and [rij]g = g[rij] = g. Therefore ∞(cid:88) ∞(cid:88) [bij][xij] · gn = ([rij] − g) gn = [rij]. Thus there is an element [cij] ∈ Md(cid:48)(B) such that n=0 n=0 [bij][xij][cij] = [rij]. Letting [aij] = [γ−1(xij)][γ−1(cij)] ∈ Md×d(cid:48)(A) we obtain [bij][γ(aij)] = [rij]. Without loss of generality we may assume that aij = T ∗ i Ti)γ(aij)R∗ j Rj). This follows by the computation [bij] · [γ(T ∗ i Tiaijγ−1(R∗ j Rj] = [bij] · [γ(tij)] · [γ(aij)] · [rij] j Tj)] · [γ(aij)] · [rij] = [bijγ(T ∗ = [bij] · [γ(aij)] · [rij] = [rij]. Let (γ−1, V ) : F → E be the adjointable map associated with [aij]. We compute d(cid:88) d(cid:48)(cid:88) d(cid:48)(cid:88) U V (ζi) = ζlblkγ(aki) = ζlδl,iR∗ i Ri = ζiR∗ i Ri = ζi k=1 l=1 l=1 hence U V = IF and the proof of Claim 2 is complete. Reasoning with Φ−1 instead of Φ, we find that there exists a surjective C*-correspondence map (cid:101)U : F → E. Denote W = (cid:101)U U : E → E. The following claim will conclude the proof of the theorem. 43 Claim 3. The C*-correspondence map W : E → E is injective. Proof of Claim 3. Denote by [wij] ∈ Md(A) the matrix representing W . Since W is surjective, there is a matrix [vij] such that [wij] · [vij] = [tij]. It suffices to show that [vij] · [wij] = [tij] also holds. For this, it is enough to show that Q[m] ⊗ Id · [vij] · [wij] = Q[m] ⊗ Id · [tij] i Ti i = 1, . . . , d), since these projections add up for every minimal projection Q[m] in C∗(T ∗ to an identity of the entries by Lemma 4.4. As above, we may assume that every wij has already been calibrated to satisfy wij = T ∗TiwijT ∗ i Ti i = 1, . . . , d), which without loss To this end let Q[m] be a minimal projection in C∗(T ∗ j Tj and similarly for vij. of generality we assume 1 T1 . . . T ∗ Using commutativity of A, we have that Q ≡ Q[m] = T ∗ k Tk(I − T ∗ k+1Tk+1) . . . (I − T ∗ d Td). implies Q ⊗ Id · [wij] · [vij] = Q ⊗ Id · [tij] (cid:20)[QwijQ]k×k 0 (cid:21)(cid:20)[QvijQ]k×k 0 (cid:21) (cid:20)[Q]k×k 0 (cid:21) 0 0 . = As QAQ is commutative, we have that 0 0 0 0 [QvijQ]k×k · [QwijQ]k×k = Q ⊗ Ik. Working backwards we find that Q ⊗ Id · [vij] · [wij] = Q ⊗ Id · [tij] as well and the proof of Claim 3 is complete. 8.3. Local piecewise conjugacy. Davidson and Katsoulis [23] introduced the notion of piecewise conjugacy for multivariable classical systems. Let X and Y be locally compact Hausdorff spaces and suppose that σ1, . . . , σd are proper continuous self-mappings of X, and τ1, . . . , τd are proper continuous self-mappings of Y . The classical systems (X, σ) ≡ (X, σ1, . . . , σd) and (Y, τ ) ≡ (Y, τ1, . . . , τd) are called piecewise conjugate if there is a homeo- morphism γs : Y → X, and for every y ∈ Y there is a permutation π ∈ Sd and a neighbour- hood Uπ such that γsτiUπ = σπ(i)γsUπ for all i = 1, . . . , d. Equivalently there is an open cover {Uπ π ∈ Sd} of Y such that the above equation holds. By local compactness we can substitute the Uπ by open subsets Vπ that are relatively compact and Vπ ⊆ Vπ ⊆ Uπ. One of the breakthrough results of Davidson and Katsoulis [23, Theorem 3.22] is that algebraic homomorphism of the tensor algebras associated with the dynamics implies piece- wise conjugacy of the dynamics. In the appendix we include an alternative proof of [23, Theorem 3.22]. We replace the nest representations of [23, 27] by the simpler compressions of the Fock representation. We hope that this analysis will shed light to a general converse of [23, Theorem 3.22] as well as will serve to clarify points as the following. 44 Remark 8.4. Davidson and Katsoulis [23, Definition 1.2] consider the algebra A(X, σ) generated by si and C0(X) separately, where the si form a row contraction and f · si = sif · σi. The algebra A(X, σ) is called the tensor algebra in [23]. It is shown that A(X, σ) is completely isometrically isomorphic to the tensor algebra T +E of a C*-correspondence E i=1 Ei where every Ei = C0(X) becomes a C*- [23, Theorem 2.10]. correspondence over C0(X) by defining In particular E = (cid:80)d f · ξ · h = (ξh)f σi and (cid:104)ξ, η(cid:105) = ξ∗η. However in general T +E is generated by sif and there is no apparent reason why the si can be isolated from the f ∈ C0(X) when X is not compact. What holds is that T +E ⊆ A(X, σ). However in the C*-correspondences literature the tensor algebra of (X, σ) is T +E and not A(X, σ). The latter could be seen as the tensor algebra of the one-point compactification of (X, σ). It is not clear to us whether isomorphism of the tensor algebras in the sense of Davidson and Katsoulis [23] implies isomorphism of the possibly smaller tensor algebras of the C*- correspondences in the sense of Muhly and Solel [63]. In the appendix we show directly that algebraic isomorphism of the tensor algebras of the C*-correspondences implies piecewise conjugacy. This alternative proof obviously works also for the tensor algebras in the sense of Davidson and Katsoulis [23]. The fact that A(X, σ) is generated by si and C0(X) separately is used in a non trivial way in [23]. Examples include [23, Lemma 3.7 and Lemma 3.9] where the characters are completely identified. Proving such remarks for T +E requires some extra work (respectively, see Claim 1 in the appendix). A second example is [23, Example 3.14] in conjunction with Claim 1 and Claim 6 of the appendix. Davidson and Roydor [27] notice that the notion of piecewise conjugacy passes naturally to the topological graphs in the sense of Katsura [49]. Let (X 0,X 1, rX , sX ) and (Υ0, Υ1, rΥ, sΥ) be topological graphs on compact spaces X 0 and Υ0. They are said to be locally conju- gate if there exists a homeomorphism γ0 : Υ0 → X 0 such that for every y ∈ Υ0 there is a neighbourhood U of y and a homeomorphism γ1 of s−1 Υ (U) onto s−1X γ0(U) such that Notice here that by definition s−1 sX γ1 = γ0sΥs−1 Υ (U ) and rX γ1 = γ0rΥs−1 Υ (U ). Υ (U) = ∅ if and only if s−1X γ0(U) = ∅. In one of the main results, Davidson and Roydor [27, Theorem 4.5] show that algebraic isomorphisms of the tensor algebras associated with compact topological graphs implies local conjugacy. We note here that classical systems form topological graphs. Therefore [27, Theorem 4.5] contains [23, Theorem 3.22] modulo the following remarks. Remark 8.5. First [23, Theorem 3.22] concerns classical systems over locally compact spaces in general. Secondly in the proof of [23, Theorem 3.22] it is shown that algebraic homomorphisms are automatically continuous. An analogous observation in the proof of [27, Theorem 4.5] is missing. However Davidson-Roydor make essential use of the fact that the isomorphism is bounded, see for example [27, Last paragraph of page 1257]. Whether algebraic isomorphisms of the tensor algebras are automatically continuous is something unknown to us. In fact even for our case we are able to show that this holds only under an assumption on the graph of the ideal (see Proposition 8.16 that will follow). Therefore one may consider the isomorphisms in the results of [27] to be bounded. 45 In Section 5.3 we observed that the C*-correspondence associated with a monomial ideal can be realised as the C*-correspondence of a topological graph. Hence the investigation of Davidson-Roydor [27] applies to our case. Our purpose however in the sequel is to identify the translation of local conjugacy in terms of our original data as well as to expose the implicit structure via the alternative proof that we offer. Recall from Section 4.4 that the quantised dynamics (A, α) of the monomial ideal I (cid:67) C(cid:104)x1, . . . , xd(cid:105) correspond to continuous mappings ϕi : ΩiI → ΩI when identifying A with C(ΩI). Similarly we denote by (ΩJ , ψ) the dynamical system associated with the monomial ideal J (cid:67) C(cid:104)y1, . . . , yd(cid:105). For the following discussion we identify the projections P[n] and Q[m] with the clopen set in the spectrum of which each one is a characteristic function. Definition 8.6. We say that the systems (ΩI, ϕ) and (ΩJ , ψ) are Q-P -locally piecewise conjugate if there exists a homeomorphism γs : ΩJ → ΩI, and for every y ∈ P[n] there is a neighbourhood U ⊆ P[n] of y and an [m] ∈ {0, 1, . . . , 2d− 1} such that supp[m] = supp[n], γs(U) ⊆ Q[m], and for a bijection π : supp[m] → supp[n]. γsϕiU = ψπ(i)γsU , Equivalently every P[n] ⊆ ΩJ has an open cover {Uπ}π indexed by the one-to-one corre- spondences π : supp[n] → {1, . . . , d} such that γs(Uπ) ⊆ Q[m] for all [m] with supp[m] = π(supp[n]), and γsτπ(i)Uπ = ϕiγsUπ . We do not exclude the case where Uπ = ∅ for some π. Remark 8.7. The terminology we use follows the geometric observation that when restrict- ing to P[n] then (locally) the systems look piecewise conjugate. The reason why the equation supp[m] = supp[n] appears in the definition is actually implemented by Proposition 8.1. We include the symbols P and Q in the definition to emphasise the use of the particular projections. Remark 8.8. Definition 8.6 suggests in particular that γs(P0) = Q0 for Q-P -locally piece- Indeed for y ∈ P0 the only [m] in {0, 1, . . . , 2d − 1} that has wise conjugate systems. supp[m] = 0 is [m] = 0, hence γs(y) ∈ Q0. Thus we also obtain that γ(A0) = B0 for the ∗-isomorphism γ : A → B implemented by γs : ΩJ → ΩI. Furthermore we can extend trivially the equality γsϕiU = ψπ(i)γsU from the i ∈ supp[n] to all the i ∈ {1, . . . , d} by extending π to an arbitrary permutation of {1, . . . , d}. Indeed let γ : A → B be the ∗-isomorphism implemented by γs. By definition i /∈ supp[n] if and only if π(i) /∈ supp[m] since π(supp[n]) = supp[m]. For all y ∈ P[n] there exists a neighbourhood U of y such that γs(U) ⊆ Q[m]. This implies that γ−1(P )T ∗ π(i)Tπ(i) = 0 for all subprojections P ⊆ P[n] and i /∈ supp[n]. Thus γ−1(P[n])T ∗ π(i)Tπ(i) = 0 when i /∈ supp[n]. Now we have to verify that βiγ(a)P[n] = P[n]γαπ(i)(a) for all a ∈ A; equivalently that i γ(a)Ri · R∗ R∗ π(i)aTπ(i)) i RiP[n] = P[n]γ(T ∗ π(i)Tπ(i)) · γ(T ∗ for all a ∈ A and i ∈ {1, . . . , d}. This holds trivially for all i /∈ supp[n] since both sides are zero, and it holds by definition when i ∈ supp[n]. In particular the equation βiγ(·)P[n] = P[n]γαπ(i)(·) is redundant for [n] = 0. The definition of local conjugacy passes to the systems implemented by (A0, α) and (B0, β) as well by restricting to the [n] (cid:54)= 0 and the [m] (cid:54)= 0. In this case we write (Ω0,I, ϕ) and 46 (Ω0,J , ψ) for the induced classical systems. There is a strong connection with (ΩI, ϕ) and (ΩJ , ψ). Proposition 8.9. Let I (cid:67) C(cid:104)x1, . . . , xd(cid:105) and J (cid:67) C(cid:104)y1, . . . , yd(cid:105) be monomial ideals. Then the corresponding systems (ΩI, ϕ) and (ΩJ , ψ) are Q-P -locally piecewise conjugate if and only if the systems (Ω0,I, ϕ) and (Ω0,J , ψ) are Q-P -locally piecewise conjugate. Proof. For the forward implication, Remark 8.8 gives that the isomorphism γ : A → B implemented by γs : ΩJ → ΩI is such that γ(A0) = B0. For the converse it suffices to isomorphism(cid:101)γ : A → B. Then Remark 8.8 provides the appropriate context to obtain the show that the ∗-isomorphism γ : A0 → B0 implemented by γs : Ω0,J → Ω0,I extends to an locally conjugate relations. To this end recall that A0 ⊆ C∗(T ) with identity e = IFX − P0 by Lemma 4.4. Similarly we have f = IFY − Q0 for the identity of B0 ⊆ C∗(R), and so γ(e) = f . We extend γ to the unitization so that(cid:101)γ(IFX ) = IFY . However A0 + C · IFX = A and B0 + C · IFX = B and by construction The extension(cid:101)γ is then the required ∗-isomorphism. (cid:101)γ : A0 + C · IFX → B + C · IFX R0 = IFY + f = γ(IFX ) + γ(e) = γ(P0). we obtain The next proposition provides the appropriate link with [27]. Proposition 8.10. Let I (cid:67) C(cid:104)x1, . . . , xd(cid:105) and J (cid:67) C(cid:104)y1, . . . , yd(cid:105) be monomial ideals. Then the topological graphs (Υ0I, Υ1I, rI, sI) and (Υ0J , Υ1J , rJ , sJ ) are locally conjugate if and only if the systems (ΩI, ϕ) and (ΩJ , ψ) are Q-P -locally piecewise conjugate. Proof. The converse implication is straightforward by applying the definitions. For the forward implication suppose that the topological graphs are locally conjugate and fix a y ∈ P[n] for some [n] ∈ {0, 1, . . . , 2d − 1}. By setting γs = γ0 we get the homeomorphism γs : ΩJ → ΩI. Fix a point y ∈ ΩJ . Then γs(y) ∈ Q[m] for some [m] ∈ {0, 1, . . . , 2d − 1}. s (Q[m]) and P[n] we may assume that the neighbourhood U of y By intersecting with γ−1 appearing in the definition of the local conjugacy satisfies both U ⊆ P[n] and γs(U) ⊆ Q[m]. Furthermore we obtain that the spaces s−1J (U) = {(i, u) i ∈ supp[n], u ∈ U} = (cid:116)i∈supp[n]U and s−1I γs(U) = {(j, γs(u)) j ∈ supp[m], u ∈ U} = (cid:116)j∈supp[m]γs(U) are homeomorphic by γ1. This implies that supp[n] = ∅ if and only if supp[m] = ∅ thus γs(y) ∈ Q0 for every y ∈ P0. In this case the local conjugacy is verified trivially. Hence from now on we turn our attention to the case where [n] (cid:54)= 0, for which we have that [m] (cid:54)= 0 as well. If γ1(i, u) = (j, γs(u(cid:48))) for some j ∈ supp[m] and some u(cid:48) ∈ U, then we obtain that γs(u) = γ0sJ (i, u) = sIγ1(i, u) = sI(j, γs(u(cid:48))) = γs(u(cid:48)), hence γ1(i, u) = (j, γs(u)) for some j ∈ supp[m]. Our aim is to show that for every fixed i we get that the equation γ1(i, u(cid:48)) = (ji, γs(u(cid:48))) holds for all u(cid:48) ∈ U and with the same ji. 47 That is, every copy of U in s−1J (U) maps exactly onto a copy γs(U) of s−1I γs(U). Then we get a bijection π : i (cid:55)→ ji of supp[n] onto supp[m] which for all u ∈ U implies the required γsψi(u) = γ0rJ (i, u) = rIγ1(i, u) = rI(ji, γs(u)) = ϕjiγs(u) = ϕπ(i)γs(u). To achieve this we will further substitute U by a V ⊆ U in the following way. Recall that γ1{1}×U is a homeomorphism onto its image. If γ1(1, y) = (i1, γs(y)) then select V1 ≡ W1 ⊆ U such that γ1({1} × V1) ⊆ {i1} × γs(U). This can be achieved by considering (the projection of) the open set ({1} × U) ∩ γ−1 1 ({i1} × γs(U)). Now repeat the arguments above for V1 in place of U. If γ1(2, y) = (i2, γs(y)) then select V2 ⊆ U such that Restrict further to W2 := V1∩V2 and repeat for W2 in place of U. By construction we obtain γ1({2} × V2) ⊆ {i2} × γs(U). γ1({1} × W2) ⊆ {i1} × γs(U) and γ1({2} × W2) ⊆ {i2} × γs(U). Proceed inductively. Without loss of generality we may assume that supp[n] = {1, 2, . . . , N}, and supp[m] = {1, 2, . . . , N(cid:48)}. The bijection i ↔ ji will be obtained after N steps, unless N(cid:48) (cid:54)= N . But if N > N(cid:48) then there would be a step N(cid:48) + 1 for which we would have that γ1({N(cid:48) + 1} × WN(cid:48)) ⊆ ∪N(cid:48) j=1{ij} × γs(WN(cid:48)) = γ1(∪N(cid:48) j=1{j} × U), thus {N(cid:48) + 1}×WN(cid:48) intersects some {j}×U for some j = 1, . . . , N(cid:48) which is a contradiction; hence N ≤ N(cid:48). On the other hand if N < N(cid:48) then we would have that the inverse image j=1{j} × WN and hence by composing with γ1 we of {N(cid:48)} × γs(WN ) under γ1 intersects ∪N derive that {N(cid:48)} × γs(WN ) intersects some of the {j} × γs(WN ) for j = 1, . . . , N , which is again a contradiction. Thus N(cid:48) = N and the pairing is the one claimed. Remark 8.11. Local conjugacy is easy to illustrate for monomial ideals of finite type. In this case the C*-algebra A is finite because of Proposition 4.9. Then the dynamical system (Ω, ϕ) can be represented by a graph with vertices being the points of Ω, and from every ω ∈ Ω there is an edge to ϕi(ω) if and only if ω ∈ Ωi. That is, every vertex ω ∈ Q[m] has {ei}i∈supp[m] emitting edges, so that if s(ei) = ω then r(ei) = ϕi(ω). We call the resulting graph GI the graph of I. Then local conjugate systems of finite type monomial ideals correspond to isomorphic graphs of the monomial ideals. 8.4. Bounded isomorphisms. The spaces ΩI and ΩJ are compact and have dimension 0. Hence the results on topological graphs from [27] apply in our case. Nevertheless we show that such a corollary can be derived by applying locally our alternative proof of [23, Theorem 3.22]. The reader is referred to the appendix which contains the alternative proof of [23, Theorem 3.22] and the required elements for obtaining the following corollary. Corollary 8.12. Let I (cid:67) C(cid:104)x1, . . . , xd(cid:105) and J (cid:67) C(cid:104)y1, . . . , yd(cid:48)(cid:105) be monomial ideals. Let AEA and BFB be the C*-correspondences associated with I and J , respectively. Furthermore let (ΩI, ϕ) and (ΩJ , ψ)) be the corresponding quantised dynamics. The following are equivalent: (i) T + (ii) T + (iii) E and T + E and T + (ΩI, ϕ) and (ΩJ , ψ) are Q-P -locally piecewise conjugate; F are completely isometrically isomorphic; F are isomorphic as topological algebras; 48 (iv) E and F are unitarily equivalent. If E0 and F0 are the minimal C*-correspondences associated with I and J , and (Ω0,I, ϕ) and (Ω0,J , ψ) are the corresponding quantised dynamics, then any of the items above is equivalent to any of the items below: E0 and T + and T + (v) T + (vi) T + (vii) (viii) E0 and F0 are unitarily equivalent. F0 are completely isometrically isomorphic; are isomorphic as topological algebras; E0 (Ω0,I, ϕ) and (Ω0,J , ψ) are Q-P -locally piecewise conjugate; F0 Proof. For the implication [(iii) ⇒ (iv)] we will construct a matrix [bij] that provides an invertible map between E and F . Let the projection P[n] for some [n] (cid:54)= 0. Without loss of generality we may assume that P[n] = R∗ k+1Rk+1) . . . (I − R∗ kRk(I − R∗ 1R1 . . . R∗ dRd). Then the P[n] is paired with some Q[m] with k = supp[n] = supp[m]. Since ΩJ is totally disconnected, so is P[n]. Hence we can replace the open cover of P[n] (given by the piecewise conjugacy on P[n]) with a cover by compact subsets. Now we can proceed as in the second part of the proof of [23, Proposition 3.20] to show that there is a partition into clopen sets Vπ such that γsϕiVπ = ψπ(i)γsVπ . As in [23, Corollary 3.28] let the k × k matrix [bij,[n]] with bij,[n] = δi,π(i)χVπ which intertwines and {ϕj}j∈supp[m]. {ψi}i∈supp[n] By direct computation we obtain [bij,[n]]∗[bij,[n]] = P[n] ⊗ Ik = [bij,[n]][bij,[n]]∗. Then the matrix [bij] = P[n] ⊗ Ik · [bij,[n]] d(cid:88) [n]=1 defines the required unitary C*-correspondence map. The implications [(iv) ⇒ (i) ⇒ (ii)] are immediate. To show that item (ii) implies item (iii) fix a bounded algebraic isomorphism Φ : T + F . Then Φ implies a homeomorphism between the character spaces, say γc. The restriction of every character of T + F to B is a point evaluation. Our first objective is to show that γc induces a homeomorphism γs : ΩJ → ΩI by collapsing the characters into the single points. To this end recall that the algebra T + is generated by the Ti and the a ∈ A separately, and that aTi = Tiαi(a). These are the requirements for applying [23, Lemmas 3.7 and 3.9] to obtain that the maximal analytic sets Bω in the character space are parameterized by the ω ∈ ΩI so that E → T + E Bω := {θ ∈ MT + θC(ΩI ) = evω}. E Furthermore each Bω is homeomorphic to a ball of dimension equal to the number of the ϕi for which ϕi(ω) = ω. Since γc maps maximal analytic sets onto maximal analytic sets, then the required γs is well defined (and a homeomorphism). In particular we get that if θ is a character of T + F such that θB = evy then θΦA = evx with γs(y) = x. 49 Fix y in the support of the P[n] with k = supp[n]. Here we do not exclude the case k = 0. Without loss of generality we may assume that kRk(I − R∗ 1R1 . . . R∗ P[n] = R∗ k+1Rk+1) . . . (I − R∗ dRd). Define the orbit representation πy : C(ΩJ ) → B(Ky) where Ky = ⊕µ∈M∗C by Define also Vy,i ≡ Vy(Ri) : Ky → Ky by πy(g) = diag{gψµ(y) µ ∈ M∗}. Vy(Ri)eν = eiν 0 if iν ∈ M∗, otherwise. (cid:40) Then (Vy × πy) defines a completely contractive representation of T + F . Let E denote the compression onto the subspace generated by {e∅, e1, . . . , ed(cid:48)} and let {Eij i, j = 0, 1, . . . , d(cid:48)} denote the standard matrix basis in Md(cid:48)(C). Then E(Vy × πy)(Rig) = gψi(y)Ei0 for all i = 1, . . . , k, whereas E(Vy × πy)(Rig) = 0 for all i = k + 1, . . . , d(cid:48). The former holds by definition and the latter holds since E(Vy × πy)(Rig) = E(Vy × πy)(RiR∗ (cid:32) 0 ... 0 i Rig) = E(Vy × πy)(Rig · R∗ = E(Vy × πy)(Rig)E(Vy × πy)(R∗ i Ri) (cid:33) i Ri) = gψi(y)Ei0 0 · E00 + kiEii = 0, d(cid:48)(cid:88) i=1 i Ri along the orbit of y. Consequently if we write Φ(Q) = g0 + ((cid:80)d(cid:48) where the ki are some values of the function R∗ i=1 wigi) + Z for Q ∈ T + 0 ... g0(y) ... E(Vy × πy)Φ(Q) = gk(y) E then we get 0 ... 0 g0ψk(y)  . . . . . . . . . . . . ... . . . 0 ... 0 . . . . . . . . . . . . . . . . . . 0 ... 0 0 ... g0ψk+1(y) ... 0 g0ψd(cid:48)(y)  . By compressing even further to the subspace generated by {e∅, e1, . . . , ek} we have set the right context to apply locally the ideas from the appendix. By proceeding as in Claim 3 of the appendix and thereon we derive item (iii). Note that if y ∈ P0 then we cannot have that γs(y) ∈ Q[m] for any [m] (cid:54)= 0. Therefore γs(y) ∈ Q0. Similar arguments show that the items (v), (vi), (vii) and (viii), are equivalent. Then Proposition 8.9 gives the equivalence of item (vii) with item (iii) and the proof is complete. Question 8.13. We ask whether Corollary 8.12 holds at the level of algebraic isomorphisms between T + E and T + F . 50 The next example shows that the commutative C*-algebras A and B cannot distinguish between the quantised dynamics related to the ideals I and J . Neither can the C*-modules EA and FB. It is only after taking into consideration all this structure and the left action that we can distinguish between the quantised dynamics at hand. Example 8.14. Consider the ideals I = (cid:104)x1x2, x2x1(cid:105) and J = (cid:104)x2 2(cid:105) in C(cid:104)x1, x2(cid:105), as in Example 4.15. Let T1, T2 (resp. R1, R2) be the operators corresponding to I (resp. to J ) and AEA (resp. BFB) be the associated C*-correspondence. Then T ∗ 2 T2 is the projection on the span of {e∅, e2n}. For every µ (cid:54)= ∅, we find that T ∗ 1 T1 if µ has no 2's in it, µ Tµ = 0 otherwise. Similarly, when µ (cid:54)= ∅ then µ Tµ = T ∗ T ∗ µRµ is the projection onto the span of {e∅, e1, e12, e121, . . .} if µ is an alternating sequence R∗ of 1's and 2's ending with 2. On the other hand R∗ µRµ is the projection onto the span of {e∅, e2, e21, e212, . . .} if µ is an alternating sequence of 1's and 2's ending with 1; and it is 0 otherwise. and R∗ inner product map (γ, U ) given by 1 T1 is the projection on the span of {e∅, e1n n ≥ 1}, and T ∗ 2 T2 if µ has no 1's in it, and T ∗ Then we may identify both A and B with the continuous functions on {0, 1, 2}, with T ∗ 1 T1 2R2 corresponding to χ{0,2}. The 1R1 corresponding to χ{0,1} and with T ∗ 2 T2 and R∗ µ Tµ = T ∗ 1, x2 and the ∗-isomorphism U (λT1 + kT2) = λR1 + kR2 1R1 + kR∗ is a unitary equivalence of the Hilbert C*-modules E and F . 2 T2) = λR∗ 1 T1 + kT ∗ γ(λT ∗ 2R2 However, these two Hilbert C*-modules are not unitarily equivalent as C*-correspondences. Indeed, the quantised dynamical systems of this example were calculated in Example 4.15, and we see that the resulting graphs obtained by Remark 8.11 are not isomorphic. Let us also provide the following direct operator theoretic argument to show this. Write where b1j ∈ R∗ U (Tj) = R1b1j + R2b2j for j = 1, 2, 1R1B and b2j ∈ R∗ γ(T ∗ 2R2B. We must also have that j Tj) = (cid:104)U Tj, U Tj(cid:105) = b∗ 1jb1j + b∗ 2jb2j. Furthermore the left covariance of U means that 1 T1 · T2) = γ(T ∗ 0 = U (T ∗ 1 T1) (R1b12 + R2b22) . It must be that γ(T ∗ 1 − R∗ 1R1R∗ 2R2. j Tj) is the sum of two minimal projection, so it is one of R∗ 1R1, R∗ 2R2 or 1 T1) = R∗ First assume that γ(T ∗ 1R1. Since 1R1 · R1 = R1β1(R∗ R∗ 1R1) = R1R∗ 1R∗ 1R1R1 = 0, and 1R1 · R2 = R2β2(R∗ R∗ 1R1) = R2R∗ 2R∗ 1R1R2 = R2, we must have that b22 = 0. Therefore we get that γ(T ∗ 2 T2) = b∗ 12b12 ∈ R∗ 1R1B. 51 2 T2) is either R∗ However γ(T ∗ contradiction shows that γ(T ∗ 1R1R∗ 2R2 or 1 − R∗ 1 T1) cannot be R∗ R1 ↔ R2 and R∗ 2R2, neither of which is in R∗ 1R1. On the other hand the flips 1R1 ↔ R∗ 2R2 1R1B. This extend to an automorphism of F since J is symmetrical. Therefore γ(T ∗ either. Finally we have that γ(T ∗ 2 T2) should be either R∗ case γ(T ∗ 1 T1) = 1 − R∗ 1R1 or R∗ 2R2. This is impossible by symmetry. 2R2, 2R2 is also not an option. Indeed in this 1 T1) cannot be R∗ 1R1R∗ 8.5. Automatic continuity. We are able to answer Question 8.13 for a specific class of monomial ideals, for which algebraic isomorphisms onto their tensor algebras are automat- ically continuous. The arguments follow a trick that Donsig, Hudson and Katsoulis [29] established following ideas of Sinclair [81]. Let φ : A → B be an algebraic epimorphism for the Banach algebras A and B. The discontinuity of φ is quantified by the ideal S(φ) := {b ∈ B ∃(an) ⊆ A such that an → 0 and φ(an) → b}. By the closed graph theorem then φ is continuous if and only if S(φ) = (0). Lemma 8.15 (Sinclair). Let φ : A → B be an algebraic epimorphism between the Banach algebras A and B, and let a sequence (bn) in B. Then there exists an N ∈ N such that b1b2 . . . bNS(φ) = b1b2 . . . bnS(φ) and S(φ)bnbn−1 . . . b1 = S(φ)bN bN−1 . . . b1 for all n ≥ N . Proposition 8.16. Let E be a C*-correspondence associated to a monomial ideal I in C(cid:104)x1, . . . , xd(cid:105). Suppose that for every ν ∈ Λ∗ there are words w, z ∈ Λ∗ such that wnzν ∈ Λ∗ for all n ∈ N. Then an algebraic epimorphism φ : A → T + E for any Banach algebra A is continuous. Proof. To reach contradiction let 0 (cid:54)= x ∈ S(φ). By the Fourier transform we may assume that x = Tµa for some µ ∈ Λ∗ and a ∈ A. Moreover we take µ to be of minimal length, i.e. if Tνa ∈ S(φ) for ν < µ then Tνa = 0. Let ν1, ν2 ∈ Λ∗ such that (cid:104)Tµaeν1, eν2(cid:105) (cid:54)= 0. By assumption let w, z ∈ Λ∗ such that wnzν2 ∈ Λ∗ for all n ∈ N. Then we obtain wnewnzν2(cid:105) = (cid:104)Tµaeν1, eν2(cid:105) (cid:54)= 0. wTzµaeν1, ewnzν2(cid:105) = (cid:104)Tµaeν1, T ∗ (cid:104)T n wTzµa (cid:54)= 0 for all n ∈ N and in particular Tzµa is a non-trivial element in Consequently T n S(φ). By applying Lemma 8.15 for bn = Tw we get that wS(φ) w Tzµa ∈ T n T N for all n ≥ N . In particular this is true for n = N +z + 1. However the Fourier coefficients wS(φ) are supported on words with length at least N + z + 1 + µ, of the elements in T n w Tzµa is supported on a word with length exactly N + z + µ. This leads to the whereas T N contradiction T N z T ∗ w Tzµa = 0. Remark 8.17. The assumption in Proposition 8.16 can be verified for monomial ideals that are either of finite or infinite type. An example in the case of infinite type is provided by the monomial ideal generated by {xynx n ∈ N} of C(cid:104)x, y(cid:105). 52 On the other hand there are monomial ideals of infinite type for which the assumption in Proposition 8.16 does not hold. For example suppose that the monomial ideal in C(cid:104)x, y(cid:105) is such that the allowable words are either of the form xyxy2xy3 . . . ynx , xyxy2xy3 . . . xyn , and for all n ∈ Z+, or sub-words of these words. Question 8.18. In view of Question 8.13 we ask whether the assumption in Proposition 8.16 holds for all monomial ideals of finite type. yxyx2yx3 . . . xny , yxyx2yx3 . . . yxn , 8.6. Local conjugacy and C*-algebras. Next we show how local conjugacy affects the related C*-algebras. Recall that local conjugacy coincides with unitary equivalence by Corol- lary 8.12. By the remarks preceding Proposition 2.2 we have that local conjugacy implies ∗-isomorphisms of the Toeplitz-Pimsner and the Cuntz-Pimsner algebras, as well as complete isometric isomorphisms of the tensor algebras. Below we show that the same is true for the C*-algebras C∗(T ) and C∗(R), and their quotients by the compacts. Corollary 8.19. Let I (cid:67) C(cid:104)x1, . . . , xd(cid:105) and J (cid:67) C(cid:104)y1, . . . , yd(cid:48)(cid:105) be monomial ideals. If the quantised dynamics are Q-P -locally piecewise conjugate then C∗(T ) (cid:39) C∗(R), and moreover C∗(T )/K(FX) (cid:39) C∗(R)/K(FY ). Proof. Let us denote by tµ and rν the generators in T + F . Local conjugacy implies a unitary equivalence (γ, U ) : E → F , in which case we get that d = d(cid:48) by Proposition 8.1. Let [bij] be the associated matrix and let the ∗-isomorphism Φ : TE → TF such that Φ(a) = γ(a) for all a ∈ A, and Φ(tj) = ribij for all j = 1, . . . , d. By Proposition 6.3 the C*-algebras C∗(T ) and C∗(R) are respectively quotients of TE and TF . By Proposition 2.2 then Φ implements a ∗-isomorphism Φ(cid:48) : C∗(T ) → C∗(R) if it maps the wrw w ∈ M∗} ideal generated by {I − t∗ in B. µtµ µ ∈ Λ∗} in A into the ideal generated by {I − r∗ E and T + d(cid:88) i=1 For µ = j ∈ {1, . . . , d} we compute j tj) = I − d(cid:88) k,i=1 ijr∗ b∗ Φ(I − t∗ However by unitary equivalence we have that r∗ Φ(I − t∗ j tj) = I − r∗ j rj + i=1 i rkbkj = I − d(cid:88) j rj =(cid:80)d ijbij − d(cid:88) d(cid:88) d(cid:88) i=1 b∗ ij (I − r∗ b∗ b∗ i=1 i=1 i=1 b∗ ijr∗ i ribkj. ijbij, therefore ijr∗ b∗ i ribij = I − r∗ j rj + i ri) bij. Hence Φ(I − t∗ j tj) is in the required ideal. If µ = jl then we may write l (I − t∗ I − t∗ l tl − t∗ j tjtl = I − t∗ l t∗ jltjl = I − t∗ l tl + t∗ l tl + t∗ j tj)tl. 53 It suffices to restrict our attention to t∗ l (I − t∗ j tj)tl. We then compute Φ(t∗ l (I − t∗ j tj)tl) = Φ(tl)∗Φ(I − t∗ j tj)Φ(tl) = b∗ ilr∗ i xrkbik, d(cid:88) i,k=1 wrw)c for some b, c ∈ B and some words w ∈ M∗. By where x is the sum of elements b∗(I − r∗ using the covariant relation we get that wrw)cribik = b∗ = −b∗ i b∗(I − r∗ ilr∗ b∗ i ri − r∗ ilβi(b)∗(r∗ i r∗ wrwri)βi(c)bik ilβi(b)∗(I − r∗ i ri)βi(c)bik+ + b∗ ilβi(b)∗(I − r∗ l (I − t∗ wirwi)βi(c)bik j tj)tl) is in the required ideal. Therefore Φ(I − t∗ which shows that Φ(t∗ jltjl) is in the required ideal. Induction on the length of the words µ then finishes the proof of C∗(T ) (cid:39) C∗(R). On the other hand Proposition 6.3 implies that the quotient C∗(T )/K(FX) is the A-relative Cuntz-Pimsner algebra of E. Trivially γ(A) = B and Proposition 2.2 finishes the proof of the second part. Remark 8.20. A similar analysis can be carried out for the q(E) and q(F ) of Remark 5.10. The reason is that the unitary equivalence mappings are defined by a family of ∗-algebraic relations. Φ : C∗(T ) → C∗(R) is such that Φ(a) = γ(a) and Φ(tj) =(cid:80)d Φ(cid:48)(q(a)) = q(γ(a)) and Φ(cid:48)(q(tj)) = (cid:80)d ∗-algebraic relations, e.g. (cid:80)d In particular we have that if E and F are unitarily equivalent then q(E) and q(F ) are unitarily equivalent as well. Indeed if E and F are injective then this is immediate by the E is completely isometric and because the ∗-isomorphism fact that the restriction of q to T + i=1 ribij. If E is not injective then so is F and in particular γ(ker φE) = ker φF . However Proposition 5.8 implies that ker φE = CP∅ and ker φF = CP∅. Consequently we obtain that γ(P∅) = P∅. Therefore Φ(K(FX)) = K(FY ) and C∗(T )/K(FX) is ∗-isomorphic to C∗(R)/K(FY ) by some Φ(cid:48) with i=1 q(ri)q(bij). Then Φ(cid:48)q(A) is a ∗-isomorphism onto q(B) and the q(bij) form a matrix that corresponds to a C*-correspondence map. The fact that [bij] defines a unitary C*-correspondence map between E and F is translated into some i ti. Applying Φ(cid:48) to these we obtain a set of ∗-algebraic relations which give that the C*-correspondence map associated with [q(bij)] is unitary. jk = δi,jt∗ k=1 bikb∗ The converse of this phenomenon does not hold. For a counterexample let E be associated with the monomial ideal I generated by(cid:18)  (cid:19)  x1x2 x2 2 x2x1 x1x2 x1x3 x2 x2x3 x2x1 2 x2 x3x1 x3x2 3 in C(cid:104)x1, x2(cid:105), and let F be associated with the monomial ideal J in C(cid:104)x1, x2, x3(cid:105). Then E and F cannot be unitarily equivalent as they are related with rings on a different number of symbols. However q(E) and q(F ) are both C over C since there 54 exists only one infinite word in both cases, i.e. the sequence (x1), and so they are trivially unitarily equivalent. 9. Encoding via subproduct systems We turn our attention to the classification of our data by using the subproduct systems. Our aim is to show that subproduct systems and their tensor algebras form a complete invariant for monomial ideals up to a permutation of the generators. We will require some useful algebraic facts. Lemma 9.1. Let I be a monomial ideal of C(cid:104)x1, . . . , xd(cid:105). Then the group of the graded automorphisms of C(cid:104)x1, . . . , xd(cid:105) /I is a linear algebraic group. Proof. Without loss of generality we may assume that I does not contain any of the xi. Otherwise we restrict our attention to C(cid:104)x1, . . . , xn(cid:105) /I(cid:48) where x1, . . . , xn /∈ I, xn+1, . . . , xd ∈ I and I(cid:48) = I ∩ C(cid:104)x1, . . . , xn(cid:105). Let zi = xi + I be the generators of the quotient. Then the embedding φ (cid:55)→ [aij], where φ(zj) = aijzi, d(cid:88) i=1 from the graded automorphisms of C(cid:104)x1, . . . , xd(cid:105) /I inside GLd(C) is an injective group homomorphism. Injectivity is immediate as the zi generate the quotient. To see that [aij] is indeed invertible, recall that φ is graded hence its restriction to the linear span of the zi is onto itself. Then we get and at the same time d(cid:88) d(cid:88) i=1 f ( f ( ai1xi, . . . , bi1xi, . . . , d(cid:88) d(cid:88) i=1 i=1 i=1 aidxi) = 0 for all f ∈ I, bidxi) = 0 for all f ∈ I, where [bij] = [aij]−1. On the other hand, if [aij] ∈ GLd(C) satisfies the above equations then it readily defines an automorphism of C(cid:104)x1, . . . , xd(cid:105) /I. Let the ring (C[y11, . . . , ydd])(cid:104)x1, . . . , xd(cid:105) of polynomials on the noncommuting variables xi with co-efficients from C[y11, . . . , ydd] so that yij · xk = xk · yij. Then for any polynomial f ∈ I we have that d(cid:88) d(cid:88) (cid:88) f ( yi1xi, . . . , yidxi) = i=1 i=1 xµ /∈I , µ≤deg f gf,µ(y11, . . . , ydd)xµ for some polynomial expressions gf,µ in C[y11, . . . , ydd]. Since the xµ form a basis we obtain that d(cid:88) d(cid:88) f ( yi1xi, . . . , yidxi) = 0 ⇐⇒ gf,µ(y11, . . . , ydd) = 0 for all gf,µ. i=1 i=1 55 Therefore we have that an invertible matrix [aij] satisfies f ( ai1xi, . . . , aidxi) = 0 for all f ∈ I, d(cid:88) i=1 d(cid:88) i=1 d(cid:88) i=1 d(cid:88) i=1 if and only if the tuple (a11, . . . , add) ∈ Cd2 is a solution for the system of polynomials {gf,µ f ∈ I, xµ ∈ I}. The latter set may be infinite. However the gf,µ are polynomials in C[y11, . . . , ydd], hence we can find a finite set of such polynomials with the same set of solutions. In a similar way if [bij] = [aij]−1 then f ( bi1xi, . . . , bidxi) = 0 for all f ∈ I, if and only if the (b11, . . . , bdd) ∈ Cd2 is a solution for the system of polynomials {g(cid:48) f,µ f ∈ I, xµ ∈ I}. This can be substituted again by a finite set of monomials with the same set of solutions. Recall that the bij are polynomial expressions of the aij. Hence we may view the g(cid:48) f,µ as polynomials on the y11, . . . , ydd as well. Therefore we have that an invertible matrix [aij] defines a graded automorphism of C(cid:104)x1, . . . , xd(cid:105) /I if and only if the d2-tuple (aij) of its entries forms a solution for a finite set of polynomials, and the proof is complete. Now we are in position to apply the arguments of [16, Theorem 5.27] to achieve the required classification. Theorem 9.2. Let X and Y be subproduct systems associated with the homogeneous ideals I (cid:67) C(cid:104)x1, . . . , xd(cid:105) and J (cid:67) C(cid:104)y1, . . . , yd(cid:48)(cid:105). Without loss of generality suppose that xi /∈ I and yj /∈ J for all i, j. The following are equivalent: (i) AX and AY are completely isometrically isomorphic; (ii) AX and AY are algebraically isomorphic; (iii) C(cid:104)x1, . . . , xd(cid:105) /I and C(cid:104)y1, . . . , yd(cid:48)(cid:105) /J are algebraically isomorphic by a graded iso- morphism; (iv) X and Y are similar; (v) X and Y are isomorphic; (vi) d = d(cid:48) and there is a permutation on the variables y1, . . . , yd such that I and J are defined by the same words. Proof. The implications [(vi) ⇒ (v) ⇒ (iv)], [(iv) ⇒ (iii)], [(iv) ⇒ (ii)] and [(vi) ⇒ (i)] are immediate. Moreover the implication [(i) ⇒ (v)] is shown in Theorem 3.4. To finish the proof we will show that [(ii) ⇒ (iii) ⇒ (vi)]. Suppose that item (ii) holds. Then by Lemma 3.3 we may suppose that there is a graded isomorphism φ : AX → AY . As in the proof of [31, Proposition 6.17] we get a family of isomorphisms Vn := pnφX(n) : X(n) → Y (n) where we identify X(n) with t∞(X(n)). In particular each Vn implies an isomorphism be- tween {f + I deg f = n} and {g + J deg g = n}. Applying to X(1) = Cd we obtain that d = d(cid:48). Because the family of Vn respects the associative product rule of the subprod- uct systems, it extends to a graded algebraic isomorphism between C(cid:104)x1, . . . , xd(cid:105) /I and C(cid:104)y1, . . . , yd(cid:105) /J . 56 Suppose that item (iii) holds for an algebraic isomorphism φ. Let TI be the subgroup of the graded automorphisms of C(cid:104)x1, . . . , xd(cid:105) /I defined by xi (cid:55)→ λixi, i.e. TI is a torus. Now TI is a maximal torus inside GLd(C), thus it is also a maximal torus inside the group of the graded automorphisms of C(cid:104)x1, . . . , xd(cid:105) /I. The group of automorphisms of C(cid:104)x1, . . . , xd(cid:105) /J con- tains two maximal tori; the TJ and the φTIφ−1. By Lemma 9.1 we can apply Borel's Theorem [14, Corollary 11.3 (1)] which states that all maximal tori in an algebraic group are conjugate. Hence we obtain a graded automorphism ρ of C(cid:104)x1, . . . , xd(cid:105) /J such that ρ−1TJ ρ = φTIφ−1. By substituting φ with ρφ we have that there exists a graded isomor- phism φ : C(cid:104)x1, . . . , xd(cid:105) /I → C(cid:104)x1, . . . , xd(cid:105) /J such that φTI = TJ φ. If [aij] is the invertible matrix related to φ then we obtain that for every diagonal matrix D1 there is a diagonal matrix D2 such that [aij]D1 = D2[aij]. By the Leibniz formula for the determinant there is a permutation π ∈ Sd such that j=1aπ(j)j (cid:54)= 0. Let Aπ be the corresponding permutation matrix and set [bij] = Aπ[aij]. Πd Since [aij]D1 = D2[aij] for the diagonal matrices D1 and D2 then we may write [bij]D1 = Aπ[aij]D1 = AπD2[aij] = AπD2A−1 π [bij]. The matrix AπD2A−1 π is again diagonal, as it is produced by permuting the diagonal elements of D2 by π. Hence for every λ = (λ1, . . . , λd) ∈ Cd there exists an r ≡ r(λ) = (r1, . . . , rd) ∈ Cd such that bijλj = ribij for all i, j = 1, . . . , d. Since bii (cid:54)= 0 we get that λi = ri for all i = 1, . . . , d. By choosing non-zero λi ∈ C such that λi (cid:54)= λj for i (cid:54)= j we get that bij = 0 for i (cid:54)= j. Hence the matrix [bij] is diagonal. Consequently the matrix [aij] associated with the graded isomorphism φ : C(cid:104)x1, . . . , xd(cid:105) /I → C(cid:104)x1, . . . , xd(cid:105) /J is diagonal up to a permutation π of the rows. Therefore we get that φ(xi+I) = aπ(i)iyπ(i)+J with aπ(i)i (cid:54)= 0, which completes the proof. Remark 9.3. Theorem 9.2 reads the same also for monomials in commuting variables, with the proof following verbatim. Remark 9.4. There is a direct proof of the implication [(iv) ⇒ (vi)] of Theorem 9.2 that does not pass through item (iii). By Theorem 3.4, this proof suffices also to give Theorem 9.2 for bounded isomorphisms. Recall that similarity in particular implies that d = d(cid:48). Let {e1, . . . , ed} and {f1, . . . , fd} be the canonical orthonormal bases for X(1) and Y (1), respectively. Fix a similarity V = (Vn) for which we obtain d(cid:88) i=1 V1ej = vijfi for j = 1, . . . , d. Now V1 may not be a permutation and we quantify this by defining Q(m) = {i ∈ {1, . . . , d} vim (cid:54)= 0} for m = 1, . . . , d. Thus V1 permutes the basis elements if and only if every Q(m) is a singleton. Claim. If m1 . . . mn is a forbidden word in X then i1 . . . in is a forbidden word in Y for all (i1, . . . , in) ∈ Q(m1) × ··· × Q(mn). 57 Proof of Claim. Recall from Section 2.3 that Vnpn = qnV ⊗n X(n) into Y (1)⊗n (cid:9) Y (n). Hence, if m1 . . . mn ∈ FX is spanned by some fν with ν ∈ FY corresponds a nonzero summand in this expansion, namely vi1m1 ··· vinmnfi1 ⊗ ··· ⊗ fin. 1 1 maps X(1)⊗n (cid:9) (em1 ⊗ ··· ⊗ emn) n . For every (i1, . . . , in) ∈ Q(m1) × ··· × Q(mn), there . Therefore V ⊗n n then we get that V ⊗n 1 It follows that fi1 ⊗ ··· ⊗ fin is not in Y (n) for each (i1, . . . , in) ∈ Q(m1) × ··· × Q(mn). Hence we must have that fi1 ⊗ ··· ⊗ fin ⊥ Y (n). Therefore i1 . . . in is a forbidden word, and the proof of the claim is complete. Since V1 is an isomorphism then [vij] is in GLd(C). By the Leibniz formula for the determinant there is a permutation π ∈ Sd such that viπ(i) (cid:54)= 0 for all i = 1, . . . , d. This (cid:54)= 0 up to a permutation defines an automorphism on Y , hence we may assume that vii permutation on the {y1, . . . , yd}. Then by the claim above we get that FX n for all n. This means that after applying a permutation on the {y1, . . . , yd} we have that I ⊆ J . By symmetry and simple dimensional considerations, it follows that I = J after permuting the variables. n ⊆ FY With a little more care, such arguments could also settle algebraic homomorphisms. This requires a combination with a weaker notion of similarity V = (Vn) where a uniform bound for the Vn is not provided. We leave the details to the interested reader. Remark 9.5. In Remark 9.4 we begin with any similarity and show that it induces a per- mutation of the elements. On the other hand, in the proof of [(iv) ⇒ (vi)] of Theorem 9.2 we isolate a similarity which is shown to be diagonal up to a permutation of the columns. Thus it is natural to ask whether all similarities are of this form. Equivalently if the auto- morphisms of a subproduct system X are toric up to a permutation. This is not true and here is a counterexample that settles this for more general homogeneous ideals beyond the monomial ones. Recall that if I = (0) then any unitary defines an automorphism of Ad. Therefore if I is generated by polynomials in the first symbols x1, . . . , xn from {x1, . . . , xd} with n < d, then any matrix V = In ⊕ U with U a unitary in Md−n(C) defines an automorphism of AX for X = XI. Then U can be chosen so that V is not diagonal up to a permutation. Remark 9.6. Theorem 9.2 implies that two tensor algebras AX and AY are isomorphic as algebras if and only if they are isomorphic as topological algebras. However we do not claim that every algebraic isomorphism is automatically continuous. We are able to show automatic continuity under the assumption that for every ν ∈ Λ∗ there are w, z ∈ Λ∗ such that wnzν ∈ Λ∗ for all n ∈ Z+. The proof follows verbatim the proof of Proposition 8.16. Remark 9.7. Item (iii) in Theorem 9.2 shows that rigidity of subproduct systems is con- nected to rigidity phenomena for the quotients C(cid:104)x1, . . . , xd(cid:105) /I and C(cid:104)y1, . . . , yd(cid:105) /J . This is an exceptional behaviour, and for general homogeneous ideals this is far from being true. In particular in the commutative case there are examples of ideals I and J in C[x1, . . . , xd] such that C[x1, . . . , xd]/I is isomorphic to C[x1, . . . , xd]/J , but AX is not (algebraically) isomorphic to AY . Such an example appears in [25, Example 8.6]. These examples can be pulled back to ideals in noncommuting variables by considering homogeneous ideals that contain the commutator ideal. 58 From Theorem 9.2 we derive that if AX and AY are isomorphic then T + F are iso- morphic as well, where E and F are the C*-correspondences related to I and J , respectively. Indeed equality of the monomial ideals up to a permutation produces a unitary equivalence by permuting the generators. The following counterexample shows that the converse fails. Example 9.8. Consider the monomial ideals I and J in C(cid:104)x1, x2, x3, x4(cid:105): E and T + x1x1 x1x2 x1x3 x2x1 x2x2 x3x1 x3x2 x3x3 x4x1 x4x2 x4x4 generating set for I x2x4  x1x1 x1x2 x1x3 x1x4 x2x1 x2x2 x3x1 x3x2 x3x3 x4x1 x4x2 x4x4 generating set for J  Then the resulting graphs from the quantised dynamics are: 3 3 2 1 0 4 3 12 2 4 1 graph for I 4 3 3 2 1 0 4 3 12 2 graph for J 4 2 4 where 0 = [∅]1, 12 = [1]1 = [2]1, 3 = [3]1 and 4 = [4]1. The only difference is in the right bottom arrow: in one graph it is labeled 1 (meaning that ϕ1 maps 4 to 1) and in the other it is labeled 2. The two graphs are isomorphic (they are the same when removing labels) hence the quantised dynamical systems are Q-P -locally piecewise conjugate. However, the dynamical systems are not conjugate, and there is no permutation taking I onto J . Thus T + and T + F are completely isometrically isomorphic, while AX and AY are not even algebraically isomorphic. E 10. Comparison with other constructions There are several possible ways in which to associate an operator algebra with a given monomial ideal I. In this section we show how some natural candidates (several of which have been considered in the literature) differ from the algebras that we are considering herein. In the last subsection we will show that when I comes from a sofic subshift, and E is the C*-correspondence that we have associated with I, then OE can be constructed as the graph C*-algebra of the follower set graph of the subshift. 10.1. Graph constructions. It is natural to associate a certain graph with every monomial ideal. Definition 10.1. The left graph GI,l of I (resp. the right graph GI,r of I) is the subgraph + consisting of the vertices Λ∗ and the of the directed left (resp. right) Cayley graph of Fd edges connecting them. Therefore the set of vertices of the graph GI,l is the set of legal words, and there is a directed edge from µ to ν if and only if there is some i ∈ {1, . . . , d} such that ν = iµ. We 59 w w ' '   w w ' '   , , ' ' l l w w , , ' ' l l w w 1, x2 E and T + caution the reader not to confuse the left graph with the graph GI constructed from the quantised dynamics in Remark 8.11. Quiver algebras. From every directed graph G one may construct the so-called quiver algebra T +(G) [46, 82] (quiver is just another way of saying directed graph). This is the tensor algebra of the C*-correspondence related to G. It was proved in [46] that two quiver algebras are isomorphic as topological algebras if and only if the underlying graphs are isomorphic. Therefore, a reasonable way to encode the information in I is to form the quiver algebra T +(GI,l) of the left graph GI,l of the ideal I. In general, the quiver algebra T +(GI,l) differs from the nonselfadjoint operator algebras T + E and AX. To see this, recall Example 8.14. For that example we consider the ideals 2(cid:105) in C(cid:104)x1, x2(cid:105) which give rise to the non unitarily equivalent I = (cid:104)x1x2, x2x1(cid:105) and J = (cid:104)x2 C*-correspondences E and F . Thus T + F are not isomorphic as topological algebras by Corollary 8.12. On the other hand, the graphs GI,l and GJ ,l are the same as they consist of two directed infinite paths with a common source. Thus the quiver algebras T +(GI,l) and T +(GJ ,l) are isometrically isomorphic. Therefore at least one of the quiver algebras must be different as a topological algebra from the tensor algebras. from a Banach algebra onto T + in general, the quiver algebra T +(GI,l) and the tensor algebra T + isomorphic. In fact Proposition 8.16 applies in this case to conclude that any algebraic isomorphism E is continuous. It then follows from the above argument that, E are not even algebraically Using the same example together with Theorem 9.2 we find that the tensor algebras AX do not coincide (even just as algebras) in general with the quiver algebras. The same is true when considering the right graphs due to the symmetry of I and J . Graph C*-algebras with the notation of [76]. On the other hand, from every graph G one may also construct its graph C*-algebra C∗(G). In this section we use the recent terminology as it appears in [76]. Again this algebra is different from the C*-algebras that we consider here. graph of GI,l can be drawn as follows •v−1 The example provided by I = (cid:104)x1x2, x2x1(cid:105) in C(cid:104)x1, x2(cid:105) does the job. Indeed, the left / ··· , •v−2 ··· •v0 / •v1 / •v2 f−3 f−2 f−1 f1 f2 f3 where the central vertex corresponds to the empty word. Let {en} be the standard o.n. basis of (cid:96)2(Z). Then a Cuntz-Krieger family {Pn, Lm n ∈ Z, m ∈ Z∗} is given by the following finite rank operators (cid:40) Pn = Θen,en for all n ∈ Z, and Lm = Θem−1,em Θem+1,em for m ≥ 1, for m ≤ −1, where Θx,y(z) = (cid:104)z, x(cid:105) y (in contrast to what the symbol Θ means for C*-correspondences). By the gauge invariant uniqueness theorem, this Cuntz-Krieger family integrates to a faithful representation of the graph algebra. The graph C*-algebra C∗(GI,l) is generated by compact operators hence it contains only compact operators. Since Θem,en ∈ C∗(GI,l) for all m, n ∈ Z, we find C∗(GI,l) = K((cid:96)2(Z)). On the other hand the graph algebra of C∗(T ) is irreducible and contains a non compact operator, thus it cannot be ∗-isomorphic to C∗(GI,l) (it is not CCR). The graph algebra is 60 o o o o o o / / / also not isomorphic to C∗(T )/K, as the latter is ∗-isomorphic to C(T)⊕C(T) being generated by two partial isometries v, u satisfying v∗v = vv∗ ⊥ u∗u = uu∗. The same is true when considering the graph algebra of GI,r which coincides with GI,l due to the symmetry of I. Graph C*-algebras with older notation. The notation in graph algebras has changed recently. Previously the role of the source in the Cuntz-Krieger equations was played by the range and vice versa. Here we mention that even in this case the C*-algebras are again different. Indeed, let again I = (cid:104)x1x2, x2x1(cid:105) in C(cid:104)x1, x2(cid:105) with the graph / •v2 •v−1 •v−2 / •v1 ··· •v0 f−3 f−2 f−1 f1 f2 / ··· , f3 where the central vertex corresponds to the empty word. Let {en} be the standard o.n. basis of (cid:96)2(Z), let H = (cid:96)2(Z) ⊕ C. Denote by f a unit vector in the summand C. Then a Cuntz-Krieger family {Pn, Lm n ∈ Z, m ∈ Z∗} on H is given by the following finite rank operators L−1 = Θe−1,f and Pn = Θe0,e0 + Θf,f Θen,en if n = 0, for n (cid:54)= 0, and Lm = Θem,em−1 Θem,em+1 for m ≥ 1, for m ≤ −2, (cid:40) (cid:40) In this case the graph algebra is again K ⊕ K. In particular this Cuntz-Krieger family integrates to a faithful representation of the graph algebra. Indeed, the right branch of the graph generates a C*-algebra isomorphic to K((cid:96)2(N)), and the left branch generates another copy of K((cid:96)2(N)) that is orthogonal to the first copy. Reasoning as above we arrive at the conclusion that the "old" graph C*-algebras are also different from the C*-algebras that we consider. 10.2. Dynamical systems. Let Λ be a two-sided subshift, as described in Section 4.3. Then we have the topological dynamical system (Λ, σ) defined by the left shift σ on Λ. Reasonable candidates to encode this system are the C*-crossed product C(Λ)×σ Z and the (nonselfadjoint) semicrossed product C(Λ) (cid:111)σ Z+ [71]. 10.2.1. Semicrossed product. The semicrossed product C(Λ)×σZ+ is defined as the universal nonselfadjoint operator algebra generated by vnf for f ∈ C(Λ) and n ∈ Z+, where v is a contraction such that f · v = vf σ. Consequently for any λ = (xn) ∈ Λ we have that the mapping (cid:20)f0(λ) (cid:21) Φ( vnfn) = 0 f1(λ) f0σ(λ) (cid:88) n defines a completely contractive representation of C(Λ) ×σ Z+. Furthermore conjugate sub- shifts have completely isometric isomorphic semicrossed products. A stronger converse is given by Davidson and Katsoulis [22]: algebraic isomorphism of semicrossed product im- plies conjugacy of the associated C*-dynamics. Recall that conjugate subshifts may be defined on a different number of symbols. Since the number of symbols is an invariant for the tensor algebra AX related to Λ, then AX cannot be algebraically isomorphic to C(Λ) ×σ Z+. E and C(Λ) ×σ Z+ are algebraically isomorphic. Then we may proceed as in Claim 6 of the appendix and find a column (c11, . . . , cd1)t ∈ Cd that On the other hand suppose that T + 61 o o o o o o / / / E and C(Λ) ×σ Z+ are simply the disc algebra A(D). is a left invertible matrix. However this is a contradiction unless d = 1. In this (only) case both algebras T + 10.2.2. Crossed product. Furthermore we compare C(Λ)(cid:111)σ Z with C∗(T ) and C∗(T )/K(FX). This provides also a comparison with OE. Since OE is a quotient of TE this provides also a comparison with TE. In particular we claim that in general the C*-crossed product differs from these C*-algebras. For the first counterexample let I be the monomial ideal in C(cid:104)x1, x2(cid:105) generated by {x1x2, x2x1}. Then the system (Λ, σ) is identified with the identity map on two points. Therefore C(Λ)(cid:111)σ Z (cid:39) C({0, 1})⊗C(T), thus it is commutative. However both OE (cid:39) C∗(T ) and TE for this example are not commutative. 1 T2 = 0 whereas T2T ∗ 1 e1 = e2. For the second counterexample let I be the trivial zero ideal in C(cid:104)x1, x2(cid:105). Then OE is the Cuntz algebra O2 on two generators. However the system (Λ, σ) contains fixed points. For example let the point (xn) with xn = 1 for all n ∈ Z. Therefore C(Λ) (cid:111)σ Z contains non-trivial (Fourier-invariant) ideals and cannot be ∗-isomorphic to the simple C*-algebra OE (cid:39) C∗(T )/K(FX). Indeed we have that T ∗ 10.3. Subshift constructions. In the following presentation we will follow as much as possible the notation of each work to facilitate comparison. We point out that it should not be confused with the notation we have fixed for our analysis. 10.3.1. Matsumoto's approach [58]. Given a two-sided subshift (Λ, σ) Matsumoto [58] builds the C*-algebras C∗(T ) and C∗(T )/K(FX). The latter is denoted by OΛ and it is the main subject in Matsumoto's work. As we have illustrated OE is not C∗(T )/K(FX) in general. Even more the restriction of the quotient map on the space generated by the Ti (or on the C*-algebra A) is not isometric unless E is injective. This follows by the remarks in Section 2.2 and Proposition 6.4. For a concrete example (and ad-hoc arguments) consider the forbidden words {12, 21} in the shift space {1, 2}. Indeed in this case we get that T ∗ 2 T2 = I + P∅ therefore (cid:107)T1 + T2(cid:107)2 = (cid:107)T ∗ 1 T1 + T ∗ 1 T1 + T ∗ 2 T2(cid:107) = 2, whereas for the quotient mapping q : C∗(T ) → C∗(T )/K(FX). (cid:107)q(T1 + T2)(cid:107)2 = (cid:107)q(T ∗ 1 T1 + T ∗ 2 T2)(cid:107) = 1, 10.3.2. Carlsen's approach [17]. Carlsen [17] revisited the C*-algebras that arise from a right subshift X. His approach is directed in giving a C*-algebra that has an additional universal property [17, Introduction]. To do this he constructs an injective C*-correspondence HX (K(HX)), of which he takes the relative Cuntz-Pimsner algebra with respect to the ideal φ−1 following the work of Pimsner [72] and Schweizer [77]. At this point we would like to inform the reader that we could not trace reference no.15 of [77]; this reference is essential for the proof of the main theorem of [77]. Carlsen's C*-correspondence HX is over the C*-algebra (cid:101)DX, which differs from A of Propo- sition 4.7; therefore HX differs from E. Let us give the definition of (cid:101)DX. On the topological HX 62 space X define the sets C(µ, ν) = {νw ∈ X µw ∈ X, w ∈ X}, and let (cid:101)DX be the C*-subalgebra of the bounded functions on X generated by the character- that X consists of two points and (cid:101)DX = C({0, 1}). However for the same subshift the In the particular case of the forbidden words {12, 21} on the symbol space {1, 2} we see istic functions on C(µ, ν). C*-algebra A of Proposition 4.7 is C({0, 1, 2}). 10.3.3. Carlsen-Matsumoto approach [18]. Matsumoto [60], and Carlsen and Matsumoto [18] focus on the C*-algebra q(A) for the canonical quotient map q : C∗(T ) → C∗(T )/K(FX). In his early work [60] Matsumoto gives a description for the compact Hausdorff space that identifies the commutative C*-algebra q(A). Later Carlsen and Matsumoto [18] revisited this claim which they show to be incorrect in general. They make several points. two-sided subshift (Λ, σ) on the symbol space {1, . . . , d} let First of all they consider right subshifts XΛ that arise in the following way. Given a XΛ = {(x1, x2, . . . ) ∃(yn) ∈ Λ.yn = xn for all n ≥ 1}. This is the "positive part" of the elements in Λ. For l ≥ 0 they define the equivalence relation ∼l on XΛ by µ ∼l ν ⇔ {w ∈ Λ∗ l wµ ∈ XΛ} = {w ∈ Λ∗ l wν ∈ XΛ}. Define Ωl := XΛ/ ∼l and note that there is an onto mapping Ωl+1 → Ωl. Then let Ω be the projective limit of Ωl with respect to these onto maps. It worths comparing these definitions with the ones in Section 4.4 where we consider elements in Λ∗ instead of XΛ. µSµ µ ∈ Λ∗) where the Si act on the Hilbert space H =(cid:10)e(xn) (xn) ∈ XΛ Secondly they show [18, Lemma 2.2] that C(Ω) coincides with C∗(S∗ (cid:11) by (cid:40) Sie(xn) = e(i,(xn)) 0 (i, (xn)) ∈ XΛ, if otherwise. Moreover they identify [18, Lemma 2.9] the C*-algebra q(A) with the continuous functions l be the set of the equivalence classes on the finite words µ in Λ∗ on another space Ω∗. Let Ω∗ for which the set {w ∈ Λ∗ w ∼l µ} is infinite. Then q(A) is identified with the continuous functions on the projective limit of the Ω∗ l . Furthermore they show [18, Corollary 3.3] that the mapping q(Ti) (cid:55)→ Si defines a ∗- isomorphism between C∗(T )/K(FΛ) and the C*-algebra C∗(S) generated by the Si, under the assumption that the subshift satisfies condition (*) and condition (I). Condition (I) is crucial for the results in [18]. For example [18, Corollary 3.3] is true for the full shift on Σ = {1, 2}. But it fails to be true in general. For the forbidden words {12, 21} on the symbol set {1, 2} we see that S1 = S∗ 1S1 since XΛ consists of the points (11 . . . ) and (22 . . . ). If there was a ∗-isomorphism such that q(Ti) → Si then C∗(S) would admit a gauge action, say {βz}z∈T. Then l 0 = βz(S1 − S∗ 1S1) = zS1 − S∗ 1S1 = (z − 1)S1, for every z ∈ T which leads to the contradiction S1 = 0. 63 The interested reader should be warned here that several results on subshifts hold for C∗(S) whereas other hold for C∗(T )/K(FΛ). Carlsen and Matsumoto provide a very illuminating discussion and description of their results in the introduction of [18]. 10.4. OE as a graph C*-algebra. Let Λ be a two-sided sofic subshift. If I is the monomial ideal generated by forbidden words in Λ, then we can form the C*-algebra OE for the associated C*-correspondence E. We thus obtain a new C*-algebra that is constructed out of a subshift Λ, and one may ask whether this algebra is a reasonable one to consider. Our goal in this section is to show that OE arises from Λ via two well known and natural constructions: roughly speaking, it is the graph C*-algbera of the follower set graph of Λ. The same analysis can be carried out for one sided sofic subshifts. Let us introduce the follower set graph of a subshift Λ. A useful reference for this material is [55, Chapter 3], but we warn the reader that we reverse some of the notation. If Λ is a two-sided subshift, then the follower set of µ ∈ Λ∗ is the set FΛ(µ) := {w ∈ Λ∗ wµ ∈ Λ∗}. Note that different allowable words can have the same follower set. In fact, when Λ is sofic then there are only finitely many follower sets [55, Theorem 3.2.10]. The follower set graph is then defined to be the labeled graph whose vertices are parameterized by the follower sets, and there is exactly one edge labeled i from F (µ) to F (iµ) when iµ is allowable. We allow the empty word to have its own follower set -- this may or may not coincide with the follower set of another word. As in Remark 4.11, when Λ is sofic then there exists some k such that Ωk = Ωn = Ω for all n ≥ k and so the quantized dynamics are defined on a discrete space. In [10, Section 5] it was observed that one may identify the follower set graph with the labeled graph of the quantised dynamics (as in Remark 4.14). That is, the quantised dynamics, when viewed as a labeled graph, give us the familiar follower set graph of a subshift. Now, from the labeled follower set graph of Λ we obtain a directed unlabeled graph by simply erasing all the labels. Note here that we do not identify different edges, we just forget about their labels. Let us call this graph the underlying graph of the follower set graph, and let it be denoted by G = (G0, G1). Thus, G0 is the finite set Ω, and G1 consists of all pairs of points (u, v) ∈ Ω × Ω, for which there is some map ϕi with ϕi(u) = v. In Section 5.3 we saw that the associated C*-correspondence E coincides with Katsura's [49] topological graph determined by the quantised dynamical system. But under the as- sumption that E comes form a sofic two-sided subshift, this topological graph is just the finite underlying graph of the follower set graph G discussed in the previous paragraph. So E is the C*-correspondence of G, and hence we conclude (using [48, Proposition 3.10]) that OE (cid:39) C∗(G), i.e, OE is the graph C*-algebra of the follower set graph. Since the complete details of the isomorphism OE (cid:39) C∗(G) are hard to find in the litera- ture, and also because our notation differs from that used by Katsura in [48], we pause to justify and make explicit this isomorphism. Proposition 10.2. Let E be the C*-correspondence of a sofic subshift Λ, and let G = (G0, G1) be the underlying unlabeled graph of the follower set graph of Λ. Then OE is ∗- isomorphic to C∗(G). 64 Proof. Let (π, t) be the universal covariant representation of (A, E) in OE and define the families P = {π(a) a is a minimal projection in A}, and S = {t(Tia) i = 1, . . . , d ; a ≤ T ∗ i Tia) = t(Tia)∗t(Tia), then t(Tia) (cid:54)= 0 if and only if a ≤ T ∗ Since π(T ∗ As the minimal projections of A sum up to the identity we have that i Ti ; a is a minimal projection in A}. i Ti, when a is minimal. (cid:88){t(Tia) a ≤ T ∗ t(Ti) = i Ti ; a is a minimal projection in A}. We will show that the family (P,S) is a Cuntz-Krieger family for G. As (P,S) is generating for OE and admits a gauge action, the gauge invariant uniquesness theorem (for graph C*- algebras) will then show that C∗(G) and OE are ∗-isomorphic. The minimal projections in the finite dimensional algebra A are precisely the characteristic functions of points in Ω = G0. Hence the family P consists of mutually orthogonal (nonzero) projections corresponding to the vertices in G0. Next, we shall show that every element in S is a nonzero partial isometry corresponding to an edge in G1. Let a be the minimal projection corresponding to a point va ∈ G0 = Ω. i Ti, then va ∈ Ωi. Thinking of Ti as in the picture given in Section 5.3, we have If a ≤ T ∗ that Ti corresponds to the characteristic function of all the edges labeled i emitted from Ωi (see the proof of Proposition 5.4 and the discussion above it). Therefore, Tia is the characteristic function of an edge e ∈ G1, (which originally had a label i) coming out of the point s(e) = va ∈ Ωi, and going into some point ϕi(va) = r(e). Moreover, as a ≤ T ∗ i Ti, we find that a∗T ∗ i Tia = a∗a = a (cid:54)= 0, so we get t(Tia)∗t(Tia) = π((cid:104)Tia, Tia(cid:105)) = π(a∗T ∗ i Tia) = π(a). This shows that all elements of S are partial isometries which satisfy the first Cuntz-Pimsner relation S∗ e Se = Ps(e). It remains to show the second Cuntz-Pimsner relation, namely r−1(v) (cid:54)= 0. for SeS∗ e = Pv Fix a vertex v ∈ G0 = Ω that is not source and let b ∈ A denote the minimal projection corresponding to this point. We need to show that (cid:88) e∈r−1(v) (cid:88) ϕi(va)=v (cid:88) t(Tia)t(Tia)∗ = π(b), (cid:88) t(Tia)t(Tia)∗ = ψt( ΘTia,Tia). (cid:88) ϕi(va)=v ΘTia,Tia = φE(b). ϕi(va)=v 65 where va ∈ G0 is the point corresponding to a minimal projection a ∈ A, such that (va, v) ∈ G1. We sum over all i and va such that φi(va) = v, and this amounts to summing over all edges e ∈ G1 for which v = r(e). Now, by definition of ψt (see Section 2.2), We will now show that ϕi(va)=v To see this, we fix ξ ∈ E, and compute(cid:88) ΘTia,Tiaξ = (cid:88) Tia(cid:104)Tia, ξ(cid:105). ϕi(va)=v ϕi(va)=v Now, by using the topological graph picture described in Section 5.3, we obtain (cid:104)Tia, ξ(cid:105)(u) = Ti(f )a(s(f ))ξ(f ). This expression will be zero when u (cid:54)= va, and when u = va, we have that (cid:104)Tia, ξ(cid:105)(va) = Ti(f )ξ(f ) = ξ(e) f∈s−1(va) (cid:88) (cid:88) f∈s−1(u) for the one edge e which is the unique edge leaving va with label i. Therefore ΘTia,Tia is the projection onto the subspace spanned by Tia. Now, recall that Tia is the characteristic function on of the unique edge leaving va with label i, and that, by assumption, this edge must go into v. This means that for every f ∈ G1,  (cid:88) (cid:40)  (f ) = ΘTia,Tia(ξ) 0 ξ(f ) r(f ) (cid:54)= v, r(f ) = v. But the left action of b on ξ is given by ϕi(u)=v (cid:80) [φE(b)(ξ)] (f ) = b(r(f ))ξ(f ), which has the same effect, since b is the characteristic function of v. We conclude that ϕi(u)=v ΘTia,Tia = φE(b), as we set out to show. Thus, by putting everything together and using covariance of the representation (π, t), we obtain (cid:88) t(Tia)t(Tia)∗ = ψt (φE(b)) = π(b) ϕi(u)=v which completes the proof. Remark 10.3. It will be interesting to investigate the dependence of OE dynamical aspects of the topological dynamical system (Λ, σ). We leave this for future work. ∼= C∗(G) on the 11. Appendix Davidson and Katsoulis [23, Theorem 3.22] show that algebraic isomorphisms of the tensor algebras of two classical multivariable systems implies piecewise conjugacy of the systems. Let us provide here an alternative proof of this breakthrough result. Our approach relies on appropriate compressions of the Fock representations. The proof follows a series of steps and let us start by describing the objective. Let (X, σ) ≡ (X, σ1, . . . , σd) be a classical system, i.e. X is a locally compact Hausdorff space and every σ : X → X is a proper continuous map. If µ = µ1 . . . µd ∈ Fd + we write σµ for the composition σµ1 ··· σµn. The tensor algebra T + (X,σ) is defined as the (universal) non- selfadjoint operator algebra generated by s1f, . . . , sdf with f ∈ C0(X), such that [s1, . . . , sd] is a row contraction, and f sj = sjf σj for all f ∈ C0(X) and j = 1, . . . , d. This is slightly different from [23, Definition 1.2] and the reader is addressed to Remark 8.4. The universal (X,σ) suggests that if π : C0(X) → B(H) is a representation, and there is a row property of T + 66 contraction [s1, . . . , sd] ∈ B(H d, H) such that f si = sif σi then the mapping sif (cid:55)→ siπ(f ) extends to a completely contractive representation of T + Let (Y, τ ) ≡ (Y, τ1, . . . , τd(cid:48)) be a second classical system. We will use the notation t1g, . . . , td(cid:48)g for the generators of T + (X,σ) in B(H). (Y,τ ). Fix an algebraic isomorphism Φ : T + (X,σ) → T + (Y,τ ). An important point to keep in mind is that Φ is automatically continuous [23, Corollary 3.6]. This is not immediate and follows by the trick of Donsig, Hudson and Katsoulis [29]. We aim to show that there is a homeomorphism γs : Y → X and that for every y ∈ Y there is a permutation π ∈ Sd and a neighbourhood Uπ such that γsτiUπ = σπ(i)γsUπ for all i = 1, . . . , d. Our first task is to find the homeomorphism γs : Y → X. We write M(X,σ) for the character (X,σ). If θ ∈ M(X,σ) then θC0(X) is a character, and thus equals to evx for some (X,σ) such that θC0(X) = evx for some x ∈ X. Then θ is space of T + x ∈ X. Claim 1. Let θ be a character of T + completely determined by for any f ∈ C0(X) such that θ(f ) = f (x) = 1, in the sense that for every µ = µ1 . . . µd ∈ Fd and g ∈ C0(X) we have + (λ1, . . . , λd) := (θ(s1f ), . . . , θ(sdf )) (cid:40) θ(sµg) = λµg(x) 0 if σµi(x) = x for all µi, otherwise, where λµ1...µn = λµ1 ··· λµn. In particular λi = 0 if σi(x) (cid:54)= x. Proof of Claim 1. First we compute θ(sig) = θ(sig)θ(f ) = θ(sif g) = θ(sif )θ(g), which implies that the tuple (λ1, . . . , λd) does not depend on the choice of the function f . Secondly observe that if σi(x) (cid:54)= x then λi = 0. Indeed let h ∈ C0(X) such that hσi(x) = 0 and h(x) = 1 for which we have λi = θ(sif ) = θ(h)θ(sif ) = θ(sif )θ(hσi) = 0. Suppose that the claim is true for all words of length less than k and let µ = µ1 . . . µk+1 be of length k + 1. If σµi(x) = x for all i then we can write µ = wν for a word w (cid:54)= ∅ of length strictly less than k + 1 such that σw(x) = x, and σν(x) = x. Then we get that θ(sµg) = θ(f )θ(sµg) = θ(swf σw)θ(sνg) = λwf (σw(x))λνg(x) = λµg(x), by the inductive hypothesis. On the other hand let µn be the first letter from left to right in the word µ for which σµ1...µn(x) (cid:54)= x. If n < k +1 then let w = µ1 . . . µn and ν = µn+1 . . . µk+1 for which we have that σw(x) (cid:54)= x. Let h ∈ C0(X) such that h(x) = 1 and hσw(x) = 0 and compute θ(sµg) = θ(h)θ(sµg) = θ(swhσw)θ(sνg) = λwh(σw(x))θ(sνg) = 0. If n = k + 1 then the word µ = µ1 . . . µk+1 is such that x = σµ1(x) = σµ1µ2(x) = ··· = σµ1...µk(x) 67 but x (cid:54)= σµ(x), so it follows that σµk+1(x) (cid:54)= x. Therefore we get that θ(sµk+1g) = 0 and so θ(sµg) = θ(f )θ(sµg) = θ(sµ1...µkf σµ1 ··· σµk)θ(sµk+1g) = 0. This ends the proof of Claim 1. (cid:4) The next claim establishes the connection between the evaluation functionals on X and the evaluation functionals on Y . The proof of the first part of the claim relies on [23] and is included for completeness. It is the second part of the claim that plays a central role in our analysis. Claim 2. The homomorphism Φ induces a homeomorphism γs : Y → X. In particular, if θ ∈ M(Y,τ ) with θC0(Y ) = evy then θΦ ∈ M(X,σ) with θΦC0(X) = evγs(y). Proof of Claim 2. The homomorphism Φ defines a homeomorphism γc : M(Y,τ ) → M(X,σ) : θ (cid:55)→ θΦ. It is not immediate but follows as in [23, Lemma 3.9] that the maximal analytic sets inside M(X,σ) are precisely the sets {θ ∈ M(X,σ) θC0(X) = evx} parameterized by the points x ∈ X. In particular each one is homeomorphic to a ball of dimension equal to the number of the σi that fix x. By the Fourier transform the elements in the operator algebras are generalised analytic polynomials. Consequently Φ is weakly biholomorphic and therefore γc maps maximal analytic sets onto maximal analytic sets. Thus γc induces a set map γs : Y → X by collapsing every set of characters θ that satisfy θC0(Y ) = evy to a single point. Therefore if γc(θ) = θΦ = θ(cid:48) and θC0(Y ) = evy then θ(cid:48)C0(X) = evx with γs(y) = x. This set map is moreover a homeomorphism. To see this, note that γs is the adjoint of C(X) i−→ T + (X,σ) Φ−→ T + (Y,τ ) E0−→ C(Y ), where i denotes inclusion and E0 the conditional expectation onto C(Y ). This completes (cid:4) the proof of Claim 2. We now turn to showing that the homeomorphism γs defined above induces the piecewise conjugacy. It suffices to show that γs implements a piecewise conjugacy in a neighbourhood of every y ∈ Y . The key is to work with germs. For a fixed y ∈ Y , we write τi ∼ τj if τi and τj are maps on Y for which there exists some neighbourhood U (cid:51) y such that τiU = τjU . The equivalence class of τi is called the germ of τi at y. Compare the germ of τ1 at y to the germs of the mappings γ−1 s σdγs and τ1, . . . , τd(cid:48) at y. After a possible re-enumeration we find that there is a neighbourhood V = V1 (cid:51) y and there are k, l so that: s σ1γs, . . . , γ−1 (i) γ−1 (ii) for any γ−1 yλ → y with γ−1 s σ1γs(z) = ··· = γ−1 s σkγs(z) = τ1(z) = ··· = τl(z) for all z ∈ V ; and s σiγs with i > k (resp. τj with j > l) there is a net yλ ∈ V such that s σiγs(yλ) (cid:54)= τ1(yλ) for all λ ∈ Λ (resp. τj(yλ) (cid:54)= τ1(yλ) for all λ ∈ Λ). The main objectives then are to show that k = l and that d = d(cid:48). Indeed in this case we have that σ1, . . . , σk and τ1, . . . , τk are conjugate on V1. After repeating the process for τ2, . . . , τd we will have that the intersection V1 ∩ ··· ∩ Vd gives the required Uπ on which up to the permutation π the tuple σ1Uπ , . . . , σdUπ is conjugate to τ1Uπ , . . . , τdUπ . The fact that d = d(cid:48) will ensure that every σj is paired with some τi (and vice versa), and that π is a permutation on d symbols. 68 We now proceed to the proof. For any point y ∈ Y we can define the orbit representation πy : C0(Y ) → B((cid:96)2(F+ d(cid:48))) by πy(g) = diag{gτw(y) w ∈ Fd(cid:48)} for all g ∈ C0(Y ). d(cid:48))) so that V (ti)(eµ) = eiµ for all i = 1, . . . d(cid:48) and µ ∈ F+ d(cid:48) we (Y,τ ). Define the algebraic homomorphism By setting V (ti) ∈ B((cid:96)2(F+ obtain the contractive homomorphism (V ×πy) of T + V ×πy−→ B((cid:96)2(F+ d(cid:48))) E−→ Md(cid:48)+1(C), where E is the compression by the projection onto the subspace of (cid:96)2(F+ i=1 tigi vectors e∅, e1, . . . , ed(cid:48). Consequently if we write Φ(Q) = g0 +(cid:0)(cid:80)d(cid:48) Φy : T + (cid:1) + Z for Q ∈ T + d(cid:48)) generated by the Φ−→ T + (X,σ) (Y,τ ) (X,σ) then we get Φy(Q) =  g0(y) g1(y) g2(y) ... gd(cid:48)(y) 0 g0τ1(y) 0 0 0 ... 0 g0τ2(y) ... 0 . . . . . . . . . . . . . . . 0 0 0 0 g0τd(cid:48)(y)  . By denoting Eij the rank one operator with one in the (i, j)-entry and zeroes elsewhere, we may alternatively write Φy(Q) = g0(y)E00 + g0τi(y)Eii + gi(y)Ei0. d(cid:48)(cid:88) i=1 We remark here that the range of Φy contains all Ei0. Indeed the range of Φy equals the range of E(V × πy) since Φ is onto. By choosing a g ∈ C0(Y ) such that g(y) = 1 we obtain that Ei0 = E(V × πy)(tig) for all i = 1, . . . , d(cid:48). Claim 3. Suppose that Ψy : T + ΨyC0(X) is diagonal on the range of Φy, in the strong sense that For a matrix A ∈ Md(cid:48)+1(C) we write [A]ij for the element in the (i, j)-entry. (X,σ) → Ran Φy is an algebraic homomorphism such that d(cid:48)(cid:88) [Φy(f )]ii · Eii for all f ∈ C0(X), Ψy(f ) = and suppose that i=0 Ψy(sjh) =  (cid:2)Ψy(f )Ψy(sjh)(cid:3) ∗ c1j c2j ... cd(cid:48)j 0 0 . . . 0 0 . . . ∗ 0 . . . 0 ∗ . . . ... 0 0 0 . . . 0 ∗ ... i0 =(cid:2)Ψy(f · sjh)(cid:3)  for some h ∈ C0(X). i0 =(cid:2)Ψy(sjh)Ψy(f σj)(cid:3) i0. If cij (cid:54)= 0 then τi(y) = γ−1 Proof of Claim 3. By the covariance relation and commutativity of C0(X) we obtain s σjγs(y). By using the form of the elements as in the assumption we obtain [Φy(f )]ii · cij = cij · [Φy(f σj)]00. 69 However, the homomorphisms Q (cid:55)→ [V × πy(Q)]00 and Q (cid:55)→ [V × πy(Q)]ii are characters of T + by the discussion on the characters we have that (Y,τ ) whose restrictions on C0(Y ) are evy and evτi(y), respectively. Hence [Φy(f )]00 = [V × πy(Φ(f ))]00 = evγs(y)(f ), and [Φy(f )]ii = [V × πy(Φ(f ))]ii = evγsτi(y)(f ). Consequently f γsτi(y) · cij = cij · f σjγs(y) for all f ∈ C0(X) which completes the proof of (cid:4) Claim 3. We will construct a homomorphism Ψy as the one appearing in the claim above. To this end we require the following remark. Let T2 denote the lower triangular 2 × 2 matrices. Claim 4. Let Φ : C0(X) → T2 be an algebraic homomorphism. If [Φ(f )]11 = [Φ(f )]22 then [Φ(f )]21 = 0. Proof of Claim 4. Let C0(K) = C∗(f ) be the commutative C*-subalgebra of C0(X) generated by f . Since g (cid:55)→ [Φ(g)]ii are characters, we obtain that [Φ(h)]11 = [Φ(h)]22 for all h ∈ C0(K). Moreover these characters correspond to an evaluation, say evx for some x ∈ K. Since Φ(hg) = Φ(h)Φ(g) for all h, g ∈ C0(K) we get that [Φ(hg)]21 = [Φ(h)]21 · [Φ(g)]11 + [Φ(h)]22 · [Φ(g)]21 = [Φ(h)]21 · g(x) + h(x) · [Φ(g)]21. Hence g (cid:55)→ [Φ(g)]21 is a point derivation at x ∈ K. Therefore [Φ(f )]21 is zero and the proof (cid:4) of Claim 4 is completed. Now we have all the required ingredients to establish the existence of a homomorphism Ψy as in the Claim 4. In particular Ψy will be similar to Φy by an invertible matrix Ay. Claim 5. Let Φy be the representation associated with a point y ∈ Y . Then there exists an algebraic homomorphism Ψy : T + (X,σ) → Ran Φy such that d(cid:48)(cid:88) Ψy(f ) = [Φy(f )]ii · Eii for all f ∈ C0(X), i=0 and the range of Ψy contains all Ei0 for i = 1, . . . , d. Proof of Claim 5. We will construct the homomorphism Ψy step by step. If γs(y) = γsτ1(y) then [Φy(f )]00 = [Φy(f )]11. Therefore we obtain  Φy(f ) = g(y) 0 ∗ ... ∗ 0 . . . g(y) 0 . . . ∗ . . . ... 0 . . . 0 0 0 . . . 0 ∗ 0 0 ... 0  by Claim 4 and we proceed to the second step. On the other hand, suppose that γs(y) (cid:54)= γsτ1(y) and choose an f ∈ C0(X) of norm less than 1 such that fU = 1 and fγsτ1(U ) = 0 for 70 an appropriate compact neighbourhood U of γs(y) with U ∩ γsτ1(U ) = ∅. We have already remarked that [Φy(f )]00 = f γs(y) and [Φy(f )]ii = f γsτi(y) for all f ∈ C0(X) so that If k = [Φy(f )]10 let A be the invertible matrix I − kE10 (with inverse A−1 = I + kE10). Then we derive [Φy(f )]00 = 1 and [Φy(f )]11 = 0.  1 0 0 . . . 0 0 0 . . . ∗ 0 ∗ . . . ... ∗ 0 0 . . . 0 0 0 . . . 0 ∗ ... ...  AΦy(f )A−1 = By commutativity we get that AΦy(f )A−1AΦy(h)A−1 = AΦy(h)A−1AΦy(f )A−1 for all h ∈ C0(X). This shows that the (1, 0)-entry of AΦy(h)A−1 is zero. Furthermore the diagonals of AΦy(h)A−1 and Φy(h) are the same. In addition AEi0A−1 = Ei0 for all i = 1, . . . , d(cid:48). homomorphism Ψy : T + We can continue inductively for the rest of the rows. In this way we form an algebraic (cid:4) The matrices constructed in Claim 5 do not depend on the choice of f . Indeed if g ∈ C0(X) (X,σ) → Ran Φy that satisfies the conclusion of Claim 5. is such that [Φy(g)]00 = 1 and [Φy(g)]11 = 0, then by using commutativity we obtain that the equality Φy(f )Φy(g) = Φy(g)Φy(f ) gives that [Φy(g)]10 = [Φy(f )]10. We will write Ay for the invertible matrix inside Md(cid:48)+1(C) that is constructed in Claim 5 and gives the homomorphism Ψy(·) = AyΦy(·)A−1 y . Claim 6. The systems have the same multiplicity, i.e. d = d(cid:48). Proof of Claim 6. Let Ψy be an algebraic homomorphism that satisfies the conclusion of Claim 5. Hence the range contains all Ei0. Select an h ∈ C0(X) such that 1 = hγs(y) = hγsτi(y) for all i = 1, . . . , d(cid:48). By Claim 4 then we have that Ψy(h) = Id(cid:48). Write cij = [Ψy(sjh)]i0 and observe that Ψy(sjhf ) = Ψy(sjf )Ψy(h) = Ψy(sjf ) for all f ∈ C0(X). As in [44, Proposition 4.3] we can show that there is a matrix [c(cid:48) ij] = Id(cid:48). This implies that d ≤ d(cid:48). This inequality holds for any y ∈ Y . By symmetry on Φ−1 and for an x ∈ X we get that d(cid:48) ≤ d, and the proof of Claim 6 is complete. (cid:4) As a consequence, Claim 6 further implies that the obtained matrix [cij] is square. There- To end the proof fix a y ∈ Y . Let the setup be as after the proof of Claim 2. Fix the map fore the matrix [cij] which is shown to be right invertible in Claim 6, is in fact invertible. ij] such that [cij][c(cid:48) τ1 and suppose that after a re-enumeration the germ τ1 contains exactly the functions γ−1 s σ1γs, . . . , γ−1 s σkγs and τ1, . . . , τl for some k, l. 71 We have just seen that d = d(cid:48). We aim to show that k = l. The next claim gives the required equality k = l, modulo a condition on the matrix [cij] of Claim 6. Claim 7. Let [cij] be the invertible matrix obtained in Claim 6 for the fixed y ∈ Y . If cij = 0 for the i, j that satisfy τi ∼ τ1 (cid:54)∼ γ−1 Proof of Claim 7. We may assume without loss of generality that k ≤ l, and it remains to prove that k < l is impossible. This follows verbatim from the last paragraph of [44, Proof of Theorem 4.9]. In short, we may write s σjγs, then k = l. (cid:20)[aij] 0 (cid:21) ∗ , ∗ [cij] = where [aij] is a (l × k)-matrix. If k < l then when applying the Gaussian elimination we will be able to eliminate the entire k + 1 row of the invertible matrix [cij], which gives the (cid:4) contradiction. Therefore in the rest of the proof we will focus in showing that cij = 0 for the i, j that satisfy τi ∼ τ1 (cid:54)∼ γ−1 The coefficients cij are obtained by the (i, 0)-th entries of Ψy(sjh) for the function h of Claim 6. For their analysis we can restrict even further to the (2× 2)-matrix representations. For z ∈ Y we define s σjγs. ψz,i := Pi0Ψz : T(X,σ) → T2, where Pi0 is the projection on the space generated by {e∅, ei}. Since we are in the lower triangular matrices we have that Pi0ΨzPi0 = Pi0Ψz, hence ψz,i is indeed a homomorphism. Then we get that We will also write φz,i = Pi0Φz and Az,i = Pi0Az. We can check that ψz,i(·) = Az,iφz,i(·)A−1 z,i . cij = [ψy,i(sjh)]10. Case 1. Suppose that y (cid:54)= τ1(y) and choose f with [φy,1(f )]11 = 0 as we do in Claim 5. We may also choose a neighbourhood U of y where [φz,1(f )]00 = 1 and [φz,1(f )]11 = 0 for all z ∈ U . Then we get the invertible matrices Az,1 = I − [φz,1(f )]10 · E10 for all z ∈ U . Consequently, invoking the automatic continuity of Φ, we get that A•,1 is s σjγs (cid:54)∼ τ1; then there exists a net yλ ∈ U with yλ → y continuous on y. Let j such that γ−1 such that γsτ1(yλ) (cid:54)= σjγs(yλ) for all λ. Consider the elements c1j,λ = [ψyλ(sjh)]10 = [Ayλ,1φyλ,i(sjh)A−1 yλ,1]10. By continuity we have that limλ Ayλ,1 = Ay,1 hence c1j = limλ c1j,λ. However c1j,λ = 0 as follows by the construction of yλ and Claim 3. Indeed if c1j,λ (cid:54)= 0 then Claim 3 indicates that γsτ1(yλ) = σjγs(yλ). We conclude that c1j = 0. Case 2. Suppose that y = τ1(y). Then φy,1C0(X) is scalar by Claim 4. Let j such that s σjγs (cid:54)∼ τ1. Let a net yλ ∈ V such that γsσjγ−1 s (yλ) (cid:54)= τ1(yλ) for all λ. Construct as above γ−1 the invertible matrices Ayλ and note that (cid:21) (cid:20) 1 72 Ayλ,1 = 0 cλ 1 . Now all Ayλ are uniformly bounded by 1 + (cid:107)Φ(cid:107) and by passing to a subsequence we may assume that they converge to an operator. Consequently the Ayλ,1 converge to a matrix (cid:20)1 0 (cid:21) c 1 , A = which is invertible. Then we obtain φy,1(sjh) = A−1(cid:0) lim λ Ayλ,1φyλ,1(sjh)A−1 yλ,1 (cid:1)A = A−1 lim λ ψλ,1(sjh)A. But by Claim 3 again we get that c1j,λ = [Ayλ,1φyλ,1(sjh)A−1 yλ,1]10 = [ψλ,1(sjh)]10 = 0, s σjγs (cid:54)∼ τ1. Hence φy,1(sjh) = A−1DA where D is diagonal. Now by assumption since γ−1 y = τ1(y) hence the range of φy,1 consists of the matrices a1I + a2E10, all of which commute with A. Consequently we obtain D = Aφy,1(sjh)A−1 = φy,1(sjh), We showed that in either case c1j = 0 for all j such that τ1 (cid:54)∼ γ−1 which shows that c1j = [φy,1(sjh)]10 = [D]10 = 0. with τi for i = 1, . . . , k we have that cij = 0 for all i, j that satisfy τi ∼ τ1 (cid:54)∼ γ−1 explained above, this concludes our proof of [23, Theorem 3.22]. s σjγs. By substituting τ1 s σjγs. As we 11.1. Applications. The above idea applies to other classes of operator algebras. The key is to analyse the maximal analytic sets of the character space, and also ensure the existence of a Fock representation. Let us describe briefly how this is achieved in three cases. We reserve a full discussion for a forthcoming project. The first case is the second part of [23, Theorem 3.22]. The semicrossed product C0(X)×σ Fd + related to (X, σ) is defined as the universal nonselfadjoint operator algebra generated by the sµf so that now the si are taken to be just contractions, and f si = sif σi. Once again we do not require the generators to be separated into si and f ∈ C0(X) as in [23, Definition 1.2]. When the semicrossed products are algebraically isomorphic then the maximal analytic sets are polydiscs, and again Claim 1 holds. Furthermore by definition there is a canonical epimorphism from the semicrossed product onto the tensor algebra. This sets the appropriate context to pass to an epimorphism of C0(X) ×σ Fd + onto T(Y,τ ) and then follow the steps from Claim 3 and thereon to obtain again piecewise conjugacy. Indeed, the only ingredient required is that the homomorphisms are onto. On the other hand, suppose that the σi : X → X commute. Then we may define the universal nonselfadjoint operator algebra C0(X)×σ Zd + and f ∈ C0(X) so that now the si are taken to be commuting contractions, and f si = sif σi. It follows that the character space of C0(X) ×σ Zd + coincides with the character space of C0(X) ×σ Fd + generated by the sxf for x ∈ Zd +. Furthermore the family (πy,{Vi}d i=1) defined by πy(f ) = diag{f σx(y) x ∈ Zd +} for f ∈ C0(X), and Viex = ei+x defines a completely contractive representation of C0(X) ×σ Zd are the appropriate ingredients to show that if C0(X) ×σ Zd C0(Y ) ×τ Zd(cid:48) in the case of the tensor algebras (even the part on automatic continuity). +. These + is algebraically isomorphic to + then the systems are piecewise conjugate. Indeed the proof reads the same as 73 Even more one may consider the universal nonselfadjoint operator algebra C0(X) ×rc Zd + and f ∈ C0(X) so that now the si form a commuting row generated by the sxf for x ∈ Zd contraction, and f si = sif σi. Now the character space coincides with the character space of T(X,σ). Furthermore we obtain a contractive representation by compressing the family (πy,{Vi}d + above by the projection p that gives the symmetric subproduct system. Indeed commutativity of the σi implies that p commutes with πy. Once more one derives piecewise conjugacy when C0(X)×rc Zd(cid:48) +. Automatic continuity can be shown to hold as well. i=1) of C0(X)×σ Zd + is algebraically isomorphic to C0(Y )×rc Zd τ σ + σ References [1] B. Abadie, S. Eilers and R. Exel, Morita equivalence for crossed products by Hilbert C*-bimodules, Trans. Amer. Math. Soc. 350:8 (1998), 3043 -- 3054. [2] L. Accardi and M. Skeide, Interacting Fock space versus full Fock module, Commun. Stoch. Anal. 2:3 (2008), 423 -- 444. [3] W. B. Arveson, Subalgebras of C*-algebras, Acta Math. 123 (1969) 141 -- 224. [4] W. B. Arveson, Subalgebras of C*-algebras II, Acta Math. 128:3-4 (1972), 271 -- 308. [5] W. B. Arveson, Subalgebras of C*-algebras III, Acta Math. 181:2 (1998), 159 -- 228. [6] W. B. Arveson, Non commutative dynamics and E-semigroups, Springer Monographs in Math., Springer- Verlag, 2003. [7] W. B. Arveson, p-Summable commutators in dimension d, J. Operator Theory 54:1 (2005), 101 -- 117. [8] W. B. Arveson, The noncommutative Choquet boundary, J. Amer. Math. Soc. 21:4 (2008), 1065 -- 1084. [9] W. B. Arveson, The noncommutative Choquet boundary II: hyperrigidity, Israel. J. Math. 184 (2011), 349 -- 385. [10] C. Barret and E. T.A. Kakariadis On the quantized dynamics of factorial languages, Q. J. Math., to appear. [11] B. V. R. Bhat and M. Mukherjee, Inclusion systems and amalgamated products of product systems, Infin. Dimens. Anal. Quantum Probab. Relat. Top. 13:1 (2010), 1 -- 26. [12] B. V. R. Bhat and M. Skeide, Tensor product systems of Hilbert modules and dilations of completely positive semigroups, Infin. Dimens. Anal. Quantum Probab. Relat. Top. 3:4 (2000), 519 -- 575. [13] D. P. Blecher and C. Le Merdy, Operator algebras and their modules -- an operator space approach. London Mathematical Society Monographs. New Series, 30. Oxford Science Publications. The Clarendon Press, Oxford University Press, Oxford, 2004. [14] A. Borel, Linear Algebraic Groups, Graduate Texts in Mathematics, Vol. 126. Springer-Verlag, New York, 1991. [15] N. P. Brown and N. Ozawa, C*-algebras and Finite-Dimensional Approximations, Graduate Studies in Mathematics, 88. American Mathematical Society, Providence, RI, 2008. [16] W. Bruns and J. Gubeladze, Polytopes, Rings, and K-Theory, Springer Monographs in Mathematics. Springer, Dordrecht, 2009. [17] T. M. Carlsen, C*-algebras associated with subshifts, Internat. J. Math. 19:1 (2008), 47 -- 70. [18] T. M. Carlsen and K. Matsumoto, Some remarks on the C*-algebras associated with subshifts, Math. Scand. 95:1 (2004), 145 -- 160. [19] J. Cuntz and W. Krieger, A class of C*-algebras and topological Markov chains, Invent. Math. 56:3 (1980), 251 -- 268. [20] K. R. Davidson, A. H. Fuller and E. T.A. Kakariadis, Semicrossed products of operator algebras by semigroups, Mem. Amer. Math. Soc. 247: 1168 (2017). [21] K. R. Davidson and E. T.A. Kakariadis, Conjugate dynamical systems on C*-algebras, IMRN 2014:5 (2014), 1289 -- 1311. [22] K. R. Davidson, E. G. Katsoulis, Isomorphisms between topological conjugacy algebras, J. reine angew. Math. 621 (2008), 29 -- 51. [23] K. R. Davidson and E. G. Katsoulis, Operator algebras for multivariable dynamics, Mem. Amer. Math. Soc. 209: 982 (2011). 74 [24] K. R. Davidson and M. Kennedy, The Choquet boundary of an operator system, Duke Mathematical Journal, to appear. [25] K. R. Davidson, C. Ramsey and O. M. Shalit, The isomorphism problem for some universal operator algebras, Adv. Math. 228:1 (2011), 167 -- 218. [26] K. R. Davidson, C. Ramsey and O. M. Shalit, Operator algebras for analytic varieties, Trans. Amer. Math. Soc. 367 (2015), 1121 -- 1150. [27] K. R. Davidson and J. Roydor, Isomorphisms of tensor algebras of topological graphs, Indiana Univ. Math. J. 60:4 (2011), 1249 -- 1266. [28] V. Deaconu, A. Kumjian, D. Pask and A. Sims, Graphs of C*-correspondences and Fell bundles, Indiana Univ. Math. J. 59:5 (2010), 1687 -- 1735. [29] A. P. Donsig, T. D. Hudson and E. G. Katsoulis, Algebraic isomorphisms of limit algebras, Trans. Amer. Math. Soc. 353:3 (2001), 1169 -- 1182 (electronic). [30] S. Doplicher, C. Pinzari and R. Zuccante, The C*-algebra of a Hilbert bimodule, Boll. Unione Mat. Ital. Sez. B Artic. Ric. Mat. (8) 1:2 (1998), 263 -- 281. [31] A. Dor-On and D. Markiewicz, Operator algebras and subproduct systems arising from stochastic ma- trices, J. Funct. Anal. 267:4 (2014), 1057 -- 1120. [32] A. Dor-On and D. Markiewicz, C*-envelope of tensor algebras arising from stochastic matrices, to appear in Integral Equations and Operator Theory (preprint arXiv:1605.03543). [33] M. A. Dritschel and S. A. McCullough, Boundary representations for families of representations of operator algebras and spaces, J. Operator Theory, 53:1 (2005), 159 -- 167. [34] K. J. Dykema and D. Shlyakhtenko, Exactness of Cuntz-Pimsner C*-algebras, Proc. Edinb. Math. Soc. (2) 44:2 (2001), 425 -- 444. [35] N. J. Fowler, P. S. Muhly and I. Raeburn, Representations of Cuntz-Pimsner algebras, Indiana Univ. Math. J. 52:3 (2003), 569 -- 605. [36] N. J. Fowler and I. Raeburn, The Toeplitz algebra of a Hilbert bimodule, Indiana Univ. Math. J. 48:1 (1999), 155 -- 181. arXiv:1402.0198. [37] M. Gerhold and M. Skeide, Discrete subproduct systems and word systems, preprint (2014), [38] M. Gurevich, Subproduct systems over N × N, J. Funct. Anal. 262:10 (2012), 4270 -- 4301. [39] M. Hamana, Injective envelopes of operator systems, Publ. Res. Inst. Math. Sci. 15:3 (1979), 773 -- 785. [40] E. T.A. Kakariadis, The Silov boundary for operator spaces, Integral Equations Operator Theory 76:1 (2013), 25 -- 38. [41] E. T.A. Kakariadis, Notes on the C*-envelope and the Silov ideal, Lecture Notes, 2012. [42] E. T.A. Kakariadis, A Note on the Gauge Invariant Uniqueness Theorem for C*-correspondences, Israel Journal of Mathematics, to appear. [43] E. T.A. Kakariadis and E. G. Katsoulis, C*-algebras and equivalences for C*-correspondences, J. Funct. Anal. 266:2 (2014), 956 -- 988. [44] E. T.A. Kakariadis and E. G. Katsoulis, Isomorphism invariants for multivariable C*-dynamics, J. Noncomm. Geom. 8:3 (2014), 771 -- 787. [45] E. T.A. Kakariadis and J. R. Peters, Representations of C*-dynamical systems implemented by Cuntz families, Munster J. Math. 6 (2013), 383 -- 411. [46] E. G. Katsoulis and D. W. Kribs, Isomorphisms of algebras associated with directed graphs, Math. Ann. 330:4 (2004), 709 -- 728. [47] E. G. Katsoulis and D. W. Kribs, Tensor algebras of C*-correspondences and their C*-envelopes, J. Funct. Anal. 234:1 (2006), 226 -- 233. [48] T. Katsura, A construction of C*-algebras from C*-correspondences, Advances in quantum dynamics (South Hadley, MA, 2002), 173 -- 182, Contemp. Math., 335, Amer. Math. Soc., Providence, RI, 2003. [49] T. Katsura, A class of C*-algebras generalizing both graph algebras and homeomorphism C*-algebras I, Fundamental results, Trans. Amer. Math. Soc. 356:11 (2004), 4287 -- 4322. [50] T. Katsura, On C*-algebras associated with C*-correspondences, J. Funct. Anal. 217:2 (2004), 366 -- 401. [51] T. Kajiwara, C. Pinzari and Y. Watatani, Ideal structure and simplicity of the C*-algebras generated by Hilbert bimodules, J. Funct. Anal. 159:2 (1998), 295 -- 322. 75 [52] M. Kennedy and O. M. Shalit, Essential normality, essential norms and hyperrigidity, J. Funct. Anal. 268:10 (2015), 2990 -- 3016; corrigendum, J. Funct. Anal. 270:7 (2016), 2812 -- 2815. [53] B. K. Kwa´sniewski, C*-algebras generalizing both relative Cuntz-Pimsner and Doplicher-Roberts alge- bras, Trans. Amer. Math. Soc. 365:4 (2013), 1809 -- 1873. [54] E. C. Lance, Hilbert C*-modules. A toolkit for operator algebraists, London Mathematical Society Lec- ture Note Series, 210. Cambridge University Press, Cambridge, 1995. [55] D. Lind and B. Marcus, An Introduction to Symbolic Dynamics and Coding, Cambridge University Press, Cambridge, 1995. [56] V. M. Manuilov and E. V. Troitsky, Hilbert C*-modules, Translated from the 2001 Russian original by the authors. Translations of Mathematical Monographs, 226. American Mathematical Society, Providence, RI, 2005. [57] D. Markiewicz, On the product system of a completely positive semigroup, J. Funct. Anal. 200:1 (2003), 237 -- 280. [58] K. Matsumoto, On C*-algebras associated with subshifts, Internat. J. Math. 8:3 (1997), 357 -- 374. [59] K. Matsumoto, K-Theory for C∗-algebras associated with subshifts, Math. Scand. 82:2 (1998), 237 -- 255. [60] K. Matsumoto, Dimension groups for subshifts and simplicity of the associated C*-algebras, J. Math. Soc. Japan 51:3 (1999), 679 -- 698. [61] K. Matsumoto, C*-algebras associated with presentations of subshifts, Doc. Math. 7 (2002), 1 -- 30. [62] P. S. Muhly, D. Pask and M. Tomforde, Strong Shift Equivalence of C*-correspondences, Israel J. Math. 167 (2008), 315 -- 345. [63] P. S. Muhly and B. Solel, Tensor algebras over C*-correspondences: representations, dilations and C*-envelopes, J. Funct. Anal. 158:2 (1998), 389 -- 457. [64] P. S. Muhly and B. Solel, On the Morita Equivalence of Tensor Algebras, Proc. London Math. Soc. (3) 81:1 (2000), 113 -- 168. [65] P. S. Muhly and B. Solel, Quantum Markov processes (correspondences and dilations), Internat. J. Math. 13:8 (2002), 863 -- 906. [66] P. S. Muhly and M. Tomforde, Adding tails to C*-correspondences, Doc. Math. 9 (2004), 79 -- 106. [67] M. Mukherjee, Index computation for amalgamated products of product systems, Banach J. Math. Anal. 5:1 (2011), 148 -- 166. [68] W. Parry and D. Sullivan, A topological invariant of flows on 1-dimensional spaces, Topology 14:4 (1975), 297 -- 299. [69] W. L. Paschke, Inner product modules over B*-algebras, Trans. Amer. Math. Soc. 182 (1973), 443 -- 468. [70] V. I. Paulsen, Completely bounded maps and operator algebras, Cambridge Studies in Advanced Math- ematics, vol. 78. Cambridge University Press, Cambridge, 2002. [71] J. R. Peters, Semicrossed products of C*-algebras, J. Funct. Anal. 59:3 (1984), 498 -- 534. [72] M. V. Pimsner, A class of C*-algebras generalizing both Cuntz-Krieger algebras and crossed products by Z, Free probability theory (Waterloo, ON, 1995), 189 -- 212, Fields Inst. Commun., 12, Amer. Math. Soc., Providence, RI, 1997. [73] G. Popescu, Von Neumann inequality for (B(H)n)1, Math. Scand. 68 (1991), 292 -- 304. [74] G. Popescu, Non-commutative disc algebras and their representations, Proc. Amer. Math. Soc. 124:7 (1996), 2137 -- 2148. [75] G. Popescu, Operator theory on noncommutative varieties, Indiana Univ. Math. J. 5:2 (2006), 389 -- 442. [76] I. Raeburn, Graph Algebras, CBMS Regional Conference Series in Mathematics, 103. American Math- ematical Society, Providence, RI, 2005. [77] J. Schweizer, Dilations of C*-correspondences and the simplicity of Cuntz-Pimsner algebras, J. Funct. Anal. 180:2 (2001), 404 -- 425. [78] O. M. Shalit and M. Skeide, Three commuting, unital, completely positive maps that have no minimal dilation, Integral Equations Operator Theory 71:1 (2011), 55 -- 63. [79] O. M. Shalit and M. Skeide, CP-Semigroups and Dilations; Subproduct Systems and Superproduct Sys- tems: The Multi-Parameter Case and Beyond, in preparation. [80] O. M. Shalit and B. Solel, Subproduct Systems, Doc. Math. 14 (2009), 801 -- 868. [81] A. M. Sinclair, Homomorphisms from C0(R), J. London Math. Soc. (2) 11:2 (1975), 165 -- 174. [82] B. Solel, You can see the arrows in a quiver operator algebra, J. Aust. Math. Soc. 77:1 (2004), 111 -- 122. 76 [83] B. Tsirelson, Graded algebras and subproduct systems: dimension two, preprint (2009), arXiv:0905.4418. [84] B. Tsirelson, Subproduct systems of Hilbert spaces: dimension two, preprint (2009), arXiv:0906.4255. [85] A. Vernik, Dilations of CP maps commuting according to a graph, Houston J. Math. 42:4 (2016), 1291 -- 1329. [86] A. Viselter, Covariant representations of subproduct systems, Proc. Lond. Math. Soc. (3) 102:4 (2011), 767 -- 800. [87] A. Viselter, Cuntz-Pimsner algebras for subproduct systems, Internat. J. Math. 23:8 (2012), 1250081, 32pp. [88] R. F. Williams, Classification of subshifts of finite type, Ann. of Math. (2) 98 (1973), 120 -- 153; erratum, Ann. of Math. (2) 99 (1974), 380 -- 381. School of Mathematics and Statistics, Newcastle University, Newcastle upon Tyne, NE1 7RU, UK E-mail address: [email protected] Department of Mathematics, Technion - Israel Institute of Technology, Haifa 3200003, Israel E-mail address: [email protected] 77
1512.07593
2
1512
2016-03-07T14:55:36
Regularity of distributions of Wigner integrals
[ "math.OA", "math.PR" ]
Wigner integrals and the corresponding Wigner chaos were introduced by P. Biane and R. Speicher in 1998 as a non-commutative counterpart of classical Wiener-It\^o integrals and the corresponding Wiener-It\^o chaos, respectively, in free probability. In the classical case, a famous result of I. Shigekawa states that non-trivial elements in the finite Wiener-It\^o chaos have an absolutely continuous distribution. We provide here a first contribution to such regularity questions for Wigner integrals by showing that the distribution of non-trivial elements in the finite Wigner chaos cannot have atoms. This answers a question of I. Nourdin and G. Peccati. For doing so, we establish the notion of directional gradients in the context of the free Malliavin calculus. These directional gradients bridge between free Malliavin calculus and the theory of non-commutative derivations as initiated by D. Voiculescu and Y. Dabrowski. Methods recently invented by R. Speicher, M. Weber, and the author for treating similar questions in the case of finitely many variables are extended, such that they apply to directional gradients. This approach also excludes zero-divisors for the considered elements in the finite Wigner chaos.
math.OA
math
REGULARITY OF DISTRIBUTIONS OF WIGNER INTEGRALS TOBIAS MAI Abstract. Wigner integrals and the corresponding Wigner chaos were introduced by P. Biane and R. Speicher in 1998 as a non-commutative counterpart of classical Wiener-Ito integrals and the corresponding Wiener-Ito chaos, respectively, in free probability. In the classical case, a famous result of I. Shigekawa states that non-trivial elements in the finite Wiener-Ito chaos have an absolutely continuous distribution. We provide here a first contribution to such regularity questions for Wigner integrals by showing that the distribution of non-trivial elements in the finite Wigner chaos cannot have atoms. This answers a question of I. Nourdin and G. Peccati. For doing so, we establish the notion of directional gradients in the context of the free Malliavin calculus. These directional gradients bridge between free Malliavin calculus and the theory of non-commutative derivations as initiated by D. Voiculescu and Y. Dabrowski. Methods recently invented by R. Speicher, M. Weber, and the author for treating similar questions in the case of finitely many variables are extended, such that they apply to direc- tional gradients. This approach also excludes zero-divisors for the considered elements in the finite Wigner chaos. 1. Introduction In 1998, P. Biane and R. Speicher established with their seminal work [BS98] a non- commutative counterpart of classical stochastic calculus and Malliavin calculus in the realm of free probability. In particular, they introduced there the so-called (multiple) Wigner integrals I S n (f ) = ZRn + f (t1, . . . , tn) dSt1 · · · dStn for f ∈ L2(Rn +) on R+ = [0,∞) as the free counterpart of the classical (multiple) Wiener- Ito integrals [Wie38, Ito51, Ito52]. Despite some clear peculiarities of these free objects, their construction proceeds to a great extend parallel to the classical case, roughly speaking by replacing the classical Brownian motion by its free relative (St)t≥0. In analogy to the classical Wiener-Ito chaos, these Wigner integrals form the so-called Wigner chaos, which likewise enjoys many properties similar to the classical Wiener-Ito chaos; e.g. [KNPS12]. Date: October 16, 2018. 2000 Mathematics Subject Classification. 46L54 (46L53, 46L57). Key words and phrases. free probability theory, Wigner integrals, commutative derivations, zero-divisors, absence of atoms. free Malliavin calculus, non- This work was supported by the ERC Advanced Grant "Non-commutative distributions in free probabil- ity" and by a grant from the DFG (SP-419-8/1). The author wants to express his thanks to Roland Speicher for many inspiring discussions on the topic and his valuable comments during the preparation of this article. Moreover, the author is grateful to Yoann Dabrowski for some useful remarks at the final stage, to Mehmet Madensoy for bringing some of the ref- erences that are listed at the end to the author's attention, and to Soren Moller, together with whom the author once started to learn free stochastic calculus. Finally, the author wants to thank Ivan Nourdin and Giovanni Peccati for very interesting and fruitful discussions on this article and beyond. 1 2 T. MAI We point out that the increments of the free Brownian motion (St)t≥0 carry the semicircular distribution as the free equivalent of the normal distribution from classical probability theory. It might seem strange at first sight that the nomenclature of Wigner integrals refers explicitly to Wigner, although his work clearly predates the birth of free stochastic calculus. However, this simply highlights the very important fact that the semicircular distribution already appeared in Wigner's famous semicircle law and that this rather surprising connection to random matrix theory, which was later clarified by Voiculescu, marks the starting point of an extremely fruitful interaction between random matrix theory and the theory of operator algebras. Classical Malliavin calculus has many important applications (cf. [Nua06, Nua09]). In particular, it became prominent for its use in treating regularity questions in different situa- tions, as e.g. for distributions of random variables in the Wiener-Ito chaos. For instance, it was used by Shigekawa [Shi80] for proving that any non-trivial element in the finite Wiener- Ito chaos, i.e. any non-constant finite sum of Wiener-Ito integrals, has a distribution which is absolutely continuous with respect to the Lebesgue measure. In contrast, in the world of free probability, distributions of non-commutative random variables that appear in the Wigner chaos are poorly understood. The aim of this paper is a first step towards a better understanding of these distributions by answering one of the fundamental questions formulated by Nourdin and Peccati in [NP13, Remark 1.6], namely: can the distribution of any non-constant self-adjoint Wigner integral have atoms or not? We will see that the answer to this question is no in full generality. Even more, we will show that the distribution of self-adjoint elements in the finite Wigner chaos, i.e. non-commutative random variables of the form I S 1 (f ) + I S 2 (f2) + · · · + I S N (fN ) with mirror-symmetric fn ∈ L2(Rn is the content of of our main Theorem 2.4. +) for n = 1, . . . , N and fN 6= 0, cannot have atoms. This Although this result is clearly in accord with the classical result of Shigekawa [Shi80], the proof of Theorem 2.4 uses completely different methods. Shigekawa's approach is based on arguments which are specially adapted to the commutative setting. In fact, he uses Malliavin's Lemma, which is a powerful result that provides a sufficient condition for a measure on Rd to be absolutely continuous with respect to Lebesgue measure. The non- commutativity in our situation forces us therefore to follow a totally different strategy, which is inspired by recently developed methods [MSW15, Shl14]. In free probability, regularity questions of this type were successfully addressed only quite recently [SS15, MSW14, Shl14, MSW15, CS15]. Our considerations here are very much based on the paper [MSW15], where it was shown that in a von Neumann algebra M, which is en- dowed with a faithful normal tracial state τ , the distribution of any non-constant self-adjoint polynomial expression P (X1, . . . , Xn) in finitely many self-adjoint variables X1, . . . , Xn ∈ M does not have atoms if the so-called non-microstates free entropy dimension δ∗(X1, . . . , Xn) is maximal, i.e., if it satisfies the condition δ∗(X1, . . . , Xn) = n. We note that the quantity δ∗(X1, . . . , Xn) has its origin among other important quantities in the work of Voiculescu. He transferred in a groundbreaking series of papers [Voi93, Voi94, Voi96, Voi97, Voi98, Voi99] (see also the survey article [Voi02]) the classical notions of entropy and Fisher information to the non-commutative world. At the base of our work are techniques from the so-called non-microstates approach presented in [Voi98, Voi99]. REGULARITY OF DISTRIBUTIONS OF WIGNER INTEGRALS 3 Formulated in general terms, so that it can be applied in our situation, the method of [MSW15] works as follows: (1) Rephrase the question of absence of atoms in more algebraic terms as a question about the absence of (certain) zero-divisors. (2) Prove that zero-divisors survive under special operations that are built on non- commutative derivations. This means that zero-divisors for some particular non- commutative random variable induce zero-divisors for some other non-commutative random variables of "lower degree", where the term "degree" refers to the degree of the considered polynomial, or in general to some natural grading on the space of non-commutative random variables under consideration. (3) Iterate the procedure of (2) until reaching a non-commutative random variable of degree zero and check that the obtained element cannot be zero under the imposed conditions on the initial non-commutative random variable. This will lead to a con- tradiction and hence excludes zero-divisors. It might be of independent interest that Step (1) establishes a very interesting relationship to the work of Linnell [Lin91, Lin92, Lin93, Lin98] on analytic versions of the zero divisor conjecture, particularly in the case of the free group. In fact, we will prove the more general statement that the product of any non-commutative random variable in the finite Wigner chaos, which is non-zero, with any non-zero element from the von Neumann algebra generated by the underlying free Brownian motion cannot be zero as well. The crucial part is Step (2), which relies in [MSW15] as well as in our considerations heavily on results of Dabrowski [Dab10, Dab14], concerning bounds for the non-commutative derivatives that underlie the non-microstates approach to free Fisher information and free entropy of [Voi97] and also for more general derivations. In contrast to the preceding studies, which especially concern the case of finitely many variables, the underlying von Neumann algebra in the setting of Wigner integrals is generated by a free Brownian motion (St)t≥0 and therefore by an uncountable family of semicircular ele- ments, indexed by the continuous parameter t ≥ 0. Accordingly, the role of non-commutative derivatives in [MSW14, MSW15] is taken over here by the directional gradient operators of free Malliavin calculus. Thus, the subsequent investigations can be seen as a continuous extension of the previous work [SS15, MSW14, Shl14, MSW15, CS15]. In [MSW14], which is an earlier version of [MSW15], the absence of atoms in the distribu- tion of P (X1, . . . , Xn) for non-constant self-adjoint polynomials P was first shown under the stronger assumption of finite non-microstates free Fisher information Φ∗(X1, . . . , Xn). Based on these ideas, Shlyakhtenko [Shl14] was able to prove a significant extension, namely to the most general case of full non-microstates entropy dimension δ∗(X1, . . . , Xn), by involving different techniques from [CS05]. However, shortly after [Shl14], the authors of [MSW14] were also able to upgrade their own methods to this generality, which led to the final version [MSW15]. Deep results of Shlyakhtenko and Skoufranis [SS15] characterize the possible sizes of atoms that can appear in distributions of polynomial expressions P (X1, . . . , Xn) in non- commutative random variables X1, . . . , Xn, which have not necessarily non-atomic distribu- tions, (and even more matrices (Pij(X1, . . . , Xn))d i,j=1 thereof) under the assumption that X1, . . . , Xn are freely independent. Since the non-microstates free entropy is additive for freely independent variables and since in the case of a single self-adjoint variable X the max- imality condition δ∗(X) = 1 holds if and only if the distribution of X has no atomic part, 4 T. MAI the results from [MSW15, Shl14] clearly generalize some parts of the statements given in [SS15]. However, the full range of regularity results presented in [SS15] is still out of reach in this generality, but nevertheless, one expects that indeed for most of these properties rather the maximality of the non-microstates free entropy dimension matters than the free independence of the involved variables. We point out that certain questions concerning the non-singularity and absolute continu- ity of distributions were addressed recently by Charlesworth and Shlyakhtenko [CS15], in continuation of [Shl14]. The paper is organized as follows. In Section 2, we state our main result Theorem 2.4 on the regularity of distributions of Wigner integrals. For reader's convenience, we recall there also the fundamental definition of a free Brownian motion and the construction of Wigner integrals, as it can be found in the seminal work [BS98]. This exposition of the foundations of free stochastic calculus will then be continued in Section 3. In particular, we will define there the main operators of free Malliavin calculus and collect some results from [BS98], which will be used later on. Section 4 is then devoted to the theory of non-commutative derivations. At first, we will put several results from [Voi98] and [Dab10] (see also [Dab14]) in a uniform framework. Based on this, we will then obtain a significant generalization of a result that was obtained in [MSW15], namely Proposition 4.14, which is at the core of Step (2) and hence crucial for the proof of Theorem 2.4. Finally, in Section 5, we will piece together these ingredients for the actual proof of Theorem 2.4. For this purpose, we will introduce the notion of directional gradients. The proof itself relies then on the fact that directional gradients, which belong by definition to free Malliavin calculus as presented Section 3, fit also nicely into the general framework of non-commutative derivations as considered in Section 4. Indeed, this will allow us to follow the aforementioned strategy in the spirit of [MSW14, MSW15]. Contents Introduction 1. 2. Wigner integrals and regularity of their distributions 2.1. Non-commutative probability spaces and distributions 2.2. Free Brownian motion 2.3. Wigner integrals 2.4. Main Theorem 3. Free stochastic calculus 3.1. Biprocesses 3.2. The free Brownian motion on the full Fock space 3.3. Free Malliavin calculus 4. Non-commutative derivations 4.1. Voiculescu's formulas for δ∗ 4.2. Dabrowski's inequalities 4.3. Survival of zero divisors 5. Proof of Theorem 2.4 5.1. Directional gradients 5.2. Reduction by directional gradients 5.3. How to control the reduction 5.4. Absence of zero divisors References 1 5 5 6 7 9 10 10 13 17 21 23 26 31 34 34 37 39 41 42 REGULARITY OF DISTRIBUTIONS OF WIGNER INTEGRALS 5 2. Wigner integrals and regularity of their distributions In this section, we provide all basic terminology and background knowledge as far as it is needed for stating our main result, Theorem 2.4. First of all, we will briefly recall some very basic concepts of free probability, before we proceed by giving the definition of a free Brownian motion and by presenting the construction of free Wigner integrals as they were introduced by Biane and Speicher in [BS98]; see also [Spe03] and [KNPS12]. The introduction to free stochastic calculus will be continued later in Section 3. At the beginning, a few words on tensor products are in order. Throughout the paper, tensor products are understood as tensor products over the complex numbers C. Moreover, we lay down here that the purely algebraic tensor product of complex vector spaces or complex algebras will be denoted by ⊙, whereas the more familiar symbol ⊗ is reserved for its "natural" closure in the corresponding analytic setting, as for instance for Hilbert spaces or von Neumann algebras. Since the tensor sign will appear mostly in its closed version, this convention saves us from decorating the tensor signs repeatedly with fancy tags and hence keeps the notation as simple as possible. 2.1. Non-commutative probability spaces and distributions. The actual amount of techniques from free probability theory that are needed explicitly below is surprisingly small. The reason is that they are mostly hidden in the quoted results from free stochastic calculus and thus the computations involving them are just outsourced to other papers. However, we prefer to give a separate introduction to the very basic concepts of free probability theory, since it supplies the right language for our considerations. Any reader, who is interested in a more detailed introduction to free probability theory, is cordially invited to have a look at [VDN92], [Voi00], or [NS06] for instance. a non-commutative probability space (which is actually commutative). At the basis of free probability are non-commutative probability spaces. A non-commutative probability space (A, φ) consists of a unital complex algebra A and a linear functional φ : A → C that satisfies φ(1) = 1. Referring to classical probability theory, elements of A are called non-commutative random variables and φ is called expectation. This nomenclature is justified by the observation that any classical probability space (Ω, Σ, P) induces by A = L∞(Ω, Σ, P) and φ(X) = RΩ X(ω)dP(ω) a standard example of In the generality of this purely algebraic setting, we can already introduce the notion of free independence. Unital subalgebras (Ai)i∈I of A are called freely independent (or just free), if for any choice of finitely many indices i1, . . . , in ∈ I, n ∈ N, satisfying i1 6= i2, i2 6= i3, . . . , in−1 6= in, and for any choice of elements Xk ∈ Aik with φ(Xk) = 0 for k = 1, . . . , n, the condition φ(X1 · · · Xn) = 0 is fulfilled. Consequently, we call non-commutative random variables (Xi)i∈I in A freely independent (or just free), if the subalgebras (Ai)i∈I are freely independent, where Ai denotes for each i ∈ I the unital subalgebra of A that is generated by Xi. Roughly speaking, free independence provides a rule to calculate mixed moments. For any tuple (X1, . . . , Xn) of non-commutative random variables in A, we refer to the collection of all moments φ(Xi1Xi2 · · · Xik ), k ≥ 0, 1 ≤ i1, . . . , ik ≤ n 6 T. MAI (including also the trivial moment φ(1) = 1) as the (joint) distribution µX1,...,Xn of (X1, . . . , Xn). If the non-commutative random variables X1, . . . , Xn are freely independent, then the distri- bution µX1,...,Xn is completely determined by the single variable distributions µX1, . . . , µXn. For the seek of completeness, we point out that one can make this relation much more ex- plicit by using the powerful combinatorial concept of free cumulants as it was introduced to free probability by Speicher. It was a fundamental observation of Voiculescu that the distribution µX+Y of two freely independent non-commutative random variables X and Y depends only on the distributions µX and µY of X and Y , respectively, and not on the concrete realization of X and Y . Consequently, he defined the free additive convolution ⊞ on abstract distributions by µX ⊞ µY := µX+Y . For our purposes, it is necessary to impose some additional analytic structure. If we consider a C∗-probability space (A, φ), i.e. a non-commutative probability space (A, φ), where A is a unital C∗-algebra and φ a state on A, then the distribution µX of any self- adjoint non-commutative random variable X in A can be identified with the compactly supported Borel probability measure µX on the real line R that is uniquely determined by the condition φ(X k) = ZR tk dµX(t) for k = 0, 1, 2, . . . . Accordingly, the free additive convolution ⊞ gives rise to a binary operation on all (compactly supported) Borel probability measures on R. Here, we will mainly work in the setting of tracial W ∗-probability spaces. A tracial W ∗- probability space (M, τ ) means a non-commutative probability space (M, τ ), where M is a von Neumann algebra and τ a faithful normal tracial state on M. If (M1, τ1) and (M2, τ2) are two tracial W ∗-probability spaces, then also their von Neumann algebra tensor product M1 ⊗ M2 becomes, endowed with the tensor product state τ1 ⊗ τ2, a tracial W ∗-probability space. Another construction that will be used repeatedly in the subsequent considerations are the non-commutative Lp-spaces. Given any tracial W ∗-probability space (M, τ ), we may introduce the non-commutative Lp-spaces Lp(M, τ ) for 1 ≤ p ≤ ∞ as the completion of M with respect to the norm kxkLp(M,τ ) := τ(cid:0)(x∗x) p , and for p = ∞ simply by L∞(M, τ ) := M where we put kxkL∞(M,τ ) := kxk. Whenever it is not necessary to indicate explicitly the underlying von Neumann algebra, we will abbreviate k · kp := k · kLp(M,τ ). 2(cid:1) 1 p 2.2. Free Brownian motion. Like the classical Brownian motion in the case of Wiener-Ito integrals, the free Brownian motion is the fundamental object in free stochastic analysis and underlies in particular the construction of Wigner integrals. Thus, we want to recall now its definition. Note that the definition itself will reflect the important fact that the role of the normal distribution in classical probability is taken over in free probability by the semicircular dis- tribution as its free counterpart. We will denote by σt the semicircular distribution with mean 0 and variance t > 0, i.e. the compactly supported probability measure σt on the real line R that is given by dσt(x) = 1 2πt √4t − x2 1[−2√t,2√t](x) dx. REGULARITY OF DISTRIBUTIONS OF WIGNER INTEGRALS 7 Note that (σt)t≥0 forms a semi-group with respect to the free additive convolution, i.e. we have that σs ⊞ σt = σs+t holds for all s, t ≥ 0. Definition 2.1. Let (M, τ ) be a tracial W ∗-probability space. A family (St)t≥0 of operators in (M, τ ) is called free Brownian motion, if there exists a filtration (Mt)t≥0 of M, i.e. a family (Mt)t≥0 of von Neumann subalgebras Mt of M with whenever s ≤ t, Ms ⊆ Mt such that the following conditions are satisfied: • We have S0 = 0 and St = S∗t ∈ Mt for all t ≥ 0. • For each t > 0, the distribution of St is the semicircular distribution σt. • For all 0 ≤ s < t, the distribution of St − Ss is the semicircular distribution σt−s. • For all 0 ≤ s < t, the increment St − Ss is free from Ms, which means more precisely that the unital subalgebra generated by St − Ss is free from Ms. A free Brownian motion can be constructed in several ways. For instance, one construction gives the free Brownian motion as the limit of matrix-valued classical Brownian motions as the dimension tends to infinity. In contrast to this certainly appealing but rather indirect approach, we will present in Subsection 3.2 a construction of the free Brownian motion on the full Fock space over the Hilbert space L2(R+) of all square-integrable functions on the positive real half-line R+ := [0,∞). This has the advantage that it will not only prove the existence of the free Brownian motion but it will also give an additional structure to this important object, which is in fact the starting point of free Malliavin calculus. However, for the moment, we take the existence of a free Brownian motion for granted. 2.3. Wigner integrals. Presuming the existence of a free Brownian motion (St)t≥0 in a W ∗-probability space (M, τ ) with respect to a filtration (Mt)t≥0 of M, we may introduce now (multiple) Wigner integrals integrals with respect to (St)t≥0. Definition 2.2. Let n ∈ N be given. We denote by Dn ⊂ Rn in Rn + the collection of all diagonals +, i.e. Dn := {(t1, . . . , tn) ∈ Rn + ti = tj for some 1 ≤ i, j ≤ n with i 6= j}. The construction of the (multiple) Wigner integral I S n (f ) for any function f ∈ L2(Rn +) proceeds as follows. • For any indicator function f = 1E of some set E = [s1, t1] × · · · × [sn, tn] ⊂ Rn + that satisfies E ∩ Dn = ∅, we define I S n (f ) by I S n (f ) = (St1 − Ss1)· · · (Stn − Ssn). • By linearity, we extend I S n to all off-diagonal step functions, i.e. to all step functions f = m Xj=1 aj1Ej on Rn +, where each set Ej ⊂ Rn + is of the form Ej = [sj,1, tj,1] × · · · × [sj,n, tj,n] and satisfies Ej ∩ Dn = ∅. 8 T. MAI • Since off-diagonal step functions are dense in L2(Rn +) (an important fact, which is actually not hard to prove, but which is definitely worth to think about for a moment) and since the Ito isometry τ (I S n (f )∗I S n (g)) = hg, fiL2(Rn +) holds for all off-diagonal step functions f and g, we may finally extend I S to L2(Rn +). n isometrically For given f ∈ L2(Rn +), we will write I S n (f ) = ZRn + f (t1, . . . , tn) dSt1 · · · dStn. Note that multiple Wigner integrals I S +) by definition ele- n (f ) belongs to M for each f ∈ L2(Rn ments of L2(M, τ ). But in fact, it turns out that I S +) (and actually, to be more precise, it belongs to the C∗-subalgebra of M that is generated by the free Brownian motion (St)t≥0). This is an immediate consequence of the fact that off-diagonal step functions are dense in L2(Rn +) and of [BS98, Theorem 5.3.4], which tells us that the operator norm can be bounded by a kind of Haagerup inequality, namely n (f ) are for general f ∈ L2(Rn (2.1) ZRn + (cid:13)(cid:13)(cid:13) f (t1, . . . , tn) dSt1 · · · dStn(cid:13)(cid:13)(cid:13) ≤ (n + 1)kfkL2(Rn +) for all f ∈ L2(Rn +). Since Wigner integrals are bounded linear operators, we are of course allowed to multiply them, and it is therefore natural to ask, whether one can describe this operation also on the level of the corresponding functions. Indeed, this turns out to be possible and it leads to a free counterpart of Ito's formula (see, for example, [Spe03, Theorem 2.11]). Although this result appears in many different formulations, it always reflects the same inherent structure that shows up, roughly speaking, under multiplication. We mention here the following version, which allows us to decompose products of Wigner integrals explicitly as linear combinations of Wigner integrals. Theorem 2.3 (Biane and Speicher, 1998, [BS98]). Let f ∈ L2(Rn any 0 ≤ p ≤ min{n, m}, we define the p'th contraction of f and g by +) and g ∈ L2(Rm + ). For f p a g(t1, . . . , tn+m−2p) = ZRp + f (t1, . . . , tn−p, s1, . . . , sp) g(sp, . . . , s1, tn−p+1, . . . , tn+m−2p) ds1 . . . dsp. Then the Ito formula holds. I S n (f )I S m(g) = min{n,m} Xp=0 I S n+m−2p(f p a g) In principle, all previously collected facts about Wigner integrals put them in the most convenient setting of non-commutative probability, such that we can already talk about their (joint) distributions in a purely combinatorial sense. However, since we work here in the regular setting of W ∗-probability spaces, we also want to study distributions of Wigner integrals in a stronger analytic sense, namely as (compactly supported) probability measures. REGULARITY OF DISTRIBUTIONS OF WIGNER INTEGRALS 9 Thus, we should have a criterion on the level of integrands that allows us to guarantee that the corresponding Wigner integral is self-adjoint. This criterion is provided by mirror symmetry. It follows immediately from the definition of Wigner integrals that n (f )∗ = I S I S holds, where the function f∗ ∈ L2(Rn for all f ∈ L2(Rn +) n (f∗) +) is determined for any f ∈ L2(Rn +) by f∗(t1, t2, . . . , tn) = f (tn, . . . , t2, t1) for Lebesgue almost all (t1, . . . , tn) ∈ Rn f = f∗ gives a self-adjoint Wigner integral I S symmetric. +. As a consequence, any f ∈ L2(Rn n (f ). We will call such f ∈ L2(Rn +) satisfying +) mirror 2.4. Main Theorem. Here, we are interested in properties of the distributions of Wigner integrals I S n (f ) = ZRn + f (t1, . . . , tn) dSt1 · · · dStn +), and, more generally, in distributions of finite for mirror symmetric functions f ∈ L2(Rn sums of such Wigner integrals like Y = I S 1 (f1) + I S 2 (f2) + · · · + I S N (fN ) for some N ∈ N and mirror symmetric functions fn ∈ L2(Rn +) for n = 1, . . . , N with fN 6= 0. Surely one of the most basic questions one can ask about distributions in general is whether their support is connected or not. Basic functional analysis yields that this question can be reformulated in more operator algebraic terms to a question about the existence of non-trivial projections in the C∗-algebra that is generated by the considered operator. Fortunately, this translation is also helpful in our situation: As we have mentioned above, Wigner integrals are in fact elements of the C∗-algebra that is generated by the free Brownian motion (St)t≥0. Hence, by quoting a results obtained by Guionnet and Shlyakhtenko in [GS09], which ex- cludes non-trivial projections in C∗({St t ≥ 0}), we may conclude without further effort that the distribution µY of any operator Y as above must have connected support. However, apart from this observation, almost nothing was known until now about regu- larity properties of these distributions. In particular, as it was formulated by Nourdin and Peccati in [NP13, Remark 1.6], it remained an open questions whether the distribution of Wigner integrals of mirror symmetric functions being non-zero (except, of course, in the chaos of order zero) may have atoms or not. We are going to answer this question here by showing that the distribution of any such Wigner integral of a non-zero mirror symmetric function (and even of any non-constant finite sum of such Wigner integrals) does not have atoms. Recall that an atom of a Borel probability measure µ on R means some α ∈ R satisfying The statement of the main theorem of this paper reads as follows. the condition µ({α}) 6= 0. Theorem 2.4. For given N ∈ N, we consider mirror symmetric functions fn ∈ L2(Rn n = 1, . . . , N, where we assume that fN 6= 0. Then, the distribution µY of +) for regarded as an element in (M, τ ), has no atoms. Y := I S 1 (f1) + I S 2 (f2) + · · · + I S N (fN ), 10 T. MAI The proof of Theorem 2.4 will be given in Section 5. We stress that the above statement clearly stays valid if we add to Y a constant multiple of the identity. In fact, this will be a direct outcome of the proof of Theorem 2.4, since we will use the chaos decomposition to deal with such shifts in a uniform way. More precisely, we can just encode constant multiples of the identity by the chaos of order zero. Furthermore, we point out that Theorem 2.4 corresponds nicely to a classical result of Shigekawa [Shi78, Shi80] (although its proof uses completely different methods for which there are by now no free analogues), which states that any non-trivial finite sum of Wiener- Ito integrals has an absolutely continuous distribution, and hence cannot have atoms. Thus, confident of the far reaching parallelism between classical and free probability, we are tempted to conjecture in accordance with [Spe13] that the analogy between Wiener-Ito integrals and Wigner integrals goes even further, namely that any Y like in Theorem 2.4 has in fact an absolutely continuous distribution. We leave this question to further investigations. 3. Free stochastic calculus One of the main pillars on which the proof of Theorem 2.4 rests is free stochastic calculus as it was introduced by Biane and Speicher in [BS98]. For readers convenience, we recall in this section the basic definitions and some results of this theory as far as necessary. First of all, we will introduce the notion of biprocesses. Secondly, we will describe the concrete realization of the free Brownian motion on the full Fock space over L2(R+). This additional structure will finally allow us to introduce the basic operators of Malliavin calculus. 3.1. Biprocesses. We broach now the theory of biprocesses. Our exposition here heavily relies on [BS98], [Spe03], and [KNPS12]. Let us first introduce a few general notions. We denote by E(R+) the space of all complex valued functions f on R+, which can be written as a finite sum f = n Xj=1 aj 1Ej for some intervals E1, . . . , En ⊆ R+ of the form Ej = [sj, tj) with 0 ≤ sj < tj < ∞ for j = 1, . . . , n and complex numbers a1, . . . , an ∈ C. As usually, 1E denotes the indicator function of a subset E ⊆ R+. It is easy to see that E(R+) is in fact a complex algebra. For any unital complex algebra A, the algebraic tensor product E(R+,A) := E(R+) ⊙ A consists of all functions f defined on R+ and taking values in A, which can be written as n f = Xj=1 Aj 1Ej for some intervals E1, . . . , En ⊆ R+ of the form Ej = [sj, tj) with 0 ≤ sj < tj < ∞ for j = 1, . . . , n and elements A1, . . . , An ∈ A. 3.1.1. Definition of biprocesses. We are prepared now to define biprocesses. For the remain- ing part of this subsection, we fix a tracial W ∗-probability space (M, τ ) for which a filtration (Mt)t≥0 exists. Definition 3.1. We distinguish several types of biprocesses, which are build on each other. Their definition proceeds as follows: REGULARITY OF DISTRIBUTIONS OF WIGNER INTEGRALS 11 (i) The elements U : R+ → M ⊙ M, t 7→ Ut of E(R+, M ⊙ M) are called simple biprocesses. (ii) A simple biprocess U : R+ → M ⊙ M is called adapted, if the condition Ut ∈ Mt ⊙ Mt is satisfied for all t ≥ 0. The set of all adapted simple biprocesses will be denoted by E a(R+, M ⊙ M). (iii) We denote by Bp for 1 ≤ p ≤ ∞ the completion of E(R+, M ⊙ M), with respect to the norm k · kBp, which is given by kUkBp := (cid:18)ZR+ kUtk2 An element of Bp is called an Lp-biprocess. by Ba p. Elements of Ba Lp(M⊗M,τ⊗τ ) dt(cid:19) 1 2 . (iv) For 1 ≤ p ≤ ∞, the closure of E a(R+, M ⊙ M) with respect to k·kBp will be denoted p are called adapted Lp-biprocesses. 3.1.2. Integration of biprocesses. For our purposes, the integration theory of biprocesses is of great importance. We focus here first on the integration of Lp-biprocesses with respect to functions in L2(R+). On the basic level of simple biprocesses, such integrals can be introduced quite easily: if U is any simple biprocess, we may write (3.1) U = n Xj=1 U (j) 1Ej for some intervals E1, . . . , En ⊆ R+ of the form Ej = [sj, tj) with 0 ≤ sj < tj < ∞ for j = 1, . . . , n and certain elements U (1), . . . , U (n) ∈ M ⊙ M. Then, we put n ZR+ Ut h(t) dt := Xj=1 h1Ej , hiL2(R+) U (j), and it is easy to see that the value of this integral does not depend on the concrete choice of the representation (3.1). i.e. disjoint. Sometimes, it is more appropriate to write a given simple biprocess U in standard form, in the form of (3.1), where the intervals E1, . . . , En ⊆ R+ are assumed to be pairwise By the construction presented above, we obtain a sesqui-linear pairing h·,·i : E(R+, M ⊙ M) × L2(R+) → M ⊙ M, which is given by Ut h(t) dt hU, hi := ZR+ for any U ∈ E(R+, M ⊙ M) and h ∈ L2(R+). Since we want to extend h·,·i to a sesqui-linear paring between Bp and L2(R+), we need to study its continuity with respect to k · kBp. This will be done in the following lemma. In the case p = ∞, this property of h·,·i was already mentioned in [BS98]. The general case is probably also well-known to experts, but for the seek of completeness, we include here the straightforward proof. 12 T. MAI Lemma 3.2. Let 1 ≤ p ≤ ∞ be given. For any U ∈ E(R+, M ⊙ M) and h ∈ L2(R+), it holds true that khU, hikLp(M⊗M,τ⊗τ ) ≤ kUkBpkhkL2(R+). Proof. Let U ∈ E(R+, M ⊙ M) and h ∈ L2(R+) be given and write U in standard form n U = Xj=1 For any fixed 1 ≤ p ≤ ∞, we may check that Lp(M⊗M,τ⊗τ ), dt(cid:19) kUkBp = (cid:18)ZR+ kUtk2 1 2 U (j) 1Ej . = (cid:18) n Xj=1 λ1(Ej)kU (j)k2 Lp(M⊗M,τ⊗τ )(cid:19) 1 2 , where λ1 denotes the Lebesgue measure on R. Thus, applying twice the Cauchy-Schwarz inequality yields as desired khU, hikLp(M⊗M,τ⊗τ ) n h1Ej , hiL2(R+) kU (j)kLp(M⊗M,τ⊗τ ) h1Ej , 1Ej hiL2(R+) kU (j)kLp(M⊗M,τ⊗τ ) k1Ej hkL2(R+)k1EjkL2(R+)kU (j)kLp(M⊗M,τ⊗τ ) n n = ≤ Xj=1 Xj=1 Xj=1 ≤ ≤ (cid:18) n Xj=1 ≤ khkL2(R+)kUkBp, k1Ej hk2 L2(R+)(cid:19) 1 2(cid:18) n Xj=1 k1Ejk2 L2(R+)kU (j)k2 Lp(M⊗M,τ⊗τ )(cid:19) 1 2 where we used in addition that due to the pairwise orthogonality of the functions {1Ej h j = 1, . . . , n} holds and that we have k1Ek2 measure. 1 2 k1Ej hk2 L2(R+)(cid:19) (cid:18) n Xj=1 L2(R+) = λ1(E) for any Borel set E ⊆ R+ with finite Lebesgue ≤ khkL2(R+) (cid:3) Due to the inequality that we have established in Lemma 3.2, the definition of h·,·i extends now naturally to Bp. Definition 3.3. For any 1 ≤ p ≤ ∞, the sesqui-linear pairing h·,·i : E(R+, M ⊙ M) × L2(R+) → M ⊙ M, extends continuously according to hU, hi = ZR+ Uth(t) dt, khU, hikLp(M⊗M,τ⊗τ ) ≤ kUkBpkhkL2(R+). REGULARITY OF DISTRIBUTIONS OF WIGNER INTEGRALS 13 to a sesqui-linear pairing h·,·i : Bp × L2(R+) → Lp(M ⊗ M, τ ⊗ τ ). RR+ 3.1.3. Stochastic integrals of biprocesses. Next, we are going to define stochastic integrals Ut♯dSt of biprocesses U with respect to the free Brownian motion (St)t≥0. For this purpose, we first have to introduce the notation ♯, which appears here and repeatedly in the non-commutative setting. Remark 3.4. Let A and B be complex algebras. If M is an A-B-bimodule, then we denote by ♯ the operation (A ⊙ B) × M → M that is determined by linear extension of (a ⊗ b)♯m := a · m · b. Even more, if we would replace here B by its opposite algebra Bop, then ♯ would give rise to a left action of the algebraic tensor product A ⊙ Bop on M. But since the multiplicative structure of A ⊙ B will play a minor role in our considerations, we will not care about this subtlety in the following. Definition 3.5. Let (St)t≥0 be a free Brownian motion in M with respect to its given filtration (Mt)t≥0. • For any simple biprocess U ∈ E(R+, M ⊙ M), we define ZR+ Ut♯dSt := n Xj=1 U (j)♯(Stj − Ssj ) = n mj Xj=1 Xi=1 A(j) i (Stj − Ssj )B(j) i , where U is written in the form (3.1) for intervals Ej = [sj, tj) with 0 ≤ sj < tj < ∞ and elements U (j) ∈ M ⊙ M of the form U (j) = Pmj • If U, V ∈ E a(R+, M⊙M) are simple adapted biprocesses, then the general Wigner-Ito i ⊗ B(j) isometry (cf. [Spe03, Proposition 2.7]) tells us that for j = 1, . . . , n. i i=1 A(j) hZR+ Ut♯dSt,ZR+ Vt♯dSti = ZR+hUt, Vti dt =: hU, V iB2 holds. Thus, we have in particular that ZR+ (cid:13)(cid:13)(cid:13) Ut♯dSt(cid:13)(cid:13)(cid:13)2 = kUkB2 for all U ∈ E a(R+, M ⊙ M). Therefore, the integral RR+ adapted biprocesses to any adapted L2-biprocess U ∈ Ba induced mapping Ut♯dSt extends from simple 2 in such a way that the U 7→ ZR+ Ut♯dSt is isometric from Ba 2 to L2(M, τ ). 3.2. The free Brownian motion on the full Fock space. We come back now to the construction of the free Brownian motion. As we announced earlier, we will do this here in an explicit way on the full Fock space over L2(R+). These techniques will be used to build up free Malliavin calculus, in the same way as classical Malliavin calculus is built on the symmetric Fock space. 14 T. MAI 3.2.1. The full Fock space and field operators. We first recall the construction of the full Fock space over an arbitrary complex Hilbert space. Recall that in the context of complex Hilbert spaces, the symbol ⊙ stands for the algebraic tensor product (over the complex numbers C), whereas its completion with respect to the canonical inner product will be denoted by ⊗. Definition 3.6. Let (H,h·,·iH) be a complex Hilbert space. We define the full Fock space F (H) associated to H as the complex Hilbert space that is given by F (H) := ∞ Mn=0 H⊗n, where L is understood as Hilbert space operation. Therein, we declare that H⊗0 := CΩ for some fixed vector Ω of norm 1, which we call the vacuum vector of F (H). More explicitly, the inner product h·,·i on F (H) is determined by the following rules: We have and in the case m = n hg1 ⊗ · · · ⊗ gm, h1 ⊗ · · · ⊗ hni = 0 if m 6= n hg1 ⊗ · · · ⊗ gm, h1 ⊗ · · · ⊗ hmi = hg1, h1iH · · ·hgm, hmiH. Later on, we will also work with some special (non-closed) subspaces of the full Fock space F (H), involving an infinite but algebraic direct sum, namely • Falg(H) := H⊙n, i.e. the subspace of F (H) that consists of finite sums of tensor products of vectors in H, and ∞ Malg n=0 ∞ Malg n=0 • Ffin(H) := H⊗n, i.e. the subspace of F (H) that consists of finite sums of elements in the Hilbert spaces H⊗n. It is clear by definition that we have the inclusions Falg(H) ⊆ Ffin(H) ⊆ F (H) and that both subspaces Falg(H) and Ffin(H) are dense in F (H). On the full Fock space F (H), we may introduce the so-called field operators. In the case H = L2(R+), these operators will provide the desired realization of the free Brownian motion. Definition 3.7. Let (H,h·,·iH) be a complex Hilbert space. For each h ∈ H we introduce the following operators on the full Fock space F (H) over H: (i) The creation operator l(h) ∈ B(F (H)) is determined by l(h) h1 ⊗ · · · ⊗ hn = h ⊗ h1 ⊗ · · · ⊗ hn, l(h) Ω = h. (ii) The annihilation operator l∗(h) ∈ B(F (H)) is given by l∗(h) h1 ⊗ · · · ⊗ hn = hh, h1iHh2 ⊗ · · · ⊗ hn, l∗(h) h1 = hh, h1iHΩ, l∗(h) Ω = 0. n ≥ 2, REGULARITY OF DISTRIBUTIONS OF WIGNER INTEGRALS 15 (iii) The field operator X(h) ∈ B(F (H)) is defined by X(h) := l(h) + l∗(h). An easy calculation shows that we have l∗(h) = l(h)∗ for all h ∈ H, as the notation suggests. As an immediate consequence, X(h) = X(h)∗ holds for each h ∈ H. In order to obtain a W ∗-probability space, in which the free Brownian lives, it is natural to consider the von Neumann algebra generated by field operators X(h) for a sufficiently large family of vectors h. As it turns out, the right choice for this purpose are the "real" vectors h. More formally, we will consider the full Fock space over the complexification HC = H ⊕ iH of any real Hilbert space (H,h·,·iH). The "real" vectors are then naturally those, which are coming from H. We shall make this more precise with the following definition. Definition 3.8. Let H be a real Hilbert space and denote by HC = H ⊕ iH its complexifi- cation. We define the von Neumann algebra S(H) ⊆ B(F (HC)) by We may endow S(H) with the vacuum expectation τ : S(H) → C given by S(H) = vN(cid:0){X(h) h ∈ H}(cid:1). τ (X) = hXΩ, Ωi. Due to the fact that H is a real Hilbert space, we are in the nice situation that τ gives a faithful normal tracial state on S(H). Thus, we have obtained a W ∗-probability space (S(H), τ ). Later on, we will also use the unital ∗-algebra Salg(H) that is given by Salg(H) := alg(cid:0){X(h) h ∈ H}(cid:1). Clearly, Salg(H) ⊆ S(H) ⊆ B(F (HC)). It is a very nice feature of (S(H), τ ) that its L2-space L2(S(H), τ ) can be identified in a natural way with the corresponding full Fock space F (HC). This important observation is at the base of free Malliavin calculus. Since we have for all X1, X2 ∈ S(H) that hX1, X2iL2(S(H),τ ) = τ (X∗2 X1) = h(X∗2 X1)Ω, ΩiF (H) = hX1Ω, X2ΩiF (H), we see that the map admits an isometric extension Φ0 : S(H) → F (HC), X 7→ XΩ Φ : L2(S(H), τ ) → F (HC). The following lemma allows us to conclude that Φ is even more surjective and hence gives the desired isometric isomorphism between L2(S(H), τ ) and F (HC). A proof can be found in [BS98, Section 5.1]. Lemma 3.9. Given h1, . . . , hn ∈ HC, then there exists a unique operator W (h1 ⊗ · · · ⊗ hn) ∈ S(H), called the Wick product of h1 ⊗ · · · ⊗ hn, such that More precisely, if (ej)j∈J is an orthonormal basis of H then W (h1 ⊗ · · · ⊗ hn)Ω = h1 ⊗ · · · ⊗ hn. W (e⊗k1 j1 ⊗ · · · ⊗ e⊗kn jn ) = Uk1(X(ej1))· · · Ukn(X(ejn)), 16 T. MAI where j1 6= j2 6= · · · 6= jn and Uk denotes the k'th (normalized) Chebyshev polynomial of the second kind. These polynomials are determined by U0(X) = 1, U1(X) = X and the recursion Uk+1(X) = XUk(X) − Uk−1(X) for k ≥ 1. Note that the lemma implies in particular that Φ0(Salg(H)) = Falg(H). 3.2.2. F (L2(R+)) and the free Brownian motion. We return now to the actual goal of this subsection, namely the construction of the free Brownian motion. This is achieved by apply- ing the foregoing constructions to the real Hilbert H = L2(R+, R), whose complexification is clearly given by HC ∼= L2(R+). In the W ∗-probability space (S, τ ) where we abbreviate S := S(L2(R+, R)), the free Brownian motion (St)t≥0 is obtained by putting The corresponding filtration (St)t≥0 of S is given by St := X(1[0,t]) for all t ≥ 0. St := vN(cid:0){X(h) h ∈ L2([0, t], R)}(cid:1), where we regard L2([0, t], R) as a subspace of L2(R+, R) via extension by zero. In fact, St is generated as a von Neumann algebra by {Ss 0 ≤ s ≤ t}, while S is generated by {Ss s ≥ 0}. The very concrete realization of the free Brownian motion in the W ∗-probability space (S, τ ) has the advantage that it carries the rich structure provided by the underlying Fock space F := F (L2(R+)) by the isometric isomorphism Φ : L2(S, τ ) → F , which was obtained by isometric extension of the map Φ0 : S → F given by Φ0(X) = XΩ. This will be used in the next subsection on free Malliavin calculus. But before continuing in this direction, we first discuss the chaos decomposition for arbi- trary elements in L2(S, τ ), which emerges from the isomorphism Φ. In the simplest case, it boils down to a nice relation between Wigner integrals and the Wick products as introduced in Lemma 3.9. More precisely, we have for all h1, . . . , hn ∈ L2(R+) that W (h1 ⊗ · · · ⊗ hn) = I S n (h1 ⊗ · · · ⊗ hn) = ZRn + h1(t1)· · · hn(tn) dSt1 · · · dStn. This observation is generalized by the following result. Proposition 3.10 (Proposition 5.3.2. L2(S, τ ) → F is given by in [BS98]). The inverse of the isomorphism Φ : where for any I S : F → L2(S, τ ), f 7→ I S(f ), I S(f ) := ∞ Xn=0 I S n (fn) f = (fn)∞n=0 ∈ ∞ Mn=0 L2(Rn +) ∼= F . REGULARITY OF DISTRIBUTIONS OF WIGNER INTEGRALS 17 This means that each element of L2(S, τ ) has a unique representation in the form I S(f ) for some f ∈ L∞n=0 L2(Rn There is a similar decomposition for L2-biprocesses. Since the mapping I S : F → L2(S, τ ) +), to which we refer as its chaos decomposition. gives rise to an isometric isomorphism we see, by using the natural isometric identifications I S ⊗ I S : F ⊗ F → L2(S, τ ) ⊗ L2(S, τ ), L2(R+,F ⊗ F ) ∼= F ⊗ L2(R+) ⊗ F and L2(R+, L2(S, τ ) ⊗ L2(S, τ )) ∼= L2(S, τ ) ⊗ L2(R+) ⊗ L2(S, τ ) ∼= B2, that I S ⊗ I S induces an isometric isomorphism I S ⊗ I S : L2(R+,F ⊗ F ) → B2, which is again denoted by I S ⊗ I S. More explicitly, this induced isomorphism sends each f : R+ → F ⊗ F , t 7→ ft that belongs to L2(R+,F ⊗ F ) to the L2-biprocess that is given by t 7→ (I S ⊗ I S)(ft). The following diagram offers a clear view on the situation described above. F ⊗ L2(R+) ⊗ F ∼= L2(R+,F ⊗ F ) I S⊗id ⊗I S I S⊗I S L2(S, τ ) ⊗ L2(R+) ⊗ L2(S, τ ) ∼= / B2 We call U = (I S ⊗ I S)(f ) for f ∈ L2(R+,F ⊗ F ) the Wigner chaos expansion of the L2-biprocess U. 3.3. Free Malliavin calculus. Like in the classical case, the basic operators of free Malli- avin calculus are constructed first on the side of the full Fock space and are then transferred to the algebra of field operators via the identification that is provided by the map X 7→ XΩ. 3.3.1. Free Malliavin calculus on F (H). As above in the construction of the free Brownian motion, we begin with the general case of an arbitrary complex Hilbert space H. On the full Fock space F (H) over H, we consider • an unbounded linear operator ∇ : F (H) ⊇ D( ∇) → F (H) ⊗ H ⊗ F (H) with domain D( ∇) = Falg(H), which is determined by the conditions ∇Ω = 0 and n ∇(h1 ⊗ · · · ⊗ hn) := Xj=1 (h1 ⊗ · · · ⊗ hj−1) ⊗ hj ⊗ (hj+1 ⊗ · · · ⊗ hn), where the tensor products appearing in the brackets are understood as Ω if the corresponding set of indices happens to be empty. / /     ✤ ✤ ✤ ✤ ✤ ✤ / 18 T. MAI • an unbounded linear operator δ : F (H) ⊗ H ⊗ F (H) ⊇ D(δ) → F (H) with domain D(δ) = Falg(H) ⊙ H ⊙ Falg(H) by linear extension of δ((h1 ⊗ · · · ⊗ hn) ⊗ h ⊗ (g1 ⊗ · · · ⊗ gm)) δ(Ω ⊗ h ⊗ (g1 ⊗ · · · ⊗ gm)) δ((h1 ⊗ · · · ⊗ hn) ⊗ h ⊗ Ω) δ(Ω ⊗ h ⊗ Ω) := h1 ⊗ · · · ⊗ hn ⊗ h ⊗ g1 ⊗ · · · ⊗ gm, := h ⊗ g1 ⊗ · · · ⊗ gm, := h1 ⊗ · · · ⊗ hn ⊗ h, := h • an unbounded linear operator N : F (H) ⊇ D(N) → F (H) with domain D( N) = Falg(H), which is defined by N Ω = 0 and N (h1 ⊗ · · · ⊗ hn) := n h1 ⊗ · · · ⊗ hn. We collect now a few observations related to the operators ∇ and δ. We grant that some of these statements might appear quite artificial at the first sight, but there actual meaning will become clear after passing from the Fock space to operators defined on it. Remark 3.11. Consider the setting that was described above. (a) A straightforward calculation shows that h ∇y, uiF (H)⊗H⊗F (H) = hy, δ(u)iF (H) (3.2) (3.3) holds for all y ∈ D( ∇) and u ∈ D(δ). (b) If we endow Falg(H) with the multiplication induced by the tensor product ⊗ (in fact, we obtain in this way the tensor algebra over H), we may easily check that ∇ satisfies a kind of product rule, namely ∇(y1 ⊗ y2) = ( ∇y1) · y2 + y1 · ( ∇y2) for all y1, y2 ∈ D( ∇), where · denotes the canonical left and right action, respectively, of Falg(H) on Falg(H) ⊗ H ⊗ Falg(H) that is induced by ⊗, i.e. y1 · (x1 ⊗ h ⊗ x2) · y2 = (y1 ⊗ x1) ⊗ h ⊗ (x2 ⊗ y2). (c) Since the range of ∇ is by definition contained in the domain of δ, the composition δ ◦ ∇ is well-defined. In fact, one has N = δ ◦ ∇. 3.3.2. Free Malliavin calculus on F (L2(R+)). We apply now the preceding construction in the special case, where the Hilbert space H is given by L2(R+). Thus, we may use the isomorphisms I S : F → L2(S, τ ) and I S ⊗ I S : L2(R+,F ⊗ F ) → B2 to pull over • the operator ∇ : F ⊇ D( ∇) → L2(R+,F ⊗ F ) to the so-called gradient operator with domain D(∇) = I S(D( ∇)), ∇ : L2(S, τ ) ⊇ D(∇) → B2 REGULARITY OF DISTRIBUTIONS OF WIGNER INTEGRALS 19 • and the operator to the so-called divergence operator δ : L2(R+,F ⊗ F ) ⊇ D(δ) → F with domain D(δ) = (I S ⊗ I S)(D(δ)), in the obvious way as shown in the following two commutative diagrams. δ : B2 ⊇ D(δ) → L2(S, τ ) I S I S F D( ∇) ∇ L2(R+,F ⊗ F ) I S⊗I S / L2(S, τ ) / D(∇) ∇ / B2 In fact, the above definitions amount to L2(R+,F ⊗ F ) I S⊗I S / B2 D(δ) I S⊗I S / D(δ) δ F δ I S / L2(S, τ ) D(∇) = Salg and D(δ) = Salg ⊙ L2(R) ⊙ Salg, where we abbreviate Salg := Salg(L2(R+, R)). Remark 3.12. We may observe that the properties of the operators ∇ and δ, which were formulated in (a) and (b) of Remark 3.11 take now a much more natural form. Indeed, • formula (3.2) reduces to h∇Y, UiB2 = hY, δ(U)iL2(S,τ ) (3.4) (3.5) for all Y ∈ D(∇) and U ∈ D(δ), • and (3.3) implies that ∇ is a derivation in the sense that a kind of Leibniz rule ∇(Y1Y2) = (∇Y1) · Y2 + Y1 · (∇Y2) holds for all Y1, Y2 ∈ D(∇), where · denotes the left and right action, respectively, of S on B2. We recall [KNPS12, Proposition 3.23], which is itself a combination of Propositions 5.3.9 and 5.3.10 in [BS98]. Proposition 3.13. The gradient operator is densely defined and closable. The domain D(∇) of the closure ∇ : L2(S, τ ) ⊇ D(∇) → B2 ∇ : L2(S, τ ) ⊇ D(∇) → B2 can be characterized by the chaos expansion in the following way D(∇) = nI S(f )(cid:12)(cid:12)(cid:12) f = (fn)∞n=0 ∈ F : ∞ Xn=0 nkfnk2 L2(Rn +) < ∞o. / ?  O   / ?  O   ✤ ✤ ✤ ✤ ✤ ✤ / / ?  O   / ?  O   ✤ ✤ ✤ ✤ ✤ ✤ / 20 T. MAI In fact, if we write Y ∈ D(∇) in the form Y = I S(f ) with f ∈ F , we have that k∇Y k2 B2 = nkfnk2 L2(Rn +). ∞ Xn=0 Moreover, the action of ∇ on its domain D(∇) is determined by ∇t(cid:16)Z f (t1, . . . , tn) dSt1 · · · dStn(cid:17) Xj=1 Z f (t1, . . . , tj−1, t, tj+1, . . . , tn) dSt1 · · · dStj−1 ⊗ dStj+1 · · · dStn = n for f ∈ L2(Rn +). Remark 3.14. We point out that Proposition 5.2.3 in [BS98] shows beyond this that ∇ is also closable as an unbounded linear operator from Lp(S, τ ) to Bp for each 1 ≤ p < ∞. The domain of its closure, which will be denoted by Dp, is given as the closure of Salg with respect to the norm k · k1,p defined by kY k1,p := (cid:0)kY kp Lp(S,τ ) + k∇Y kp Bp(cid:1) 1 p . We will use this observation only in the case p = 2, where D(∇) = D2 gives an alterna- tive description of the domain D(∇) of the closure of the gradient operator ∇, which was characterized in Proposition 3.13 in terms of the chaos decomposition. Concerning now the divergence operator, we record here [KNPS12, Proposition 3.25], which combines Propositions 5.3.9 and 5.3.11 of [BS98]. Proposition 3.15. The divergence operator is densely defined and closable. The domain D(δ) of its closure δ : B2 ⊇ D(δ) → L2(S, τ ) contains all adapted L2-biprocesses Ba 2, we have δ : B2 ⊇ D(δ) → L2(S, τ ) 2 and for each U ∈ Ba δ(U) = ZR+ Ut♯dSt. In general, the action of δ on its domain D(δ) is determined by δ(cid:16)Z ft(t1, . . . , tn; s1, . . . , sm) dSt1 · · · dStn ⊗ dSs1 · · · dSsm(cid:17) = Z ft(t1, . . . , tn; s1, . . . , sm) dSt1 · · · dStndStdSs1 · · · dSsm for any f ∈ L2(R+, L2(Rn +) ⊗ L2(Rm + )). Finally, we also take the operator N into account. This operator induces the so-called number operator N : L2(S, τ ) ⊇ D(N) → L2(S, τ ) REGULARITY OF DISTRIBUTIONS OF WIGNER INTEGRALS 21 with domain D(N) := I S(Falg) = Salg as shown in the following commutative diagram. F D( N) N F I S I S I S / L2(S, τ ) / D(N) N / L2(S, τ ) Remark 3.16. The relation N = δ ◦ ∇ on Falg, which was recorded in part (c) of Remark 3.11, translates by definition immediately to the relation N = δ ◦ ∇ on Salg. We recall now [KNPS12, Remark 3.24]. Proposition 3.17. The number operator N : L2(S, τ ) ⊇ D(N) → L2(S, τ ) is densely defined and closable. The domain D(N) of its closure can be characterized by using the chaos expansion in the following way D(N) = nI S(f )(cid:12)(cid:12)(cid:12) f = (fn)∞n=0 ∈ F : ∞ Xn=0 n2kfnk2 L2(Rn +) < ∞o. In particular, the closure of the gradient ∇ maps D(N) into D(δ), and on D(N), it holds true that N = δ ◦ ∇D(N ). 4. Non-commutative derivations The second pillar of the proof of Theorem 2.4 is the theory of non-commutative derivations as it arises from the work of Voiculescu [Voi98, Voi99] and of Dabrowski [Dab10, Dab14], and the corresponding generalization of methods originating from [MSW15]. Derivations are mainly characterized by the Leibniz rule, which is a straightforward gen- eralization of the Leibniz rule for usual derivatives. Hence, these objects can be introduced and studied in a purely algebraic setting: if A is any unital complex algebra and if M is an arbitrary A-bimodule, we call a linear mapping δ : A → M a M-valued derivation on A, if it satisfies the Leibniz rule δ(x1x2) = δ(x1) · x2 + x1 · δ(x2) for all x1, x2 ∈ A. But since we are interested more in the analytic rather than the purely algebraic properties of derivations, we will impose here some additional conditions on the algebra A and the A- bimodule M. For doing this, we clearly have a lot of flexibility. The most general notion of such analytic derivations is probably the one that is presented in [CS03, Definition 4.1]. However, the feasibility of our arguments here depends strongly on more restrictive assumptions, due to which those derivations will behave pretty much like the usual non-commutative derivatives, / ?  O   / ?  O   ✤ ✤ ✤ ✤ ✤ ✤ / 22 T. MAI as they appear for instance in [Voi98]. Accordingly, we shall call them non-commutative derivations. Throughout this section, let (M, τ ) be a tracial W ∗-probability space. Definition 4.1. A linear map δ : M ⊇ D(δ) → L2(M, τ ) ⊗ L2(M, τ ) is called a non-commutative derivation on M if the following two conditions are satisfied: in M. • The domain D(δ) of δ is a unital ∗-subalgebra of M, which is moreover weakly dense • The linear map δ satisfies the Leibniz rule (or product rule) δ(x1x2) = δ(x1) · x2 + x1 · δ(x2) for all x1, x2 ∈ D(δ), where · denotes the natural bimodule operation of M on the Hilbert space L2(M, τ ) ⊗ L2(M, τ ) ∼= L2(M ⊗ M, τ ⊗ τ ). Remark 4.2. Let Chx1, . . . , xni denote the ∗-algebra of non-commutative polynomials in formal self-adjoint variables x1, . . . , xn. If P is any monomial in the variables x1, . . . , xn, we put for any fixed i = 1, . . . , n ∂iP := XP =P1xiP2 P1 ⊗ P2 where the sum runs over all decompositions of P in the form P = P1xiP2 with some mono- mials P1, P2. In particular, we have ∂ixj = δi,j1 ⊗ 1 for i, j = 1, . . . , n. It is easy to check that ∂i extends by linearity to a Chx1, . . . , xni⊙2-valued derivation on Chx1, . . . , xni. Conversely, as a linear map ∂i : Chx1, . . . , xni → Chx1, . . . , xni ⊙ Chx1, . . . , xni, the non-commutative derivative ∂i is uniquely determined by the Leibniz rule and the prop- erty ∂ixj = δi,j1 ⊗ 1 for i, j = 1, . . . , n. In the case n = 1, we will abbreviate ∂1 simply by ∂. Assume now that δ : M ⊇ D(δ) → L2(M, τ ) ⊗ L2(M, τ ) is any non-commutative deriva- tion in the sense of Definition 4.1. If X1, . . . , Xn are self-adjoint elements in D(δ) and if P ∈ Chx1, . . . , xni is any non-commutative polynomial, then the evaluation P (X1, . . . , Xn) belongs clearly to D(δ) and we have the formula (4.1) δ(P (X1, . . . , Xn)) = n Xi=1 (∂iP )(X1, . . . , Xn)♯δ(Xi), where we abbreviate by Q(X) the natural evaluation of any Q ∈ Chx1, . . . , xni⊙2 at (X1, . . . , Xn). In other words, the non-commutative derivatives ∂1, . . . , ∂n are universal in the sense that they provide an explicit expression for the restriction of any non-commutative derivation δ to a subalgebra ChX1, . . . , Xni of its domain D(δ) in terms of its values on the generators X1, . . . , Xn. REGULARITY OF DISTRIBUTIONS OF WIGNER INTEGRALS 23 Following [Voi98, Voi99], we change now our point of view by considering any non- commutative derivation δ : M ⊇ D(δ) → L2(M, τ ) ⊗ L2(M, τ ) in the sense of Definition 4.1 as an unbounded linear operator δ : L2(M, τ ) ⊃ D(δ) → L2(M, τ ) ⊗ L2(M, τ ). Since D(δ) is clearly dense in L2(M, τ ) with respect to the L2-norm k · k2 induced by τ , we can also consider its adjoint operator δ∗ : L2(M, τ ) ⊗ L2(M, τ ) ⊇ D(δ∗) → L2(M, τ ). The theory that we are going to presented in the next subsections concerns properties of δ and its adjoint δ∗. More precisely, we will discuss the question of closability for δ and we will show that δ and δ∗, which are unbounded operators by definition, can nevertheless be controlled in appropriate norms. For most of these results, the condition 1⊗ 1 ∈ D(δ∗) turns out to be crucial. 4.1. Voiculescu's formulas for δ∗. In [Voi98], Voiculescu deduced formulas for the adjoint operator δ∗ of a non-commutative derivation δ under the assumption that 1⊗1 ∈ D(δ∗). This was shown in [Voi98] only in the case of the non-commutative derivatives that are defined on the algebra of finitely many generators, but it was noted and worked out in [Voi99] that the same arguments apply in more general situations. Although this is commonly accepted as a well-known fact, we give here for reader's convenience a complete introduction to this circle of ideas, since these beautiful results are of great importance for our considerations. For the rest of this subsection, let δ : M ⊇ D(δ) → L2(M, τ ) ⊗ L2(M, τ ) be some fixed non-commutative derivation in the sense of Definition 4.1, viewed as an unbounded linear operator δ : L2(M, τ ) ⊃ D(δ) → L2(M, τ ) ⊗ L2(M, τ ). Following Voiculescu's strategy, we begin by deducing some very useful product rules for its adjoint operator δ∗. Clearly, we may extend the involution ∗ on M from M uniquely to an involution on L2(M, τ ), and the canonical involution ∗ on M ⊗ M from M ⊗ M uniquely to an involution L2(M, τ ) ⊗ L2(M, τ ). Consequently, (4.2) hx, yi = hy∗, x∗i holds for all x, y ∈ L2(M, τ ). Lemma 4.3. Let u ∈ D(δ∗) ∩ (M ⊙ M) and x ∈ D(δ) be given. Then (4.3) δ∗(x · u) = xδ∗(u) − (τ ⊗ id)(u♯δ(x∗)∗), δ∗(u · x) = δ∗(u)x − (id⊗τ )(u♯δ(x∗)∗), where ♯ is defined according to Remark 3.4 with respect to the M-M-bimodule L2(M, τ ) ⊗ L2(M, τ ). In particular, for any u ∈ D(δ∗) ∩ (M ⊙ M), we have {x1 · u · x2 x1, x2 ∈ D(δ)} ⊆ D(δ∗). from which x · u ∈ D(δ∗) and the first formula in (4.3) follows. Analogously, we obtain by hδ(y), x · ui = hx∗ · δ(y), ui = hδ(x∗y), ui − hδ(x∗) · y, ui = hx∗y, δ∗(u)i − h1 ⊗ y, u♯δ(x∗)∗i = hy, xδ∗(u)i − hy, (τ ⊗ id)(u♯δ(x∗)∗)i = hy, xδ∗(u) − (τ ⊗ id)(u♯δ(x∗)∗)i, hδ(y), u · xi = hδ(y) · x∗, ui = hδ(yx∗), ui − hy · δ(x∗), ui = hyx∗, δ∗(u)i − hy ⊗ 1, u♯δ(x∗)∗i = hy, δ∗(u)xi − hy, (id⊗τ )(u♯δ(x∗)∗)i = hy, δ∗(u)x − (id⊗τ )(u♯δ(x∗)∗)i 24 T. MAI Proof. Let u ∈ D(δ∗)∩ (M ⊙ M) and x ∈ D(δ) be given. For any y ∈ D(δ), we observe that that u · x ∈ D(δ∗) and the second formula in (4.3). A combination of both observations immediately yields the stated inclusion for any u ∈ D(δ∗) ∩ (M ⊙ M). {x1 · u · x2 x1, x2 ∈ D(δ)} ⊆ D(δ∗) (cid:3) In the case 1 ⊗ 1 ∈ D(δ∗), Lemma 4.3 yields an explicit formula for δ∗ on D(δ) ⊙ D(δ) in terms of δ∗(1 ⊗ 1) and δ. It takes its nicest form if we require an additional property of δ. In fact, we will assume a certain compatibility between the involution ∗ on M and the involution † on M ⊗ M, where the latter is determined by (x1 ⊗ x2)† := x∗2 ⊗ x∗1. Note that † differs from the canonical involution ∗ on M ⊗ M only by the flip mapping σ : M ⊗ M → M ⊗ M, i.e., we have u† = σ(u∗). Clearly, we may extend the involution † from M ⊗ M uniquely to an involution L2(M, τ )⊗ L2(M, τ ). Accordingly, for all u, v ∈ L2(M, τ ) ⊗ L2(M, τ ), it holds true that (4.4) hu, vi = hv†, u†i. Definition 4.4. A non-commutative derivation on (M, τ ) is called real, if it satisfies δ : M ⊇ D(δ) → L2(M, τ ) ⊗ L2(M, τ ) (4.5) δ(x)† = δ(x∗) for all x ∈ D(δ). Often, condition (4.5) can be weakened. We record this here as a remark. Remark 4.5. We point out that condition (4.5) is automatically satisfied if the unital ∗- algebra D(δ) is generated by self-adjoint elements xi, i ∈ I, for some index set I 6= ∅, such that δ(xi)† = δ(xi) holds for all i ∈ I. Indeed, if we define δ with D(δ) := D(δ) by δ : D(δ) → L2(M, τ ) ⊗ L2(M, τ ), x 7→ δ(x∗)†, REGULARITY OF DISTRIBUTIONS OF WIGNER INTEGRALS 25 we can easily check that δ is a non-commutative derivation as well. Thus, the set D := {x ∈ D(δ) δ(x) = δ(x)} is closed under multiplication, i.e. x1, x2 ∈ D implies x1x2 ∈ D. Since it contains the generators {xi i ∈ I} by assumption, we must have that D = D(δ), from which it follows by construction that δ(x)† = δ(x∗) holds for all x ∈ D(δ). The following lemma collects some useful formulas for real non-commutative derivations. Lemma 4.6. Let δ : M ⊇ D(δ) → L2(M, τ )⊗L2(M, τ ) be a real non-commutative derivation on (M, τ ), Then, for all x ∈ D(δ), it holds true that (id⊗τ )(δ(x))∗ = (τ ⊗ id)(δ(x∗)), (τ ⊗ id)(δ(x))∗ = (id⊗τ )(δ(x∗)). Furthermore, for any u ∈ D(δ∗), we have also u† ∈ D(δ∗) and it holds true that δ∗(u†) = δ∗(u)∗. In particular, if 1 ⊗ 1 ∈ D(δ∗), we have δ∗(1 ⊗ 1) = δ∗(1 ⊗ 1)∗. Proof. The first statement is an immediate consequence of the defining property of real derivations, since in general (4.6) (id⊗τ )(u)∗ = (τ ⊗ id)(u†), (τ ⊗ id)(u)∗ = (id⊗τ )(u†) holds for each u ∈ L2(M, τ ) ⊗ L2(M, τ ). For seeing the second statement, we take any y ∈ D(δ) and we observe by using (4.4) that hu†, δ(y)i = hδ(y)†, ui = hδ(y∗), ui = hy∗, δ∗(u)i = hδ∗(u)∗, yi. This yields u† ∈ D(δ∗) with δ∗(u†) = δ∗(u)∗, as desired. Now, we can combine formulas (4.3) of Lemma 4.3. (cid:3) Lemma 4.7. If the condition 1 ⊗ 1 ∈ D(δ∗) is satisfied, then D(δ) ⊙ D(δ) ⊆ D(δ∗). If δ is a real derivation in the sense of Definition 4.4, then we have more explicitly for all u ∈ D(δ) ⊙ D(δ) that (4.7) where, in general, we denote by mη for any η ∈ L2(M, τ ) the linear mapping mη : M ⊙ M → L2(M, τ ) that is determined by mη(v) = v♯η, so that m1 is nothing else than the multiplication map m1(x1 ⊗ x2) = x1x2. δ∗(u) = u♯δ∗(1 ⊗ 1) − m1(id⊗τ ⊗ id)(δ ⊗ id + id⊗δ)(u), The formula (4.7) given in Lemma 4.7 immediately implies that in particular (4.8) δ∗(x ⊗ 1) = xδ∗(1 ⊗ 1) − (id⊗τ )(δ(x)), δ∗(1 ⊗ x) = δ∗(1 ⊗ 1)x − (τ ⊗ id)(δ(x)), which we record here for later reference. 26 T. MAI Proof of Lemma 4.7. The first assertion, namely that D(δ) ⊙ D(δ) ⊆ D(δ∗) holds under the condition 1 ⊗ 1 ∈ D(δ∗), is an immediate consequence of Lemma 4.3. Note that we did not use for this conclusion the assumption that δ is real. For seeing (4.7), we proceed as follows. First of all, we note that the validity of (4.5) guarantees according to Lemma 4.6 that (id⊗τ )(δ(x∗)∗) = (τ ⊗ id)(δ(x)), (τ ⊗ id)(δ(x∗)∗) = (id⊗τ )(δ(x)) for each x ∈ D(δ). Next, for any u = x1 ⊗ x2 with x1, x2 ∈ D(δ), we check by using consecutively both formulas of (4.3) and Lemma 4.6 that δ∗(u) = δ∗(x1 · (1 ⊗ x2)) = x1δ∗((1 ⊗ 1) · x2) − (τ ⊗ id)((1 ⊗ x2)♯δ(x∗1)∗) = x1δ∗(1 ⊗ 1)x2 − x1(id⊗τ )(δ(x∗2)∗) − (τ ⊗ id)((1 ⊗ x2)♯δ(x∗1)∗) = u♯δ∗(1 ⊗ 1) − x1(τ ⊗ id)(δ(x2)) − (id⊗τ )(δ(x1))x2 = u♯δ∗(1 ⊗ 1) − m1(id⊗τ ⊗ id)(δ ⊗ id + id⊗δ)(u). By linearity, this shows (4.7) for all u ∈ D(δ) ⊙ D(δ). This concludes the proof. 4.2. Dabrowski's inequalities. Based on Voiculescu's formulas, Dabrowski deduced in [Dab10] a collection of interesting inequalities concerning the boundedness of the non- commutative derivatives, which are very surprising from a classical point of view. In [Dab14], he noted that the same arguments also apply in a more general setting. More precisely, he observed (without carrying out the proof) that his result remain valid for any real derivation, which satisfies in addition the so-called coassociativity relation. Definition 4.8. Let δ : M ⊇ D(δ) → L2(M, τ )⊗ L2(M, τ ) be a non-commutative derivation on (M, τ ). We say that δ satisfies the coassociativity relation, (cid:3) • if δ takes its values in D(δ) ⊙ D(δ), • and if δ has the property that (δ ⊗ id) ◦ δ = (id⊗δ) ◦ δ. (4.9) For reader's convenience, we state here those of Dabrowski's formulas, which we need for our purposes. Since it is instructive, we also include a slightly simplified proof thereof. Theorem 4.9. Let δ : M ⊇ D(δ) → L2(M, τ ) ⊗ L2(M, τ ) be a non-commutative derivation on a tracial W ∗-probability space (M, τ ), which • is real in the sense of Definition 4.4 • and satisfies the coassociativity relation as formulated in Definition 4.8. If the condition 1 ⊗ 1 ∈ D(δ∗) is satisfied, we have for all x ∈ D(δ) that (4.10) kδ∗(x ⊗ 1)k2 ≤ kδ∗(1 ⊗ 1)k2kxk kδ∗(1 ⊗ x)k2 ≤ kδ∗(1 ⊗ 1)k2kxk and (4.11) k(id⊗τ )(δ(x))k2 ≤ 2kδ∗(1 ⊗ 1)k2kxk k(τ ⊗ id)(δ(x))k2 ≤ 2kδ∗(1 ⊗ 1)k2kxk REGULARITY OF DISTRIBUTIONS OF WIGNER INTEGRALS 27 Before proceeding with to the proof of Theorem 4.9, we record here the following formula for its later use therein. Lemma 4.10. In the situation of Theorem 4.9, let x ∈ D(δ) be given and put Then y ∈ D(δ) holds and we have that (id⊗τ )(δ(y)) = (id⊗h·, δ∗(1 ⊗ 1)i)(δ(x)). y := (id⊗τ )(δ(x)). Proof. Since δ is assumed to satisfy the coassociativity relation, we know by Definition 4.8 that in particular D(δ) ⊙ D(δ) holds, which gives y ∈ D(δ). Furthermore, according to the coassociativity relation formulated in (4.9), we see that holds. Since we have on D(δ) the identity (τ ⊗ τ ) ◦ δ = h·, δ∗(1 ⊗ 1)i, we get δ(y) = (id⊗ id⊗τ )(cid:0)(δ ⊗ id)(δ(x))(cid:1) = (id⊗ id⊗τ )(cid:0)(id⊗δ)(δ(x))(cid:1) (id⊗τ )(δ(y)) = (id⊗τ ⊗ τ )(cid:0)(id⊗δ)(δ(x))(cid:1) = (cid:0) id⊗((τ ⊗ τ ) ◦ δ)(cid:1)(δ(x)) = (cid:0) id⊗h·, δ∗(1 ⊗ 1)i(cid:1)(δ(x)), which is the desired formula. (cid:3) Additionally, the proof of Theorem 4.9 will be based on the following observation. Lemma 4.11. Let (M, τ ) be a W ∗-probability space and let T : D(T ) → M be a linear operator on a unital ∗-subalgebra D(T ) of M. Assume that the following conditions are satisfied: (i) There exists a constant C > 0 such that 2 ≤ CkT (x∗x)k2 kT (x)k2 (ii) For each x ∈ D(T ), we have that lim sup 1 m for all x ∈ D(T ). m→∞ kT (xm)k Then T satisfies kT (x)k2 ≤ Ckxk for all x ∈ D(T ). Proof. Let x ∈ D(T ) be given. For each n ∈ N0, we define zn := (x∗x)2n ∈ D(T ). By assumption (i), we see that 2 ≤ kxk. which yields inductively kT (zn)k2 2 ≤ CkT (zn+1)k2 for all n ∈ N0, Since lim sup n→∞ kT (zn)k due to (ii), it follows that kT (z0)k2 ≤ C 1 2 +···+ 1 1 2n 2 = lim sup 1 2n 2 2n kT (zn)k n→∞ kT ((x∗x)2n kT (z0)k2 ≤ Ckxk2. for all n ∈ N0. 1 2n 2 ≤ kx∗xk = kxk2 )k 28 T. MAI By using (ii) once again, we obtain kT (x)k2 and hence kT (x)k2 ≤ Ckxk, as stated. Proof of Theorem 4.9. First of all, we note that it suffices to prove (4.10), since (4.11) follows from (4.10) and Voiculescu's formula (4.8) by an application of the triangle inequality. 2 ≤ CkT (z0)k2 ≤ C 2kxk2 (cid:3) For proving (4.10), we want to use Lemma 4.11. We consider the linear mapping T : D(T ) → M on D(T ) := D(δ) given by T (x) := δ∗(x ⊗ 1) for all x ∈ D(δ). Since Lemma 4.7 guarantees D(δ) ⊙ D(δ) ⊆ D(δ∗), the mapping T is indeed well-defined. Now, we just have to follow the receipt given in Lemma 4.11. (i) For any given x ∈ D(δ), we have to compare kT (x∗x)k2 and kT (x)k2. In fact, we will show that (4.12) from which kT (x)k2 2 = hT (x∗x), δ∗(1 ⊗ 1)i kT (x)k2 2 ≤ kδ∗(1 ⊗ 1)k2kT (x∗x)k2 immediately follows by an application of the Cauchy-Schwarz inequality. Formula (4.12) can be shown as follows. Let x ∈ D(δ) be given and put y := (id⊗τ )(δ(x)). Since y∗ = (id⊗τ )(δ(x))∗ = (τ ⊗ id)(δ(x∗)) according to Lemma 4.6, we may observe by using in turn Lemma 4.10 and Lemma 4.6 in the version (4.8) that kyk2 2 = hy, (id⊗τ )(δ(x))i = hy ⊗ 1, δ(x)i = hδ∗(y ⊗ 1), xi = hyδ∗(1 ⊗ 1), xi − h(id⊗τ )(δ(y)), xi = hδ∗(1 ⊗ 1)x∗, y∗i − h(id⊗h·, δ∗(1 ⊗ 1)i)(δ(x)), xi = h1 ⊗ δ∗(1 ⊗ 1)x∗, δ(x∗)i − hδ(x), x ⊗ δ∗(1 ⊗ 1)i. Because moreover hδ(x),x ⊗ δ∗(1 ⊗ 1)i = hx∗ · δ(x), 1 ⊗ δ∗(1 ⊗ 1)i = hδ(x∗x), 1 ⊗ δ∗(1 ⊗ 1)i − hδ(x∗) · x, 1 ⊗ δ∗(1 ⊗ 1)i = hδ(x∗x), 1 ⊗ δ∗(1 ⊗ 1)i − hδ(x∗), 1 ⊗ δ∗(1 ⊗ 1)x∗i, we may conclude kyk2 2 = 2ℜ(cid:0)h1 ⊗ δ∗(1 ⊗ 1)x∗, δ(x∗)i(cid:1) − hδ(x∗x), 1 ⊗ δ∗(1 ⊗ 1)i. REGULARITY OF DISTRIBUTIONS OF WIGNER INTEGRALS 29 Furthermore, since T (x) = xδ∗(1 ⊗ 1) − y due to (4.8), we get that kT (x)k2 2 = hxδ∗(1 ⊗ 1) − y, xδ∗(1 ⊗ 1) − yi = kxδ∗(1 ⊗ 1)k2 = kxδ∗(1 ⊗ 1)k2 2 + kyk2 2 + kyk2 2 − 2ℜ(cid:0)hxδ∗(1 ⊗ 1), yi(cid:1) 2 − 2ℜ(cid:0)hxδ∗(1 ⊗ 1) ⊗ 1, δ(x)i(cid:1). We check now hxδ∗(1 ⊗ 1) ⊗ 1, δ(x)i = hδ∗(1 ⊗ 1), (id⊗τ )(x∗ · δ(x))i = h(id⊗τ )(x∗ · δ(x))∗, δ∗(1 ⊗ 1)i = h(id⊗τ )(δ(x))∗x, δ∗(1 ⊗ 1)i = h(τ ⊗ id)(δ(x∗))x, δ∗(1 ⊗ 1)i = h(τ ⊗ id)(δ(x∗)), δ∗(1 ⊗ 1)x∗i = hδ(x∗), 1 ⊗ δ∗(1 ⊗ 1)x∗i, (by (4.2)) (by Lemma 4.6) so that A combination of our previous computations leads us to ℜ(cid:0)hxδ∗(1 ⊗ 1) ⊗ 1, δ(x)i(cid:1) = ℜ(cid:0)h1 ⊗ δ∗(1 ⊗ 1)x∗, δ(x∗)i(cid:1). (4.13) kT (x)k2 2 = kxδ∗(1 ⊗ 1)k2 2 − hδ(x∗x), 1 ⊗ δ∗(1 ⊗ 1)i Furthermore, due to (4.8), we have T (x∗x) = x∗xδ∗(1 ⊗ 1) − (id⊗τ )(δ(x∗x)), and hence (4.14) hT (x∗x), δ∗(1 ⊗ 1)i = kxδ∗(1 ⊗ 1)k2 2 − hδ(x∗x), δ∗(1 ⊗ 1) ⊗ 1i. Since (4.13) implies that hδ(x∗x), 1⊗ δ∗(1⊗1)i must be real, we get by using (4.4), Lemma 4.6, and (4.5) that hδ(x∗x), δ∗(1 ⊗ 1) ⊗ 1i = h1 ⊗ δ∗(1 ⊗ 1), δ(x∗x)i = hδ(x∗x), 1 ⊗ δ∗(1 ⊗ 1)i. Thus, comparing (4.13) and (4.14) gives kT (x)k2 2 = hT (x∗x), δ∗(1 ⊗ 1)i, which is the stated formula (4.12). (ii) To begin with, we observe that for any polynomial P and any x ∈ D(δ) kT (P (x))k2 ≤ kP (x)kkδ∗(1 ⊗ 1)k2 + k(∂P )(x)kπkδ(x)k2, (4.15) where k · kπ denotes the projective norm on D(δ) ⊙ D(δ), which is given by kukπ := inf n N Xj=1 kajkkbjk(cid:12)(cid:12)(cid:12) for any u ∈ D(δ) ⊙ D(δ). N ∈ N, a1, . . . , aN , b1, . . . , bN ∈ D(δ) : u = N Xj=1 aj ⊗ bjo 30 T. MAI Indeed, according to (4.8), we have for each polynomial P and x ∈ D(δ) T (P (x)) = δ∗(P (x) ⊗ 1) = P (x)δ∗(1 ⊗ 1) − (id⊗τ )(δ(P (x))) = P (x)δ∗(1 ⊗ 1) − (id⊗τ )((∂P )(x)♯δ(x)), where we used that δ(P (x)) = (∂P )(x)♯δ(x) according to formula (4.1), which was given in Remark 4.2. This yields as desired kT (P (x))k2 ≤ kP (x)δ∗(1 ⊗ 1)k2 + k(id⊗τ )((∂P )(x)♯δ(x))k2 ≤ kP (x)δ∗(1 ⊗ 1)k2 + k(∂P )(x)♯δ(x)k2 ≤ kP (x)kkδ∗(1 ⊗ 1)k2 + k(∂P )(x)kπkδ(x)k2. If we apply (4.15) to the polynomial P (x) = xm for any m ∈ N, we may deduce that kT (xm)k2 ≤ kxkmkδ∗(1 ⊗ 1)k2 + mkxkm−1kδ(x)k2 since k(∂P )(x)kπ ≤ mkxkm−1 holds. From this, we immediately get that lim sup m→∞ kT (xm)k 1 m 2 ≤ kxk. Thus, condition (ii) of Lemma 4.11 is satisfied. Lemma 4.11 tells us now that kT (x)k2 ≤ kδ∗(1 ⊗ 1)k2kxk, which is by definition of T exactly the first inequality in (4.10). The second one can simply be deduced from the first one by using that δ∗(u†) = δ∗(u)∗ holds for any u ∈ D(δ∗) according to Lemma 4.6, since δ was assumed to be real. (cid:3) Combining Theorem 4.9 with Lemma 4.7 yields the following corollary. Corollary 4.12. Let δ : M ⊇ D(δ) → L2(M, τ ) ⊗ L2(M, τ ) a non-commutative derivation on a tracial W ∗-probability space (M, τ ). We assume that δ is a real derivation in the sense of 4.4 and that is satisfies the coassociativity relation formulated in 4.8. Then, for all x1, x2 ∈ D(δ), it holds true that (4.16) kδ∗(x1 ⊗ x2)k2 ≤ 3kδ∗(1 ⊗ 1)k2kx1kkx2k and (4.17) k(id⊗τ )(δ(x1) · x2)k2 ≤ 4kδ∗(1 ⊗ 1)k2kx1kkx2k, k(τ ⊗ id)(x1 · δ(x2))k2 ≤ 4kδ∗(1 ⊗ 1)k2kx1kkx2k. Proof. According to Lemma 4.7, we have for all x1, x2 ∈ D(δ) that δ∗(x1 ⊗ x2) = x1δ∗(1 ⊗ 1)x2 − m1(id⊗τ ⊗ id)(δ ⊗ id + id⊗δ)(x1 ⊗ x2) = x1δ∗(1 ⊗ 1)x2 − (id⊗τ )(δ(x1))x2 − x1(τ ⊗ id)(δ(x2)) = δ∗(x1 ⊗ 1)x2 − x1(τ ⊗ id)(δ(x2)) and thus, by applying the estimates (4.11) and (4.10), that kδ∗(x1 ⊗ x2)k2 ≤ kδ∗(x1 ⊗ 1)k2kx2k + kx1kk(τ ⊗ id)(δ(x2))k2 ≤ 3kδ∗(1 ⊗ 1)k2kx1kkx2k. REGULARITY OF DISTRIBUTIONS OF WIGNER INTEGRALS 31 This shows the validity of (4.16). For proving (4.17), we first use integration by parts in order to obtain (id⊗τ )(δ(x1) · x2) = (id⊗τ )(δ(x1x2)) − (id⊗τ )(x1 · δ(x2)) = (id⊗τ )(δ(x1x2)) − x1(id⊗τ )(δ(x2)) for arbitrary x1, x2 ∈ D(δ). From this, we can easily deduce by using (4.11) that k(id⊗τ )(δ(x1) · x2)k2 ≤ k(id⊗τ )(δ(x1x2))k2 + kx1kk(id⊗τ )(δ(x2))k2 ≤ 4kδ∗(1 ⊗ 1)k2kx1kkx2k which is the first inequality of (4.17). The second inequality can either be proven similarly or can be deduced from the first one by using that δ is real. (cid:3) We conclude this subsection by highlighting Formula (4.12), which was obtained in the proof of Theorem 4.9. Since we think that this observation might be of independent interest and could be helpful for future investigations, we record (4.12) here by the following corollary. Corollary 4.13. Let δ : M ⊇ D(δ) → L2(M, τ ) ⊗ L2(M, τ ) be a non-commutative deriva- tion on a tracial W ∗-probability space (M, τ ), which is real and satisfies the coassociativity relation. Assume additionally that 1 ⊗ 1 ∈ D(δ∗). Then, for each x ∈ D(δ), it holds true that Assume, for instance, that in the situation of Corollary 4.13 the conditions kδ∗(x ⊗ 1)k2 2 = hδ∗((x∗x) ⊗ 1), δ∗(1 ⊗ 1)i. δ∗(1 ⊗ 1) ∈ D(δ) ∩ M and δ(δ∗(1 ⊗ 1)) ∈ M ⊗ M are satisfied in addition. Corollary 4.13 allows us then to conclude that for any x ∈ D(δ) 2 = hδ∗((x∗x)⊗ 1), δ∗(1 ⊗ 1)i = h(x∗x) ⊗ 1, δ(δ∗(1⊗ 1))i = hx ⊗ 1, x· δ(δ∗(1⊗ 1))i kδ∗(x⊗ 1)k2 and hence kδ∗(x ⊗ 1)k2 ≤ kδ(δ∗(1 ⊗ 1))k1/2kxk2 holds. Like in Theorem 4.9, we can use this in combination with (4.8) in order to deduce that k(id⊗τ )(δ(x))k2 ≤ (cid:0)kδ∗(1 ⊗ 1)k + kδ(δ∗(1 ⊗ 1))k1/2(cid:1)kxk2 holds for each x ∈ D(δ). Analogous inequalities can of course be proven for δ∗(1 ⊗ x) and (τ ⊗ id)(δ(x)). In other words, we can strengthen the bounds that were obtained in Theorem 4.9 by imposing some stronger "regularity conditions" on δ∗(1 ⊗ 1). Note that this in fact slightly improves similar estimates that were deduced in [Dab14]. 4.3. Survival of zero divisors. We are mainly interested here in applications of the theory of non-commutative derivations to regularity questions for certain distributions. The basic idea that originates in [MSW14, MSW15] is that, in order to exclude atoms, one should reformulate this question in more algebraic terms as a question about the existence of zero- divisors, where the latter can be excluded by a successive reduction of the degree by applying non-commutative derivations. The key for this purpose is a certain inequality which allows the conclusion that zero- divisors xu = 0 survive under applying operators of the form ∆p(x) := (τ ⊗ id)(p · δ(x)) 32 T. MAI for any non-commutative derivation δ satisfying certain conditions and some non-trivial projection p. This inequality will be given below in Proposition 4.14. As we will see, it will more generally relate products xu and x∗v for elements x in the domain of the given non- commutative derivation δ and arbitrary elements u, v in the corresponding von Neumann algebra with an expression of the form v∗ · δ(x) · u. We point out that although the inequality itself holds in a considerably large generality, the feasibility of the whole strategy for excluding zero-divisors relies heavily on the structure of the given non-commutative derivation. Roughly speaking, applying δ has to "reduce the degree" of the given element x. More formally, one should think of a grading on the space of distributions under consideration that is compatible with δ. We do not want to give a definition in full generality, however we want to mention that the grading that was used in [MSW14, MSW15] was given by the monomials of fixed degree. As we will see in Section 5, where we present the proof of Theorem 2.4, there is a closely related grading on the space of finite Wigner integrals. The crucial inequality will now be formulated in the following proposition. Proposition 4.14. Let δ : L2(M, τ ) ⊇ D(δ) → L2(M, τ ) ⊗ L2(M, τ ) be a non-commutative derivation. We assume that δ is real and satisfies the coassociativity relation. Then, if in addition 1 ⊗ 1 ∈ D(δ∗) holds, we have for all x ∈ D(δ), where δ denotes the closure of δ, and u, v ∈ M the inequality (4.18) for all y1, y2 ∈ D(δ). also v∗ · δ(x) · u = 0 holds. hv∗ · δ(x) · u, y1 ⊗ y2i ≤ 4kδ∗(1 ⊗ 1)k2(cid:0)kvkkxuk2 + kukkx∗vk2(cid:1)ky1kky2k In particular, if we have both xu = 0 and x∗v = 0 for x ∈ D(δ) with some u, v ∈ M, then Before giving the proof of Proposition 4.14, we first mention an easy but useful application of Kaplansky's density theorem. Lemma 4.15. In the given setting of a tracial W ∗-probability space (M, τ ), let D be a ∗- subalgebra of M, which is weakly dense in M. Then, for each w ∈ M, there exists a sequence (wk)k∈N of elements in D such that k∈N kwkk ≤ kwk, (i) sup (ii) kwk − wk2 → 0 as k → ∞. If w = w∗, then we may assume in addition that wk = w∗k for all k ∈ N. Proof. First of all, we note that for proving the existence of a sequence (wk)k∈N of elements in D, which satisfies conditions (i) and (ii), it suffices to find a net (wλ)λ∈Λ of elements in D, which satisfies (i)' sup λ∈Λ kwλk ≤ kwk, λ∈Λ−→ 0. (ii)' kwλ − wk2 Indeed, given such a net (wλ)λ∈Λ, we may choose a sequence (λk)k∈N in Λ, such that kwλk − wk2 < 1 k holds for all k ∈ N. Hence, the sequence (wλk)k∈N satisfies (i) and (ii), as desired. Now, for finding a net of elements in D, which satisfies (i)' and (ii)', we apply Kaplansky's density theorem. Indeed, this theorem guarantees the existence of a net (wλ)λ∈Λ of elements in D, such that kwλk ≤ kwk holds for all λ ∈ Λ, and which converges to w in the strong REGULARITY OF DISTRIBUTIONS OF WIGNER INTEGRALS 33 operator topology. Thus, the net (wλ)λ∈Λ already satisfies condition (i)' and it remains to show the validity of (ii)'. For seeing (ii)', we note that with respect to the weak operator topology, w∗λw λ∈Λ−→ w∗w, w∗wλ λ∈Λ−→ w∗w, and w∗λwλ λ∈Λ−→ w∗w, such that according to the continuity of τ kwλ − wk2 2 = τ ((wλ − w)∗(wλ − w)) = τ (w∗λwλ) − τ (w∗λw) − τ (w∗wλ) + τ (w∗w) λ∈Λ−→ 0, as claimed in (ii)'. This concludes the proof of the first part of the lemma. For proving the additional statement, we just have to observe that in the case w = w∗, we can take any sequence (wk)k∈N that satisfies (i) and (ii), and replace each wk by its real part ℜ(wk) = 1 2(wk + w∗k). Indeed, for the sequence (wk)k∈N obtained in this way, conditions (i) and (ii) are still valid, but we have achieved wk = w∗k for all k ∈ N in addition. (cid:3) Now, we may proceed by Proof of Proposition 4.14. Firstly, we assume that x ∈ D(δ) as well as u, v ∈ D(δ). In this particular case, we may compute hxu, δ∗(vy1 ⊗ y2)i = hδ(xu), vy1 ⊗ y2i = hδ(x) · u, vy1 ⊗ y2i + hx · δ(u), vy1 ⊗ y2i = hv∗ · δ(x) · u, y1 ⊗ y2i + hδ(u) · y∗2, x∗vy1 ⊗ 1i = hv∗ · δ(x) · u, y1 ⊗ y2i + h(id⊗τ )(δ(u) · y∗2), x∗vy1i. Rearranging the terms yields hv∗ · δ(x) · u, y1 ⊗ y2i = hxu, δ∗(vy1 ⊗ y2)i − h(id⊗τ )(δ(u) · y∗2), x∗vy1i, from which we deduce by the inequalities in Corollary 4.12 that hv∗ · δ(x) · u, y1 ⊗ y2i ≤ hxu, δ∗(vy1 ⊗ y2)i + h(id⊗τ )(δ(u) · y∗2), x∗vy1i ≤ kxuk2kδ∗(vy1 ⊗ y2)k2 + k(id⊗τ )(δ(u) · y∗2)k2kx∗vy1k2 ≤ 4kδ∗(1 ⊗ 1)k2(cid:0)kvkkxuk2 + kukkx∗vk2(cid:1)ky1kky2k, as desired. Due to Lemma 4.15, this inequality extends to arbitrary u, v ∈ M. extend it from x ∈ D(δ) to x ∈ D(δ). Thus, we have proven (4.18) for x ∈ D(δ) and u, v ∈ M. It remains to show that we may Since D(δ) turns out to be the closure of D(δ) with respect to the norm k·k2,1 defined by kxk2,1 := (cid:0)kxk2 2 + kδ(x)k2 2(cid:1) 1 2 for any x ∈ D(δ), we can find for any x ∈ D(δ) a sequence (xk)k∈N in D(δ) such that both conditions kxk − xk2 → 0 and kδ(xk) − δ(x)k2 → 0 as k → ∞ are satisfied. Hence, for given u, v ∈ M, we observe lim k→∞hv∗ · δ(xk) · u, y1 ⊗ y2i = hv∗ · δ(xk) · u, y1 ⊗ y2i 34 and T. MAI lim k→∞(cid:0)kvkkxkuk2 + kukkx∗kvk2(cid:1) = kvkkxuk2 + kukkx∗vk2, from which (4.18) immediately follows in full generality. Finally, if we have xu = 0 and x∗v = 0, then (4.18) implies that hv∗ · δ(x) · u, y1 ⊗ y2i = 0 for all y1, y2 ∈ D(δ) and hence by linearity hv∗ · δ(x) · u, wi = 0 for all w ∈ D(δ) ⊙ D(δ). Since D(δ) ⊙ D(δ) is dense in L2(S, τ ) ⊗ L2(S, τ ), we obtain v∗ · δ(x) · u = 0, as stated. (cid:3) 5. Proof of Theorem 2.4 We are prepared now to build the proof of Theorem 2.4 on its pillars raised in the previous sections. In the light of free Malliavin calculus, it seems natural that methods from Section 4 could be used for a proof of Theorem 2.4 based on the same reduction method as in [MSW14, MSW15]. Nevertheless, there is the fundamental obstacle that in the world of free stochastic calculus, the role of non-commutative derivatives which were used in the "discrete setting" of [MSW14, MSW15], is taken over by the Malliavin operators as their "continuous counterparts". These operators are seemingly of completely different nature. But on closer inspection, it turns out that the right object for this purpose, which bridges -- somehow as an architrave, if one wants to strain the architecture language again -- be- tween free stochastic calculus and the theory of non-commutative derivatives are directional gradients. We will introduce this concept in the following subsection. 5.1. Directional gradients. Roughly speaking, directional gradients are obtained from the gradient operator by integrating out the (for us obstructive) time dependence against any function in L2(R+). More formally, we shall introduce these objects as follows. Definition 5.1. For each h ∈ L2(R+), we define an unbounded linear operator ∇h : L2(S, τ ) ⊇ D(∇h) → L2(S, τ ) ⊗ L2(S, τ ) with domain D(∇h) := D(∇) = Salg by ∇hY := h∇Y, hi = ZR+ ∇tY h(t) dt, where we refer to the pairing h·,·i that was introduced in Definition 3.3. We call ∇h the directional gradient (in the direction h). This terminology goes in fact parallel to classical Malliavin calculus, where corresponding expressions are also interpreted as directional derivatives. We collect some basic but very important properties of directional gradients in the follow- ing lemma. Lemma 5.2. Let h ∈ L2(R+) be given. REGULARITY OF DISTRIBUTIONS OF WIGNER INTEGRALS 35 (a) If · denotes the left and right action of S on L2(S, τ ) ⊗ L2(S, τ ), respectively, then the Leibniz rule ∇h(Y1Y2) = (∇hY1) · Y2 + Y1 · (∇hY2) holds for all Y1, Y2 ∈ D(∇h) = Salg. (b) For all Y ∈ D(∇h), it holds true that Thus, if h ∈ L2(R+, R), we have in particular that ∇h(Y ∗) = (∇hY )†. ∇h(Y ∗) = (∇hY )† holds for all Y ∈ D(∇h). (c) The directional gradient ∇h takes its values in Salg ⊙ Salg and we have that (∇h ⊗ id)∇h = (id⊗∇h)∇h. More generally, it holds true for all h1, h2 ∈ L2(R+) that (∇h1 ⊗ id)∇h2 = (id⊗∇h2)∇h1. Proof. The fact that ∇h satisfies the Leibniz rule stated in (a) follows immediately from the Leibniz rule (3.5) for ∇ on D(∇), since the domains D(∇) and D(∇h) agree. For seeing (b), we consider Y = X(h1) . . . X(hn) ∈ Salg for h1, . . . , hn ∈ L2(R+, R). A straightforward calculation confirms that n ∇h(Y ∗) = n Xj=1 = (cid:16) Xj=1 = (∇hY )†. hhj, hiX(hn)· · · X(hj+1) ⊗ X(hj−1)· · · X(h1) hhj, hiX(h1)· · · X(hj−1) ⊗ X(hj+1)· · · X(hn)(cid:17)† Because h = h holds for any h ∈ L2(R+, R), the additional statement in (b) is an immediate consequence of the formula ∇h(Y ∗) = (∇hY )†. Alternatively, by referring to Remark 4.5, it suffices to check ∇h(Y ∗) = (∇hY )† on the algebraic generators (X(g))g∈L2(R+,R) of Salg. But in this case, the statement is obvious since X(g) is self-adjoint and since we have ∇hX(g) = hg, hiL2(R+) 1 ⊗ 1 for any g ∈ L2(R+, R). stated formula. For doing this, it suffices by linearity to prove (∇h1 ⊗ id)∇h2Y = (id⊗∇h2)∇h1Y for all h1, h2 ∈ L2(R+) and any element Y ∈ Salg of the form Y = X(g1)X(g2)· · · X(gn). For proving (c), since ∇h clearly takes its values in Salg ⊙Salg, it only remains to show the If 1 ≤ j1 < j2 ≤ n are given, we will abbreviate in the following Xj1,j2 := X(g1)· · · X(gj1−1) ⊗ X(gj1+1)· · · X(gj2−1) ⊗ X(gj2+1)· · · X(gn), where as usually empty products are understood as 1. Firstly, we compute hgj2, h2iX(g1)· · · X(gj2−1) ⊗ X(gj2+1)· · · X(gn), ∇h2Y = X1≤j2≤n 36 which yields T. MAI (∇h1 ⊗ id)∇h2Y = X1≤j1<j2≤n Similarly, we compute hgj1, h1ihgj2, h2i Xj1,j2 ∇h1Y = X1≤j1≤n hgj1, h1iX(g1)· · · X(gj1−1) ⊗ X(gj1+1)· · · X(gn), which yields (id⊗∇h2)∇h1Y = X1≤j1<j2≤n hgj1, h1ihgj2, h2i Xj1,j2 Because the right hand sides of both results agree, we finally obtain the desired equality. This concludes the proof. (cid:3) Combining the properties of directional gradients that we have established in the previous Lemma 5.2 leads us immediately to the following crucial observation. Corollary 5.3. For any h ∈ L2(R+), the directional gradient ∇h : L2(S, τ ) ⊇ D(∇h) → L2(S, τ ) ⊗ L2(S, τ ), induces a non-commutative derivation on S in the sense of Definition 4.1, which satisfies additionally the coassociativity relation that was formulated in Definition 4.8. If we choose particularly any h ∈ L2(R+, R), then ∇h is also a real derivation in the sense of Definition 4.4. The importance of this observations is perfectly clear now, since it puts directional gra- dients in the setting non-commutative derivations and gives therefore access to the general theory that was presented in Section 4. However, there is still one key property missing that is needed to fully open this powerful toolbox, namely the condition 1 ⊗ 1 ∈ D(δh), where δh denotes the adjoint operator of ∇h, i.e. We shall call δh the directional divergence operator (in the direction h) in the following. δh := (∇h)∗ : L2(S, τ ) ⊗ L2(S, τ ) ⊇ D(δh) → L2(S, τ ), The condition 1 ⊗ 1 ∈ D(δh) would in particular guarantee according to Proposition 4.7 that δh is densely defined and hence that ∇h is closable. But there is actually a shortcut in our situation. We insert here the following lemma which expresses the directional divergence operator δh in terms of the divergence operator δ and which will allow us to conclude directly that the domain of δh is sufficiently large. Lemma 5.4. For any h ∈ L2(R+), the domain D(δh) of the directional divergence operator δh contains Salg ⊙ Salg and we have explicitly δh(U) = δ(U ♯1 ⊗ h ⊗ 1) for all U ∈ Salg ⊙ Salg. In particular, δh is densely defined and we have that 1 ⊗ 1 ∈ D(δh) with δh(1 ⊗ 1) = X(h). Proof. We just have to note that by definition U ♯1 ⊗ h ⊗ 1 ∈ D(δ) for any U ∈ Salg ⊙ Salg and that the corresponding element δ(U ♯1 ⊗ h ⊗ 1) ∈ L2(S, τ ) satisfies hY, δ(U ♯1 ⊗ h ⊗ 1)i = h∇Y, U ♯1 ⊗ h ⊗ 1iB2 = h∇hY, Ui. ∇h : L2(S, τ ) ⊇ D(∇h) → L2(S, τ ) ⊗ L2(S, τ ) is densely defined and closable. The domain D(∇ h ) of its closure h ∇ : L2(S, τ ) ⊇ D(∇ ) → L2(S, τ ) ⊗ L2(S, τ ) h contains the domain D(∇) of ∇. Proof. Basic functional analysis tells us that in this case D(∇ Salg with respect to the norm k · kh 2,1 := (cid:0)kY k2 2,1 that is given by 2 + k∇hY k2 2(cid:1) kY kh 1 2 h for all Y ∈ Salg, whereas the domain D(∇) of ∇ is obtained as the closure of Salg with respect to the norm kY k2,1 = (cid:0)kY k2 2 + k∇Y k2 B2(cid:1) 1 2 for all Y ∈ Salg, as we pointed out in Remark 3.14. Therefore, the desired inclusion D(∇) ⊆ D(∇ as soon as we have established that ) is obtained as the closure of REGULARITY OF DISTRIBUTIONS OF WIGNER INTEGRALS 37 This means that Salg ⊙ Salg ⊆ D(δh) and even more explicit δh(U) = δ(U ♯1 ⊗ h ⊗ 1) for all U ∈ Salg ⊙ Salg. In particular, we may deduce that δh is densely defined and that 1 ⊗ 1 ∈ D(δh) holds true with δh(1 ⊗ 1) = δ(1 ⊗ h ⊗ 1) = X(h). The closability of ∇h, which is implied by the lemma above, will be recorded in the following proposition. But we discuss there in addition that the domain of the closure of ∇h contains the domain of the closure of ∇. Proposition 5.5. Given h ∈ L2(R+). The directional gradient (cid:3) h ) follows (cid:3) (5.1) kY kh 2,1 ≤ max{1,khkL2(R+)}kY k2,1 for all Y ∈ Salg. For that purpose, we make use of Lemma 3.2. This yields Now, the desired inequality (5.1) immediately follows. k∇hY k2 = kh∇Y, hik2 ≤ khkL2(R+)k∇Y kB2. 5.2. Reduction by directional gradients. In the previous subsection, we have seen that directional gradients fit nicely into the general frame of non-commutative derivations. The following proposition, which will be a the core of our reduction method, is therefore an immediate consequence of Proposition 4.14. Proposition 5.6. Take any Y ∈ Sfin. If there are u, v ∈ S such that the conditions Y u = 0 and Y ∗v = 0 are satisfied, then it holds true that h v∗ · (∇ Y ) · u = 0 for all h ∈ L2(R+, R). Proof. Let h ∈ L2(R+, R) be given. Firstly, we recall that the directional gradient ∇ : L2(S, τ ) ⊇ D(∇h) → L2(S, τ ) ⊗ L2(S, τ ), induces according to Corollary 5.3 a real non-commutative derivation, which satisfies in addi- tion the coassociativity relation. Furthermore, its adjont operator, the directional divergence operator δh, satisfies due to Lemma 5.4 the condition 1 ⊗ 1 ∈ D(δh). Thus, we can apply Proposition 4.14, which yields the desired statement. (cid:3) 38 T. MAI Remark 5.7. In the proof of Proposition 5.6 above, we used crucially the properties of directional gradients, which put them nicely in the setting of non-commutative derivations and which therefore allowed us to turn on by Proposition 4.14 the powerful machinery that was build up in Section 4. But recall that one of the crucial ingredients in the proof of Proposition 4.14 were Dabrowski's inequalities 4.9. Thus, concealed in the larger apparatus, we deduced particularly for any Y ∈ D(∇h) according to the inequalities (4.10) that (5.2) kδh(Y ⊗ 1)k2 ≤ khkL2(R+)kY k, kδh(1 ⊗ Y )k2 ≤ khkL2(R+)kY k, and according to the inequalities (4.11) that (5.3) k(id⊗τ )(∇hY )k2 ≤ 2khkL2(R+)kY k, k(τ ⊗ id)(∇hY )k2 ≤ 2khkL2(R+)kY k, However, the semicircular generators that underlie our situation force in fact a much since we have kδh(1 ⊗ 1)k2 = khkL2(R+). stronger result than the inequalities above. In fact, for any Y ∈ Sfin, we have that (5.4) and (5.5) kδh(Y ⊗ 1)k2 = khkL2(R+)kY k2, kδh(1 ⊗ Y )k2 = khkL2(R+)kY k2, k(id⊗τ )(∇hY )k2 ≤ khkL2(R+)kY k2, k(τ ⊗ id)(∇hY )k2 ≤ khkL2(R+)kY k2. This can be seen by considering the chaos decomposition of Y and by using the formulas (5.6) and (5.7) δh(I S δh(1 ⊗ I S n (f ) ⊗ 1) = I S n (f )) = I S n+1(f ⊗ h), n+1(h ⊗ f ) (id⊗τ )(∇hI S (τ ⊗ id)(∇hI S n (f )) = I S n (f )) = I S 1 a h), 1 a f ). n−1(f n−1(h The author is grateful to Yoann Dabrowski for pointing out that this fact should be included for reasons of clarity. Of course, one could argue now that in view of this observation, the discussion around Theorem 4.9 becomes superfluous in the context of this paper. But since there is absolutely no chance to avoid completely a detour through the realm of non-commutative derivations -- even by taking this shortcut -- we decided to present the theory of non-commutative derivations (and in particular the result of Proposition 4.14) in full generality, in order to show the complete picture and to make it ready for its possible use in future investigations. REGULARITY OF DISTRIBUTIONS OF WIGNER INTEGRALS 39 f = (fn)∞n=0 ∈ Ffino 5.3. How to control the reduction. Because Theorem 2.4 is a statement about elements f ∈ F , which break off after finitely many non-zero terms, namely about elements in Ffin := Ffin(L2(R)), we shall take now a closer look on Sfin := nI S(f )(cid:12)(cid:12)(cid:12) as the corresponding space of Wigner integrals, called the finite Wigner chaos. By definition, Sfin is only a subset of L2(S, τ ), but due to (2.1) and Proposition 2.3 it turns out to be in fact a ∗-subalgebra of S. Combining this with the easy observation that Salg is contained in Sfin, we may localize Sfin as intermediate ∗-algebra Salg ⊆ Sfin ⊆ S. Following the lines of the proof in [MSW14, MSW15], we shall introduce now certain operators, which will later allow us to reduce any zero-divisor in Sfin in a controllable way to a zero-divisor in the chaos of order zero by means of Proposition 5.6. Definition 5.8. For any h ∈ L2(R+) and any projection p ∈ S, we consider the linear operator ∆p,h : Sfin → Sfin that is defined by ∆p,hY := (τ ⊗ id)(cid:0)p ⊗ 1 (∇ Y )(cid:1) h for all Y ∈ Sfin. Note that these operators are indeed well-defined since Sfin ⊆ D(∇) ⊆ D(∇ ) holds by Proposition 5.5. The fact that ∆p,h takes its values in Sfin and is made more precise in the following lemma, which moreover shows that ∆p,h "reduces the degree" with respect to the natural grading on Sfin, which is induced by Ffin. Lemma 5.9. Let h ∈ L2(R+) and any projection p ∈ S be given. Let τp be the bounded linear functional on F that is given by h τp : F → C, f 7→ τ (pI S(f )). Now, let f ∈ L2(Rn In fact, if we make use of the chaos decomposition of p, we can write p = I S(g) for some g = (gn)∞n=0 ∈ F , so that τp(f ) = hf, giF holds for all f ∈ F . in L2(R+, L2(Rk−1 +) be given. For 1 ≤ k ≤ n, we may regard f as an element f (k−1,n−k) + ) ⊗ L2(Rn−k + )) ⊂ L2(R+,F ⊗ F ). Using this notation, it holds true that Xk=1 ∆p,hI S n (f ) = n−k(cid:16)(τp ⊗ idF )(cid:16)ZR+ h(t) dt(cid:17)(cid:17). f (k−1,n−k) t (5.8) I S n Proof. It is very easy to check the validity of the formula under question in the case f = f1 ⊗ · · · ⊗ fn. Indeed, we have Xk=1 k−1(f1 ⊗ · · · ⊗ fk−1) ⊗ I S n−k(fk+1 ⊗ · · · ⊗ fn) hfk, hiI S I S n (f ) = ∇ n h and hence n ∆p,hI S n (f ) = = = n Xk=1 Xk=1 Xk=1 n n−k(fk+1 ⊗ · · · ⊗ fn) I S k−1(f1 ⊗ · · · ⊗ fk−1))I S hfk, hiτ (pI S n−k(cid:16)hfk, hiτp(f1 ⊗ · · · ⊗ fk−1) fk+1 ⊗ · · · ⊗ fn(cid:17) n−k(cid:16)(τp ⊗ idF )(cid:16)ZR+ h(t) dt(cid:17)(cid:17), f (k−1,n−k) t I S 40 T. MAI which confirms the desired formula (5.8) in the case f = f1 ⊗ · · · ⊗ fn. By linearity of both of its sides, we conclude that formula (5.8) also holds for any function in the linear span of {f1 ⊗ · · · ⊗ fn f1, . . . , fn ∈ L2(R+)}, +) with respect +), it remains to note that (5.8) stays valid under taking limits with respect to +), which means that we prove the continuity of the left and the right hand side of i.e. for any function in L2(R+)⊙n. Since this linear space is dense in L2(Rn to k · kL2(Rn k · kL2(Rn the formula under question with respect to k · kL2(Rn +). Concerning first the left hand side, we note that n (f )k2 = √nkfkL2(Rn k∇I S +). k∇I S n (f )k2 Indeed, we have according to Proposition 3.17 and the Ito isometry that n (f )k2 n (f )k2 = √nkpkkfkL2(Rn Thus, we obtain the desired bound n (f )i = nkI S n (f ),∇I S 2 = h∇I S k∆p,hI S n (f ), I S +). 2 = nkfk2 L2(Rn +). Concerning now the right hand side of the formula under question, we note that n (f )i = h(cid:0)δ ∇(cid:1)I S n (f )k2 ≤ kpkk∇I S h(t) dt(cid:13)(cid:13)(cid:13)L2(Rk−1 ZR+ (cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13) f (k−1,n−k) t + )⊗L2(Rn−k holds for 1 ≤ k ≤ n, which yields by the Ito isometry + ) ≤ khkL2(R+)kfkL2(Rn +) I S n−k(cid:16)(τp⊗ idF )(cid:16)ZR+ f (k−1,n−k) t h(t) dt(cid:17)(cid:17)(cid:13)(cid:13)(cid:13)2 f (k−1,n−k) t (τp ⊗ idF )(cid:16)ZR+ ≤ (cid:13)(cid:13)(cid:13) ≤ kpk2(cid:13)(cid:13)(cid:13) ZR+ f (k−1,n−k) t ≤ kpk2khkL2(R+)kfkL2(Rn +). h(t) dt(cid:17)(cid:13)(cid:13)(cid:13)L2(Rn−k + ) + )⊗L2(Rn−k + ) h(t) dt(cid:13)(cid:13)(cid:13)L2(Rk−1 This concludes the proof. (cid:3) By applying iteratively operators of the form ∆p,h to a fixed element in the finite Wigner chaos Sfin, we will therefore reach the chaos of order zero after finitely many steps. The following proposition provides an explicit formula for the output of this procedure. Proposition 5.10. Let f = (fn)∞n=0 ∈ Ffin be given and let N ∈ N be chosen such that fn = 0 for all n ≥ N + 1. Then, for any choice of functions h1, . . . , hN ∈ L2(R+) and projections p1, . . . , pN , it holds true that ∆pN ,hN · · · ∆p1,h1I S(f ) = τ (p1)· · · τ (pN )hfN , h1 ⊗ · · · ⊗ hNi 1. Before continuing with the proof of the general statement, we first focus on the special case of simple functions. Remark 5.11. We note that for any Y ∈ Salg ∆pN ,hN · · · ∆p1,h1Y = (τ⊗N ⊗ id)(cid:0)p1 ⊗ · · · ⊗ pN ⊗ 1 (∇ h1,...,hN Y )(cid:1) REGULARITY OF DISTRIBUTIONS OF WIGNER INTEGRALS 41 holds, where the iterated gradient ∇ h1,...,hN : Salg → S⊙(N +1) hN ) . . . (id⊗∇ alg h1,...,hN := (id⊗(N−1) ⊗∇ ∇ h1. h2)∇ is defined by Thus, the statement of Proposition 5.10 becomes apparent in the case where f = (fn)∞n=0 ∈ +). Indeed, since I S(f ) decomposes by the condi- Ffin consists of simple functions fn ∈ E(Rn tions that are imposed on f as and because obviously we see that I S(f ) = I S 0 (f0) + I S 1 (f1) + · · · + I S N (fN ) h1,...,hN I S k (fk) = 0 ∇ h1,...,hN I S(f ) = ∇ ∇ for 0 ≤ k ≤ N − 1, h1,...,hN I S N (fN ). By using Proposition 3.13, we get h1,...,hN I S Combining these observations yields ∇ N (fN ) = hf, h1 ⊗ · · · ⊗ hNi 1⊗(N +1). ∆pN ,hN · · · ∆p1,h1I S(f ) = τ (p1)· · · τ (pN )hfN , h1 ⊗ · · · ⊗ hNi 1, which is the stated formula. Proof of Proposition 5.10. For the general case, we make use of Lemma 5.9. Applying for- mula (5.8) iteratively, yields that for 1 ≤ m ≤ N for some f (m) ∈ Ffin, where f (m) we see that ∆pm,hm . . . ∆p1,h1I S(f ) = I S(f (m)), n = 0 for all n ≥ N − m + 1. Moreover, if we put f (0) := f , f (m) N−m(tm+1, . . . , tN ) = τ (pm)ZR+ f (m−1) N−m+1(tm, tm+1, . . . , tN ) hm(tm) dtm for all 1 ≤ m ≤ N − 1 and f (N ) 0 = τ (pN )ZR+ f (N−1) 1 (tN ) hN (tN ) dtN . Hence, the only term that survives in ∆pN ,hN . . . ∆p1,h1I S(f ) is induced by which gives the stated formula. f (N ) 0 = τ (p1)· · · τ (pN )hfN , h1 ⊗ · · · ⊗ hNi, (cid:3) 5.4. Absence of zero divisors. Our discussion in the previous subsections has shown that directional gradients allow us to transfer tools from the theory of non-commutative derivations as presented in Section 4 to the setting of free stochastic calculus. Moreover, we have convinced ourselves that directional gradients ∇h induce operators ∆p,h, which satisfy the general conditions for performing our reduction method. Putting things together, we obtain the following theorem, of which the desired Theorem 2.4 will be a corollary. Theorem 5.12. There are no zero divisors in Sfin. More precisely, if 0 6= Y ∈ Sfin is given, then there is no 0 6= u ∈ S such that Y u = 0. 42 T. MAI Proof. Contrarily, assume that there are 0 6= Y ∈ Sfin and 0 6= u ∈ S such that Y u = 0. We may write Y = I S(f ) for some f ∈ Ffin of the form f = (fn)∞n=0. Moreover, we may choose N ∈ N such that fN 6= 0 but fn = 0 for all n ≥ N + 1. Now, we fix arbitrary functions h1, . . . , hN ∈ L2(R+, R). Recall that whenever we have an element X ∈ S such that Xu = 0 holds, then there exists (since we assumed that u 6= 0) a non-zero projection p ∈ S such that X∗p = 0. This is in fact an easy consequence of the Murray-von Neumann equivalence of the left and right support projections of X; see also [MSW15, Lemma 3.14]. Thus, by applying Proposition 5.6 iteratively, we may find non-zero projections p1, . . . , pN ∈ S such that (∆pN ,hN . . . ∆p1,h1Y )u = 0. According to Proposition 5.10, this means that τ (p1)· · · τ (pN )hfN , h1 ⊗ · · · ⊗ hNi u = 0. Since we have by assumption u 6= 0 and furthermore τ (p1)· · · τ (pN ) 6= 0, because p1, . . . , pN are non-zero projections, it follows Inasmuch as the linear span of hfN , h1 ⊗ · · · ⊗ hNi = 0. {h1 ⊗ · · · ⊗ hN h1, . . . , hN ∈ L2(R+, R)} is dense in L2(RN + ), the previous insight yields fN = 0, which contradicts the condition according to which N was chosen. Thus, the assumption made above was wrong, so that the statement of the theorem must be true. (cid:3) + ) with respect to k · kL2(RN We finish by showing that Theorem 2.4 is indeed a consequence of Theorem 5.12 above. In fact, we will deduce Theorem 2.4 exactly in the same way as it was done for the analogous statement in [MSW15]. Proof of Theorem 2.4. More generally, by allowing right from the beginning a constant sum- mand I S 0 (f0), we show the following: the distribution µY of any self-adjoint element Y ∈ Sfin, which does not belong to the chaos of order zero, cannot have atoms. Let Y ∈ Sfin be given. If Y does not belong to the chaos of order zero, we can write it as Y = I S(f ) = I S 0 (f0) + I S 1 (f1) + · · · + I S N (fN ) for some f = (fn)∞n=0 ∈ Ffin, which is stationary zero after fN 6= 0 for some N ∈ N. (Note that N 6= 0 means abstractly speaking that Y is not constant, as it was assumed in [MSW15].) Then, we observe that any atom α of the distribution µY of Y , i.e. any α ∈ R satisfying µY ({α}) 6= 0, leads by the spectral theorem for bounded self-adjoint operators on Hilbert spaces to a non-zero projection u satisfying (Y − α1)u = 0. Now, Theorem 5.12 tells us that Y = α1, which contradicts fN 6= 0. (cid:3) References [BS98] P. Biane and R. Speicher, Stochastic calculus with respect to free Brownian motion and analysis on Wigner space., Probab. Theory Relat. Fields 112 (1998), no. 3, 373 -- 409. [CS03] F. Cipriani and J.-L. Sauvageot, Derivations as square roots of Dirichlet forms, J. Funct. Anal. 201 (2003), no. 1, 78 -- 120. [CS05] A. Connes and D. Shlyakhtenko, L2-cohomology for von Neumann algebras., J. Reine Angew. Math. 586 (2005), 125 -- 168. REGULARITY OF DISTRIBUTIONS OF WIGNER INTEGRALS 43 [CS15] I. Charlesworth and D. Shlyakhtenko, Regularity of Polynomials arXiv:1408.0580v2 (2015). in Free Variables, [Dab10] Y. Dabrowski, A note about proving non-Γ under a finite non-microstates free Fisher information assumption, J. Funct. Anal. 258 (2010), no. 11, 3662 -- 3674. [Dab14] , A free stochastic partial differential equation, Ann. Inst. Henri Poincar´e, Probab. Stat. 50 (2014), no. 4, 1404 -- 1455. [GS09] A. Guionnet and D. Shlyakhtenko, Free diffusions and matrix models with strictly convex interaction, Geom. Funct. Anal. 18 (2009), no. 6, 1875 -- 1916. [Ito51] K. Ito, Multiple Wiener integral, J. Math. Soc. Japan 3 (1951), 157 -- 169. [Ito52] [KNPS12] T. Kemp, I. Nourdin, G. Peccati, and R. Speicher, Wigner chaos and the fourth moment, The , Complex multiple Wiener integral, Jpn. J. Math. 22 (1952), 63 -- 86. Annals of Probability 40 (2012), no. 4, 1577 -- 1635. [Lin91] P. A. Linnell, Zero divisors and group von Neumann algebras, Pac. J. Math. 149 (1991), no. 2, 349 -- 363. [Lin92] [Lin93] [Lin98] , Zero divisors and L2(G), C. R. Acad. Sci., Paris, S´er. I 315 (1992), no. 1, 49 -- 53. , Division rings and group von Neumann algebras, Forum Math. 5 (1993), no. 6, 561 -- 576. , Analytic versions of the zero divisor conjecture, Geometry and cohomology in group theory, Cambridge: Cambridge University Press, 1998, pp. 209 -- 248. [MSW14] T. Mai, R. Speicher, and M. Weber, Absence of algebraic relations and of zero divisors under the assumption of finite non-microstates free Fisher information, arXiv:1407.5715v2 (2014). [MSW15] , Absence of algebraic relations and of zero divisors under the assumption of full non- [NP13] microstates free entropy dimension, arXiv:1502.06357v2 (2015). I. Nourdin and G. Peccati, Poisson approximations on the free Wigner chaos, Ann. Probab. 41 (2013), no. 4, 2709 -- 2723. [NS06] A. Nica and R. Speicher, Lectures on the combinatorics of free probability, Cambridge: Cambridge University Press, 2006. [Nua06] D. Nualart, The Malliavin calculus and related topics. 2nd ed., 2nd ed. ed., Berlin: Springer, 2006. , Malliavin calculus and its applications, Providence, RI: American Mathematical Society [Nua09] [Shi78] (AMS), 2009. I. Shigekawa, Absolute continuity of probability laws of Wiener functionals, Proc. Japan Acad., Ser. A 54 (1978), 230 -- 233. [Shi80] , Derivatives of Wiener functionals and absolute continuity of induced measures, J. Math. Kyoto Univ. 20 (1980), 263 -- 289. [Shl14] D. Shlyakhtenko, Free Entropy Dimension and Atoms, arXiv:1408.0580v1 (2014). [Spe03] R. Speicher, Free calculus, Quantum probability communications. Vol XII. Lectures from the Greno- ble summer school, Grenoble, France, June 1998, River Edge, NJ: World Scientific, 2003, pp. 209 -- 235. [Spe13] , Asymptotic eigenvalue distribution of random matrices and free stochastic analysis, Ran- dom matrices and iterated random functions. Selected papers based on the presentations at the workshop, Munster, Germany, October 4 -- 7, 2011, Berlin: Springer, 2013, pp. 31 -- 44. [SS15] D. Shlyakhtenko and P. Skoufranis, Freely independent random variables with non-atomic distribu- tions., Trans. Am. Math. Soc. 367 (2015), no. 9, 6267 -- 6291. [VDN92] D. Voiculescu, K.J. Dykema, and A. Nica, Free random variables. A noncommutative probability approach to free products with applications to random matrices, operator algebras and harmonic analysis on free groups, Providence, RI: American Mathematical Society, 1992. [Voi93] D. Voiculescu, The analogues of entropy and of Fisher's information measure in free probability theory, I, Commun. Math. Phys. 155 (1993), no. 1, 71 -- 92. [Voi94] , The analogues of entropy and of Fisher's information measure in free probability theory, II, Invent. Math. 118 (1994), no. 3, 411 -- 440. [Voi96] , The analogues of entropy and of Fisher's information measure in free probability theory, III: The absence of Cartan subalgebras, Geom. Funct. Anal. 6 (1996), no. 1, 172 -- 199. [Voi97] , The analogues of entropy and Fisher's information measure in free probability theory, IV: Maxiumum entropy and freeness, Fields Inst. Commun. 12 (1997), 293 -- 302. 44 T. MAI [Voi98] , The analogues of entropy and of Fisher's information measure in free probability theory, V: Noncommutative Hilbert transforms, Invent. Math. 132 (1998), no. 1, 189 -- 227. [Voi99] [Voi00] , The analogues of entropy and of Fisher's information measure in free probability the- ory, VI: Liberation and mutual free information, Adv. Math. 146 (1999), no. 2, 101 -- 166, art. no. aima.1998.1819. , Lectures on free probability, Lectures on probability theory and statistics. ´Ecole d'´Et´e de Probabilit´es de Saint-Flour XXVIII - 1998. Summer school, Saint-Flour, France, August 17 -- September 3, 1998, Berlin: Springer, 2000, pp. 279 -- 349. , Free entropy, Bull. Lond. Math. Soc. 34 (2002), no. 3, 257 -- 278. [Voi02] [Wie38] N. Wiener, The homogeneous chaos, Am. J. Math. 60 (1938), 897 -- 936. Universitat des Saarlandes, FR 6.1−Mathematik, 66123 Saarbrucken, Germany E-mail address: [email protected]
1003.5156
4
1003
2011-12-28T10:14:49
Classification of spin structures on the noncommutative n-torus
[ "math.OA", "math-ph", "math-ph", "math.QA" ]
We classify spin structures on the noncommutative torus, and find that the noncommutative n-torus has 2^n spin structures, corresponding to isospectral deformations of spin structures on the commutative n-torus. For n>3 the classification depends on Connes' spin manifold theorem. In addition, we study unitary equivalences of these spin structures.
math.OA
math
CLASSIFICATION OF SPIN STRUCTURES ON THE NONCOMMUTATIVE n-TORUS JAN JITSE VENSELAAR Abstract. We classify spin structures on the noncommutative torus, and find that the noncommutative n-torus has 2n spin structures, corresponding to isospectral deformations of spin structures on the commutative n-torus. For n ≥ 4 the classification depends on Connes' spin manifold theorem. In addition, we study unitary equivalences of these spin structures. 1. Introduction The different possible spin structures on a differentiable manifold were classified in the work of Milnor [27]; for example, on a (commutative) n-torus, there exist 2n inequivalent spin structures. No such general classification of spin structures in currently know in noncommutative geometry -- this amounts to classifying the possible real spectral triple structures on a C∗-algebra. In this paper we prove that there exist precisely 2n different real spectral triples on a noncommutative n-torus, and that these structures are isospectral deformations of spin structures on the commutative n-torus. The noncommutative torus A(Tn θ ), or irrational rotation algebra, is one of the first nontrivial examples of a noncommutative topological manifold, given as a deformation of the usual commutative torus [17] [31] [16]. The parameter θ is a number for a noncommutative 2-torus, and an antisymmetric n × n matrix for higher dimensional tori. The analog of putting a spin structure and a metric on this algebra is to enhance it into a real spectral triple in the sense of Connes [8] [9]. This introduces a set of extra parameters τ i, which are the analogue of the size of the torus. The non- commutative n-torus, both topologically and with spin structure, has found many applications in mathematical physics, for example [4], [25] and [3]. A noncommu- tative spin structure can certainly be constructed by deforming a spin structure on the commutative n-torus [13], so the question becomes whether this deformation gives all possible spin structures on the noncommutative n-torus. In dimension 2, the problem was solved by Paschke and Sitarz [28, Theorem 2.5], who showed that a noncommutative 2-torus admits exactly 4 different real spectral triples (which are deformations of spin structures on the commutative torus). This result can be reformulated as follows: any real spectral triple which is equivariant with respect to a 2-torus action in the sense of [34] (see section 2.1), Date: November 6, 2018. 2010 Mathematics Subject Classification. 58B34; 46L87. Key words and phrases. spectral geometry, noncommutative geometry, real spectral triples. The author thanks Gunther Cornelissen for discussions and encouragement, and the referee for pertinent remarks. 1 2 JAN JITSE VENSELAAR is an isospectral deformation of a spin structure on a commutative 2-torus. Note that an equivariant action of n-torus is different from an n-torus action as in [11], the former is a condition on the spectral triple, the latter is an action along which the algebra is deformed. Our first result is that the theorem of Paschke and Sitarz holds true in arbitrary dimension: Theorem A. All irreducible real spectral triples with an equivariant n-torus actions are isospectral deformations of spin structures on an n-torus. The proof of [28, Theorem 2.5] does not generalize readily to higher dimensions. Rather than working with the grading operator, which only gives nontrivial con- ditions in the even-dimensional case, we use the reality operator first. Then we establish that out of several possible candidate structures only one satisfies the growth condition and compact resolvent condition on the Dirac operator. Also, our proof uses at a crucial point Connes' reconstruction theorem [10, Theorem 11.5]. We describe the spin structures explicitly in Theorem C. In a celebrated paper [1] (see also [2, Theorem 1]), Adams used the classification of independent vector fields on spheres to deduce elementary results on Radon- Hurwitz numbers of certain classes of matrices. Similarly, our Theorem A can be used to prove the following elementary result on Hermitian matrices, for which we do not know an elementary proof: Corollary 1. A set of 2b × 2b Hermitian matrices {Ai}n that the equation i=1, where n = 2b + 1, such det xiAi = 0, (cid:33) (cid:32)(cid:88) n(cid:89) i i i=1 in Rn, generate a Clifford algebra if and only only has the zero solution (xi = 0)n if (cid:88) for some nonzero λ ∈ R. σ∈Sn sign(σ) Aσ(i) = λIdk, After we have obtained a classification, we study equivalences between different real spectral triples on the same noncommutative n-torus. For a commutative n- torus, the diffeomorphisms of a torus act affine on the set of spin structures identified with the vector space Zn 2 , as shown in [15]. In particular, for the commutative 2- torus, there are two orbits, one consisting of one element, and the other consisting of three elements. In the case of the noncommutative torus the full diffeomorphism group is not known when n > 2. Restricting to inner automorphisms of the algebra, we can show the the following. Theorem B. Except for a set of θ of measure 0, the different spin structures of the smooth noncommutative n-torus A(Tn θ ) cannot be unitarily equivalent by an inner automorphism of the algebra. We also compute the action of unitary transformations induced by some outer automorphisms on the spin structures. These are the action of SL(2, Z) on the noncommutative 2-torus and the flip automorphism on the noncommutative n-torus for n > 2. CLASSIFICATION OF SPIN STRUCTURES ON THE NONCOMMUTATIVE n-TORUS 3 The set of θ of measure 0 in Theorem B is determined by some Diophantine approximation conditions given in [5], and includes the θ with only rational entries. In addition to unitary equivalences of real spectral triples, one could also look at Morita equivalences of real spectral triples [37, Chapter 7.2], even though it is not a true equivalence since it is not symmetric in general [12, Remark 1.143]. It would be interesting to study Morita equivalences of spin structures on noncommutative tori, especially since all Morita equivalences of the algebra of functions on the smooth noncommutative tori are known [32] [20, Theorem 1.1]. We hope to return to this question in the future. 2. Definition of a real spectral triple Since there are various, slightly different, definitions of real spectral triples, cf. [9, Pages 159 -- 162],[37, Chapter 3],[22, Chapter 10], and we need to refer to the axioms unambiguously, we will explicitly state the definition of real spectral triple we use. For the definition of an equivariant spectral triple, see section 2.1. A spin structure on a manifold M , of dimension n, is a nontrivial double cover of the principal SO(n) bundle of orthogonal frames of the tangent bundle T M [26, Chapter II.1]. After [10], we know that a real spectral triple is the right noncommutative geometry analog of a spin structure on a manifold. For a real spectral triple we have the following data: operators, • A unital, Fr´echet algebra A, ∗-isomorphic to a proper dense subalgebra of a C∗-algebra which is stable with respect to the holomorphic functional calculus, • A separable Hilbert space H with a representation of A acting as bounded • An unbounded self-adjoint operator D, called 'Dirac operator', • An antilinear isometry J of H onto itself, called 'reality operator', • An integer n ≥ 0, called 'dimension', • If n is even, a self-adjoint unitary operator Γ of H onto itself, such that Γ2 = Id, called 'grading operator'. We call the the spectral triple even in this case. These objects also satisfy the Axioms 1 to 9 in order for them to be called a real spectral triple of dimension n. Axiom 1 (Compact resolvent). The Dirac operator D has compact resolvent, that is, D has finite dimensional kernel and D−1 (defined on the orthogonal complement of the kernel) is a compact operator. Furthermore, for all a ∈ A, [D, π(a)] is a bounded operator. Axiom 2 (Grading operator). If n is even, the Z2 grading operator Γ splits the Hilbert space H as H+ and H−, where H± is the (±) eigenspace of Γ. The operator D is odd with respect to this operator, DΓ = −ΓD, and the representation π of A on H is even, so we can write (cid:18)a 0 (cid:19) (cid:18) 0 D− (cid:19) π(a) = D = where D+ and D− are adjoint to each other. Axiom 3. The operators J, D and Γ satisfy the commutation relations from Ta- ble 1, and the operator J is unitary: J† = J−1. 0 a D+ , 0 We will write bo = Jb∗J†. The above formulas establish that the opposite algebra: lies within the commutator of A. Recall that a Hochschild k-chain is defined as an element c of Ck(M,A) = M ⊗ A⊗k, with M a bimodule over A. A boundary map b : Ck → Ck−1 is defined as [a, Jb∗J†] = 0. [[D, a], Jb∗J†] = 0. Ao = {ao = J†a∗Ja ∈ A}, (cid:88) i=0,k b = (−1)idi, 4 JAN JITSE VENSELAAR Table 1. Signs of the spectral triple 0 n mod 8 7 J 2 = ±Id (J ) + + − − − − + + JD = ±DJ (D) + − + + + − + + JΓ = ±ΓJ (Γ) + − − + 1 2 3 4 5 6 Axiom 4 (Dimension). The eigenvalues µk of D−1, arranged in decreasing order, grow asymptotically as µk = O(k−n), for an integer n (called the dimension). Axiom 5 (First order condition). For all a, b ∈ A the following commutation relations hold: (1) (2) d0(m ⊗ a1 ⊗ ··· ⊗ ak) = ma1 ⊗ a2 ··· ⊗ ak, di(m ⊗ a1 ⊗ ··· ⊗ ak) = m ⊗ a1 ⊗ ··· ⊗ aiai+1 ⊗ ··· ⊗ ak, dk(m ⊗ a1 ⊗ ··· ⊗ ak) = akm ⊗ a1 ⊗ ··· ⊗ ak−1. Since b2 = 0, this makes Ck(M,A) into a chain complex. Axiom 5 gives a representation of Hochschild k-chains Ck(A,A ⊗ Ao) on H by πD((a ⊗ bo) ⊗ a1 ⊗ ··· ⊗ ak) = abo[D, a1] . . . [D, ak]. (3) Axiom 6 (Orientability). There is a Hochschild cycle c ∈ Zn(A,A⊗Ao) such that πD(c) = Γ when n is even, and πD(c) = Id when n is odd. Axiom 7 (Regularity). For all a ∈ A, both a and [D, a] belong to the domain of D = k=1 Dom(cid:0)δk(cid:1), where the derivation δ is given by δ(T ) = [D, T ], with smoothness(cid:84)∞ k=1 Dom(cid:0)Dk(cid:1) is a Axiom 8 (Finiteness). The space of smooth vectors H∞ = (cid:84)∞ where −(cid:82) is the noncommutative integral (defined for example in [37, Chapter 5]). finitely generated projective left A module. Also, there is a Hermitian pairing (ηξ) on this module, given by (ηξ)D−n, (cid:104)η, ξ(cid:105) = − D∗D. (cid:90) √ Axiom 9 (Poincar´e duality). The Fredholm index of the operator D yields a non- degenerate intersection form on the K-theory ring of the algebra A ⊗ Ao. CLASSIFICATION OF SPIN STRUCTURES ON THE NONCOMMUTATIVE n-TORUS 5 Finally, we restrict our attention to irreducible spectral triples, that is, the only operators commuting with the action of the algebra and D are the scalars. In the case A is commutative, this is equivalent to demanding the manifold is connected, see [9, Remark 6 on p.163]. 2.1. Equivariant spectral triples. There are different candidates for the notion of symmetries of noncommutative geometries. One obvious candidate is the group of automorphisms of the algebra A, but for noncommutative algebras, this group can be very small, while it seems that there should be more symmetries available. An attempt to enlarge this group of symmetries in an interesting way is the notion of equivariant spectral triples. Equivariant spectral triples were introduced in [34]. One describes symmetries of spectral triples in the form of Hopf algebras. In this paper we are interested in n-dimensional spectral triples, which are equivariant with respect to a Hopf algebra with n different commuting derivations. This, together with the irreducibility condition, is the analog of a connected compact homogeneous space in commutative geometry. In the context of Hopf algebras, we shall use Sweedler's notation for the coprod- uct: ∆h = h(1) ⊗ h(2). See for example [24, Chapter 3] for an introduction to Hopf algebras, and some standard notation. An equivariant real spectral triple is a real spectral triple (A,H, D, J) together with a Hopf algebra H, with multiplication µ, unit η, comultiplication ∆, counit  and antipode S, and an antilinear involution ∗ making H into a ∗-algebra such that ∆h∗ = (∆h) ∗⊗∗ (h∗) = (h) (S ◦ ∗)2 = Id. Recall that an H-module algebra is an algebra A with a complex linear repre- sentation ρ of H on A such that A is a linear space, and ρ respects the algebra structure: for all h ∈ H, a1, a2 ∈ A. When A is an H-module algebra we define an equivariant (left,right) A-module to be a (left,right) A-module M such that (cid:1) , ρ(h)(a1a2) =(cid:0)ρ(h(1))a1 (cid:1)(cid:0)ρ(h(2))a2 ρM (h)(am) =(cid:0)ρA(h(1))a(cid:1)(cid:0)ρM (h(2))m(cid:1) , for all h ∈ H, a ∈ A, m ∈ M . The objects of the equivariant real spectral triple transform in a compatible way under the action of the Hopf algebra: module. • The algebra A is an H-module algebra. • There is a dense subspace V ⊂ H such that V is an equivariant left A- • For every h ∈ H, the Dirac operator is equivariant, [D, h] = 0 on the • The action of H op is well defined on the opposite algebra Ao via the equality • If the spectral triple is even, [Γ, h] = 0. (dense) intersection of the domain of D and V . J−1hJ = (Sh) ∗ . In the case of equivariance with respect to a torus action, we take the universal enveloping algebra U (tn) of the familiar Lie algebra tn of the n-torus. This means that we have a basis of derivations δi, with a representation ρ on H such that for 6 JAN JITSE VENSELAAR a ∈ A(Tn θ ), v ∈ H: (4a) (4b) (4c) (4d) (4e) δiδj = δjδi, ∆(δi) = δi ⊗ Id + Id ⊗ δi, ρ(δi)π(a)v = (π(δia) + π(a)) v, ρ(δi)Dv = Dρ(δi)v, ∗ ρ(δi)Jv = −Jρ(δi) v. 3. Outline of the classification The outline of the proof of Theorem A is as follows. First in section 4 we determine the action of the algebra on the Hilbert space, such that the equivariance condition is met. This action is already well-known, but we derive it to show that there are no other possibilities. In section 5 we move to real spectral triples, and determine possible forms of the reality operator J, by considering the anti- isomorphism A(Tn θ )o and the equivariance condition (4e). We find several possible families of real spectral triples, only one of which consists of isospectral deformations of spin structures on the commutative torus. In the next section, section 6, we determine the classes of possible Dirac operators for each candidate family of real spectral triples using equivariance of the Dirac operator and the first- order condition, and show that only the isospectral deformation family is compatible with the compact resolvent condition. θ ) (cid:55)→ A(Tn In section 7 we determine that the parameters τ in the Dirac operators must be linearly independent vectors spanning Rn, using the Hochschild homology condi- tion and earlier results on the Hochschild homology on noncommutative tori. The last step in the classification is done in section 8, where we use the spin manifold reconstruction theorem to show that the Dirac operator is really a Dirac operator in the sense of spin geometry. After the classification, we give in section 9 an explicit description of the con- structed real spectral triples on the noncommutative torus. Finally, in section 10, we study unitary equivalences of the real spectral triples, and show that unitary equivalences induced by inner automorphisms of the algebra do not change the spin structure for almost all θ. When n = 2 we show that the known outer automorphisms do change the spin structure, if the real spectral triple is an isospectral deformation of a nontrivial spin structure. When n > 2, we show that the flip automorphism cannot change the spin structure. 4. Hilbert space and algebra We look for possible equivariant representations of the algebra of functions of the smooth noncommutative torus, and give a basis of the Hilbert space H for which equivariance is obvious. We do not use any special conditions from the definition of a real spectral triple, except that the Hilbert space should be separable. We denote the noncommutative torus, or more precisely, the algebra of con- tinuous functions on the noncommutative n-torus, by A(Tn θ ), where θ is an anti- symmetric real n × n matrix. As Hopf algebra symmetry for which the algebra representation must be equivariant we take the Hopf algebra generated by n inde- pendent commuting elements δ1, . . . , δn. CLASSIFICATION OF SPIN STRUCTURES ON THE NONCOMMUTATIVE n-TORUS 7 The algebra of smooth or continuous functions on the noncommutative torus is generated by unitary elements Ue1, . . . , Uen such that Uek Uel = exp (2πiθkl) Uel Uek , with θkl the component of the matrix at position (k, l). As a short-hand notation, we will write e(·) = exp(2πi·). The Hopf algebra action of our basis elements on the unitary generators is [34]: δiUej = Uej if i = j, and 0 otherwise. If we interpret the ej as the j-th basis vector of Zn, we can write more generally unitary elements of the algebra as:  1 2 (cid:88)  (Ue1)x1 (Ue2 )x2 ··· (Uen)xn, Ux = e for x =(cid:80) xjej. The Fr´echet algebra of smooth functions on the noncommutative xjθjkxk k>j torus is a dense subalgebra of this algebra. Just as for the algebra, we will write δx for the derivation given by δx = δx1δx2 ··· δxn. From the definitions, it is immediate that δxUy = (x · y) Uy, (5) where (x · y) is the standard inner product on Zn. a basis of H0 we choose mutual eigenvectors eµ of the derivations: Now we look for a Hilbert space H0 which is an equivariant left Aθ-module. As where the µ form a countable subset of Rn. ρ(δi)eµ = µieµ. In order for the spectral triple to be a noncommutative torus, we demand that the action of the algebra is equivariant with respect to a torus action, as in equa- tion (4c). Written out for the unitary generators Ux, we see: π0(Ux)eµ = ux,µeµ+x, with ux,µ ∈ C to be determined. Thus for the minimal irreducible equivariant representation the µ will lie in a translate of a lattice: (cid:77) m∈Zn H0 = Hµ0+m, ∼= C. There are no restrictions yet on µ0, these will be deter- (6) where each Hµ0+m mined later. Since the Ux should be unitary, we have that (7) (cid:104)eν , ux,µeµ+x(cid:105) = (cid:104)u−x,ν eν−x, eµ(cid:105) ⇒ ux,µδν,µ+x = u−x,ν δν−x,µ. Finally the definition of Ux in terms of Ui gives the relations (8) ux+y,µ = e( x · θy)ux,µ+yuy,µ. 1 2 8 JAN JITSE VENSELAAR Lemma 1. Up to unitary transformations of H any unitary equivariant represen- tation of A(Tn θ ) on H is given by (cid:18) 1 2 (cid:19) (9) with A any n × n matrix such that A − At = θ. πA 0 (Ux)eµ = e x · Ax + x · Aµ eµ+x, Since the representations πA 0 , given by different matrices A such that A−At = θ are equivalent, we drop the A from the notation, and just write π0 for a represen- tation defined by equation (9). condition (4c) we can write any element w ∈ H as a unique sum(cid:80) Proof. It is clear that for any matrix A the representation given above satisfies the relations (7) and (8). Given two representations π and π(cid:48) satisfying the equivariance x∈Zn λxπ(Ux)e0 with (λ)x∈Zn ∈ (cid:96)2(Zn). In other words, e0 is a cyclic vector with respect to the algebra action. Now construct an operator T : H → H by setting T e0 = e0 and extending by T w = T λxπ(Ux)e0 = λxπ(cid:48)(Ux)e0. This is well defined if π and π(cid:48) are representations of the same algebra A(Tn θ ), since they satisfy the same algebra relation (8), and it is an invertible map on H, because both π(Ux)e0 and π(cid:48)(Ux)e0 span the Hilbert space if we take all x ∈ Zn. By construction T −1π(cid:48)(Ux)T = π(Ux), and it is a unitary transformation, because we can calculate: (cid:88) x∈Zn (cid:88) x∈Zn (cid:104)T v, T w(cid:105) = = ¯λxµy(cid:104)π(cid:48)(Ux)e0, π(cid:48)(Uy)e0(cid:105) ¯λxµy(cid:104)π(cid:48)(Ux)e0, π(cid:48)(Uy)e0(cid:105) (cid:88) (cid:88) (cid:88) x,y∈Zn x−y=0 = x∈Zn = (cid:104)v, w(cid:105), ¯λxµ−x since (cid:104)ex, ey(cid:105) = 0 if x (cid:54)= y, and the representations π and π(cid:48) are unitary. (cid:3) There might be more unitary equivalences of the algebra, a question which we explore in section 10, but for now we were only interested in possible representations of the algebra. 5. Reality operator In this section, we derive conditions on the equivariant reality operator J, to give us a real spectral triple. We will only consider conditions following from relations between A, H and J and the equivariance condition (4e). No use is made of the Dirac operator in this section. The equivariance condition for J is given in (4e): ρ(l)Jv = −Jρ(l∗)v, (10) for v ∈ H, and l an element of the Hopf algebra of the symmetries of the non- commutative n-torus. For our basic derivations δi, we have δ∗ i = δi, so this just CLASSIFICATION OF SPIN STRUCTURES ON THE NONCOMMUTATIVE n-TORUS 9 means (11) δiJeµ = −µiJeµ, for all i. Since we are working with the representation given in (6), we see that J must map an element eµ of the basis to e−µ. The Tomita involution J0(a) = a∗ [35] gives a commuting representation, but it does not follow the commutation relations in Table 1, so we will have to enlarge our Hilbert space. Define (12) H := Hj, (cid:77) j∈I (cid:88) with each Hj as in (6) and I an index set. Every nondegenerate representation of an involutive Banach algebra is a direct sum of cyclic representations [36, Proposition I.9.17], and because of the equivariance condition (4c), we get that the only cyclic representations we can consider are the ones given by Lemma 1. We write basis vectors of this Hilbert space as eµ,j with µ ∈ Zn and j ∈ I. For different j ∈ I, lattices spanned by µ could a priori be shifted by a different amount, satisfying (11), but we will see in section 6 that this cannot be the case if the spectral triple is irreducible. Table 1, and so that for every a ∈ A(Tn satisfying equation (1). The image of A(Tn is denoted by A(Tn We look for an antilinear operator J such that J 2 = J Id, with signs as in θ ), Ja∗J−1 commutes with all b ∈ A(Tn θ ), θ ) under the isomorphism a (cid:55)→ Ja∗J−1 x for the image of Ux. θ )o. We write U o Firstly, equation (10) has as a consequence that J acts as follows on elements of the basis eµ,j: (13) We can thus write Jeµ,j = Λ(−µ)J0eµ,j, with Λ(−µ) some unitary or skew unitary bounded linear functional, and J0 an antilinear diagonal operator: Jeµ,j = akj(µ)e−µ,k. J0aeµ,j = a∗e−µ,j. If we apply J twice, we should get J Id, which can be written as J · Jeµ,j = JΛ(−µ)e−µ,j = Λ(µ)Λ∗(−µ)e−µ,j, and so we see that (14) ∗ Λ(µ)Λ(−µ) = J Id. By applying J on a unitary generator Ux of Aθ we get the following condition on Λ(µ): Lemma 2. The map a (cid:55)→ Ja∗J† is an isomorphism into the commutant, if and only if for all x, y ∈ Zn: Λ(x + y) = e(x · Ay + y · Ax)Λ(x)Λ(0) (15a) Λ(y), where Jeµ,j = Λ(−µ)J0eµ,j, with J0 the Tomita involution, and † (15b) (15c) † Λ(x)Λ(x) ∗ Λ(x)Λ(−x) = Id, = J Id. 10 JAN JITSE VENSELAAR As a consequence xeµ,j = e(µ · Ax + U o x · Ax)Λ(cid:48)(x)Λ(0) † (16) where Λ(µ) = e(µ · Aµ)Λ(cid:48)(µ), and Λ(cid:48)(µ)ij = cije(φij(µ)) with cij ∈ C and φij : eµ+x,j, Rn → R such that(cid:80) j∈I cije(φij(−µ))c∗ jke(−φij(µ)) = J δik. 1 2 xJ† using equation (13): x = JU∗ Proof. First we calculate U o xeµ,j = JU∗ xJ†eµ,j U o (cid:18) = JU∗ xΛ(µ)te−µ,j x · Aµ + 1 2 = Je (cid:19) x · Ax (cid:18) Λ(µ)te−µ−x,j (cid:19) = Λ(µ + x)e −x · Aµ − 1 2 x · Ax † Λ(µ) eµ+x,j. We compute the commutator [Uy, U o x]: UyU o xeµ,j = e U o xUyeµ,j = e y · A(µ + x + 1 2 −x · A(µ + y + y) − x · A(µ + 1 2 x) + y · A(µ + 1 2 (cid:18) (cid:18) (cid:19) x) 1 2 y) (cid:19) Λ(µ + x)Λ†(µ)eµ+x+y,j Λ(µ + x + y)Λ†(µ + y)eµ+x+y,j. If the commutator vanishes, we see that by canceling common terms we must have: e (y · Ax) Λ(µ + x)Λ†(µ) = e (−x · Ay) Λ(µ + x + y)Λ†(µ + y). This has as a consequence that Λ(x)ij consists of e(fij(x)) with fij(x) a function of the form x· Bijx + φij(x) + νij. We see that the quadratic part must be the same for each component, and that Bij = A. The constant part νij can be absorbed into by unitary transformation. We can thus write Λ(x) = e(x · Ax)Λ(cid:48)(x), (17) where Λ(cid:48)(x) consists of functions cije(φij(x)) such that Λ(cid:48)(x) = J Λ(cid:48)(−x)t and Λ(cid:48)(x)Λ(cid:48)(x) = Id. If we then calculate U o x: † (cid:18) (cid:19) U o xeµ = Λ(µ + x)e −x · Aµ − 1 2 x · Ax † Λ(µ) eµ+x,j (cid:19) (µ + x) · A(µ + x) − x · Aµ − 1 2 x · Ax − µ · Aµ Λ(cid:48)(µ + x)Λ(cid:48)(µ) † eµ+x,j (cid:18) (cid:18) (cid:18) = e = e = e (cid:19) (cid:19) µ · Ax + µ · Ax + x · Ax x · Ax 1 2 1 2 Λ(cid:48)(µ + x)Λ(cid:48)(µ) † eµ+x,j † Λ(cid:48)(x)Λ(cid:48)(0) eµ+x,j. By definition (17) of Λ(cid:48), we have Λ(cid:48)(0) = Λ(0), so this is exactly equation (16). (cid:3) CLASSIFICATION OF SPIN STRUCTURES ON THE NONCOMMUTATIVE n-TORUS 11 6. Dirac operator The remaining piece of the real spectral triple is the Dirac operator. In this section, using the results from the previous section, the equivariance condition (4d), and Axioms 1 and 5, we derive conditions for the Dirac operator D. We see that for different forms of Λ as given in Lemma 2, only the form Λ(cid:48)(x)2 = Id can lead to isospectral deformations of spin structures on the commutative n-torus. By applying the compact resolvent condition of Axiom 1, we show that this is the only possibility compatible with the definition of noncommutative geometry. An equivariant Dirac operator D commutes with the basic derivations δi as described in (4d): Deµ,j = dk µ,jeµ,k. Since D should be self-adjoint, we have that dk µ,j = (18) Deµ,j = D(µ)eµ,j. (cid:88) k∈I (cid:17)∗ (cid:16) dj µ,k . We will write This means that D(µ) is the operator D restricted to the eigenspace of derivations with eigenvalue µ. This is well defined due to the equivariance condition (4d). Given an element µ in a shifted lattice Zn, we define the Hilbert space Hµ as the span of eigenvectors vj of the basic derivations δi such that δivj = µivj for each i. As a consequence of the equivariance of the Dirac operator and the irreducibility condition, all the µ must lie in the same lattice, by the following argument. Between two Hilbert spaces Hµ and Hν we have an isometry if µ − ν ∈ Zn, given by the unitary element Uµ−ν ofthe algebra. Now consider the projector Pµ that is Id on Hν for which µ − ν ∈ Zn, and 0 otherwise. This projection clearly commutes with the algebra, and, because of (18), also with the Dirac operator. If the spectral triple is irreducible, only scalars may commute with both the algebra and the Dirac operator, so Pµ is the identity on the whole Hilbert space H, hence all lattices are shifted by the same vector . Because of (11), we can then conclude that  consists of elements which are either 0 or 1 2 . From the first order condition in equation (2), we deduce: Lemma 3. An equivariant Dirac operator D that satisfies the first order condition must be of the form: (19) D(x + y) = Λ(cid:48)(x)Λ(0) (D(y) − D(0)) + D(x). (cid:16) †(cid:17)2 Proof. To check the the first order condition, it is sufficient to check it only for the unitary generators of Aθ: [[D, Ux], U o y − UxDU o y] = DUxU 0 yDUx + U o y − U o yUxD = 0, for all x, y ∈ Zn. Using Lemma 2, we write out the first order condition: † yeµ,j = a(x, y, µ)D(x + y + µ)Λ(cid:48)(y)Λ(cid:48)(0) † yeµ,j = a(x, y, µ)D(y + µ)Λ(cid:48)(y)Λ(cid:48)(0) DUxU 0 UxDU 0 † yDUxeµ,j = a(x, y, µ)Λ(cid:48)(y)Λ(cid:48)(0) U 0 † yUxDeµ,j = a(x, y, µ)Λ(cid:48)(y)Λ(cid:48)(0) U 0 D(x + µ)eµ+x+y,j D(µ)eµ+x+y,j, eµ+x+y,j eµ+x+y,j 12 JAN JITSE VENSELAAR where a(x, y, µ) is the common factor (cid:18) x · Aµ + µ · Ay + x · Ay + (cid:19) (x · Ax + y · Ay) . 1 2 a(x, y, µ) = e This gives the relation † (D(x + y + µ) − D(y + µ)) Λ(cid:48)(y)Λ(cid:48)(0) † = Λ(cid:48)(y)Λ(cid:48)(0) (D(x + µ) − D(µ)) . Since D is self-adjoint and Λ(cid:48)(x) unitary, we can rewrite this as (D(x + y + µ) − D(y + µ)) = Λ(cid:48)(y)Λ(cid:48)(0) (cid:16) †(cid:17)2 (D(x + µ) − D(µ)) . (cid:3) For y = x and µ = 0, the solution to the defining equation (19) is (cid:88) (cid:16) †(cid:17)2 (τ i · x) Ai + Λ(cid:48)(x)Λ(0) i (20) D(x) = where the Ai, B and C are bounded operators such that Ceµ,j =(cid:80) similarly for B and Ai, and ker ((cid:80) (cid:16) †(cid:17)2 (cid:54)= Id. This is the unique solution, since we see from equation (19) k cjkeµ,k and i τ i · xAi) contains at least the x ∈ Zn such that Λ(cid:48)(x)Λ(0) that D is fully determined after we choose suitable D(ei) for 1 ≤ i ≤ n. B + C, If (Λ(cid:48)(x)Λ(0) † 2 ) = Id, this gives a linear Dirac operator, familiar from commu- tative geometry. However, at first glance it seems that there might be other spin structures, not corresponding to commutative spin geometries. These other candi- 2 (cid:54)= Id, will however drop out because they are date geometries, where (Λ(cid:48)(x)Λ(0) ) incompatible with the compact resolvent condition on D. Lemma 4. Only if (Λ(cid:48)(x)Λ(0) a compact resolvent. Hence D is of the form: = Id can the equivariant Dirac operator D have † † ) 2 (cid:32)(cid:88) (cid:33) (21) Deµ,j = (τ i · µ) Ai + C eµ,j. i As a corollary, we have that J is of the form: (22) Jeµ,j = e (µ · Aµ) Λe−µ,j, with Λ a constant isometry such that Λ2 = J Id. Proof. By [29, Theorem XIII.64], we see that an unbounded self-adjoint operator D bounded away from 0 has a compact resolvent if and only if the set Fb = {ψ ∈ Dom(D) : ψ ≤ 1;Dψ ≤ b}, know that x ∈ ker ((cid:80) is compact for all b ∈ R. However, if (Λ(cid:48)(x)Λ(0) 2 (cid:54)= Id for some direction x, we i τ i · xAi) and then it follows from (20) that the norm of Deλx is bounded by B + C for all λ ∈ Z. When we take b > B + C, Fb contains at least eλx for all λ ∈ Z (cid:54)= 0, so Fb cannot be compact. (cid:3) ) † CLASSIFICATION OF SPIN STRUCTURES ON THE NONCOMMUTATIVE n-TORUS 13 7. Grading and Hochschild homology In this section we investigate what extra conditions on the the spectral triple of the noncommutative torus come from the grading operator and Hochschild con- ditions, Axioms 2 and 6. We find that the parameters τ i introduced in section 6 must be linearly independent vectors spanning Rn. We start by investigating the Hochschild cycle condition, which states that there is a Hochschild cocycle c ∈ Zn(A,A ⊗ Ao) whose representative on H is Γ when n is even, or Id if n is odd. The Γ operator is an isometry, with eigenvalues 1 and −1, and so by Axiom 2 and the diagonal action of the algebra on the Hilbert space, we see that where Γ+ and Γ− are unitary self-adjoint operators which have only eigenvalues +1 and −1 respectively. Using a unitary transformation, we can assume these operators to be diagonal, thus (cid:18)Γ+ 0 (cid:19) 0 Γ− Γ = (cid:18)Id+ , (cid:19) Γ = 0 0 −Id− , where Id+ and Id− are the identity operators on the positive and negative eigenspaces of Γ. An obvious candidate for the Hochschild cycle in Zn(A,A ⊗ Ao) is the straight- forward generalization of the unique such cycle for the noncommutative 2-torus [28, Page 324], which is (1,0)U∗ (1,0) ⊗ U(1,0) ⊗ U(0,1). c2 = U∗ (0,1)U∗ A candidate generator of the n-th Hochschild homology of the noncommutative n-torus is given by (0,1) ⊗ U(0,1) ⊗ U(1,0) − U∗ (cid:88) (cid:32) n(cid:89) (cid:32) sign(σ) Ueσ(i) (cid:33)∗ n(cid:79) (cid:1)(cid:33) , (cid:0)Ueσ(i) i=1 i=1 (23) cn = σ∈Sn with ei an orthonormal basis of Zn. It is known [22, Lemma 12.15] that this a Hochschild cycle. Due to [38, Theorem 1.1], the n-th Hochschild homology of the n-torus is 1-dimensional. Together with Lemmas 5 and 6 below, this means (23) generates the n-th Hochschild homology. Lemma 5. For the noncommutative n-torus, only nontrivial cycles can be mapped to Γ when n is even, and to Id when n is odd, by the map πD. Proof. Since the Hochschild cycle consists of polynomial expressions, it is enough to prove the result for individual homogeneous polynomials, since any cycle can be written as the sum of homogeneous polynomials. Define c(cid:48) = Ux0 ⊗ U o y ⊗ Ux1 ⊗ ··· ⊗ Uxn ∈ Zn+1(A,A ⊗ Ao), with xi, y ∈ Zn. As in [37, Chapter 3.5], we see that (24) Since [D, Ux] = CxUx with Cx some operator depending on x, πD(bc(cid:48)) is propor- tional to πD(bc(cid:48)) = (−1)n [πo (Uy) π (Ux0 ) [D, π (Ux1 )] . . . [D, π (Uxn )], π (Uxn+1)] . Cx1,...,xn π (Ux0 ) πo (Uy) π (Ux1 ) . . . π (Uxn ) π (Uxn+1) . 14 A(Tn JAN JITSE VENSELAAR θ ) and A(Tn θ )o have a trivial intersection, U o When n is even, Γ maps eµ,j to ±eµ,j, the total sum y +(cid:80)n+1 total(cid:80)n+1 i=0 xi must be 0. If y must have total degree 0. The i=0 xi is 0, and since Ux and U−x commute, the commutator (24) vanishes. When n is odd, the argument goes the same. θ ) and A(Tn θ )o have nontrivial intersection, there are y for which UxUy = UyUx for all x ∈ Zn. In that case, by the same arguments as for the trivial intersection case above, the Uy must lie entirely within the intersection of A(Tn θ ) and A(Tn i=0 xi, the commutator (24) vanishes: If A(Tn θ )o, and since xn+1 = −y −(cid:80)n (cid:2)U o (cid:32) (cid:3) = yUx1 ··· Uxn , Uxn+1 (cid:33) y [Ux1 ··· Uxn , Uxn+1 ] = n(cid:88) U o 1 − e(xn+1 · θ xi) U o y (cid:0)1 − e(xn+1 · θ(−y − xn+1))(cid:1) = 0. = i U o y Since xn+1 · θxn+1 = 0 and Uy commutes with Uxn+1. (cid:3) Lemma 6. The Dirac operator given in Lemma 4 satisfies the Hochschild condition only if the vectors τ i are linearly independent. The Dirac operator is (cid:0)τ i · δ(cid:1) Ai + C, D = (cid:88) (cid:89) i i (cid:88) σ∈Sn where the Ai are bounded operators such that sign(σ) Aσ(i) = det(τ 1τ 2 . . . τ n)Γ, (25) (26) when n is even, and det(τ 1τ 2 . . . τ n)Id when n is odd. Proof. In order to deduce the representative of the Hochschild cycle (23) on the Hilbert space H, we calculate [D, Ux]eµ,j = De(x · Aµ + 1 2 x · Ax)eµ+x,j − Ux τ k · µAkeµ,j (cid:33) (cid:32)(cid:88) k τ k · xAkeµ+x,j (cid:33) . (cid:32)(cid:88) k (27) = e(x · Aµ + x · Ax) 1 2 We suppress the e(x · Aµ + 1 by the U∗ factors, and expand (23): ei 2 x · Ax) factors, since these are canceled from the left (cid:32) (cid:33)∗ n(cid:79) (cid:1)(cid:33)(cid:33) πD(c) = πD sign(σ) (cid:32)(cid:88) (cid:32) (cid:88) (cid:32) (cid:88) σ∈Sn σ∈Sn σ∈Sn = = sign(σ) (cid:0)Ueσ(i) (cid:33) (cid:32) n(cid:89) i=1 (cid:32) n(cid:89) (cid:88) n(cid:89) i=1 k i=1 Ueσ(i) (cid:33)∗ n(cid:89) i=1 τ k · eσ(i)Ak i=1 (cid:33) . sign(σ) Ueσ(i) [D, Ueσ(i) ] CLASSIFICATION OF SPIN STRUCTURES ON THE NONCOMMUTATIVE n-TORUS 15 This expression should be some constant times Γ or Id, depending on the dimension. Due to [10, Proposition 4.2], we can write this as (cid:88) σ∈Sn (cid:89) i πD(c) = det(τ 1τ 2 . . . τ n) sign(σ) Aσ(i), where we view (τ 1τ 2 . . . τ n) an n × n matrix with τ i as the columns, from which (cid:3) it is immediately clear that the vectors must be linearly independent. 8. Dimension, finiteness and regularity Here we establish, using the conditions of dimension, regularity and finiteness (Axioms 4, 7 and 8), that the real spectral triple must be an isospectral deformation of a spin structure on a noncommutative torus. The result follows from Connes' spin manifold theorem. In the course of proving this, we find a proof for an elementary fact about Hermitian matrices generating a Clifford algebra, for which we do not know an elementary proof. Lemma 7. If the τ i span Rn, for arbitrary ai ∈ R and all  > 0, there is a t ∈ R and a set of µj ∈ Zn, with Zn a shifted lattice as in section 5 such that (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)tai − τ i ·(cid:88) j (cid:88) i (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) < . µj Proof. Every vector p ∈ Zn can be written as a sum of at most 2 vectors µj ∈ Zn. By Dirichlet's theorem of simultaneous Diophantine approximation [33, Theorem i) ∈ Rn we can find integers q, p1, . . . pn with q < N , II.1B], for all N > 1 and (a(cid:48) such that qa(cid:48) i − pi < N−1/n, for 1 ≤ i ≤ n, and ai ∈ R. on the τ i such that (Rp)i = τ i · p. Set a(cid:48) with the maximal absolute value τmax, then clearly we have Since the τ i span Rn, there is a transformation R ∈ GL(R, n) depending only i = (R−1a)i. Call the eigenvalue of R µj = q(Ra(cid:48))i − (Rp)i ≤ τmaxqa(cid:48) i − pi < τmaxN−1/n. qai − τ i ·(cid:88) j Choosing N such that nτmaxN−1/n < , we get the asked result. Lemma 8. In order for D to be satisfy the compact resolvent condition of Axiom 1, (cid:3) and the dimension condition of Axiom 4,(cid:88) xiAi, i needs to be invertible for all xi ∈ R except when all xi vanish. In particular, all Ai should be invertible operators. i xiAi is not invertible at a point where xi = τ i · µ for all i, then clearly this is also the case for µ(cid:48) = λµ with λ ∈ Z. Thus the kernel cannot be finite dimensional in this case. i xiAi is not invertible for some non-trivial xi, but xi (cid:54)= τ i · µ for all µ ∈ Zn Proof. If(cid:80) If(cid:80) we have the following. 16 JAN JITSE VENSELAAR of D−1 is unbounded, hence D−1 cannot be compact. i (cid:80) of vectors µj ∈ Zn such that(cid:80) kernel of(cid:80) i xiAi. Then ((cid:80) eigenvalue of(cid:80) Since the τ i span Rn by Lemma 6, by Lemma 7 we have for every  > 0 a set j τ i · µj − xi < . Take an element y in the i,j τ i · µjAi) · y < y. Hence there is at least one i,j τ i· µjAi smaller or equal to . But this means that the spectrum (cid:3) We have so far assumed nothing about the size of the Hilbert space H compared to the basic irreducible representation of the algebra, H0 as defined in equation (6). By construction, H is a left A(Tn θ )-module. According to Axiom 8, a certain sub- module H∞ of H should be a finitely generated projective left A(Tn θ )-module. This has the following consequence: Lemma 9. The spectral triple (A,H, D) only satisfies the finiteness condition of Axiom 8 if the Hilbert space H is a finite direct sum of copies of H0. If we assume the algebra A to be closed under the collection of seminorms δk(·) defined in Axiom 7, it is given by (cid:41) (28) A(Tn θ ) = a(x)Ux a(x) ∈ S(Zn) , (cid:40)(cid:88) x∈Zn (cid:88) (cid:88) x∈Zn x∈Zn where S(Zn) is the set of Schwartz functions over Zn: (cid:26) S(Zn) = a : Zn → C sup x∈Zn (1 + x2) ka(x)2 < ∞ for all k ∈ N (cid:27) . We can write δk(a)2 =(cid:80) mas 6 and 8, we know that(cid:80) Proof. By the arguments in [7, Lemma III.6.α.2], we know that the intersection of the domain of the δk is stable under the holomorphic functional calculus. The collection of seminorms δk(·) is equivalent to the collection of seminorms qk = supx∈Zn (1 + x2)ka(x) because of the following argument. (cid:80) x∈Zn Ai2kτ i · x2ka(x)2. Because of Lem- i Ai2kτ i · x2k can be bounded between c1,kx2k and c2,kx2k for constants c1,k, c2,k > 0. Thus the family of seminorms δk(a) x∈Zn x2ka(x)2. Now if an ele- is equivalent to the family of seminorms q(cid:48) k, then clearly supx∈Zn (1 + x2)ka(x)2 < ment a of the C∗-algebra A has a finite norm in all q(cid:48) ∞ for all k ∈ N. Conversely, if a ∈ A(Tn x2ka(x)2 k(a) =(cid:80) θ ), we can write i q(cid:48) k(a) = = (1 + x2) p (1 + x2) −px2la(x)2 and this last sum converges if p is big enough, since (1 + x2) tion less than some finite constant cp and (cid:80) pa(x)2 is by assump- −p converges when p > n/2 + k. By Axiom 8, we have that H∞ is a finitely generated projective left A(Tn θ ) module. We already knew that H must be a direct sum of copies of H0 due to the discussion following (12). The finiteness condition then ensures that the sum must be finite. All conditions stated in Axioms 7 and 8 are then easily seen to be (cid:3) fulfilled. x∈Zn cpx2k(1 + x) CLASSIFICATION OF SPIN STRUCTURES ON THE NONCOMMUTATIVE n-TORUS 17 Remark 1. The choice of smooth structure is not unique even in the commutative case, given just the algebra C∗-algebra A of the noncommutative torus. See for example [23]. However, given the equivariance condition of the Dirac operator, and the assumption that the smooth algebra consists of all elements such that δk(·) < ∞ for all k ∈ N, the algebra is uniquely determined by the above argument. The number of generators of H in terms of H0 is still undetermined, but a lower bound is given by [2, Theorem 1]. This theorem states that for complex Hermitian k × k matrices, with k = (2a + 1)2b, there exists at most 2b + 1 matrices satisfying the non-invertibility property of Lemma 8. So in order to have at least n such matrices, k should be at least 2(cid:98)n/2(cid:99). Hermitian matrices generating an irreducible representation of a Clifford algebra Cln,0 are an example of a set of matrices attaining this lower bound. matrices Ai is a Clifford algebra. What remains to be shown is that ((cid:80) Remark 2. It does not follow from Lemma 8 that the algebra generated by the i xiAi)2 lies in the center of the bounded operators on H for all x ∈ Rn. Only the condition of Lemma 8 is not enough to show this. Consider a set of self-adjoint matrices {Bi} generating a Clifford algebra. These satisfy the invertibility condition of Lemma 8, and the Hochschild condition (26). Since they are self-adjoint, we can diagonalize B1 as B1 = U ΛU† with Λ a diagonal matrix with real elements. If we rescale the elements of Λ each by a different nonzero amount to Λ(cid:48), the set {U Λ(cid:48)U†}∪{Bi}i≥2 still has the invertibility property, but not necessarily the Hochschild property, as a calculation for any case n ≥ 4 will show. If n = 2, 3, the invertibility property does imply the Hochschild property however. However, a weaker form of Connes' reconstruction theorem [10, Theorem 11.5], implies the following result: θ ),H, D, J) to satisfy both Lemma 10. In order for the candidate structure (A(Tn the Hochschild condition of Axiom 6 and the dimension condition of Axiom 4, the matrices Ai of Lemma 4 must generate a Clifford algebra. θ ), H, D and J, we see that none of them Proof. If we look at our conditions on A(Tn depend on the antisymmetric matrix θ. Also, the action of the Dirac operator on the Hilbert space is independent of θ. This means that we can just set θ = 0, where we have the real spectral triple (A0,H, D, J) of smooth functions on the n-torus. Due to the results of Connes' spin manifold theorem (see for example [22, Lemma 11.6], [10, Remark 5.12]), this implies that the Ai generate a Clifford algebra. (cid:3) Because the size of the maximal set of matrices which satisfy the invertibility condition of Lemma 8 is odd, due to [2, Theorem 1], we have as a corollary: Corollary 1. A set of 2b × 2b Hermitian matrices {Ai}n that the equation i=1, where n = 2b + 1, such only has the zero solution (xi = 0)n if i=1 in Rn generate a Clifford algebra if and only (cid:33) (cid:32)(cid:88) n(cid:89) i i (cid:88) σ∈Sn det xiAi = 0, sign(σ) Aσ(i) = λIdk, 18 JAN JITSE VENSELAAR for some nonzero λ ∈ R. If we assume the spectral triple to be irreducible, we need that the matrices generate an irreducible representation of the Clifford algebra. This restricts the size of the matrices to be exactly 2(cid:98)n/2(cid:99). From the results in sections 4, 5 and 6 it now follows that the remaining condi- tions, the Poincar´e duality of Axiom 9 and the Hermitian pairing of Axiom 8 are also satisfied, since they are satisfied by isospectral deformations. This completes the proof of Theorem A: Theorem A. All irreducible real spectral triples with an equivariant n-torus actions are isospectral deformations of spin structures on an n-torus. 9. Description of the real spectral triples Finally, we show that given a Dirac operator D that satisfies all conditions so far, the reality operator J is uniquely determined, and list all the ingredients that constitute all real spectral triples of the noncommutative n-torus. Also, we show in some low dimensional cases what freedom there still exist in the definition of the Dirac operator. Recall that the Clifford algebra Cln,0 is the algebra over R generated by Rn, 1 and a positive definite quadratic form q, subject to the relation v·v = −q(v)Id. The Clifford group is defined as the group generated by the image of an orthonormal basis of Rn together with −1. (cid:0)τ j · δ(cid:1) Aj with Aj representatives of the Clifford Lemma 11. If D is given by(cid:80) group of Cln,0 there is a unique J operator for each n, up to multiplication with a complex number of norm 1. j Proof. If d = 1, 2, 3 or 4 this can be done for example by calculations, see Remark 3 below. We proceed by induction. First we prove existence. Recall that there are isomorphisms Cln+2,0 (cid:39) Cl0,n ⊗ Cl2,0 and Cl0,n+2 (cid:39) Cln,0 ⊗ Cl0,2 [26, Theorem I.4.1]. Let d > 4, and assume it has been proven for n − 4. The operator Jn = ∼= Cln−4,0 ⊗ Cl4,0 has precisely the right commutation Jn−4 ⊗ J4, acting on Cln,0 relations, except for n ≡ 1 mod 4, as can easily be calculated by looking at Table 1, and taking into account the periodicity mod 4 of the table, except for the first row, 4 = −1. In case n ≡ 1 mod 4, we can achieve the same by setting where we use J 2 Jn = Jn−4 ⊗ Γ4J4. n for the representation of the i-th basis vector of Rn in Cln,0. An explicit isomorphism Cln,0 ≡ Cln−4,0 ⊗ Cl4,0 can be chosen, for example Now we prove the uniqueness. Write γi n = Idn−4 ⊗ γi γi n and γ3 4 for i ≤ 4, γi n = γi−4 n commute with γi nγ4 n−4 ⊗ γ1 4 γ2 4 γ3 4 γ4 4 for i > 4. The operators γ1 γ1 n, γ2 n, γ4 n and γ3 nγ2 n respectively. They square to −1, so the operator n for i > 4 and anticommute with (cid:0)1 + iγ1 nγ2 n (cid:1)(cid:0)1 + iγ3 nγ4 n (cid:1) , P + := 1 4 n for i > 4. The projection P + does not com- is a projection, which commutes with γi mute with Jn, but does commute with γ2γ3Jn. Also, the projection P + projects onto a subspace of dimension 1/4 times the dimension of the irreducible repre- sentation of Cln,0, and the operator γ2γ3Jn has the same commutation relations CLASSIFICATION OF SPIN STRUCTURES ON THE NONCOMMUTATIVE n-TORUS 19 n for i > 4 as Jn, and since (γ2γ3Jn)2 = −J 2 with γi n it has the right signs for an (n − 4)-dimensional J operator. This means that P + projects onto a Hilbert space belonging to an (n−4)-dimensional spectral triple, where we have a unique J opera- tor by the induction hypothesis. On the complement of the P + eigenspace, we have the unique J4 operator with the right commutation relations with γi, i ≤ 4. (cid:3) Stated more elaborately, we have the following result: Theorem C. The following give all 2n irreducible real spectral triples on the smooth noncommutative n-torus A(Tn θ ): • A Hilbert space H constructed as follows: (cid:77) m∈Zn+ 2}. with  = (1, . . . , n) ∈ Rn, and i ∈ {0, 1 2(cid:98)n/2(cid:99)(cid:77) Hi Hi = H = i C, • An involutive algebra A with unitary generators Ux with x ∈ Zn A(Tn θ ) := {A = a(x)Ux : a ∈ S(Zn)}, with Ux acting on a basis vector eµ,i ∈ Hi by x · Ax + x · Aµ Uxeµ,i = e eµ+x,i, (cid:19) (29a) (29b) (29c) (cid:88) x (cid:18) 1 2 n(cid:88) for any matrix A such that A − At = θ. • An unbounded, densely defined, self-adjoint first order operator D (29d) D = (τ j · δ)Aj + C, j=1 acting on H with C a bounded self-adjoint operator commuting with the algebra satisfying JCJ−1 = DC, and ΓC = −CΓ if n is even, for τ j n linearly independent vectors in Rn, n matrices Aj of size 2(cid:98)n/2(cid:99) × 2(cid:98)n/2(cid:99) generating an irreducible representation of the Clifford group Cln,0, and δ the derivations δieµ = µieµ. • If n is even, the grading operator Γ is given by (cid:80) i Aσ(i), • A unique (up to multiplication with a complex number of modulus 1) anti- sign(σ)(cid:81) with Aj the matrices given above. σ∈Sn linear isometry J that acts as (29e) Jeµ,j = e(µ · Aµ)Λe−µ,j. with Λ a bounded linear operator such that ΛΛ† = Id and DΛ = −DΛD∗. In lower dimensional cases we can explicitly calculate what form C in (29d) can take. In [28, Lemma 2.3, Theorem 2.5] it is proven that C = 0 if n = 2. For n = 3 we have the following: Proposition 1. If n = 3 the constant matrix C must have the form qId where q ∈ R arbitrary. If n = 4, the matrix C must have the form 0 0 a 0 −¯b 0 ¯a −b 0 ¯b 0 a  , b ¯a 0 0 20 where a, b ∈ C. JAN JITSE VENSELAAR Proof. If n = 3 it follows from Theorem C that (cid:88) k∈I J = e(µ · Aµ) ajke−µ,k, (cid:18)0 −1 (cid:19) 1 0 , J = e(µ · Aµ) with I = {1, 2}. We choose a particular form of D, given by the particular repre- sentation of the Clifford group Cl3,0 known as the Pauli matrices. We see that just as in the n = 2 case in [28, Theorem 2.5]. Using the appropriate values for J and D found in Table 1, we see that the defining equation is JCJ = C and by a calculation this shows that C = qId where q ∈ R arbitrary. Similarly for n = 4, we can just check what the conditions are for C to satisfy the equations ΓD = −DΓ and JC = CJ, and this gives the possibilities given in the proposition. It is trivial to calculate similar conditions for higher dimensions. Due to the increase in the size of matrices, and relaxation of the commutation relation with Γ when going from (cid:3) n = 2k to n = 2k + 1, the number of parameters will increase when n grows. Remark 3. If we choose a representation for Cl1,0, Cl2,0 and Cl0,2 all Clifford algebras Cln,0 can be constructed by the basic isomorphisms [26, Theorem I.4.1]: Cln,0 ⊗ Cl0,2 Cl0,n ⊗ Cl2,0 ∼= Cl0,n+2, ∼= Cln+2,0. We choose a representation: the (unique up to multiplication with a complex number of norm 1) matrix part of the J operator can easily be calculated: Cl1,0 = 1, Cl2,0 = Cl0,2 = (cid:19) (cid:19) , , 0 0 −1 0 0 −i , , 0 (cid:19) (cid:18) 0 (cid:18)1 (cid:18)0 −1 (cid:19) (cid:18)i i−i (cid:18) 0 (cid:19) 1−1 0 1 0 J2 = J3 =  0 0 −1 0 0 1 0 0 0 0 0 0 −1 0 0 1 0 0 −1 0 0 0 0 1 0 0 −1 0 0 1 0 0   . J4 = J5 = (30a) (30b) (30c) In this representation, we also see that for all n, the matrix component of J has precisely one nonzero element in every column or row. CLASSIFICATION OF SPIN STRUCTURES ON THE NONCOMMUTATIVE n-TORUS 21 10. Unitary equivalences After establishing that all real spectral triples on the noncommutative n-torus are isospectral deformations of spin structures on the commutative n-torus, we show that there is a big difference between their groups of symmetries. In the commutative case, a diffeomorphism acting from the torus can transform a spin structure into another according to the action of the diffeomorphism group as cal- culated in [15]. Here we show that in the noncommutative case when n = 2 most spin structures, except the isospectral deformation of the trivial spin structure, are equivalent. When n > 2 the result is less conclusive, due to insufficient knowledge of the full automorphism group of the C∗-algebra in that case. Recall the definition of a unitary equivalence between two spectral triples: Definition (Unitary equivalence). A unitary equivalence between two spectral triples (A, H, D, J, (Γ)) and (A, H, D(cid:48), J(cid:48), (Γ(cid:48))) is given by a unitary operator W acting on the Hilbert space H such that (31a) (31b) (31c) W π(a)W −1 = π(σ(a)) ∀a ∈ A, W DW −1 = D(cid:48), W JW −1 = J(cid:48), W ΓW −1 = Γ(cid:48), (31d) where σ is a ∗-automorphism of the C∗-algebra A such that the algebra A is mapped into itself. We first recall what is known about the automorphisms of the C∗-algebra A(Tn θ ). The main tool for understanding the automorphism group of a general noncommu- tative n-torus is [5, Theorem I], which tells us that for θ in a set which has full measure in the space of all antisymmetric matrices, the algebra A(Tn θ ) is an induc- tive limit of direct sums of circle algebras. This results allows one to generalize a lot of results on noncommutative two-tori to higher dimensional tori. For a noncommutative n-torus which is an inductive limit of direct sums of circle algebras, the automorphism group fits in the following exact sequence [21, Theorem 2.1]: 1 → Inn(A(Tn θ )) → Aut(A(Tn θ )) → Aut(K(A(Tn θ ))) → 1, where an automorphism of K0(A(Tn [1A(Tn θ )] and the order structure (K0(A(Tn θ ))⊕ K1(A(Tn θ )) ⊕ K1(A(Tn θ )))+. θ )) should preserve the order unit θ )) = Inn0(A(Tn The left side of the exact sequence can be further specified by [21, Corollary 4.6], which states that for algebras which are inductive limits of direct sums of circle algebras Inn(A(Tn θ )). This means that an inner automorphism of a noncommutative torus can only give a unitary equivalence of two spin structures if the identity automorphism gives a unitary equivalence between the two spin structures. By [30, Theorem 6.1] the order structure (K0(A(Tn θ )))+ consists precisely of those elements for which the normalized trace is positive, and by [18, Theorem 3.1], the image of this trace on K0 is equal to the range of the exterior exponential of θ: (cid:94) exp θ = 1 ⊕ θ ⊕ 1 2 (θ ∧ θ) ⊕ . . . : even(cid:94) Zn → R. 22 JAN JITSE VENSELAAR θ)) must be the identity on K0(A(T2 For noncommutative 2-tori with θ irrational, this means that all automorphisms θ)) = Z + θZ. The K1 groups for of K(A(T2 noncommutative n-tori are also known to be isomorphic to Z2n−1 . For n = 2 a partial lifting of Aut(Z2) = GL(2, Z) is known [6], and given by the action of SL(2, Z) on the lattice Z2 of unitary generators Ux. In [19] it is proven that in fact the whole automorphism group of the algebra A(T2 θ) for irrational θ with certain extra Diophantine conditions is given by a semidirect product of this action, the (cid:55)→ λiUei and the projectivized group of unitaries of canonical torus action: Uei A(T2 θ) in the connected component of the identity: P U (A(T2 θ)) 0 (cid:111) (T2 (cid:111) SL(2, Z)). For n > 2 the situation is less clear, since the action on the K0 group need not be trivial anymore. When n > 2, if θ is a matrix such that all θij are independent over Z, one can easily calculate that an outer automorphism cannot simply map basic unitaries to other basic unitaries. An automorphism σ of this form must satisfy e(σ(x) · θσ(y)) = e(x · θy) for all x, y ∈ Zn. If the θij are independent over Z, we ∈ SL(2, Z). see that σ(x) = α11x + α12y and σ(y) = α21x + α22y with When n > 2 the only solution for all x and y is σ(x) = ±x with sign the same for all x. (cid:18)α11 α12 α21 α22 (cid:19) With this knowledge of the automorphism group, we can proceed to our main theorem of this section. Theorem B. Except for a set of θ of measure 0, the different spin structures of the smooth noncommutative n-torus A(Tn θ ) cannot be unitarily equivalent by an inner automorphism of the algebra. In the case n = 2 the theorem was proven in [28, Theorem 2.5]. Proof. While for the proof of Theorem B it is only necessary to consider inner automorphisms, we get Corollary 2 if we also consider automorphisms induced by an action of SL(2, Z) on the algebra. θ )) = Inn0(A(Tn By [21, Corollary 4.6] and [5, Theorem I], we have that for almost all noncom- mutative tori Inn(A(Tn θ )). This means that if an inner automor- phism changes the spin structure, than so does the identity automorphism. We will assume in the following that the components of θ in the upper right corner are independent over Z. The set of θ all of whose components in the upper right corner are independent over Z is of full measure, and so is the set of θ for which Inn(A(Tn θ )), so their intersection also has full measure. θ )) = Inn0(A(Tn We label the basis of the Hilbert space for the different spin structures by the same labels m, so µ = m + . We see that for a spin structure  the operator J, written as in equation (22), acts as Jem,i = Λije ((m + ) · A(m + )) e−m+2,j. We consider a unitary transformation W , induced by an automorphism σ ∈ (cid:88) (cid:88) k j SL(2, Z), and denote W e0,i = wk,ijek,j, CLASSIFICATION OF SPIN STRUCTURES ON THE NONCOMMUTATIVE n-TORUS 23 for the action on the e0,i, and write σ(x) for the obvious action of the automorphism σ (either an element of SL(2, Z) when n = 2, or the action x (cid:55)→ ±x when n > 2) on the vector x ∈ Zn. We first show that the action of the unitary transformation W on the Hilbert space is fully determined by the action on the different e0,i. Next we show that the condition (31c) implies that the action of the unitary transformation on the basis vectors must be such that a basis vector e0,i is mapped to a linear combination of vectors ek,l with k =  − σ(), where  is the original spin structure and  the new spin structure. If σ() = ±, this means that the spin structure is unchanged, since k must lie in Zn. For the noncommutative 2-torus, we have σ ∈ SL(2, Z), and we see that if  ∈ Z2, the spin structure is unchanged. All spin structures on the noncommutative 2-torus for which  /∈ Z2 are unitarily equivalent. ex,i = Uxe(−x · Ax/2 − x · A)e0,i and using (31a) we can write Consider a unitary transformation W that maps a spin structure  to . Since W em,i = W Ume(−m · Am/2 − m · A)e0,i = Uσ(m)e(−m · Am/2 − m · A) (cid:88)  wk,ijek,j (cid:18) (cid:88) k,j = e k,j σ(m) · A(k + ) + 1 2 σ(m) · Aσ(m) − m · Am/2 − m · A (cid:19) wk,ijek+σ(m),j. So the action of W on the Hilbert space is fully determined by the action on the e0,i. Requirement (31c) gives the following equations: W Jem,i = Λije ((m + ) · A(m + )) W e−m+2,j (cid:18) Λije (m + ) · A(m + ) + σ(−m + 2) · A(k +  + (cid:19) σ(−m + 2)) 1 2 (cid:19) −(2 − m) · A( + Λije ( · A(−3 + 2m) + σ() · A(2( + k) + σ(2 − m))) wk,jlek+σ(−m+2),l (2 − m)) 1 2 (cid:19) 1 2 m) + σ(m) · A(− − k + σ( m · A(3 + e (σ(m) · A(k +  + σ(m)/2) − m · A( + m/2)) wk,ijek+σ(m),j m − )) 1 2 wk,jlek+σ(−m+2),l = · e = · e JW em,i = k,l k,l (cid:88) (cid:18) (cid:88) (cid:18) (cid:88) (cid:88) (cid:88) (cid:18) 1 k,j k,j 2 k,j Λjle ((k + σ(m) + ) · A(k + σ(m) + )) = · e (−σ(m) · A(k +  + σ(m)/2) + m · A( + m/2)) w∗ = Λjle (( + k) · A( + k + σ(m)) + m · A( + m/2)) k,ije−k−σ(m)+2,j (cid:19) · e σ(m) · Aσ(m) w∗ k,ije−k−σ(m)+2,j. 24 JAN JITSE VENSELAAR Collecting the vectors with the same indices and in the same Hilbert space Hj we see that for indices k + σ(−m + 2) = −k(cid:48) − σ(m) + 2, or k(cid:48) = −k + 2 − 2σ(): (cid:19) (cid:19) (cid:18) that (cid:18) σ(−m + 2) · A(k +  + e σ(−m + 2)) − (2 − m) · A( + 1 2 1 2 (2 − m)) wk,jl = Λjle Since Λ−1 = Λ† and (cid:80) ( + k) · A( + k + σ(m)) + m · A( + m/2) + w∗ k,ij. ji = ±δil with δil the Kronecker delta, this implies σ(m) · Aσ(m) 1 2 j ΛljΛ∗ wk,ij = ± e (2 · A(3 − 2m) + 3 · A(3 − k + σ(−2 + m))) · e (k · A(−3 + k + σ(2 − m)) + σ() · A(−4 + 4k + 3σ(2 − m))) · e (σ(m) · A( + k + σ()) − 2m · A) w∗ (32) Applying this same formula again for w∗ −k+2−2σ(),ij = e ( · A(3 − 2m) − ( + k) · A( + k + σ(m)) − m · A(4 + m)) w∗ (33) · e (σ() · A(6 − 2k − σ(2 − m)) + σ(m) · A(3 − k − σ( + m))) wk,jk. −k+2−2σ(),ij, we get −k+2−2σ(),ij. Filling in the expression of (33) in (32), we see wk,jk = e (2 · A(4 − 2k + σ(m − 3)) + 2k · A(σ( − m) − 2)) · e (2σ() · A(3k − 5 + σ(4 − m)) + 2σ(m) · A(k −  + σ())) wk,jk. Collecting all terms which contain m we see that these add up to 2σ(m) · θ(k + σ() − ). Since the equality above should hold for all m, this means that either wk,ij = 0, or m · θ(k + σ() − ) = 0 for all m. If all components of θ in the upper right corner are independent over Z this can only be the case if k + σ()−  = 0, hence k = − σ(). Since k must lie in Zn, we see that if σ() = ± the spin structure cannot change. (cid:3) From the proof it also follows: Corollary 2. Let σ ∈ SL(2, Z). The automorphism σ of Z2 induces an automor- phism of the noncommutative 2-torus of the form Ux (cid:55)→ Uσ(x). This automorphism induces a unitary equivalence of real spectral triples which maps a spin structure  to  = σ(). In particular, real spectral triples on the noncommutative 2-torus which are not isospectral deformations of the trivial spin structure on the commutative 2-torus are unitary equivalent to each other, via the following unitary map W : W eµ = eσ(µ), W UxW −1 = Uσ(x), (cid:88) D(cid:48) = σ−1(τ j) · δ  ⊗ Aj, j J(cid:48) = J, composed with an additional unitary map given by Lemma 1 that maps the repre- sentation πA(cid:48) with A(cid:48) = σtAσ to the original πA. CLASSIFICATION OF SPIN STRUCTURES ON THE NONCOMMUTATIVE n-TORUS 25 (cid:19) (cid:18) 1 Proof. The statement about the spin structures follows from the statements at the end of the proof of Theorem B. Since σ ∈ SL(2, Z) is an automorphism of the lattice Z2, it cannot map an  /∈ Z2 to one in Z2 and vice-versa. On the other hand, set 10 = in SL(2, Z). We see (cid:18) 1 (cid:19) 0−1 (cid:18)1 −1 (cid:18)0 , and M = (cid:18) 1 , 11 = , 01 = (cid:19) (cid:19) (cid:19) , N = 1 0 1 2 0 1 2 2 1 2 11 M−1 N M NM−1 N−1 10 01 M N−1 where the matrices M and N act by left multiplication, so there exist σ ∈ SL(2, Z) such that the  /∈ Z2 are mapped to one another. Given a unitary operator W eµ = eσ(µ), we can calculate its action on the Dirac operator by using the definition (31b): D(cid:48)eµ,k = W DW −1eµ,k = W Deσ−1(µ),k j = W (cid:88) (cid:0)τ j · (σ−1(µ))(cid:1) ⊗ Ajeσ−1(µ),k (cid:88) (cid:0)σ−1(τ j) · µ(cid:1) ⊗ Ajeµ,k. (cid:18) 1 (cid:19) (cid:18) 0 (cid:19) = j The action of the automorphisms σ ∈ SL(2, Z) on the Dirac operator when  = 0 was also determined for  = 0 in [37, Section 7.1]. with the change in notation that (cid:3) our τ 1 is given there by and our τ 2 is there . Imτ Reτ Some questions left unanswered by these results are the effects of the outer auto- morphisms for n > 2 and Morita equivalences of noncommutative tori, as described in [32] and [20], on the spin structure. Also, the definition of an equivariant spec- tral triple can be generalized to allow [D, h] (cid:54)= 0, but bounded. This was used to construct equivariant Uq(SU (2)) spectral triples in [14]. It would be interesting to investigate what possibilities this would open up for the noncommutative torus. We hope to return to these questions in the future. References [1] J. Frank Adams, Vector fields on spheres, Ann. of Math. (2) 75 (1962), 603 -- 632. MR0139178 (25 #2614) [2] J. Frank Adams, Peter D. Lax, and Ralph S. Phillips, On matrices whose real linear com- binations are non-singular, Proc. Amer. Math. Soc. 16 (1965), 318 -- 322. MR0179183 (31 #3432) [3] Hajime Aoki, Jun Nishimura, and Yoshiaki Susaki, The index of the overlap Dirac operator on a discretized 2d non-commutative torus, Journal of High Energy Physics 2007 (2007), no. 02, 033. [4] Jean Bellissard, Andreas van Elst, and Hermann Schulz-Baldes, The noncommutative geom- etry of the quantum Hall effect, Journal of Mathematical Physics 35 (1994), no. 10, 5373 -- 5451. x x   9 9 - - m m [ [ 26 JAN JITSE VENSELAAR [5] Florin P. Boca, The structure of higher-dimensional noncommutative tori and metric Diophantine approximation, J. Reine Angew. Math. 492 (1997), 179 -- 219. MR1488068 (98k:46096) [6] Berndt A. Brenken, Representations and automorphisms of the irrational rotation algebra, Pacific J. Math. 111 (1984), no. 2, 257 -- 282. MR734854 (86a:46089) [7] Alain Connes, Noncommutative geometry, Academic Press Inc., San Diego, CA, 1994. MR1303779 (95j:46063) [8] [9] , Noncommutative geometry and reality, J. Math. Phys. 36 (1995), no. 11, 6194 -- 6231. MR1355905 (96g:58014) , Gravity coupled with matter and the foundation of non-commutative geometry, Comm. Math. Phys. 182 (1996), no. 1, 155 -- 176. MR1441908 (98f:58024) [10] , On the spectral characterization of manifolds (October 2008), available at math.OA: 0810.2088. http://arxiv.org/abs/0810.2088. [11] Alain Connes and Michel Dubois-Violette, Noncommutative finite-dimensional manifolds. I. Spherical manifolds and related examples, Comm. Math. Phys. 230 (2002), no. 3, 539 -- 579. MR1937657 (2004a:58006) [12] Alain Connes and Matilde Marcolli, Noncommutative geometry, quantum fields and motives, American Mathematical Society Colloquium Publications, vol. 55, American Mathematical Society, Providence, RI, 2008. MR2371808 (2009b:58015) [13] Ludwik D¸abrowski, Spinors and theta deformations, Russ. J. Math. Phys. 16 (2009), no. 3, 404 -- 408. MR2551887 (2011b:58060) [14] Ludwik D¸abrowski, Francesco D'Andrea, Giovanni Landi, and Elmar Wagner, Dirac operators on all Podle´s quantum spheres, J. Noncommut. Geom. 1 (2007), no. 2, 213 -- 239. MR2308305 (2008h:58051) [15] Ludwik D¸abrowski and Roberto Percacci, Spinors and diffeomorphisms, Comm. Math. Phys. 106 (1986), no. 4, 691 -- 704. MR860317 (88d:81054) [16] Shaun Disney, George A. Elliott, Alexander Kumjian, and Iain Raeburn, On the classification of noncommutative tori, C. R. Math. Rep. Acad. Sci. Canada 7 (1985), no. 2, 137 -- 141. MR781813 (86j:46064a) [17] Edward G. Effros and Frank Hahn, Locally compact transformation groups and C∗- algebras, Memoirs of the American Mathematical Society, No. 75, American Mathematical Society, Providence, R.I., 1967. MR0227310 (37 #2895) [18] George A. Elliott, On the K-theory of the C∗-algebra generated by a projective representation of a torsion-free discrete abelian group, Operator algebras and group representations, Vol. I (Neptun, 1980), 1984, pp. 157 -- 184. MR731772 (85m:46067) , The diffeomorphism group of the irrational rotation C∗-algebra, C. R. Math. Rep. [19] Acad. Sci. Canada 8 (1986), no. 5, 329 -- 334. MR859436 (87m:46122) [20] George A. Elliott and Hanfeng Li, Morita equivalence of smooth noncommutative tori, Acta [21] George A. Elliott and Mikael Rørdam, The automorphism group of the irrational rotation Math. 199 (2007), no. 1, 1 -- 27. MR2350069 (2008k:58023) C∗-algebra, Comm. Math. Phys. 155 (1993), no. 1, 3 -- 26. MR1228523 (94j:46059) [22] Jos´e M. Gracia-Bond´ıa, Joseph C. V´arilly, and H´ector Figueroa, Elements of noncommutative geometry, Birkhauser Advanced Texts: Basel Textbooks, Birkhauser Boston Inc., Boston, MA, 2001. MR1789831 (2001h:58038) [23] Wu Chung Hsiang and C. Terry C. Wall, On homotopy tori. II, Bull. London Math. Soc. 1 (1969), 341 -- 342. MR0258044 (41 #2691) [24] Christian Kassel, Quantum groups, Graduate Texts in Mathematics, vol. 155, Springer-Verlag, New York, 1995. MR1321145 (96e:17041) [25] Giovanni Landi, Fedele Lizzi, and Richard J. Szabo, String geometry and the noncommutative torus, Comm. Math. Phys. 206 (1999), no. 3, 603 -- 637. MR1721895 (2001c:81212) [26] H. Blaine Lawson Jr. and Marie-Louise Michelsohn, Spin geometry, Princeton Mathematical Series, vol. 38, Princeton University Press, Princeton, NJ, 1989. MR1031992 (91g:53001) [27] John W. Milnor, Spin structures on manifolds, Enseignement Math. (2) 9 (1963), 198 -- 203. MR0157388 (28 #622) [28] Mario Paschke and Andrzej Sitarz, On Spin structures and Dirac operators on the noncom- mutative torus, Lett. Math. Phys. 77 (2006), no. 3, 317 -- 327. MR2260377 (2007e:58010) CLASSIFICATION OF SPIN STRUCTURES ON THE NONCOMMUTATIVE n-TORUS 27 [29] Michael Reed and Barry Simon, Methods of modern mathematical physics. IV. Analysis of operators, Academic Press [Harcourt Brace Jovanovich Publishers], New York, 1978. MR0493421 (58 #12429c) [30] Marc A. Rieffel, Projective modules over higher-dimensional noncommutative tori, Canadian Journal of Mathematics. Journal Canadien de Math´ematiques 40 (1988), no. 2, 257338. [31] , Noncommutative tori -- a case study of noncommutative differentiable manifolds, Geometric and topological invariants of elliptic operators (Brunswick, ME, 1988), 1990, pp. 191 -- 211. MR1047281 (91d:58012) [32] Marc A. Rieffel and Albert Schwarz, Morita equivalence of multidimensional noncommutative tori, Internat. J. Math. 10 (1999), no. 2, 289 -- 299. MR1687145 (2000c:46135) [33] Wolfgang M. Schmidt, Diophantine approximations and Diophantine equations, Lecture Notes in Mathematics, vol. 1467, Springer-Verlag, Berlin, 1991. MR1176315 (94f:11059) [34] Andrzej Sitarz, Equivariant spectral triples, Noncommutative geometry and quantum groups (Warsaw, 2001), 2003, pp. 231 -- 263. MR2024433 (2005g:58058) [35] Masamichi Takesaki, Tomita's theory of modular Hilbert algebras and its applications, Lecture Notes in Mathematics, Vol. 128, Springer-Verlag, Berlin, 1970. MR0270168 (42 #5061) [36] , Theory of operator algebras. I, Springer-Verlag, New York, 1979. MR548728 (81e:46038) [37] Joseph C. V´arilly, An introduction to noncommutative geometry, EMS Series of Lec- tures in Mathematics, European Mathematical Society (EMS), Zurich, 2006. MR2239597 (2007e:58011) [38] Marc Wambst, Hochschild and cyclic homology of the quantum multiparametric torus, J. Pure Appl. Algebra 114 (1997), no. 3, 321 -- 329. MR1426492 (98c:16008) Mathematical Institute, Utrecht University, PO Box 80010, 3508 TA Utrecht, The Netherlands E-mail address: [email protected]
1006.3115
1
1006
2010-06-16T00:36:50
Strength of convergence in the orbit space of a groupoid
[ "math.OA" ]
Let G be a second-countable locally-compact Hausdorff groupoid with a Haar system, and let {x_n} be a sequence in the unit space of G. We show that the notions of strength of convergence of {x_n} in the orbit space and measure-theoretic accumulation along the orbits are equivalent ways of realising multiplicity numbers associated to a sequence of induced representation of the groupoid C*-algebra.
math.OA
math
STRENGTH OF CONVERGENCE IN THE ORBIT SPACE OF A GROUPOID ROBERT HAZLEWOOD AND ASTRID AN HUEF Abstract. Let G be a second-countable locally-compact Hausdorff groupoid with a Haar system, and let {xn} be a sequence in the unit space G(0) of G. We show that the notions of strength of convergence of {xn} in the orbit space G(0)/G and measure-theoretic accumulation along the orbits are equivalent ways of realising multiplicity numbers associated to a sequence of induced representation of the groupoid C∗-algebra. 1. Introduction (i) s−1 (ii) tns−1 subsequence. Suppose H is a locally-compact Hausdorff group acting freely and con- tinuously on a locally-compact Hausdorff space X, so that (H, X) is a free transformation group. In [13, pp. 95 -- 96] Green gives an example of a free non-proper action of H = R on a subset X of R3; the non-properness comes down to the existence of z ∈ X, {xn} ⊂ X, and two sequences {sn} and {tn} in H such that n · xn → z; and n · xn → z and t−1 n → ∞ as n → ∞, in the sense that {tns−1 n } has no convergent In [2, Definition 2.2]), and subsequently in [3, p. 2], the sequence {xn} is said to converge 2-times in the orbit space to z ∈ X. Each orbit H · x gives an induced representation Ind x of the associated transformation-group C∗- algebra C0(X) (cid:111) H which is irreducible, and the k-times convergence of {xn} in the orbit space to z ∈ X translates into statements about various multiplicity numbers associated to Ind z in the spectrum of C0(X) (cid:111) H, as in [2, Theorem 2.5], [3, Theorem 1.1] and [4, Theorem 2.1]. Upper and lower multiplicity numbers associated to irreducible represen- tations π of a C∗-algebra A were introduced by Archbold [1] and extended to multiplicity numbers relative to a net of irreducible representations by Archbold and Spielberg [9]. The upper multiplicity MU (π) of π, for exam- ple, counts 'the number of nets of orthogonal equivalent pure states which can converge to a common pure state associated to π' [6, p. 26]. The defini- tion of k-times convergence and [2, Theorem 2.5] were very much motivated by a notion of k-times convergence in the dual space of a nilpotent Lie group [16] and its connection with relative multiplicity numbers (see, for example, [6, Theorem 2.4] and [7, Theorem 5.8]). Date: 16 June 2010. This research was supported by the Australian Research Council. Robert Hazlewood thanks the Department of Mathematics and Statistics at the University of Otago for their kind hospitality. 1 2 ROBERT HAZLEWOOD AND ASTRID AN HUEF Theorem 1.1 of [3] shows that the topological property of a sequence {xn} converging k-times in the orbit space to z ∈ X is equivalent to (1) a measure theoretic accumulation along the orbits G · xn and (2) that the lower multiplicity of Ind z relative to the sequence {Ind xn} is at least k. In this paper we prove that the results of [3] generalise to principal groupoids. In our main arguments we have tried to preserve as much as possible the structure of those in [3], although the arguments presented here are often more complicated in order to cope with the partially defined product in a groupoid and the set of measures that is a Haar system compared to the fixed Haar measure used in the transformation-group case. Our theorems have led us to a new class of examples exhibiting k-times convergence in groupoids that are not based on transformation groups, thus justifying our level of generality. Given a row-finite directed graph E, Kumjian, Pask, Raeburn and Renault in [15] used the set of all infinite paths in E to construct an r-discrete groupoid GE, called a path groupoid. We prove that GE is principal if and only if E contains no cycles (Proposi- tion 8.1). We then exhibit principal GE with Hausdorff and non-Hausdorff orbits space, respectively, both with a k-times converging sequence in the orbit space. In particular, our examples can be used to find a groupoid GE whose C∗-algebra has non-Hausdorff spectrum and distinct upper and lower multiplicity counts among its irreducible representations. 2. Preliminaries call the set r(cid:0)s−1({x})(cid:1) = s(cid:0)r−1({x})(cid:1) the orbit of x and denote it by We denote the unit space of a groupoid G by G(0). For x ∈ G(0) we [x]. For a subset U of G(0) we define GU := s−1(U ), GU := r−1(U ), and GU := s−1(U )∩ r−1(U ). We denote the set of all positive integers by P and the set of all non-negative integers by N. Definition 2.1. A right Haar system on G is a set {λx : x ∈ G(0)} of non- negative Radon measures on G such that (ii) for f ∈ Cc(G), the function x (cid:55)→(cid:82) f dλx on G(0) is in Cc(G(0)); and for all x ∈ G(0); (i) supp λx = Gx (iii) for f ∈ Cc(G) and γ ∈ G, (cid:0) = s−1({x})(cid:1) (cid:90) (cid:90) f (αγ) dλr(γ)(α) = f (α) dλs(γ)(α). We will refer to (ii) as the continuity of the Haar system and to (iii) as Haar-system invariance. The collection {λx : x ∈ G(0)} of measures where λx(E) := λx(E−1) is a left Haar system, which is a system of measures such that supp λx = Gx and, for f ∈ Cc(G), x (cid:55)→ (cid:82) f dλx is continuous and (cid:82) f (γα) dλs(γ)(α) = (cid:82) f (α) dλr(γ)(α). Given that we can easily convert a right Haar system {λx} into a left Haar system {λx} and vice versa, we will simply refer to a Haar system λ and use subscripts to refer to elements of the right Haar system {λx} and superscripts to refer to elements of the left Haar system {λx}. The following lemma follows from the invariance of the Haar system and the Dominated Convergence Theorem; we omit the proof. STRENGTH OF CONVERGENCE IN THE ORBIT SPACE OF A GROUPOID 3 Lemma 2.2 (Haar-system invariance). Suppose G is a locally-compact Haus- dorff groupoid with Haar system λ. If K ⊂ G is compact and γ ∈ G with s(γ) = x and r(γ) = y, then λx(Kγ) = λy(K) and λx(γ−1K) = λy(K). Definition 2.3 below is Definition 2.45 in the unpublished book [17]. Al- ternative descriptions of the induced representation may be found in [19, pp. 234] and [23, pp. 81 -- 82]. (cid:82) Definition 2.3. Suppose G is a second-countable locally-compact Hausdorff groupoid with Haar system λ and let µ be a Radon measure on G(0). (i) We write ν = µ ◦ λ = (cid:82) λx dµ for the measure on G defined for every Borel-measurable function f : G → C by (cid:82) (cid:82) the formula(cid:0)Ind µ(f )ξ(cid:1)(γ) = G f (γ) dν(γ) = G f (γ) dλx(γ) dµ(x). We call ν the measure induced by µ, and we write ν−1 for the image of ν under the homeomorphism γ (cid:55)→ γ−1. (ii) For f ∈ Cc(G), Ind µ(f ) is the operator on L2(G, ν−1) defined by f (α)ξ(α−1γ) dλr(γ)(α) G(0) (cid:90) (cid:90) G (cid:90) G = f (γα)ξ(α−1) dλs(γ)(α). G In this paper we are interested in representations that are induced by point-mass measures δx on G(0). We denote Ind δx by Lx for all x ∈ G(0) as in [19] and [11]. Remark 2.4. It follows from the definition of the induced mesasure that for x ∈ G(0), the measure ν induced by δx is equal to λx. In particular we have ν−1 = λx, so Lx acts on L2(G, λx). The operator Lx is then given by (cid:0)Lx(f )ξ(cid:1)(γ) = f (γα−1)ξ(α) dλx(α) for all ξ ∈ L2(G, λx) and all γ ∈ G. There is a close relationship between the convolution on Cc(G) and these induced representations: recall that for f, g ∈ Cc(G), the convolution f ∗ g ∈ Cc(G) is given by f ∗ g(γ) = f (γα−1)g(α) dλs(γ)(α) for all γ ∈ G, (cid:90) so that G (cid:0)Lx(f )g(cid:1)(γ) = f ∗ g(γ) for any x ∈ G(0) and γ ∈ Gx. We denote the norm in L2(G, λx) by (cid:107) · (cid:107)x. Finally note that when G is a second-countable locally-compact principal groupoid that admits a Haar system, each Lx is irreducible by [19, Lemma 2.4]. Remark 2.5. If G = (H, X) is a second-countable free transformation group, then the representations Lx defined above are unitarily equivalent to the representations Ind x used in [3]. Specifically, let ν be a choice of right Haar measure on H and ∆ the associated modular function. The map ι : Cc(H × X) → Cc(H × X) defined by ι(f )(t, x) = f (t, x)∆(t)1/2 4 ROBERT HAZLEWOOD AND ASTRID AN HUEF extends to an isomorphism ι of the groupoid C∗-algebra C∗(H × X) onto the transformation-group C∗-algebra C0(X) (cid:111) H [23, p. 58]. Fix x ∈ X. Then there is a unitary Ux : L2(H, ν) → L2(H × X, λx), characterised by U (ξ)(h, y) = ξ(h)δx(h−1 · y) for ξ ∈ Cc(H), and U (Ind x(ι(f ))U∗ = Lx(f ) for f ∈ C∗(H × X). Let A be a C∗-algebra. We write θ for the canonical surjection from the space P (A) of pure states of A to the spectrum A of A. We frequently identify an irreducible representation π with its equivalence class in A and we write Hπ for the Hilbert space on which π(A) acts. Let π ∈ A and let {πα} be a net in A. We now recall the definitions of upper and lower multiplicity MU(π) and ML(π) from [1], and relative upper and relative lower multiplicity MU(π,{πα}) and ML(π,{πα}) from [9]. Let N be the weak∗-neighbourhood base at zero in the dual A∗ of A consisting of all open sets of the form N = {ψ ∈ A∗ : ψ(ai) < , 1 (cid:54) i (cid:54) n}, where  > 0 and a1, a2, . . . , an ∈ A. Suppose φ is a pure state of A associated with π and let N ∈ N . Define V (φ, N ) = θ(cid:0)(φ + N ) ∩ P (A)(cid:1), an open neighbourhood of π in A. For σ ∈ A let Vec(σ, φ, N ) = {η ∈ Hσ : (cid:107)η(cid:107) = 1, (σ(·)η η) ∈ φ + N}. Note that Vec(σ, φ, N ) is non-empty if and only if σ ∈ V (φ, N ). For any σ ∈ V (φ, N ) define d(σ, φ, N ) to be the supremum in P∪{∞} of the cardinalities of finite orthonormal subsets of Vec(σ, φ, N ). Write d(σ, φ, N ) = 0 when Vec(σ, φ, N ) is empty. Define MU(φ, N ) = sup σ∈V (φ,N ) d(σ, φ, N ) ∈ P ∪ {∞}. Note that if N(cid:48) ∈ N and N ⊂ N , then MU(φ, N(cid:48)) (cid:54) MU(φ, N ). Now define MU(φ) = inf N∈N MU(φ, N ) ∈ P ∪ {∞}. By [1, Lemma 2.1], the value of MU(φ) is independent of the pure state φ associated to π. Thus MU(π) := MU(φ) is well-defined. For lower multi- plicity, assume that {π} is not open in A, and using [1, Lemma 2.1] again, define (cid:17) ∈ P ∪ {∞}. (cid:16) ML(π) := inf N∈N lim inf σ→π,σ(cid:54)=π d(σ, φ, N ) Now suppose that {πα}α∈Λ is a net in A. For N ∈ N let (cid:0)φ, N,{πα}(cid:1) = lim sup (cid:0)π,{πα}(cid:1) := inf α∈Λ N∈N MU MU MU Note that if N(cid:48) ∈ N and N(cid:48) ⊂ N then MU Then d(πα, φ, N ) ∈ N ∪ {∞}. (cid:0)φ, N(cid:48),{πα}(cid:1) (cid:54) MU (cid:0)φ, N,{πα}(cid:1) ∈ N ∪ {∞}, (cid:0)φ, N,{πα}(cid:1). STRENGTH OF CONVERGENCE IN THE ORBIT SPACE OF A GROUPOID 5 is well-defined because the right-hand side is independent of the choice of φ by an argument similar to the proof of [1, Lemma 2.1]. Similarly define ML and let ML α∈Λ (cid:0)φ, N,{πα}(cid:1) := lim inf (cid:0)π,{πα}(cid:1) = inf (cid:0)π,{πα}(cid:1) (cid:54) MU N∈N ML ML It follows that for any irreducible representation π and any net {πα}α∈Λ of irreducible representations, d(πα, φ, N ) ∈ N ∪ {∞}, (cid:0)φ, N,{πα}(cid:1) ∈ N ∪ {∞}. (cid:0)π,{πα}(cid:1) (cid:54) MU(π) (cid:0)π,{πα}(cid:1). (cid:0)π,{πβ}(cid:1) (cid:54) MU (cid:0)π,{πα}(cid:1). and, if {πα} converges to π with πα (cid:54)= π eventually, ML(π) (cid:54) ML Finally, if {πβ} is a subnet of {πα}, then (cid:0)π,{πα}(cid:1) (cid:54) ML (cid:0)π,{πβ}(cid:1) (cid:54) MU ML 3. Lower multiplicity and k-times convergence I A key goal for this paper is to describe the relationship between multi- plicities of induced representations and strength of convergence in the orbit space. We start this section by recalling the definition of k-times conver- gence in a groupoid from [11]. We then show that if a sequence converges k-times in the orbit space of a principal groupoid, then the lower multiplicity of the associated sequence of representations is at least k; the converse will be shown in Section 6. Recall that a sequence {γn} ⊂ G tends to infinity if it admits no conver- gent subsequence. Definition 3.1. Let k ∈ P. A sequence {xn} in G(0) is k-times convergent in n } ⊂ G G(0)/G to z ∈ G(0) if there exist k sequences {γ(1) such that n }, . . . ,{γ(k) n },{γ(2) n ) = xn for all n and 1 (cid:54) i (cid:54) k; n ) → z as n → ∞ for 1 (cid:54) i (cid:54) k; and (i) s(γ(i) (ii) r(γ(i) (iii) if 1 (cid:54) i < j (cid:54) k then γ(j) The proof of the following proposition is based on [2, Theorem 2.3] and n )−1 → ∞ as n → ∞. n (γ(i) a part of [3, Theorem 1.1]. Proposition 3.2. Suppose G is a second-countable locally-compact Haus- dorff principal groupoid with Haar system λ. Let z ∈ G(0) and suppose that {xn} is a sequence in G(0) that converges k-times to z in G(0)/G. Then ML(Lz,{Lxn}) (cid:62) k. Proof. We will use a contradiction argument. Suppose that ML(Lz,{Lxn}) = r < k. Fix a real-valued g ∈ Cc(G) so that (cid:107)g(cid:107)z > 0. Define η ∈ L2(G, λz) by η(α) = (cid:107)g(cid:107)−1 z g(α) for all α ∈ G. Then (cid:107)η(cid:107)2 z = (cid:107)g(cid:107)−2 z g(α)2 dλz(α) = (cid:107)g(cid:107)−2 z (cid:107)g(cid:107)2 z = 1, (cid:90) 6 ROBERT HAZLEWOOD AND ASTRID AN HUEF so η is a unit vector in L2(G, λz) and the GNS construction of φ := (Lz(·)η η) is unitarily equivalent to Lz. By the definition of lower multiplicity we now have ML(Lz,{Lxn}) = inf N∈N ML(φ, N,{Lxn}) = r, so there exists N ∈ N such that ML(φ, N,{Lxn}) = lim inf n d(Lxn, φ, N ) = r, and consequently there exists a subsequence {ym} of {xn} such that d(Lym, φ, N ) = r for all m. (3.1) Since any subsequence of a sequence that is k-times convergent is also k- times convergent, we know that {ym} converges k-times to z in G(0)/G. We will now use the k-times convergence of {ym} to construct k sequences of unit vectors with sufficient properties to establish our contradiction. By the k-times convergence of {ym} there exist k sequences {γ(1) m },{γ(2) m }, . . . ,{γ(k) m } ⊂ G satisfying m ) = ym for all m and 1 (cid:54) i (cid:54) k; m ) → z as m → ∞ for 1 (cid:54) i (cid:54) k; and (i) s(γ(i) (ii) r(γ(i) (iii) if 1 (cid:54) i < j (cid:54) k then γ(j) For each 1 (cid:54) i (cid:54) k and m (cid:62) 1, define η(i) m (γ(i) m by m )−1 → ∞ as m → ∞. It follows from Haar-system invariance that m (α) = (cid:107)g(cid:107)−1 η(i) r(γ(i) m ) (cid:107)η(i) m (cid:107)2 ym = (cid:107)g(cid:107)−2 = (cid:107)g(cid:107)−2 = (cid:107)g(cid:107)−2 r(γ(i) m ) r(γ(i) m ) r(γ(i) m ) g(cid:0)α(γ(i) m )−1(cid:1) (cid:90) g(cid:0)α(γ(i) (cid:90) for all α ∈ G. m )−1(cid:1)2 dλym(α) g(α)2 dλ (α) r(γ(i) m ) (cid:107)g(cid:107)2 r(γ(i) m ) = 1, so the η(i) m are unit vectors in L2(G, λym). Now suppose that 1 (cid:54) i < j (cid:54) k. Then (cid:90) g(cid:0)α(γ(i) m )−1(cid:1)g(cid:0)α(γ(j) m )−1(cid:1) dλym(α) m η(j) (η(i) m )ym = (cid:107)g(cid:107)−1 r(γ(i) m ) (cid:107)g(cid:107)−1 m (γ(j) Since γ(i) (supp g)(supp g)−1, and so there exists m0 such that if m (cid:62) m0, then m (γ(j) m )−1 → ∞, γ(i) r(γ(j) m ) (3.2) m )−1 is eventually not in the compact set (supp g)γ(i) m ∩ (supp g)γ(j) m = ∅. (To see this claim, note that if (supp g)γ(i) exist α, β ∈ supp g such that αγ(i) m = βγ(j) (supp g)−1(supp g).) For the integrand of (3.2) to be non-zero, both α(γ(i) and α(γ(j) m )−1 must be in supp g, so α must be in (supp g)γ(i) m (cid:54)= ∅ then there m )−1 = α−1β ∈ m )−1 m ∩ (supp g)γ(j) m . m ∩ (supp g)γ(j) m (γ(j) m , and so γ(i) STRENGTH OF CONVERGENCE IN THE ORBIT SPACE OF A GROUPOID 7 But this is not possible if m (cid:62) m0. Thus, for any distinct i, j, we will eventually have η(i) m ⊥ η(j) m . For the last main component of this proof we will establish that (cid:1) →(cid:0)Lz(·)η(cid:12)(cid:12) η(cid:1) = φ as m → ∞ in the dual of C∗(G) with the weak∗ topology for each i. Fix f ∈ Cc(G). (cid:0)Lym(·)η(i) (cid:90) We have(cid:0)Lz(f )η(cid:12)(cid:12) η(cid:1) = (cid:90) m G m (cid:12)(cid:12) η(i) (cid:0)Lz(f )η(cid:1)(α)η(α) dλz(α) (cid:90) (cid:90) (cid:90) G f (αβ−1)g(β)g(α) dλz(β) dλz(α) Now fix 1 (cid:54) i (cid:54) k. By the invariance of the Haar system we have = (cid:107)g(cid:107)−2 G z (3.3) f (αβ−1)η(β)η(α) dλz(β) dλz(α) G (cid:90) f (αβ−1)g(cid:0)α(γ(i) f (αβ−1)η(i) G m (β)η(i) m (α) dλym(β) dλym(α) m )−1(cid:1)g(cid:0)β(γ(i) m )−1(cid:1) dλym(β) dλym(α) = G (cid:90) (cid:90) (cid:90) G G G (cid:1) = (cid:90) (cid:90) (cid:90) G G (cid:0)Lym(f )η(i) m (cid:12)(cid:12) η(i) m = (cid:107)g(cid:107)−2 r(γ(i) m ) = (cid:107)g(cid:107)−2 r(γ(i) m ) = (cid:107)g(cid:107)−2 f (αβ−1)g(α)g(β) dλ r(γ(i) m ) (β) dλ r(γ(i) m ) (α) f ∗ g(α)g(α) dλ r(γ(i) m ) G We know that r(γ(i) system, (cid:107)g(cid:107) continuity of the Haar system with (3.3) and (3.4) to see that (3.4) m ) → z as m → ∞ so, by the continuity of the Haar → (cid:107)g(cid:107)z as m → ∞. Since f ∗ g ∈ Cc(G) we can apply the r(γ(i) m ) r(γ(i) m ) (α). (cid:0)Lym(f )η(i) m (cid:12)(cid:12) η(i) m (cid:1) = (cid:107)g(cid:107)−2 G (α) r(γ(i) m ) f ∗ g(α)g(α) dλ (cid:90) f ∗ g(α)g(α) dλz(α) =(cid:0)Lz(f )η(cid:12)(cid:12) η(cid:1) (cid:1) →(cid:0)Lz(·)η(cid:12)(cid:12) η(cid:1) = φ (cid:90) (cid:12)(cid:12) η(i) G r(γ(i) m ) → (cid:107)g(cid:107)−2 z (cid:0)Lym(·)η(i) as m → ∞. We have thus shown that, for each i, in the dual of C∗(G) equipped with the weak∗ topology. Thus there exists m1 such that for any m (cid:62) m1 and any 1 (cid:54) i (cid:54) k, the pure state(cid:0)Lym(·)η(i) (cid:12)(cid:12) η(i) (cid:1) m m m m is in φ + N . We have now established that every η(i) m with m (cid:62) max{m0, m1} is in m for i (cid:54)= j, so d(Lym, φ, N ) (cid:62) k for all m (cid:62) Vec(Lym, φ, N ) with η(i) max{m0, m1}, contradicting our earlier observation (3.1) that d(Lym, φ, N ) = (cid:3) r < k for all m. m ⊥ η(j) 4. Measure ratios and k-times convergence In this section we show that lower bounds on measure ratios along orbits gives strength of convergence in the orbit space. We begin by generalising [3, Proposition 4.1]. 8 ROBERT HAZLEWOOD AND ASTRID AN HUEF Proposition 4.1. Let G be a second-countable locally-compact Hausdorff principal groupoid with Haar system λ. Let k ∈ P and z ∈ G(0) with [z] locally closed in G(0). Assume that {xn} is a sequence in G(0) such that [xn] → [z] uniquely in G(0)/G. Suppose {Wm} is a basic decreasing sequence of compact neighbourhoods of z such that each m satisfies λxn(GWm) > (k − 1)λz(GWm). lim inf Then {xn} converges k-times in G(0)/G to z. Proof. Let {Km} be an increasing sequence of compact subsets of G such m(cid:62)1 Int Km. By the regularity of λz, for each m (cid:62) 1 there exist that G =(cid:83) n δm > 0 and an open neighbourhood Um of GWm such that λxn(GWm) > (k − 1)λz(Um) + δm. z lim inf n (4.1) We will construct, by induction, a strictly increasing sequence of positive integers {nm} such that, for all n (cid:62) nm, λxn(Kmα ∩ GWm) < λz(Um) + δm/k for all α ∈ GWm xn , λxn(GWm) > (k − 1)λz(Um) + δm. (4.3) By applying Lemma 5.5 with δ = λz(U1) − λz(GW1) + δ1/k there exists (4.2) and n1 such that n (cid:62) n1 implies λxn(K1α ∩ GW1) < λz(U1) + δ1/k for all α ∈ GW1 xn , establishing (4.2) for m = 1. If necessary we can increase n1 to ensure (4.3) holds for m = 1 by considering (4.1). Assuming that we have constructed n1 < n2 < ··· < nm−1, we apply Lemma 5.5 with δ = λz(Um) − λz(GWm) + δm/k to obtain nm > nm−1 such that (4.2) holds, and again, if necessary, increase nm to obtain (4.3). n = xn. For each n (cid:62) n1 there is a unique m such that nm (cid:54) n < nm+1. For every such n and m choose γ(1) (which is always non-empty by (4.3)). Using (4.2) and (4.3) we have λxn(GWm\Kmγ(1) If n1 > 1 then, for each 1 (cid:54) n < n1 and 1 (cid:54) i (cid:54) k, let γ(i) n ) = λxn(GWm) − λxn(GWm ∩ Kmγ(1) n ) n ∈ GWm xn >(cid:0)(k − 1)λz(Um) + δm (cid:1) −(cid:0)λz(Um) + δm/k(cid:1) = (k − 2)λz(Um) + (k − 1) δm. k λxn So for each n (cid:62) n1 and its associated m we can choose γ(2) We now have n ∈ GWm n )(cid:1) n ∪ Kmγ(2) (cid:0)GWm\(Kmγ(1) = λxn(GWm\Kmγ(1) (cid:16) (cid:62) λxn(GWm\Kmγ(1) (k − 2)λz(Um) + > = (k − 3)λz(Um) + (k − 1) k (k − 2) δm, k (cid:0)(GWm\Kmγ(1) (cid:17) −(cid:16) n ) − λxn n ) ∩ Kmγ(2) (cid:17) n ) − λxn(GWm ∩ Kmγ(2) n ) λz(Um) + δm/k δm n xn \Kmγ(1) n . (cid:1) STRENGTH OF CONVERGENCE IN THE ORBIT SPACE OF A GROUPOID 9 n ∪ Kmγ(2) enabling us to choose γ(3) process, for each j = 3, . . . , k and each n (cid:62) n1 we have xn \(Kmγ(1) n ∈ GWm n ). By continuing this (cid:32) (cid:18) j−1(cid:91) GWm\ λxn Kmγ(i) n (cid:19)(cid:33) i=1 enabling us to choose > (k − j)λz(Um) + (k − j − 1)δm k , (cid:19) . Kmγ(i) n (cid:18) j−1(cid:91) i=1 n ∈ GWm xn \ γ(j) (4.4) Note that for nm (cid:54) n < nm+1 we have γ(j) n /∈ Kmγ(i) n for 1 (cid:54) i < j (cid:54) k. n . Note that s(γ(i) We will now establish that xn converges k-times to z in G(0)/G by con- sidering the γ(i) n ) = xn for all n and i by our choice of the n ) → z as n → ∞ for 1 (cid:54) i (cid:54) k. To γ(i) n . We will now establish that r(γ(i) see this, fix i and let V be an open neighbourhood of z. Since Wm → {z} there exists m0 such that m (cid:62) m0 implies Wm ⊂ V . For each n (cid:62) nm0 there exists a m (cid:62) m0 such that nm (cid:54) n < nm+1, and so r(γ(i) n )−1 → ∞ as n → ∞ for 1 (cid:54) i < j (cid:54) k. Fix i < j and let K be a compact subset of G. There exists m0 such that K ⊂ Km for all m (cid:62) m0. Then for n (cid:62) nm0 there exists m (cid:62) m0 such that nm (cid:54) n < nm+1. By (4.4) we know Finally we claim that γ(j) n ) ∈ Wm ⊂ V . n (γ(i) n ∈ GWm γ(j) =(cid:0)GWm =(cid:0)(GWm xn \(Kmγ(i) n ) xn (γ(i) xn (γ(i) n )−1γ(i) n )−1)\Km (cid:1)\(Kmγ(i) (cid:1)γ(i) n n ) n )−1 ∈(cid:0)GWm n )−1(cid:1)\Km ⊂ G\Km ⊂ G\K, enabling us to n , n (γ(i) and so γ(j) conclude that {xn} converges k-times in G(0)/G to z. xn (γ(i) (cid:3) In Proposition 4.4 below we prove a generalisation of a part of [3, Propo- sition 4.2]; to do this we need the following two lemmas. Lemma 4.2. Suppose G is a second-countable groupoid with Haar system λ and let K be a compact subset of G. If {xn} ⊂ G(0) is a sequence that converges to z ∈ G(0), then λxn(K) (cid:54) λz(K). lim sup n Proof. Fix  > 0. By the outer regularity of λz, there exists an open neigh- bourhood U of K such that λz(K) (cid:54) λz(U ) < λz(K) + /2. By Urysohn's Lemma there exists f ∈ Cc(G) with 0 (cid:54) f (cid:54) 1 such that f is identically one on K and zero off U . In particular we have (cid:90) λz(K) (cid:54) f dλz < λz(K) + /2. (4.5) 10 ROBERT HAZLEWOOD AND ASTRID AN HUEF The continuity of the Haar system implies (cid:82) f dλxn →(cid:82) f dλz, so there (cid:90) By our choice of f we have λxn(K) (cid:54)(cid:82) f dλxn, so exists n0 such that n (cid:62) n0 implies f dλz − /2 < f dλz + /2. f dλxn < (cid:90) (cid:90) (cid:90) (cid:90) λxn(K) (cid:54) f dλxn < f dλz + /2. Combining this with (4.5) enables us to observe that for n (cid:62) n0, λxn(K) < (cid:3) λz(K) + , completing the proof. Lemma 4.3. Suppose G is a second-countable groupoid with Haar system λ and let K be a compact subset of G. For every  > 0 and z ∈ G(0) there exists a neighbourhood U of z ∈ G(0) such that x ∈ U implies λx(K) < λz(K) + . Proof. Fix  > 0 and z ∈ G(0). Let {Un} be a decreasing neighbourhood basis for z in G(0). If our claim is false, then each Un contains an element xn such that λxn(K) (cid:62) λz(K) + . But since each xn ∈ Un, xn → z, and so by Lemma 4.2 there exists n0 such that n (cid:62) n0 implies λxn(K) < λz(K) + , (cid:3) a contradiction. Proposition 4.4. Suppose G is a second-countable locally-compact Haus- dorff groupoid with Haar system λ. Suppose that z ∈ G(0) with [z] locally closed in G(0) and suppose {xn} is a sequence in G(0). Assume that for every open neighbourhood V of z in G(0) such that GV is relatively com- z pact, λxn(GV ) → ∞ as n → ∞. Then, for every k (cid:62) 1, the sequence {xn} converges k-times in G(0)/G to z. Proof. Let {Km} be an increasing sequence of compact subsets of G such that G = ∪m(cid:62)1 Int Km. By Lemma 4.3, for each Km there exists an open neighbourhood Vm of z such that x ∈ Vm implies λx(Km) < λz(Km) + 1. Since [z] is locally closed, by Lemma 4.1(1) in [12] we can crop V1 if necessary to ensure that GV1 is relatively compact. By further cropping each Vm we may assume that {Vm} is a decreasing neighbourhood basis of z. By our z hypothesis, for each m there exists nm such that n (cid:62) nm implies λxn(GVm) > k(cid:0)λz(Km) + 1(cid:1). (4.6) Note that for any γ ∈ GVm xn with n (cid:62) nm, we have r(γ) ∈ Vm, and so λr(γ)(Km) < λz(Km) + 1. By Haar-system invariance we know that λr(γ)(Km) = λxn(Kmγ), which shows us that (4.7) If necessary we can increase the elements of {nm} so that it is a strictly increasing sequence. λxn(Kmγ) < λz(Km) + 1. We now proceed as in the proof of Proposition 4.1. For all n < n1 and 1 (cid:54) i (cid:54) k let γ(i) n = xn. For each n (cid:62) n1 there exists a unique number m(n) such that nm(n) (cid:54) n < nm(n)+1. For the remainder of this proof we STRENGTH OF CONVERGENCE IN THE ORBIT SPACE OF A GROUPOID 11 will write m instead of m(n) because the specific n will be clear from the context. For each n (cid:62) n1 choose n ∈ GVm γ(1) xn . Then by (4.6) and (4.7) we have We can thus choose γ(2) λxn(GVm\(Kmγ(1) for each n (cid:62) n1. This now gives us n λxn(GVm\Kmγ(1) n ) = λxn(GVm) − λxn(GVm ∩ Kmγ(1) n ) n ∈ GVm (cid:62) λxn(GVm) − λxn(Kmγ(1) n ) > k(cid:0)λz(Km) + 1(cid:1) −(cid:0)λz(Km) + 1(cid:1) = (k − 1)(cid:0)λz(Km) + 1(cid:1). (cid:0)(GVm\Kmγ(1) > (k − 1)(cid:0)λz(Km) + 1(cid:1) −(cid:0)λz(Km) + 1(cid:1) = (k − 2)(cid:0)λz(Km) + 1(cid:1). (cid:19) xx \Kmγ(1) n ∪ Kmγ(2) n )) n ) − λxn n ) − λxn(Kmγ(2) n ) = λxn(GVm\Kmγ(1) (cid:62) λxn(GVm\Kmγ(1) n ) ∩ Kmγ(1) n (cid:1) Continuing in this manner we can choose (cid:18) j−1(cid:91) xn \ n ∈ GVm γ(j) Kmγ(i) n for every n (cid:62) n1 and j = 3, . . . , k. The tail of the proof of Proposition 4.1 (cid:3) establishes our desired result. i=1 5. Measure ratios and bounds on lower multiplicity In this section we show that upper bounds on measure ratios along orbits gives upper bounds on multiplicities. A subset S of a topological space X is locally closed if there exist an open set U of X and a closed set V of X such that S = U ∩ V ; this is equivalent to S being open in the closure of S with the subspace topology by, for example, [24, Lemma 1.25]. Lemma 5.1. Suppose G is a second-countable locally-compact Hausdorff groupoid. Suppose z ∈ G(0) and [z] is locally closed. Then the restriction of r to Gz/(G{z}) is a homeomorphism onto [z]. If in addition G is principal, then the restriction of r to Gz is a homeomorphism onto [z]. Proof. We consider the transitive groupoid G[z]. Since [z] is locally closed, G[z] is a second-countable locally-compact Hausdorff groupoid. Thus G[z] is Polish by, for example, [24, Lemma 6.5]. Now [22, Theorem 2.1] applies (cid:3) to give the result. Theorem 5.2 is based on [3, Theorem 3.1]; it is only an intermediary result which will be used to prove a sharper bound in Theorem 5.8. Theorem 5.2. Suppose G is a second-countable locally-compact Hausdorff principal groupoid with Haar system λ. Let M ∈ R with M (cid:62) 1, suppose z ∈ G(0) such that [z] is locally closed and let {xn} be a sequence in G(0). 12 ROBERT HAZLEWOOD AND ASTRID AN HUEF Suppose there exists an open neighbourhood V of z in G(0) such that GV relatively compact and z is λxn(GV ) (cid:54) M λz(GV ) frequently. Then ML(Lz,{Lxn}) (cid:54) (cid:98)M 2(cid:99). Proofish. Fix  > 0 such that M 2(1 + )2 < (cid:98)M 2(cid:99) + 1. We will build a function D ∈ Cc(G) such that Lz(D∗ ∗ D) is a rank-one projection and Tr(cid:0)Lxn(D∗ ∗ D)(cid:1) < M 2(1 + )2 < (cid:98)M 2(cid:99) + 1 frequently. By the generalised lower semi-continuity result of [9, Theo- rem 4.3] we will have lim inf Tr(cid:0)Lxn(D∗ ∗ D)(cid:1) (cid:62) ML(Lz,{Lxn}) Tr(cid:0)Lz(D∗ ∗ D)(cid:1) = ML(Lz,{Lxn}), and the result will follow. For the next few paragraphs we will be working with Gz equipped with the subspace topology. Note that λz can be thought of as a Radon measure on Gz with λz(S ∩ Gz) = λz(S) for any λz-measurable subset S of G. Fix δ > 0 such that δ < λz(GV ) 1 +  < λz(GV ). Since λz is inner regular on open sets and GV z Gz-compact subset W of GV z such that is Gz-open, there exists a Since W is Gz-compact there exists a Gz-compact neighbourhood W1 of W z and there exists a continuous function g : Gz → [0, 1] that is contained in GV that is identically one on W and zero off the interior of W1. We have g(t)2 dλz(t) = (cid:107)g(cid:107)2 z, λz(GV ) − δ = λz(GV z ) − δ < λz(W ) (cid:54) 0 < λz(GV z ) − δ < λz(W ). (cid:90) Gz δ and hence λz(GV ) (cid:107)g(cid:107)2 z δ (cid:107)g(cid:107)2 z < 1 + < 1 + < 1 + . (5.1) λz(GV ) − δ (cid:0)r(γ)(cid:1) = g(γ) for all γ ∈ W1. Thus r(W1) is [z]-compact, which implies By Lemma 5.1 the restriction r of r to Gz is a homeomorphism onto [z]. So there exists a continuous function g1 : r(W1) → [0, 1] such that g1 that r(W1) is G(0)-compact. Since we know that G(0) is second countable and Hausdorff, Tietze's Extension Theorem can be applied to extend g1 to a continuous map g2 : G(0) → [0, 1]. Because r(W1) is a compact subset of the open set V , there exist a compact neighbourhood P of r(W1) contained in V and a continuous function h : G(0) → [0, 1] that is identically one on r(W1) and zero off the interior of P . Note that h has compact support that is contained in P . We set f (x) = h(x)g2(x). Then f ∈ Cc(G(0)) with 0 (cid:54) f (cid:54) 1 and suppf ⊂ supph ⊂ P ⊂ V. (5.2) STRENGTH OF CONVERGENCE IN THE ORBIT SPACE OF A GROUPOID 13 Note that (cid:107)f ◦ r(cid:107)2 z = = (cid:62) (cid:90) (cid:90) (cid:90) (cid:90) Gz Gz W1 = = (cid:107)g(cid:107)2 W1 f(cid:0)r(γ)(cid:1)2 dλz(γ) h(cid:0)r(γ)(cid:1)2g2 (cid:0)r(γ)(cid:1)2 dλz(γ) h(cid:0)r(γ)(cid:1)2g(γ)2 dλz(γ) g(γ)2 dλz(γ) (5.3) since supp g ⊂ W1 and h is identically one on r(W1). We now define F ∈ Cc(G(0)) by z Then (cid:107)F ◦ r(cid:107)z = 1 and F ◦ r(γ) (cid:54)= 0 =⇒ h(cid:0)r(γ)(cid:1) (cid:54)= 0 =⇒ r(γ) ∈ V =⇒ γ ∈ GV . F (x) = f (x) (cid:107)f ◦ r(cid:107)z . (5.4) (5.5) Let N = supp F so that N = supp f ⊂ V by (5.2) and (5.4). Since GV relatively compact by our hypothesis, the set GN be a function that is identically one on (GN z )(GN in [0, 1]. We can assume that b is self-adjoint by considering 1 necessary. Define D ∈ Cc(G) by z is z is compact. Let b ∈ Cc(G) z )−1 and has range contained 2 (b + b∗) if For ξ ∈ L2(G, λu) and γ ∈ G we have (cid:0)Lu(D)ξ(cid:1)(γ) = G D(γα−1)ξ(α) dλu(α) D(γ) := F(cid:0)r(γ)(cid:1)F(cid:0)s(γ)(cid:1)b(γ). (cid:90) (cid:90) = F(cid:0)r(γ)(cid:1)(cid:90) (cid:90) (cid:0)Lz(D)ξ(cid:1)(γ) = F(cid:0)r(γ)(cid:1)F(cid:0)s(α−1)(cid:1)b(γα−1)ξ(α) dλu(α) F(cid:0)r(α)(cid:1)b(γα−1)ξ(α) dλu(α). F(cid:0)r(α)(cid:1)ξ(α) dλz(α) z . This implies b(γα−1) = 1, so = G G In the case where u = z, if α, γ ∈ supp F ◦ r ∩ s−1(z), then r(α), r(γ) ∈ supp F = N and γ, α ∈ GN G = (ξ F ◦ r)zF ◦ r(γ), and Lz(D) is a rank-one projection. that By the hypothesis on V there exists a subsequence {xni} of {xn} such for all i (cid:62) 1. If we define E := {γ ∈ G : F(cid:0)r(γ)(cid:1) (cid:54)= 0} then E is open with (GV ) (cid:54) M λz(GV ) λxni λxni (E) (cid:54) λxni (GV ) (cid:54) M λz(GV ) (using (5.5)) (5.6) 14 and (cid:90) G ROBERT HAZLEWOOD AND ASTRID AN HUEF (cid:0)F ◦ r(γ)(cid:1)2 dλxni T (α, β) := F(cid:0)r(α)(cid:1)F(cid:0)r(β)(cid:1)b(αβ−1). (E) (cid:107)f ◦ r(cid:107)2 (γ) (cid:54) λxni z (cid:54) M λz(GV ) (cid:107)g(cid:107)2 z by (5.3). Consider the continuous function . (5.7) Note that(cid:90) G (cid:90) × λxni )(α, β) T (α, β)2 d(λxni F(cid:0)r(α)(cid:1)2F(cid:0)r(β)(cid:1)2b(αβ−1)2 d(λxni (cid:90) = (cid:54) (cid:107)F(cid:107)4∞ = (cid:107)F(cid:107)4∞λxni χE×E(α, β) d(λxni (E)2, × λxni G G )(α, β) × λxni )(α, β) which is finite by (5.6). Thus T ∈ L2(G × G, λxni × λxni ), and since T is conjugate symmetric, [20, Proposition 3.4.16] implies that Lxni (D) is the self-adjoint Hilbert-Schmidt operator on L2(G, λxni ) with kernel T . It follows that Lxni (D∗ ∗ D) is a trace-class operator, and since we equip the Hilbert-Schmidt operators with the trace norm, we have Tr Lxni (D∗ ∗ D) = (cid:107)T(cid:107)2 L2(λxni ×λxni ). Applying Fubini's Theorem to T now gives Tr Lxni (D∗ ∗ D) (cid:90) (cid:90) (cid:18)(cid:90) G G = (cid:54) F(cid:0)r(α)(cid:1)2F(cid:0)r(β)(cid:1)2b(αβ−1)2 dλxni F(cid:0)r(α)(cid:1)2 dλxni (cid:19)2 (α) G (α) dλxni (β) (cid:54) M 2λz(GV )2 (cid:107)g(cid:107)4 z (using (5.7)) (using (5.1)). < M 2(1 + )2 Now ML(Lz,{Lxn}) (cid:54) lim inf Tr(cid:0)Lxn(D∗ ∗ D)(cid:1) n (cid:54) M 2(1 + )2 < (cid:98)M 2(cid:99) + 1, (5.8) and hence ML(Lz,{Lxn}) (cid:54) (cid:98)M 2(cid:99), completing the proof. (cid:3) The following proposition is an immediate consequence of Theorem 5.2 and Proposition 4.4. This result will be strengthened later in Corollary 6.5, where we will show that these three items are in fact equivalent. STRENGTH OF CONVERGENCE IN THE ORBIT SPACE OF A GROUPOID 15 Proposition 5.3. Suppose G is a second-countable locally-compact Haus- dorff principal groupoid with Haar system λ. Let z ∈ G(0) and let {xn} be a sequence in G(0). Assume that [z] is locally closed in G(0). Consider the following properties. (1) ML(Lz,{Lxn}) = ∞. (2) For every open neighbourhood V of z such that GV z is relatively com- (3) For each k (cid:62) 1, the sequence {xn} converges k-times in G(0)/G to pact, λxn(GV ) → ∞ as n → ∞. z. Then (1) implies (2) and (2) implies (3). Our next goal is to sharpen the (cid:98)M 2(cid:99) bound in Theorem 5.2. This strengthened theorem appears later on as Theorem 5.8. We will first es- tablish several results to assist in strengthening this bound. Lemma 5.4. Suppose G is a second-countable groupoid and x, y ∈ G(0). If [x] = [y] and [x] is locally closed, then [x] = [y]. Proof. We have x ∈ [y], so there exists {γn} ⊂ G such that s(γn) = y and r(γn) → x. Since [x] is locally closed, there exists an open subset U of G such that [x] = U ∩ [x]. Then r(γn) is eventually in U , so eventually r(γn) ∈ U ∩ [y] = U ∩ [x] = [x]. Thus there exists γ ∈ G with s(γ) = y and r(γ) ∈ [x], as required. (cid:3) Lemma 5.5. Suppose G is a second-countable groupoid with Haar system λ. Let W be a compact neighbourhood of z ∈ G(0) and let K be a compact subset of G. Let {xn} be a sequence in G(0) such that [xn] → [z] uniquely in G(0)/G. Then for every δ > 0 there exists n0 such that, for every n (cid:62) n0 and every γ ∈ GW xn, λxn(Kγ ∩ GW ) < λz(GW ) + δ. Proof. Suppose not. Then, by passing to a subsequence if necessary, for each n there exists γn ∈ GW xn such that λxn(Kγn ∩ GW ) (cid:62) λz(GW ) + δ. (5.9) Since each r(γn) is in the compact set W , we can pass to a subsequence so that r(γn) → y for some y ∈ G(0). This implies [r(γn)] → [y], but [r(γn)] = [s(γn)] = [xn] and [xn] → [z] uniquely, so [y] = [z]. Choose ψ ∈ G with s(ψ) = z and r(ψ) = y. By Haar-system invariance λxn(Kγn ∩ GW ) = λr(γn)(K ∩ GW ), so by applying Lemma 4.2 with the compact space K ∩ GW and {r(γn)} converging to y, λxn(Kγn ∩ GW ) = lim sup λr(γn)(K ∩ GW ) lim sup n n (cid:54) λy(K ∩ GW ) = λz(Kψ ∩ GW ) (cid:54) λz(GW ). (by Lemma 4.2) (Haar-system invariance) (cid:3) This contradicts our assertion (5.9). 16 ROBERT HAZLEWOOD AND ASTRID AN HUEF The following is a generalisation of [3, Lemma 3.3]. Lemma 5.6. Suppose G is a groupoid with Haar system λ. Fix  > 0, z ∈ G(0) and let V be an open neighbourhood of z ∈ G(0) such that λz(GV ) < ∞. Then there exists an open relatively-compact neighbourhood V1 of z such that V1 ⊂ V and λz(GV ) −  < λz(GV1) (cid:54) λz(GV1) (cid:54) λz(GV ) < λz(GV1) + . Proof. We use Gz equipped with the subspace topology to find a compact subset λz-estimate of V . This estimate is then used to obtain the required open set V1. Since GV is Gz-open, by the regularity of λz there exists a z z ) − . Then r(W ) is compact subset W of GV z such that λz(W ) > λz(GV compact and contained in V , so there exists an open relatively-compact neighbourhood V1 of r(W ) such that V1 ⊂ V . Then λz(GV ) −  < λz(W ) (cid:54) λz(GV1) (cid:54) λz(GV1) (cid:54) λz(GV ) < λz(W ) +  (cid:54) λz(GV1) + , as required. (cid:3) The following lemma is equivalent to the claim in [10, Proposition 3.6] that [x] (cid:55)→ [Lx] from G(0)/G to the spectrum of C∗(G) is open. Lemma 5.7. Suppose G is a second-countable locally-compact Hausdorff groupoid with Haar system λ. If {xn} is a sequence in G(0) with Lxn → Lz, then [xn] → [z]. Proof. We prove the contrapositive. Suppose [xn] (cid:57) [z]. Then there exists an open neighbourhood U0 of [z] in G(0)/G such that [xn] is frequently not in U0. Let q : G(0) → G(0)/G be the quotient map x (cid:55)→ [x]. Then U1 := q−1(U0) is an open invariant neighbourhood of z and xn /∈ U1 frequently. Note that C∗(GU1) is isomorphic to a closed two-sided ideal I of C∗(G) (see [18, Lemma 2.10]). We now claim that I ⊂ ker Lxn whenever xn /∈ U1. Suppose xn /∈ U1 and recall from Remark 2.4 that Lxn acts on L2(G, λxn). Fix f ∈ Cc(G) such that f (γ) = 0 whenever γ /∈ GU1 and fix ξ ∈ L2(G, λxn). Then by Remark 2.4 we have (cid:90) (cid:18)(cid:90) (cid:19)2 f (γα−1)ξ(α) dλxn(α) When evaluating the inner integrand, we have s(α) = s(γ) = xn, so γα−1 ∈ G[xn]. Since U1 is invariant with xn /∈ U1, it follows that γα−1 /∈ GU1, and so f (γα−1) = 0. Thus (cid:107)Lxn(f )ξ(cid:107)2 xn = G G dλxn(γ). (cid:107)Lxn(f )ξ(cid:107)xn = 0, and since ξ was fixed arbitrarily, Lxn(f ) = 0. This implies that I ⊂ ker Lxn. We now conclude by observing that since I ⊂ ker Lxn frequently, Lxn /∈ I (cid:3) We may now proceed to strengthening the (cid:98)M 2(cid:99) bound in Theorem 5.2. frequently. But I is an open neighbourhood of Lz, so Lxn (cid:57) Lz. This theorem is a generalisation of [3, Theorem 3.5]. STRENGTH OF CONVERGENCE IN THE ORBIT SPACE OF A GROUPOID 17 Theorem 5.8. Suppose G is a second-countable locally-compact Hausdorff principal groupoid with Haar system λ. Let M ∈ R with M (cid:62) 1, suppose z ∈ G(0) such that [z] is locally closed and let {xn} be a sequence in G(0). Suppose there exists an open neighbourhood V of z in G(0) such that GV z is relatively compact and λxn(GV ) (cid:54) M λz(GV ) frequently. Then ML(Lz,{Lxn}) (cid:54) (cid:98)M(cid:99). Proof. If Lxn does not converge to Lz, then ML(Lz,{Lxn}) = 0 < (cid:98)M(cid:99). So we assume from now on that Lxn → Lz. Lemma 5.7 now shows that [xn] → [z]. Next we claim that we may assume, without loss of generality, that [z] is the unique limit of {[xn]} in G(0)/G. To see this, note that ML(Lz,{Lxn}) (cid:54) (cid:98)M 2(cid:99) < ∞ by Theorem 5.2. Hence, by [3, Proposition 3.4], {Lz} is open in the set of limits of {Lxn}. So there exists an open neighbourhood U2 of Lz in C∗(G)∧ such that Lz is the unique limit of {Lxn} in U2. By [19, Proposition 2.5] there is a continuous function L : G(0)/G → C∗(G)∧ such that [x] (cid:55)→ Lx for all x ∈ G(0). Define p : G(0) → G(0)/G by p(x) = [x] for all x ∈ G(0). Then p is continuous, and Y := (L ◦ p)−1(U2). is an open G-saturated neighbourhood of z in G(0). Note that xn ∈ Y eventually. Now suppose that, for some y ∈ Y , [xn] → [y] in Y /G and hence in G(0)/G. Then Lxn → Ly by [19, Proposition 2.5], and Ly ∈ U2 since y ∈ (L◦p)−1(U2). But {Lxn} has the unique limit Lz in U2, so Lz = Ly and hence [z] = [y]. Since [z] is locally closed, Lemma 5.4 shows that [z] = [y] in G(0) and hence in Y . We know Y is an open saturated subset of G(0), so C∗(GY ) is isomorphic to a closed two-sided ideal J of C∗(G). We can apply [8, Proposition 5.3] with the C∗-subalgebra J to see that ML(Lz,{Lxn}) is the same whether we z = GV ∩Y compute it in the ideal J or in C∗(G). Since Y is G-invariant, GV . We may thus consider GY instead of G and and eventually GV therefore assume that [z] is the unique limit of [xn] in G(0)/G as claimed. xn = GV ∩Y xn z As in [3], the idea for the rest of the proof is the same as in Theorem 5.2, although more precise estimates are used. Fix  > 0 such that M (1 + )2 < (cid:98)M(cid:99) + 1 and choose κ > 0 such that λz(GV ) κ < 1 +  < λz(GV ). (5.10) By Lemma 5.6 there exists an open relatively compact neighbourhood V1 of z such that V1 ⊂ V and 0 < λz(GV ) − κ < λz(GV1) (cid:54) λz(GV1) (cid:54) λz(GV ) < λz(GV1) + κ. Choose a subsequence {xni} of {xn} such that (GV ) (cid:54) M λz(GV ) λxni 18 ROBERT HAZLEWOOD AND ASTRID AN HUEF for all i (cid:62) 1. Then λxni (GV1) (cid:54) λxni (GV ) (cid:54) M λz(GV ) < M(cid:0)λz(GV1) + κ(cid:1) < M λz(GV1) + M (cid:0)λz(GV ) − κ(cid:1) < M λz(GV1) + M λz(GV1) = M (1 + )λz(GV1) for all i. Since λz(GV1)(cid:0)λz(GV1) + κ + 1/j(cid:1) (cid:0)λz(GV1) − 1/j(cid:1)2 λz(GV1) as j → ∞, there exists δ > 0 such that δ < λz(GV1) and λz(GV1)(cid:0)λz(GV1) + δ(cid:1) (cid:0)λz(GV1) − δ(cid:1)2 λz(GV1)(cid:0)λz(GV1) + κ + δ(cid:1) (cid:0)λz(GV1) − δ(cid:1)2 < → 1 + κ (by (5.10)) (5.11) < 1 +  < 1 + . (5.12) We will now construct a function F ∈ Cc(G(0)) with support inside V1. is Gz-open, there exists a Since λz is inner regular on open sets and GV1 z Gz-compact subset W of GV1 z such that 0 < λz(GV1 z ) − δ < λz(W ). Since W is Gz-compact there exists a Gz-compact neighbourhood W1 of W z and there exists a continuous function g : Gz → [0, 1] that is contained in GV1 that is identically one on W and zero off the interior of W1. We have λz(GV1) − δ < λz(W ) (cid:54) g(t)2 dλz(t) = (cid:107)g(cid:107)2 z, (cid:90) Gz (5.13) (cid:0)r(γ)(cid:1) = By Lemma 5.1 the restriction r of r to Gz is a homeomorphism onto [z]. So there exists a continuous function g1 : r(W1) → [0, 1] such that g1 g(γ) for all γ ∈ W1. Thus r(W1) is [z]-compact, which implies that r(W1) is G(0)-compact. Since we know that G(0) is second countable and Hausdorff, Tietze's Extension Theorem can be applied to show that g1 can be extended to a continuous map g2 : G(0) → [0, 1]. Because r(W1) is a compact subset of the open set V1, there exist a compact neighbourhood P of r(W1) contained in V1 and a continuous function h : G(0) → [0, 1] that is identically one on r(W1) and zero off the interior of P . Note that h has compact support that is contained in P . We set f (x) = h(x)g2(x). Then f ∈ Cc(G(0)) with 0 (cid:54) f (cid:54) 1 and suppf ⊂ supph ⊂ P ⊂ V1. (5.14) STRENGTH OF CONVERGENCE IN THE ORBIT SPACE OF A GROUPOID 19 Note that (cid:107)f ◦ r(cid:107)2 z = = (cid:62) f(cid:0)r(γ)(cid:1)2 dλz(γ) h(cid:0)r(γ)(cid:1)2g2 (cid:0)r(γ)(cid:1)2 dλz(γ) h(cid:0)r(γ)(cid:1)2g(γ)2 dλz(γ) (cid:90) (cid:90) (cid:90) (cid:90) Gz Gz W1 g(γ)2 dλz(γ) = = (cid:107)g(cid:107)2 W1 (5.15) since supp g ⊂ W1 and h is identically one on r(W1). We now define F ∈ Cc(G(0)) by z F (x) = f (x) (cid:107)f ◦ r(cid:107)z . Then (cid:107)F ◦ r(cid:107)z = 1 and F ◦ r(γ) (cid:54)= 0 =⇒ h(cid:0)r(γ)(cid:1) (cid:54)= 0 =⇒ r(γ) ∈ V1 =⇒ γ ∈ GV1. (5.16) (5.17) z )(GN z )(GN Let N = supp F . Suppose K is an open relatively compact symmetric z )−1 in G and choose b ∈ Cc(G) such that b is neighbourhood of (GN z )−1 and identically zero off K. As in Theorem identically one on (GN 2 (b + b∗). Define 5.2 we may assume that b is self-adjoint by considering 1 in Theorem 5.2, Lz(D), and hence Lz(D∗ ∗ D), is the rank one projection determined by the unit vector F ◦ r ∈ L2(G, λz). From (5.8) we have D ∈ Cc(G) by D(γ) := F(cid:0)r(γ)(cid:1)F(cid:0)s(γ)(cid:1)b(γ). By the same argument as Tr(cid:0)Lxni (D∗ ∗ D)(cid:1) (cid:90) (cid:18)(cid:90) F(cid:0)r(β)(cid:1)2 = G F(cid:0)r(α)(cid:1)2b(αβ−1)2 dλxni G Since b is identically zero off K, the inner integrand is zero unless αβ−1 ∈ K. enables us Combining this with (5.14) and the fact that supp λxni to see that this inner integrand is zero unless α ∈ GV1 xni (α) dλxni (β). (cid:19) Tr(cid:0)Lxni (D∗ ∗ D)(cid:1) ⊂ Gxni ∩ Kβ. Thus (cid:19) (α) dλxni (β). F(cid:0)r(α)(cid:1)2 dλxni (cid:19) (cid:90) (cid:54) (cid:54) β∈GV1 xni 1 (cid:107)f ◦ r(cid:107)4 z F(cid:0)r(β)(cid:1)2 (cid:90) (cid:18)(cid:90) (cid:18)(cid:90) 1 β∈GV1 xni α∈GV1 xni ∩Kβ 1 dλxni (α) dλxni (β). α∈GV1 xni ∩Kβ Since V1 and K are compact, by Lemma 5.5 there exists i0 such that for every i (cid:62) i0 and any β ∈ GV1 xni (Kβ ∩ GV1) < λz(GV1) + δ. , λxni So, provided i (cid:62) i0, Tr(cid:0)Lxni (D∗ ∗ D)(cid:1) (cid:54) (cid:90) 1 (cid:107)f ◦ r(cid:107)4 z β∈GV1 xni (Kβ ∩ GV1 xni λxni ) dλxni (β) 20 ROBERT HAZLEWOOD AND ASTRID AN HUEF (β) (cid:90) 1 z β∈GV1 xni (cid:107)f ◦ r(cid:107)4 (cid:0)λz(GV1 z ) + δ(cid:1) dλxni (cid:0)λz(GV1) + δ(cid:1)λxni M (1 + )(cid:0)λz(GV1) + δ(cid:1)λz(GV1) M (1 + )(cid:0)λz(GV1) + δ(cid:1)λz(GV1) (cid:107)f ◦ r(cid:107)4 (GV1) z (cid:107)g(cid:107)4 z (cid:54) < < < (λz(GV1) − δ)2 (by (5.12)). < M (1 + )2 (by (5.13)) (by (5.11) and (5.15)) lower semi-continuity [9, Theorem 4.3], We can now make our conclusion as in [3, Theorem 3.5]: by generalised Tr(cid:0)Lxn(D∗ ∗ D)(cid:1) (cid:62) ML(Lz,{Lxn}) Tr(cid:0)Lz(D∗ ∗ D)(cid:1) lim inf n We now have ML(Lz,{Lxn}) (cid:54) lim inf n (cid:54) M (1 + )2 < (cid:98)M(cid:99) + 1, and so ML(Lz,{Lxn}) (cid:54) (cid:98)M(cid:99), as required. = ML(Lz,{Lxn}). Tr(cid:0)Lxn(D∗ ∗ D)(cid:1) (cid:3) 6. Lower multiplicity and k-times convergence II We proved in Proposition 3.2 that if a sequence converges k-times in the orbit space of a principal groupoid, then the lower multiplicity of the associated sequence of representations is at least k. In this section we will prove the converse. Our next lemma generalises [3, Lemma 5.1]; with the exception of notation changes, the proof is the same as the proof in [3]. Lemma 6.1. Suppose G is a second-countable locally-compact Hausdorff principal groupoid. Let k ∈ P, z ∈ G(0), and {xn} be a sequence in G(0). Assume that [z] is locally closed in G(0) and that there exists R > k − 1 such that for every open neighbourhood U of z with GU z relatively compact we have λxn(GU ) (cid:62) Rλz(GU ). Given an open neighbourhood V of z such that GV there exists a compact neighborhood N of z with N ⊂ V such that z lim inf n is relatively compact, λxn(GN ) > (k − 1)λz(GN ). lim inf n Proof. Apply Lemma 5.6 to V with 0 <  < R−k+1 relatively-compact neighbourhood V1 of z with V1 ⊂ V and R λz(GV ) to get an open λz(GV ) −  < λz(GV1) (cid:54) λz(GV1) (cid:54) λz(GV ) < λz(GV1) + . STRENGTH OF CONVERGENCE IN THE ORBIT SPACE OF A GROUPOID 21 Since GV1 z is relatively compact we have lim inf n λxn(GV1) (cid:62) lim inf λxn(GV1) n (cid:62) Rλz(GV1) > R(cid:0)λz(GV ) − (cid:1) > (k − 1)λz(GV ) (cid:62) (k − 1)λz(GV1). (by hypothesis) (by our choice of ) So we may take N = V1. (cid:3) Remark 6.2. The preceding Lemma also holds when lim inf is replaced by lim sup. No modification of the proof is needed beyond replacing the two occurrences of lim inf with lim sup. We may now proceed to our main theorem. Theorem 6.3. Suppose G is a second-countable locally-compact Hausdorff principal groupoid that admits a Haar system λ. Let k be a positive integer, let z ∈ G(0) and let {xn} be a sequence in G(0). Assume that [z] is locally closed in G(0). Then the following are equivalent: (1) the sequence {xn} converges k-times in G(0)/G to z; (2) ML(Lz,{Lxn}) (cid:62) k; (3) for every open neighbourhood V of z in G(0) such that GV z is relatively compact we have λxn(GV ) (cid:62) kλz(GV ); lim inf n (4) there exists a real number R > k − 1 such that for every open neigh- bourhood V of z in G(0) with GV z relatively compact we have λxn(GV ) (cid:62) Rλz(GV ); and lim inf n (5) there exists a basic decreasing sequence of compact neighbourhoods {Wm} of z in G(0) such that, for each m (cid:62) 1, λxn(GWm) > (k − 1)λz(GWm). lim inf n Proof. We know that (1) implies (2) by Proposition 3.2. Suppose (2). If ML(Lz,{Lxn}) (cid:62) k, then ML(Lz,{Lxn}) > (cid:98)k − (cid:99) for all  > 0. By Theorem 5.8, for every G(0)-open neighborhood V of z such that GV z is relatively compact, λxn(GV ) > (k − )λz(GV ) eventually, and hence (3) holds. It is immediately true that (3) implies (4). Suppose (4). We will construct the sequence {Wm} of compact neigh- bourhoods inductively. Let {Vj} be a basic decreasing sequence of open neighborhoods of z such that GV1 is relatively compact (such neighborhoods z exist by [12, Lemma 4.1(1)]). By Lemma 6.1 there exists a compact neigh- bourhood W1 of z such that W1 ⊂ V1 and λxn(GW1) > (k − 1)λz(GW1). 22 ROBERT HAZLEWOOD AND ASTRID AN HUEF Now assume there are compact neighbourhoods W1, W2, . . . , Wm of z with W1 ⊃ W2 ⊃ ··· ⊃ Wm such that Wi ⊂ Vi and λxn(GWi) > (k − 1)λz(GWi) (6.1) for all 1 (cid:54) i (cid:54) m. Apply Lemma 6.1 to (Int m) ∩ Vm+1 to obtain a compact neighbourhood Wm+1 of z such that Wm+1 ⊂ (Int Wm) ∩ Vm+1 and (6.1) holds for i = m + 1, establishing (5). Suppose (5). We begin by showing that [xn] → [z] in G(0)/G. Let q : G(0) → G(0)/G be the quotient map. Let U be a neighbourhood of [z] in G(0)/G and V = q−1(U ). There exists m such that Wm ⊂ V . Since (cid:54)= ∅ for all n (cid:62) n0. lim inf n λxn(GWm) > 0 there exists n0 such that GWm xn Thus, for n (cid:62) n0, [xn] = q(xn) ∈ q(Wm) ⊂ q(V ) = U. Thus [xn] is eventually in every neighbourhood of [z] in G(0)/G. Now suppose that ML(Lz,{Lxn}) < ∞. Then, as in the proof of Theorem 5.8, we may localise to an open invariant neighbourhood Y of z such that [z] is the unique limit in Y /G of [xn]. Eventually Wm ⊂ Y , and so the sequence {xn} converges k-times in Y /(GY ) = Y /G to z by Proposition 4.1 applied to the groupoid GY . This implies that the sequence {xn} converges k-times in G(0)/G. Finally, if ML(Lz,{Lxn}) = ∞, then {xn} converges k-times in G(0)/G to (cid:3) z by Proposition 5.3, establishing (1) and completing the proof. Corollary 6.4. Suppose that G is a second-countable locally-compact Haus- dorff principal groupoid such that all the orbits are locally closed. Let k ∈ P and let z ∈ G(0) such that [z] is not open in G(0). Then the following are equivalent: (1) whenever {xn} is a sequence in G(0) which converges to z with [xn] (cid:54)= (2) ML(Lz) (cid:62) k. [z] eventually, then {xn} is k-times convergent in G(0)/G to z; Proof. Assume (1). We must first establish that {Lz} is not open in C∗(G)∧. If this is not the case, then {Lz} is open and we can apply [10, Proposi- tion 3.6] to see that {[z]} is open in G(0)/G, and so [z] is open in G(0), contradicting our assumption. Since {Lz} is not open in C∗(G)∧, we can apply [3, Lemma A.2] to see that there exists a sequence {πi} of irreducible representations of C∗(G) such that each πi is not unitarily equivalent to Lz, πi → Lz in C∗(G)∧, and ML(Lz) = ML(Lz,{πi}) = MU(Lz,{πi}). (6.2) Since the orbits are locally closed, the map G(0)/G → C∗(G)∧ such that [x] (cid:55)→ Lx is a homeomorphism by [10, Proposition 5.1]1. It follows that the mapping G(0) → C∗(G)∧ such that x (cid:55)→ Lx is an open surjection, so by [24, 1Proposition 5.1 in [10] states that if a principal groupoid has locally closed orbits, then the map from G(0)/G to C∗(G)∧ where [x] (cid:55)→ Lx is a 'homeomorphism from G(0)/G into C∗(G)∧'. The proof explicitely shows that this map is a surjection. STRENGTH OF CONVERGENCE IN THE ORBIT SPACE OF A GROUPOID 23 Proposition 1.15] there is a sequence {xn} in G(0) such that xn → z and {Lxn} is unitarily equivalent to a subsequence of {πi}. By (6.2), ML(Lz) = MU(Lz,{πi}) (cid:62) MU(Lz,{Lxn}) (cid:62) ML(Lz,{Lxn}). We know by (1) that {xn} converges k-times to z in G(0)/G, so it follows from Theorem 6.3 that ML(Lz) (cid:62) ML(Lz,{Lxn}) (cid:62) k. Assume (2). If {xn} is a sequence in G(0) which converges to z such that [xn] (cid:54)= [z] eventually, then ML(Lz,{Lxn}) (cid:62) ML(Lz) (cid:62) k. By Theorem 6.3, {xn} is k-times convergent to z in G(0)/G. (cid:3) The next corollary improves Proposition 5.3 and is an immediate conse- quence of Proposition 5.3 and Theorem 6.3. Corollary 6.5. Suppose that G is a second-countable locally-compact Haus- dorff principal groupoid with Haar system λ. Let z ∈ G(0) and let {xn} be a sequence in G(0). Assume that [z] is locally closed. Then the following are equivalent: (1) ML(Lz,{Lxn}) = ∞. (2) For every open neighbourhood V of z such that GV z is relatively com- (3) For each k (cid:62) 1, the sequence {xn} converges k-times in G(0)/G to pact, λxn(GV ) → ∞ as n → ∞. z. 7. Upper multiplicity and k-times convergence The results in this section are corollaries of Theorems 5.8 and 6.3: they relate k-times convergence, measure ratios and upper multiplicity numbers, generalising all the upper-multiplicity results of [3]. We begin with the upper-multiplicity analogue of Theorem 5.8. Theorem 7.1. Suppose that G is a second-countable locally-compact Haus- dorff principal groupoid with Haar system λ. Let M ∈ R with M (cid:62) 1, let z ∈ G(0) and let {xn} be a sequence in G(0). Assume that [z] is locally closed. Suppose that there exists an open neighbourhood V of z in G(0) such that GV z is relatively compact and λxn(GV ) (cid:54) M λz(GV ) < ∞ eventually. Then MU(Lz,{Lxn}) (cid:54) (cid:98)M(cid:99). Proof. Since G is second countable, C∗(G) is separable. By [3, Lemma A.1] there exists a sequence {Lxni} such that MU(Lz,{Lxn}) = MU(Lz,{Lxni}) = ML(Lz,{Lxni}). By Theorem 5.8, ML(Lz,{Lxni}) (cid:54) (cid:98)M(cid:99), so MU(Lz,{Lxn}) (cid:54) (cid:98)M(cid:99). Corollary 7.2. Suppose that G is a second-countable locally-compact Haus- dorff principal groupoid with Haar system λ such that all the orbits are locally closed. Let M ∈ R with M (cid:62) 1 and let z ∈ G(0). If for every sequence {xn} (cid:3) 24 ROBERT HAZLEWOOD AND ASTRID AN HUEF in G(0) which converges to z there exists an open neighbourhood V of z in G(0) such that GV z is relatively compact and λxn(GV ) (cid:54) M λz(GV ) < ∞ frequently, then MU(Lz) (cid:54) (cid:98)M(cid:99). Proof. Since G is second countable, C∗(G) is separable, and so we can apply [5, Lemma 1.2] to see that there exists a sequence {πn} in C∗(G)∧ that converges to Lz such that ML(Lz,{πn}) = MU(Lz,{πn}) = MU(Lz). Since the orbits are locally closed, the map G(0)/G → C∗(G)∧ such that [x] (cid:55)→ Lx is a homeomorphism by [10, Proposition 5.1]. In particular, the mapping G(0) → C∗(G)∧ such that x (cid:55)→ Lx is an open surjection, so by [24, Proposition 1.15] there exists a sequence {xi} in G(0) converging to z such that {[Lxi]} is a subsequence of {[πn]}. By Theorem 5.8, ML(Lz,{Lxn}) (cid:54) (cid:98)M(cid:99). Since MU(Lz) = ML(Lz,{πn}) (cid:54) ML(Lz,{Lxi}) (cid:54) MU(Lz,{Lxi}) (cid:54) MU(Lz,{πn}) = MU(Lz), we obtain MU(Lz) (cid:54) (cid:98)M(cid:99), as required. (cid:3) In Proposition 4.1 we generalised the first part of [3, Proposition 4.1]. We will now generalise the second part. The argument we use is similar to that used in Proposition 4.1. Proposition 7.3. Let G be a second-countable locally-compact Hausdorff principal groupoid with Haar system λ. Let k ∈ P and z ∈ G(0) with [z] locally closed in G(0). Assume that {xn} is a sequence in G(0) such that [xn] → [z] uniquely in G(0)/G. Suppose {Wm} is a basic decreasing sequence of compact neighbourhoods of z such that each m satisfies λxn(GWm) > (k − 1)λz(GWm). lim sup Then there exists a subsequence of {xn} which converges k-times in G(0)/G to z. Proof. Let {Km} be an increasing sequence of compact subsets of G such m(cid:62)1 Int Km. By the regularity of λz, for each m (cid:62) 1 there exist that G =(cid:83) n δm > 0 and an open neighbourhood Um of GWm such that λxn(GWm) > (k − 1)λz(Um) + δm. z lim sup n (7.1) We will construct, by induction, a strictly increasing sequence of positive integers {im} such that, for all m, λxim (Kmα ∩ GWm) < λz(Um) + δm/k for all α ∈ GWm λxim (GWm) > (k − 1)λz(Um) + δm. (7.3) By Lemma 5.5 with δ = λz(U1) − λz(GW1) + δ1/k, there exists n1 such (7.2) and , xim that n (cid:62) n1 implies λxn(K1α ∩ GW1) < λz(U1) + δ1/k for all α ∈ GWm xn . STRENGTH OF CONVERGENCE IN THE ORBIT SPACE OF A GROUPOID 25 By considering (7.1) with m = 1 we can choose i1 (cid:62) n1 such that (GW1) > (k − 1)λz(U1) + δ1. λxi1 Assuming that i1 < i2 < ··· < im−1 have been chosen, we can apply Lemma 5.5 with δ = λz(Um) − λz(GWm) + δm/k to obtain nm > im−1 such that n (cid:62) nm implies λxn(Kmα ∩ GWm) < λz(Um) + δm/k for all α ∈ GWm xn , and then by (7.1) we can choose im (cid:62) nm such that λxim (GWm) > (k − 1)λz(Um) + δm. For each m ∈ P choose γ(1) ∈ GWm xim im (which is non-empty by (7.3)). By (7.2) and (7.3) we have λxim (GWm\Kmγ(1) im ) = λxim (GWm) − λxim (GWm ∩ Kmγ(1) > (k − 1)λz(Um) + δm −(cid:0)λz(Um) + δm/k(cid:1) im ) = (k − 2)λz(Um) + k − 1 k δm. ∈ GWm xim \Kmγ(1) . This implies, as in the proof of im ∪ Kmγ(2) ∈ GWm xim γ(j) im ∈ GWm xim \ im )(cid:1) > (k − 3)λz(Um) + (cid:18) j−1(cid:91) ∩ Kmγ(2) (cid:19) im Kmγ(l) im . l=1 (k − 2) k δm, ). Continuing in this (7.4) So we can choose γ(2) im Proposition 4.1, that (cid:0)GWm\(Kmγ(1) im λxim \(Kmγ(1) enabling us to choose γ(3) im im way for j = 3, . . . , k, for each im we choose Note that γ(j) im /∈ Kmγ(l) im We claim that r(γ(l) im for 1 (cid:54) l < j (cid:54) k. ) → z as m → ∞ for 1 (cid:54) l (cid:54) k. To see this, fix l and let V be an open neighbourhood of z. Since {Wm} is a decreasing neighbourhood basis for z there exists m0 such that m (cid:62) m0 implies Wm ⊂ V , and so r(γ(l) im )−1 → ∞ as m → ∞ for 1 (cid:54) l < j (cid:54) k. Fix l < j and let K be a compact subset of G. There exists m0 such that K ⊂ Km for all m (cid:62) m0. By (7.4) we know ) ) ∈ Wm ⊂ V . Finally we claim that γ(j) im (γ(l) im γ(j) im im ∈ GWm xim =(cid:0)GWm =(cid:0)(GWm \(Kmγ(l) (γ(l) im (γ(l) im )−1γ(l) )−1)\Km (γ(l) im )−1 ∈(cid:0)GWm (cid:1)\(Kmγ(l) (cid:1)γ(l) )−1(cid:1)\Km ⊂ G\Km ⊂ G\K, xim xim im im im ) . im (γ(l) im So provided m (cid:62) m0, γ(j) enabling us to conclude that {xim} converges k-times in G(0)/G to z. Theorem 7.4. Suppose that G is a second-countable locally-compact Haus- dorff principal groupoid with Haar system λ. Let k ∈ P, let z ∈ G(0), and let {xn} be a sequence in G(0) such that [xn] converges to [z] in G(0)/G. Assume that [z] is locally closed. Then the following are equivalent: xim (cid:3) pact we have lim sup n λxn(GV ) (cid:62) kλz(GV ); 26 ROBERT HAZLEWOOD AND ASTRID AN HUEF (1) there exists a subsequence {xni} of {xn} which converges k-times in (2) MU(Lz,{Lxn}) (cid:62) k; (3) for every open neighbourhood V of z such that GV G(0)/G to z; z is relatively com- (4) there exists a real number R > k − 1 such that for every open neigh- bourhood V of z in G(0) with GV z relatively compact we have λxn(GV ) (cid:62) Rλz(GV ); and lim sup n (5) there exists a basic decreasing sequence of compact neighbourhoods {Wm} of z in G(0) such that, for each m (cid:62) 1, λxn(GWm) > (k − 1)λz(GWm). lim sup n Proof. If (1) holds then ML(Lz,{Lxni}) (cid:62) k by Theorem 6.3, and so MU(Lz,{Lxn} (cid:62) MU(Lz,{Lxni}) (cid:62) ML(Lz,{Lxni}) (cid:62) k. If (2) holds then by [3, Lemma A.1] there is a subsequence {xnr} such that ML(Lz,{Lxnr}) = MU(Lz,{Lxn}) so that ML(Lz,{Lxnr}) (cid:62) k. Let V be any open neighbourhood of z in G(0) such that GV z is relatively compact. Then lim sup n λxn(GV ) (cid:62) lim sup r λxnr (GV ) (cid:62) lim inf r λxnr (GV ) (cid:62) kλz(GV ), using Theorem 6.3 for the last step. That (3) implies (4) is immediate. That (4) implies (5) follows by making references to Remark 6.2 rather than Lemma 6.1 in the (4) implies (5) component of the proof of Theorem 6.3. Assume (5). First suppose that ML(Lz,{Lxn}) < ∞. Since [xn] → [z], we can use an argument found at the beginning of the proof of Theorem 5.8 to obtain an open G-invariant neighborhood Y of z in G(0) so that if we define H := GY , there exists a subsequence {xni} of {xn} such that [xni] → [z] uniquely in H (0)/H. Proposition 7.3 now shows us that there exists a subsequence {xnij } of {xni} that converges k-times in H (0)/H to z. It follows that {xnij } converges k-times in G(0)/G to z. When ML(Lz,{Lxn}) = ∞, {xn} converges k-times in G(0)/G to z by (cid:3) Corollary 6.5, establishing (1). Corollary 7.5. Suppose that G is a second-countable locally-compact Haus- dorff principal groupoid such that all the orbits are locally closed. Let k ∈ P and let z ∈ G(0). Then the following are equivalent: (1) there exists a sequence {xn} in G(0) which is k-times convergent in G(0)/G to z; (2) MU(Lz) (cid:62) k. STRENGTH OF CONVERGENCE IN THE ORBIT SPACE OF A GROUPOID 27 Proof. Assume (1). By the definitions of upper and lower multiplicity, MU(Lz) (cid:62) MU(Lz,{Lxn}) (cid:62) ML(Lz,{Lxn}). By Theorem 6.3 we know that ML(Lz,{Lxn}) (cid:62) k, establishing (2). Assume (2). By [5, Lemma 1.2] there exists a sequence {πn} converging to Lz such that ML(Lz,{πn}) = MU(Lz,{πn}) = MU(Lz). Since the orbits are locally closed, by [10, Proposition 5.1] the mapping G(0) → C∗(G)∧ : x (cid:55)→ Lx is a surjection. So there is a sequence {Lxn} in C∗(G)∧ such that Lxn is unitarily equivalent to πn for each n. Then ML(Lz,{Lxn}) (cid:62) ML(Lz,{πn}) = MU(Lz) (cid:62) k, and it follows from Theorem 6.3 that {xn} is k-times convergent in G(0)/G (cid:3) to z. Corollary 7.6. Suppose that G is a secound-countable locally-compact Haus- dorff principal groupoid with Haar system λ. Let z ∈ G(0) and let {xn} ⊂ G(0) be a sequence converging to z. Assume that [z] is locally closed. Then the following are equivalent: (1) there exists an open neighbourhood V of z such that GV z is relatively compact and (2) MU(Lz,{Lxn}) < ∞. n lim sup λxn(GV ) < ∞; Proof. Suppose that (1) holds. Since C∗(G) is separable, it follows from [3, Lemma A.1] that there exists a subsequence {xnj} of {xn} such that ML(Lz,{Lxnj}) = MU(Lz,{Lxnj}) = MU(Lz,{Lxn}). By (1) and Corollary 6.5, ML(Lz,{Lxn}) < ∞. Hence MU(Lz,{Lxn}) < ∞, as required. Suppose that (1) fails. Let {Vi} be a basic decreasing sequence of open is relatively compact (such neighborhoods neighbourhoods of z such that GV1 z exist by [12, Lemma 4.1(1)]). Then λxn(GVi) = ∞ for each i lim sup n and we may choose a subsequence {xni} of {xn} such that λxni as i → ∞. There exists i0 such that Vi ⊂ V for all i (cid:62) i0. Then, for i (cid:62) i0, Let V be any open neighbourhood of z such that GV z is relatively compact. (GVi) → ∞ (GVi) (cid:54) λxni λxni (GV ). (GV ) → ∞ as i → ∞. By Corollary 6.5, ML(Lz,{Lxn}) = ∞. (cid:3) Thus λxni Hence MU(Lz,{Lxn}) = ∞, that is (2) fails. Corollary 7.7. Suppose G is a second-countable locally-compact Hausdorff principal groupoid with Haar system λ such that all the orbits are locally closed. Let z ∈ G(0). Then the following are equivalent: (1) MU(Lz) < ∞; 28 ROBERT HAZLEWOOD AND ASTRID AN HUEF (2) there exists an open neighbourhood V of z such that GV z is relatively compact and λx(GV ) < ∞. sup x∈V Proof. If (2) holds then (1) holds by Corollary 7.2. Let {Vi} be a basic decreasing sequence of open neighbourhoods of z such is relatively compact. If (2) fails then supx∈Vi{λx(GVi)} = ∞ for that GV1 each i and we may choose a sequence {xi} such that xi ∈ Vi for all i and z λxi(GVi) → ∞. Since {Vi} is a basic decreasing sequence, xi → z. There exists i0 such that Vi ⊂ V for all i (cid:62) i0. Then, for i (cid:62) i0, Let V be an open neighbourhood of z such that GV z is relatively compact. λxi(GVi) (cid:54) λxi(GV ). Thus λxi(GV ) → ∞. By Corollary 7.6, MU(Lz,{Lxi}) = ∞. Hence MU(Lz) = ∞, and so (1) fails. (cid:3) 8. Graph algebra examples We begin this section by introducing the notion of a directed graph as well as some related concepts as in the expository book [21], although some notation is also taken from [15]. A directed graph E = (E0, E1, r, s) consists of two countable sets E0, E1 and functions r, s : E1 → E0. The elements of E0 and E1 are called vertices and edges respectively. For each edge e, call s(e) the source of e and r(e) the range of e. A directed graph E is row finite if r−1(v) is finite for every v ∈ E0. A finite path in a directed graph E is a finite sequence α = α1α2 ··· αk of edges αi with s(αj) = r(αj+1) for 1 (cid:54) j (cid:54) k − 1; write s(α) = s(αk) and r(α) = r(α1), and call α := k the length of α. An infinite path x = x1x2 ··· is defined similarly, although s(x) remains undefined. Let E∗ and E∞ denote the set of all finite paths and infinite paths in E respectively. If α = α1 ··· αk and β = β1 ··· βj are finite paths then, provided s(α) = r(β), let αβ be the path α1 ··· αkβ1 ··· βj. When x ∈ E∞ with s(α) = r(x) define αx similarly. A cycle is a finite path α of non-zero length such that s(α) = r(α). When v is a vertex, f is an edge, and there is exactly one infinite path with range v that includes the edge f , then we denote this infinite path by [v, f ]∞. When there is exactly one finite path α with r(α) = v and αα = f , we denote α by [v, f ]∗. In [15] two paths x, y ∈ E∞ are defined to be shift equivalent with lag k ∈ Z (written x ∼k y) if there exists N ∈ N such that xi = yi+k for all i (cid:62) N . Suppose E is a row-finite directed graph. We refer to the groupoid con- structed from E by Kumjian, Pask, Raeburn and Renault in [15] as the path groupoid. Before describing this construction we caution that we are using the now standard notation for directed graphs which has the range and source swapped from the notation used in [15]. This new convention is due to the development of the higher-rank graphs, where edges become morphisms in a category and the new convention ensures that "composition of morphisms is compatible with multiplication of operators in B(H)" [21, p. 2]. The path groupoid G = GE constructed from E is defined as follows: G := {(x, k, y) ∈ E∞ × Z × E∞ : x ∼k y}. STRENGTH OF CONVERGENCE IN THE ORBIT SPACE OF A GROUPOID 29 For elements of G(2) := {(cid:0)(x, k, y), (y, l, z)(cid:1) : (x, k, y), (y, l, z) ∈ G}, Kumjian, Pask, Raeburn, and Renault defined (x, k, y) · (y, l, z) := (x, k + l, z), and for arbitrary (x, k, y) ∈ G, defined (x, k, y)−1 := (y,−k, x). For each α, β ∈ E∗ with s(α) = s(β), let Z(α, β) be the set {(x, k, y) : x ∈ Z(α), y ∈ Z(β), k = β − α, xi = yi+k for i > α}. By [15, Proposition 2.6], the collection of sets {Z(α, β) : α, β ∈ E∗, s(α) = s(β)} is a basis of compact open sets for a second-countable locally-compact Haus- dorff topology on G that makes G r-discrete. Kumjian, Pask, Raeburn and Renault equipped G with the Haar system consisting of counting measures, which they observe is possible by first showing that a Haar system exists for the groupoid with [23, Proposition I.2.8], and then using [23, Lemma I.2.7] to show that they can choose the system of counting measures. By [15, Corollary 2.2], the cylinder sets Z(α) := {x ∈ E∞ : x1 = α1, . . . , xα = αα} parameterised by α ∈ E∗ form a basis of compact open sets for a locally- compact σ-compact totally-disconnected Hausdorff topology on E∞. After identifying each (x, 0, x) ∈ G(0) with x ∈ E∞, [15, Proposition 2.6] tells us that the topology on G(0) is identical to the topology on E∞. For a row-finite directed graph E, Kumjian, Pask, Raeburn and Renault use the path groupoid G to construct the usual groupoid C∗-algebra C∗(G), and show how a collection of partial isometries subject to some relations derived from E generate C∗(G). More recently, a C∗-algebra C∗(E) is con- structed from a collection of partial isometries subject to slightly weakened relations derived from E. The slightly weakened relations permit non-zero partial isometries to be related to sources in the graph, and as a result C∗(E) is isomorphic to C∗(G) only when E contains no sources. It turns out that C∗(E) and C∗(G) can be substantially different: an example in [14] de- scribes a graph with sources where C∗(G) has continuous trace while C∗(E) does not. In this paper we are only interested in groupoid C∗-algebras, so we will make no further mention of the graph algebra C∗(E). Since we wish to apply Theorem 6.3 to path groupoids, we must be able to show that the path groupoids we consider are principal. Proposition 8.1. Suppose E is a row-finite directed graph. The path groupoid G constructed from E is principal if and only if E contains no cycles. Proof. We first show that if E contains no cycles then G is principal. Sup- pose G is not principal. Then there exist x, y ∈ E∞ and distinct γ, δ ∈ G such that r(γ) = r(δ) = x and s(γ) = s(δ) = y. It follows that there exist a, b ∈ Z such that γ = (x, a, y) and δ = (x, b, y). Notice that since γ (cid:54)= δ, a (cid:54)= b. We may assume without loss of generality that a > b. 30 ROBERT HAZLEWOOD AND ASTRID AN HUEF Now γ = (x, a, y) implies x ∼a y and δ = (x, b, y) implies x ∼b y, so there exists N such that n (cid:62) N =⇒ xn = yn+a = yn+b, and so xn = yn+a = yn+a−b+b = xn+a−b. Thus E contains a cycle of length at most a − b. We now show that if G is principal then E contains no cycles. Suppose E contains the cycle α = α1α2 ··· αk. Then x := αα··· is in E∞ with x ∼k x, so both (x, 0, x) and (x, k, x) are in G. It follows that G is not principal. (cid:3) Example 8.2 (2-times convergence in a path groupoid). Let E be the graph v1 v2 v3 v4 f (1) 1 f (2) 1 f (1) 2 f (2) 2 f (1) 3 f (2) 3 f (1) 4 f (2) 4 n ) = x(n) = s(γ(2) n := (x(n), 0, x(n)) and γ(2) n ) for all n and that both r(γ(1) n :=(cid:0)[v1, f (2) n ]∞, 0, x(n)(cid:1). It follows n ) converge to z as n → ∞. It remains to show that γ(2) n ]∞ and let G be the path groupoid. For each n (cid:62) 1 define x(n) := [v1, f (1) and let z be the infinite path with range v1 that passes through each vn. Then {x(n)} converges 2-times in G(0)/G to z. Proof. We will describe two sequences in G as in Definition 3.1. For each n (cid:62) 1 define γ(1) immediately that s(γ(1) n ) and n )−1 → ∞ r(γ(2) as n → ∞. n )−1 = Let K be a compact subset of G. Our goal is to show that γ(2) n is eventually not in K. Since sets of the form Z(α, β) for some α, β ∈ E∗ γ(2) form a basis for the topology on the path groupoid, for each γ ∈ K there exist α(γ), β(γ) ∈ E∗ with s(α(γ)) = s(β(γ)) so that Z(α(γ), β(γ)) is an open neighbourhood of γ in G. Thus ∪γ∈KZ(α(γ), β(γ)) is an open cover of the compact set K, and so admits a finite subcover ∪I We now claim that for any fixed n ∈ P, if there exists i with 1 (cid:54) i (cid:54) I such that γ(2) and suppose there exists i with 1 (cid:54) i (cid:54) I such that γ(2) n ∈ Z(α(i), β(i)), then (cid:12)(cid:12)[v1, f (2) Suppose the converse: that α(i) <(cid:12)(cid:12)[v1, f (2) 1 (cid:54) p (cid:54) α(i). By examining the graph we can see that s(cid:0)[v1, f (2) n ]∗(cid:12)(cid:12). Since we also know that α(i) <(cid:12)(cid:12)[v1, f (2) for all 1 (cid:54) p <(cid:12)(cid:12)[v1, f (2) n ]∗(cid:12)(cid:12) (cid:54) α(i). Temporarily fix n ∈ P n ]∗(cid:12)(cid:12). Since γ(2) (cid:1) = vp+1 n ]∗(cid:12)(cid:12), we n ∈ Z(α(i), β(i)). n ∈ Z(α(i), β(i)), it p for every n ]∞ ∈ Z(α(i)), and so α(i) follows that r(γ(2) n ) = [v1, f (2) p = [v1, f (2) n ]∞ n ]∞ p n (γ(1) n (γ(1) i=1Z(α(i), β(i)). can deduce that s(α(i)) = vj for some j. Furthermore since s(α(i)) = s(β(i)), s(β(i)) = vj. There is only one path with source vj and range v1, so α(i) = β(i). Note that when k = β(i) − α(i), the set Z(α(i), β(i)) is by definition equal to {(x, k, y) : x ∈ Z(α(i)), y ∈ Z(β(i)), xp = yp+k for p > α(i)}, so since γ(2) r(γ(2) n ∈ Z(α(i), β(i)) and α(i) = β(i), we can see that s(γ(2) n ]∞ and r(γ(2) n )p for all p > α(i). We know s(γ(2) n ) = [v1, f (1) n )p = n ) = STRENGTH OF CONVERGENCE IN THE ORBIT SPACE OF A GROUPOID 31 n ]∞ p for all p > α(i). In particular, since n ]∞ [v1, f (2) p = [v1, f (1) n ]∞, so [v1, f (2) we assumed that(cid:12)(cid:12)[v1, f (2) n ]∗(cid:12)(cid:12) > α(i), we have n ]∞(cid:12)(cid:12)[v1,f (2) n ]∗(cid:12)(cid:12) = [v1, f (1) contradiction, and we must have(cid:12)(cid:12)[v1, f (2) n ]∞(cid:12)(cid:12)[v1,f (2) n ]∗(cid:12)(cid:12), n ]∗(cid:12)(cid:12) (cid:54) α(i). n . But f (1) n so that f (2) n = f (1) and f (2) n [v1, f (2) are distinct, so we have found a Our next goal is to show that each Z(α(i), β(i)) contains at most one γ(2) n . m are in Z(α(i), β(i)) for some n ]∞ ∈ n ]∗x ∈ Z(α(i)) and, Fix n, m ∈ P and suppose that both γ(2) i. We will show that n = m. Since γ(2) Z(α(i)). Thus there exists x ∈ E∞ such that [v1, f (2) n ]∗(cid:12)(cid:12) (cid:54) α(i), we can crop x to form a finite  ∈ E∗ such that n and γ(2) n ∈ Z(α(i) β(i)), r(γ(2) since (cid:12)(cid:12)[v1, f (2) n ) = [v1, f (2) n ]∗ = α(i). Similarly there exists δ ∈ E∗ such that [v1, f (2) m ]∗δ = α(i). [v1, f (2) Then [v1, f (2) n ]∗ = α(i) = [v1, f (2) m ]∗δ, n and γ(2) which we can see by looking at the graph is only possible if n = m. We have thus shown that if γ(2) m are in Z(α(i), β(i)), then γ(2) n ∈ K}. Since K ⊂ ∪I Let S = {n ∈ P : γ(2) i=1Z(α(i), β(i)) and since m ∈ Z(α(i), β(i)) implies n = m, S can contain at most I elements. γ(2) n , γ(2) /∈ K provided n > n0. Thus Then S has a maximal element n0 and γ(2) n γn → ∞ as n → ∞, and we have shown that x(n) converges 2-times to z in (cid:3) G(0)/G. n = γ(2) m . Example 8.3 (k-times convergence in a path groupoid). For any fixed positive integer k, let E be the graph v1 v2 v3 v4 f(1) 1 ,...,f(k) 1 f(1) 2 ,...,f(k) 2 f(1) 3 ,...,f(k) 3 f(1) 4 ,...,f(k) 4 n ]∞ and and let G be the path groupoid. For each n (cid:62) 1 define x(n) := [v1, f (1) let z be the infinite path that passes through each vn. Then the sequence {x(n)} converges k-times in G(0)/G to z. Proof. After defining γ(i) n argument similar to that in Example 8.2 establishes the k-times convergence. (cid:3) n ]∞, 0, x(n)(cid:1) for each 1 (cid:54) i (cid:54) k, an := (cid:0)[v1, f (i) Example 8.4. [Lower multiplicity 2 and upper multiplicity 3] Consider the graph E described by f (1) 1 v1 w1 f (1) 2 v2 w2 f (1) 3 v3 w3 f (1) 4 v4 w4 f (1) 5 v5 w5 32 ROBERT HAZLEWOOD AND ASTRID AN HUEF where for each odd n (cid:62) 1 there are exactly two paths f (1) n with source wn and range vn, and for each even n (cid:62) 2 there are exactly three paths f (1) n , f (2) n with source wn and range vn. Let G be the path groupoid, n ]∞ for every n (cid:62) 1, and let z be the infinite path that define x(n) := [v1, f (1) meets every vertex vn (so z has range v1). Then n , f (2) n , f (3) ML(Lz,{Lx(n)}) = 2 and MU(Lz,{Lx(n)}) = 3. Proof. We know that {x(n)} converges 2-times to z in G(0)/G by the argu- ment in Example 8.2, so we can apply Theorem 6.3 to see that ML(Lz,{Lx(n)}) (cid:62) 2. We can see that the subsequence {x2n} of {x(n)} converges 3-times to z in G(0)/G by Example 8.3. Theorem 7.4 now tells us that MU(Lz,{Lx(n)}) (cid:62) 3. Now suppose ML(Lz,{Lx(n)}) (cid:62) 3. Then by Theorem 6.3, {x(n)} con- n },{γ(2) n }, verges 3-times to z in G(0)/G, so there must exist three sequences {γ(1) n } as in the definition of k-times convergence (Definition 3.1). For and {γ(3) each odd n, there are only two elements in G with source x(n), so there must n )−1 = exist 1 (cid:54) i < j (cid:54) 3 such that γ(i) n )−1} admits a convergent r(γ(i) n ) frequently and, since r(γ(i) n )−1 (cid:57) ∞, contradicting the definition of k-times subsequence. Thus γ(j) convergence. If MU(Lz,{Lx(n)}) (cid:62) 4, then by Theorem 7.4 there is a subsequence of {x(n)} that converges 4-times to z in G(0)/G. A similar argument to that in the preceding paragraph shows that this is not possible since there are at most 3 edges between any vn and vm. It follows that ML(Lz,{Lx(n)}) = 2 and MU(Lz,{Lx(n)}) = 3. (cid:3) frequently. Then γ(j) n (γ(i) n = γ(j) n n ) → z, {γ(j) n (γ(i) n (γ(i) Lemma 8.5. In Example 8.2, ML(Lz,{Lx(n)}) = MU(Lz,{Lx(n)}) = 2; and in Example 8.3, ML(Lz,{Lx(n)}) = MU(Lz,{Lx(n)}) = k. Proof. The same argument as that found in Example 8.4 can be used to demonstrate this lemma. The explicit proof was given for Example 8.4 since it covers the case where the upper and lower multiplicities are distinct. (cid:3) In the next example we will add some structure to the graph from Ex- ample 8.2 to create a path groupoid G with non-Hausdorff orbit space that continues to exhibit 2-times convergence. Example 8.6. Let E be the directed graph STRENGTH OF CONVERGENCE IN THE ORBIT SPACE OF A GROUPOID 33 v1 v2 v3 v4 w1 w2 w3 w4 f (1) 1 f (2) 1 f (1) 2 f (2) 2 f (1) 3 f (2) 3 f (1) 4 f (2) 4 and let G be the path groupoid. For every n (cid:62) 1 let x(n) be the infinite path n ]∞. Let x be the infinite path with range v1 that passes through each [v1, f (1) vn and let y be the infinite path with range w1 that passes through each wn. Then the orbit space G(0)/G is not Hausdorff and {x(n)} converges 2-times in G(0)/G to both x and y. Proof. To see that {x(n)} converges 2-times to x in G(0)/G, consider the se- quences {([v1, f (2) n ]∞, 0, x(n))} and {(x(n), 0, x(n))} and follow the argument as in Example 8.2. To see that {x(n)} converges 2-times to y in G(0)/G, con- sider the sequences {([w1, f (1) n ]∞, 0, x(n))}. While it is tempting to think that this example exhibits 4-times convergence (or even 3-times convergence), this is not the case (see Example 8.4 for an argu- ment demonstrating this). We know x(n) converges k-times to x in G(0)/G, so [x(n)] → [x] in G(0)/G, and similarly [x(n)] → [y] in G(0)/G. It follows that G(0)/G is not Hausdorff since [x] (cid:54)= [y]. (cid:3) In all of the examples above, the orbits in G(0) are closed and hence C∗(G) and G(0)/G are homeomorphic by [10, Proposition 5.1]. By combining the features of the graphs in Examples 8.4 and 8.6 we obtain a principal groupoid whose C∗-algebra has non-Hausdorff spectrum and distinct upper and lower multiplicities among its irreducible representations. n ]∞, 0, x(n))} and {([w1, f (2) References C∗-algebras. Proc. London Math. Soc., 69(1):121 -- 143, 1994. [1] Robert J. Archbold. Upper and lower multiplicity for irreducible representations of [2] Robert J. Archbold and Klaus Deicke. Bounded trace C∗-algebras and integrable actions. Math. Z., 250(2):393 -- 410, 2005. [3] Robert J. Archbold and Astrid an Huef. Strength of convergence in the orbit space of a transformation group. J. Funct. Anal., 235(1):90 -- 121, 2006. [4] Robert J. Archbold and Astrid an Huef. Strength of convergence and multiplicities in the spectrum of a C∗-dynamical system. Proc. London Math. Soc., 96(3):545 -- 581, 2008. [5] Robert J. Archbold and Eberhard Kaniuth. Upper and lower multiplicity for irre- ducible representations of SIN-groups. Illinois J. Math., 43(4):692 -- 706, 1999. [6] Robert J. Archbold, Eberhard Kaniuth, Jean Ludwig, Gunter Schlichting, and Dou- glas W. B. Somerset. Strength of convergence in duals of C∗-algebras and nilpotent Lie groups. Adv. Math., 158(1):26 -- 65, 2001. [7] Robert J. Archbold, Jean Ludwig, and Gunter Schlichting. Limit sets and strengths of convergence for sequences in the duals of thread-like Lie groups. Math. Z., 255(2):245 -- 282, 2007. 34 ROBERT HAZLEWOOD AND ASTRID AN HUEF [8] Robert J. Archbold, Douglas W. B. Somerset, and Jack S. Spielberg. Upper mul- tiplicity and bounded trace ideals in C∗-algebras. J. Funct. Anal., 146(2):430 -- 463, 1997. representations of C∗-algebras. II. J. Operator Theory, 36(2):201 -- 231, 1996. [9] Robert J. Archbold and Jack S. Spielberg. Upper and lower multiplicity for irreducible [10] Lisa O. Clark. Classifying the types of principal groupoid C∗-algebras. J. Operator [11] Lisa O. Clark and Astrid an Huef. Principal groupoid C∗-algebras with bounded [12] Lisa O. Clark and Astrid an Huef. The representation theory of C∗-algebras associated [13] Philip Green. C∗-algebras of transformation groups with smooth orbit space. Pacific trace. Proc. Amer. Math. Soc., 136(2):623 -- 634, 2008. Theory, 57(2):251 -- 266, 2007. to groupoids. Preprint, 2010. J. Math., 72(1):71 -- 97, 1977. [14] Robert Hazlewood. Continuous trace, Fell, bounded trace, liminal and postliminal graph algebras. In preparation. [15] Alex Kumjian, David Pask, Iain Raeburn, and Jean Renault. Graphs, groupoids, and Cuntz-Krieger algebras. J. Funct. Anal., 144(2):505 -- 541, 1997. [16] Jean Ludwig. On the behaviour of sequences in the dual of a nilpotent Lie group. Math. Ann., 287(2):239 -- 257, 1990. [17] Paul S. Muhly. Coordinates in operator algebra. (Book in preparation). [18] Paul S. Muhly, Jean N. Renault, and Dana P. Williams. Continuous-trace groupoid [19] Paul S. Muhly and Dana P. Williams. Continuous trace groupoid C∗-algebras. Math. C∗-algebras. III. Trans. Amer. Math. Soc., 348(9):3621 -- 3641, 1996. Scand., 66(2):231 -- 241, 1990. [20] Gert K. Pedersen. Analysis now, volume 118 of Graduate Texts in Mathematics. Springer-Verlag, New York, 1989. [21] Iain Raeburn. Graph algebras, volume 103 of CBMS Regional Conference Series in Mathematics. Published for the Conference Board of the Mathematical Sciences, Washington, DC, 2005. groupoids. J. Funct. Anal., 94(2):358 -- 374, 1990. [22] Arlan Ramsay. The Mackey-Glimm dichotomy for foliations and other Polish [23] Jean Renault. A groupoid approach to C∗-algebras, volume 793 of Lecture Notes in [24] Dana P. Williams. Crossed products of C∗-algebras, volume 134 of Mathematical Sur- Mathematics. Springer, Berlin, 1980. veys and Monographs. American Mathematical Society, Providence, RI, 2007. School of Mathematics and Statistics, The University of New South Wales, Sydney, NSW 2052, Australia E-mail address: [email protected] Department of Mathematics and Statistics, University of Otago, PO Box 56, Dunedin 9054, New Zealand E-mail address: [email protected]
1002.4104
1
1002
2010-02-22T12:23:28
Spatial discretization of restricted group algebras
[ "math.OA", "math.NA" ]
We consider spatial discretizations by the finite section method of the restricted group algebra of a finitely generated discrete group, which is represented as a concrete operator algebra via its left-regular representation. Special emphasis is paid to the quasicommutator ideal of the algebra generated by the finite sections sequences and to the stability of sequences in that algebra. For both problems, the sequence of the discrete boundaries plays an essential role. Finally, for commutative groups and for free non-commutative groups, the algebras of the finite sections sequences are shown to be fractal.
math.OA
math
Spatial discretization of restricted group algebras Steffen Roch Abstract We consider spatial discretizations by the finite section method of the restricted group algebra of a finitely generated discrete group, which is represented as a concrete operator algebra via its left-regular representa- tion. Special emphasis is paid to the quasicommutator ideal of the algebra generated by the finite sections sequences and to the stability of sequences in that algebra. For both problems, the sequence of the discrete bound- aries plays an essential role. Finally, for commutative groups and for free non-commutative groups, the algebras of the finite sections sequences are shown to be fractal. 1 Introduction Approximately finite algebras and quasi-diagonal algebras are examples of C ∗- algebras which are distinguished by intrinsic finiteness properties. These prop- erties can be used in principle to approximate the elements of the algebra by finite-dimensional (or discrete) objects and, thus, to discretize the algebra in a sense. In this paper we consider a completely different kind of discretization, called spatial discretization, the main idea of which is as follows: We represent a given C ∗-algebra A faithfully as an algebra A of linear bounded operators on a separable Hilbert space with basis {ei}i∈N. Then we let Pn stand for the or- thogonal projection from H onto the linear span of e1, . . . , en, associate with each operator A ∈ A the sequence (PnAPn) of its finite sections, and consider the C ∗-algebra S(A) which is generated by all sequences (PnAPn) with A ∈ A. There is a natural homomorphism from S(A) onto A which associates with each sequence in S(A) its strong limit. Thus, the algebra A appears as a quotient of S(A) by the ideal of all sequences tending strongly to zero. The idea of spatial discretization has its origins in numerical analysis, where the numerical solution of an operator equation Au = f is a basic problem. Nu- merical analysis provides a huge arsenal of methods to discretize this equation for several classes of operators. The perhaps simplest (from the conceptual point of view) and most universal (applicable to each operator) method is the finite sections method which replaces the equation Au = f by the sequence of the 1 finite-dimensional linear systems PnAPnun = Pnf , n = 1, 2, . . .. The basic ques- tion is if these systems are uniquely solvable for sufficiently large n and if their solutions un tend to a solution of Au = f . The central aspect of this question is if the operators (= n × n-matrices) PnAPn are invertible for sufficiently large n and if the norms of their inverses are uniformly bounded. In this case, the sequence (PnAPn) is called stable. A Neumann series argument shows that the sequence (PnAPn) with A ∈ A is stable if and only if its coset is invertible in the quotient of the algebra S(A) by the ideal of all sequences which tend to zero in the norm. This observation due to Kozak brings numerical analysis into the realm of C ∗-algebras (and con- versely). It was soon realized that, for instance, Gelfand theory and its several non-commutative generalizations provide effective tools to study stability prob- lems for the finite sections method for convolution type equations; see [9] for an overview. In the consequence, the algebras S(A) were examined for several classes of operator algebras A. The pioneering example was the Toeplitz alge- bra, T(C), generated all Toeplitz operators on l2(N) with continuous generating function. This algebra can be viewed as a faithful representation of the univer- sal C ∗-algebra generated by one isometry (Coburn's theorem, [6]). The algebra S(T(C)) of the finite sections method is very well understood; for several aspects of finite sections of Toeplitz operators as well as for the rich history of the field see [3, 4]. These results were later extended to algebras generated by Toeplitz operators with piecewise continuous (and even "more discontinuous") symbols and to algebras of singular integral operators, see [8]. The algebra S(BDO) of the finite sections of band-dominated operators was subject of [18, 15] (note that the algebra BDO of the band-dominated operators is a faithful representation of the reduced crossed product algebra l∞(Z) ×αr Z), and the algebra S(ON ) where ON is a concrete representation of the Cuntz algebra ON was considered in [19]. The present paper is devoted to the spatial discretization of restricted group algebras C ∗ r (Γ) where Γ is a finitely generated discrete and exact group. Basic properties of group algebras can be found, e.g., in [2, 5, 7]. Restricted group alge- bras come with a natural representation, the so-called left-regular representation, which makes C ∗ r (Γ) isomorphic to the algebra Sh(Γ) of shift operators on l2(Γ). It is this algebra to which spatial discretization is applied in what follows. The paper is organized as follows. In Section 2 we provide some preliminaries on spatial discretization of represented C ∗-algebras. Section 3 is devoted to the spatial discretization of Sh(Γ). For we choose a family Y = (Yn) of finite subsets of Γ and consider the sequence of the finite sections PYnAPYn of A ∈ Sh(Γ). We show that the algebra SY (Sh(Γ)) generated by these sequences splits into the direct sum of Sh(Γ) and of an ideal which can be characterized as the quasicommutator ideal of the algebra. A main result is that the sequence (P∂Yn) of the discrete boundaries always belongs to the algebra SY(Sh(Γ)), and that this sequence already generates the quasicommutator ideal. This surprising fact has been already observed in other settings, for example for the algebra S(T(C)) of the finite sections method 2 for the Toeplitz operators (a classical result, closely related to the present paper), but also for the algebra S(ON ) related with Cuntz algebra (see [19]). In Section 4 we derive a necessary and sufficient criterion for the stability of sequences in SY(Sh(Γ)). The criterion is formulated of terms of limit operators (see [14, 18]). It turns out that it is sufficient to consider limit operators with respect to sequences η such that each ηn belongs to the boundary of some set Ykn, which gives another hint to the exceptional role of the discrete boundaries. In two special settings (commutative groups and free non-commutative groups) we show moreover that one can restrict to the case when η is an (inverse) geodesic path, which implies the fractality of the algebra SY(Sh(Γ)) for these groups. We will not present the details, but it should be at least mentioned here that one consequence of fractality is the excellent convergence properties of certain spectral quantities. For example, if a sequence (An) belongs to a fractal algebra, then the sets of the singular values (the points in the ǫ-pseudospectrum, the points in the numerical range, respectively) of the An converge with respect to the Hausdorff metric. For these and other applications of fractality, see [9, 17, 18, 20]. 2 Spatial discretization 2.1 Hilbert spaces and projections For a non-empty finite or countable set X, let l2(X) stand for the Hilbert space of all functions f : X → C with kxk2 := X x∈X f (x)2 < ∞. For X = ∅, we define l2(X) as the space {0} consisting of the zero element only. For each subset Y of X, we consider l2(Y ) as a closed subspace of l2(X) in a natural way. The orthogonal projection from l2(X) to l2(Y ) will be denoted by PY . Thus, PX and P∅ are the identity and the zero operator, respectively. For x ∈ X, let δx be the function on X which is 1 at x and 0 at all other points. If X is non-empty, then the family (δx)x∈X forms an orthonormal basis of l2(X), to which we refer as the standard basis. For each sequence (Yn)n≥1 of subsets of X, define its upper and lower limit as lim sup Yn := ∩k≥1 ∪n≥k Yn and lim inf Yn := ∪k≥1 ∩n≥k Yn. Thus, lim sup Yn is the set of all x ∈ X with x ∈ Yn for infinitely many n, whereas lim inf Yn contains all x ∈ X such that x ∈ Yn for all but finitely many n. A set sequence (Yn) is said to converge if lim sup Yn = lim inf Yn. In this case we denote the upper and lower limit by lim Yn. The following assertions are easy to check. 3 Proposition 2.1 (a) The sequence (PYn) of projections converges strongly if and only if the set sequence (Yn) converges. In this case, s-lim PYn = Plim Yn. (b) The sequence (PYn) converges strongly to the identity operator if and only if lim inf Yn = X. Corollary 2.2 (a) If Yn ⊆ Yn+1 for all n, then the sequence (PYn) converges strongly to P∪n≥1Yn. (b) If Ym ∩ Yn = ∅ for all m 6= n, then the sequence (PYn) converges strongly to 0. 2.2 Algebras of matrix sequences Let X be as before. Given a sequence Y := (Yn) of subsets of X, let FY denote the set of all bounded sequences A = (An) of operators An : im PYn → im PYn. Equipped with the operations (An) + (Bn) := (An + Bn), (An)(Bn) := (AnBn), (An)∗ := (A∗ n) and the norm kAkFY := kAnk, the set FY becomes a C ∗-algebra with identity, and the set GY of all sequences (An) ∈ FY with lim kAnk = 0 forms a closed ideal of FY. The relevance of the algebra FY and its ideal GY in our context stems from the fact (following from a simple Neumann series argument) that a sequence A ∈ FY is stable if, and only if, the coset A + GY is invertible in the quotient algebra FY/GY. Thus, every stability problem is equivalent to an invertibility problem in a suitably chosen C ∗-algebra. Let further stand F C Y for the set of all sequences A = (An) of operators An : im PYn → im PYn with the property that the sequences (AnPYn) and (A∗ nPYn) con- verge strongly. By the uniform boundedness principle, the quantity sup kAnPYnk is finite for every sequence (An) in F C Y is a closed and symmetric subalgebra of FY which contains G. Note that the mapping Y . Thus, F C W : F C Y → L(l2(X)), A 7→ s-lim AnPYn (1) is a ∗-homomorphism. 2.3 Spatial discretization of represented algebras Let A be a C ∗-subalgebra of L(l2(X)) (i.e., a represented C ∗-algebra), and let Y := {Yn} be a sequence of subsets of X. Write D for the mapping of spatial (= finite sections) discretization, i.e., D : L(l2(X)) → FY, A 7→ (PYnAPYn), (2) and let SY (A) stand for the smallest closed C ∗-subalgebra of the algebra FY which contains all sequences D(A) with A ∈ A. Clearly, SY(A) is contained in F C Y , and 4 the mapping W in (1) induces a ∗-homomorphism from SY(A) onto A. On this level, one cannot say much about the algebra SY (A). The little one can say will follow from the following simple facts. A proof is in [19]. Proposition 2.3 Let A and B be C ∗-algebras, D : A → B a symmetric linear contraction, and W : B → A a ∗-homomorphism such that W (D(A)) = A for every A ∈ A. Then (a) D is an isometry, D(A) is a closed linear subspace of B, and alg D(A), the smallest closed subalgebra of B which contains D(A), splits into the direct sum alg D(A) = D(A) ⊕ (ker W ∩ alg D(A)). Moreover, for every A ∈ A, kD(A)k = min K∈ker W kD(A) + Kk. (3) (4) (b) If B = alg D(A), then ker W coincides with the quasicommutator ideal of B, i.e., with the smallest closed ideal of B which contains all quasicommutators D(A1)D(A2) − D(A1A2) with A1, A2 ∈ A. We shall apply this proposition in the following context: A is a C ∗-subalgebra of L(l2(X)), B is the algebra SY (A), D is the restriction of the discretization (2) to A, and W is the restriction of the homomorphism (1) to SY (A). Then Proposition 2.3 specializes to the following. Proposition 2.4 Let A be a C ∗-subalgebra of L(l2(X)). Then the finite sections discretization D : A → FY is an isometry, and D(A) is a closed subspace of the algebra SY (A). This algebra splits into the direct sum SY (A) = D(A) ⊕ (ker W ∩ SY (A)), and for every operator A ∈ A one has kD(A)k = min K∈ker W kD(A) + Kk. Finally, ker W ∩ SY (A) is equal to the quasicommutator ideal of SY (A), i.e., to the smallest closed ideal of SY (A) which contains all sequences (PYnA1PYnA2PYn − PYnA1A2PYn) with operators A1, A2 ∈ A. We denote the ideal ker W ∩ SY (A) by J (A). Since the first item in the decom- position D(A) ⊕ J (A) of SY (A) is isomorphic (as a linear space) to A, a main part of the description of the algebra SY (A) is to identify the ideal J (A). Here is a first result which describes J (A) in terms of generators of A. Abbreviate I − PA =: QA. 5 Proposition 2.5 Let A be a C ∗-subalgebra of L(l2(X)) and let E be a subset of A which generates A as a Banach algebra, i.e., the smallest closed subalgebra of A which contains E is A. Then, for each m ≥ 2 and each choice of operators Ai ∈ E, the sequence (PYnA1QYnA2QYn . . . QYnAmPYn)n≥1 (5) belongs to J (A), and J (A) is the smallest closed ideal of SY (A) which contains all sequences of the form (5). Proof. First we show per induction that all sequences of the form (5) belong to the quasicommutator ideal J (A). This is evident for m = 2: (PYnA1QYnA2PYn) = (PYnA1A2PYn) − (PYnA1PYnA2PYn). Suppose the assertion is proved for sequences (5) of length less than m. Then (PYnA1QYn . . . QYnAm−1QYnAmPYn) = (PYnA1QYn . . . QYnAm−1AmPYn) −(PYnA1QYn . . . QYnAm−1PYn) (PYnAmPYn). The second sequence on the right-hand side of this equality is in J (A) by as- sumption. Write the first sequence as (PYnA1QYn . . . QYnAm−2QYnAm−1AmPYn) = (PYnA1QYn . . . QYnAm−2Am−1AmPYn) −(PYnA1QYn . . . QYnAm−2PYn) (PYnAm−1AmPYn). Again, the second sequence on the right-hand side is in J (A). We continue in this way to arrive finally at (PYnA1QYnA2A3 . . . AmPYn) = (PYnA1A2 . . . AmPYn) −(PYnA1PYn) (PYnA2A3 . . . AmPYn) which is in J (A) by the definition of the quasicommutator ideal. Conversely, we are going to show that the sequences (5) generate J (A) as a closed ideal of SY (A). Let J refer to the smallest closed ideal of SY (A) which contains all sequences (5). From the first part of this proof we infer that J ⊆ J (A). For the reverse inclusion it is sufficient to show that (PYnAQYnBPYn) ∈ J for all A, B ∈ A. Since J is a closed linear space, it is sufficient to verify this claim in case A and B are finite products of operators in E. Thus, we have to prove that (PYnA1 . . . AmQYnB1 . . . BlPYn) ∈ J (6) 6 for arbitrary operators Ai, Bj ∈ E and integers l, m ≥ 1. Again we use induction. The assertion is evident in case m = l = 1. For the general step we write (PYnA1 . . . AmQYnB1 . . . BlPYn) = (PYnA1PYn) (PYnA2 . . . AmQYnB1 . . . BlPYn) + (PYnA1QYnA2 . . . AmQYnB1 . . . BlPYn). The first summand on the right-hand side is a product of a sequence in SY (A) and a sequence of the form (6), but with less factors. By assumption, this summand is in J . The second summand can be again written as a sum by inserting I = PYn + QYn after A2. We continue in this way and arrive finally at the sequence (PYnA1QYnA2QYn . . . QYnBl−1QYnBlPYn) which is in J by definition. 3 Spatial discretization of restricted group C ∗- algebras 3.1 Left regular representations Let Γ be a (not necessarily commutative) discrete group. We write the group operation as multiplication and let e stand for the identity element. With Γ we associate the Hilbert space l2(Γ) with its canonical basis (δs)s∈Γ. The left regular representation L : Γ → L(l2(Γ)) associates with every group element r a unitary operator Lr such that Lrδs = δrs for s ∈ Γ. Since δrs(t) = δs(r−1t), one has (Lru)(t) = u(r−1t) for every u ∈ l2(Γ). Hence, r 7→ Lr is a group isomorphism. We define Sh(Γ) as the smallest closed subalgebra of L(l2(Γ)) which contains all operators Lt with t ∈ Γ. The algebra Sh(Γ) is ∗- isomorphic to the restricted group C ∗-algebra C ∗ r (Γ) in a natural way (see Section 2.5 in [5]). It can thus be considered as a concrete representation of C ∗ r (Γ). Note also that the restricted group C ∗-algebra coincides with the universal group C ∗- algebra C ∗(Γ) if the group Γ is amenable. For this and further characterizations of amenable groups, see Theorem 2.6.8 in [5]. We have seen above that every restricted group C ∗-algebra C ∗ r (Γ) comes with a canonical faithful representation as the concrete operator algebra Sh(Γ) on l2(Γ). We will take this representation as the basis for the spatial discretization of C ∗ r (Γ) by a finite sections method in the following sections. The existence of a canonical representation is only one reason why we con- sider spatial discretizations only for restricted group C ∗-algebras in what fol- lows. Another reason is that universal group C ∗-algebras sometimes own intrinsic finiteness properties which can be used to approximate their elements by finite dimensional objects, but which are not shared by the associated restricted group C ∗-algebras. For example, if Γ is the free non-commutative group F2 of two gen- erators, then the universal group C ∗-algebra C ∗(F2) is known to be quasidiagonal, whereas C ∗ r (F2) fails to have this property (see Sections VII.6 and VII.7 in [7]). 7 3.2 Discretization of Sh(Γ) To discretize the algebra Sh(Γ) by the finite sections method we choose a sequence Y = (Yn) of finite subsets of Γ and consider the sequences (PYnAPYn) of the finite sections of A ∈ Sh(Γ). Usually we will assume that the set limit lim Yn exists and is equal to Γ, in which case the PYn converge strongly to the identity operator, but some of the following results will hold without this assumption. In accordance with earlier notation, let SY (Sh(Γ)) stand for the smallest closed C ∗-subalgebra of the algebra FY which contains all sequences (PYnAPYn) with A ∈ Sh(Γ). The associated quasicommutator ideal is denoted by J (Sh(Γ)). In the next section, we shall present some characterizations of J (Sh(Γ)). For we have to introduce some notions of topological type. Note that the standard topology on Γ is the discrete one; so every subset of Γ is open with respect to this topology. Let Ω be a finite subset of Γ which contains the identity element e and which generates Γ as a semi-group, i.e., if Ωn denotes the set of all words of length at most n with letters in Ω, then ∪n≥0Ωn = Γ. By convention, Ω0 := {e}. Note also that the sequence (Ωn) is increasing; so the operators PΩn can play the role of the finite sections projections PYn, and in fact we will obtain some of the subsequent results exactly for this sequence. With respect to Ω, we define the following "algebro-topological" notions. Let A ⊆ Γ. A point a ∈ A is called an Ω-inner point of A if Ωa := {ωa : ω ∈ Ω} ⊆ A. The set intΩA of all Ω-inner points of A is called the Ω-interior of A, and the set ∂ΩA := A \ intΩA is the Ω-boundary of A. Note that by this definition, the Ω-boundary of a set is always a part of that set. (In this point, the present definition of a boundary differs from other definitions used in the literature, see, e.g., [1].) One easily checks that the Ω-interior and the Ω-boundary of a set are invariant with respect to multiplication from the right-hand side: (intΩA)s = intΩ(As) and (∂ΩA)s = ∂Ω(As) for s ∈ Γ. One also has Ωn−1 ⊆ intΩΩn ⊆ Ωn for each n ≥ 1, whence ∂ΩΩn ⊆ Ωn \ Ωn−1 for each n ≥ 1. In many concrete settings, one has equality in (8). (7) (8) 3.3 The structure of the quasicommutator ideal Let Ω and Y := (Yn) be as in the previous section. We will derive two results on the structure of the quasicommutator ideal J (Sh(Γ)). 8 Theorem 3.1 J (Sh(Γ)) is the smallest closed ideal of SY (Sh(Γ)) which contains all sequences (PYnLω−1QYnLωPYn)n≥1 with ω ∈ Ω. Proof. First note that, for arbitrary ω ∈ Ω and A ⊆ Γ, QALωPA = QALωPALω−1QALωPA. (9) (10) Indeed, QALωPA = QALωPALω−1LωPA = QALωPALω−1QALωPA + QALωPALω−1PALωPA. The second summand on the right-hand side vanishes since LωPALω−1 = PωA commutes with PA. Let now, for a moment, J denote the smallest closed ideal of SY (Sh(Γ)) which contains all sequences (9). Clearly, J ⊆ J (Sh(Γ)). The reverse implication will follow via Proposition 2.5 once we have shown that each sequence (PYnLω1QYnLω2QYn . . . QYnLωmPYn)n≥1 (11) with m ≥ 2 and ωi ∈ Ω belongs to J . Write the sequence (11) as (AnQYnLωmPYn). By (10), (AnQYnLωmPYn) = (AnQYnLωmPYn) (PYnLω−1QYnLωmPYn). Since the sequence (11) belongs to SY (Sh(Γ)) and J is an ideal of that algebra, the sequence (11) is in J . Lemma 3.2 Let A ⊆ Γ. Then ∩ω∈Ω(A ∩ ω−1A) = intΩA. Proof. Let a ∈ intΩA. Then, for each ω ∈ Ω, a = ω−1ωa ∈ ω−1Ωa ⊆ ω−1A ⊆ A ∩ ω−1A, whence the inclusion ⊇. For the reverse inclusion, let a ∈ A \ intΩA = ∂ΩA. By definition of the Ω-boundary, there is an ω0 ∈ Ω such that ω0a 6∈ A. Hence, a 6∈ A ∩ ω−1 0 A, which implies a 6∈ ∩ω∈Ω(A ∩ ω−1A). Lemma 3.3 Let A be a subalgebra of L(l2(Γ)) and A ⊆ Γ. If the operators PALω−1PALωPA belong to A for each ω ∈ Ω, then the operators PA, PintΩA and P∂ΩA belong to A, too. Proof. Since e ∈ Ω, the assertion is evident for PA. Further we have PALω−1PALωPA = PAPω−1A = PA∩ω−1A ∈ A 9 for each ω ∈ Ω. Since A is an algebra, this implies Y ω∈Ω PA∩ω−1A = P∩ω∈Ω(A∩ω−1A) ∈ A. By Lemma 3.2, this is the assertion for PintΩA. The assertion for P∂ΩA follows since PA = PintΩA + P∂ΩA. We call (P∂ΩYn)n≥1 the sequence of the discrete boundaries of the finite section method with respect to (Yn). Note that the assumptions in the following theorem are satisfied if Yn = Ωn due to (7). Theorem 3.4 Assume that Yn−1 ⊆ intΩYn ⊆ Yn for all n ≥ 2 and that lim Yn = Γ. Then the sequence (P∂ΩYn)n≥1 of the discrete boundaries belongs to the algebra SY (Sh(Γ)), and the quasicommutator ideal is generated by this sequence, i.e., J (Sh(Γ)) is the smallest closed ideal of SY (Sh(Γ)) which contains (P∂ΩYn)n≥1. Proof. By definition, the sequence (PYnLω−1PYnLωPYn)n≥1 is in SY (Sh(Γ)) for each ω ∈ Ω. From Lemma 3.3 we then conclude that the sequence (P∂ΩYn) is in SY (Sh(Γ)), too. That this sequence is even in the quasicommutator ideal, is a consequence of the assumptions. Indeed, from Yn−1 ⊆ intΩYn ⊆ Yn we conclude that lim n→∞ intΩYn = lim n→∞ Yn = Γ whence s-limP∂ΩYn = 0. By Proposition 2.4, this implies (P∂ΩYn) ∈ J (Sh(Γ)). It remains to show that the sequence (P∂ΩYn) generates J (Sh(Γ)). Let J denote the smallest closed ideal of SY(Sh(Γ)) which contains the sequence (P∂ΩYn). By what we have just seen, J ⊆ J (Sh(Γ)). The reverse inclusion will follow from Theorem 3.1 once we have shown that (PYnLω−1QYnLωPYn)n≥1 ∈ J for each ω ∈ Ω. (12) Note that PYnLω−1QYnLωPYn = PYn − PYnLω−1PYnLωPYn = PYn\(Yn∩ω−1Yn). From Lemma 3.2 we know that intΩYn ⊆ Yn ∩ ω−1Yn. Hence, Yn \ (Yn ∩ ω−1Yn) ⊆ Yn \ intΩYn = ∂ΩYn which implies that PYnLω−1QYnLωPYn = PYn\(Yn∩ω−1Yn) = PYn\(Yn∩ω−1Yn) P∂ΩYn. This verifies (12) and finishes the proof of the theorem. 10 4 Stability In this section, we are going to study the stability of sequences in SY (Sh(Γ)) via the limit operators method. The key ingredients are the facts that the stability of a sequence A in that algebra is equivalent to the Fredholmness of a certain associated operator and that the Fredholmness of that operator can be studied by means of its limit operators due to a result of Roe. 4.1 Fredholmness vs. stability Let Y := (Yn) be a sequence of finite subsets of Γ. A sequence (vn) ⊆ Γ is called an inflating sequence for Y if Ymv−1 n = ∅ for m 6= n. The existence of inflating sequences is a consequence of the following lemma. m ∩ Ynv−1 Lemma 4.1 Let A, B ⊂ Γ be finite and V ⊂ Γ be infinite. Then there is a v ∈ V such that A ∩ Bv−1 = ∅. Indeed, let A ∩ Bv−1 6= ∅ for every v ∈ V . Then, for each v ∈ V , there is a bv ∈ B such that bvv−1 =: av ∈ A. Thus, v = bva−1 v . But since A and B are finite, there are only finitely many products bva−1 v . Hence V is finite, a contradiction. Corollary 4.2 Let Y = (Yn) be a sequence of finite subsets of Γ and V an infinite subset of Γ. Then there is an inflating sequence for Y in V . Proof. Let v1 ∈ V . Then Y1v−1 1 that Y1v−1 such that 1 ∩ Y2v−1 2 = ∅. Further, since Y1v−1 is finite. By the lemma, there is a v2 ∈ V such is finite, there is a v3 ∈ V 1 ∪ Y2v−1 2 (cid:0)Y1v−1 1 ∪ Y2v−1 2 (cid:1) ∩ Y3v−1 3 = ∅. We proceed in this way to find the desired inflating sequence. In what follows let Y as above and choose and fix an inflating sequence (vn) for Y. Further set Γ′ := Γ \ ∪∞ n=1Ynv−1 n . (13) For s ∈ Γ, let Rs : l2(Γ) → l2(Γ) refer to the operator (Rsf )(t) := f (ts). Evi- dently, the mapping R : s 7→ Rs is a group isomorphism from Γ into the group of the unitary operators on l2(Γ). Moreover, RsLt = LtRs for s, t ∈ Γ. The proof of the following theorem is adapted from [18]. Theorem 4.3 Let A = (An) ∈ FY . Then (a) the series ∞ X RvnAnR−1 vn (14) converges strongly on l2(Γ). The sum of this series is denoted by Op (A). n=1 11 (b) the sequence (An) is stable if and only if the operator Op (A)+PΓ′ is Fredholm on l2(Γ). (c) The mapping Op is a continuous homomorphism from FY to L(l2(Γ)). Proof. (a) It is convenient to identify the operator An acting on im PYn with the operator PYnAnPYn acting on all of l2(Γ). Since RvnPYnR−1 , one can then identify the operator vn = PYnv−1 n RvnAnR−1 vn : im PYnv−1 n → im PYnv−1 n RvnAnR−1 with the operator PYnv−1 inflating property ensures that the vectors RvnAnR−1 in l2(Γ). Consequently, the series P∞ n=1 RvnAnR−1 series vn PYnv−1 on l2(Γ). Thus, for x ∈ l2(Γ), the vn x form an orthogonal system vn x converges if and only if the n n ∞ X kRvnAnR−1 vn xk2 (15) n=1 converges. Set M := sup kAnk. Employing the orthogonality of the vectors PYnv−1 x, we get n ∞ X n=1 kRvnAnR−1 vn xk2 ≤ M 2 ∞ X n=1 kPYnv−1 n xk2 ≤ M 2kxk2. Thus, the series (15) converges for every x, whence assertion (a). (b) Let A = (An) be a stable sequence, i.e., there is an n0 ∈ N such that the operators An : im PYn → im PYn are invertible for n ≥ n0 and that the norms of their inverses are uniformly bounded. Then the operator B := is invertible with inverse B−1 = n0−1 X n=1 n0−1 X n=1 PYnv−1 n + PYnv−1 n + ∞ X n=n0 ∞ X n=n0 RvnAnR−1 vn + PΓ′ RvnA−1 n R−1 vn + PΓ′. Since Op (A) + PΓ′ is a compact perturbation of B, Op (A) + PΓ′ is a Fredholm operator (with Fredholm index 0). Let, conversely, Op (A) + PΓ′ be a Fredholm operator. Then there are an operator B ∈ L(l2(Γ)) and a compact operator K on l2(Γ) such that B · (Op (A) + PΓ′) = I + K. 12 Since the projections PYnv−1 find n commute with Op (A), and since Ynv−1 n ∩ Γ′ = ∅, we PYnv−1 n BPYnv−1 n whence · RvnAnR−1 n vn = PYnv−1 = PYnv−1 = PYnv−1 = PYnv−1 n n n n Op (A)PYnv−1 n n n · PYnv−1 BPYnv−1 BOp (A)PYnv−1 B(Op (A) + PΓ′)PYnv−1 + PYnv−1 KPYnv−1 , n n n R−1 vn PYnv−1 n BPYnv−1 n Rvn · An = PYn + R−1 vn PYnv−1 n KPYnv−1 n Rvn. (16) Since PYnv−1 since K is compact and kRvnk = 1, we further conclude → 0 strongly by the inflating property and by Corollary 2.2 (b) and n kR−1 vn PYnv−1 n KPYnv−1 n Rvnk → 0 as n → ∞. Hence, the operators on the right-hand side of (16) (considered as acting on im PYn) are invertible for n large enough, and the norms of their inverses are uniformly bounded with respect to n. This implies the uniform boundedness of the operators Bn := (cid:0)PYn + R−1 vn PYnv−1 n KPYnv−1 n Rvn(cid:1)−1 R−1 vn PYnv−1 n BPYnv−1 n Rvn, also considered as acting on im PYn. Since BnAn = PYn for all sufficiently large n and the An act on a finite-dimensional space, the stability of the sequence (An) follows. Assertion (c) is an immediate consequence of the inflating property. 4.2 Band-dominated operators Theorem 4.3 translates the stability problem for a bounded sequence of finite- rank operators into a Fredholm problem for an associated operator. In case of the finite sections sequence of an operator in Sh(Γ), the associated operator is a band-dominated operator in the sense defined below. Since there is an effective criterion to verify the Fredholm property (which we will recall in the subsequent section) of band-dominated operators, this observation offers a way to study the stability of the finite sections method for operators in Sh(Γ). Consider functions k ∈ l∞(Γ × Γ) with the property that there is a finite subset Γ0 of Γ such that k(t, s) = 0 whenever ts−1 6∈ Γ0. Then (Au)(t) := X s∈Γ k(t, s) u(s), t ∈ Γ, (17) defines a linear operator A on the linear space of all functions u : Γ → C, since the occurring series is finite for every t ∈ G. We call operators of this form band 13 operators and the set Γ0 a band-width of A. It is not hard to see that the band operators form a symmetric algebra of bounded operators on l2(Γ). Operators in the norm closure of that algebra are called band-dominated operators. Thus, the band-dominated operators form a C ∗-subalgebra BDO(Γ) of L(l2(Γ)). It turns out that band operators on Γ are constituted by two kinds of "ele- mentary" band operators: the unitary operators Lt of left shift by t ∈ Γ, and the operators bI of multiplication by a function b ∈ l∞(G), bI : l2(Γ) → l2(Γ), (bu)(s) = b(s) u(s). Proposition 4.4 A operator in L(l2(Γ)) is a band operator if and only if it can be written as a finite sum P biLti where bi ∈ l∞(Γ) and ti ∈ Γ. Proof. Let A be an operator of the form (17) and let Γ0 := {t1, t2, . . . , tr} be a finite subset of Γ such that k(t, s) = 0 if ts−1 6∈ Γ0 or, equivalently, if s is not of the form t−1 t for some i. Thus, i (Au)(t) = r X i=1 k(t, t−1 i t) u(t−1 i t) for all t ∈ Γ. Set bi(t) := k(t, t−1 i t). The functions bi are in l∞(Γ), and one has A = r X i=1 biLti. (18) Conversely, one easily checks that each operator Lt with t ∈ Γ is a band operator with band width {t} and that each operator bI with b ∈ l∞(Γ) is a band operator with band width {e}. Since the band operators form an algebra, each finite sum P biLti is a band operator. It is easy to see that the representation of a band operator on Γ in the form (18) is unique. The functions bi are called the diagonals of the operator A. In particular, operators in Sh(Γ) can be considered as band-dominated operators with constant coefficients. As before, let Y := (Yn) be a sequence of finite subsets of Γ and (vn) an associated inflating sequence. Note that the following proposition remains valid if the algebra Sh(Γ) is replaced by the C ∗-algebra BDO(Γ) of all band-dominated operators. Proposition 4.5 Let A = (An) be a sequence in the finite sections algebra SY (Sh(Γ)). Then Op (A) is a band-dominated operator. 14 Proof. First let A ∈ Sh(Γ) be a band operator (i.e., A is a linear combination of a finite number of the Lt) and let Γ0 be a band width of A. It is easy to check that then RvnPYnAPYnR−1 vn is a band operator with the same band width for every n. The inflating property ensures that Op ((PYnAPYn)) is a band operator with band width Γ0, too. Now Theorem 4.3 (c) yields the assertion. To define limit operators, let h : N → Γ be a sequence tending to infinity in the sense that for each finite subset Γ0 of Γ, there is an n0 ∈ N such that h(n) 6∈ Γ0 if n ≥ n0. Clearly, if h tends to infinity, then the inverse sequence h−1 tends to infinity, too. We say that an operator Ah ∈ L(l2(Γ)) is a limit operator of A ∈ L(l2(Γ)) defined by the sequence h if R−1 h(m)ARh(m) → Ah and R−1 h(m)A∗Rh(m) → A∗ h strongly as m → ∞. Clearly, every operator has at most one limit operator with respect to a given sequence h. Note that the generating function of the shifted operator R−1 r ARr is related with the generating function of A by kR−1 r ARr (t, s) = kA(tr−1, sr−1) (19) and that the generating functions of R−1 to the generating function of the limit operator Ah (if the latter exists). h(m)ARh(m) converge pointwise on Γ × Γ It is an important property of band-dominated operators that they always possess limit operators. More general, the following result can be proved by a standard Cantor diagonal argument (see [12, 13, 14]). Proposition 4.6 Let A be a band-dominated operator on l2(Γ). Then every sequence h : N → Γ which tends to infinity possesses a subsequence g such that the limit operator Ag of A with respect to g exists. Let A be a band-dominated operator and h : N → Γ a sequence tending to infinity for which the limit operator Ah of A exists. Let B be another band-dominated operator. By Proposition 4.6 we can choose a subsequence g of h such that the limit operator Bg exists. Then the limit operators of A, A + B and AB with respect to g exist, and Ag = Ah, (A + B)g = Ag + Bg, (AB)g = AgBg. Thus, the mapping A 7→ Ah acts, at least partially, as an algebra homomorphism. The following theorem is due to Roe [22], see also [11]. Recall in this connec- tion that a group Γ is said to be exact, if its reduced translation algebra is an exact C ∗-algebra. The latter is defined as the reduced crossed product of l∞(Γ) by Γ and coincides in our setting with the C ∗-algebra of all band-dominated operators on l2(Γ). The class of exact groups is extremely rich. It includes all amenable groups (hence, all solvable groups such as the discrete Heisenberg group and the commutative groups) and all hyperbolic groups (in particular, all free groups with finitely many generators) (see [21], Chapter 3). 15 Theorem 4.7 (Roe) Let Γ be a finitely generated discrete and exact group, and let A be a band-dominated operator on l2(Γ). Then the operator A is Fredholm on l2(Γ) if and only if all limit operators of A are invertible and if the norms of their inverses are uniformly bounded. Note that this result holds as well if the left regular representation is replaced by the right regular one and if, thus, the operators Ls and Rt change their roles. In fact, in [11, 22] the results are presented in this symmetric setting. In [11] we showed moreover that the uniform boundedness condition in Theorem 4.7 is redundant for band operators if the group Γ has sub-exponential growth and if not every element of Γ is cyclic in the sense that wn = e for some positive integer n. For details see [11]. Note that the condition of sub-exponential growth is satisfied by the abelian groups ZN , the discrete Heisenberg group and, more general, by nilpotent groups (in fact, these groups have polynomial growth), whereas the growth of the free groups FN is exponential. Theorem 4.8 Let Γ be a finitely generated discrete and exact group with sub- exponential growth which possesses at least one non-cyclic element, and let A be a band operator on l2(Γ). Then the operator A is Fredholm on l2(Γ) if and only if all limit operators of A are invertible. 4.3 Limit operators and stability Let Y = (Yn) be a sequence of finite subsets of Γ. To verify the stability of a sequence A = (An) in SY (Sh(Γ)) via the results of the previous section, we have to choose an inflating sequence for Y and to compute the limit operators of Op (A) + PΓ′. Note that the exactness of Γ is not relevant in this computation. Note also that large parts of this computation hold for sequences in SY (BDO(Γ)), too. We will consider the finite sections method for operators in BDO(Γ) in detail in a forthcoming paper. Let Ω be a finite subset of Γ with e ∈ Ω which generates Γ as a semi-group. Let Ωn denote the set of all words with letters in Ω of length at most n. Thus Γ = ∪n≥1Ωn = limn→1 Ωn. By Theorem 4.3, the Fredholmness of the operator Op (A) is independent of the concrete choice of the inflating sequence. For technical reasons, we choose an inflating sequence (vn) for the sequence (cid:0)(Yn ∪ Ωn)(Yn ∪ Ωn)−1(Yn ∪ Ωn)(cid:1)n≥1 instead of (Yn). Since Yn ∪ Ωn ⊂ (Yn ∪ Ωn)(Yn ∪ Ωn)−1 ⊂ (Yn ∪ Ωn)(Yn ∪ Ωn)−1(Yn ∪ Ωn), (vn) is also an inflating sequence for (Yn). Moreover, since lim Ωn = Γ, one also has lim (Yn ∪ Ωn)(Yn ∪ Ωn)−1 = Γ. (20) 16 Let now A = (An) ∈ SY(Sh(Γ)), set as before Op (A) = ∞ X n=1 RvnAnR−1 vn and Γ′ = Γ \ ∪∞ n=1Ynv−1 n , and let h : N → Γ be a sequence tending infinity for which the limit operator (Op (A) + PΓ′)h := s-limn→∞R−1 h(n)(Op (A) + PΓ′)Rh(n) exists. Then the limit operator (Op (A) + PΓ′)g exists for every subsequence g of h, and it coincides with (Op (A) + PΓ′)h. So we can pass freely to subsequences of h. By passing to subsequences, we can restrict the computation of the limit operator to the following cases: Case 1: All elements h(n) belong to ∪k≥1 vkY −1 Case 2: No element h(n) belong to ∪k≥1 vkY −1 . k k . Consider Case 1. Passing again to a subsequence of h we can further suppose kn , and that vknY −1 that each h(n) belongs to one of the sets vkY −1 kn contains no other element of the sequence h besides h(n). For each n, let rn denote the smallest non-negative integer such that h(n) ∈ vkn(∂ΩYkn)−1Ωrn. Thus, rn measures the distance of h(n) to the Ω-boundary of vknY −1 kn . Finally, let r∗ := lim inf n→∞ rn. Again we distinguish two cases. Case 1.1: r∗ is finite. Then there are infinitely many n ∈ N such that rn = r∗. Thus, there is a subsequence of h (denoted by h again) such that , say to vknY −1 k h(n) ∈ vknY −1 kn ∩ vkn(∂ΩYkn)−1Ωr∗ for all n. n ∈ Ωr∗ such that h(n) ∈ vkn(∂ΩYkn)−1w∗ Further, for each n there is an w∗ n. Since Ωr∗ is a finite set, one of its elements w∗ n occurs for infinitely many n. Let w∗ be an element of Ωr∗ with this property. Consider the subsequence of h which contains all elements h(n) with w∗ n = w∗. Denoting this subsequence by h again, we can hence assume that h(n) ∈ vknY −1 kn ∩ vkn(∂ΩYkn)−1w∗ (21) for all n. With respect to this sequence h we obtain R−1 h(n)(Op (A) + PΓ′)Rh(n) = ∞ X k=1 R−1 h(n)Rvk AkR−1 vk Rh(n) + R−1 h(n)PΓ′Rh(n) = X k6=kn R−1 h(n)Rvk AkR−1 vk Rh(n) + R−1 h(n)PΓ′Rh(n) + R−1 h(n)Rvkn AknR−1 vkn Rh(n) (22) 17 with Γ′ as in (13). By (21), h(n) = vknηknw∗ with ηkn ∈ (∂ΩYkn)−1. Thus, the last item in (22) becomes Rw−1 ∗ Rη−1 kn AknRηkn Rw∗. (23) Set Πn := P(Ykn ∪Ωkn )(Ykn ∪Ωkn )−1w∗. By (20), Πn → I strongly. Since Akn acts on im PYkn , the operator (23) acts on im PYkn ηkn w∗. The evident inclusion Yknηknw∗ ⊆ (Ykn ∪ Ωkn)(Ykn ∪ Ωkn)−1w∗ implies that ΠnR−1 h(n)Rvkn AkR−1 vkn Rh(n) = R−1 h(n)Rvkn AkR−1 vkn Rh(n)Πn = R−1 h(n)Rvkn AkR−1 vkn Rh(n). Let now k 6= kn. Then, by the inflating property, (Yk ∪ Ωk)(Yk ∪ Ωk)−1(Yk ∪ Ωk)v−1 k ∩(Ykn ∪ Ωkn)(Ykn ∪ Ωkn)−1(Ykn ∪ Ωkn)v−1 kn = ∅. (24) Since Ykv−1 k ⊆ (Yk ∪ Ωk)(Yk ∪ Ωk)−1(Yk ∪ Ωk)v−1 k and (Ykn ∪ Ωkn)(Ykn ∪ Ωkn)−1η−1 kn v−1 kn ⊆ (Ykn ∪ Ωkn)(Ykn ∪ Ωkn)−1(Ykn ∪ Ωkn)v−1 kn we conclude from (24) that Ykv−1 k ∩ (Ykn ∪ Ωkn)(Ykn ∪ Ωkn)−1η−1 kn v−1 kn = ∅ whence Ykv−1 h(n)Rvk AkR−1 k vknηknw∗ ∩ (Ykn ∪ Ωkn)(Ykn ∪ Ωkn)−1w∗ = ∅. vk Rh(n) is an operator living on im PYkv−1 k vkn ηkn w∗, we conclude Since R−1 that R−1 for k 6= kn. Hence, h(n)Rvk AkR−1 vk Rh(n)Πn = ΠnR−1 h(n)Rvk AkR−1 vk Rh(n) = 0 R−1 h(n)(Op (A) + PΓ′)Rh(n) h(n)RvkAkR−1 = X R−1 vk Rh(n)(I − Πn) + R−1 h(n)PΓ′Rh(n) k6=kn + R−1 w∗ R−1 ηkn AknRηkn Rw∗Πn. (25) Since Πn → I strongly, the first summand on the right-hand side of (25) converges strongly (and even ∗-strongly since Πn commutes with that sum) to zero. Thus, s-lim R−1 h(n)(Op (A) + PΓ′)Rh(n) = s-lim R−1 w∗ R−1 ηkn AknRηkn Rw∗Πn + s-lim R−1 h(n)PΓ′Rh(n), provided that the strong limits on the right-hand side exist. The existence of the second strong limit can always be forced by passing to a suitable subsequence of h. Collecting these facts, we arrive at the following. 18 Theorem 4.9 Let h be a sequence such that the limit operator Op (A) + PΓ′ exists. In Case 1.1, there is a subsequence g of h such that the limit operator (PΓ′)g exists, and there are a monotonically increasing sequence (kn) in N, for each n a vector ηkn ∈ (∂ΩYkn)−1, and a w∗ ∈ Γ such that (Op (A) + PΓ′)h = s-lim R−1 w∗ R−1 ηkn AknRηkn Rw∗ + (PΓ′)g. Thus, the operator Akn living on im PYkn is shifted by a vector ηkn ∈ (∂ΩYkn)−1 and by another vector w∗ independent of n. It is only a matter of taste to consider Akn as shifted by the vector η−1 kn belonging to the Ω-boundary of Ykn. In particular, every limit operator of Op (A) is a shift by some vector w∗ of a strong limit of operators Akn, shifted by vectors in the boundary of Ykn. This is well known for the group Z and intervals Yk = [−k, k] ∩ Z, and it was observed by Lindner [10] in case Γ = ZN and Yk = Ωk is a polygon with integer vertices. Before turning to the other cases, let us specify Theorem 4.9 to pure finite sections sequences for operators in Sh(Γ). The existence of the limit operator (PΓ′)h is guaranteed if the strong limit s-lim R−1 w∗ R−1 ηkn PYkn Rηkn Rw∗ = s-lim PYkn ηkn w∗ exists, i.e., if the set limit lim Yknηknw∗ =: Y (h) (26) exists. In this case, (PΓ′)g = I − PY (h). Corollary 4.10 Let A ∈ Sh(Γ), and let h be a sequence such that the limit operator Op (A)h for the sequence (PYnAPYn) exists. In Case 1.1, there are kn, ηkn and w∗ as in Theorem 4.9 such that the set limit (26) exists. Then (Op (A) + PΓ′)h = PY (h)APY (h) + (I − PY (h)). (27) Conversely, if the limit (26) exists for a certain choice of kn, ηkn and w∗ as in Theorem 4.9, then the limit operator Op (A)h exists for the sequence h(n) := vknηknw∗, and (27) holds. The proof of the first assertion follows immediately from Theorem 4.9 and from the shift invariance of the operator A: Rw−1 ∗ η−1 kn PYkn APYkn Rηkn w∗ = Rw−1 ∗ η−1 kn PYkn Rηkn w∗ · A · Rw−1 ∗ η−1 kn PYkn Rηkn w∗. The second assertion is evident. Case 1.2: r∗ is infinite. Recall that h(n) ∈ vknY −1 kn and h(n) 6∈ vkn(∂ΩYkn)−1Ωrn−1 (28) 19 for all n ∈ N. The second assertion in (28) implies that h(n)Ω−1 rn−1 ∩ vkn(∂ΩYkn)−1 = ∅. Hence, we can rewrite (28) as e ∈ Yknv−1 kn h(n) and Ωrn−1 ∩ (∂ΩYkn)v−1 kn h(n) = ∅. We claim that this implies that Ωrn−1 ⊆ Yknv−1 kn h(n). (29) (30) Suppose (30) is wrong. Then Ωrn−1 has at least one point outside Yknv−1 kn h(n), say a, but it also has points inside this set, for example the point e due to the first assumption of (29). Write a as a product a = wrn−1 . . . w1w0 of elements wi ∈ Ω with w0 := e, and let 0 ≤ j < rn − 1 be the smallest integer such that wj . . . w1w0 ∈ Yknv−1 kn h(n), but wj+1wj . . . w1w0 6∈ Yknv−1 kn h(n). Then Ωwj . . . w1w0 6⊆ Yknv−1 kn h(n), hence wj . . . w1w0 ∈ ∂Ω(Yknv−1 kn h(n)). Since wj . . . w1w0 ∈ Ωrn−1, this contradicts the second assertion of (29), and the claim (30) follows. Roughly speaking, we used the fact that Ω-boundaries do not have gaps. Since PΩn → I strongly, we conclude from (30) that PYkn v−1 kn h(n) → I strongly. (31) Theorem 4.11 Let A ∈ SY (Sh(Γ)), and let h be a sequence such that the limit operator Op (A)h exists. Then in Case 1.2, Op (A)h = A with A := s-limAnPYn. (32) Proof. (PYnAPYn) with A ∈ Sh(Γ). For these sequences, one has It is sufficient to prove (32) for pure finite sections sequences A = h(n)(Op (A) + PΓ′)Rh(n) = X R−1 k6=kn R−1 h(n)Rvk PYkAPYkR−1 vk Rh(n)(I − PYkn v−1 h(n)) kn + R−1 + PYkn v−1 kn h(n)PΓ′Rh(n)(I − PYkn v−1 h(n)) kn h(n)APYkn v−1 kn h(n). Letting n go to infinity the assertion follows due to (31). 20 Thus, in Case 1.2, the invertibility of the limit operators of Op (A) + PΓ′ follows already from the invertibility of A. Now consider Case 2, i.e., suppose that none of the h(n) belongs to ∪vkY −1 . For n ∈ N, let rn stand for the smallest non-negative integer such that there is a kn ∈ N with h(n) ∈ vkn(∂ΩYkn)−1Ωrn. Consequently, k h(n) 6∈ vkn(∂ΩYkn)−1Ωrn−1 for all n. Again we set r∗ := lim inf rn and distinguish two cases. Case 2.1: r∗ is finite. We proceed as in Case 1.1 and find a subsequence of h (denoted by h again) and an element w∗ ∈ Γ such that h(n) ∈ vkn(∂ΩYkn)−1w∗. Since the inclusion h(n) ∈ vknY −1 kn in (21) had not been used in Case 1.1 we can continue exactly as in that case to obtain that Theorem 4.9 and its corollary hold verbatim in the case at hand, too. Case 2.2: r∗ is infinite. As in Case 1.2, we choose the sequence (rn) as strongly monotonically increasing. Then we have h(n) 6∈ vkY −1 k for all k, n, h(n) 6∈ vk(∂ΩYk)−1Ωrn−1 for all k, n. We claim that these two facts imply that Ωrn−1 ∩ Ykv−1 k h(n) = ∅ for all k, n. (33) (34) (35) Indeed, from (33) we conclude that e 6∈ Ykv−1 Ωrn−1 contains points from the complement of Ykv−1 e. Suppose that Ωrn−1 also contains points in Ykv−1 from Case 1.2 imply that Ωrn−1 contains points in the Ω-boundary of Ykv−1 But (34) implies that Ωrn−1 ∩ (∂ΩYk)v−1 located in the complement of Ykv−1 k h(n). Thus, for each k and n, k h(n), for instance the point k h(n). Then the arguments k h(n). k h(n) = ∅. Thus, Ωrn−1 is completely k h(n), whence (35). h(n)Rvk AkR−1 vk Rh(n) lives on im PYkv−1 k h(n), we obtain from Since the operator R−1 (35) h(n)(Op (A) + PΓ′)Rh(n) = X R−1 k≥1 R−1 h(n)Rvk AkR−1 vk Rh(n)(I − PΩrn−1) + R−1 h(n)PΓ′Rh(n)(I − PΩrn−1) + PΩrn−1. The first two summands on the right-hand side of this equality tend strongly to zero as n → ∞, whereas the third one tends strongly to the identity. Thus, the identity operator is the only limit operator of Op (A) + PΓ′ in Case 2.2. The following theorem summarizes the results from Cases 1.1 - 2.2. 21 Theorem 4.12 Let A ∈ SY (Sh(Γ)). Then the limit operators of Op (A) + PΓ′ are the identity operator I, the operator A := s-lim AnPYn, and all operators of the form s-lim R−1 w∗ R−1 ηkn AknRηkn Rw∗ + (PΓ′)g with a suitable subsequence g of h and with elements ηkn ∈ (∂ΩYkn)−1 and w∗ ∈ Γ. Combining this theorem with Theorems 4.3 (b) and 4.7 we arrive at the following stability results. Theorem 4.13 Let Γ be an exact discrete group, and let A ∈ SY (Sh(Γ)). The sequence A is stable if and only if the operator A := s-lim AnPYn and all operators of the form s-lim R−1 ηkn AknRηkn + Rw∗(PΓ′)gR−1 w∗ with a suitable subsequence g of h and with elements ηkn ∈ (∂ΩYkn)−1 and w∗ ∈ Γ are invertible and if the norms of their inverses are uniformly bounded. Corollary 4.14 Let Γ be an exact discrete group, and let A ∈ Sh(Γ). The se- quence A = (PYnAPYn) is stable if and only if the operator A and all operators PY (h)APY (h) : im PY (h) → im PY (h) where (36) with certain elements ηkn ∈ (∂ΩYkn)−1 are invertible and if the norms of their inverses are uniformly bounded. Y (h) := lim Yknηkn Theorem 4.8 allows us to remove the uniform boundedness condition in the pre- vious corollary. Corollary 4.15 Let Γ be a finitely generated discrete and exact group with sub- exponential growth which possesses at least one non-cyclic element, and let A ∈ Sh(Γ) be a band operator. Then the sequence A = (PYnAPYn) is stable if and only if the operators mentioned in the previous corollary are invertible. 4.4 Geodesic paths Now we turn to special sequences Y = (Yn) and η : N → Γ for which the existence of the set limit (36) can be guaranteed. Let again Ωn refer to the set of all products of at most n elements of Ω and set Ω0 := {e}. A sequence (νn) in Γ is called a geodesic path (with respect to Ω) if there is a sequence (wn) in Ω \ {e} such that νn = w1w2 . . . wn and νn ∈ Ωn \ Ωn−1 for each n ≥ 1. Note that this condition implies that each νn is in the right Ω-boundary of Ωn, which is the set of all w ∈ Ωn for which wΩ is not a subset of Ωn. We will see now that the lim Ωnηn exists if η is an inverse geodesic path, i.e., if ηn = ν−1 n for a geodesic path ν. 22 Lemma 4.16 Let (wn)n≥1 be a sequence in Ω and set ηn := w−1 n ≥ 1. Then the set limit lim Ωnηn exists, and n w−1 n−1 . . . w−1 1 for lim Ωnηn = ∪n≥1Ωnηn. (37) Proof. For n ≥ 1, one has Ωnηn = Ωnwn+1w−1 inclusions imply the existence of the set limit and the equality (37). n . . . w−1 n+1w−1 1 ⊆ Ωn+1ηn+1. These The natural question arises whether every sequence η : N → Γ for which the set limit (36) exists has a subsequence which is a subsequence of an inverse geodesic path. If the answer is affirmative, then it would prove sufficient to consider strong limits with respect to inverse geodesic paths in Theorem 4.13 and its corollary. We are going to answer this question for two special families of groups. 4.5 Commutative groups Let Γ be a commutative group which is generated, as a semi-group, by the finite set Ω with e ∈ Ω. Define Ωn as in the previous section. Proposition 4.17 Let Γ be commutative, and let µ = (µn)n∈N be a sequence in Γ which has a subsequence (µn)n∈N0 with µn ∈ Ωn \ Ωn−1 for each n ∈ N0. Then (µn)n∈N0 has a subsequence which is a subsequence of a geodesic path. . . . ωekn Proof. Let Ω = {e, ω1, . . . , ωk}. Each µn can be written as ωe1n where e1n + e2n + . . . + ekn = n for n ∈ N0. (We do not claim that this repre- sentation of µn is unique.) Consider the sequence (e1n)n∈N0. This sequence has a constant subsequence or a strongly monotonically increasing subsequence. Let (e1n)n∈N1 with an infinite subset N1 of N0 be a subsequence of (e1n)n∈N0 which owns one of these properties. Then consider (e2n)n∈N1 and choose a subsequence (e2n)n∈N2 which is constant or strongly monotonically increasing. We proceed in this way. After k steps we arrive at a subsequence (µn)n∈Nk of (µn)n∈N0 with µn = ωf1n and f1n + f2n + . . . + fkn = n for n ∈ Nk and where each of the sequences (fin) is either constant or strongly monotonically increasing. 1 ωf2n 1 ωe2n . . . ωfkn 2 2 k k For n ∈ Nk let νn := µn, and set ν0 := e. Let (kn) be the enumeration of the elements of Nk in increasing order, and set k0 := 0. In order to define νn for kr < n < kr+1 we proceed as follows. Let i1 be the smallest positive integer such that fi1kr < fi1kr+1. For l = 1, . . . , fi1kr+1 − fi1kr , set νkr+l := ωf1kr 1 . . . ω fi1−1,kr i1−1 ω fi1,kr +l i1 ω fi1+1,kr i1+1 . . . ωfkkr k . Now we are looking for the next subscript, say i2, for which the exponents at ωi2 of νkn and νkn+1 are different and proceed in the same way. After a finite number of steps, we arrive at a sequence ν = (νn)n∈N with νn ∈ Ωn for each n ∈ N. It remains to show that the sequence ν is a geodesic path, i.e. that νn ∈ Ωn \ Ωn−1 for each n. Suppose that νk 6∈ Ωk \ Ωk−1 for some k ≥ 2. Then νk is 23 a product of l < k elements from Ω \ {e}. Choose n such that kn > k and let a ∈ Ωkn−k such that µkn = aνk. Then µkn ∈ Ωkn−kΩl = Ωkn−k+l with kn − k + l < kn, a contradiction to the hypothesis that µkn ∈ Ωkn \ Ωkn−1 for each n. Since commutative groups are exact, one has the following consequences. Corollary 4.18 Let Γ be a commutative discrete group, and let Ω be a finite subset of Γ which generates Γ as a semi-group. Set Yn := Ωn, and let A ∈ SY (Sh(Γ)). The sequence A is stable if and only if the operator A := s-lim AnPΩn and, for each inverse geodesic path η, the operator s-lim R−1 ηn AnRηn : im P∪Ωnηn → im P∪Ωnηn are invertible and if the norms of the inverses of these operators are uniformly bounded. Corollary 4.19 Let Γ and Ω be as in Corollary 4.18, and let A ∈ Sh(Γ). The sequence A = (PΩnAPΩn) is stable if and only if the operator A and, for each inverse geodesic path η, the operator P∪ΩnηnAP∪Ωnηn : im P∪Ωnηn → im P∪Ωnηn are invertible and if the norms of their inverses are uniformly bounded. In many cases, there will be only finitely many different set limits lim Ωnηn; then the uniform boundedness condition in the previous corollaries is redundant. The same happens if A is a band operator by Theorem 4.8. The perhaps most important consequence of Corollary 4.18 is that the finite sections method for operators in Sh(Γ) is fractal. More general, one has the following. Corollary 4.20 Let Γ, Ω and Y be as in Corollary 4.18. Then the algebra SY (Sh(Γ)) is fractal. Roughly saying, an algebra of matrix sequences is fractal if each sequence in the algebra can be reconstructed from each of its (infinite) subsequences modulo a sequence tending to zero in the norm. For an exact definition and some properties of sequences in fractal algebras, see [9, 17]. The proof of Corollary 4.20 follows immediately from Corollary 4.18. See Theorem 1.69 in [9] and its corollary for the argument. 24 4.6 The free non-commutative group FN Proposition 4.17 does certainly not hold for all discrete groups. For example, let Γ = F2 with generators u and v, set Ω := {e, u±1, v±1}, and let Ωn stand for the set of all products of at most n elements of Ω. Consider ηn := vun−1. It is easy to see (indeed, drawing pictures will help a lot in what follows) that the set limit lim Ωnηn exists, but the sequence η has no subsequence which is a subsequence of an inverse geodesic path. On the other hand, a simple calculation gives lim Ωnηn = lim Ωn−1un−1; thus, the set limit lim Ωnηn coincides with another set limit which is taken with respect to an inverse geodesic path. We will see now that this observation is archetypal for the free non-commutative groups FN . Still for a moment, let Γ be a general discrete group with a finite set Ω of generators. Let (ηkn) be a sequence with ηkn ∈ (Ωkn \ Ωkn−1)−1 for each n. Write η−1 kn as (38) kn = ω(n) η−1 1 ω(n) 2 . . . ω(n) kn with ω(n) i ∈ Ω \ {e} for each i = 1, . . . , kn. Again, we do not claim that this representation is unique. Since Ω is finite, there is an ω1 ∈ Ω such that ω(n) 1 = ω1 for infinitely many n ∈ N, say for all n ∈ N1. By the same argument, there is an ω2 ∈ Ω such that ω(n) 2 = ω2 for infinitely many n ∈ N1, say for all n ∈ N2. We proceed in that way to obtain a sequence (ωn)n∈N in Ω \ {e} having the property that, for each r ∈ N, there are infinitely many elements ηkn with For r ≥ 1, set kn = ω1 ω2 . . . ωr ω(n) η−1 r+1 . . . ω(n) kn . ηr := (ω1 ω2 . . . ωr)−1. By Lemma 4.16, the set limit lim Ωr ηr exists. Lemma 4.21 Let ηkn and ηr be as in (38) and (40), respectively. Then lim r→∞ Ωr ηr ⊆ lim sup n→∞ Ωknηkn. Proof. Let x ∈ Ωr ηr for some r, and let η−1 kn be as in (39). Then (39) (40) (41) x ∈ Ωr ηr = Ωrω(n) r+1 . . . ω(n) kn (ω(n) kn )−1 . . . (ω(n) r+1)−1 ω−1 r . . . ω−1 1 ⊆ Ωknηkn. Since there are infinitely many elements as in (39), this inclusion implies that x ∈ lim sup Ωknηkn, whence ∪r≥1Ωr ηr ⊆ lim sup Ωknηkn. This is the assertion. It is one consequence of the lemma that the set limits lim Ωknηkn cannot be too 25 small. In particular, they contain a shifted copy of Ωr for each r and are, thus, growing sets in the sense of Shteinberg (see [23] and Definition 2.4.8 in [14]). In general, one cannot expect that equality holds in (41). For example, let Γ be the (additively written) group Z2 with Ω = {(0, 0), (±1, 0), (0, ±1)} and consider the sequence η2n = (−n, −n). If we write −η2n as −η2n = (1, 0) + . . . + (1, 0) + (0, 1) + . . . + (0, 1) with each summand occurring n times, then the above construction yields ηr := (−r, 0). In this setting, both set limits lim Ω2nη2n and lim Ωr ηr exist, but they do not coincide (the first one is the intersection of Z2 with a half plane, the second one with a quadrant). It turns out that, in case of the free non-commutative groups FN , equality holds in (41). Theorem 4.22 For N > 1, let FN be the free group generated by its elements ω1, . . . , ωN , set Ω := {e, ω±1 N }, and let Ωn be the set of all products of elements of Ω of length at most n. Further let (ηkn) be a sequence with 1 , . . . , ω±1 ηkn ∈ (Ωkn \ Ωkn−1)−1 which we write as in (39) and let (ηr) be the associated sequence as in (40). Then lim inf n→∞ Ωknηkn ⊆ lim r→∞ Ωr ηr. In particular, if the set limit limn→∞ Ωknηkn exists, then lim n→∞ Ωknηkn = lim r→∞ Ωr ηr. (42) (43) Proof. Let x ∈ lim inf Ωknηkn. Then there is an n0 such that x ∈ Ωknηkn for n ≥ n0. Thus, for each n ≥ n0, x ∈ Ωkn(ω(n) kn )−1 . . . (ω(n) n+1)−1 ω−1 n . . . ω−1 1 . Choose elements ν(n) i in Ω such that x = ν(n) 1 . . . ν(n) kn {z } (∗∗) (ω(n) kn )−1 . . . (ω(n) n+1)−1 ω−1 {z n . . . ω−1 1 } (∗) . (44) The assumption ηkn ∈ (Ωkn \ Ωkn−1)−1 guarantees that there is no cancelation possible inside part (∗) of the representation (44) but, of course, there might be cancelation inside part (∗∗) as well as between the most right of the ν and the most left of the ω−1. For each n ≥ n0, we cancel the representation (44) of x as far as possible. k )−1 remains in Suppose that, after complete cancelation, at least one factor (ω(n) 26 each representation. Then, for each n ≥ 1, we can represent x as a word without n . . . ω−1 further cancelation which starts from the right-hand side with . . . ω−1 1 and, hence, has length at most n. This is impossible since each x ∈ FN can be uniquely represented as a reduced word of finite length. This contradiction shows that there is at least one n ≥ n0 such that all factors (ω(n) k )−1 in the representation (44) can be canceled. Thus, x ∈ Ωk ω−1 1 = Ωk ηk for some k ≥ n0. Since k the set sequence (Ωk ηk) is monotonically increasing, this implies . . . ω−1 x ∈ ∪k≥1 Ωk ηk = lim Ωk ηk whence the first assertion. Combining this result with Lemma 4.21, the second assertion follows. Thus, each set limit lim Ωknηkn can be obtained as a set limit along an inverse geodesic path. Since free groups are exact, this leads to the same consequences as for commutative groups. Corollary 4.23 Let Γ = FN and Ω and Ωn as in Theorem 4.22. Set Yn := Ωn, and let (An) ∈ SY (Sh(FN )). The sequence (An) is stable if and only if the operator A := s-lim AnPΩn and, for each inverse geodesic path η, the operator s-lim R−1 ηn AnRηn : im P∪Ωnηn → im P∪Ωnηn are invertible and if the norms of the inverses of these operators are uniformly bounded. Corollary 4.24 Let Ω be as in Corollary 4.23 and let A ∈ Sh(FN ). The sequence A = (PΩnAPΩn) is stable if and only if the operators A and, for each inverse geodesic path η, P∪ΩnηnAP∪Ωnηn : im P∪Ωnηn → im P∪Ωnηn are invertible and if the norms of their inverses are uniformly bounded. Corollary 4.25 Let Ω, Y be as in Corollary 4.23. Then the algebra SY (Sh(FN )) is fractal. References [1] T. Adachi, A note on the Følner condition for amenability. -- Nagoya Math. J. 131(1993), 67 -- 74. [2] B. Blackadar, Operator Algebras. Theory of C ∗-Algebras and von Neu- mann Algebras. -- Springer, Berlin, Heidelberg 2006. 27 [3] A. Bottcher, S. M. Grudsky, Spectral Properties of Banded Toeplitz Matrices. -- siam, Philadelphia 2005. [4] A. Bottcher, B. Silbermann, Introduction to Large Truncated Toeplitz Matrices. -- Springer-Verlag, Berlin, Heidelberg 1999. [5] N. P. Brown, N. Ozawa, C ∗-Algebras and Finite-Dimensional Approxi- mations. -- Graduate Studies Math. 88, Amer. Math. Soc., Providence, R. I., 2008. [6] L. A. Coburn, The C ∗-algebra generated by an isometry. -- Bull. Am. Math. Soc. 73(1967), 722 -- 726. [7] K. R. Davidson, C ∗-Algebras by Example. -- Fields Institute Monographs Vol. 6, Providence, R. I., 1996. [8] R. Hagen, S. Roch, B. Silbermann, Spectral Theory of Approximation Methods for Convolution Equations. -- Birkhauser Verlag, Basel, Boston, Berlin 1995. [9] R. Hagen, S. Roch, B. Silbermann, C ∗-Algebras and Numerical Anal- ysis. -- Marcel Dekker, Inc., New York, Basel 2001. [10] M. Lindner The finite section method and stable subsequences. -- Preprint 15/2008 TU Chemnitz. [11] V. S. Rabinovich, S. Roch, Fredholm properties of band-dominated op- erators on periodic discrete structures. -- to appear in Complex Anal. Oper. Theory. [12] V. S. Rabinovich, S. Roch, B. Silbermann, Fredholm theory and finite section method for band-dominated operators. -- Integral Equations Oper. Theory 30(1998), 452 -- 495. [13] V. S. Rabinovich, S. Roch, B. Silbermann, Band-dominated opera- tors with operator-valued coefficients, their Fredholm properties and finite sections. -- Integral Eq. Oper. Theory 40(2001), 3, 342 -- 381. [14] V. S. Rabinovich, S. Roch, B. Silbermann, Limit Operators and their Applications in Operator Theory. -- Oper. Theory: Adv. Appl. 150, Birkhauser, Basel 2004. [15] V. S. Rabinovich, S. Roch, B. Silbermann, On finite sections of band- dominated operators. -- Preprint 2486 TU Darmstadt, December 2006, 7 p., submitted to Proc. WOAT Lisbon 2006. 28 [16] M. Reed, B. Simon, Methods of Modern Mathematical Physics. Volume 1: Functional Analysis. -- Academic Press, New York, London 1972. [17] S. Roch, Algebras of approximation sequences: Fractality. -- In: Oper. Theory: Adv. Appl. 121, Birkhauser, Basel 2001, 471 -- 497. [18] S. Roch, Finite sections of band-dominated operators. -- Memoirs AMS Vol. 191, 895, Providence, R.I., 2008. [19] S. Roch, Spatial discretization of Cuntz algebras. -- Houston Math. J., to appear. [20] S. Roch, B. Silbermann, C ∗-algebra techniques in numerical analysis. -- J. Oper. Theory 35(1996), 2, 241 -- 280. [21] J. Roe, Lectures on Coarse Geometry. Univ. Lecture Ser. 31, Amer. Math. Soc., Providence, R. I., 2003. [22] J. Roe, Band-dominated Fredholm operators on discrete groups. -- Integral Equations Oper. Theory 51(2005), 3, 411 -- 416. [23] B. Ya. Shteinberg, Compactification of locally compact groups and Fred- holmness of convolution operators with coefficients in factor groups. -- Tr. St-Peterbg. Mat. Obshch. 6(1998), 242 -- 260 (Russian). Author's address: Steffen Roch, Technische Universitat Darmstadt, Fachbereich Mathematik, Schloss- gartenstrasse 7, 64289 Darmstadt, Germany. E-mail: [email protected] 29
1608.06426
2
1608
2019-02-04T06:49:03
On fundamental groups of tensor product $\rm II_1$ factors
[ "math.OA" ]
Let $M$ be a $\rm II_1$ factor and let $\mathcal{F}(M)$ denote the fundamental group of $M$. In this article, we study the following property of $M$: for arbitrary $\rm II_1$ factor $B$, we have $\mathcal{F}(M \overline{\otimes} B)=\mathcal{F}(M)\mathcal{F}(B)$. We prove that for any subgroup $G\leq \mathbb{R}^*_+$ which is realized as a fundamental group of a $\rm II_1$ factor, there exists a $\rm II_1$ factor $M$ which satisfies this property and whose fundamental group is $G$. Using this, we deduce that if $G,H \leq \mathbb{R}^*_+$ are realized as fundamental groups of $\rm II_1$ factors (with separable predual), then so are groups $G \cdot H$ and $G \cap H$.
math.OA
math
On fundamental groups of tensor product II1 factors Yusuke Isono∗ Abstract Let M be a II1 factor and let F (M ) denote the fundamental group of M . In for any II1 factor B, we have this article, we study the following property of M : F (M ⊗ B) = F (M )F (B). We prove that for any subgroup G ≤ R∗ + which is realized as a fundamental group of a II1 factor, there exists a II1 factor M which satisfies this property and whose fundamental group is G. Using this, we deduce that if G, H ≤ R∗ + are realized as fundamental groups of II1 factors, then so are groups G · H and G ∩ H. 1 Introduction and main theorems In their pioneering work, Murray and von Neumann introduced the fundamental group as an invariant of II1 factors [MV43]. For a II1 factor M with trace τ , the fundamental group is defined as F(M ) :=(cid:26) τ (p) τ (q) ∈ R∗ +(cid:12)(cid:12)(cid:12)(cid:12) p, q are projections in M with pM p ≃ qM q(cid:27) . Murray and von Neumann proved the hyperfinite (or amenable) II1 factor has the full fundamental group R∗ Indeed, the fundamental group is the most well known invariant for II1 factors, and to determine which subgroup of R∗ + appears as a fundamental group is a long-standing open problem in the von Neumann algebra theory. +, and then asked the general behavior of this invariant. Computation of fundamental groups, however, is a hard problem. Indeed, II1 factors pM p and qM q share a lot of properties in common, so it is very difficult to distinguish them. Thus very few computations have been done until recently. Connes proved that LΓ, where Γ is an ICC property (T) group, has a countable fundamental group [Co80], which is the first example of a II1 factor with fundamental group not equal to R∗ +. Voiculescu and Radulescu proved F(LF∞) has the full fundamental group R∗ + [Vo89, Ra91]. In 2001, Popa introduced a new framework to study this problem [Po01]. He developed a way of identifying Cartan subalgebras and then reduced this computation problem for a certain class of II1 factors to the one for corresponding orbit equivalence relations. Thus combined with Gaboriau's work on orbit equivalence relations [Ga99, Ga01], Popa obtained the first example of a II1 factor which has the trivial fundamental group. Much progress has been made by this new technology in the von Neumann algebra theory. The study in this new framework is now called the deformation/rigidity theory. Thus, a lot of computations of fundamental groups have been done in the last decade. We say that a subgroup G ≤ R∗ + is in the class Sfactor if there is a II1 factor M with separable predual such that F(M ) = G. Popa proved that any countable subgroup ∗Research Institute for Mathematical Sciences, Kyoto University, 606-8502, Kyoto, Japan E-mail: [email protected] 1 of R∗ + is contained in Sfactor [Po03]. Popa and Vaes proved that Sfactor contains many uncountable subgroups in R∗ + [PV08a]. See [Po04, IPP05, Po06a, Ho07, PV08b, De10] for other calculations of fundamental groups. We note that, while this new theory provides a lot of concrete examples, very few general properties for the class Sfactor are known so far. (See Proposition 2.1 below.) The aim of this article is to study fundamental groups of tensor product II1 factors. For this, recall that for a II1 factor M and t > 0, the amplification M t is defined (up to ∗-isomorphism) as pM p ⊗ Mn for any n ∈ N with t ≤ n and any projection p ∈ M with trace t/n. It is then easy to verify that • F(M ) = {t ∈ R∗ + M ≃ M t}; • (M1 ⊗ M2)st ≃ M s 1 ⊗ M t 2 for II1 factors Mi and s, t > 0. They particularly imply the following inclusion: F(M1 ⊗ M2) ⊃ F(M1)F(M2) for any II1 factors M1 and M2. It is likely that the converse inclusion also holds true at the first glance. However, as we emphasized, the computation of fundamental groups is a hard problem, and so we know very little about this converse inclusion until recently. To study this inclusion is actually the main purpose of this article. This is a natural question, since it provides a quite useful formula for fundamental groups of tensor product II1 factors. Indeed, if the converse inclusion holds, then one actually has an equation, so the computation of fundamental groups for the tensor product can be reduced to the one for each tensor component. Here we state this problem in the following precise form. Question. For which II1 factors M1 and M2, do we obtain the equation F(M1 ⊗ M2) = F(M1)F(M2)? We do not believe that all II1 factors satisfy it, although, to the best of our knowledge, any example of II1 factors which do not satisfy this equation is not known. In the deformation/rigidity theory, Ozawa and Popa provided the first class of II1 factors that satisfy this equality. They proved that if each Mi is a free group factor, then the tensor product satisfies a unique prime factorization result and particularly the equation above holds true [OP03]. See [Pe06, Sa09, CSU11, SW11, Is14, CKP14, HI15, Ho15] for other classes of factors which satisfy the unique prime factorization result. In this article, we further develop Ozawa -- Popa's strategy. We particularly study the following property for a II1 factor M : F(M ⊗ B) = F(M )F(B) for any II1 factor B. Obviously this condition is stronger than the one Ozawa-Popa obtained for free groups factors, since the factor B in the condition can be arbitrary. If this condition holds, we say that M satisfies the tensor factorization property for fundamental groups (say, property (TFF) in short). Our first theorem provides examples of II1 factors having property (TFF). See [BO08, Definitions 12.3.1 and 15.1.2] for definitions of weak amenability and bi-exactness (and note that free groups, more generally hyperbolic groups, satisfy them). Theorem A. Let M be one of the following II1 factors. • A group II1 factor LΓ, where Γ is an ICC, non-amenable, weakly amenable, and bi-exact group. 2 • A free product II1 factor M1 ∗ M2, where M1 and M2 are diffuse (and tracial). • A group II1 factor L(∆ ≀ Λ), where ∆ is a non-trivial amenable group and Λ is a non-amenable group. Then M satisfies the property (TFF). As a corollary of this theorem, we obtain the main observation of this article. In fact, the following corollary states general properties for the class Sfactor. Although it is not enough to answer the aforementioned question by Murray and von Neumann, this is an interesting consequence since there are very few general properties for the class Sfactor as we mentioned. Corollary B. For any G ∈ Sfactor, there is a II1 factor M with separable predual and with the property (TFF) such that F(M ) = G. The class Sfactor admits the following properties. • Stability under multiplication: for any G, H ∈ Sfactor, the group G · H is in Sfactor. • Stability under countable intersection: for any Gn ∈ Sfactor (possibly Gn = Gm for n 6= m), n ∈ N, the group Tn Gn is in Sfactor. We note that the proof of the first statement in this corollary in fact shows the following: if we put N := LFn ∗ L(Z2 ⋊ SL(2, Z)), then for any II1 factor B we have F(B) = F(∗N(B ⊗ N )). Thus combined with Theorem A, the free product II1 factor ∗N(B ⊗ N ) does the work. The proof of Theorem A uses the idea in our previous paper [Is14], in which we introduced another notion of primeness for II1 factors. Recall that a II1 factor M is said to be prime if it does not have a tensor decomposition as II1 factors, namely, if it has a decomposition M = M1 ⊗ M2, then at least one Mi must be of type I. Obviously this definition comes from the notion of prime numbers in the number theory. Actually there are two equivalent notions of prime numbers. Recall that a number p ∈ N is irreducible if for any q, r ∈ N with p = qr, we have q = 1 or r = 1; and is prime if for any q, r, s ∈ N with pq = rs, we have p r or p s. In the von Neumann algebra theory, we adapt the first one (i.e. irreducibility) as a definition of primeness. In [Is14, Section 5], we introduced a different notion of primeness, which corresponds to the second one as follows. To distinguish two primeness, we name it strongly prime. • We say a II1 factor M is strongly prime if for any II1 factors B, K and L with M ⊗ B = K ⊗ L, there is a unitary u ∈ U (M ⊗ B) and t > 0 such that, under the identification K ⊗ L = K t ⊗ L1/t, we have uM u∗ ⊂ K t or uM u∗ ⊂ L1/t. Here we identify each tensor component as a subalgbera (e.g. M = M ⊗ C ⊂ M ⊗ B). Our second main theorem treats examples of strongly prime factors. Note that the first item in this theorem was already obtained in our previous article [Is14, Theorem 5.1]. We also note that the first and the second item in the theorem treat exactly the same ones as in Theorem A. Theorem C. Let M be one of the following II1 factors. • A group II1 factor LΓ, where Γ is a non-amenable, ICC, weakly amenable, and bi-exact group. • A free product II1 factor M1 ∗ M2, where M1 and M2 are diffuse. 3 • A group II1 factor L(∆≀Λ), where ∆ is a non-trivial amenable group and Λ = Λ1×Λ2 is a direct product of any group (possibly trivial) Λ1 and a non-amenable, weakly amenable, and bi-exact group Λ2. Then M is strongly prime. In section 3, we will show that the property (TFF) has a sufficient condition similar to strong primeness (Lemma 3.4), and hence strong primeness is actually a sufficient condition to the property (TFF) (Proposition 3.6). We note that strong primeness implies primeness, but the converse fails (Propositions 3.6 and 3.7). We will also discuss unique prime factorization result, using the strong primeness. This particularly provides the first example of unique prime factorization result for infi- nite tensor products. Below we say that a II1 factor M is semiprime if for any tensor decomposition M = M1 ⊗ M2, at least one Mi is amenable. The reason we use semiprime- ness is that any infinite tensor product factor M is McDuff (i.e. M ≃ M ⊗ R for the hyperfinite II1 factor R), so tensor components are determined up to tensor product with R. Proposition D. Let m, n ∈ N∪{∞}. Let Mi be strongly prime II1 factors, and Nj any II1 factors such that ⊗m j=1Nj =: M . Then there is a unique map σ : {1, 2, . . . , m} → {1, 2, . . . , n} such that Mi (cid:22)M Nσ(i) for all i ∈ {1, 2, . . . , m}. In this case, the following statements hold true. i=1Mi = ⊗n • The map σ is surjective if and only if all Nj are non-amenable. • The map σ is injective if and only if all Nσ(i) are semiprime. Thus the map σ is bijective if all Nj are non-amenable and semiprime. In this case for each i ∈ {1, 2, . . . , m}, Nσ(i) is isomorphic to M ti i ⊗ Pi for some ti > 0 and some amenable factor Pi. In the proposition, if we assume all Nj are prime, then the map σ is bijective and Mi and Nσ(i) are stably isomorphic for all i. We note that the map σ in the proposition is surjective whenever m < ∞, since M = M1 ⊗ · · · ⊗ Mm is full (Proposition 3.7 and [Co75, Corollary 2.3]) and so Nj can not be amenable. Acknowledgement. The author would like to thank R´emi Boutonnet, Cyril Houdayer, Adrian Ioana, Narutaka Ozawa and Stefaan Vaes for fruitful conversations. He also thank the referee for pointing out that a part of the proof of Theorem C can be simplified using Lemma 2.6. He was supported by JSPS, Research Fellow of the Japan Society for the Promotion of Science. 2 Preliminaries In this article, all von Neumann algebras that we consider are assumed to be finite and σ-finite, namely, they admit faithful normal tracial states. General properties for the class Sfactor Let M be a II1 factor and Tr a trace on M ⊗ B(ℓ2). Then since the trace on M ⊗ B(ℓ2) is unique up to scalars, there is a homomorphism Mod : Aut(M ⊗ B(ℓ2)) → F(M ); Tr ◦ α = Mod(α)Tr, α ∈ Aut(M ). 4 It is then not difficult to see that Mod is surjective and continuous (with respect to the u-topology on Aut(M ⊗ B(ℓ2))). Since Aut(M ⊗ B(ℓ2)) with the u-topology is a Polish group when M has separable predual, we get the following proposition. This is the only known general property for the class Sfactor so far. Proposition 2.1. For any group G ∈ Sfactor, there is a Polish group P and a continuous surjective homomorphism from P onto G. Using this, one can show for example that any group in Sfactor is a Borel subset of R∗ +. Our main observation will provide the second general property for Sfactor. Popa's intertwining technique We recall Popa's intertwining theorem. This is the main tool in the deformation/rigidity theory. Theorem 2.2 ([Po01, Po03]). Let M be a finite von Neumann algebra with trace τ , p and q projections in M , A ⊂ pM p and B ⊂ qM q von Neumann subalgebras with τ -preserving conditional expectations EA and EB. Then the following conditions are equivalent. (i) There exist non-zero projections e ∈ A, f ∈ B, a unital normal ∗-homomorphism θ : eAe → f Bf , and a partial isometry v ∈ eM f such that vθ(x) = xv for all x ∈ eAe. (i)′ There exist a nonzero normal ∗-homomorphism ψ : A → B ⊗Mn for some n ∈ N and a nonzero partial isometry w ∈ (p ⊗ e1,1)(M ⊗ Mn) such that wψ(x) = (x ⊗ e1,1)w for all x ∈ A, where (ei,j)i,j is a fixed matrix unit in Mn. (ii) There exists no net (wi)i of unitaries in A such that kEB(b∗wia)k2,τ → 0 for any a, b ∈ pM q. (iii) There exists a positive element d ∈ phM, Bip ∩ A′ such that 0 < TrhM,Bi(d) < ∞, where TrhM,Bi is the canonical trace on hM, Bi (with respect to τ ). We write A (cid:22)M B if one of these conditions holds. Note that when B = C, A 6(cid:22)M C if and only if A is diffuse. We next observe some elementary lemmas. Lemma 2.3 ([HI15, Lemma 4.6]). Let M and N be finite von Neumann algebras, p a projection in M , A ⊂ M , N0 ⊂ N and B ⊂ M finite von Neumann subalgebras. Then A (cid:22)M B if and only if A ⊗ C1N (cid:22)M ⊗N B ⊗ C1N if and only if A ⊗ N0 (cid:22)M ⊗N B ⊗ N . Lemma 2.4. Let B be a finite von Neumann algebra and Γ a discrete group acting on B as a trace preserving action. Write M := B ⋊ Γ. Then M (cid:22)M B if and only if LΓ (cid:22)M B if and only if Γ is a finite group. Proof. If Γ is a finite group, then the canonical trace of the basic construction hM, Bi is finite. So by Theorem 2.2(iii), we get M (cid:22)M B. If Γ is infinite, then we can find a sequence gn ∈ Γ such that all gn are distinct with each other. Then it satisfies Theorem 2.2(ii) and hence LΓ 6(cid:22)M B. Finally by Theorem 2.2(i)′, it is obvious that M (cid:22)M B implies LΓ (cid:22)M B. Lemma 2.5. Let M = M1 ∗ M2 be a tracial free product von Neumann algebra, p ∈ M1 a projection, and let A ⊂ pM1p be a diffuse von Neumann subalgebra. Then we have A 6(cid:22)M M2. 5 Proof. We may assume M2 6= C. Let (un)n be a sequence of unitaries in A which converges to 0 weakly. By simple calculations, one can show that if each a, b ∈ M is a scalar or a reduced word, then kEM2(b∗una)k2 converges to 0 as n → ∞. Hence by Theorem 2.2 (ii), A 6(cid:22)M M2 holds. In the lemma below, we denote the normalizer for an inclusion B ⊂ M by NM (B) := {u ∈ U (M ) uBu∗ = B}. Lemma 2.6. Let B ⊂ M be finite von Neumann algebras, p a projection in M , and A, P ⊂ pM p von Neumann subalgebras. Assume that A and P commute. Assume A (cid:22)M B and P (cid:22)M B. If B is regular (i.e. NM (B)′′ = M ) and NpM p(A)′ ∩ pM p = Cp, then we have (A ∨ P ) (cid:22)M B. Proof. We follow the proof of [Sa09, Lemma 33]. By Theorem 2.2 (iii), we find a positive element dA ∈ A′ ∩ phM, Bip with 0 < TrhM,Bi(d) < ∞. Taking a spectral projection, we may assume dA is a projection. Ob- serve that for any u ∈ NpM p(A) and v ∈ NM (B), the element vopudAu∗(vop)∗ satisfies the same condition as the one on dA, where vop is the right action of v on L2(M ) (we indeed have TrhM,Bi ◦ Ad vop = TrhM,Bi, since (vop)∗veBvopv∗ = eB. See [BO08, Exercise F.6] for the construction of TrhM,Bi and use the fact that TrhM,Bi is uniquely determined by TrhM,Bi(x∗eBx) = τ (x∗x) for x ∈ M ). So the element d := sup{vopudAu∗(vop)∗ u ∈ NpM p(A), v ∈ NM (B)} is contained in A′ ∩ NpM p(A)′ ∩ phM, Bip ∩ p(NM (B)op)′p = NpM p(A)′ ∩ phM, Bip ∩ pM p = NpM p(A)′ ∩ pM p = Cp. Hence we get d = p. Let now dP be a non-zero trace finite projection in P ′ ∩ phM, Bip. Then since d = p, there are finite subsets E ⊂ NpM p(A) and F ⊂ NM (B) satisfying that ∨u∈E,v∈F vopudAu∗(vop)∗ is not orthogonal to dP . Thus up to exchanging dA with this element, we can assume dAdP 6= 0. Consider a convex subset K := cow{udAu∗ u ∈ U (P )} ⊂ hM, Bi and observe that K is regarded as a subset in L2(hM, Bi, TrhM,Bi) which is L2-norm bounded (e.g. [BO08, Exercise F.3]). Take the unique minimal L2-norm element ed in K. We have uedu∗ = ed for any u ∈ U (P ) by the uniqueness, and hence ed is contained in P ′ ∩ A′ ∩ phM, Bip = (A ∨ P )′ ∩ phM, Bip. Observe that ed is trace finite in hM, Bi since so is dA (and TrhM,Bi is normal). Finally ed is non-zero since for any u ∈ U (P ), hudAu∗, dP i = TrhM,Bi(udAu∗dP ) = TrhM,Bi(dAdP ) > 0, and so any a ∈ K satisfies ha, dP i = TrhM,Bi(dAdP ) > 0. Thus we obtain (A ∨ P ) (cid:22)M B. Relative amenability We next recall relative amenability introduced in [OP07]. Definition 2.7 ([OP07, Definition 2.2]). Let M be a finite von Neumann algebra with trace τ . Let p ∈ M be a projection and A ⊂ pM p and B ⊂ M von Neumann subalgebras. We say A is amenable relative to B in M , and write as A⋖M B, if there exists a conditional expectation from phM, Bip onto A which restricts to a τ -preserving expectation on pM p. Proposition 2.8 ([OP07, Proposition 2.4(3)]). Let B ⊂ M and A ⊂ pM p as above, and let N ⊂ M be another von Neumann subalgebra. If A ⋖M B and B ⋖M N , then A ⋖M N . 6 We record the following elementary lemma. Lemma 2.9. Let M and B be finite von Neumann algebras. Then M ⊗ B ⋖ only if M is amenable. M ⊗B B if and 3 Some observations on tensor product II1 factors In this section, we briefly review fundamental properties of tensor product II1 factors to study the property (TFF) and strong primeness. We say M1 ⊗· · ·⊗Mm = N1 ⊗· · · ⊗Nn is a tensor decomposition as II1 factors if each Mi and Nj is a II1 factor. Property (TFF) We first recall the following observation of Ozawa and Popa. This shows that, to see a unitary embedding on tensor products, we have only to find Popa's conjugacy "(cid:22)" introduced in Theorem 2.2. This allows us to reformulate strong primeness (Lemma 3.5), so that we can make use of results in the deformation/rigidity theory. Lemma 3.1 ([OP03, Proposition 12]). Let M1 ⊗ M2 = N1 ⊗ N2 (=: M ) be a tensor decomposition as II1 factors. Then N1 (cid:22)M M1 if and only if there is a unitary element u ∈ M and a decomposition M = M t for some t > 0 such that uN1u∗ ⊂ M t 1. 1 ⊗ M 1/t 2 Remark 3.2. In this lemma, we are having an identification M = M t , using a non-canonical isomorphism M1 ⊗ M2 ≃ M t . Since this isomorphism is given at the level of a partial isometry conjugacy of M1 ⊗ M2 ⊗ Mn (for some large n ∈ N), one can show that N1 (cid:22)M M1 if and only if N1 (cid:22)M M t 1 for any t > 0 and any such an identification M1 ⊗ M2 = M t . So we do not need to be careful to identify M1 ⊗ M2 with M t in the study of Popa's conjugacy. 1 ⊗ M 1/t 1 ⊗ M 1/t 2 2 1 ⊗ M 1/t 2 1 ⊗ M 1/t 2 Here we record a simple but very useful lemma on tensor product factors. Lemma 3.3. Let M1 ⊗ M2 = N1 ⊗ N2 (=: M ) be a tensor decomposition as II1 factors and assume that M1 ⊂ N1. Then M ′ 1 ∩ N1 is a factor and satisfies M2 = (M ′ 1 ∩ N1) ⊗ N2 and N1 = M1 ⊗ (M ′ 1 ∩ N1). Proof. Since M1 ⊂ N1, we have M2 = M ′ 1 ∩ M = M ′ 1 ∩ (N1 ⊗ N2) = (M ′ 1 ∩ N1) ⊗ N2. 1 ∩ N1 is a factor. We have M = M1 ⊗ (M ′ 1 ∩ N1) ⊗ N2 and hence N1 = N ′ 2 ∩ M = So M ′ M1 ⊗ (M ′ 1 ∩ N1). The following lemma is a key observation in this paper, which states a sufficient condi- tion to the property (TFF) in terms of Popa's conjugacy. Although its proof is easy, this lemma plays significant roles in our study. Lemma 3.4. Let M be a prime II1 factor satisfying the following condition. • For any II1 factor B and any t > 0 such that M ⊗ B ≃ M ⊗ Bt (=: K ⊗ L), under the identification M ⊗ B = K ⊗ L, we have either K (cid:22)M ⊗B B, L (cid:22)M ⊗B B, M (cid:22)M ⊗B L, or B (cid:22)M ⊗B L. 7 Then M has the property (TFF). Proof. Fix a II1 factor B and take t ∈ F(M ⊗ B). We will show t ∈ F(M )F(B). Fix an isomorphism M ⊗ B ≃ (M ⊗ B)t ≃ M ⊗ Bt (=: K ⊗ L). By assumption, regarding N := M ⊗ Bt = K ⊗ L, we have either K (cid:22)N B, L (cid:22)N B, M (cid:22)N L, or B (cid:22)N L. If L (cid:22)N B, then by Lemma 3.1 there exists s > 0 and u ∈ U (N ) such that uLu∗ ⊂ Bs under the isomorphism M ⊗ B = M 1/s ⊗ Bs. For simplicity we assume u = 1. Then by Lemma 3.3, putting P := L′ ∩ Bs, it holds that Bs = L ⊗ P and K = P ⊗ M 1/s. Since K(= M ) is prime, P is finite dimensional. Write P = Mn for some n ∈ N and we obtain Bs = Ln and K = M n/s. Since Ln = Btn and K = M , this implies that s/tn ∈ F(B) and n/s ∈ F(M ), and hence 1 t = s tn · n s ∈ F(B) F(M ). Thus t ∈ F(B) F(M ). Next assume K (cid:22)N B. Then by the same reasoning as above, there exists s > 0 and u ∈ U (N ) such that uKu∗ ⊂ Bs. We assume u = 1. Putting Q := K ′ ∩ Bs it holds that Bs = K ⊗ Q and L = Q ⊗ M 1/s. Since K = M and L = Bt, these equations imply Bs = M ⊗ Q and Bt = Q ⊗ M 1/s, and hence Bs = M ⊗ Q ≃ Q ⊗ M = Bts. This implies t ∈ F(B) and we obtain the conclusion. Finally assume that M (cid:22)N L or B (cid:22)N L. Then since K = M and L = Bt, if we et, and eK ⊗ eL = M ⊗ eB, we can apply exactly put eB := Bt, et := 1/t, eK := M , eL := eB the same argument as in the previous two cases, and obtain et = 1/t ∈ F(M )F(eB). Since F(eB) = F(Bt) = F(B), we obtain the conclusion. Strong primeness We study fundamental properties on strong primeness. We first give a reformulation of strong primeness in terms of Popa's conjugacy. Lemma 3.5. A II1 factor M is strongly prime if and only if for any tensor decomposition M ⊗ B = K ⊗ L as II1 factors, we have either K (cid:22) B or L (cid:22) B. Proof. Use Lemma 3.1 and [Va08, Lemma 3.5]. We deduce primeness from strong primeness. This is not entirely trivial since, in the definition of strong primeness, we mention only a decomposition as II1 factors. Proposition 3.6. Strong primeness implies primeness. In particular any strongly prime II1 factor satisfies the property (TFF). 8 Proof. Let M be a non-prime II1 factor with a decomposition M = M1 ⊗ M2 as II1 factors. Fix any II1 factor B and put K := M1, L := M2 ⊗ B, and N := M ⊗ B = K ⊗ L. Then if M is strongly prime, we have either M (cid:22)N K or M (cid:22)N L. By Lemma 2.3, the first one is equivalent to M2 (cid:22)M2 C. Thus in each case, we get a contradiction. Use Lemmas 3.4 and 3.5 for the second assertion. C and the second one is to M1 (cid:22)M1 We observe the difference between the two notions of primeness discussed above. This follows from [Ho15, Theorem B]. Proposition 3.7. Any strongly prime II1 factor is full. In particular there is a prime II1 factor, which is not strongly prime. Proof. Let M and B be non-full II1 factors. Then by [Ho15, Theorem B], there is an automorphism φ on M ⊗ B such that φ(M ) 6(cid:22)M ⊗B B and φ(B) 6(cid:22)M ⊗B B. Thus the decomposition M ⊗ B = φ(M ) ⊗ φ(B) shows that M is not strongly prime. Let F2 y X be a free, ergodic, and measure preserving action of the free group on a standard probability space. Assume that it is not strongly ergodic. Then the crossed product M := L∞(X)⋊F2 is a prime II1 factor by [Oz04, Theorem 4.6], and is not strongly prime since it is not full. 4 Proof of Proposition D We study a unique prime factorization phenomena, by using our strong primeness. This was already mentioned in our previous paper [Is14, Corollary 5.1.3], that shows strongly prime factors behave like prime numbers with respect to von Neumann algebra tensor products. We only discussed the case of tensor products with finitely many strongly prime factors. So in this paper, we study the case of infinite tensor products. We start with several lemmas. Lemma 4.1. Let M be a strongly prime II1 factor and let M ⊗ B = N1 ⊗ · · · ⊗ Nn (=: N ) be a tensor decomposition as II1 factors with n ≥ 2. Then there is i such that M (cid:22)N Ni. Proof. We prove it by induction on n. The case n = 2 is obvious by the definition of strong primeness. So assuming n − 1 ≥ 2 is proven, we show the case n holds. 1 ∩ N . Then since M ⊗ B = N1 ⊗ N ′ 1 := N ′ 1. Since M (cid:22)N N1 implies the conclusion, we may assume M (cid:22)N N ′ Put N ′ 1, we have either M (cid:22)N N1 or M (cid:22)N N ′ 1. By 1)t. Then by Lemma 3.3 we Lemma 3.1 we find u ∈ U (N ) and t > 0 such that uM u∗ ⊂ (N ′ have uM u∗ ⊗ P = (N ′ 1)t. Observe that P is a II1 factor. In fact, if P is finite dimensional, then because M is prime, n must be 2 which contradicts our assumption. 1)t, where P = (uM u∗)′ ∩ (N ′ Now we can apply strong primeness of M and the assumption on the induction to the decomposition uM u∗ ⊗ P = N2 ⊗ · · · ⊗ Nn and get that uM u∗ (cid:22)uM u∗⊗P Ni for some i ≥ 2. Then take θ, p, q, v as in Theorem 2.2(i), and observe that θ ◦ Ad u, u∗pu, q, u∗v gives the condition M (cid:22)N Ni. Thus we get the conclusion. Lemma 4.2. Let M ⊗ B = N1 ⊗ · · · ⊗ Nn (=: N ) be a tensor decomposition as II1 factors with n ≥ 2. If M (cid:22)N Ni and M (cid:22)N Nj, then i = j. Proof. Suppose by contradiction that i 6= j, and put i = 1 and j = 2 for simplicity. Then by Lemmas 3.1 and 3.3, one has uM u∗ ⊗ P = N t 1, where u ∈ U (N ), t > 0, and P := (uM u∗)′ ∩ N t 1, that gives a decomposition N = uM u∗ ⊗ P ⊗ N 1/t 2 ⊗ N3 ⊗ · · · ⊗ Nn. 9 Observe by Lemma 3.1 that the given condition M (cid:22)N N2 is equivalent to uM u∗ (cid:22)N N 1/t By Lemma 2.3, we get uM u∗ (cid:22) C, which contradicts the diffuseness of M . 2 . N ∩(N 1/t 2 )′ Lemma 4.3. Let M1 ⊗ M2 ⊗ B = N1 ⊗ · · · ⊗ Nn (=: N ) be a tensor decomposition as II1 factors with n ≥ 2. If M1 (cid:22)N N1 and M2 (cid:22)N N1, then M1 ⊗ M2 (cid:22)N N1. In this case, N1 is not prime. Proof. The first assertion is immediate by Lemma 2.6. For the second one, by Lemmas 3.1 and 3.3, take u ∈ U (N ) and t > 0 such that u(M1 ⊗ M2)u∗ ⊂ N t 1 = u(M1 ⊗ M2)u∗ ⊗ P , where P := u(M1 ⊗ M2)′u∗ ∩ N t 1 ⊗ M2 ⊗ P , N1 is not prime. 1. Since N1 ≃ M 1/t 1 and N t Lemma 4.4. Let ⊗m m = ∞. If B (cid:22)M ⊗m i=1Mi = N ⊗ B (=: M ) be a tensor decomposition as II1 factors with i=kMi for all k ∈ N, then B is amenable. Proof. We follow the idea in [HU15, Proposition 4.2] due to Ioana. In the proof, for any subset F ⊂ N we put MF := ⊗i∈F Mi ⊂ M . Put M := M ⊗ M and we regard the left M as the original one. Let Σ be the flip map on M given by Σ(a⊗b) = b⊗a. For any F ⊂ N, put MF := MF ⊗MF with the flip ΣF . We regard ΣF ∈ Aut(M) by putting ΣF MF c := id. Observe that weak- limF ΣF (x) = Σ(x) for all x ∈ M, where the limit is taken over all finite subsets F ⊂ N. Observe next B (cid:22)M MF c for any finite F ⊂ N by assumption, so there is a unitary F c by Lemma 3.1. In this case, we may assume F c ⊂ Mmax F ⊗ MF c where max F := max{i i ∈ F} (recall that we are fixing F ⊗ M tF F c, so applying again a partial isometry conjugacy at the level of M ⊗ Mn F ⊗ 1) = F ⊗1)ΣF ◦(vF ⊗1) ∈ U (M) vF ∈ M and tF > 0 such that vF Bv∗ that M tF M = M 1/tF for some n ∈ N we may assume this condition). In particular we have ΣF ◦(vF bv∗ vF bv∗ and calculate that for all b ∈ B, F ⊗1 for all b ∈ B, where F ◦ := F \max F. We put uF := (v∗ F ⊂ M tF uF ΣF ◦(b ⊗ 1) = (v∗ = (v∗ = (v∗ = (v∗ = (b ⊗ 1)uF . F ⊗ 1)ΣF ◦(vF ⊗ 1)ΣF ◦ (b ⊗ 1) F ⊗ 1)ΣF ◦(vF b ⊗ 1) F ⊗ 1)ΣF ◦(vF bv∗ F ⊗ 1)(vF bv∗ F ⊗ 1)ΣF ◦(vF ⊗ 1) F ⊗ 1)ΣF ◦ (vF ⊗ 1) Define a state Ω on B(L2(M)) by Ω(X) := limF hXuF , uF iL2(M), where the limit is taken over all finite F. It satisfies for x ∈ M Ω(a) = lim F hauF , uF iL2(M) = lim F τM(u∗ F auF ) = τM(a). For all b ∈ U (B), regarding L2(M) = L2(M ) ⊗ L2(M ) with the right M -action given by M ∋ x 7→ 1⊗ JM x∗JM where JM is the anti-unitary map JM (y) = y∗ for y ∈ M ⊂ L2(M ), since uF ΣF ◦(b ⊗ 1) = (b ⊗ 1)uF and ΣF ◦(b ⊗ 1) → Σ(b ⊗ 1) weakly for all b ∈ B, we have Ω(b ⊗ JM bJM ) = lim F = lim F = lim F = lim F h(b ⊗ JM bJM )uF , uF iL2(M) h(b ⊗ 1)uF (1 ⊗ b∗), uF iL2(M) huF ΣF ◦(b ⊗ 1)(1 ⊗ b∗), uF iL2(M) τM(ΣF ◦(b ⊗ 1)(1 ⊗ b∗)) = τM(Σ(b ⊗ 1)(1 ⊗ b∗)) = 1. 10 So the state Ω satisfies Ω((b ⊗ JM bJM )(X ⊗ 1)) = Ω(X ⊗ 1) and hence Ω(bXb∗ ⊗ 1) = Ω((b ⊗ JM bJM )(X ⊗ 1)(b ⊗ JM bJM )∗) = Ω(X) for all X ∈ B(L2(M )) and b ∈ U (B). Thus the restriction of Ω on B(L2(M )) ⊗ C1L2(M ) is a B-central state which is the trace on B. This means B is amenable. Proof of Proposition D. We fix i ∈ N with 1 ≤ i ≤ m. Then since Mi is non-amenable, strong primeness and Lemma 4.4 imply that there is k ∈ N with 1 ≤ k ≤ n such that Mi (cid:22)M ⊗k j=1Nj (this is obvious if n 6= ∞). By Lemmas 3.1 and 3.3, one has uMiu∗ ⊗ P = N t 1 ⊗ N2 ⊗ · · · ⊗ Nk for a factor P , u ∈ U (M ), and t > 0. Then if P is of type I, then k = 1 by the primeness of Mi and hence Mi (cid:22)M N1. If P is a II1 factor, then by Lemma 4.1 there is some j ∈ N with 1 ≤ j ≤ k such that Mi (cid:22)M Nj. Thus in any case there is j such that Mi (cid:22)M Nj. We put σ(i) := j, and σ is uniquely determined by Lemma 4.2. Surjectivity of σ. Assume that σ is surjective. Then since the condition Mi (cid:22)M Nσ(i) implies non- amenability of Nσ(i), we have that all Nj are non-amenable. To see the converse direction, we show the following claim. Claim. Assume that there is j0 ∈ N with 1 ≤ j0 ≤ n such that Mi 6(cid:22)M Nj0 for all i ∈ N with 1 ≤ i ≤ n. Then we have (i) a contradiction if m 6= ∞, and (ii) Nj0 is amenable if m = ∞. Proof. We fix k ∈ N with 1 ≤ k ≤ m. Observe that Mi (cid:22)M ⊗k and then Lemma 4.3 implies ⊗k have j=1Nσ(j) for all 1 ≤ i ≤ k, j=1Nσ(j). By taking relative commutants, we i=1Mi (cid:22)M ⊗k Nj0 ⊂ (⊗k j=1Nσ(j))′ ∩ M (cid:22)M (⊗k i=1Mi)′ ∩ M = ⊗m i=k+1Mi. If m 6= ∞, one can put k = m and obtain Nj0 (cid:22)M C, a contradiction. If m = ∞, then we have Nj0 (cid:22)M ⊗m i=k+1Mi for all k ∈ N that implies amenability of Nj0 by Lemma 4.4. Observe now that j0 6∈ Imσ if and only if Mi 6(cid:22)M Nj0 for all i ∈ N with 1 ≤ i ≤ n (since Mi (cid:22)M Nj0 exactly means σ(i) = j0 by the uniqueness of σ). So this claim shows that (i) σ is always surjective if m 6= ∞, and (ii) if m = ∞ non-surjectivity of σ implies amenability of Nj0 for some j0. This completes the statement for surjectivity (note that Nj can not be amenable if m 6= ∞ as we mentioned in Introduction). Injectivity of σ. Assume next that σ is not injective. Then there are i 6= i′ such that σ(i) = σ(i′) =: j, that means Mi (cid:22)M Nj and Mi′ (cid:22)M Nj. By (the proof of) Lemma 4.3, Nj is isomorphic to M t i ⊗ Mi′ ⊗ P for some t > 0 and a factor P . Since Mi and Mi′ are non-amenable, Nj is not semiprime. j0 ⊗ N 2 j0 with non-amenable II1 factors N 1 Conversely assume Nσ(i) is not semiprime for some i, so there is a tensor decomposition j0 for all i, then j0 is amenable if j0 implies there is k, l ∈ N such that Mk (cid:22)M N 1 j0 j0 ⊂ Nj0, we have Nj0 = N 1 j0. If Mi 6(cid:22)M N 1 the claim above shows that we have (i) a contradiction if m 6= ∞, and (ii) N 1 m = ∞. So non-amenability of N 1 and Ml (cid:22)M N 2 Mk (cid:22)M Nj0 and Ml (cid:22)M Nj0 that means σ(k) = σ(l). So σ is not injective. j0. We know k 6= l by Lemma 4.2. Finally since N 1 j0 and N 2 j0 and N 2 j0, N 2 Finally we assume that each Nj is non-amenable and semiprime. Then σ is bijective i ⊗ Pi by previous arguments. By Lemmas 3.1 and 3.3, Mi (cid:22)M Nσ(i) implies Nσ(i) ≃ M ti for some ti > 0 and a factor Pi. Since Nσ(i) is semiprime, Pi must be amenable. 11 5 Proofs of main theorems In the proofs of main theorems, we will make use of the following three structural the- orems. Note that all of them are formulated with relative amenability, and this relativity is crucial to our proofs. Theorem 5.1 ([PV12, Theorem 1.4]). Let B be any finite von Neumann algebra and Γ be weakly amenable and bi-exact group acting on B. Put M := B ⋊ Γ. Then for any von Neumann subalgebra A ⊂ M which is amenable relative to B in M , we have either (i) A (cid:22)M B or (ii) NM (A)′′ is amenable relative to B in M . Theorem 5.2 ([Io12, Theorem 1.6][Va13, Theorem A]). Let M = M1 ∗B M2 be an amal- gamated free product of tracial von Neumann algebras (Mi, τ ) with common von Neu- mann subalgebra B ⊂ Mi w.r.t. the unique trace preserving conditional expectations. Let A ⊂ M be a von Neumann subalgebra that is amenable relative to B inside M and satisfies A 6(cid:22)M B. Then we have either (i) NM (A)′′ (cid:22)M Mi for some i or (ii) NM (A)′′ is amenable relative to B inside M . Theorem 5.3 ([SW11, Theorem 2.2]). Let Γ be a wreath product group of a non-trivial amenable group by a non-amenable group and let B be a finite von Neumann algebra. Put M := B ⊗ LΓ. Let Q ⊂ M be a von Neumann subalgebra which is not amenable relative to B. If Q′ ∩ M is a regular subfactor in M , then we have Q′ ∩ M (cid:22)M B. Proof of Theorem C. The first case was already proved in [Is14, Theorem 5.1.1]. Suppose by contradiction that M is not strongly prime. Then by Lemma 3.5, there are II1 factors B, K, and L such that B ⊗ M = K ⊗ L (=: N ) with K 6(cid:22)N B and L 6(cid:22)N B. Case 1. M is a free product M1 ∗ M2. By [BO08, Corollary F.14], there is a diffuse abelian subalgebra A ⊂ K such that A 6(cid:22)N B. Then regarding N = (B ⊗ M1) ∗B (B ⊗ M2), we apply Theorem 5.2 to A ⊂ N and get either (i) NN (A)′′ (cid:22)N (B ⊗ Mi) for some i or (ii) NN (A)′′ is amenable relative to B in N . Assume first that (ii) happens. Since L ⊂ NN (A)′′, L is amenable relative to B in N . By Theorem 5.2, we get either (i)′ N (cid:22)N (B ⊗Mi) for some i or (ii)′ N is amenable relative to B inside N . If (i)′, by Lemma 2.3, one has M (cid:22)M Mi which contradicts diffuseness of Mj (where i 6= j) by Lemma 2.5. If (ii)′, then we get that M is amenable by Lemma 2.9, which is a contradiction. Thus the condition (ii) does not happen. Assume next condition (i). We have two conditions L (cid:22)N (B ⊗ Mi) and L 6(cid:22)N B, and it is known that they imply N = K ⊗ L (cid:22)N (B ⊗ Mi) [IPP05]. Here we give a sketch of this argument in the paragraphs below for reader's convenience. Once we obtain it, then by Lemma 2.3, this means M (cid:22)M Mi which contradicts diffuseness of Mj (where i 6= j) by Lemma 2.5, and hence we can end the proof. Suppose now that L (cid:22)N (B ⊗ M1). Then there is a ∗-homomorphism θ : pLp → q(B ⊗ M1)q for some projections p ∈ L, q ∈ B ⊗ M1, and a partial isometry v ∈ N such that vθ(x) = xv for x ∈ pLp. We may replace q with the support projection of EB⊗M1(v∗v). Put D := θ(pLp). If D (cid:22)B⊗M1 B, then by the choice of q, we can deduce L (cid:22)B⊗M B (e.g. [Va08, Remark 3.8]) and hence a contradiction. So we have D 6(cid:22)B⊗M1 B. By [IPP05, Theorem 1.1], any quasi-normalizer of D in q(B ⊗ M )q is contained in B ⊗ M1. In particular we have v∗v ∈ B ⊗ M1. We put q := v∗v, θ := θ(·)q, and eD := Dq, and observe that eD 6(cid:22)B⊗M1 B. Write vv∗ = pp′ for some p′ ∈ L′ ∩ N = K. Then we get a ∗-homomorphism Adv∗ : pLpp′ → q(B ⊗ M1)q. Since v∗pLpp′v = eD 6(cid:22)B⊗M1 B, again by [IPP05, Theorem 1.1], any quasi-normalizer of v∗pLpp′v is contained in B ⊗ M1. Hence 12 we have v∗pp′Kpp′v ⊂ q(B ⊗ M1)q. Thus we obtain v∗pp′(K ⊗ L)pp′v ⊂ q(B ⊗ M1)q and K ⊗ L (cid:22)N B ⊗ M1. This is the desired condition. Case 2. M is a wreath product group factor L(∆ ≀ Λ). Write Λ = Λ1 × Λ2 as in the statement. For simplicity, we also write as Γ := ∆ ≀ Λ = ∆Λ ⋊ Λ and Γi := ∆Λ ⋊ Λi for all i. Since K and L are regular subfactors, and K = L′ ∩ N and L = K ′ ∩ N , by Theorem 5.3, it holds that K and L are amenable relative to B in M . In particular, K and L are amenable relative to B ⊗LΓ1 (since B ⊂ B ⊗LΓ1). Regarding B ⊗LΓ as a crossed product of B ⊗ LΓ1 by Λ2, by Theorem 5.1, we get K (cid:22)N B ⊗ LΓ1 and L (cid:22)N B ⊗ LΓ1. We can then apply Lemma 2.6 and obtain that N = K ∨ L (cid:22)N B ⊗ Γ1. However by Lemma 2.4, this contradicts the fact that Λ2 is an infinite group. Proof of Theorem A. We consider only the case that M is the wreath product group factor, and other cases are proved by Theorem C and Proposition 3.6. Put Γ := ∆ ≀ Λ and M := LΓ. We will verify the sufficient condition in Lemma 3.4. Let B be a II1 factor and t > 0 such that M ⊗ B ≃ M ⊗ Bt (=: K ⊗ L). Regarding M ⊗ B = K ⊗ L, we will show that either K (cid:22)M ⊗B B, L (cid:22)M ⊗B B, M (cid:22)M ⊗B L, or B (cid:22)M ⊗B L. So suppose by contradiction that any of them does not hold and we will deduce amenability of M , which is a contradiction. We apply Theorem 5.3 to K and get either (i) K is amenable relative to B in M ⊗ B or (ii) K ′ ∩(M ⊗B) = L (cid:22)M ⊗B B. So by assumption, we have that K is amenable relative to B in M ⊗ B. By the same reason, L is also amenable relative to B in M ⊗ B. Exchanging the roles of M ⊗ B and M ⊗ Bt, it further holds that M and B are amenable relative to L in M ⊗ B. Hence using M ⋖ M ⊗B B together with Proposition 2.8, we obtain that M is amenable relative to B in M ⊗ B. This means that M is amenable by Lemma 2.9 and thus a contradiction. M ⊗B L and L ⋖ 6 Proof of Corollary B Let Gn ∈ Sfactor for n ∈ N (possibly Gn = Gm for different n, m), and take II1 factors Bn with separable predual such that F(Bn) = Gn. We may assume Bn = Bm whenever Gn = Gm. Let N be a free product II1 factor given by N := LF2 ∗ L(Z2 ⋊ SL(2, Z)). Observe that F(N ) = {1} by [IPP05, Corollary 6.4] and hence F(N ⊗ Bn) = F(Bn) by Theorem A. Define an infinite free product II1 factor M := ∗∞ n=1Mn, where Mn := N ⊗ Bn for all n ∈ N. We first show that it satisfies F(M ) =Tn∈N F(Bn) =Tn∈N Gn. Recall first from [DR99, Theorem 1.5] that for any 0 < t ≤ 1, one has M t = ∗∞ n=1M t n and this implies Tn∈N F(Mn) ⊂ F(M ) [DR99, Corollary 1.6]. Since F(Bn) = F(Mn) for all n ∈ N, we get an inclusion Tn∈N F(Bn) ⊂ F(M ). We next see the reverse inclusion. Fix t ∈ F(M ). Up to replacing with 1/t if necessary, we may assume 0 < t ≤ 1 and so we have an isomorphism ∗∞ n=1Mn = M ≃ M t = ∗∞ n=1M t n. Since each Mn is a tensor product of non-amenable II1 factors, we can apply [HU15, Main Theorem] (see also [Oz04, IPP05, Po06b] for the case of finitely many free components). So there is a bijection α on N such that Mn and M t α(n) are isomorphic for all n ∈ N. Indeed, [HU15, Main Theorem] actually shows Mn (cid:22)M M t α(n) (cid:22)M Mn. Once α(n) and M t 13 we get this condition, then by the proof of unique factorization of free products II1 factors (e.g. [Oz04, Theorem 3.3]) one can show that Mn and M t α(n) are unitary conjugate in M , namely, there is u ∈ U (M ) such that uMnu∗ = M t α(n) (under the given isomorphism). This particularly implies Gn = F(Bn) = F(Mn) = F(M t α(n)) = F(Mα(n)) = F(Bα(n)) = Gα(n) and hence Bn = Bα(n) by our choice of {Bk}k∈N. Thus the above isomorphism Mn ≃ M t α(n) means t ∈ F(Mn) = F(Bn) for each n ∈ N, and so t ∈ Tn∈N F(Bn). We conclude F(M ) ⊂Tn∈N F(Bn). Now we start the proof of Corollary B. The stability for intersection was already proved above. Let G ∈ Sfactor and take a II1 factor B with separable predual such that F(B) = G. Then by putting Bn := B for all n ∈ N, the above argument shows that F(B) = F(M ) for M := ∗n∈N(B ⊗N ), which is exactly the formula we mentioned in Introduction. (Note that even in the case Bn = B for all n ∈ N, one needs infinitely many free product components, since [DR99, Theorem 1.5] holds only for infinite free products.) Since M is a free product, it satisfies the property (TFF), so the first assertion of Corollary B holds. The stability for multiplication is then an immediate consequence of the first assertion and the definition of the property (TFF). 7 Some partial results It would be interesting to know whether L(Z2⋊SL(2, Z)) satisfies the property (TFF) or not. However we can not apply Theorem A because of the lacking of the weak amenability. In this section, we study some partial answers to this problem. Observe that L(Z2 ⋊SL(2, Z)) has two structures: one is the crossed product L∞(T2)⋊ SL(2, Z) coming from a strongly ergodic action of a bi-exact weakly amenable group; and the other is a bi-exact group factor [Oz08]. From these viewpoints, we give partial answers to the property (TFF) as follows. See [BO08, Definition 12.3.9] for the definition of the W∗CMAP (or equivalently, the W∗CBAP with Cowling -- Haagerup constant 1). Proposition 7.1. The following statements hold true. (1) Let Γ be a non-amenable, weakly amenable, and bi-exact group acting on a standard probability space X as a free, strongly ergodic, and p.m.p. action. Put M := L∞(X)⋊ Γ. Then for any full II1 factor B, one has F(B ⊗ M ) = F(B)F(M ). (2) Let Γ be a non-amenable bi-exact ICC group. Then for any II1 factor B with the W∗CMAP, one has F(B ⊗ LΓ) = F(B)F(LΓ). The first assertion of this proposition will be proved by combining the proof of [Is14, Theorem 5.1.1] with the following lemma. Lemma 7.2 ([Ho15, Proposition 6.3]). Let N = M ⊗B = K ⊗L be a tensor decomposition as II1 factors, and let A ⊂ M be a Cartan subalgebra. If K (cid:22)N A ⊗ B and K is full, then we have K (cid:22)N B. Proof of Proposition 7.1(1). We show that for any tensor decomposition M ⊗ B = K ⊗ L with B full, one has K (cid:22)M ⊗B B or L (cid:22)M ⊗B B. This gives the conclusion by Lemma 3.4. Observe that M is full since the action is strongly ergodic. So by [Co75, Corollary 2.3], the tensor product M ⊗ B is full, and hence so are K and L. By Theorem 5.1 and the proof of [Is14, Theorem 5.1.1], one has K (cid:22)M ⊗B B ⊗ L∞(X) or L (cid:22)M ⊗B B ⊗ L∞(X). Then we can apply Lemma 7.2, and obtain K (cid:22)M ⊗B B or L (cid:22)M ⊗B B. 14 For the second assertion of Proposition 7.1, we prove the following proposition. This should be regarded as a "relativization" of Ozawa's semisolidity theorem [Oz04, Theorem 4.6]. Actually we can not give a complete generalization of Ozawa's theorem, since local reflexivity (or exactness) of C ∗ λ(Γ) is not enough as an extension property in this setting. We will use the W∗CMAP on B to avoid this problem. Proposition 7.3. Let Γ be a bi-exact group and B a finite von Neumann algebra. Put M := LΓ ⊗ B. Then for any von Neumann subalgebra A ⊂ M with A 6(cid:22)M B, there is a u.c.p. map from hM, Bi into A′ ∩ M , which restricts to the conditional expectation EA′∩M on LΓ ⊗min B. If B has the W∗CMAP, then the resulting u.c.p. map can be taken as the one restricting EA′∩M on LΓ ⊗ B, and thus A′ ∩ M is amenable relative to B. Proof. Since most parts of the proof are straightforward "relativization" of the one of [Oz04, Theorem 4.6], we give only a sketch. Our proof here is very similar to the one of [BO08, Theorem 15.1.5] (and its generalization [Is12, Theorem 5.3.3]). We will use the Hilbert space H := L2(M )⊗B L2(M ) = ℓ2(Γ)⊗L2(B)⊗ℓ2(Γ), in stead of L2(M )⊗L2(M ). The first part is exactly the same as the one of [BO08, Theorem 15.1.5] (and [Is12, Theorem 5.3.3]). Assume A 6(cid:22)M B. By [BO08, Corollary F.14], we may assume A is abelian. Then one can define a proper conditional expectation ΨA : B(L2(M )) −→ A′ ∩ B(L2(M )). The condition A 6(cid:22)M B implies ΨA(K(ℓ2(Γ)) ⊗min B(L2(B))) = 0. From now on, we use the relative tensor product. In particular, we will not use [Oz04, Proposition 4.2] but use a characterization of bi-exactness [BO08, Lemma 15.1.4]. Let πH and θH be left and right actions of M on H, and denote by ν the algebraic ∗- homomorphism from πH(M )θH (M op) to B(L2(M )) given by ν(πH(a)θH (bop)) = abop. Let λ(Γ)op → B(ℓ2(Γ)) be a u.c.p. map such that Θ(a ⊗ bop) − abop ∈ K(ℓ2(Γ)) Θ : C ∗ [BO08, Lemma 15.1.4]. Put M0 := C ∗ Identifying C∗{πH (M0), θH(M0)} as C ∗ λ(Γ)op, we may define Θ on this algebra, which is the identity on B and Bop. Observe that at the C∗-algebra level, Θ and ν coincide modulo K(ℓ2(Γ)) ⊗min B(L2(B)), that is, λ(Γ) ⊗min C∗{B, Bop} ⊗min C ∗ λ(Γ) ⊗min C ∗ λ(Γ) ⊗min B. Θ(πH(a)θH (bop)) − abop ∈ K(ℓ2(Γ)) ⊗min B(L2(B)), a, b ∈ M0. Thus on C∗{πH(M0), θH (M0)} the composition ΦA ◦ ν coincides with ΦA ◦ Θ, and hence is a bounded u.c.p. map. Observe that ΦAM is the unique trace preserving conditional expectation EA′∩M : M → A′ ∩M , and hence in particular normal on M . So the map ΦA ◦ν is a normal u.c.p. map on πH(M ). Regarding again C∗{πH (M0), θH (M0)} = C ∗ λ(Γ)op, we can apply the local reflexivity of C ∗ λ(Γ) (this comes from exactness of Γ) and extend λ(Γ)op which is normal on LΓ (see Lemma 9.4.1, ΦA ◦ ν on LΓ ⊗min C∗{B, Bop} ⊗min C ∗ Proposition 9.2.5, and the proof of Lemma 9.2.9 in [BO08] for these facts). Finally by Arveson's extension theorem, we again extend ΦA ◦ ν on C∗{πH(hM, Bi), θH (M0)}. Then the restriction on πH(hM, Bi) of the resulting map defines a u.c.p. map from hM, Bi into A′ ∩ (M op 0 )′ = A′ ∩ M . By construction, this is a desired item. λ(Γ) ⊗min C∗{B, Bop} ⊗min C ∗ Finally assume that B has the W∗CMAP, and take a net (ψi)i of normal finite rank c.c. maps on B converging to idB point weakly. We extend these maps to hM, Bi = B(ℓ2(Γ)) ⊗ B by id ⊗ ψi =: eψi. Observe that eψi(M ) ⊂ LΓ ⊗min B for all i. Let Φ be the u.c.p. map constructed in the first half of the proof. If we take a cluster point eΦ of 15 (Φ ◦ eψi)i, then this is a c.c. map from hM, Bi into A′ ∩ M which restricts to EA′∩M on M . In fact, for any x ∈ M ⊗ B, one has Φ ◦ eψi(x) = EA′∩M ◦ eψi(x) → EA′∩M (x), as i → ∞. Hence eΦM = EA′∩M and eΦ is a conditional expectation onto A′ ∩ M . Proof of Proposition 7.1(2). Take t ∈ F(LΓ ⊗ B) and fix M := LΓ ⊗ B = LΓ ⊗ Bt(=: K ⊗ L). By (the proof of) Lemma 3.4, we have only to show that K (cid:22)M B, L (cid:22)M B, LΓ (cid:22)M L, or B (cid:22)M L. So suppose by contradiction that each of them does not happen. We apply Proposition 7.3 to K (actually an abelian subalgebra of K by [BO08, Corol- lary F.14]), and get that L ⋖M B. By exchanging the roles, we also have that LΓ ⋖M L and hence LΓ ⋖M B by Proposition 2.8. Thus by Lemma 2.9, we obtain amenability of LΓ which is a contradiction. Reference [BO08] N. P. Brown and N. Ozawa, C ∗-algebras and finite-dimensional approximations. Graduate Studies in Mathematics, 88. American Mathematical Society, Providence, RI, 2008. [CKP14] I. Chifan, Y. Kida and S. Pant, Primeness results for von Neumann algebras associated with surface braid groups. Int. Math. Res. Not. (2015), article id:rnv271, 42pp. [CSU11] I. Chifan, T. Sinclair and B. Udrea, On the structural theory of II1 factors of negatively curved groups, II. Actions by product groups. Adv. Math. 245 (2013), 208 -- 236. [Co75] A. Connes, Classification of injective factors. Cases II1, II∞, IIIλ, λ 6= 1. Ann. of Math. (2) 104 (1976), 73 -- 115. [Co80] A. Connes, A factor of type II1 with countable fundamental group. J. Operator Theory 4 (1980), 151 -- 153. [De10] S. Deprez, Explicit examples of equivalence relations and factors with prescribed funda- mental group and outer automorphism group. Preprint, arXiv:1010.3612. [DR99] K. J. Dykema and F. Radulescu, Compressions of free products of von Neumann algebras. Math. Ann., 316 (2000), 61 -- 82. [Ga99] D. Gaboriau, Cout des relations d'´equivalence et des groupes. Invent. Math. 139 (2000), no. 1, 41 -- 98. [Ga01] D. Gaboriau, Invariants ℓ2 de relations d'´equivalence et de groupes. Publ. Math. Inst. Hautes ´Etudes Sci. No. 95 (2002), 93 -- 150. [Ho15] D.J. Hoff, Von Neumann algebras of equivalence relations with nontrivial one-cohomology. J. Funct. Anal. 270 (2016), no. 4, 1501 -- 1536. [Ho07] [HI15] C. Houdayer, Construction of type II1 factors with prescribed countable fundamental group. J. Reine Angew Math. 634 (2009), 169 -- 207. C. Houdayer and Y .Isono, Unique prime factorization and bicentralizer problem for a class of type III factors. Adv. Math. 305 (2017), 402 -- 455. [HU15] C. Houdayer and Y. Ueda, Rigidity of free product von Neumann algebras. Compos. Math. 152 (2016), 2461 -- 2492. [Io12] A. Ioana, Cartan subalgebras of amalgamated free product II1 factors (with an appendix joint with S. Vaes). Ann. Sci. ´Ecole Norm. Sup. 48 (2015), 71 -- 130. [IPP05] A. Ioana, J. Peterson and S. Popa, Amalgamated free products of weakly rigid factors and calculation of their symmetry groups. Acta Math. 200 (2008), 85 -- 153. [Is12] Y. Isono, Weak Exactness for C∗-algebras and Application to Condition (AO), J. Funct. Anal. 264 (2013), 964 -- 998. 16 [Is14] Y. Isono, Some prime factorization results for free quantum group factors. J. Reine Angew. Math. 722 (2017), 215 -- 250. [MV43] F.J. Murray and J. Von Neumann, On rings of operators IV. Ann. Math. 44 (1943), 716 -- 808. [Oz04] N. Ozawa, A Kurosh type theorem for type II1 factors. Int. Math. Res. Not. (2006), Art. ID 97560, 21 pp. [Oz08] N. Ozawa, An example of a solid von Neumann algebra. Hokkaido Math. J., 38 (2009), 557 -- 561. [OP03] N. Ozawa and S. Popa, Some prime factorization results for type II1 factors. Invent. Math. 156 (2004), 223 -- 234. [OP07] N. Ozawa and S. Popa, On a class of II1 factors with at most one Cartan subalgebra. Ann. of Math. (2), 172 (2010), 713 -- 749. [Pe06] J. Peterson, L2-rigidity in von Neumann algebras, Invent. Math. 175 (2009), 417 -- 433. [Po01] [Po03] [Po04] S. Popa, On a class of type II1 factors with Betti numbers invariants. Ann. of Math. 163 (2006), 809 -- 899. S. Popa, Strong rigidity of II1 factors arising from malleable actions of w-rigid groups I. Invent. Math. 165 (2006), 369 -- 408. S. Popa, Strong rigidity of II1 factors arising from malleable actions of w-rigid groups, II. Invent. Math. 165 (2006), 409 -- 452. [Po06a] S. Popa, On the superrigidity of malleable actions with spectral gap. J. Amer. Math. Soc. 21 (2008), 981 -- 1000. [Po06b] S. Popa, On Ozawa's property for free group factors. Int. Math. Res. Notices. Vol. 2007 : article ID rnm036, 10 pages. [PV08a] S. Popa and S. Vaes, Actions of F∞ whose II1 factors and orbit equivalence relations have prescribed fundamental group. J. Amer. Math. Soc. 23 (2010), 383 -- 403. [PV08b] S. Popa and S. Vaes, On the fundamental group of II1 factors and equivalence relations arising from group actions. In Quanta of Maths, Proceedings of the Conference in honor of A. Connes' 60th birthday, Clay Mathematics Institute Proceedings, 11 (2011), pp. 519 -- 541. [PV12] S. Popa and S. Vaes, Unique Cartan decomposition for II1 factors arising from arbitrary actions of hyperbolic groups. J. Reine Angew. Math. 694 (2014), 215 -- 239. [Ra91] [Sa09] F. Radulescu, The fundamental group of the von Neumann algebra of a free group with infinitely many generators is R∗ +. J. Amer. Math. Soc. 5 (1992), 517 -- 532. H. Sako, Measure equivalence rigidity and bi-exactness of groups. J. Funct. Anal. 257 (2009) 3167 -- 3202. [SW11] J. O. Sizemore and A. Winchester, A unique prime decomposition result for wreath prod- uct factors. Pacific J. Math. 265 (2013), no. 1, 221 -- 232. [Va08] [Va13] [Vo89] S. Vaes, Explicit computations of all finite index bimodules for a family of II1 factors. Ann. Sci. ´Ecole Norm. Sup. 41 (2008), 743 -- 788. S. Vaes, Normalizers inside amalgamated free product von Neumann algebras. Publ. Res. Inst. Math. Sci. 50 (2014), 695 -- 721. D.V. Voiculescu, Circular and semicircular systems and free product factors. In Operator algebras, unitary representations, enveloping algebras, and invariant theory (Paris, 1989), Progr. Math. 92, Birkhauser, Boston, 1990, p. 45 -- 60. 17
1712.09192
2
1712
2018-10-22T07:04:27
Tingley's problem through the facial structure of operator algebras
[ "math.OA", "math.FA" ]
Tingley's problem asks whether every surjective isometry between the unit spheres of two Banach spaces admits an extension to a real linear surjective isometry between the whole spaces. In this paper, we give an affirmative answer to Tingley's problem when both spaces are preduals of von Neumann algebras, the spaces of self-adjoint operators in von Neumann algebras or the spaces of self-adjoint normal functionals on von Neumann algebras. We also show that every surjective isometry between the unit spheres of unital C$^*$-algebras restricts to a bijection between their unitary groups. In addition, we show that every surjective isometry between the normal state spaces or the normal quasi-state spaces of two von Neumann algebras extends to a linear surjective isometry.
math.OA
math
TINGLEY'S PROBLEM THROUGH THE FACIAL STRUCTURE OF OPERATOR ALGEBRAS MICHIYA MORI Abstract. Tingley's problem asks whether every surjective isometry between the unit spheres of two Banach spaces admits an extension to a real linear sur- jective isometry between the whole spaces. In this paper, we give an affirmative answer to Tingley's problem when both spaces are preduals of von Neumann algebras, the spaces of self-adjoint operators in von Neumann algebras or the spaces of self-adjoint normal functionals on von Neumann algebras. We also show that every surjective isometry between the unit spheres of unital C∗- algebras restricts to a bijection between their unitary groups. In addition, we show that every surjective isometry between the normal state spaces or the normal quasi-state spaces of two von Neumann algebras extends to a linear surjective isometry. 1. Introduction In 1951, the study of isometries between operator algebras began in [10]. In that paper, Kadison proved that every complex linear surjective isometry between two unital C∗-algebras can be decomposed as the composition of a Jordan ∗- isomorphism and the multiplication by a unitary. Since then, linear isometries between operator algebras have been considered in various settings by many re- searchers. For example, see [7] and [20], which contain results and references con- cerning generalizations of Kadison's theorem to thoroughly different directions. On the other hand, the Mazur-Ulam theorem states that every surjective isom- etry between two real normed spaces is affine. This result attracted many mathe- maticians, and isometries without assuming affinity were considered in many cases. The symbol S(X) denotes the unit sphere (i.e. the subset of the elements with norm 1) of a Banach space X, while the notation BX means the closed unit ball of X. What we focus on in this paper is the following problem, which is closely related to the Mazur-Ulam theorem. Problem 1.1. Let X and Y be two Banach spaces and T : S(X) → S(Y ) be a surjective isometry. Does T admit an extension to a real linear surjective isometry eT : X → Y ? The first contribution to this problem dates back to 1987, and it is due to Tingley [24], so this problem is nowadays called Tingley's problem (or the surjective iso- metric extension problem). More than 30 years have passed since the birth of this problem, but the answer in general situations is yet far from having been achieved. Indeed, it is said that Tingley's problem is unsolved even in the case X = Y and X is two dimensional. However, until now, no counterexamples for Tingley's problem 2010 Mathematics Subject Classification. Primary 46B04, Secondary 46B20, 47B49. Key words and phrases. Tingley's problem; isometric extension; operator algebra. 2 M. MORI have been found. Moreover, in many cases (including the cases of most of classi- cal real Banach spaces and some special Banach spaces), affirmative answers have been given for Tingley's problem. The survey [3] contains good expositions and references on Tingley's problem. Tingley's problem in the setting of operator algebras was first considered by Tanaka [22], and he later solved Tingley's problem affirmatively when X and Y are finite von Neumann algebras [23]. Recently, Fern´andez-Polo and Peralta generalized this result to the cases of general von Neumann algebras [8]. On the other side, Fern´andez-Polo, Garc´es, Peralta and Villanueva solved Tingley's problem positively when X and Y are the spaces of trace class operators on complex Hilbert spaces [6]. See Introduction of [8] for the latest developments in this field. It is common to use the following strategy to solve Tingley's problem for operator algebras. First we detect some substructures of the unit spheres such as unitary groups and minimal or maximal partial isometries. In this step, the facial structure of unit balls plays a crucial role. Second we construct the only one candidate for the real linear extension which is determined by such substructures. And finally we show that this linear mapping is the extension we wanted. In this paper, applying some versions of this strategy, we give several new results concerning Tingley's problem in the setting of operator algebras. In Section 2, we summarize some known results about the facial structure of operator algebras and (pre)duals (due to Akemann and Pedersen [1]) and its appli- cation to Tingley's problem, which are used in the later sections. In Section 3, we show that every surjective isometry between the unit spheres of two unital C∗-algebras restricts to a bijection between their unitary groups. In the proof, we detect the unitary group from extreme points in the unit ball. Using the surjective isometry between unitary groups and the result due to Hatori and Moln´ar [9], we construct the only one candidate for the real linear isometric extension. Although the author does not know whether this linear mapping actually extends the original mapping, we show that Tingley's problem for unital C∗-algebras is equivalent to Problem 6.1. In Section 4, we give a positive answer to Tingley's problem when X and Y are preduals, M∗ and N∗ of von Neumann algebras M and N , respectively. In the proof, we use the structure of maximal faces, and calculate Hausdorff distances between them to construct a surjective isometry between the unitary groups of M and N . By the theorem of Hatori and Moln´ar, this mapping extends to a real linear surjective isometry from M onto N . This linear mapping canonically determines a real linear surjective isometry from N∗ onto M∗, whose inverse mapping is shown to be the extension we wanted. In Section 5, we show that Tingley's problem has an affirmative answer when X and Y are the spaces Msa and Nsa of self-adjoint operators in von Neumann algebras M and N , respectively. In this case, some techniques used in sections before cannot be applied. Instead, we use the structure of projection lattices and orthogonality combined with a theorem of Dye [4]. We also solve Tingley's problem positively when X and Y are the spaces M∗sa and N∗sa of self-adjoint elements in preduals of von Neumann algebras M and N , respectively. Additionally, applying some discussions in this paper, we show that every surjective isometry T : S(M∗)∩M∗+ → S(N∗)∩ N∗+ (resp. T : BM∗ ∩ M∗+ → BN∗ ∩ N∗+) between the normal state spaces TINGLEY'S PROBLEM FOR OPERATOR ALGEBRAS 3 (resp. between the normal quasi-state spaces) of two von Neumann algebras M and N admits a linear surjective isometric extension from Msa onto Nsa. In Section 6, along the line of this paper, we list problems which seem to be open and new, with some comments. 2. Facial structure of operator algebras and its use in Tingley's problem Recall that a nonempty convex subset F of a convex set C in a Banach space is called a face in C if F has the following property: if x, y ∈ C and λx + (1− λ)y ∈ F for some 0 < λ < 1, then x, y ∈ F . It can be easily proved by Hahn-Banach theorem that for a Banach space X, a subset F of BX is a maximal norm-closed proper face in BX if and only if F is a maximal convex subset of S(X) (see [23, Lemma 3.2]). In order to attack Tingley's problem, nowadays the following geometric property is known: every surjective isometry between the unit spheres of two Banach spaces preserves maximal convex sets of the spheres ([2, Lemma 5.1(ii)], [21, Lemma 3.5]). On the other hand, the facial structure of the unit ball of operator algebras and (pre)duals were thoroughly studied by Akemann and Pedersen [1]. Let X be a real or complex Banach space and F ⊂ X, G ⊂ X ∗ be subsets. We define F ′ := {f ∈ BX ∗ f (x) = 1 for any x ∈ F}, G′ := {x ∈ BX f (x) = 1 for any f ∈ G}. Theorem 2.1 (Akemann and Pedersen [1, Theorem 5.3]). Let X be one of the following Banach spaces: a C∗-algebra, the space of self-adjoint operators in a C∗- algebra, the predual of a von Neumann algebra, or the space of self-adjoint elements in the predual of a von Neumann algebra. (Consider X as a complex Banach space in the first or the third case, and real in the other cases.) Then the mapping F 7→ F ′ is an order-reversing bijection from the class of norm-closed faces in BX onto the class of weak∗-closed faces in BX ∗. The inverse mapping is given by G 7→ G′. Using this theorem as in the proof of Corollary 3.4 in [23] (or by Corollary 2.5 of [6], which can also be applied in the situations of real Banach spaces), we obtain the following proposition. For the convenience of the readers, we add a proof. Proposition 2.2 (A version of [23, Corollary 3.4] or [6, Corollary 2.5]). Let A and B be C∗-algebras, M and N be von Neumann algebras and the pair (X, Y ) be one of the following pairs: (A, B), (Asa, Bsa), (M∗, N∗) or (M∗sa, N∗sa). Suppose T : S(X) → S(Y ) is a surjective isometry. Then for a subset F ⊂ S(X), F is a norm-closed proper face in BX if and only if T (F ) is in BY . In particular, x ∈ S(X) is an extreme point in BX if and only if T (x) is in BY . Proof. Let F be a norm-closed proper face in BX . By the preceding theorem and the Krein-Milman theorem, we have F = (F ′)′ = \f ∈F ′{f}′ = \f ∈ext(F ′) {f}′. Since F ′ is a face, it follows that ext(F ′) ⊂ ext(BX ∗ ). Again by the preceding theorem, for every f ∈ ext(BX ∗ ), {f}′ is a maximal norm-closed proper face in 4 M. MORI S(X). By the fact that T gives a bijection between the classes of maximal norm- closed proper faces in unit balls, it follows that the set T (F ) = T(cid:18) \f ∈ext(F ′) {f}′(cid:19) = \f ∈ext(F ′) T ({f}′) (cid:3) is a norm-closed proper face in BY . We add a little more to these results. Proposition 2.3 (See also [24, Section 4]). Let X and Y be Banach spaces and suppose that T : S(X) → S(Y ) is a surjective isometry. (a) Let F ⊂ S(X) be a maximal convex set. Then T (−F ) = −T (F ). (b) Suppose (X, Y ) is a pair as in the preceding proposition and let F ⊂ BX be a norm-closed proper face. Then T (−F ) = −T (F ). Proof. (a) It suffices to show that −F = {x ∈ S(X) kx − yk = 2 for any y ∈ F}. Let y1, y2 ∈ F . Then (y1 + y2)/2 ∈ F . In particular, k − y1 − y2k = ky1 + y2k = 2. Thus we obtain −F ⊂ {x ∈ S(X) kx − yk = 2 for any y ∈ F}. Let x ∈ S(X) and suppose kx − yk = 2 for all y ∈ F . Then the open convex sets S1 := {z1 ∈ X dist(z1,F ) < 1} = F + intBX and S2 := {z2 ∈ X kz2 − xk < 1} = x + intBX do not have a common element. By the Hahn-Banach theorem, we obtain a functional f ∈ S(X ∗) and a real number c ∈ R such that Re f (z1) > c for every z1 ∈ S1 and Re f (z2) < c for every z2 ∈ S2. Since Re f (x), Re f (y) ∈ [−1, 1] for every y ∈ F and Re f (intBX ) = (−1, 1), we have c = 0, Re f (x) = −1, Re f (y) = 1 and thus f (x) = −1 and f (y) = 1. It follows that f −1(1) ∩ BX ⊂ BX is a norm-closed face which contains F . By the maximality of F , we have f −1(1) ∩ BX = F . Thus x ∈ f −1(−1) ∩ BX = (−f −1(1)) ∩ BX = −F . (b) follows by (a) and the fact that every norm-closed face is the intersection of some maximal convex sets in S(X) (see the proof of the preceding proposition). (cid:3) In fact, Akemann and Pedersen concretely described the facial structure of op- erator algebras and (pre)duals in order to prove Theorem 2.1. Let A be a (not necessarily unital) C∗-algebra. The partial order in the set of partial isometries in A is given by the following: u majorizes (or extends) v if u = v + (1 − vv∗)u(1 − v∗v). A projection p in the bidual A∗∗ (considered as the enveloping von Neumann algebra) is said to be open if there exists an increasing net of positive elements in A converging to p in the σ-strong topology of A∗∗. A projection p ∈ A∗∗ is said to be closed if 1 − p is open. A closed projection p in A∗∗ is compact if p ≤ a for some norm-one positive element a ∈ A. A partial isometry v ∈ A∗∗ belongs locally to A if v∗v is a compact projection and there exists a norm-one element x in A such that xv∗ = vv∗. See [1] for more information. Theorem 2.4 (Akemann and Pedersen [1]). Let A be a C∗-algebra and M be a von Neumann algebra. (a) For each norm-closed face F of BA, there exists a unique partial isometry v (b) For each norm-closed face F of BAsa, there exists a unique pair of compact (c) For each weak∗-closed proper face G of BA∗ , there exists a unique nonzero partial belonging locally to A such that F = {x ∈ BA xv∗ = vv∗}. projections p, q such that pq = 0 and F = {x ∈ BAsa x(p − q) = p + q}. isometry v belonging locally to A such that G = {v}′. TINGLEY'S PROBLEM FOR OPERATOR ALGEBRAS 5 sa, there exists a unique pair of com- (d) For each weak∗-closed proper face G of BA∗ (e) For each σ-weakly closed face G of BM (resp. BMsa), there exists a unique pact projections p, q such that p + q 6= 0, pq = 0 and G = {p − q}′. partial isometry (resp. self-adjoint partial isometry) v in M such that G = {x ∈ BM xv∗ = vv∗} = v + (1 − vv∗)BM (1 − v∗v) (resp. G = {x ∈ BMsa xv = v2} = v + (1 − v2)BMsa(1 − v2)). (f ) For each norm-closed proper face F of BM∗ (resp. BM∗sa), there exists a unique nonzero partial isometry (resp. self-adjoint partial isometry) v in M such that F = {v}′. See also [5] for a variant of this result in the setting of JBW∗-triples. 3. On Tingley's problem between unital C∗-algebras For a unital C∗-algebra A, the symbol U(A) will denote the group of unitaries in A, and P(A) stands for the set of projections in A. These substructures contain a lot of information about A. What we focus on in this section is the group U(A). In the proof of [23, Theorem 4.12], Tanaka showed that if T : S(M ) → S(N ) is a surjective isometry between the unit spheres of two finite von Neumann algebras, then T restricts to a bijection between their unitary groups, i.e. T (U(M )) = U(N ). Recently, this result was extended to the case of general von Neumann algebras by Fern´andez-Polo and Peralta [8, Theorem 3.2]. We further extend these results to the case of arbitrary unital C∗-algebras using somewhat a different method. We would like to use the notation E(X) := ext(BX ) for the set of extreme points of BX where X is a Banach space. Recall that, if A is a unital C∗-algebra, then E(A) = {x ∈ S(A) (1 − xx∗)A(1 − x∗x) = {0}} is the set of maximal partial isometries in A and in particular U(A) ⊂ E(A) (see for example [13, Theorem 7.3.1]). Lemma 3.1. Let A be a unital C∗-algebra and x ∈ E(A). Then x is in U(A) if and only if the set Ax := {y ∈ E(A) kx ± yk = √2} has an isolated point as a metric space. The idea of this lemma comes from the easiest case A = C: U(A) = {z ∈ C z = 1}, we see Ax = {ix,−ix}. Proof. First realize A as a unital C∗-subalgebra of some B(H) (the algebra of bounded linear operators on a complex Hilbert space H). Suppose x is in U(A). For y ∈ Ax, we have 2 = kx±yk2 = k1+y∗y±(x∗y+y∗x)k. Decompose H in the form H = y∗yH ⊕ (1 − y∗y)H. Using this decomposition, we express for x ∈ E(A) = By the same decomposition, we can express z2 0 0(cid:19) : y∗yH ⊕ x∗y =(cid:18)z1 1 + y∗y ± (x∗y + y∗x) =(cid:18)2 ± (z1 + z∗ (1 − y∗y)H ±z2 → y∗yH ⊕ . (1 − y∗y)H 1 ) ±z∗ 2 1 (cid:19) . 6 M. MORI Since 2 ∈ 2E(B(y∗yH)), by the norm condition we obtain z1 + z∗ Since x ∈ U(A), it follows that x∗y ∈ E(A). Combining this with the equation 1 = 0 and z2 = 0. x∗y =(cid:18)z1 0 0 0(cid:19) =(cid:18)−z∗ 0 1 0 0(cid:19) , we have x∗y ∈ U(A) and the spectrum σ(x∗y) of x∗y is a subset of {i,−i}. It follows that Ax = ix(1 − 2P(A)) = i(1 − 2P(A))x, which has isolated points ±ix. Next suppose x /∈ U(A) and y ∈ Ax. We show y is not isolated in Ax. We may assume xx∗ 6= 1. Suppose (1 − xx∗)y 6= 0. For c ∈ T := {z ∈ C z = 1}, set c := (xx∗ + c(1 − xx∗))y (∈ E(A)). Then we have y′ ck = kx± (xx∗ + c(1 − xx∗))yk = k(xx∗ + c(1− xx∗))x ± yk = kx± yk = √2. kx± y′ Hence y′ c ∈ Ax. Since y′ c → y (c → 1), y is not isolated in Ax. Similarly, y is not isolated in Ax if (1 − x∗x)y∗ 6= 0. In what follows, we assume (1 − xx∗)y = 0 = (1 − x∗x)y∗. Then we obtain xx∗ ≥ yy∗ and x∗x ≥ y∗y. Since y ∈ E(A), we have (1 − yy∗)A(1 − y∗y) = 0. Taking the closure in the sot of B(H), we also have (1 − yy∗)A (1 − y∗y) = 0. By the theory of von Neumann algebras, there exists a central projection p in A such that y1 := yp is an isometry on pH and y2 := y(1 − p) is a coisometry (i.e. the adjoint operator of an isometry) on (1 − p)H. Set x1 := xp and x2 := x(1 − p). Then it follows that x∗ 1x1 = y∗ 1 or 2y2, then we have √2 = kx±yk = maxn=1,2 kxn±ynk = 2, a contradiction. 2x2 6= y∗ x∗ It follows that xx∗ = yy∗ and x∗x = y∗y. The same discussion as in the first half of this proof shows that there exists a projection q in A with q ≤ xx∗ such that y = i(1 − 2q)x = i(xx∗ − 2q)x. Suppose that q is isolated in P(xx∗Axx∗). Let a ∈ (xx∗Axx∗)sa. Since the map- ping R ∋ t 7→ eitaqe−ita ∈ P(xx∗Axx∗) is norm-continuous, we obtain eiaqe−ia = q. By the Russo-Dye theorem (see for example Exercise 10.5.4 of [13]) it follows that q is central in xx∗Axx∗. In this case, we have y′′ θ := yxx∗ + (y cos θ + sin θ)(1− xx∗) ∈ E(A) for θ ∈ R, and simple calculations show that 2x2 ≥ y∗ 1 6= y1y∗ 1 ≥ y1y∗ 2y2. If x1x∗ 1 and x2x∗ 2 = y2y∗ 2, x∗ 1y1, x1x∗ sot sot 1 √2 (x ± y′′ θ ) = 1 √2 ((x ± y)xx∗ + (x ± (y cos θ + sin θ))(1 − xx∗)) θ ∈ Ax. Since y′′ are partial isometries. In particular, y′′ θ → y (θ → 0), y is not isolated in Ax. If q is not isolated in P(xx∗Axx∗), take qn ∈ P(xx∗Axx∗), n ∈ N = {1, 2, . . .} such that q 6= qn → q (n → ∞). Then we have y 6= i(xx∗ − 2qn)x =: y′′′ n ∈ Ax and y′′′ n → y (n → ∞). Now we can prove the main theorem of this section. (cid:3) Theorem 3.2. Let A and B be unital C∗-algebras and T : S(A) → S(B) be a surjective isometry. Then T (U(A)) = U(B). Proof. We know by Proposition 2.2 that T (E(A)) = E(B) and by (b) of Proposition 2.3 that T (−x) = −T (x) for every x ∈ E(A). It follows that T (Ax) = AT (x) for every x ∈ E(A). Therefore the preceding lemma implies T (U(A)) = U(B). (cid:3) Recall the following theorem due to Hatori and Moln´ar [9]: TINGLEY'S PROBLEM FOR OPERATOR ALGEBRAS 7 Theorem 3.3 (Hatori and Moln´ar [9, Theorem 1]). Let A and B be unital C∗- algebras and T : U(A) → U(B) be a surjective isometry. Then there exists a real a ∈ Asa. In fact, there exists a Jordan ∗-isomorphism J : A → B and a central linear surjective isometry eT : A → B which satisfies T (eia) = eT (eia) for every projection p ∈ B such that eT (x) = T (1)(pJ(x) + (1 − p)J(x)∗) for all x ∈ A. Φ := eT −1 ◦ T : S(A) → S(A) (in the sense of the preceding theorems) is equal to The Russo-Dye theorem shows that such a linear isometry is unique. In order to solve Tingley's problem between unital C∗-algebras, it suffices to show that the identity mapping on S(A). Before we end this section we give an additional partial result. The notation A−1 means the set of invertible elements for a unital C∗-algebra A. Proposition 3.4. Let A be a unital C∗-algebra and Φ : S(A) → S(A) be a surjective isometry such that Φ(eia) = eia for every a ∈ Asa. Then Φ(x) = x for every x ∈ S(A) ∩ A−1. Proof. First we show Φ(u) = u for an arbitrary unitary u ∈ U(A). Consider the functional calculus in A∗∗, and set v1 = uχ{Re z≥0}(u), v2 = uχ{Re z≤0}(u) ∈ A∗∗, which are partial isometries belonging locally to A. Take continuous functions f, g : T → {z ∈ T Re z ≥ 0} which satisfy the following two properties: f (z) = z = g(z) for every z ∈ T with Re z ≥ 0, and Im f (z) > Im g(z) for every z ∈ T with Re z < 0. It follows that v1 is the maximum partial isometry in the collection of every partial isometry v0 ∈ A∗∗ which satisfies f (u)v∗ 0 = v0v∗ 0. Thus the minimum norm-closed face in BA which contains both f (u) and g(u) is the face {x ∈ S(A) xv∗ 1}. Since f (u), g(u) ∈ eiAsa , it follows that Φ({x ∈ S(A) xv∗ 1}. Similarly, Φ({x ∈ S(A) 2}) = {x ∈ S(A) xv∗ 2 = v2v∗ xv∗ 1} ∩ {x ∈ S(A) xv∗ 2 = v2v∗ A−1 ∩ A+. Set c := min(σ(a)) (> 0) and Next we show Φ(a) = a for an arbitrary positive invertible element a ∈ S(A) ∩ 1}) = {x ∈ S(A) xv∗ 2} = {u}, we obtain Φ(u) = u. 2}. Since {x ∈ S(A) xv∗ 0 = v0v∗ 0 and g(u)v∗ 1 = v1v∗ 2 = v2v∗ 1 = v1v∗ S := {u ∈ U(A) ku − ak = 1 − c} = {u ∈ U(A) ku − Φ(a)k = 1 − c}. We see Re λ ≥ c/2 for every λ ∈ σ(u), u ∈ S. Assume there exists a λ ∈ σ(u) such that Re λ < c/2. Realizing A ⊂ B(H), we obtain unit vectors ξn ∈ H, n ∈ N such that kuξn − λξnk → 0 (n → ∞). Then it follows that limn→∞huξn, ξni = λ and haξn, ξni ≥ c for every n ∈ N. We have ku − ak ≥ 1, a contradiction. We consider the surjective isometry u 7→ u∗ on S. By the observation above, it follows that 1 = v1v∗ 1 = v1v∗ ku∗ − uk = k1 − u2k = k(1 + u)(1 − u)k ≥(cid:16)1 + c 2(cid:17) k1 − uk for every u ∈ S. Since 1∗ = 1 ∈ S, 1 + c/2 > 1 and S is bounded, it follows by [25, Theorem 1.2] that τ (1) = 1 for every surjective isometry τ : S → S. Since k1− Φ(a)k = k1− ak = 1 − c < 1, the polar decomposition Φ(a) = vΦ(a) satisfies v ∈ U(A). For u ∈ S, we have kvu∗v − Φ(a)k = ku∗v − Φ(a)k = kv∗u − Φ(a)k = ku − Φ(a)k = 1 − c. Thus the mapping u 7→ vu∗v is a surjective isometry on S. Therefore, by the commented result in [25], it follows that 1 = v1∗v = v2. Combining this with the equation kv∗ − Φ(a)k = k1 − Φ(a)k = 1 − c, we obtain v = 1. i.e. Φ(a) is positive. 8 M. MORI Take the continuous function f0 : [c, 1] → {z ∈ T Im z ≥ 0} which is uniquely determined by the condition t − f0(t) = 1 + c, t ∈ [c, 1]. Put w := f0(a). Then (a − w)/(1 + c) is a unitary. Assume Φ(a) 6≤ a. Then there exist λ > 0 and unit vectors ηn ∈ H, n ∈ N such that k(Φ(a) − a)ηn − ληnk → 0 (n → ∞). We have h(Φ(a) − w)ηn, (a − w)ηni = h(Φ(a) − a)ηn, (a − w)ηni + (1 + c)2. We know that limn→∞h(Φ(a) − a − λ)ηn, (a − w)ηni = 0. Since Re(t − f0(t)) ≥ √c2 + 2c for every t ∈ [c, 1], we also know that λ 2 hηn, ((a − w) + (a − w)∗)ηni ≥ λpc2 + 2c > 0 Re λhηn, (a − w)ηni = for every n ∈ N. We have (1 + c)2 = kΦ(a) − wkka − wk ≥ lim n→∞ = lim n→∞ Reh(Φ(a) − w)ηn, (a − w)ηni Re λhηn, (a − w)ηni + (1 + c)2 > (1 + c)2, a contradiction. Therefore we obtain Φ(a) ≤ a and similarly a ≤ Φ(a). Lastly we show Φ(x) = x for an arbitrary x ∈ S(A) ∩ A−1. The polar decom- position x = u0x satisfies u0 ∈ U(A) and x ∈ S(A) ∩ A−1 ∩ A+. Consider the surjective isometry Ψ : S(A) → S(A) which is defined by Ψ(y) := u−1 0 Φ(u0y), y ∈ S(A). Then the first part of this proof shows Ψ(u) = u for every u ∈ U(A). The second part of this proof shows x = Ψ(x) = u−1 0 Φ(u0x), hence Φ(x) = Φ(u0x) = u0x = x. (cid:3) 4. Tingley's problem between preduals of von Neumann algebras In this section, we present an affirmative answer to Tingley's problem when the two spaces are preduals of von Neumann algebras. Our theorem extends the result of Fern´andez-Polo, Garc´es, Peralta and Villanueva [6], in which Tingley's problem for the spaces of trace class operators on complex Hilbert spaces is solved affirmatively. Let M be a von Neumann algebra. By (f ) of Theorem 2.4 we know that for every norm-closed proper face F ⊂ BM∗ there exists a unique nonzero partial isometry v ∈ M such that F = {v}′. Recall that for a metric space (X, d) and nonempty subsets X1, X2 ⊂ X, the Hausdorff distance between X1 and X2 is defined by dH (X1, X2) := max{ sup x∈X1 inf y∈X2 d(x, y), sup y∈X2 inf x∈X1 d(x, y)}. Endow the space of nonzero partial isometries in M with the distance δH (v, w) := dH ({v}′,{w}′). (It is easy to show that δH actually satisfies the axioms of distance.) Lemma 4.1. Let M ⊂ B(H) be a von Neumann algebra. (a) Let w1, w2 ∈ M be nonzero partial isometries with w∗ (b) ku − vk = δH (u, v) for every u ∈ U(M ) and every v ∈ E(M ). Proof. (a) Suppose w∗ ϕ(w1w∗ 2w2. Let ϕ ∈ {w1}′. Then defining ψ(x) := 2. Then kw1 − w2k ≥ δH (w1, w2). 2w2 or w1w∗ 1w1 = w∗ 1w1 = w∗ w2w∗ 1 = 2x) (x ∈ M ), we have ψ ∈ {w2}′ and kϕ − ψk = kϕ((w1w∗ 1 − w1w∗ 2)· )k ≤ kw1w∗ 1 − w1w∗ 2k ≤ kw1 − w2k. TINGLEY'S PROBLEM FOR OPERATOR ALGEBRAS 9 sup inf inf sup ϕ∈{w1}′ Therefore we obtain ψ∈{w2}′ kϕ − ψk ≤ kw1 − w2k. Similarly we can see ϕ∈{w1}′ kϕ − ψk ≤ kw1 − w2k, and therefore δH (w1, w2) ≤ kw1 − w2k. A ψ∈{w2}′ similar discussion can be applied in the case w1w∗ (b) Suppose first that u, v ∈ U(M ). The inequality δH (u, v) ≤ ku − vk follows from (a). We may assume u = 1. In that case, we have ku − vk = k1 − vk = supλ∈σ(v) 1 − λ. Take λ0 ∈ σ(v) which attains this supremum. Since λ0 ∈ σ(v), there exist ξn ∈ H with kξnk = 1, n ∈ N such that kvξn − λ0ξnk → 0 (n → ∞). Define ϕn := ωξn,vξn = h· ξn, vξni (∈ {v}′). Then for every ψ ∈ {1}′ we have 1 = w2w∗ 2, too. kψ − ϕnk ≥ ψ(1) − ϕn(1) = 1 − hξn, vξni → 1 − λ0 = k1 − vk (n → ∞). Pn k=1(−1)kvkξ and ϕ′ Therefore we obtain δH (1, v) ≥ k1− vk. The proof when u, v ∈ U(M ) is completed. Let us assume next v /∈ U(M ). We may assume u = 1 and vv∗ 6= 1. In that case, it follows that ku − vk = k1 − vk = 2. Take a unit vector ξ ∈ (1 − vv∗)H. Since v∗v(1 − vv∗) = 1 − vv∗, the system {vnξ}n∈N is orthonormal. Define ηn := n := n−1ωηn,vηn (∈ {v}′) for n ∈ N. Then for every ψ ∈ {1}′ we have kϕ′ n → 2 It follows that δH (1, v) ≥ 2. The inequality δH (1, v) ≤ 2 is trivial. n(1) − ψ(1) =(cid:12)(cid:12)(cid:12)(cid:12)− n − 1(cid:12)(cid:12)(cid:12)(cid:12) = n − ψk ≥ ϕ′ (n → ∞). 2n − 1 n − 1 (cid:3) for every pair of nonzero partial isometries w1, w2 ∈ M with w∗ w1w∗ Note that using the same discussion as in (b), we also gain kw1−w2k = δH (w1, w2) 2w2 and 1 = w2w∗ 2. The author does not know whether kv − wk = δH(v, w) holds for every pair v, w ∈ E(M ), but the following lemma which is similar to Lemma 3.1 holds. Lemma 4.2. Let M be a von Neumann algebra and x be in E(M ). Then x is in point with respect to the metric δH . U(M ) if and only if the set bAx := {y ∈ E(M ) δH (x,±y) ≤ √2} has an isolated 1w1 = w∗ Proof. The proof is parallel to that of Lemma 3.1. Suppose x is in U(M ). The preceding lemma shows that δH (x, y) = kx − yk for every y ∈ E(M ). By the same discussion as in the proof of Lemma 3.1, we obtain proof of the Lemma 3.1. Note that the operators y′ n have the same initial spaces as y. Hence it is not difficult to see that the preceding lemma shows y is not (cid:3) bAx = i(2P(M ) − 1)x (⊂ U(M )), which have isolated points ±ix. Next suppose x /∈ U(M ) and y ∈ bAx. We again use the argument as in the isolated in bAx with respect to the metric δH. We state the main theorem of this section: θ , y′′′ c, y′′ Theorem 4.3. Let M and N be von Neumann algebras and T : S(M∗) → S(N∗) be a surjective isometry. Then there exists a unique real linear surjective isometry We start proving. Since T gives a bijection between the classes of maximal convex sets in unit spheres, a bijection T1 : E(M ) → E(N ) is determined by T ({v}′) = {T1(v)}′, eT : M∗ → N∗ which extends T . v ∈ E(M ) (see Proposition 2.2). We also have T1(bAv) = bAT1(v) for any v ∈ E(M ). 10 M. MORI By the preceding lemma and (a) of Proposition 2.3, T1 restricts to a bijection between unitary groups. Moreover, by (b) of Lemma 4.1, this is a surjective isometry between unitary groups. By the theorem of Hatori and Moln´ar, there exists a −1 are σ-weakly continuous since they can be expressed by Jordan ∗-isomorphisms. unique real linear surjective isometryfT1 : M → N such that T1(u) =fT1(u) for all u ∈ eMsa = U(M ). Note thatfT1 andfT1 Now we can construct a real linear surjective isometry T2 : N ∗ → M ∗ which is canonically determined byfT1 as the following: By the σ-weak continuity of fT1 and fT1 , T2 restricts to a real linear surjective −1 : M∗ → N∗ is the isometry from N∗ onto M∗. We would like to show that T2 extension we wanted. In order to show this, it suffices to show that the surjective isometry Φ := T2 ◦ T : S(M∗) → S(M∗) is equal to the identity mapping on S(M∗). We know that Φ({u}′) = {u}′ for every u ∈ U(M ). (T2ϕ)(x) = (Re ϕ)(fT1(x)) − i(Re ϕ)(fT1(ix)), ϕ ∈ N ∗, x ∈ M. −1 Let v ∈ M be a nonzero partial isometry which has a unitary extension u. Then {v}′ = {u}′∩{2v−u}′, and since u and 2v−u are unitaries we have Φ({v}′) = {v}′. Let v ∈ M be a nonzero partial isometry which does not admit a unitary ex- tension. Then there exist nonzero sub-partial isometries v1, v2 ∈ M of v which have unitary extensions and satisfy v = v1 + v2. (Indeed, we can take v1 and v2 as follows. Decompose the projection v∗v to the sum of a finite projection p1 and a properly infinite projection p2. Since v does not admit a unitary extension, we have p2 6= 0. Decompose p2 into the sum of mutually Murray-von Neumann equivalent projections p21 and p22. Then v1 = v(p1 + p21) and v2 = vp22 satisfy the condition. See for example [13, Chapter 6] for information about the comparison theory of projections.) Since {v}′ is the minimum norm-closed face in BM∗ which contains both {v1}′ and {v2}′, we obtain Φ({v}′) = {v}′. Therefore, in order to show that Φ is an identity mapping, it suffices to show Φ(ϕ) = ϕ for every normal state ϕ on M (i.e. for every ϕ ∈ {1}′ = S(M∗) ∩ M∗+). Restricting our attention to ((supp ϕ)M (supp ϕ))∗ which can be identified canonically with a subspace of M∗, we may also assume ϕ is faithful (i.e. ϕ(a) 6= 0 for an arbitrary nonzero positive element a in M ). We need some more preparations. Lemma 4.4. Let M be a von Neumann algebra, ϕ ∈ M∗ be a self-adjoint element and p ∈ P(M ). Then kϕk ≥q(ϕ(p) − ϕ(p⊥))2 + 4kϕ(p · p⊥)k2. Proof. Take a partial isometry v ∈ M such that vv∗ ≤ p, v∗v ≤ p⊥ and ϕ(v) = kϕ(p · p⊥)k. Then for every θ ∈ R we have p cos θ − p⊥ cos θ + (v + v∗) sin θ ∈ BMsa. Take the supremum of ϕ(p cos θ − p⊥ cos θ + (v + v∗) sin θ) = (ϕ(p) − ϕ(p⊥)) cos θ + 2kϕ(p · p⊥)k sin θ with respect to θ ∈ R. Lemma 4.5. Let M be a von Neumann algebra, ϕ be a normal state on M and p be in P(M ). Suppose 0 < ϕ(p) < 1. Put λ := ϕ(p). Then the following two conditions are equivalent: (a) There exist ψ1 ∈ {p}′ and ψ2 ∈ {p⊥}′ such that kϕ − ψ1k = 2(1 − λ) and (cid:3) kϕ − ψ2k = 2λ. TINGLEY'S PROBLEM FOR OPERATOR ALGEBRAS 11 (b) ϕ(p · p⊥) = 0 = ϕ(p⊥ · p). Proof. (b) ⇒ (a) Put ψ1 := λ−1ϕ(p · p) and ψ2 := (1 − λ)−1ϕ(p⊥ · p⊥). (a) ⇒ (b) If (b) is not true, then kϕ(p · p⊥)k > 0. Therefore, by the preceding lemma, we have kψ1 − ϕk ≥q((1 − ϕ(p)) + ϕ(p⊥))2 + 4kϕ(p · p⊥)k2 > (1 − ϕ(p)) + ϕ(p⊥), kϕ − ψ2k ≥q(ϕ(p) + (1 − ϕ(p⊥)))2 + 4kϕ(p · p⊥)k2 > ϕ(p) + (1 − ϕ(p⊥)) for every ψ1 ∈ {p}′ and every ψ2 ∈ {p⊥}′. It follows that kϕ− ψ1k +kϕ − ψ2k > 2, so (a) is not true. (cid:3) We return to the proof of Theorem 4.3. Our task is to show Φ(ϕ) = ϕ for every normal faithful state ϕ on M . Set ϕ0 := Φ(ϕ) (∈ Φ({1}′) = {1}′). Assume ϕ 6= ϕ0. Consider the Jordan decomposition of ϕ−ϕ0 (6= 0). We obtain a nonzero projection p ∈ P(M ) such that ϕ(p) < ϕ0(p) and ϕ(p · p) ≤ ϕ0(p · p), ϕ(p · p⊥) = ϕ0(p · p⊥), ϕ(p⊥ · p⊥) ≥ ϕ0(p⊥ · p⊥). Put λ := ϕ(p). Then 0 < λ < 1. Set S0 := {ψ ∈ {1}′ ψ(p) = λ, ψ(p · p⊥) = 0 = ψ(p⊥ · p)}. By the preceding lemma, S0 is equal to {ψ ∈ {1}′ kψ − ψ1k = 2(1 − λ), kψ − ψ2k = 2λ for some ψ1 ∈ {p}′, ψ2 ∈ {p⊥}′}. Thus the equations Φ({p}′) = {p}′, Φ({p⊥}′) = {p⊥}′ imply Φ(S0) = S0. In particular, we have inf ψ∈S0 kϕ − ψk = inf ψ∈S0 kϕ0 − ψk. However, Lemma 4.4 implies ψ∈S0 kϕ − ψk = kϕ − (ϕ(p · p) + ϕ(p⊥ · p⊥))k = 2kϕ(p · p⊥)k inf and ψ∈S0 kϕ0 − ψk ≥ inf inf ψ∈S0q((ϕ0(p) − ψ(p)) − (ϕ0(p⊥) − ψ(p⊥))2 + 4kϕ(p · p⊥))k2 =q4(ϕ0(p) − ϕ(p))2 + 4kϕ(p · p⊥)k2. We have a contradiction. The proof of Theorem 4.3 is completed. (cid:3) Corollary 4.6. Let A and B be C∗-algebras and T : S(A∗) → S(B∗) be a surjective isometry. (We do not assume A or B is unital.) Then there exists a unique real linear surjective isometry eT : A∗ → B∗ which extends T . Proof. We know that A∗ and B∗ can be considered as the preduals of the enveloping von Neumann algebras A∗∗ and B∗∗, respectively. Thus we can apply Theorem 4.3. (cid:3) 12 M. MORI 5. Tingley's problem between the spaces of self-adjoint elements To solve Tingley's problem between the spaces of self-adjoint elements in (pre- duals of) von Neumann algebras, it seems to be difficult to make use of the set of self-adjoint unitaries because the theorem of Hatori and Moln´ar (Theorem 3.3) can- not be applied in this case. What we use in this section is the structure of projection lattices of von Neumann algebras, but note that in general a surjective isometry between projection lattices cannot be extended to a linear surjective isometry. For example, every bijection from P(ℓ∞) onto itself is automatically isometric. However, combining the metric condition with a condition about orthogonal- ity, we see that a mapping between projection lattices can be extended linearly. We rely on the following theorem due to Dye [4]. Let M, N be von Neumann algebras. A bijection T : P(M ) → P(N ) (or P(M ) \ {0} → P(N ) \ {0}) is called an orthoisomorphism if for any projections p, q ∈ P(M ) (or P(M ) \ {0}), pq = 0 ⇐⇒ T (p)T (q) = 0. Theorem 5.1 (Dye [4, Corollary of Theorem 1]). Let M and N be von Neumann algebras and T : P(M ) → P(N ) be an orthoisomorphism. Suppose M does not have a type I2 summand. Then there exists a unique linear surjective isometry eT : Msa → Nsa which extends Φ. The condition M does not have a I2 summand is inevitable in general cases. In order to drop this condition, we add another condition. Proposition 5.2. Let M and N be von Neumann algebras and T : P(M ) → P(N ) be an orthoisomorphism. Suppose kp−qk = kΦ(p)−Φ(q)k for every pair of maximal abelian projections p, q in the type I2 summand of M . Then there exists a unique 0 0 0 0 0 0 projection in M2(A) is fixed under T . We also have in N (∼= M2(A)). Taking an appropriate ∗-isomorphism from N onto M2(A), we Proof. It suffices to show this proposition when M and N are of type I2. Since T restricts to a bijection between the classes of central projections, it follows that M is ∗-isomorphic to N . We decompose M as M = M2(A) using an abelian von linear surjective isometry eT : Msa → Nsa which extends T . Neumann algebra A. Then the element T(cid:18)1 0(cid:19) is a maximal abelian projection may assume M = N = M2(A), T(cid:18)1 0(cid:19) = (cid:18)1 0(cid:19) and every nonzero central 1(cid:19) = T (cid:18)1 0 0 0(cid:19)⊥! = T(cid:18)1 0 0 0(cid:19)⊥ =(cid:18)0 0 0 1(cid:19) . =(cid:13)(cid:13)(cid:13)(cid:13)p −(cid:18)0 0 0 1(cid:19)(cid:13)(cid:13)(cid:13)(cid:13)(cid:27) =(cid:26) 1 1(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)u ∈ U(A)(cid:27) 0(cid:19)(cid:13)(cid:13)(cid:13)(cid:13) = 2(cid:18) 1 2(cid:18) 1 2(cid:18) 1 2(cid:18) 1 −i 1(cid:19)(cid:19) = 1(cid:19) . T(cid:18) 1 1(cid:19) , −i 1(cid:19)(cid:13)(cid:13)(cid:13)(cid:13) = 1(cid:19)(cid:13)(cid:13)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:13)(cid:13) 2(cid:18) 1 2(cid:18)1 1 2(cid:18) 1 1 1(cid:19) − 1(cid:19) − T(cid:18)0 (cid:26)p ∈ P(M2(A))(cid:12)(cid:12)(cid:12)(cid:12)(cid:13)(cid:13)(cid:13)(cid:13)p −(cid:18)1 2(cid:18)1 1 1 1(cid:19)(cid:19) = T(cid:18) 1 =(cid:13)(cid:13)(cid:13)(cid:13) 2(cid:18) 1 Thus T restricts to a bijection from 1 √2 0 0 onto itself. There exist u1, ui ∈ U(A) such that Since ku1 − uik 2 ui i u∗ u∗ i 0 0 1 √2 1 u1 1 ui 1 1 1 u1 u∗ 1 u u∗ 1 u∗ i i 1 TINGLEY'S PROBLEM FOR OPERATOR ALGEBRAS 13 and 1 1 1 1 2 ui u1 u∗ i u∗ 1 ku1 + uik 2(cid:18)1 2(cid:18) 1 2(cid:18) 1 1(cid:19) − =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 1(cid:19)⊥(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) −i 1(cid:19)⊥(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 2(cid:18) 1 it follows that (u1 ± ui)/√2 ∈ U(A). We define a linear surjective isometry eT : eT(cid:18) a1 M2(A)sa → M2(A)sa by a2 + a3i a4 (cid:19) =(cid:18) a1, a2, a3, a4 ∈ Asa. =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) a1 1 + a3u∗ 1(cid:19) − a2 − a3i a2u1 + a3ui (cid:19) , 1 √2 a2u∗ a4 = 1 1 i i , Let c ∈ T. Consider the distance from c c 1 1 −1 1(cid:19) to 2(cid:18)1 1 (cid:19) =(cid:18) 1 1 1(cid:19)(cid:19)⊥! , 2(cid:18)1 1 2(cid:18) 1 −1 1(cid:19) =(cid:18) 1 −i 1(cid:19)(cid:19)⊥! . 2(cid:18) 1 2(cid:18)1 −i 2(cid:18)1 (cid:19) = eT(cid:18) 1 u1 Re c + ui Im c i Im c 1 1 1 c i i Then we easily obtain i 1 1 2(cid:18)1 1 1 1(cid:19) , 2(cid:18) 1 −i 1(cid:19) and 2(cid:18)1 2(cid:18) T(cid:18) 1 1(cid:19)(cid:19) = c2(cid:19)(cid:19) = 1 + c2(cid:18)1 T(cid:18) 1 1 c c c c A similar consideration shows that 1 Re c + u∗ u∗ c 1(cid:19)(cid:19) . (cid:19) u1 Re c + ui Im c c2 cnqn cn2qn(cid:19)! c c 1 1 cnqn i Im c 1 Re c + u∗ u∗ 1 + c2(cid:18) 1 + c2(cid:18)1 c2(cid:19)(cid:19) = eT(cid:18) 1 cn2qn(cid:19)! = eT NXn=1 1 + cn2(cid:18) qn cn2qn(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) cnqn 1 cnqn for an arbitrary c ∈ C. Since T is an orthoisomorphism, we obtain T NXn=1 1 1 + cn2(cid:18) qn cnqn n=1 ⊂ P(A) with q1 + for arbitrary numbers c1, . . . , cN ∈ C and projections {qn}N ··· + qN = 1 ∈ A. The set 1 cnqn N ∈ N, c1, . . . , cN ∈ C, 1 + cn2(cid:18) qn is norm-dense in the class of maximal abelian projections in P(M2(A)), so we ( NXn=1 q1,··· , qN ∈ P(A), q1 + ··· + qN = 1 ∈ A) have T = eT on this class. Let p ∈ M2(A). Then p can be decomposed as p = q1(cid:19) p0, where q0 and q1 are mutually orthogonal projections in A q0(cid:19) +(cid:18)q1 (cid:18)q0 we obtain T (p) = eT (p). and p0 is a maximal abelian projection in M2(A). Since T is an orthoisomorphism, (cid:3) We also make use of the following proposition, whose proof can be found in the 0 0 0 0 paper of Akemann and Pedersen [1]. 14 M. MORI Proposition 5.3 (See [1, Lemma 2.7]). Let A be a C∗-algebra, p be a compact projection, q be an open projection with p ≤ q. Then there exist a decreasing net (xα) and an increasing net (yα) in A+ such that p ≤ xα, yα ≤ q with the property xα converges to p and yα converges to q σ-strongly in A∗∗. Using this, we obtain the following proposition. Proposition 5.4. Let M be a von Neumann algebra and F ⊂ BMsa be a norm- closed proper face. Then F is σ-weakly closed if and only if there exists a unique element xF ∈ F such that kxF − yk ≤ 1 for every y ∈ F . Proof. Suppose F is σ-weakly closed. Then by (e) of Theorem 2.4 there exists a unique pair of projections p, q ∈ P(M ) such that pq = 0 and F = p − q + (1 − p − q)BMsa (1 − p − q) = p − q + B((1−p−q)M(1−p−q))sa . Then x := p − q is the only element which satisfies given conditions. pair of compact projections p, q such that pq = 0 and Suppose F is not σ-weakly closed. By (b) of Theorem 2.4, there exists a unique F = {y ∈ Msa y(p − q) = p + q} = {y ∈ Msa 2p − 1 ≤ y ≤ 1 − 2q}. Since F is not σ-weakly closed, at least one of p and q is not an element of P(M ). Let x be in F . Then 0 6= x−(p−q) ∈ (1−p−q)M ∗∗ sa (1−p−q). However, by the preceding lemma, there exist nets (aα), (bα) ∈ F such that aα ց 2p − 1 and bα ր 1 − 2q σ-strongly in M ∗∗. Hence we have x−aα → x−2p+1 = (x−(p−q))+(1−p−q) and x− bα → x− 1 + 2q = (x− (p− q))− (1− p− q) (σ-strongly), thus lim kx− aαk > 1 or limkx − bαk > 1. Therefore, there exists no element x which satisfies the given conditions. (cid:3) Therefore, we can detect σ-weakly closed faces in BMsa from the class of norm- closed faces only by the metric structure of them. Recall that Mankiewicz's generalization of the Mazur-Ulam theorem states that every surjective isometry between open connected nonempty subsets of real Banach spaces extends to an affine surjective isometry between the whole spaces [15]. Now we can prove the following theorem: Theorem 5.5. Let M, N be von Neumann algebras and T : S(Msa) → S(Nsa) be a surjective isometry. Then there exists a unique linear surjective isometry eT : Msa → Nsa which extends T . Proof. Propositions 2.2 and 5.4 imply that for a σ-weakly closed proper face F1 ⊂ BMsa, F2 := T (F1) is a σ-weakly closed proper face in BNsa and T (xF1) = xF2. Therefore T restricts to a bijection between the classes of nonzero self-adjoint partial isometries. We also have T (−v) = −T (v) for every nonzero self-adjoint partial isometry v ∈ M by (b) of Proposition 2.3. We know that T restricts to a bijection from E(Msa) (= {2p − 1 p ∈ P(M )}, which is the collection of self-adjoint unitaries in M ) onto E(Nsa). Since u ∈ E(Msa) is central if and only if u is isolated in E(Msa) (see the last two paragraphs of the proof of Lemma 3.1), it follows that T (1) is central in N . Define T1 : S(Msa) → S(Nsa) by T1(x) := T (1)−1T (x), x ∈ S(Msa). What we have to do is to show that the mapping T1 admits a linear extension. We first show that T1 restricts to an orthoisomorphism from P(M ) \ {0} onto P(N ) \ {0}. We already know that T1 restricts to an order-preserving bijection from P(M ) \ {0} TINGLEY'S PROBLEM FOR OPERATOR ALGEBRAS 15 onto P(N ) \ {0}. Hence it suffices to show T1(p⊥) = T1(p)⊥ for an arbitrary projection p ∈ P(M )\{0, 1}. Let p ∈ P(M )\{0, 1}. Then T1 restricts to a bijection from p + p⊥BMsap⊥ onto T1(p) + T1(p)⊥BNsaT1(p)⊥. Identify p + p⊥BMsap⊥ with p⊥BMsa p⊥ = B(p⊥M p⊥)sa and T1(p) + T1(p)⊥BNsaT1(p)⊥ with B(T (p)⊥N T (p)⊥)sa. It follows by Mankiewicz's theorem that T1(p) = T1(cid:18) 1 2 ((p − p⊥) + 1)(cid:19) = 1 2 (T1(p − p⊥) + T1(1)) = 1 2 (T1(p − p⊥) + 1) Similarly we obtain T1(p⊥) = (T1(p⊥ − p) + 1)/2. Since T1(p − p⊥) = −T1(p⊥ − p), we obtain T1(p) + T1(p⊥) = 1. i.e. T1(p⊥) = T1(p)⊥. T (p) + T (p)⊥BNsaT (p)⊥. Moreover, they coincide on {p0 ∈ P(M ) \ {0} p ≤ p0}, which is total in B(p⊥M p⊥)sa identified with p + p⊥BMsap⊥. Hence Mankiewicz's Thus Proposition 5.2 implies that there exists a linear surjective isometry fT1 : Msa → Nsa such that T1(p) =fT1(p) for every p ∈ P(M )\{0}. Let p ∈ P(M )\{0}. Then both T1 and fT1 restrict to surjective isometries from p + p⊥BMsap⊥ onto theorem implies that T1(x) =fT1(x) for every x ∈ p + p⊥BMsap⊥. Similarly we have T1(x) =fT1(x) for every x ∈ −p + p⊥BMsap⊥. By the functional calculus, we know that the set [p∈P(M)\{0}(cid:0)(cid:0)p + p⊥BMsap⊥(cid:1) ∪(cid:0)−p + p⊥BMsap⊥(cid:1)(cid:1) is norm-dense in S(Msa). Thus we obtain T1(x) =fT1(x) for every x ∈ S(Msa). (cid:3) Remark 5.6. In [18, Theorem 5.8], using the Bunce-Wright-Mackey-Gleason the- orem instead of Dye's theorem, Peralta gave another way to show the preceding theorem. In order to think about the space of self-adjoint elements in the preduals of von Neumann algebras, we again use the Hausdorff distance as in Section 4. Lemma 5.7. Let M be a von Neumann algebra of type I2. Then for arbitrary maximal abelian projections p, q ∈ M we have 2kp − qk = δH (p, q). Proof. As in the proof of Proposition 5.2, we can decompose M as M = M2(A) us- 0 0(cid:19) ∈ P(M2(A)). Hence we may also assume that q can be decomposed to the following form: Let q1, q2 ∈ P(M2(A)). If kq1 − q2k is sufficiently small, for every ϕ ∈ {q1}′, it is not difficult to see that ϕ(q2 · q2) kϕ(q2 · q2)k ∈ {q2}′ Thus the mapping P(M2(A)) ∋ q 7→ δH (p, q) is continuous in the norm metric. there exist N ∈ N, q1, . . . , qN ∈ P(A) and c1, . . . , cN ∈ C such that ing an abelian von Neumann algebra A. We may assume p =(cid:18)1 0 kϕ(q2 · q2)k(cid:13)(cid:13)(cid:13)(cid:13) is small. cn2qn(cid:19) ∈ M2(A). cn2(cid:19)(cid:13)(cid:13)(cid:13)(cid:13) = max (cid:13)(cid:13)(cid:13)(cid:13)ϕ − 1 + cn2(cid:18) qn NXn=1 1 + cn2(cid:18) 1 0(cid:19) − qn = 1 ∈ A, In this case, we easily have 1≤n≤N(cid:13)(cid:13)(cid:13)(cid:13)(cid:18)1 cnp1 + cn2 kp − qk = max 1 cn cn ϕ(q2 · q2) NXn=1 q = 1 cnqn cnqn 0 0 and 1≤n≤N 16 and δH (p, q) = max 1≤n≤N δH(cid:18)(cid:18)1 0 0 0(cid:19) , M. MORI 1 1 + cn2(cid:18) 1 cn In particular, we obtain 2kp − qk = δH (p, q). Let us recall the following well-known fact. cn cn2(cid:19)(cid:19) = max 1≤n≤N . 2cn p1 + cn2 (cid:3) Lemma 5.8. Let M be a von Neumann algebra and ϕ, ψ be normal states on M . Then we have kϕ − ψk = 2 if and only if supp ϕ ⊥ supp ψ. Proof. Suppose kϕ − ψk = 2. There exists a self-adjoint partial isometry v ∈ M such that ϕ(v) − ψ(v) = 2. We decompose as v = p − q, where p, q ∈ P(M ) are mutually orthogonal projections. Since ϕ and ψ are states, by the equation ϕ(v) − ψ(v) = 2, we have ϕ(p) = 1 and ψ(q) = 1, thus supp ϕ ≤ p ⊥ q ≥ supp ψ. The other implication is clear. (cid:3) Now we are ready to prove the following theorem. Theorem 5.9. Let M, N be von Neumann algebras and T : S(M∗sa) → S(N∗sa) be a surjective isometry. Then there exists a unique linear surjective isometry eT : M∗sa → N∗sa which extends T . Proof. We know T ({1}′) is a maximal convex set in S(N∗sa), so it can be written as {u}′, where u ∈ N is a self-adjoint unitary. Consider the space E(Msa) endowed with the metric δH. By Lemma 4.1, this metric is equal to the norm metric. Thus the same discussion as in the second paragraph in the proof of Theorem 5.5 shows that u is central. It suffices to show that the surjective isometry T1 defined by T1(ϕ) := (T ϕ)(· u), ϕ ∈ S(M∗sa) admits a linear isometric extension. We can define T2 : P(M )\{0} → P(N ) \ {0} by T1({p}′) = {T2(p)}′, p ∈ P(M ) \ {0}. By the preceding lemma, it is easy to see that for p, q ∈ P(M )\{0}, pq = 0 ⇐⇒ dist({p}′,{q}′) = 2. It follows that T2 is an orthoisomorphism. Since every orthoisomorphism restricts to a bijection between the classes of max- imal abelian projections of the type I2 summands, the preceding lemma implies that kp − qk = kT2(p) − T2(q)k for arbitrary maximal abelian projections p, q in the I2 summand of M . Therefore by Proposition 5.2, there exists a linear surjective isom- etry fT2 : Msa → Nsa such that T2(p) = fT2(p) for every p ∈ P(M ) \ {0}. Then we can show that (fT2∗)−1 is the linear surjective isometry we wanted, using an argument similar to the proof of Theorem 4.3. (cid:3) Like Corollary 4.6, we have the following corollary. Corollary 5.10. Let A and B be C∗-algebras and T : S(A∗ sa) be a surjective isometry. (We do not assume A or B is unital.) Then there exists a sa) → S(B∗ unique linear surjective isometry eT : A∗ A similar discussion can be applied to prove the next theorem. sa → B∗ sa which extends T . Theorem 5.11. Let M and N be von Neumann algebras. (a) Suppose T : S(M∗) ∩ M∗+ → S(N∗) ∩ N∗+ is a surjective isometry between normal state spaces. Then there exists a unique linear surjective isometry eT : M∗sa → N∗sa which extends T . TINGLEY'S PROBLEM FOR OPERATOR ALGEBRAS 17 (b) Suppose the dimension of N is larger than one and T : BM∗∩M∗+ → BN∗∩N∗+ is a surjective isometry between normal quasi-state spaces. Then there exists a Proof. (a) For p ∈ P(M ) \ {0}, we easily see that unique linear surjective isometry eT : M∗sa → N∗sa which extends T . {p}′ = {ϕ ∈ S(M∗) ∩ M∗+ kϕ − ψk = 2 for any ψ ∈ {p⊥}′}. It follows that T ({p}′) = {ϕ ∈ S(N∗) ∩ N∗+ kϕ − T (ψ)k = 2 for any ψ ∈ {p⊥}′} = {ϕ ∈ S(N∗) ∩ N∗+ supp ϕ ⊥ supp T (ψ) for any ψ ∈ {p⊥}′}. show that T1 admits a unique linear surjective isometric extensionfT1 : Msa → Nsa. Thus there exists an orthoisomorphism T1 : P(M )\{0} → P(N )\{0} which satisfies T ({p}′) = {T1(p)}′ for every p ∈ P(M )\{0}. Then Lemma 5.7 and Proposition 5.2 Use discussions in Section 4 to show that this linear mapping is what we wanted. (b) First we see T (0) = 0. By [14, Lemma 3.6], we have T (0) = 0 unless N is equal to C ⊕ C or M2(C). It is easy to show T (0) = 0 if N = C ⊕ C or N = M2(C). Thus T restricts to a bijection from S(M∗) ∩ M∗+ onto S(N∗) ∩ N∗+. Hence (a) shows that there exists a unique linear surjective isometry eT : Msa → Nsa such that eT (ϕ) = T (ϕ) for all ϕ ∈ S(M∗) ∩ M∗+. It suffices to show that Φ := eT −1 ◦ T : BM∗ ∩ M∗+ → BM∗ ∩ M∗+ is equal to the identity mapping. Let ϕ ∈ BM∗ ∩ M∗+. Since Φ is a surjective isometry and Φ(ψ) = ψ for every ψ ∈ S(M∗) ∩ M∗+, the set Sϕ = {ψ ∈ S(M∗) ∩ M∗+ ϕ ≤ ψ} = {ψ ∈ S(M∗) ∩ M∗+ kψ − ϕk = 1 − kϕk} is equal to SΦ(ϕ) = {ψ ∈ S(M∗) ∩ M∗+ Φ(ϕ) ≤ ψ} = {ψ ∈ S(M∗) ∩ M∗+ kψ − Φ(ϕ)k = 1 − kΦ(ϕ)k}. Since {ϕ0 ∈ BM∗ ∩ M∗+ kϕ0k = kϕk and ϕ0 ≤ ψ for any ψ ∈ Sϕ} = {ϕ}, we obtain Φ(ϕ) = ϕ. (cid:3) Note that (b) answers the question in [14, Remark 3.11] positively in the case p = 1. See also [12, Theorem (4.5)], in which Kadison proved (a) with an additional assumption of affinity, and [16, Theorem 4], in which the case M = N = B(H) for (a) is solved. 6. Problems The results in this paper may be extended to Tingley's problem between various types of Banach spaces concerning operator algebras. In this section, we give some problems which seem to have new perspectives for the study of Tingley's problem in the setting of operator algebras. In Section 3, we showed that Tingley's problem between unital C∗-algebras has a positive answer if and only if the following problem has a positive answer. Problem 6.1. Let A be a unital C∗-algebra and Φ : S(A) → S(A) be a surjective isometry. Suppose that Φ(x) = x for every x ∈ S(A) ∩ A−1. Is Φ equal to the identity mapping on S(A)? 18 M. MORI In Section 5, we used the projection lattice of a von Neumann algebra. Such a discussion is generally impossible in the cases of general unital C∗-algebras. So we propose the following problem. Problem 6.2 (Tingley's problem for the spaces of self-adjoint operators in uni- tal C∗-algebras). Let A, B be unital C∗-algebras and T : S(Asa) → S(Bsa) be a surjective isometry. Does T admit an extension to a linear surjective isometry Since T (1) is isolated in E(Bsa), we obtain that T (1) is a central self-adjoint unitary in B. By the theorem due to Kadison in the paper in 1952 [11, Theorem eT : Asa → Bsa? 2], if such a eT exists, then the mapping T (1)−1eT (·) is the restriction of a Jordan ∗-isomorphism from A onto B. As another direction, we present the problem below. Problem 6.3 (Tingley's problem for noncommutative Lp-spaces). Let 1 < p < ∞, p 6= 2, M, N be von Neumann algebras and T : S(Lp(M )) → S(Lp(N )) be a surjective isometry between the unit spheres of (Haagerup) noncommutative Lp- spaces (with respect to fixed normal semifinite faithful weights). Does T admit an extension to a real linear surjective isometry eT : Lp(M ) → Lp(N )? See [19] for information about noncommutative Lp-spaces. We mention that the noncommutative Lp-space Lp(M ) can be considered as the complex interpolation space between L1(M ) = M∗ and L∞(M ) = M . Noncommutative Lp-spaces are strictly convex, so it is completely impossible to apply the facial method which is wholly used in this paper. However, as the first step to challenge this problem, one may make use of the following property. Proposition 6.4 (Equality condition of the noncommutative Clarkson inequality, see [20, Theorem 2.3] for references). Let 1 ≤ p < ∞, p 6= 2, M, N be von Neumann algebras and ξ, η ∈ Lp(M ). Then kξ + ηkp + kξ − ηkp = 2(kξkp + kηkp) ⇐⇒ ξη∗ = 0 = ξ∗η. Therefore, in the setting of Problem 6.3, for ξ, η ∈ S(Lp(M )), ξη∗ = 0 = ξ∗η if and only if T (ξ)T (η)∗ = 0 = T (ξ)∗T (η). See [20] for a result about complex linear surjective isometry between noncommutative Lp-spaces. Note added in proof After the submission of this paper, the author and Ozawa announced several results on Tingleys problem in [17] which contain an affirmative solution of Problem 6.1. Acknowledgements This paper is written for master's thesis of the author. The author appreciates Yasuyuki Kawahigashi who is the advisor of the author. This work was supported by Leading Graduate Course for Frontiers of Mathemat- ical Sciences and Physics, MEXT, Japan. References [1] C.A. Akemann and G.K. Pedersen, Facial structure in operator algebra theory, Proc. London Math. Soc. (3) 64 (1992), 418 -- 448. [2] L. Cheng and Y. Dong, On a generalized Mazur-Ulam question: extension of isometries between unit spheres of Banach spaces, J. Math. Anal. Appl. 377 (2011), 464 -- 470. TINGLEY'S PROBLEM FOR OPERATOR ALGEBRAS 19 [3] G. Ding, On isometric extension problem between two unit spheres, Sci. China Ser. A 52 (2009), 2069 -- 2083. [4] H.A. Dye, On the geometry of projections in certain operator algebras, Ann. of Math. (2) 61 (1955), 73 -- 89. [5] C.M. Edwards and G.T. Ruttimann, On the facial structure of the unit balls in a JBW∗-triple and its predual, J. London Math. Soc. (2) 38 (1988), 317 -- 332. [6] F.J. Fern´andez-Polo, J.J. Garc´es, A.M. Peralta, I. Villanueva, Tingley's problem for spaces of trace class operators, Linear Algebra Appl. 529 (2017), 294 -- 323. [7] F.J. Fern´andez-Polo, J. Mart´ınez, A.M. Peralta, Surjective isometries between real JB∗- triples, Math. Proc. Cambridge Philos. Soc. 137 (2004), 709 -- 723. [8] F.J. Fern´andez-Polo and A.M. Peralta, On the extension of isometries between the unit spheres of von Neumann algebras, J. Math. Anal. Appl. 466 (2018), no. 1, 127 -- 143. [9] O. Hatori and L. Moln´ar, Isometries of the unitary groups and Thompson isometries of the spaces of invertible positive elements in C ∗-algebras, J. Math. Anal. Appl. 409 (2014), no. 1, 158 -- 167. [10] R.V. Kadison, Isometries of operator algebras, Ann. of Math. (2) 54 (1951), 325 -- 338. [11] R.V. Kadison, A generalized Schwarz inequality and algebraic invariants for operator algebras, Ann. of Math. (2) 56 (1952), 494 -- 503. [12] R.V. Kadison, Transformations of states in operator theory and dynamics, Topology 3 (1965), suppl. 2, 177 -- 198. [13] R.V. Kadison and J.R. Ringrose, "Fundamentals of the theory of operator algebras. Vol. II", Academic Press, Inc., Orlando, FL (1986). [14] C.-W. Leung, C.-K. Ng, N.-C. Wong, The positive contractive part of a noncommutative Lp-space is a complete Jordan invariant, Linear Algebra Appl. 519 (2017), 102 -- 110. [15] P. Mankiewicz, On extension of isometries in normed linear spaces, Bull. Acad. Polon. Sci., S´er. Sci. Math. Astronom. Phys. 20 (1972), 367 -- 371. [16] L. Moln´ar and W. Timmermann, Isometries of quantum states, J. Phys. A 36 (2003), no. 1, 267 -- 273. [17] M. Mori and N. Ozawa, Mankiewicz's Theorem and the Mazur -- Ulam property for C∗- algebras, preprint, arXiv:1804.10674. [18] A.M. Peralta, A survey on Tingley's problem for operator algebras, Acta Sci. Math. (Szeged) 84 (2018), no. 1 -- 2, 81 -- 123. [19] G. Pisier and Q. Xu, Non-commutative Lp-spaces, in: "Handbook of the geometry of Banach spaces, Vol. 2", North-Holland, Amsterdam (2003), 1459 -- 1517. [20] D. Sherman, Noncommutative Lp structure encodes exactly Jordan structure. J. Funct. Anal. 221 (2005), no. 1, 150 -- 166. [21] R. Tanaka, A further property of spherical isometries, Bull. Aust. Math. Soc. 90 (2014), 304 -- 310. [22] R. Tanaka, The solution of Tingley's problem for the operator norm unit sphere of complex n × n matrices, Linear Algebra Appl. 494 (2016), 274 -- 285. [23] R. Tanaka, Tingley's problem on finite von Neumann algebras, J. Math. Anal. Appl. 451 (2017), 319 -- 326. [24] D. Tingley, Isometries of the unit sphere, Geom. Dedicata 22 (1987), 371 -- 378. [25] A. Vogt, Maps which preserve equality of distance, Studia Math. 45 (1973), 43 -- 48. Graduate School of Mathematical Sciences, the University of Tokyo, Komaba, Tokyo, 153-8914, Japan. E-mail address: [email protected]
1702.07629
1
1702
2017-02-24T15:34:39
2-Local derivations on matrix algebras and algebras of measurable operators
[ "math.OA" ]
Let \(\mathcal{A}\) be a unital Banach algebra such that any Jordan derivation from \(\mathcal{A}\) into any \(\mathcal{A}\)-bimodule \(\mathcal{M}\) is a derivation. We prove that any 2-local derivation from the algebra $M_n(\mathcal{A})$ into $M_n(\mathcal{M})$ $(n\geq 3)$ is a derivation. We apply this result to show that any 2-local derivation on the algebra of locally measurable operators affiliated with a von Neumann algebra without direct abelian summands is a derivation.
math.OA
math
2-LOCAL DERIVATIONS ON MATRIX ALGEBRAS AND ALGEBRAS OF MEASURABLE OPERATORS SHAVKAT AYUPOV1, KARIMBERGEN KUDAYBERGENOV2 and AMIR ALAUADINOV3 Abstract. Let A be a unital Banach algebra such that any Jordan derivation from A into any A-bimodule M is a derivation. We prove that any 2-local derivation from the algebra Mn(A) into Mn(M) (n ≥ 3) is a derivation. We apply this result to show that any 2-local derivation on the algebra of locally measurable operators affiliated with a von Neumann algebra without direct abelian summands is a derivation. 1. Introduction Let A be an associative algebra over C the field of complex numbers and let M be an A-bimodule. A linear map D from A to M is called a derivation if D(xy) = D(x)y + xD(y) for all x, y ∈ A. If it satisfies a weaker condition D(x2) = D(x)x + xD(x) for every x ∈ A then it is called a Jordan derivation. It is easy to verify that each element a ∈ M implements a derivation Da from A into M by Da(x) = ax − xa, x ∈ A. Such derivations Da are called inner derivations. In 1990, Kadison [12] and Larson and Sourour [15] independently introduced the concept of local derivation. A linear map ∆ : A → M is called a local deriva- tion if for every x ∈ A there exists a derivation Dx (depending on x) such that ∆(x) = Dx(x). It would be interesting to consider under which conditions local derivations automatically become derivations. Many partial results have been done in this problem. In [12] Kadison shows that every norm-continuous local derivation from a von Neumann algebra M into a dual M-bimodule is a deriva- tion. In [11] Johnson extends Kadison's result and proves every local derivation from a C ∗-algebra A into any Banach A-bimodule is a derivation. Similar problems for local derivations on algebras of measurable operators S(M) and locally measurable operators LS(M), affiliated with a von Neumann algebra M, have been considered in [4] and [9]. Namely, it was proved that if M is a von Neumann algebra without abelian direct summand then every local deriva- tion on LS(M) is a derivation. Moreover, for abelian von Neumann algebras M necessary and sufficient condition are given in [5] for S(M) = LS(M) to admit local derivations which are not derivations (see for details the survey [4, Section 5]). Date: September 23, 2018. 1991 Mathematics Subject Classification. Primary 46L57; 47B47; Secondary 47C15; 16W25. Key words and phrases. matrix algebra; derivation; inner derivation; 2-local derivation; mea- surable operator. 1 2 SH. A. AYUPOV, K. K. KUDAYBERGENOV and A.K. ALAUADINOV In 1997, Semrl [17] initiated the study of so-called 2-local derivations and 2- local automorphisms on algebras. Namely, he described such maps on the algebra B(H) of all bounded linear operators on an infinite dimensional separable Hilbert space H. In the above notations, map ∆ : A → M (not necessarily linear) is called a 2-local derivation if, for every x, y ∈ A, there exists a derivation Dx,y : A → M such that Dx,y(x) = ∆(x) and Dx,y(y) = ∆(y). Afterwards local derivations and 2-local derivations have been investigated by many authors on different algebras and many results have been obtained in [1 -- 3, 5, 12, 14, 17]. Recall that an algebra A is called a regular (in the sense of von Neumann) if for each a ∈ A there exists b ∈ A such that a = aba. Let Mn(A) be the algebra of all n × n matrices over a unital commutative regular algebra A. In [5], we prove that every 2-local derivation on Mn(A), n ≥ 2, is a derivation. We applied this result to a description of 2-local derivations on the algebras of measurable operators S(M) and locally measurable operators LS(M) affiliated with a type I finite von Neumann algebra M. Further this result was extended to type I∞ von Neumann algebras: it was proved that in this case every 2-local derivations on the algebra of locally measurable operators is a derivation (see [4, Theorem 6,7]). Moreover in [5] we also gave necessary and sufficient conditions for a commutative regular algebra, in particular for the algebra S(M) of measurable operators affiliated with an abelian von Neumann algebra M, to admit 2-local derivations which are not derivations. In [3] we considered a unital semi-prime Banach algebra A with the inner derivation property and proved that any 2-local derivation on the algebra M2n(A), n ≥ 2, is a derivation. We have applied this result to AW ∗-algebras and proved that any 2-local derivation on an arbitrary AW ∗-algebra is a derivation. In [10], W. Huang, J. Li and W. Qian, have characterized derivations and 2-local derivations from Mn(A) into Mn(M), n ≥ 2, where A is a unital algebra over C and M is a unital A-bimodule. They considered a unital Banach algebra such that any Jordan derivation from the algebra A into any A-bimodule M is an inner derivation and proved that any 2-local derivation from the algebra Mn(A) into Mn(M) (n ≥ 3) is a derivation, when A is commutative and commutes with M. In the present paper we shall consider matrix algebras over unital (non commu- tative in general) Banach algebras and describe 2-local derivations from Mn(A) into Mn(M), where A is a unital Banach algebra such that any Jordan derivation from the algebra A into any A-bimodule M is a derivation. The main result of Section 2 asserts that under the above conditions every 2-local derivation from the algebra Mn(A) into Mn(M) (n ≥ 3) is a derivation. In Section 3, we apply the main result of the previous section to algebras of locally measurable operators affiliated with von Neumann algebras. Namely, we extend all above mentioned results from [3 -- 5, 10] and prove that for an arbi- trary von Neumann algebra M without abelian direct summands every 2-local derivation on each subalgebra A of the algebra LS(M), such that M ⊆ A, is a 2-LOCAL DERIVATIONS ON MATRIX ALGEBRAS 3 derivation. A similar result for local derivation is obtained in [9, Theorem 1] (see also [4, Theorem 5.5]). 2. 2-local derivations on matrix algebras If ∆ : A → M is a 2-local derivation, then from the definition it easily follows that ∆ is homogenous. At the same time, ∆(x2) = ∆(x)x + x∆(x) for each x ∈ A. This means that additive (and hence, linear) 2-local derivation is a Jordan derivation. In [8] Bresar suggested various conditions on an algebra A under which any Jordan derivation from A into any A-bimodule M is a derivation. In the present paper we shall consider algebras with the following property: (J): any Jordan derivation from the algebra A into any A-bimodule M is a derivation. Therefore, in the case of algebras with the property (J) in order to prove that a 2-local derivation ∆ : A → M is a derivation it is sufficient to prove that ∆ : A → M is additive. Throughout this paper, A is a unital Banach algebra over C, M is an A- bimodule with 1x = x1 = x for all x ∈ M, where 1 is the unit element of A. The following theorem is the main result of this section. Theorem 2.1. Let A be a unital Banach algebra with the property (J), M be a unital A-bimodule and let Mn(A) be the algebra of all n × n-matrices over A, where n ≥ 3. Then any 2-local derivation ∆ from Mn(A) into Mn(M) is a derivation. The proof of Theorem 2.1 consists of two steps. In the first step we shall show additivity of ∆ on the subalgebra of diagonal matrices from Mn(A). Let {ei,j}n i,j=1 be the system of matrix units in Mn(A). For x ∈ Mn(A) by xi,j we denote the (i, j)-entry of x, where 1 ≤ i, j ≤ n. We shall, if necessary, identify this element with the matrix from Mn(A) whose (i, j)-entry is xi,j, other entries are zero, i.e. xi,j = ei,ixej,j. Each element x ∈ Mn(A) has the form x = n X i,j=1 xijeij, xij ∈ A, i, j ∈ 1, n. Let δ : A → M be a derivation. Setting δ(x) = n X i,j=1 δ(xij)eij, xij ∈ A, i, j ∈ 1, n (2.1) we obtain a well-defined linear operator δ from Mn(A) into Mn(M). Moreover δ is a derivation from Mn(A) into Mn(M). 4 SH. A. AYUPOV, K. K. KUDAYBERGENOV and A.K. ALAUADINOV It is known [10, Theorem 2.1] that every derivation D from Mn(A) into Mn(M) can be represented as a sum D = ad(a) + δ, (2.2) where ad(a) is an inner derivation implemented by an element a ∈ Mn(M), while δ is the derivation of the form (2.1) generated by a derivation δ from A into M. Consider the following two matrices: u = n X i=1 1 2i ei,i, v = n X i=2 ei−1,i. (2.3) It is easy to see that an element x ∈ Mn(M) commutes with u if and only if it is diagonal, and if an element a ∈ Mn(M) commutes with v, then a is of the form a =   a1 a2 a3 0 a1 a2 0 0 a1 ... ... ... 0 0 . . . 0 0 . . . . . . an . . . . an−1 . . . . an−2 . ... ... ... a2 . a1 0 a1 .   . (2.4) A result, similar to the following one, was proved in [5, Lemma 4.4] for matrix algebras over commutative regular algebras. Further in Lemmata 2.2 -- 2.5 we assume that n ≥ 2. Lemma 2.2. For every 2-local derivation ∆ from Mn(A) into Mn(M) there exists a derivation D such that ∆sp{ei,j }n is the linear span of the set {ei,j}n i,j=1 = Dsp{ei,j}n i,j=1 , where sp{ei,j}n i,j=1 i,j=1. Proof. Take a derivation D from Mn(A) into Mn(M) such that ∆(u) = D(u), ∆(v) = D(v), where u, v are the elements from (2.3). Replacing ∆ by ∆ − D, if necessary, we can assume that ∆(u) = ∆(v) = 0. Let i, j ∈ 1, n. Take a derivation D = ad(h) + δ of the form (2.2) such that ∆(ei,j) = [h, ei,j] + δ(eij), ∆(u) = [h, u] + δ(u). Since ∆(u) = 0 and δ(u) = 0, it follows that [h, u] = 0, and therefore h has a n diagonal form, i.e. h = hses,s, hs ∈ A, s ∈ 1, n. Ps=1 In the same way, but starting with the element v instead of u, we obtain where b has the form (2.4), depending on ei,j. So ∆(ei,j) = bei,j − ei,jb, Since ∆(ei,j) = hei,j − ei,jh = bei,j − ei,jb. hei,j − ei,jh = (hi − hj)ei,j 2-LOCAL DERIVATIONS ON MATRIX ALGEBRAS 5 and [bei,j − ei,jb]i,j = 0, it follows that ∆(ei,j) = 0. Now let us take a matrix x = n Pi,j=1 λi,jei,j ∈ Mn(C). Then ei,j∆(x)ei,j = ei,jDei,j ,x(x)ei,j = = Dei,j ,x(ei,jxei,j) − Dei,j ,x(ei,j)xei,j − ei,jxDei,j ,x(ei,j) = = Dei,j ,x(λj,iei,j) − ∆(ei,j)xei,j − ei,jx∆(ei,j) = = λj,iDei,j ,x(ei,j) − 0 − 0 = λj,i∆(ei,j) = 0, i.e. ei,j∆(x)ei,j = 0 for all i, j ∈ 1, n. This means that ∆(x) = 0. The proof is complete. (cid:3) Further in Lemmata 2.3 -- 2.8 we assume that ∆ is a 2-local derivation from Mn(A) into Mn(M) such that ∆sp{ei,j }n i,j=1 = 0. Let ∆i,j be the restriction of ∆ onto Ai,j = ei,iMn(A)ej,j, where 1 ≤ i, j ≤ n. Lemma 2.3. ∆i,j maps Ai,j into itself. Proof. Let us show that ∆i,j(x) = ei,i∆(x)ej,j (2.5) for all x ∈ Ai,j. Take x = xi,j ∈ Ai,j, and consider a derivation D = ad(h) + δ of the form (2.2) such that ∆(x) = [h, x] + δ(x), ∆(u) = [h, u] + δ(u), where u is the element from (2.3). Since ∆(u) = 0 and δ(u) = 0, it follows that [h, u] = 0, and therefore h has a diagonal form. Then ∆(x) = (hi − hj)eij + δ(xij)eij. This means that ∆(x) ∈ Ai,j. The proof is complete. (cid:3) Lemma 2.4. Let x = n Pi=1 xi,i be a diagonal matrix. Then ek,k∆(x)ek,k = ∆(xk,k) (2.6) for all k ∈ 1, n. Proof. Take a derivation D = ad(a) + δ of the form (2.2) such that ∆(x) = [a, x] + δ(x) and ∆(xk,k) = [a, xk,k] + δ(xkk). Using equality (2.5), we obtain that ∆(xk,k) = ek,k∆(xk,k)ek,k = ek,k[a, xk,k]ek,k + ek,kδ(xk,k)ek,k = = [ak,k, xk,k] + δ(xk,k). Since x is a diagonal matrix, we get ek,k∆(x)ek,k = ek,k[a, x]ek,k + ekkδ(x)ek,k = [ak,k, xk,k] + δ(xk,k). Thus ek,k∆(x)ek,k = ∆(xk,k). The proof is complete. (cid:3) 6 SH. A. AYUPOV, K. K. KUDAYBERGENOV and A.K. ALAUADINOV Lemma 2.5. Let x = xi,i ∈ Ai,i. Then ej,i∆(x)ei,j = ∆(ej,ixei,j) (2.7) for every j ∈ {1, · · · , n}. Proof. For i = j we have already proved (see Lemma 2.4). Suppose that i 6= j. For an arbitrary element x = xi,i ∈ Ai,i , consider y = x + ej,ixei,j ∈ Ai,i + Aj,j. Take a derivation D = ad(a) + δ such that ∆(y) = [a, y] + δ(y) and ∆(v) = [a, v] + δ(v), where v is the element from (2.3). Since ∆(v) = 0 and δ(v) = 0, it follows that a has the form (2.4). By Lemma 2.4 we obtain that ej,i∆(x)ei,j = ej,iei,i∆(y)ei,iei,j = ej,i[a, y]ei,j + ej,iδ(y)ei,j = = ([a1, x] + δ(x)) ej,j, ∆(ej,ixei,j) = ej,j∆(y)ej,j = ej,j[a, y]ej,j + ej,jδ(y)ej,j = = ej,j[a, x + ej,ixei,j]ej,j + ej,jδ(x)ej,j = ([a1, x] + δ(x)) ej,j. The proof is complete. (cid:3) Further in Lemmata 2.6 -- 2.13 we assume that n ≥ 3. Lemma 2.6. ∆i,i is additive for all i ∈ 1, n. Proof. Let i ∈ 1, n. Since n ≥ 3, we can take different numbers k, s such that (k − i)(s − i) 6= 0. For arbitrary x, y ∈ Ai,i consider the diagonal element z ∈ Ai,i + Ak,k + As,s such that zi,i = x + y, zk,k = x, zs,s = y. Take a derivation D = ad(a) + δ such that ∆(z) = [a, z] + δ(z) and ∆(v) = [a, v] + δ(v), where v is the element from (2.3). Since ∆(v) = 0 and δ(v) = 0, it follows that a has the form (2.4). Using Lemmata 2.4 and 2.5 we obtain that ∆i,i(x + y) ∆i,i(x) ∆i,i(y) (2.6) = ei,i∆(z)ei,i = ei,i[a, z]ei,i + ei,iδ(z)ei,i = = ([a1, x + y] + δ(x + y)) ei,i, (2.7) = ei,k∆(ek,ixei,k)ek,i = ei,k[a, z]ek,i + ei,kδ(z)ek,i = ([a1, x] + δ(x)) ei,i, (2.7) = ei,s∆(es,iyei,s)es,i = ei,s[a, z]es,i + ei,sδ(z)es,i = ([a1, y] + δ(y)) ei,i. (2.6) = ei,kek,k∆(z)ek,kek,i = (2.6) = ei,ses,s∆(z)es,ses,i = Hence The proof is complete. ∆i,i(x + y) = ∆i,i(x) + ∆i,i(y). (cid:3) As it was mentioned in the beginning of the section any additive 2-local deriva- ∼= A has the property (J), Lemma 2.6 tion is a Jordan derivation. Since Ai,i implies the following result. 2-LOCAL DERIVATIONS ON MATRIX ALGEBRAS 7 Lemma 2.7. ∆i,i is a derivation for all i ∈ 1, n. Denote by Dn(A) the set of all diagonal matrices from Mn(A), i.e. the set of all matrices of the following form x =   0 x1 0 x2 ... ... 0 0 0 0 0 . . . 0 . . . ... ... . . . xn−1 . . . 0 0 0 ... 0 xn   . Let us consider a derivation ∆1,1 of the form (2.1). By Lemmata 2.4 and 2.5 we obtain that Lemma 2.8. ∆Dn(A) = ∆1,1Dn(A) and ∆1,1sp{ei,j}n i,j=1 = 0. Now we are in position to pass to the second step of our proof. In this step we show that if a 2-local derivation ∆ satisfies the following conditions ∆Dn(A) ≡ 0 and ∆sp{ei,j}n i,j=1 ≡ 0, then it is identically equal to zero. Below in the five Lemmata we shall consider 2-local derivations which satisfy the latter equalities. We denote by e the unit of the algebra A. Lemma 2.9. Let x ∈ Mn(A). Then ∆(x)k,k = 0 for all k ∈ 1, n. Proof. Let x ∈ Mn(A), and fix k ∈ 1, n. Since ∆ is homogeneous, we can assume that kxk,kk < 1, where k · k is the norm on A. Take a diagonal element y in Mn(A) with yk,k = e + xk,k and yi,i = 0 otherwise. Since kxk,kk < 1, it follows that e + xk,k is invertible in A. Take a derivation D = ad(a) + δ of the form (2.2) such that ∆(x) = [a, x] + δ(x), ∆(y) = [a, y] + δ(y). Since y ∈ Dn(A) we have that 0 = ∆(y) = [a, y] + δ(y), and therefore 0 = ∆(y)k,k = ak,k(e + xk,k) − (e + xk,k)ak,k + δ(e + xk,k) = 0, 0 = ∆(y)i,k = ai,k(e + xk,k) = 0, 0 = ∆(y)k,i = −(e + xk,k)ak,i = 0 for all i 6= k. Thus and ak,kxk,k − xk,kak,k + δ(xk,k) = 0 for all i 6= k. The above equalities imply that ai,k = ak,i = 0 ∆(x)k,k = ak,kxk,k − xk,kak,k + δ(xk,k) = ∆(y)k,k = 0. The proof is complete. (cid:3) Lemma 2.10. Let x be a matrix with xk,s = e. Then ∆(x)k,s = 0. 8 SH. A. AYUPOV, K. K. KUDAYBERGENOV and A.K. ALAUADINOV Proof. We have es,k∆(x)es,k = es,kDes,k,x(x)es,k = = Des,k,x(es,kxes,k) − Des,k,x(es,k)xes,k − es,kxDes,k,x(es,k) = = Des,k,x(es,k) − ∆(es,k)xes,k − es,kx∆(es,k) = = ∆(es,k) − 0 − 0 = 0. Thus ek,k∆(x)es,s = ek,ses,k∆(x)es,kek,s = 0. This means that ∆(x)k,s = 0. The proof is complete. (cid:3) Lemma 2.11. Let k, s be numbers such that k 6= s and let x be a matrix with xk,s = e. Then ∆(x)s,k = 0. Proof. Take a diagonal element y such that yk,k = xs,k and yi,i = λie otherwise, where λi (i 6= k) are distinct numbers with λi > kxs,kk. Take a derivation D = ad(a) + δ such that ∆(x) = [a, x] + δ(x) and ∆(y) = [a, y] + δ(y). Then 0 = ∆(y)ij = λjai,j − λiai,j = ai,j(λj − λi), i 6= j, (i − k)(j − k) 6= 0, 0 = ∆(y)i,k = ai,kyk,k − λiai,k = ai,k(xs,k − λi), i 6= k, 0 = ∆(y)k,j = ak,jλj − ykkakj = (λj − xs,k)ak,j, j 6= k. Thus ai,j = 0 for all i 6= j, i.e. a is a diagonal element. Since 0 = ∆(x)ks = akk − ass, it follows that ak,k = as,s. Finally, ∆(x)s,k = as,sxs,k − xs,kak,k + δ(xs,k) = = ak,kxs,k − xs,kak,k + δ(yk,k) = ∆(y)k,k = 0. The proof is complete. (cid:3) Lemma 2.12. Let k 6= s and let x, y be matrices with xi,j = yi,j for all (i, j) 6= (s, k). Then ∆(x)k,s = ∆(y)k,s. Proof. Take a derivation D = ad(a) + δ such that ∆(x) = [a, x] + δ(x) and ∆(y) = [a, y] + δ(y). Then ∆(x)k,s = = n X j=1 n X j=1 (ak,jxj,s − xk,jaj,s) + δ(xks) = (ak,jyj,s − yk,jaj,s) + δ(yks) = ∆(y)k,s. The proof is complete. Lemma 2.13. Let k 6= s. Then ∆(x)k,s = 0. (cid:3) 2-LOCAL DERIVATIONS ON MATRIX ALGEBRAS 9 Proof. Take a matrix y with ys,k = e and yi,j = xi,j otherwise. By Lemma 2.11 we have that ∆(y)k,s = 0. Further Lemma 2.12 implies that The proof is complete. (cid:3) ∆(x)k,s = ∆(y)k,s = 0. Now we are in position to prove Theorem 2.1. Proof of Theorem 2.1. Let ∆ be a 2-local derivation from Mn(A) into Mn(M), i,j=1 = . Consider a 2-local derivation Θ = ∆ − D. Since Θ is equal to zero i,j=1, by Lemma 2.8 we obtain that ΘDn(A) = Θ11Dn(A), where Θ11 is where n ≥ 3. By Lemma 2.2 there exists a derivation D such that ∆sp{ei,j }n Dsp{ei,j }n on sp{ei,j}n the derivation defined by (2.1). As in Lemma 2.8 we have that i,j=1 (cid:0)Θ − Θ11(cid:1) sp{ei,j}n i,j=1 ≡ 0 and (cid:0)Θ − Θ11(cid:1) Dn(A) ≡ 0. Now for an arbitrary element x ∈ Mn(A), by Lemmata 2.9 and 2.13 we obtain that (cid:0)Θ − Θ11(cid:1) (x)k,s = 0 for all k, s. Thus (cid:0)Θ − Θ11(cid:1) (x) = 0, i.e., Θ = Θ11. So, ∆ = Θ11 + D is a derivation. The proof is complete. (cid:3) 3. An application to 2-local derivations on algebras of locally measurable operators In this section we apply Theorem 2.1 to the description of 2-local derivations on the algebra of locally measurable operators affiliated with a von Neumann algebra and on its subalgebras. In [8, Corollary 3.11] it was proved that if an associative algebra (ring) A con- tains a noncommutative simple subalgebra (subring) A0 which contains the unit of A, then every Jordan derivation from A into any A-bimodule is a derivation, i.e. A satisfies the property (J). In particular, if there exists a subalgebra A0 of A which is isomorphic to Mn(C) (n ≥ 2) and contains the unit of A, then A has the property (J). Let M be a von Neumann algebra and denote by S(M) the algebra of all mea- surable operators and by LS(M) the algebra of all locally measurable operators affiliated with M (see for example [16, 18]). Theorem 3.1. Let M be an arbitrary von Neumann algebra without abelian di- rect summands and let LS(M) be the algebra of all locally measurable operators affiliated with M. Then any 2-local derivation ∆ from M into LS(M) is a deriva- tion. Proof. Let z be a central projection in M. Since D(z) = 0 for an arbitrary deriva- tion D, it is clear that ∆(z) = 0 for any 2-local derivation ∆ from M into LS(M). Take x ∈ M and let D be a derivation from M into LS(M) such that ∆(zx) = D(zx), ∆(x) = D(x). Then we have ∆(zx) = D(zx) = D(z)x + zD(x) = z∆(x). This means that every 2-local derivation ∆ maps zM into zLS(M) ∼= LS(zM) for each central projection z ∈ M. So, we may consider the restriction of ∆ onto zM. Since an arbitrary von Neumann algebra without abelian direct summands 10 SH. A. AYUPOV, K. K. KUDAYBERGENOV and A.K. ALAUADINOV can be decomposed along a central projection into the direct sum of von Neu- mann algebras of type In, n ≥ 2, type I∞, type II and type III, we may consider these cases separately. If M is a von Neumann algebra of type In, n ≥ 2, [10, Corollary 3.12] implies that any 2-local derivation from M into LS(M) ≡ S(M) is a derivation. Let the von Neumann algebra M have one of the types I∞, II or III. Then the halving Lemma [13, Lemma 6.3.3] for type I∞-algebras and [13, Lemma 6.5.6] for type II or III algebras, imply that the unit of the algebra M can be represented as a sum of mutually equivalent orthogonal projections e1, e2, e3 from M. Then the map x 7→ eixej defines an isomorphism between the algebra M and the 3 Pi,j=1 matrix algebra M3(A), where A = e1,1M e1,1. Further, the algebra LS(M) is isomorphic to the algebra M3(LS(A)). Moreover, the algebra A has same type as the algebra M, and therefore contains a subalgebra isomorphic to M3(C). This means that the algebra A satisfies the property (J). Therefore Theorem 2.1 implies that any 2-local derivation from M into LS(M) is a derivation. The proof is complete. (cid:3) Taking into account that any derivation on an abelian von Neumann algebra is trivial, Theorem 3.1 implies the following result (cf. [2, Theorem 2.1] and [3, Theorem 3.1]). Corollary 3.2. Let M be an arbitrary von Neumann algebra. Then any 2-local derivation ∆ on M is a derivation. For each x ∈ LS(M) set s(x) = l(x) ∨ r(x), where l(x) is the left and r(x) is the right support of x. Lemma 3.3. Let B be a subalgebra of LS(M) such that M ⊆ B and let ∆ : B → LS(M) be a 2-local derivation such that ∆M ≡ 0. Then ∆ ≡ 0. ∞ R0 Proof. Let us first take an arbitrary element x ∈ B ∩ S(M). Let x = λ deλ be the spectral resolution of x. Since x ∈ S(M), it follows that e⊥ n is a finite projection for a sufficiently large n. Take a derivation Dx,xen such that ∆(x) = Dx,xen(x) and ∆(xen) = Dx,xen(xen), n ∈ N. Since xen ∈ M, it follows that ∆(xen) = 0 for all n ∈ N. We have ∆(x) = ∆(x) − ∆(xen) = Dx,xen(x) − Dx,xen(xen) = = Dx,xen(x − xen) = Dx,xen(xe⊥ n ). Let D be a dimension function on the lattice P (M) of all projections from M (see [18]). Using [6, Lemma 4.3] we obtain that D(s(∆(x))) = D(s(Dx,xen(xe⊥ n ))) ≤ 3D(s(xe⊥ n )) = 3D(l(xe⊥ n ) ∨ r(xe⊥ n )) ≤ ≤ 3D (cid:0)l(xe⊥ n )(cid:1) + 3D(r(xe⊥ n )) ≤ 6D(e⊥ n ) ↓ 0, and therefore ∆(x) = 0. Now let take an element x ∈ B. By the definition of locally measurable operator there exists a sequence {zn} of central projections in M such that zn ↑ 1 and 2-LOCAL DERIVATIONS ON MATRIX ALGEBRAS 11 xzn ∈ S(M) for all n ∈ N (see [16]). Taking into account the previous case we obtain that zn∆(x) = znDx,znx(x) = Dx,znx(znx) − Dx,znx(zn)x = = Dx,znx(znx) = ∆(znx) = 0, i.e., zn∆(x) = 0 for all n ∈ N. Hence ∆(x) = 0. The proof is complete. (cid:3) Theorem 3.4. (cf. [4, Theorem 5.5]). Let M be an arbitrary von Neumann algebra without abelian direct summands and let B be a subalgebra of LS(M) such that M ⊆ B. Then any 2-local derivation ∆ on B is a derivation. Proof. By Theorem 3.1 the restriction ∆M of ∆, is a derivation from M into LS(M). By [6, Theorem 4.8] the derivation ∆M can be extended to a derivation from B into LS(M), which we denote by D. Since the 2-local derivation ∆ − D is equal to zero on M, Lemma 3.3 implies that ∆ ≡ D. The proof is complete. (cid:3) Remark 3.5. As it was mentioned in the introduction, the paper [5] gives neces- sary and sufficient conditions on a commutative regular algebra to admit 2-local derivations which are not derivations. In particular, for an arbitrary abelian von Neumann algebra M with a non atomic lattice of projections P (M) the algebras S(M) and LS(M) always admit a 2-local derivation which is not a derivation. A complete description of derivations on the algebra LS(M) for type I von Neumann algebras M is given in [4, Section 3]). Moreover, for general von Neu- mann algebras every derivation on the algebra LS(M) is inner, provided that M is a properly infinite von Neumann algebra [4, 7]. But for type II1 von Neumann algebra M description of structure of derivations on the algebra S(M) ≡ LS(M) is still an open problem (see [4]). In this connection it should be noted that Theorem 3.4 is one of the first results on 2-local derivations without information on the general form of derivations on these algebras. References [1] R. Alizadeh, M. J. Bitarafan, Local derivations of full matrix rings, Acta Mathematica Hungarica, 145 (2015) 433 -- 439. [2] Sh. A. Ayupov and K. K. Kudaybergenov, 2-Local derivations on von Neumann algebras, Positivity, 19 (2015) 445 -- 455. [3] Sh. A. Ayupov and K. K. Kudaybergenov, 2-Local derivations on matrix algebras over semi-prime Banach algebras and on AW ∗-algebras, Journal of Physics: Conference Series, 697 (2016) 1 -- 10. [4] Sh. A. Ayupov, K. K. Kudaybergenov, Derivations, local and 2-local derivations on algebras of measurable operators, in Topics in Functional Analysis and Algebra, Contemporary Mathematics, vol. 672, Amer. Math. Soc., Providence, RI, 2016, pp. 51-72. [5] Sh. A. Ayupov, K. K. Kudaybergenov, A. K. Alauadinov, 2-Local derivations on matrix algebras over commutative regular algebras, Linear Alg. Appl. 439 (2013) 1294 -- 1311. [6] A. F. Ber, V. I. Chilin, F. A. Sukochev, Continuity of derivations of algebras of locally measurable operators, Integral Equations and Operator Theory, 75 4 (2013) 527 -- 557. [7] A. F. Ber, V. I. Chilin, F. A. Sukochev, Continuous derivations on algebras of locally measurable operators are inner, Proc. London Math. Soc. 109 (2014) 65 -- 89. [8] M. Bresar, Jordan derivations revisited, Math. Proc. Camb. Phil. Soc. 139, 411 -- 425 (2005). [9] D. Hadwin, J. Li, Q. Li, X. Ma, Local derivations on rings containing a von Neumann algebra and a question of Kadison, arXiv:1311.0030. 12 SH. A. AYUPOV, K. K. KUDAYBERGENOV and A.K. ALAUADINOV [10] W. Huang, J. Li and W. Qian, Derivations and 2-local derivations on matrix algebras over commutative algebras, arXiv:1611.00871v1. [11] B. E. Johnson, Local derivations on C ∗-algebras are derivations, Trans. Amer. Math. Soc., 353 (200) 313 -- 325. [12] R. V. Kadison, Local derivations, J. Algebra, 130 (1990) 494 -- 509. [13] R.V. Kadison, J.R. Ringrose, Fundamentals of the theory of operator algebras, Vol. II, Birkhauser Boston, 1986. [14] S.O. Kim, J.S. Kim, Local automorphisms and derivations on Mn, Proc. Amer. Math. Soc. 132, no. 5, 1389-1392 (2004). [15] D. R. Larson and A. R. Sourour, Local derivations and local automorphisms of B(X), Operator theory: operator algebras and applications, part 2 (Durham,NH, 1988), 187 -- 194, Proc. Sympos. Pure Math. 51, Part 2, Amer.Math.Soc., Providence, RI, (1990). [16] M. Muratov, V. Chilin, *-Algebras of unbounded operators affiliated with a von Neumann algebra, J. Math. Sci., 140 (2007), 445-451. [17] P. Semrl, Local automorphisms and derivations on B(H), Proc. Amer. Math. Soc. 125, 2677-2680 (1997). [18] I.E.Segal, A non-commutative extension of abstract integration, Ann. of Math. 57 (1953), 401-457. 1Institute of Mathematics, National University of Uzbekistan, Dormon yoli 29, 100125 Tashkent, Uzbekistan E-mail address: sh−[email protected] 2Department of Mathematics, Karakalpak State University, Ch. Abdirov 1, Nukus 230113, Uzbekistan E-mail address: [email protected] 3Department of Mathematics, Karakalpak State University, Ch. Abdirov 1, Nukus 230113, Uzbekistan E-mail address: amir−[email protected]
1501.06983
1
1501
2015-01-28T04:12:48
Centre-valued index for Toeplitz operators with noncommuting symbols
[ "math.OA" ]
We consider an action of the real line on a C*-algebra for which there is a centre-valued invariant trace. We define a family of Toeplitz operators with symbols in the original algebra. When the symbol is invertible, the Toeplitz operator is Fredholm in an appropriate sense, and we give a formula for the index using a notion of centre-valued winding number defined using the trace. The results and techniques generalise previous results of the authors for scalar-valued traces.
math.OA
math
CENTRE-VALUED INDEX FOR TOEPLITZ OPERATORS WITH NONCOMMUTING SYMBOLS by John Phillips Department of Mathematics and Statistics University of Victoria Victoria, B.C. V8W 3P4, CANADA and Iain Raeburn Department of Mathematics and Statistics University of Otago, PO Box 56, Dunedin 9054, NEW ZEALAND This research was supported by the Natural Sciences and Engineering Research Council of Canada, The Australian Research Council, and the University of Otago. Abstract. We formulate and prove a "winding number" index theorem for certain "Toeplitz" operators in the same spirit as the Gohberg-Krein Theorem and generalizing previous work of Lesch and others. The "number" in "winding number" is replaced by a self-adjoint oper- ator in a subalgebra Z ⊆ Z(A) of a unital C ∗-algebra, A. We assume that there is a faithful Z-valued trace τ on A which is left invariant under an action α : R → Aut(A) which leaves Z pointwise fixed. If δ is the infinitesimal generator of α and u is an invertible element in dom(δ) then the "winding operator" of u is 1 2πi τ (δ(u)u−1) ∈ Zsa. By a careful choice of rep- resentations we can extend the data (A, Z, τ, α) to a von Neumann setting (A, Z, ¯τ , ¯α) where A = A′′ and Z = Z ′′. Then, A ⊂ A ⊂ A ⋊ R, the von Neumann crossed product, and there is a faithful, dual Z-trace on A ⋊ R. If P is the projection in A ⋊ R corresponding to the non- negative spectrum of the generator of the representation of R in A ⋊ R and π : A → A ⋊ R is the embedding then we define for u ∈ A−1, Tu = P π(u)P and show that it is Fredholm in an appropriate sense and the Z-valued index of Tu is the negative of the winding operator, i.e., −1 2πi τ (δ(u)u−1) ∈ Zsa. In outline the proof follows the proof of the scalar case done previously by the authors. The difficulties arise in making sense of the various constructions when the scalars are replaced by Z in the von Neumann setting. In particular, the construction of the dual Z-trace on A ⋊ R required the nontrivial development of a Z-Hilbert Algebra theory. We show that certain of these Fredholm operators fiber as a "section" of Fredholm operators with scalar-valued index and the centre-valued index fibers as a section of the scalar-valued indices. 1. WINDING OPERATOR Objects of Study: We consider a unital C ∗-algebra, A with a unital C ∗-subalgebra Z of the centre of A; Z(A). We also assume that there exists a faithful, unital, tracial, conditional expectation τ : A → Z (a "faithful Z-trace") and a continuous action α : R → Aut(A) which 1 leaves τ invariant. That is , τ ◦ αt = τ for all t ∈ R. That is, our Objects of Study are 4-tuples (A, Z, τ, α) satisfying these conditions. Under these hypotheses we show that the "winding number theorem" of [PhR] holds. We will often refer to this as a "winding operator". Theorem 1.1. Let (A, Z, τ, α) be a 4-tuple; so that A is a unital C ∗-algebra and Z ⊆ Z(A) is a unital C ∗-subalgebra of the centre of A; τ : A → Z is a faithful, unital, tracial, conditional expectation; and α : R → Aut(A) is a continuous action leaving τ invariant. Let δ be the infinitesimal generator of α. Then, a 7→ 1 2πi τ (δ(a)a−1) : dom(δ)−1 → Zsa is a group homomorphism which is constant on connected components and so extends uniquely to a group homomorphism A−1 → Zsa which is constant on connected components and is 0 on Z −1. We denote this map by windα(a). Proof. It is an easy calculation to see that a 7→ τ (δ(a)a−1) : dom(δ)−1 → (Z, +) is a homomorphism. We next calculate that αt(z) = z for all z ∈ Z and t ∈ R : τ ((αt(z) − z)∗(αt(z) − z)) = · · · = τ (z∗z) − τ (z∗)z − z∗τ (z) + τ (z∗z) = 0. Therefore, αt(z) − z = 0 since τ is faithful. So, Z ⊆ dom(δ) and δ(Z) = {0}. But then for each z ∈ Z −1 we have τ (δ(z)z−1) = 0. Now, for any a ∈ dom(δ), we have Hence, by the Leibnitz rule, for each n ≥ 1 h h→0 αh(a) − a τ (δ(a)) = τ(cid:18)lim 0 = τ (δ(an)) = τ n−1Xk=0 = τ (akδ(a)a(n−1)−k) = n−1Xk=0 = nτ (an−1δ(a)). (cid:19) = lim h→0 1 h τ (αh(a) − a) = 0. akδ(a)a(n−1)−k! τ (an−1δ(a)) n−1Xk=0 Thus, for each a ∈ dom(δ) and each k ≥ 0 we have τ (δ(a)ak) = τ (akδ(a)) = 0. 2 converges in norm. Since δ(1) = 0 we have: Now, if a ∈ dom(δ) and k1 − ak < 1 then a is invertible and a−1 =P∞ τ (δ(a)a−1) = −τ (δ(1− a)a−1) = −τ δ(1 − a) (1 − a)k! = − ∞Xk=0 k=0(1 − a)k which ∞Xk=0 τ (δ(1− a)(1− a)k) = 0. To see that the map is constant on connected components, we use the previous paragraph to show that it is locally constant. So we fix a ∈ dom(δ)−1 and suppose b ∈ dom(δ)−1 where kb − ak < 1/ka−1k. Then, kba−1 − 1k ≤ kb − ak ka−1k < 1 so that 0 = τ (δ(ba−1)(ba−1)−1) = τ (δ(b)b−1) + τ (δ(a−1)a) = τ (δ(b)b−1) − τ (δ(a)a−1) as required. Finally, to see that τ (δ(a)a−1) ∈ iZsa, we observe that since dom(δ) is a ∗-subalgebra of A that a ∈ dom(δ)−1 implies that a∗a ∈ dom(δ)−1 and so t 7→ t1 + (1 − t)a∗a is a path of invertible elements in dom(δ)−1 connecting 1 to a∗a. Hence, τ (δ(a∗a)(a∗a)−1) = τ (δ(1)1) = 0. Since the map is a homomorphism, this implies that τ (δ(a∗)(a∗)−1) = −τ (δ(a)a−1). But, then: [τ (δ(a)a−1)]∗ = τ ((a∗)−1δ(a∗)) = τ (δ(a∗)(a∗)−1) = −τ (δ(a)a−1) as required. Since dom(δ) is a dense ∗-subalgebra of A and A−1 is open, dom(δ)−1 is dense in A−1 and so the map extends uniquely to A−1. (cid:3) Definition 1.2. (Morphism) For i = 1, 2 let (Ai, Zi, τi, αi) be two such 4-tuples where Ai is a unital C ∗-algebra and Zi is a unital C ∗-subalgebras of the centre of Ai, etc. A morphism from (A1, Z1, τ1, α1) to (A2, Z2, τ2, α2) is given by a unital ∗-homomorphism ϕ : A1 → A2 which maps Z1 → Z2 and makes all the appropriate diagrams commute: A1 ϕ / A2 τ1 τ2 Z1 / Z2 ϕ A1 ϕ / A2 α1 t α2 t A1 / A2 ϕ Proposition 1.3. If ϕ : A1 → A2 defines a morphism from (A1, Z1, τ1, α1) to (A2, Z2, τ2, α2) and if a ∈ A−1 1 ∩ (dom(δ1)) then ϕ(a) ∈ A−1 2 ∩ (dom(δ2)) and windα1(a) ∈ (Z1)sa while windα2(ϕ(a)) ∈ (Z2)sa and also ϕ(windα1(a)) = windα2(ϕ(a)).   /   /   /   / Proof. We first show that a ∈ dom(δ1) implies that ϕ(a) ∈ dom(δ2) and that ϕ(δ1(a)) = δ2(ϕ(a)). So if a ∈ dom(δ1) then ϕ(δ1(a)) = ϕ(cid:18)lim t→0 α1 t (a) − a t (cid:19) = lim t→0 ϕ(cid:18)α1 t (a) − a t (cid:19) = lim t→0 α2 t (ϕ(a)) − ϕ(a) . t 3 So the right hand limit exists and defines δ2(ϕ(a)). That is ϕ(δ1(a)) = δ2(ϕ(a)). Now for a ∈ A−1 1 ∩ (dom(δ1)): ϕ(windα1(a)) = ϕ(τ1(δ1(a)a−1)) = 1 2πi τ2(ϕ(δ1(a))ϕ(a)−1)) = 1 2πi 1 2πi = τ2(ϕ(δ1(a)a−1)) 1 2πi τ2(δ2(ϕ(a))ϕ(a)−1)) = windα2(ϕ(a)). (cid:3) 2. EXTENSION to an ENVELOPING von NEUMANN ALGEBRA Key Idea 1. Since the range of our C ∗-algebra trace, Z (an abelian C ∗-algebra), is no longer restricted to being the scalars, the index of our generalized Toeplitz operators will not be scalar-valued either, but will necessarily take values in an abelian von Neumann algebra, say Z, containing Z. Unless, Z is finite-dimensional (a relatively trivial extension of the scalar-valued trace) we will generally have Z 6= Z (if Z is separable but not finite- dimensional we must have Z 6= Z). We want our unital C ∗-algebra, A, to be concretely represented on a Hilbert space, H in such a way that the following nontrivial conditions hold. Let A = A′′ and Z = Z ′′. (1) There exists a necessarily unique faithful, tracial, uw-continuous conditional expecta- tion, ¯τ : A → Z extending τ. We will refer to this as a Z-trace. (2) The continuous action α : R → Aut(A) which leaves τ invariant extends to an ultra- weakly continuous action ¯α : R → Aut(A) which leaves ¯τ invariant. To achieve this we will assume that Z has a faithful state, ω (this is automatically true if Z is separable). We will use the following Proposition to define a concrete representation where these conditions obtain. We emphasize that the extension depends on the choice of the faithful state on Z. However, our notation will not show the dependence on this state. Of course if Z = C the state is unique. If ϕ is a morphism from (A1, Z1, τ1, α1) to (A2, Z2, τ2, α2), we will assume that ϕ carries the faithful state ω1 on Z1 to ω2 on Z2 : that is ω1 = ω2 ◦ ϕ restricted to Z1. Proposition 2.1. Let (A, Z, τ, α) be a 4-tuple and let ω be a faithful state on Z. Then ¯ω := ω ◦ τ is a faithful tracial state on A which is left invariant by the action α. If we 4 let (π,H, ξ0) be the GNS representation of A afforded by ¯ω, with cyclic separating trace vector ξ0, then there is a continuous unitary representation {Ut} of R on H so that (π, U) is covariant for α on A. Then {Ut} implements an uw-continuous extension of α to ¯α acting on A = π(A)′′. Morover, letting Z = π(Z)′′, there exists a unique faithful unital, uw-continuous Z-trace ¯τ : A → Z extending τ, and ¯α leaves ¯τ invariant. Proof. Denoting the image of a ∈ A in H¯ω := H by a, it is completely standard that Ut(a) := [αt(a) defines a continuous unitary representation of R on A so that (π, U) is covariant for α. Hence, {Ut} implements an uw-continuous extension of α to ¯α acting on A = π(A)′′. It is also standard that the cyclic and separating vector ξ0 = 1 gives an extension of the trace ¯ω to a faithful uw-continuous trace on A. By an abuse of notation we will drop the notation "π" for the representation of A and just assume that A acts directly on H. In this way we will also write the extended scalar trace (given by ξ0) on A as ¯ω. With this notation, we invoke [U] to obtain an uw-continuous conditional expectation E : A → Z defined by the equation: For x = a ∈ A and y = z ∈ Z, we have: ¯ω(E(x)y) = ¯ω(xy) f or x ∈ A, y ∈ Z. ¯ω(τ (a)z) = ω(τ (τ (a)z)) = ω(τ (a)z) = ω(τ (az)) = ¯ω(az). Since Z is uw-dense in Z we can replace the z ∈ Z by any y ∈ Z in the previous equation. That is, for a ∈ A we have E(a) = τ (a) and so E is just an extension of τ by uw-continuity. We now use the notation ¯τ in place of E, and observe that since τ is tracial, so is ¯τ . To see that ¯τ is faithful, suppose x ∈ A and ¯τ (x∗x) = 0. Then, by the defining equation for ¯τ we have 0 = ¯ω(¯τ (x∗x)1) = ¯ω(x∗x), and since ¯ω is faithful, x = 0. Finally to see that ¯α leaves ¯τ invariant, we let x ∈ A and t ∈ R. Choose a bounded net {ai} in A which converges to x ultraweakly. Then since ¯αt is spatial, we have αt(ai) = ¯αt(ai) → ¯αt(x) ultraweakly. Hence, ¯τ ( ¯αt(x)) = limi ¯τ (αt(ai)) = limi τ (αt(ai)) = limi τ (ai) = limi ¯τ (ai) = ¯τ (x). (cid:3) Examples. 4-tuples 1. Kronecker (scalar trace) Example. Let A = C(T2), the C ∗-algebra of continuous functions on the 2-torus, with the usual scalar trace τ0 given by integration against the Haar measure on T2. We let αµ : R → Aut(A) be the Kronecker flow on A determined by the real number, µ (note that µ is not a power merely a superscript). That is, for s ∈ R, f ∈ A, and (z, w) ∈ T2 we have: (αµ s (f ))(z, w) = f(cid:0)e−2πis z, e−2πiµs w(cid:1) . In terms of the two commuting unitaries which generate A = C(T2), namely U(z, w) = z and V (z, w) = w we have s (U) = e−2πisU, αµ αµ s (V ) = e−2πisµV. Of course, this action leaves our scalar trace τ0 invariant. In this case where Z = C the faithful state ω on Z = C is just the identity mapping and so ¯ω := ω ◦ τ0 = τ0. That is, H¯ω = Hτ0 = L2(T2) with the obvious representation of A on Hτ0. In this case, Z = Z = C and so A = L∞(T2). Clearly τ0 and α extend to ¯τ0 and ¯α as required. 5 One easily calculates the winding numbers of the generators: windαµ(U) = −1 and windαµ(V) = −µ. 1.a. Noncommutative Tori. We quickly observe that the previous construction can be carried over almost verbatim to noncommutative tori. For θ ∈ [0, 1) let Aθ = C ∗(U, V V U = e2πiθUV ) be the universal C ∗-algebra generated by two unitaries, U, V satisfying the above relation. For θ = 0 the algebra Aθ is naturally isomorphic to A = C(T2) with U(z, w) = z and V (z, w) = w. For θ irrational, these algebras are of course the irrational rotation algebras which are simple C ∗-algebras. We let αµ : R → Aut(A) be the flow on Aθ determined by the real number, µ. That is, for s ∈ R, and U, V the generators of Aθ we have: s (U) = e−2πisU, αµ αµ s (V ) = e−2πisµV. Since αs(U) and αs(V ) satisfy the same relation as U and V this is a well-defined flow on Aθ. The scalar trace, τθ on Aθ on the dense subalgebra of finite linear combinations of U nV m for m, n in Z satisfies: τθ(U nV m) =(cid:26) 0 if n 6= 0 or m 6= 0 1 if n = 0 = m. Again, one easily calculates the winding numbers of the generators: windαµ(U) = −1 and windαµ(V) = −µ. 2. Generalized Kronecker and Generalized Noncommutative tori Examples. We show that any self-adjoint element in any unital commutative C ∗-algebra (with a faithful state) can be used as a replacement for the scalar µ in Examples 1 and 1.a to obtain a non-scalar example. Let Z = C(X) be any commutative unital C ∗-algebra with a faithful state and let η ∈ Zsa be any self-adjoint element in Z. Let A = Z ⊗ C(T2) = C(X, C(T2)) (respectively , A = Z ⊗ Aθ) = C(X, Aθ)) and let τ : A → Z be given by the "slice-map" τ = idZ ⊗ τθ where τθ for θ = 0 is the standard trace on C(T2) given by Haar measure (respectively, the usual scalar trace τθ on Aθ defined above). Then, τ is a faithful, tracial 6 have conditional expectation of A onto Z. In particular, for f ∈ A = Z ⊗ C(T2) ∼= C(T2, Z) we τ (f ) =ZT2 f (z, w)d(z, w) ∈ Z. In this case we note that for A = Z ⊗ C(T2), we have Z(A) = A and hence Z is strictly contained in Z(A). On the other hand, for θ irrational, Z(A) = Z(Z ⊗ Aθ) = Z since Aθ is simple. In either case we use the element η ∈ Zsa to define a τ -invariant action {αη t} of R on A: for f ∈ A, t ∈ R, x ∈ X (again, η and η(x) are not powers, but merely superscripts). It is clear that (A, Z, τ, αη) is a 4-tuple. t (f )(x) = αη(x) αη t (f (x)), In both these cases one calculates the following winding operators: windαη(1 ⊗ U) = −1 ⊗ 1 and windαη (1 ⊗ V) = −η ⊗ 1. Using the faithful state ω on Z, we define a faithful (tracial) state ¯ω on A via ¯ω := ω ◦ τ. By Proposition 2.1, ¯ω is a faithful (tracial) state on A which is left invariant by α and if (π,H) is the GNS representation of A induced by ¯ω then there is a continuous unitary rep- resentation {Ut} of R on H so that (π, U) is covariant for α on A. Also, {Ut} implements an uw-continuous extension of α to ¯α acting on A := π(A)′′. Morover, letting Z := π(Z)′′, there exists a unique faithful unital, uw-continuous Z-trace ¯τ : A → Z extending τ, and ¯α leaves ¯τ invariant. 3. C ∗-algebra of the Integer Heisenberg group Let A be the C ∗-algebra C ∗(H) of the integer Heisenberg group, H: H =  1 m p 0 1 n 0 0 1  m, n, p ∈ Z . We view A = C ∗(H) as the universal C ∗-algebra generated by three unitaries U, V, W satis- fying: W U = UW, W V = V W, and UV = W V U. Here U, V, W correspond respectively to the three generators of H: u = 1 1 0 0 1 0 0 0 1  , v = 1 0 0 0 1 1 0 0 1  , w = 1 0 1 0 1 0 0 0 1  . Proposition 2.2. If H is a discrete group with subgroup C, then the map l1(H) → l1(C) ∗(C). If defined by f 7→ fC extends to faithful, conditional expectation τ from Cr ∗(H) → Cr 7 ∗(H) we see that we can also view τ as a trace on C ∗(H). C is the centre of H then τ is also tracial. Combining τ with the canonical ∗-homomorphism: C ∗(H) → Cr Proof. Let f 7→ πH(f ) and g 7→ πC(g) denote the left regular representations of l1(H) and l1(C) on l2(H) and l2(C) respectively. Then for η ∈ l2(C) ⊆ l2(H) we have: πH (f )(η)(c) =Xh∈H f (ch−1)η(h) =Xh∈C f (ch−1)η(h) =Xh∈C fC (ch−1)η(h) = πC(fC )(η)(c). In other words, for each η ∈ l2(C), πH(f )(η) = πC(fC )(η) so that πH (f )l2(C) = πC(fC). We let E : l2(H) → l2(C) denote the canonical projection. then all η ∈ l2(C) have the form η = E(ξ) for ξ ∈ l2(H) and we have πC(fC )E(ξ) = EπC(fC )E(ξ) = EπH (f )E(ξ). We now define τ (πH (f )) = πC(fC ). To see that τ is bounded in operator norm, kπC(fC )k = kEπH(f )Ek ≤ kπH (f )k. ∗(H) we have Thus τ extends by continuity to τ : Cr τ (x) = EπH (x)E so that the extended τ is clearly completely positive, onto and has norm 1: that is, it is a conditional expectation by Tomiyama's theorem. ∗(C). For general x ∈ Cr ∗(H) → Cr Now for f ∈ l1(H) we have τ (πH (f )) = πC(fC ) so that, if C is the centre of H, then in order to see that τ is tracial it suffices to see that for f, g ∈ l1(H) that (f ∗ g)C = (g ∗ f )C . So for c ∈ C we have: (f ∗ g)(c) =Xh∈H f (ch−1)g(h) =Xh∈H g(h)f (h−1c) = (g ∗ f )(c). (cid:3) In our example where H is the Heisenberg group, its centre is C = {wn n ∈ Z}. In our realization of A = C ∗(H) as a universal C ∗-algebra, the centre of A is Z = C ∗(W ). Now the dense ∗-subalgebra of A generated by U, V, W has as a basis all elements of the form W pV nU m each of which corresponds uniquely to the group element wpvnum = in H. In this notation τ : A → Z is given by: 1 m p 1 n 0 0 0 1  τ (W pV nU m) =(cid:26) 0 W p if n 6= 0 or m 6= 0 if n = 0 = m. In order to define our action α : R → Aut(A), we first fix an element η ∈ Zsa. For an explicit example, we arbitrarily choose η = (µ/3)(W + 1 + W ∗) where µ is a fixed real number . For this fixed η we define the action α via: αt(U) = e−2πitU; αt(V ) = e−2πitηV ; αt(W ) = W. 8 So on the basis elements we get αt(W pV nU m) = e−2πintηe−2πimtW pV nU m = e−2πit(nη+m)W pV nU m. One easily checks that for fixed t the operators Ut := αt(U), Vt := αt(V ), and Wt := W ,satisfy the same relations as U, V, W , namely: WtUt = UtWt ; WtVt = VtWt ; UtVt = WtVtUt. Hence, αt defines a ∗-representation of H in A = C ∗(H) and so extends to a ∗-representation of C ∗(H) inside C ∗(H). Now W is in the range of this ∗-representation and so C ∗(W ) is in the range of this ∗-representation and hence e2πitη is in the range of this ∗-representation for any t ∈ R. Hence V = e2πitηVt is in the range also. Similarly, U is in the range so that αt(C ∗(H)) = C ∗(H) since it is dense and closed. Since α−t is the inverse of αt, αt is one-to-one and hence an automorphism of C ∗(H). One easily checks that αt+s = αtαs using the fact that e−2πisη is in the centre. The point-norm continuity of t 7→ αt(a) is clear. Thus we have an action α : R → Aut(A), that fixes Z = C ∗(W ) = C ∗(C) and leaves the Z-valued trace τ invariant. That is, (C ∗(H), C ∗(C), τ, α) is a 4-tuple. Now the left regular representation of C ∗(C) on l2(C) gives a faithful vector state ω(x) = hx(δ1), δ1i, which for x ∈ l1(C) is just ω(x) = x(1). Then the state ¯ω on C ∗(H) is given for x ∈ l1(H) by: ¯ω(x) = (ω ◦ τ )(x) = ω(xC ) = xC (1) = x(1). Now if x, y ∈ l1(H) then the inner product induced by ¯ω is: x(h)y(h) = hx, yi. That is, H¯ω = l2(H) and the representation of C ∗(H) on H¯ω = l2(H) is just the left regular representation, so in this case, A = W ∗ r (H) the left regular von Neumann algebra of H. x(1h)y∗(h−1) =Xh∈H hx, yi¯ω = ¯ω(x · y∗) = (x · y∗)(1) =Xh∈H Now l2(H) = LX l2(C · X) over all the cosets C · X of C. Moreover, each coset, that l2(H) = L(n,m) l2(C · V nU m). Clearly the left action of C (and hence, of C ∗(C)) on C · (W pV nU m) = C · (V nU m) is uniquely determined by the pair of integers (n, m), so each coset space is unitarily equivalent to the left regular representation of C ∗(C) on l2(C). Hence, the left action of C ∗(C) on l2(H) is just a countably infinite multiple of the left regular representation of C ∗(C) on l2(C). That is, Z = 1Z2 ⊗ W ∗ Thus the map τ : C ∗(H) → C ∗(C) with both acting on l2(H) becomes τ (x) = 1Z2⊗ExE where E is the projection from l2(H) onto l2(C). It is clear that this map is weak − operator continuous and so extends by the same formula to a tracial expectation ¯τ : A → Z. It is also clear that α extends to ¯α as needed. r (C). In this example one calculates the following winding operators in Z = C ∗(W ) windα(U) = −1; windα(V ) = −µ/3(W + 1 + W ∗); windα(W ) = 0. 9 Examples. Morphisms 1. Generalized Kronecker to Kronecker Morphisms. We let A1 = C(X) ⊗ C(T2) and Z1 = C(X) ⊗ 1. We let τ1 = idC(X) ⊗ τ0 where τ0 : C(T2) → C is given by integration with respect to Haar measure on T2. We arbitrarily fix a η ∈ (Z1)sa = (C(X) ⊗ 1)sa. We also define α1 : R → Aut(A1) via: t (h)(x, z, w) = h(x, e−2πitz, e−2πitη(x)w). α1 As before we let u ∈ A1 be the unitary u(x, z, w) = w. We let A2 = C(T2) and Z2 = C1 and τ2 = τ0 : A2 → Z2. We arbitrarily fix an x0 ∈ X and define the evaluation ∗-homomorphism ϕ : A1 → A2 via ϕ(h)(z, w) = h(x0, z, w). We let µ = η(x0) and define t (h)(z, w) = h(e−2πitz, e−2πitµw). α2 One easily checks that ϕ defines a morphism from (A1, Z1, τ1, α1) to (A2, Z2, τ2, α2), and that ϕ(u) = v where v(z, w) = w. 1a. Generalized Noncommutative tori to Kronecker Morphisms. We previously defined A = C(X) ⊗ Aθ and Z = C(X) ⊗ 1. We also let τ1 = idC(X) ⊗ τθ where τθ : Aθ → C is defined above. We arbitrarily fixed an η ∈ (Z)sa = (C(X) ⊗ 1)sa. And then defined α : R → Aut(A) via: (αt(f ))(x) = αη(x) t (f (x)) for f ∈ A, t ∈ R, x ∈ X. We let v ∈ A1 be the constant unitary v(x) = V. We now consider Aθ and Z = C1 and τθ : Aθ → Z. We arbitrarily fix an x0 ∈ X and consider the action of R on Aθ defined by the real number η(x0), that is, αη(x0). This gives us a 4-tuple, (Aθ, C, τθ, αη(x0)). We now the evaluation ∗-homomorphism ϕ : A → Aθ via ϕ(h) = h(x0). One easily checks that ϕ defines a morphism from (A, Z, τ1, α) to (Aθ, C, τθ, αη(x0)). Moreover, ϕ(v) = V. 2. Heisenberg to Kronecker Morphisms. We let A1 = C ∗(H) and Z1 = C ∗(W ) = C ∗(C) ∼= C ∗(Z) ∼= C(T) and recall τ1(W pV nU m) =(cid:26) 0 W p if n 6= 0 or m 6= 0 if n = 0 = m defines a trace τ1 : A1 → Z1. Recall that we (randomly) chose θ = (µ/3)(W +1+W ∗) ∈ (Z1)sa and defined our automorphism group by t (W pV nU m) = e−2πintθe−2πimtW pV nU m = e−2πit(nθ+m)W pV nU m. α1 10 We let A2 = C ∗(H/C) ∼= C ∗(Z2) ∼= C(T2) where the two isomorphisms are given by: Coset(W pV nU m) = C · (W pV nU m) = C · (V nU m) 7→ (n, m) 7→ znwm. We let Z2 = C1 ⊂ A2 and define τ2 : A2 → Z2 = C1 to be the composition of these isomorphisms with the trace on C(T2) given by the Haar integral. This clearly implies that τ2(C · (V nU m)) =(cid:26) 0 if n 6= 0 or m 6= 0 1 if n = 0 = m We now define α2 t ∈ Aut(A2) via t ((C · V )n(C · U)m) = e−2πitnµ(C · V )ne−2πitm(C · U)m = e−2πit(nµ+m)(C · V )n(C · U)m. α2 Clearly, (A2, Z2, τ2, α2) is isomorphic to the Kronecker example with scalar µ. We now define a ∗-homomorphism ϕ : A1 = C ∗(H) → A2 = C ∗(H/C) as the unique extension of the canonical group homomorphism H → H/C. So, ϕ(W pV nU m) = (C · V )n(C · U)m in particular, ϕ(W p) = (C · 1) = 1 ∈ H/C. One easily checks that ϕ defines a morphism from (A1, Z1, τ1, α1) to (A2, Z2, τ2, α2), and that ϕ(W pV nU m) = (C · V )n(C · U)m. Hence, ϕ(θ) = ϕ((µ/3)(W −1 + 1 + W )) = µ by our choice of θ. 3. HILBERT ALGEBRAS OVER an ABELIAN von NEUMANN ALGEBRA Key Idea 2. While centre-valued traces are well-known (eg., the Traces Op´eratorielles of [Dix]) a completely general construction of such traces suitable for use with crossed-products has not (to our knowledge) been attempted before now. In this section we combine the theory of Hilbert modules ([Pa], [R]) with the theory of Hilbert Algebras [Dix] in order to construct centre-valued traces on certain crossed product von Neumann algebras. Although the outline is similar to the usual Hilbert Algebra theory, the details are rather subtle. The main difficulties arise because the usual norm completion of these new "Hilbert Algebras" is not self-dual in the sense of Paschke [Pa]. Definition 3.1. Let B be a von Neumann algebra. A complex vector space X is a (right) pre-Hilbert B-module if there exists a B-valued inner product h·,·i which is linear in the second co-ordinate satisfying: (i) hx, xi ≥ 0 and hx, xi = 0 ⇐⇒ x = 0 for each x ∈ X. (ii) hx, yi∗ = hy, xi for all x, y ∈ X. (iii) hx, yai = hx, yia for all x, y ∈ X and a ∈ B. (iv) span{hx, yi x, y ∈ X} is uw-dense in B. 11 Key Idea 3. In the following we do not assume that our bounded module mappings are adjointable: as pointed out by Lance [L] this yields a rather trivial result that for Hilbert modules all such maps arise from inner products. However, most Hilbert modules are not self-dual: self-dual modules Y have the property that L(Y ) is a von Neumann algebra. In the examples that we use later, the Paschke dual X † of a pre-Hilbert B-module X is a self-dual module that is usually much larger than X. We need these self-dual modules in order to work in the von Neumann algebra, L(X †). Definition 3.2. We follow Paschke [Pa] by defining the dual of a pre-Hilbert B-module X to be the space: X† = {θ : X → B θ is a bounded B-module map}. In order to make the embedding of X into X† linear, Paschke defines scalar multiplication on X† by: Similarly, module multiplication on X† is given by: (λθ)(x) := ¯λθ(x) f or λ ∈ C, θ ∈ X†, and x ∈ X. (θ · a)(x) := (a∗θ(x)) f or θ ∈ X†, a ∈ B, and x ∈ X. Therefore, we can identify X in X† via x 7→ x where x(y) = hx, yi for x, y ∈ X. Since B is a von Neumann algebra, Paschke shows how to extend the B-valued inner product on X to an inner product on X† so that X† becomes self-dual [Pa] Theorem 3.2. This theorem is not trivial. We recall Paschke's construction on page 450 of [Pa]. Let B∗ be the space of ultraweakly continuous linear functionals on B: that is, the predual of B. Now for each positive functional ω in B∗ we have that for Nω = {x ∈ X ω(hx, xi) = 0}, the space X/Nω is a pre-Hilbert space with inner product: hx + Nω, y + Nωiω = ω(hx, yi). Moreover, for each θ ∈ X†, the mapping x + Nω 7→ ω(θ(x)) is a well-defined bounded linear functional on X/Nω satisfying ω(θ(x)) ≤ kωk1/2kθkkx + Nωkω. Hence, there exists a unique vector θω in Hω, the Hilbert space completion of X/Nω, with ω(θ(x)) = hθω, x + Nωiω f or all x ∈ X, and kθωkω ≤ kωk1/2kθk. Thus, kxkω := ω(hx, xi)1/2 is a well-defined seminorm on X which extends naturally to X† via kθkω = hθω, θωi1/2 ∗ , θ ∈ X†, x ∈ X we have: ω . Moreover, for all ω ∈ B+ hθω, x + Nωiω ≤ kθωkωkx + Nωkω ≤ kωk1/2kθkkωk1/2kxk = kωkkθkkxk. We recall from Proposition 3.8 of [Pa] that X† is a dual space with the weak∗-topology given by the linear functionals: θ 7→ ω(hτ, θi) f or ω ∈ B∗ τ ∈ X†. 12 Proposition 3.3. Let B be a von Neumann algebra and let X be a pre-Hilbert B-module. Then, (i) the unit ball of X† is complete in the topology given by the family of seminorms, {k · kω ω ∈ B+ ∗ }; (ii) X is dense in X† in this topology; and hence (iii) X is weak∗ dense in X†. (iv) For each ω ∈ B+ ∗ , θ ∈ X†, and ǫ > 0 there exists an x ∈ X with: kθ − xk2 ω = ω(hθ − x, θ − xi) < ǫ2. Proof. (i) Let {θα} be a Cauchy net in the unit ball of X†. Then, for a fixed ω ∈ B+ ∗ , the net {(θα)ω} is a Cauchy net in the norm k · kω on Hω by definition. Hence, there exists an element θω ∈ Hω with k(θα)ω − θωk → 0. Moreover, kθωk ≤ lim sup α k(θα)ωk ≤ kωk1/2kθαk ≤ kωk1/2. Now, for fixed x ∈ X, {θα(x)} is a bounded net in B. Moreover, for each ω ∈ B+ ∗ lim α ω(θα(x)) = lim α h(θα)ω, x + Nωiω = hθω, x + Nωiω. Thus for every ω ∈ B∗, the net {ω(θα(x))} converges in C. Clearly, this limit is linear in ω: that is, the bounded net {θα(x)} of linear functionals on B∗ converges pointwise to a linear functional on B∗ which is therefore bounded by the same bound, kxk. That is, the ∗ }) defines an element in (B∗)∗ = B via ω 7→ hθω, x + Nωiω. If we call pair (x,{θω ω ∈ B+ this element θ(x), then by definition, ω(θ(x)) = hθω, x + Nωiω = lim and kθ(x)k ≤ kxk. α ω(θα(x)), By this formula, θ(x) is clearly linear in x, and so θ : X → B is linear. By construction, θα(x) converges ultraweakly to θ(x) and since each θα is a B-module map, so is θ. Clearly, kθk ≤ 1, so θ is in the unit ball of X†, and θα converges to θ. That is, the unit ball of X† is complete as claimed. (ii) To see that X is dense in X†, fix θ ∈ X† and ǫ > 0. Let {ω1, ω2, ..., ωm} be a finite set of functionals in B+ ∗ . Given this data we let ω = ω1 + · · · + ωm. Now, ω ≥ ωi for each i = 1, 2, ..., m and so by Proposition 3.1 of [Pa], the map x + Nω 7→ x + Nωi is a well-defined contraction which extends to a contraction Hω → Hωi carrying θω to θωi. We choose x ∈ X so that k(x + Nω) − θωkω < ǫ. Then, for each i = 1, 2, ..., m, we have: kx − θkωi := k(x + Nωi) − θωikωi ≤ k(x + Nω) − θωkω < ǫ. 13 (iii) Now fix θ ∈ X† and let ǫ > 0, {τ1, ...τn} ⊆ X†, {ω1, ..., ωm} ⊆ B∗ define a basic weak∗-neighbourhood of θ. Since every element of B∗ is expressible as a linear combination of four elements in B+ ∗ we can assume that ω1, ..., ωm are positive. Let ω = ω1 + · · · + ωm and choose x ∈ X with k(x + Nω) − θωkω < ǫ kτ1k + · · · + kτnk . Then, for each i = 1, ..., m and k = 1, ..., n, we have: ωihτk, x − θi = hτk, x − θiωi ≤ kτkkωikx − θkωi ≤ kτkkωkx − θkω ≤ kτkkk(x + Nω) − θωkω < ǫ. (iv) This is just a restatement of the fact that X/Nω is dense in its Hilbert space completion (cid:3) Hω as described above in the remarks after Definition 3.2. Remark. In the following class of examples we can more or less explicitly calculate X †. Example 3.4. Let H be a Hilbert space with orthonormal basis {ξn}, let B be a von Neumann algebra, and let X be the algebraic tensor product X = H ⊗ B, with the obvious B-valued inner product. Then, X is a pre-Hilbert B-module and we can identify X † as: X † =(Xn ξn ⊗ bn bn ∈ B and ∃M > 0 with kXn∈F nbnk ≤ M, ∀ f inite F) . b∗ Such a formal sum defines a bounded B-module mapping θ on X as follows: θ NXk=1 ηk ⊗ ak! = NXk=1Xn hξn, ηkib∗ nak, where the right hand side converges in norm. Proof. First, let θ denote an arbitrary element in X †. Define b∗ n := θ(ξn ⊗ 1). Since θ is also defined on the norm closure of X, we see that θ is defined on each element of the nan converges in norm in B. In particular, if η ∈ H, so that form, Pn ξn ⊗ an where Pn a∗ η =Pnhξn, ηiξn converges in norm then, η ⊗ a =Pn ξn ⊗hξn, ηia converges in norm, and so θ(η ⊗ a) =Xn hξn, ηib∗ k=1Pn ξn⊗hξn, ηkiak converges Hence for any element x =PN hξn, ηkib∗ hξn, ηiθ(ξn ⊗ 1)a =Xn k=1 ηk⊗ak ∈ X we have x =PN ηk ⊗ ak! = NXk=1Xn θ(ξn ⊗ hξn, ηia) =Xn θ NXk=1 NXk=1 na =Xn θ(ηk ⊗ ak) = in norm and: hξn, ηib∗ na. nak, 14 as claimed. To see that the bn's satisfy the boundedness condition, let F be any finite set of indices. Then, b∗ kXn∈F = kθk · khXn∈F nbnk = kθ(Xn∈F ξn ⊗ bn,Xn∈F ξn ⊗ bn)k ≤ kθk · kXn∈F ξn ⊗ bniBk1/2 = kθk · kXn∈F ξn ⊗ bnk b∗ nbnk1/2. nbnk1/2 ≤ kθk for all finite F , so we can choose M = kθk2. That is, kPn∈F b∗ On the other hand if we have such a formal sum,Pn ξn ⊗ bn, then we will show that the finite partial sums Pn∈F ξn ⊗ bn form a Cauchy net (in the family of seminorms of Prop. 3.3) in the ball of radius √M in X, and invoke the previous proposition to conclude that they converge pointwise ultraweakly to an element in X † of norm at most √M. ∗ and let ǫ > 0. Since the finite sums, (cid:8)Pn∈F b∗ nbn(cid:9)F form a nbn)(cid:9)F converges to a finite nonnegative number. Thus, there bounded increasing net of positive operators in B, they converge strongly to an element of To this end let ω ∈ B+ nbn) < ǫ/2. B. Hence the net(cid:8)Pn∈F ω(b∗ exists a large finite set F0 so that if F0 ∩ F = φ thenPF ω(b∗ ξn ⊗ bn − XF2∼F1 ξn ⊗ bn −XF2 ξn ⊗ bn − XF2∼F1 Thus if F0 ⊂ F1 and F0 ⊂ F2, we have ξn ⊗ bnk2 ξn ⊗ bn − XF2∼F1 nbn! + ω XF2∼F1 ω = k XF1∼F2 ξn ⊗ bn), (XF1∼F2 nbn! < ǫ/2 + ǫ/2 = ǫ. kXF1 = ω *(XF1∼F2 = ω XF1∼F2 Hence, the finite sumsPF ξn ⊗ bn converge to an element θ ∈ X † : that is, for each x ∈ X, θ(x) = uw − limFhPF ξn ⊗ bn, xi. Now, for x = PN k=1Pn ξn ⊗ hξn, ηkiak converges in norm. Since θ is bounded, part of the proof that x =PN θ(x) =PN k=1Pnhξn, ηkiθ(ξn ⊗ ak) also converges in norm. But then, ξm ⊗ bm, ξn ⊗ ak+B k=1 ηk ⊗ ak ∈ X we have by the first = b∗ nak. ξn ⊗ bn)+B! b∗ ξn ⊗ bnk2 ω θ(ξn ⊗ ak) = uw − limF*Xm∈F k=1 ηk ⊗ ak) =PN k=1Pnhξn, ηkib∗ And so, indeed, θ(PN nak converges in norm. (cid:3) b∗ Key Idea 4. In the definition below of a Z-Hilbert algebra, A, a key idea is the use of the topology given by the seminorms in Proposition 3.3 to replace the norm topology on 15 HA := A† when Z is not C. Hence, axiom (viii) below seems to us the most natural replacement for the usual axiom of the norm-density of A2 in A. When we come to apply this axiom to the crossed product examples that we construct we are actually able to show that a stronger condition holds. However, in order to prove that the algebra of bounded elements Ab also satisfies axiom (viii) we need the weaker version below. Moreover, in the converse construction of a Z-Hilbert Algebra from a given Z-trace one also needs the weaker version of axiom (viii) below. Definition 3.5. Let Z be an abelian von Neumann algebra. A complex ∗-algebra A is called a Z-Hilbert algebra if A is a right pre-Hilbert Z-module which satisfies the further four axioms: (v) ha∗, b∗i = hb, ai f or a, b ∈ A. (vi) hab, ci = hb, a∗ci f or a, b, c ∈ A. (vii) b 7→ ab : A → A is bounded in the Z-module norm for each fixed a ∈ A. (viii) The space A2 = span{ab a, b ∈ A} is dense in A in the topology given by the family of seminorms {k · kω ω ∈ Z+ Remark. It is easy to see that if A2 is norm-dense in A in the Z-module norm, kak2 = kha, aik then axiom (viii) is satisfied. Example 3.6. Let A be a von Neumann algebra and let Z be a von Neumann subalgebra of the centre of A. Suppose τ : A → Z is a faithful, unital uw-continuous Z-trace. Then, for a, b ∈ A, the following inner product makes A into a Z-Hilbert algebra: ∗ }, defined above. ha, biZ := τ (a∗b). Proof. The only axioms that are not completely trivial are (iii) and (vii). Axiom (iii) follows from lemma 1.1, while Axiom (vii) follows from the calculation: A = khab, abiZkZ = kτ (b∗a∗ab)kZ opkbk2 A. ≤ kτ (ka∗akopb∗b)kZ = kak2 kabk2 Since τ is unital, it is easy to see that k1kA = 1 and so kakA ≤ kakop for all a ∈ A. (cid:3) Of course, even if Z = C one usually has strict containment A ⊂ A† := HA. Remarks. We denote by π(a) the operator "left multiplication by a" and note that by axioms (vi) and (vii) π(a) is adjointable with adjoint π(a∗) and hence π(a) is a Z-module mapping. That is, a(bz) = (ab)z f or a, b ∈ A , z ∈ Z. We denote by π′(a) the operator "right multiplication by a" and note that by axioms (v),(vi), and (vii) that π′(a) is also bounded and adjointable with adjoint π′(a∗) and therefore 16 is also a Z-module mapping. That is, (bz)a = (ba)z f or a, b ∈ A , z ∈ Z. A little playing with the axioms and using the fact that Z is abelian yields the further useful identity: (az)∗ = a∗z∗ f or a ∈ A , z ∈ Z. Whenever A is a Z-Hilbert algebra, we will use the suggestive notation HA in place of A† for the Paschke dual of A. That is, HA = A† = {θ : A → Z θ is a bounded Z-module map}. By Theorem 3.2 of [Pa], HA is a self-dual Hilbert Z-module. For ξ ∈ HA and a ∈ A we have ξ(a) = hξ, ai where a ∈ HA is given by a(b) = ha, bi for b ∈ A. We identify a with a ∈ HA so that A ⊆ HA and so, of course, A− ⊆ HA. By Corollary 3.7 of [Pa] each π(a) (respectively, π′(a)) extends uniquely to an element of L(HA) which we will also denote by π(a) (respectively, π′(a)) and moreover, the map: A π→ L(HA) is a ∗-monomorphism. Similarly, the map: A π′ → L(HA) is a ∗-anti-monomorphism. We note that with this notation, axiom (viii) ensures that A2 is weak∗-dense in HA by Proposition 3.3 part (iii). Proposition 3.7. Let A be a Z-Hilbert algebra where Z is an abelian von Neumann algebra. For z ∈ Z and ξ ∈ HA the mapping ξ 7→ z · ξ := ξz embeds Z into L(HA). With this embedding we have the centre of L(HA). Moreover, L(HA) is a Type I von Neumann algebra. Z = Z(L(HA)), Proof. It is easy to check that this mapping embeds Z into L(HA) and since each T ∈ L(HA) is Z-linear we have that Z ֒→ Z(L(HA)). Now by Corollary 7.10 of [R], Z and L(HA) are Morita equivalent in the sense of [R] and so by Theorem 8.11 of [R], L(HA) is a Type I von Neumann algebra. Now by the construction of Corollary 7.10 of [R], HA becomes a left Hilbert L(HA)-module with the inner product: hξ, ηiL(HA)(µ) = ξhη, µiZ f or ξ, η, µ ∈ HA. That is, hξ, ηiL(HA) is the "finite-rank" operator ξ ⊗ η in L(HA). Then, for T ∈ Z(L(HA)), 17 hT ξ, ηiL(HA) = (T ξ) ⊗ η = T (ξ ⊗ η) = (ξ ⊗ η)T = ξ ⊗ T ∗η = hξ, T ∗ηiL(HA). Thus, such a T is adjointable and clearly L(HA)-linear. By Corollary 7.10 of [R], T must be of the form T ξ = ξz = z · ξ for some z ∈ Z. That is, Z = Z(L(HA)). Key Idea 5. The fact that L(HA) is a type I von Neumann algebra with centre Z is one key idea which makes the theory of Z-Hilbert algebras possible. That is, if R is a ∗-subalgebra of L(HA) which contains Z, then R is uw-closed if and only if R = R′′ where ′ denotes commutant within L(HA). This follows from compl´ement 13, III.7 of [Dix] and allows us to use commutation (pure algebra) to determine inclusion or equality of certain algebras. (cid:3) 4. COMMUTATION THEOREM for Z-HILBERT ALGEBRAS Throughout this section Z is an Abelian von Neumann algebra and A is a Z-Hilbert Algebra with Paschke dual HA. Given the machinery we have developed for Z-Hilbert Algebras, the proof of the commutation theorem below follows the outline of the classical case quite closely. Lemma 4.1. If T is a nonzero operator in L(HA) then there exists a ∈ A with T π(a) 6= 0. Proof. If T (A2) = {0}, then for all ξ ∈ HA, hT ∗ξ, abi = hξ, T (ab)i = 0. Hence, for each positive ω ∈ Z∗ we have Then by Definition 3.5 part (viii) and Proposition 3.3 part (ii) we must have T ∗ξ = 0 for all ξ ∈ HA. That is, T ∗ = 0 and hence T = 0. 0 = ω(hab, T ∗ξi) = hab, T ∗ξiω. Therefore, there exists a, b ∈ A with 0 6= T (ab) = T (π(a)b) = (T π(a))(b), so T π(a) 6= 0. (cid:3) Since L(HA) is a von Neumann algebra it has a God-given ultraweak (uw) topology. This is the topology we refer to in the following lemma. Lemma 4.2. With the standing assumptions of this section, we have (i) (π(A))−uw = (π(A))′′ and (ii) Z ⊆ (π(A))−uw. Proof. Since Z is the centre of L(HA) by Proposition 3.7 we see that (π(A))′ = [alg{π(A), Z}]′. 18 Moreover, since L(HA) is type I with centre Z and Z ⊆ alg(π(A), Z), we have by compl´ement 13, III.7 of [Dix] that Hence, (1) [alg(π(A), Z)]′′ = [alg(π(A), Z)]−uw. (π(A))′′ = [alg(π(A), Z)]′′ = [alg(π(A), Z)]−uw. Now, π(A) is a ∗-ideal in the ∗-algebra alg(π(A), Z) so that (π(A))−uw is a ∗-ideal in [alg(π(A), Z)]−uw so that there exists a central projection E in [alg(π(A), Z)]−uw with (π(A))−uw = E[alg(π(A), Z)]−uw. If E 6= 1 then 1 − E 6= 0 but (1 − E)π(A) = {0}, contradicting the previous lemma. Hence, (2) (π(A))−uw = [alg(π(A), Z)]−uw. Equations (1) and (2) imply part (i). Part (ii) follows since Z is contained in any commu- (cid:3) tant. Lemma 4.3. The map ∗ extends to a conjugate-linear isometry of HA (also denoted by ∗) by defining ξ∗(a) := (ξ(a∗))∗ for ξ ∈ HA and a ∈ A. This extension satisfies hξ, ηi∗ = hξ∗, η∗i = hη, ξi, for all ξ, η ∈ HA. Proof. It is easy to see that ξ∗ is a bounded Z-module map and that kξ∗k ≤ kξk. Since ξ∗∗ = ξ we see that ∗ is isometric on HA. By axioms (ii) and (v) we have for a, b ∈ A, (b)∗(a) = (b(a∗))∗ = hb, a∗i∗ = ha∗, bi = hb∗, ai = bb∗(a), so that this ∗ really is an extension from A to HA. Moreover, using the definition of module multiplication given in Definition 3.2 it is easy to check that (ξz)∗ = ξ∗z∗ for all z ∈ Z and ξ ∈ HA. We observe that Z is a self-dual Hilbert Z-module with the inner product hz1, z2i = z∗ for, if θ : Z → Z is a bounded Z-module map then θ(z) = θ(1)z = hθ(1)∗, zi. 1z2 : Now if ξ ∈ HA, then by Proposition 3.6 of [Pa], ξ extends uniquely to a bounded Z-module mapping: HA → Z. But, using the first paragraph of the proof one checks that η 7→ hξ, ηi and η 7→ hξ∗, η∗i∗ are two such extensions. Hence, hξ, ηi = hξ∗, η∗i∗ as claimed. The equality, hξ, ηi∗ = hη, ξi follows from axiom (ii) since HA is a (self-dual) Hilbert Z-module by Theorem 3.2 of [Pa]. (cid:3) Definition 4.4. The isometry η 7→ η∗ : HA → HA of the previous lemma will be denoted by J. That is, J(η) = η∗ for all η ∈ H. Remarks. The unique extension of Proposition 3.6 of [Pa] used in the previous proof will be used several more times in this paper under the name "unique extension property." Lemma 4.5. With the standing assumptions of this section, 19 (1) Z ⊆ (π′(A))−uw = (π′(A))′′, (2) π(A) ⊆ (π′(A))′ and (3) π′(A) ⊆ (π(A))′. Proof. (1) This is the same proof as Lemma 4.2. (2) and (3) By the unique extension property, it suffices to see that π′(a)π(b) = π(b)π′(a) (cid:3) on the space A ⊆ HA. This is trivial to check. 4.1. Bounded elements in HA. Let ξ ∈ HA and suppose that the map a 7→ π′(a)ξ : A → HA is bounded. We note that by the remarks following example 3.6, π(az) = π(a)z = zπ(a) and π′(az) = π′(a)z = zπ′(a), for all a ∈ A and z ∈ Z. Therefore, (az) 7→ π′(az)ξ = zπ′(a)ξ = (π′(a)ξ)z so that this bounded map is also Z-linear. Hence by the unique extension property this map extends uniquely to a bounded module mapping HA → HA which we denote by π(ξ). That is, π(ξ)a = π′(a)ξ for all a ∈ A. By Proposition 3.4 of [Pa] π(ξ) is adjointable and π(ξ) ∈ L(HA). Such an element ξ ∈ HA is called lef t − bounded and the set of all such elements is denoted Al. Clearly, A ⊆ Al. Similarly, we let Ar = {η ∈ HA π′(η) ∈ L(HA)}. Where, of course, π′(η)a = π(a)η for all a ∈ A. Proposition 4.6. With the standing assumptions of this section, (1) π(Al) ⊆ (π′(A))′ and similarly π′(Ar) ⊆ (π(A))′, (2) π(Al) is a left ideal in (π′(A))′ and T π(ξ) = π(T ξ) f or ξ ∈ Al and T ∈ (π′(A))′. In particular, π(η)π(ξ) = π(π(η)ξ) f or η, ξ ∈ Al. Similarly, π′(Ar) is a left ideal in (π(A))′, etc. (3) Al is an associative algebra with the multiplication ξη = π(ξ)η and π : Al → L(HA) is a monomorphism. Similarly, Ar is an associative algebra with the multiplication ξη = π′(η)ξ, and π′ is an anti-monomorphism. 20 (4) Al is invariant under ∗ and π(ξ∗) = π(ξ)∗ so that π(Al) is a ∗-ideal in (π′(A))′ and π is a ∗-monomorphism. A similar statement holds for Ar. Proof. (1) By the unique extension property, it suffices to check that if ξ ∈ Al, and b ∈ A then π(ξ)π′(b) = π′(b)π(ξ) on the space A. To this end let a ∈ A, then: (π(ξ)π′(b))(a) = π(ξ)(ab) = π′(ab)(ξ) = π′(b)π′(a)(ξ) = π′(b)π(ξ)(a), as required. (2)If ξ ∈ Al, T ∈ (π′(A))′ and a ∈ A, then: π(T ξ)a = π′(a)T ξ = T π′(a)ξ = T π(ξ)a. That is, T ξ ∈ Al and π(T ξ) = T π(ξ) by the unique extension property. (3)By (2), ξη := π(ξ)η is in Al if ξ, η ∈ Al. Moreover, by (2) π(ξη) = π(ξ)π(η). Since π : Al → L(HA) is clearly linear, it suffices to see that π is also one-to-one. But if π(ξ) = 0, then for all a, b ∈ A we have 0 = hπ(ξ)a, biω = hπ′(a)ξ, biω = hξ, ba∗iω for all positive ω ∈ Z∗. That is, ξ = 0 by axiom (viii) and Proposition 3.3. (4)Let ξ ∈ Al and let a, b ∈ A. Using Lemma 4.3 and the fact that HA is a Hilbert Z-module, we get the following calculation: hπ(ξ)∗a, bi = hb, π(ξ)∗ai∗ = hπ(ξ)b, ai∗ = hπ′(b)ξ, ai∗ = hξ, ab∗i∗ = hξ∗, ba∗i = hξ∗, π′(a∗)bi = hπ′(a)ξ∗, bi = hπ(ξ∗)a, bi. Thus, as module maps π(ξ)∗a and π(ξ∗)a agree for all b ∈ A and so π(ξ)∗a = π(ξ∗)a for all a ∈ A. That is, ξ∗ is left-bounded and π(ξ∗) = π(ξ)∗. Moreover, for ξ, η ∈ Al π((ξη)∗) = [π(ξη)]∗ = [π(ξ)π(η)]∗ = π(η)∗π(ξ)∗ = π(η∗)π(ξ∗) = π(η∗ξ∗) and so (ξη)∗ = η∗ξ∗ as π is one-to-one. (cid:3) Corollary 4.7. With the standing assumptions of this section , (1) (π(Al))′′ = π(Al)−uw = (π′(A))′, and (2) (π′(Ar))′′ = π′(Ar)−uw = (π(A))′. Proof. (1) By Proposition 4.6, π(Al)−uw is a ∗-ideal in (π′(A))′. But by Lemma 4.2, 1 ∈ Z ⊆ π(A)−uw ⊆ π(Al)−uw and so π(Al)−uw = (π′(A))′. Now, since Z ⊆ (π(Al))−uw we have by compl´ement 13, III.7 of [Dix] that (π(Al)−uw)′′ = π(Al)−uw. But then, since commutants are always ultraweakly closed: (π(Al))′′ = (π(Al)′′)−uw ⊇ (π(Al))−uw = (π(Al)−uw)′′ ⊇ (π(Al))′′. The proof of (2) is similar. Proposition 4.8. With the standing assumptions of this section, Al = Ar and (1) π′(ξ)a = [π(ξ∗)a∗]∗ for ξ ∈ Al, a ∈ A. (2) π(ξ)a = [π′(ξ∗)a∗]∗ for ξ ∈ Ar, a ∈ A. Proof. (1) Let ξ ∈ Al. Then for a, b ∈ A, hπ′(ξ)a, bi = hπ(a)ξ, bi = hξ, a∗bi = hξ∗, b∗ai∗ = hπ′(a∗)ξ∗, b∗i∗ = hπ(ξ∗)a∗, b∗i∗ = h[π(ξ∗)a∗]∗, bi. 21 (cid:3) (cid:3) Therefore, ξ ∈ Ar so that Al ⊆ Ar and (1) holds. Similarly, Ar ⊆ Al and (2) holds. Corollary 4.9. For all ξ ∈ Al = Ar and η ∈ HA, (1) π′(ξ)η = [π(ξ∗)η∗]∗ and (2) π(ξ)η = [π′(ξ∗)η∗]∗. Proof. (1) Recall J : HA → HA is the conjugate-linear isometry Jη = η∗. As noted in the proof of Lemma 4.3, J(ηz) = (Jη)z∗ for z ∈ Z. Now, by part (1) of the previous proposition, we see that for ξ ∈ Al = Ar, π′(ξ) and Jπ(ξ∗)J agree on A. Since both of these maps are bounded Z-module maps they agree on HA by uniqueness. This proves part (1). The proof of part (2) is similar. Proposition 4.10. Let ξ, η ∈ Al = Ar, then we have: (cid:3) (1) π(ξ)η = π′(η)ξ so that the two multiplications of Proposition 4.6 agree, and (2) π(ξ)π′(η) = π′(η)π(ξ). Proof. (1) Fix a ∈ A, then: hπ(ξ)η, ai = h(π(ξ)η)∗, a∗i∗ = hπ′(ξ∗)η∗, a∗i∗ = hη∗, π′(ξ)a∗i∗ = hη∗, π(a∗)ξi∗ = hπ(a)η∗, ξi∗ = hπ′(η∗)a, ξi∗ = ha, π′(η)ξi∗ = hπ′(η)ξ, ai so that (1) holds. (2) Again fix a ∈ A then, π(ξ)π′(η)a = π(ξ)π(a)η = π(π(ξ)a)η by 4.6(2) = π′(η)(π(ξ)a) = π′(η)π(ξ)a. 22 Notation. Since Al = Ar (even as ∗-algebras) we now use the notation Ab to denote the ∗-algebra of bounded elements in HA. Theorem 4.11. [Commutation Theorem] Let A be a Z-Hilbert Algebra over the abelian von Neumann algebra Z. Then, (cid:3) (1) π(A)−uw = (π(A))′′ = (π(Ab))′′ = π(Ab)−uw = (π′(Ab))′ = (π′(A))′ and (2) π′(A)−uw = (π′(A))′′ = (π′(Ab))′′ = π′(Ab)−uw = (π(Ab))′ = (π(A))′. Proof. (1) By part (1) of Corollary 4.7, we have However, by part (2) of the previous corollary, we have (π(Ab))−uw = (π(Ab))′′ = (π′(A))′ ⊇ (π′(Ab))′. (π(Ab))′′ ⊆ (π′(Ab))′′′ = (π′(Ab))′. Hence, On the other hand, by part (2) of Corollary 4.7: (π(Ab))−uw = (π(A))′′ = (π′(A))′ = (π′(Ab))′. Since π(A)−uw = (π(A))′′ by Lemma 4.2, we are done. (π(A))′′ = (π′(Ab))′′′ = (π′(Ab))′. The proof of (2) is similar. (cid:3) Definition 4.12. We define the left von Neumann algebra of A to be We define the right von Neumann algebra of A to be U(A) := (π(A))′′. V(A) := (π′(A))′′. Corollary 4.13. Let A be a Z-Hilbert algebra over the abelian von Neumann algebra Z. Then, for all ξ, η ∈ Ab, with J as in Definition 4.4 (1) Jπ(ξ)J = π′(Jξ) and Jπ′(ξ)J = π(Jξ). JU(A)J = V(A) and JV(A)J = U(A). (2) Proof. Item (1) is just Corollary 4.7. To see item (2), let T ∈ U(A) = (π′(Ab))′. Then for ξ ∈ Ab and η ∈ HA we get: JT Jπ(ξ)η = JT Jπ(ξ)Jη∗ = JT π′(Jξ)η∗ = Jπ′(Jξ)T η∗ = Jπ′(Jξ)JJT Jη = π(ξ)JT Jη. Therefore, JU(A)J ⊆ (π(Ab))′ = V(A). Similarly, JV(A)J ⊆ U(A). Since J 2 = 1, we're done. Remarks. At this point we could show that Ab is a Z-Hilbert algebra satisfying HAb = HA, U(Ab) = U(A), and V(Ab) = V(A). Since we don't appear to need this now, we defer the statement and proof to Proposition 6.4. (cid:3) 23 5. CENTRE-VALUED TRACES With the same hypotheses and notation of the previous section we show how to construct a natural Z-valued trace on the von Neumann algebra, U(A). We first remind the reader of Paschke's results that both HA and L(HA) are dual spaces, and that since L(HA) is a von Neumann algebra, its weak∗-topology must also be its uw-topology since pre-duals for von Neumann algebras are unique. Key Idea 6. The problem of convergence is one of our main headaches. The topology of Proposition 3.3 (closely related to a topology introduced by Paschke [Pa]) and Proposition 3.10 of [Pa] are exactly what is needed to prove the following result which is used several times in the remainder of this paper. Proposition 5.1. If A is a pre-Hilbert Z-module (not necessarily a Z-Hilbert Algebra) with Paschke dual HA, then: (1) A bounded net {ξα} in HA converges weak∗ to ξ ∈ HA ⇐⇒ hη, ξαi → hη, ξi ultraweakly in Z for all η ∈ HA. (2) A net {Tα} in L(HA) converges ultraweakly to T ∈ L(HA) ⇐⇒ hTαξ, ηi → hT ξ, ηi ultraweakly in Z for all ξ, η ∈ HA. (3) A bounded net {Tα} in L(HA) converges ultraweakly to T ∈ L(HA) ⇐⇒ hTαa, bi → hT a, bi ultraweakly in Z for all a, b ∈ A. Proof. Item (1) is just Remark 3.9 of [Pa] and works for any self-dual Hilbert module over a von Neumann algebra. Item (2) follows immediately from the definition of the weak∗ topology on L(HA) in Remark 3.9 and the proof of Proposition 3.10 of [Pa]. This result also holds for any self-dual Hilbert module over a von Neumann algebra. Item (3) follows from item (2) and the usual ǫ/3-argument using item (iv) of Proposition (cid:3) 3.3. Since π(A2 b) is going to be the domain of definition of our Z-valued trace on U(A), we need a condition on an operator T ∈ U(A) (involving Z-valued inner products) to be an element of π(Ab). 24 Remark. In Example 3.6 where our Z-Hilbert algebra is itself a von Neumann algebra A with Z ⊆ Z(A) and a faithful, tracial, uw-continuous Z-trace τ : A → Z, one can use item (3) in Proposition 5.1 to show that π(A) = (π(A))′′, as expected. Proposition 5.2. If T ∈ U(A) then T ∈ π(Ab) if and only if {hT ξ, T ξi ξ ∈ Ab and kπ(ξ)k ≤ 1} is bounded above in Z+. In this case, T = π(η) where z = hη, ηi, and z is the supremum of this set in Z+. Proof. (⇐=) Let z be an upper bound for this set in Z+. Let {π(ξα)} be a net in π(Ab) converging ultraweakly to 1 and norm bounded by 1. Then, kT ξαk = khT ξα, T ξαik1/2 ≤ kzk1/2 so that {T ξα} is a bounded net in the dual space HA and so we can assume that it converges weak∗ to some η ∈ HA. That is, T ξα → η and π(T ξα) = T π(ξα) uw→ T. w∗ By Proposition 5.1 we see that for all a ∈ A and all µ ∈ HA: hT a, µi = lim α hπ(T ξα)a, µi = lim α hπ′(a)T ξα, µi = lim α hT ξα, π′(a∗)µi = hη, π′(a∗)µi = hπ(η)a, µi. So, T a = π(η)a for all a ∈ A and hence T = π(η) where η ∈ Ab. (=⇒) On the other hand, if T = π(η) for some η ∈ Ab, then for all ξ ∈ Ab with kπ(ξ)k ≤ 1 we get by Proposition 2.6 of [Pa]: hT ξ, T ξi = hηξ, ηξi = hξ∗η∗, ξ∗η∗i = hπ(ξξ∗)η∗, η∗i ≤ kπ(ξξ∗)khη, ηi ≤ hη, ηi ∈ Z. Now, since Z is abelian, the supremum of any finite set of self-adjoint elements exists and so the supremum of the bounded set, {hT ξ, T ξi ξ ∈ Ab and kπ(ξ)k ≤ 1} can be written as the limit of a bounded increasing net of elements in Z+ which exists (in Z+) by Vigier's Theorem. We let z0 be this supremum. Then, if T = π(η) for η ∈ Ab we see by the second part of the above argument that z0 ≤ hη, ηi. On the other hand, If we choose the net {ξα} as in the first part of the above argument to also satisfy ξ∗ α = ξα, then: hT ξα, T ξαi = hηξα, ηξαi = hξαη∗, ξαη∗i = hπ(ξα)2η∗, η∗i uw−→ hη∗, η∗i = hη, ηi. That is hη, ηi ≥ z0, and we're done. Lemma 5.3. Let I = π(Ab)2 := span{π(ξ)π(η) ξ, η ∈ Ab}. Then I is an uw dense ∗-ideal in U(A) and I+ = {π(ξ∗)π(ξ) ξ ∈ Ab}. (cid:3) Proof. It follows from Proposition 4.6 and Theorem 4.11 that I is an uw dense ∗-ideal in U(A). Let I0 = {π(ξ∗)π(ξ) ξ ∈ Ab}. We verify that I0 satisfies the conditions of Lemme 1 of I.1.6 of [Dix]. 25 (i) I0 is unitarily invariant in U(A) since π(Ab) is an ideal in U(A). (ii) Let η ∈ Ab and let T ∈ U(A)+ with 0 ≤ T ≤ π(η∗)π(η). Then for each ξ ∈ Ab with kπ(ξ)k ≤ 1 we get: hT 1/2ξ, T 1/2ξi = hT ξ, ξi ≤ hπ(η∗)π(η)ξ, ξi = hηξ, ηξi = hξ∗η∗, ξ∗η∗i ≤ kπ(ξ∗)k2hη∗, η∗i ≤ hη, ηi. By Proposition 5.2, T 1/2 = π(µ) for some µ ∈ Ab. That is, T = π(µ∗)π(µ) ∈ I0. (iii) If S = π(η∗η) and T = π(µ∗µ) are in I0, then for all ξ ∈ Ab with kπ(ξ)k ≤ 1 we have: h(S + T )1/2ξ, (S + T )1/2ξi = hSξ, ξi + hT ξ, ξi = hπ(η∗η)ξ, ξi + hπ(µ∗µ)ξ, ξi ≤ · · · ≤ hη, ηi + hµ, µi. Again by Proposition 5.2, (S + T )1/2 = π(γ) for some γ ∈ Ab, and so S + T = π(γ∗γ) ∈ I0. Hence, I0 = J+ the positive part of an ideal J and J = spanI0. Clearly, J ⊆ I. On the other hand, if ξ, η ∈ Ab then π(ξ)π(η∗) = 1 4 3Xk=0 ikπ(ξ + ikη)π((ξ + ikη)∗) is in J . Thus, I ⊆ J , and so they are equal. That is, {π(ξ∗)π(ξ) ξ ∈ Ab} = I0 = J+ = I+. (cid:3) (cid:3) Corollary 5.4. With the above hypotheses, I := span{π(ξ)π(η) ξ, η ∈ Ab} = {π(ξ)π(η) ξ, η ∈ Ab}. Proof. Let T ∈ I and let T = V T be the polar decomposition of T in U(A). Then T = V ∗T ∈ I+. Hence, T = V T = V π(ξ)π(ξ∗) = π(V ξ)π(ξ∗) by part (2) of Proposition 4.6. Remarks. At this point we can define a "trace" on the ideal I in the usual way: τ (π(ξη)) := hξ∗, ηi, as in the following theorem. However, in order to connect this up with Dixmier's "trace op´eratorielle" [Dix] which includes unbounded operators affiliated with Z in its range (and also includes a notion of normal) we are forced to work a little harder. 26 Theorem 5.5. Let A be a Z-Hilbert algebra over the abelian von Neumann algebra Z. Let b) be the canonical uw dense ∗-ideal in U(A) = (π(A))′′, the left von Neumann I = π(A2 algebra of A. Then, τ : I → Z defined by τ (π(ξη)) = hξ∗, ηi is a well-defined positive Z-linear mapping which is: (1) faithful, i.e., τ (T ) = 0 and T ≥ 0 =⇒ T = 0 and, (2) tracial, i.e., τ (T S) = τ (ST ) f or T ∈ U(A) and S ∈ I. Proof. To see that τ is well-defined, fix a net {ξα} in Ab with π′(ξα) → 1 ultraweakly. Let T = π(ξη) ∈ I. Then the element ξη ∈ A2 b is unique since π is one-to-one (of course, its representation as a product is not unique). Now, τ (T ) = hξ∗, ηi = uw lim α hπ′(ξα)ξ∗, ηi = uw lim α hξα, ξηi. That is, τ (T ) is uniquely determined by T . Thus, τ (T ) is well-defined and Z-linear. If T ∈ I+, then T = π(ξ∗ξ) by Lemma 5.3 and τ (T ) = hξ, ξi ≥ 0 so that τ is positive. Clearly, τ (T ) = 0 =⇒ ξ = 0 =⇒ π(ξ) = 0 =⇒ T = 0. That is, τ is faithful. To see that τ is tracial, let S = π(ξη) ∈ I and let T ∈ U(A). Then, τ (T S) = τ (T π(ξ)π(η)) = τ (π(T ξ)π(η)) = h(T ξ)∗, ηi = hT ξ, η∗i∗ = hξ, T ∗(η∗)i∗ = hξ∗, (T ∗(η∗))∗i = τ (π(ξ)π(T ∗(η∗))∗) = τ (π(ξ)[T ∗π(η∗)]∗) = τ (π(ξ)π(η)T ) = τ (ST ). (cid:3) 6. TRACES OP ´ERATORIELLES We recall here J. Dixmier's definition of a "Z-trace" [Dix]. We begin by paraphrasing (and translating) Dixmier's discussion of the formal set-up. Let A be a von Neumann algebra and let Z be a von Neumann subalgebra of the centre of A. In this section we fix a locally compact Hausdorff space X, a positive measure ν on X, and an isomorphism of L∞(X, ν) with Z (see th´eor`eme 1 of I.7 of [Dix]). Then Z+ is embedded in the set, Z+, of nonnegative measureable functions on X which are not necessarily finite- valued. Of course, we identify functions in Z+ which are equal ν-almost everywhere. As mentioned before, any bounded increasing net in Z+ has a supremum in Z+. It is clear that the same thing holds for the set Z+. Definition 6.1. With the above notation, we define a Z-trace on A+ to be a mapping φ : A+ → Z+ which satisfies: 27 (i) If S, T ∈ A+ then φ(S + T ) = φ(S) + φ(T ), (ii) If S ∈ A+ and T ∈ Z+ then φ(T S) = T φ(S), and (iii) If S ∈ A+ and U is a unitary in A then φ(USU ∗) = φ(S). We call φ faithful if S ∈ A+ and φ(S) = 0 =⇒ S = 0. We call φ finite if φ(S) ∈ Z+ for all S ∈ A+. We call φ semifinite if for each nonzero S ∈ A+ there exists a nonzero T ∈ A+ with T ≤ S and φ(T ) ∈ Z+. We call φ normal if for every bounded increasing net {Sα} in A+ with supremum S ∈ A+, φ(S) is the supremum of the increasing net {φ(Sα)} in Z+. We now show that if A is a Z-Hilbert algebra then there is a natural Z-trace on the von Neumann algebra U(A) constructed in the usual way. Theorem 6.2. (cf., Th´eor`eme 1, I.6.2 of [Dix]) Let A be a Z-Hilbert algebra over the abelian von Neumann algebra Z and let τ : I = π(A2 b) → Z be the tracial mapping defined in Theorem 5.5. Then τ restricted to I+ extends to a mapping ¯τ : U(A)+ → Z+ via: ¯τ (T ) = sup{τ (S) S ∈ I+, S ≤ T}. This extension is a faithful, normal, semifinite Z-trace in the sense of Dixmier and moreover, Clearly, ¯τ is the unique normal extension of τ. {T ∈ U(A)+ ¯τ(T ) ∈ Z+} = I+. Proof. This proof is similar in outline to Th´eor`eme 1, I.6.2 of [Dix]. However, there are many complications (some subtle) in this degree of generality. At least it is clear that ¯τ extends τ. (i) ¯τ is additive. Trivially, we have for T1, T2 ∈ U(A)+ ¯τ (T1) + ¯τ (T2) ≤ ¯τ (T1 + T2). 1 = AT 1/2 2 = BT 1/2 for A, B ∈ U(A) and E = A∗A + B∗B is the range projection of T . Now, On the other hand, let T = T1 + T2 for T1, T2 ∈ U(A). Then by p. 86 of [Dix], T 1/2 and T 1/2 if 0 ≤ S ≤ T with S ∈ M+ then ASA∗ ≤ AT A∗ = (AT 1/2)(AT 1/2)∗ = T 1/2 1 T 1/2 1 = T1, 28 and similarly, BSB∗ ≤ T2. Since I is an ideal, ASA∗ and BSB∗ are in I+. Thus, since ES = S, τ (S) = τ (ES) = τ (A∗AS) + τ (B∗BS) Taking the supremum over all such S yields the other inequality: = τ (ASA∗) + τ (BSB∗) ≤ ¯τ (T1) + ¯τ (T2). ¯τ (T ) ≤ ¯τ (T1) + ¯τ (T2). (ii) ¯τ is Z+-linear. Unlike the scalar case this is not completely trivial. If E is a projection in Z+ and T ∈ U(A)+, then one easily checks that: (S ∈ I+ and S ≤ ET ) ⇐⇒ (S = ER f or R ∈ I+ with R ≤ T ). Applying the definition of ¯τ , we get ¯τ (ET ) = E ¯τ (T ). Now, if z0 ∈ Z+ and if there exists z1 ∈ Z+ with z1z0 = E the range projection of z0 then again one shows that: (S ∈ I+ and S ≤ z0T ) ⇐⇒ (S = z0R f or R ∈ I+ with R ≤ T ). Hence, ¯τ (z0T ) = z0¯τ (T ) if z0 is bounded away from 0 on its range projection. Now for an arbitrary z0 ∈ Z+ and T ∈ U(A)+ we work pointwise on X where we have identified Z = L∞(X.ν). So, fix x ∈ X. There are two cases. If z0(x) = 0, then [z0¯τ (T )](x) = z0(x)¯τ (T )(x) = 0. On the other hand, if S ≤ z0T and S ∈ I+ then S = ES where E, the range projection of z0, satisfies E(x) = 0, then: τ (S)(x) = τ (ES)(x) = (Eτ (S))(x) = E(x)τ (S)(x) = 0. Taking the supremum over such S we get ¯τ (z0T )(x) = 0 That is, if z0(x) = 0, then ¯τ (z0T )(x) = [z0¯τ (T )](x) = 0. In the second case, z0(x) > 0, so that we can write z0 = z1 + z2 in Z+ where z1 is bounded away from 0 on its support (which contains x) and z2(x) = 0. Then: ¯τ (z0T )(x) = [¯τ (z1T ) + ¯τ (z2T )](x) = [z1¯τ (T ) + ¯τ (z2T )](x) = z1(x)¯τ (T )(x) + ¯τ (z2T )(x) = z0(x)¯τ (T )(x) + 0 = [z0¯τ (T )](x). Hence, ¯τ (z0T ) = z0¯τ (T ). (iii) ¯τ is unitarily invariant. This follows easily from Theorem 5.5 part (2). (iv) ¯τ is faithful. If ¯τ (T ) = 0, then the only S ∈ I+ with S ≤ T is S = 0. However, if {π(ξα)} is a net in π(Ab) converging ultraweakly to 1 and having norm ≤ 1 then: 0 ≤ T 1/2π(ξαξ∗ α)T 1/2 ≤ T. But, T 1/2π(ξαξ∗ α)T 1/2 is in I+ and converges ultraweakly to T . Hence, T = 0. 29 (v) ¯τ is semifinite. This is the same argument as in part (iv). (vi) {T ∈ U(A)+ ¯τ (T ) ∈ Z+} = I+. Clearly, I+ is contained in this set. So, suppose ¯τ (T ) = z ∈ Z+. We apply Proposition 5.2. That is, let ξ ∈ Ab satisfy kπ(ξ)k ≤ 1. Then, and so, But, π(cid:2)(T 1/2(ξ))(T 1/2(ξ))∗(cid:3) = T 1/2π(ξξ∗)T 1/2 ≤ T τ(cid:0)π(cid:2)(T 1/2(ξ))(T 1/2(ξ))∗(cid:3)(cid:1) ≤ ¯τ (T ) = z. τ(cid:0)π(cid:2)(T 1/2(ξ))(T 1/2(ξ))∗(cid:3)(cid:1) = h(T 1/2(ξ))∗, (T 1/2(ξ))∗i = hT 1/2(ξ), T 1/2(ξ)i. Therefore, by Proposition 5.2, T 1/2 = π(η) for some η ∈ Ab and so T = π(η∗η) ∈ I+. (vii) ¯τ is normal. We first show that ¯τ satisfies the normality condition when the relevant operators are all in I+. That is, suppose that {π(ξ∗ αξα)} is an increasing net in I+ with least upper bound π(ξ∗ξ) also in I+. Now for any η ∈ Ab we have by the polar decomposition theorem that π(η) = V π(η) = π(V η) and that V η ∈ Ab. Hence, for any η ∈ Ab, π(η∗η) = π(η)2 = π((V η)2) and π(V η) ≥ 0. Thus we can assume that ξα and ξ are self-adjoint and that π(ξα) ≥ 0 and π(ξ) ≥ 0. Then, π(ξα) = (π(ξ∗ αξα))1/2 and π(ξ) = (π(ξ∗ξ))1/2. α) → π(ξ2) in the strong operator topology by Vigier's theorem and by the proof of Th´eor`eme 1 of I.6.2 of [Dix] we also have π(ξα) → π(ξ) in the strong operator topology. As the square root function is operator monotone, this implies that π(ξ) = supα π(ξα). Now, π(ξ2 It easily follows that kξαk ≤ kξk for all α. Since HA is a dual space, we can find a subnet {ξβ} which converges weak∗ to some ζ ∈ HA. To see that ζ = ξ, let λ, µ ∈ Ab then by Proposition 5.1: hζ, λµi = lim β hξβ, λµi = lim β hπ(ξβ)µ∗, λi = hπ(ξ)µ∗, λi = hξ, λµi. Thus, ζ and ξ define the same Z-valued mapping on A2 on A. That is, ζ = ξ. b ⊇ A2 and therefore the same mapping Now, since τ is positive we have τ (π(ξ∗ξ)) ≥ sup α τ (π(ξ∗ αξα)). On the other hand, by Kaplansky's Cauchy-Schwarz inequality [K] (which holds since Z is abelian) we have: Since ξ and ξβ are self-adjoint it is seen that hξβ, ξi is also self-adjoint and so in fact hξβ, ξi ≤ hξβ, ξβi1/2hξ, ξi1/2 f or all β. hξβ, ξi ≤ hξβ, ξβi1/2hξ, ξi1/2 f or all β. 30 Hence, hξ, ξi = uw lim β hξβ, ξi ≤ sup β hξβ, ξβi1/2hξ, ξi1/2 ≤ (sup α hξα, ξαi1/2)hξ, ξi1/2. Since Z is abelian this implies that hξ, ξi1/2 ≤ sup α hξα, ξαi1/2 and so hξ, ξi ≤ sup α hξα, ξαi. That is, τ (π(ξ∗ξ)) ≤ sup α τ (π(ξ∗ αξα)), and so they are equal. Now, we let {Tα} be an increasing net in U(A)+ with supremum T ∈ U(A)+. We define f = supα(¯τ (Tα)), in Z+. Let E = {x ∈ X f (x) = +∞}. Since ¯τ (Tα) ≤ ¯τ (T ) for all α, we have f ≤ ¯τ (T ). Hence f agrees with ¯τ (T ) on the measureable set E. The complement of E is the countable union of the measureable sets EN := {x ∈ X f (x) ≤ N}, so it suffices to see that f agrees with ¯τ (T ) (almost everywhere) on each EN . To this end, let zN be the characteristic function of EN . Clearly, zN ∈ Z+ and zN T = supα zN Tα in U(A)+. Now, for each α, ¯τ (zN Tα) = zN ¯τ (Tα) ≤ zN f ≤ NzN ∈ Z+. So, by an earlier part of the proof, there exists ξα = ξ∗ hξα, ξαi ≤ NzN . Now, for each η ∈ Ab with kπ(η)k ≤ 1 we have: α hzN Tαη, ηi = lim hzN T 1/2η, zN T 1/2ηi = hzN T η, ηi = lim α ∈ Ab with zN Tα = π(ξ∗ αξα) and = lim α hη∗ξα, η∗ξαi = lim α hξαη, ξαηi α hπ(ηη∗)ξα, ξαi ≤ sup α hξα, ξαi ≤ NzN . Therefore, by Proposition 5.2 there exists a ζ ∈ Ab with zN T 1/2 = π(ζ). Moreover, sup α π(ξ∗ αξα) = sup α zN Tα = zN T = π(ζ ∗ζ). Hence by the first part of the proof of normality of ¯τ , ¯τ (π(ξ∗ ¯τ (zN T ) = ¯τ (π(ζ ∗ζ)) = sup α αξα)) = sup α ¯τ (zN Tα). That is, for x ∈ EN we have: f (x) = (zN f )(x) = (zN sup α ¯τ (Tα))(x) = (sup α ¯τ (zN Tα))(x) = (¯τ (zN T ))(x) = (zN ¯τ (T ))(x) = ¯τ (T )(x) as required. Remarks. In the above setting we want to observe that Ab is also a Z-Hilbert algebra and that U(A) = U(Ab), etc. It turns out that the only subtle point is the fact that HA = HAb! (cid:3) Lemma 6.3. Suppose X ⊆ Y ⊆ X† as pre-Hilbert B-modules where B is a von Neumann algebra. Then, in fact, X† = Y†. 31 Proof. If θ ∈ X† then y 7→ hθ, yiX† : Y → B is a bounded B-module map and so there is a unique θ ∈ Y† so that: (1) That is, θ 7→ θ embeds X† in Y†. We first show that this embedding preserves inner products. Now, given η ∈ X†, then θ 7→ hη, θiY† : X† → B is an element of X†† = X† and so there hθ, yiY† = hθ, yiX† f or all y ∈ Y. exists a unique γ ∈ X† so that hγ, θiX† = hη, θiY† f or all θ ∈ X†. (2) In particular, for all x ∈ X we get hγ, xiX† = hη, xiY† = hη, xiX† by equation (1). Hence, γ = η, and equation (2) becomes: hη, θiX† = hη, θiY† f or all η, θ ∈ X†. That is, X† is a pre-Hilbert B-submodule of Y† and we have: as pre-Hilbert B-modules. Y ⊆ X† ⊆ Y† Now, for each µ ∈ Y† the map θ 7→ hµ, θiY† : X† → B defines a unique element µ ∈ X† satisfying: hµ, θiY† = hµ, θiX† = hµ, θiY† f or all θ ∈ X†. µ = µ. But since Y ⊆ X† we must have That is,: X† → Y† is onto. Proposition 6.4. Let A be a Z-Hilbert algebra over the abelian von Neumann algebra Z. Then, Ab is also a Z-Hilbert algebra and (1) HAb = HA, (2) U(Ab) = U(A) and V(Ab) = V(A), (3) (Ab)b = Ab. Proof. Since Z ⊆ L(HA) and π(Ab) is a left ideal in L(HA), we see that Ab is a pre-Hilbert Z-submodule of HA containing A. Hence, by the previous lemma, HAb = HA. (cid:3) Thus, axioms (i), (ii), (iii), and (iv) are automatically satisfied. 32 That Ab is a ∗-algebra follows from Proposition 4.6. Now, axiom (v) follows from Lemma 4.3. Axiom (vi) follows from part (4) of Proposition 4.6 since π(ξ∗) = π(ξ)∗ for ξ ∈ Ab. Axiom (vii) follows from the definition of Ab and part (3) of Proposition 4.6. To see axiom (viii), we first note that A2 ⊆ A2 b ⊆ Ab ⊆ HAb = HA. Since A2 is dense in A by definition and A is dense in HA by Proposition 3.3, it follows that A2 b is dense in HAb and hence in Ab. Thus, Ab is also a Z-Hilbert algebra, and items (2) and (3) follow easily. (cid:3) 7. Z-HILBERT ALGEBRAS from Z-TRACES Here we suppose that φ is a faithful normal semifinite Z-trace (in Dixmier's sense) on the von Neumann algebra A, where Z is a von Neumann subalgebra of the centre of A. We abuse notation and also let φ denote the unique linear extension of the original φ from I+ = {x ∈ A φ(x) ∈ Z+} to the ideal I = spanI+, defined in Proposition 1 of III.4.1 of [Dix]. Then, by I.1.6 of [Dix] the space A = {x ∈ A φ(x∗x) ∈ Z+} is an ideal in A with A2 = I. Proposition 7.1. With the above hypotheses, the ideal A = {x ∈ A φ(x∗x) ∈ Z+} is a Z-Hilbert algebra, with the Z-valued inner product hx, yi = φ(x∗y). Proof. Since A is an ideal in A it is certainly a right Z-module. Axiom (i) is just the statement that φ is faithful. Axiom (ii) follows since the extended φ is clearly self-adjoint. Axiom (iii) follows as the original φ is Z+-linear. To see that Axiom (iv) holds requires a little thought. First, it is clear that span(φ(A2)) is an ideal in Z. Therefore, its u.w.-closure is an ideal in Z of the form EZ for some projection E ∈ Z. If (1 − E) 6= 0 then since φ is semifinite there exists x ∈ A+ with 0 6= x ≤ (1 − E) and φ(x) ∈ Z+ so that x1/2 ∈ A. But then, 0 6= φ(x) = φ((1 − E)x) = (1 − E)φ(x) lies in EZ, a contradiction. Hence E = 1 and the span of the inner products is u.w.-dense in Z. Axiom (v) follows from the tracial property of Proposition 1 of III.4.1 of [Dix]. Axiom (vi) is trivial, and Axiom (vii) is proved as in Example 3.6. To see Axiom (viii) we first show that A is u.w.-dense in A. Now the ultraweak closure of A is an u.w. closed ideal in A and so has the form F A for some projection F in Z(A). If (1 − F ) 6= 0 then since φ is semifinite there exists y ∈ A+ with 0 6= y ≤ (1 − F ) and φ(y) ∈ Z+ so that y1/2 ∈ A. But then y ∈ A and so y ≤ F , a contradiction as y 6= 0. Thus F = 1 and A is u.w.-dense in A. 33 Now, given ω ≥ 0 in the predual of Z, we have that φω := ω ◦ φ is a normal, semifinite trace on A by Proposition 2 of III.4.3 of [Dix]. Moreover, the GNS Hilbert space of the normal representation πω of A induced by φω is the same as the Hilbert space Hω of section 3. For a, b ∈ A, we have πω(a)(b + Nω) = ab + Nω. Since πω is normal, πω(A) is u.w.-dense in πω(A). Therefore, it is also s.o.-dense and hence given any b ∈ A and ǫ > 0 there exists a ∈ A with (cid:3) That is, kab − bkω < ǫ, and Axiom (viii) is satisfied. kπω(a)(b + Nω) − (b + Nω)kω < ǫ. In this setting, each x ∈ A defines an operator, x, on the ideal A = {a ∈ A φ(a∗a) ∈ Z+} via x(a) = xa. Clearly, x is Z-linear, and it is easy to check that x is a bounded Z-module map on A, and therefore extends uniquely to a bounded module map on HA, also denoted by x. As left multiplications commute with right multiplications, we see that x ∈ (π′(A))′ = U(A), by the Commutation Theorem 4.11. Lemma 7.2. Let A be an u.w.-dense ∗-ideal in the von Neumann algebra A. Then, each T ∈ A+ is the increasing limit of a net in A+. Proof. It follows from the proof of Theorem 1.4.2 of [Ped] that {a ∈ A+ kak < 1} is an increasing net in the usual ordering of positive elements and hence converges in A+ by Vigier's Theorem. By the Kaplansky Density Theorem there is a subnet of this one converging ultraweakly to the identity in A, and therefore this net converges ultraweakly to 1 ∈ A. Thus, if T ∈ A+, the net {T 1/2aT 1/2 a ∈ A+ and kak < 1} is an increasing net in A+ converging ultraweakly to T . (cid:3) Theorem 7.3. Let φ be a faithful normal semifinite Z-trace on the von Neumann algebra A, where Z is a von Neumann subalgebra of the centre of A. Let A = {a ∈ A φ(a∗a) ∈ Z+} be the corresponding Z-Hilbert algebra. Then the mapping x 7→ x : A → U(A) is an isomorphism of von Neumann algebras. Proof. It is clear the the mapping is a ∗-homomorphism. Since A is u.w.-dense in A, the mapping is also one-to-one. Hence, it suffices to see that the mapping is onto U(A). So, let T ∈ U(A)+. Since π(A) is an u.w.-dense ∗-ideal in U(A), there is a net, {bα} in A+ with π(bα) increasing to T in U(A) ⊆ L(HA). Since, {bα} is an increasing net in A+ ⊆ A+ bounded by kTk, it converges to an element x ∈ A+. To see that x = T it suffices to see that ω(hT a, ci) = ω(hxa, ci) for all a, c ∈ A and ω ≥ 0 in Z∗. 34 Now, since ω◦ φ is a normal scalar trace on A by Proposition 2 of III.4.3 of [Dix] and since ca∗ ∈ A2 = I is contained in the ideal of definition of this normal scalar trace, the map is a normal (and so u.w.-continuous) linear functional on A. Hence, y 7→ ω ◦ φ(yca∗) : A → C ω(hxa, ci) = ω(φ(a∗xc)) = ω(φ(xca∗)) = lim α ω(φ(bαca∗)) = lim α ω(hπ(bα)a, ci). But, by Proposition 5.1 part (2) this last term equals ω(hT a, ci) since π(bα) uw→ T. (cid:3) 8. The Z-TRACE on the CROSSED PRODUCT von NEUMANN ALGEBRA Let (A, Z, τ, α) be a 4-tuple as in Section 1. We also assume that Z has a faithful state, ω to apply Proposition 2.1 so that ¯ω = ω ◦ τ is a faithful tracial state on A and representing A on the GNS Hilbert space H¯ω we obtain A = A′′ and Z = Z ′′ and a Z-trace ¯τ : A → Z extending τ and an extension of α to an ultraweakly continuous action ¯α : R → Aut(A) which leaves ¯τ invariant. Remark 8.1. The following construction of the Z-trace on the crossed product algebra works in much greater generality: the action of R on A leaving τ invariant can be replaced by an action of a unimodular locally compact group G on A leaving τ invariant. We leave the minor modifications to the interested reader. All the results up to the end of section 8.5 work in this generality. We let AZ denote the C ∗-subalgebra of A generated by A and Z. Clearly, AZ =( nXi=1 aiziai ∈ A, zi ∈ Z)−k·k . It is clear that: (1) AZ contains A and Z and is therefore ultraweakly dense in A. (2) ¯τ : AZ → Z is a faithful, unital Z-trace, and (3) ¯α : R → Aut(AZ) is a norm-continuous action on AZ leaving ¯τ invariant and leaving Z pointwise fixed. Key Idea 7. The introduction of this hybrid algebra AZ allows us to treat Z as scalars and use norm-continuity in most of our calculations. This permits the use of C ∗-algebra crossed products and is a considerable simplification. We note also that one cannot simply use the space of norm-continuous functions Cc(R, A) below since ¯α-twisting the multiplication might take us out of the realm of norm-continuity. However, as a vector space (and pre- Hilbert Z-module), Cc(R, A) will have its uses. With this set-up and notation, we define: Definition 8.2. 35 the space of norm-continuous compactly supported functions from R to AZ. We require norm-continuity so that A becomes a ∗-algebra with the usual ¯α-twisted multiplication: A = Cc(R, AZ), x · y(s) =Z x(t) ¯αt(y(s − t))dt, x∗(s) = ¯αs((x(−s))∗). hx, yi =Z ¯τ (x(s)∗y(s))ds (xz)(s) = x(s)z. and involution: and Z-action: Moreover, A becomes a (right) pre-Hilbert Z-module with the inner product: Axioms (i), (ii), and (iii) are routine calculations. To see axiom (iv) we observe that the set of inner products {hx, yi x, y ∈ A} is exactly equal to Z. It comes as no surprise that A is, in fact, a Z-Hilbert algebra. Remark. We will also have occasion to use the completion of A in the vector-valued Banach L2 norm: kxk2 =(cid:18)Z kx(s)k2ds(cid:19)1/2 . We define this completion to be L2(R, AZ) and observe that since kxkA ≤ kxk2, we have a natural inclusion: Proposition 8.3. With the above inner product and Z-action, the ∗-algebra A is a Z-Hilbert algebra. L2(R, AZ) ֒→ A−k·kA ⊂ HA. Proof. Axioms (v) and (vi) are routine calculations. Since A contains all the scalar-valued functions in Cc(R), it is easy to see that A2 is dense in A in the vector-valued L2 norm: Since kxkA ≤ kxk2 , A2 is dense in A in the Z-Hilbert algebra norm and so axiom (viii) is satisfied by the Remark after Definition 3.5. Axiom (vii) requires a little more thought. We will show that the left regular representaion of the ∗-algebra A on the pre-Hilbert Z-module A is the integrated form of a covariant pair 36 of representations (πA, U) of the system (AZ, R, ¯α) inside the von Neumann algebra, L(HA). To this end we represent AZ on the Z-module A = Cc(R, AZ) via: [πA(a)x](s) = ax(s) f or a ∈ AZ, x ∈ A, s ∈ R. Similarly, we represent R on A via: [Ut(x)](s) = ¯αt(x(s − t)) f or t, s ∈ R, x ∈ A. One easily checks that these are representations as bounded, adjointable Z-module map- pings. Now, for fixed x ∈ A the map t 7→ Ut(x) is k · k2-norm continuous and so k · kA-norm continuous: by item (3) of Proposition 5.1 this easily implies that is an ultraweakly continuous representation. Morever, the following are easily verified: t 7→ Ut : R → L(HA) (1) kπA(a)k ≤ kak for a ∈ AZ, (2) hUt(x), Ut(y)i = hx, yi for t ∈ R, x, y ∈ A, (3) πA(a)∗ = πA(a∗) and U ∗ t = U−t for a ∈ AZ, t ∈ R, and t = πA( ¯αt(a)) for t ∈ R and a ∈ AZ. This is the covariance condition. (4) UtπA(a)U ∗ Combining this covariant pair of representations of the system, (AZ, R, ¯α) in L(A) with the ∗-monomorphism embedding L(A) ֒→ L(HA) (by Corollary 3.7 of [Pa]) we obtain a representation πA × U of the C ∗-algebra AZ ⋊ R in the von Neumann algebra L(HA). One then easily checks that for x ∈ A ⊂ AZ ⋊ R and y ∈ A ⊂ HA that: [(πA × U)(x)(y)] (s) =Z x(t) ¯αt(y(s − t))dt = (x · y)(s). That is, left-multiplication by x on the Z-module A is bounded in the Z-module norm and axiom (vii) is satisfied. Lemma 8.4. If A = Cc(R, AZ) as above, then the following hold. (1) The norm-decreasing embedding: (A,k · k2) → (HA,k · kZ) extends by continuity to a norm-decreasing embedding of L2(R, AZ) into HA. Moreover, L2(R, AZ) is a Z-module and the Z-valued inner product on HA restricts to L2(R, AZ) so that it is, in fact, a pre-Hilbert Z-module. (2) If x ∈ L2(R, AZ) ⊆ HA and y ∈ A then in the Z-Hilbert algebra notation, the element: (cid:3) is identical to the element x · y ∈ L2(R, AZ) given by the twisted convolution: π(x)y := π′(y)x ∈ HA (x · y)(s) =Z x(t) ¯αt(y(s − t))dt. 37 (3) If x, y ∈ L2(R, AZ) and if π(x) and π(y) are bounded, then the operator π(x)∗π(y) is in the ideal of definition of the Z-trace, σ on U(A), and σ[π(x)∗π(y)] = hx, yi =Z ¯τ (x(t)∗y(t))dt. Proof. The first statement of item (1) follows trivially from the inequality kxkA ≤ kxk2. To see the second statement of item (1), suppose {xn} is a sequence in A which is Cauchy in the k·k2 norm and that z ∈ Z. Then kxnz−xmzk2 ≤ kxn−xmk2kzk → 0, so that L2(R, AZ) is a Z-module. Similarly, if {xn} and {yn} are sequences in A which are Cauchy in the k· k2 norm, then by the Cauchy-Schwarz inequality: khxn, yni − hxm, ymik = khxn − xm, yni − hxm, ym − ynik ≤ kxn − xmkAkynkA + kxmkAkym − ynkA ≤ kxn − xmk2kynk2 + kxmk2kym − ynk2. Therefore, the Z-valued inner product on HA restricts to a Z-valued inner product on L2(R, AZ). To see the item (2), let {xn} be a sequence in A with kxn − xk2 → 0. Then: kxn · y − x · ykA ≤ kxn · y − x · yk2 ≤ kxn − xk2kyk1 → 0. On the other hand, since xn and y are both in A we have that π′(y)xn = xn · y by definition and so: kxn · y − π(x)ykA = kπ′(y)xn − π′(y)xkA ≤ kπ′(y)kkxn − xkA ≤ kπ′(y)kkxn − xk2 → 0. So, π(x)y = x · y. Item (3) follows from from the definition of the trace (Theorem 5.5) and item (1). (cid:3) Lemma 8.5. The representation πA : AZ → L(HA) extends to an ultraweakly continuous representation (also denoted πA) of A in L(HA). Proof. We first observe that the space of norm-continuous functions, Cc(R, A) ⊂ HA in a natural way. That is if x ∈ Cc(R, A), then for y ∈ A the map: is a bounded Z-module mapping from A to Z and so defines a unique element in HA. If we abuse notation and denote this element in HA by x, then we get the formula: y 7→Z ¯τ ((x(t))∗y(t))dt hx, yi =Z ¯τ ((x(t))∗y(t))dt. Clearly, A = Cc(R, AZ) ⊂ Cc(R, A) ⊂ HA. The extension of πA to A is now obvious: [πA(a)x](s) = ax(s) f or a ∈ A, x ∈ Cc(R, A), s ∈ R. 38 It is easy to check that this is a well-defined extension to A as Z-module mappings on the Z-submodule Cc(R, A) ⊂ HA. These πA(a) extend uniquely to Z-module mappings on HA since HA is also the Paschke dual of Cc(R, A) by Lemma 6.3. To see that πA : A → L(HA) is normal, it suffices to see that πA(A) is ultraweakly closed in L(HA) by Cor. I.4.1 of [Dix]. To this end, it suffices to see that the unit ball in πA(A) is ultraweakly closed. So, let {an} be a net in A with kank = kπA(an)k ≤ 1 and πA(an) → T ultraweakly in L(HA). Since the unit ball in A is ultraweakly compact we can assume (by choosing a subnet if necessary) that there is an a ∈ A such that an → a ultraweakly. By Proposition 5.1 part (3), we have for all x, y ∈ Cc(R, A) hx, πA(an)yi → hx, T yi ultraweakly in Z. On the other hand, if x = cf and y = bg for c, b ∈ A and f, g ∈ Cc(R) then one easily calculates that: hx, πA(an)yi = ¯τ (anbc∗)Z ¯f (t)g(t)dt which converges ultraweakly in Z to hx, πA(a)yi. Thus, for all such x, y we have: hx, πA(a)yi = hx, T yi. Clearly, the same equation holds for all finite linear combinations of such x and y. Since such combinations are k · k2-dense in Cc(R, A) (and so k · kZ-dense) we have the equation holding for all x, y ∈ Cc(R, A). Hence, for all y ∈ Cc(R, A) we have: πA(a)y = T y. Since πA(a) leaves the pre-Hilbert Z-module Cc(R, A) invariant, Proposition 3.6 of [Pa] implies that T = πA(a) as required. (cid:3) Key Idea 8. Now, the natural embedding of the Z-module, L2(R) ⊗alg AZ into L2(R, AZ) induces an embedding: L2(R, AZ) ֒→ L2(R) ⊗Z AZ where the latter is defined to be the completion of the algebraic tensor product in the pre-Hilbert Z-module norm, [L]. Thus we get a series of inclusions of pre-Hilbert Z-modules each of which is strict unless A is finite- dimensional: L2(R) ⊗alg AZ ⊂ L2(R, AZ) ⊂ L2(R) ⊗Z AZ ⊂ HA. One could insert another (generally strict) series of containments: L2(R) ⊗Z AZ ⊂ L2(R) ⊗Z A ⊂ HA. Or, even the diagram of containments: Cc(R, AZ) = A ∪ ∩ 39 = Cc(R, AZ) Cc(R) ⊗alg AZ ⊂ L2(R) ⊗alg AZ ⊂ L2(R, AZ) In general, one might be able to realize HA as some sort of collection of measurable L2- functions from R into the Z-module HAZ = HA; however, this does not seem particularly useful, so we refrain from exploring this idea further. The important point is that each of these Z-modules has the same Paschke dual HA and so we can define operators in L(HA) by defining bounded adjointable Z-module mappings on any one of them by Corollary 3.7 of [Pa]. Of course any one such operator may or may not leave the other Z-modules invariant. Proposition 8.6. Let A = Cc(R, AZ). Then, (1) For x ∈ A we have π(x) = (πA × U)(x) =R πA(x(t))Utdt, where the integral converges in the norm of L(HA). (2) U(A) = [(πA × U)(AZ ⋊ R)]′′ = [πA(A) ∪ {Ut}t∈R]′′. (3) U(A) = [(πA × U)(A ⋊ R)]′′. Proof. To see item (1) we note that in the proof of Proposition 8.1 it was shown that for x, y ∈ A: π(x)y = (πA × U)(x)y. By Proposition 3.6 of [Pa] this implies that π(x) = (πA × U)(x) as elements of L(HA). The second equality in item (1) is true for any crossed product when x is a compactly supported continuous function from the group into the C ∗-algebra. To see item (2) we first note that by item (1): (πA × U)(AZ ⋊ R) = (πA × U)(Cc(R, AZ))−k·k = (πA × U)(A)−k·k = π(A)−k·k. Hence, Now, by the Commutation Theorem (4.11): U(A) = [π(A)]′′ = [π(A)−k·k]′′ = [(πA × U)(AZ ⋊ R)]′′. and it is an easy calculation that πA(AZ) ⊂ (π′(A))′ . Since the representation πA is ultra- weakly continuous on A and AZ is ultraweakly dense in A we see that: U(A) = (π′(A))′ . πA(A) = πA(AZ)−u.w. ⊂ (π′(A))′ = U(A). 40 It is a straightforward calculation (since the operators Ut leave A invariant) that : {Ut}t∈R ⊂ (π′(A))′ = U(A). Thus, [πA(A) ∪ {Ut}t∈R]′′ ⊂ U(A). On the other hand, if T ∈ [πA(A) ∪ {Ut}t∈R]′, then T ∈ [πA(AZ) ∪ {Ut}t∈R]′ and by the full force of item (1), we see that T ∈ (π(A))′ = U(A)′ by Theorem 4.11. That is, [πA(A) ∪ {Ut}t∈R]′ ⊂ U(A)′ or [πA(A) ∪ {Ut}t∈R]′′ ⊃ U(A) as required. To see item (3), we observe that since A is ultraweakly dense in A, Lemma 8.5 implies that πA(A) = πA(A)′′ ⊂ [(πA × U)(A ⋊ R)]′′. Since {Ut}t∈R ⊂ [(πA × U)(A ⋊ R)]′′, we have by item (2) that U(A) ⊂ [(πA × U)(A ⋊ R)]′′. The other containment is trivial. Definition 8.7. The Induced Representation. Now, there is another representaion of A = Cc(R, AZ) (and hence AZ ⋊ R) on HA which is unitarily equivalent to π = πA × U. In the remainder of the paper we will use the standard notation for this representation, namely Ind : see below. Later when we define the notion of index, we will use the notation Index to avoid confusion. To define the representation Ind we first define a single unitary V ∈ L(HA) via: (cid:3) (V ξ)(t) = ¯α−1 t (ξ(t)) f or ξ ∈ L2(R, AZ). One easily checks that V is a bounded, adjointable, Z-module mapping on the Z-module L2(R, AZ) and therefore on HL2(R,AZ) = HA by the previous remarks. One easily checks that for a ∈ AZ , t ∈ R and ξ ∈ L2(R, AZ) V πA(a)V ∗ = π(a) and V UtV ∗ = λt, where (π(a)ξ)(s) = ¯α−1 s (a)ξ(s) and Another straightforward calculation shows that for x, ξ ∈ A (λtξ)(s) = ξ(s − t). (V π(x)V ∗ξ)(s) =Z ¯α−1 s (x(t))ξ(s − t)dt, and that this formula easily extends to ξ ∈ L2(R, AZ). Now, if x ∈ L2(R, AZ), π(x) is bounded and ξ ∈ A, then using the formula of item (2) in lemma 8.4 one easily calculates that we obtain the same formula, namely (V π(x)V ∗ξ)(s) =Z ¯α−1 s (x(t))ξ(s − t)dt. Since this representation of AZ ⋊ R, x 7→ V π(x)V ∗ is induced from the left multiplication of AZ on itself via the action of R on AZ, we denote it by Ind(x). That is, Ind(x) := V π(x)V ∗. Now, the von Neumann algebra, U(A) contains the representations πA of AZ and U of R which integrate to give the representation π = πA × U of A (and hence of AZ ⋊ R) in U(A). We define the von Neumann algebra 41 M = V U(A)V ∗ in L(HA) which also has centre Z and is unitarily equivalent to U(A) but for which the machinery of Z-Hilbert algebras is not directly applicable. M is generated by the representa- tions, π(·) := V πA(·)V ∗ of AZ and λ(·) := V U(·)V ∗ of R. The integrated representation π× λ is, of course, Ind. The trace on M is denoted by τ and is defined on Mτ := V U(A)σV ∗ via: τ (T ) := σ(V ∗T V ). It follows from item (3) of Lemma 8.4 that if x, y ∈ L2(R, AZ) and if π(x) and π(y) are bounded, then the operator Ind(x)∗Ind(y) is in the ideal of definition of the Z-trace, τ on M, and τ [Ind(x)∗Ind(y)] = τ [V π(x)∗π(y)V ∗] = hx, yi =Z ¯τ (x(t)∗y(t))dt. Definition 8.8. The Hilbert Transform. The Hilbert Transform, HR on L2(R) is defined for ξ ∈ L2(R) by: HR(ξ) = ( ξsgn), where,are the usual Fourier transform and inverse transform and sgn is the usual signum function on R. Then, HR is a self-adjoint unitary, so that H 2 2(HR + 1) is the projection onto the Hardy space, H2(R). By [L], H := HR ⊗ 1 and P := PR ⊗ 1 define bounded adjointable Z-module maps on L2(R)⊗alg AZ (and therefore on HA) with the same properties. That is, H 2 = 1 and P = 1 2(H + 1) satisfies P = P ∗ = P 2. R = 1 and PR := 1 In the lemma below, we identify L2(R) with L2(R) · 1A inside L2(R, AZ). Lemma 8.9. The operators H and P are in M. In fact, if we define for ǫ > 0 the function fǫ in L2(R) ⊂ L2(R, AZ) ⊂ HA via: fǫ(t) = 1 πit f or t ≥ ǫ then the π(fǫ) (technically, π(fǫ · 1A)) are uniformly bounded and as ǫ → 0 Ind(fǫ) = V π(fǫ)V ∗ → H strongly on L2(R) ⊗ ¯AZ, Ind(fǫ) = V π(fǫ)V ∗ → H ultraweakly on HA. and so 42 Proof. It follows from [DM] that left convolution by the functions fǫ, λ(fǫ), are uniformly bounded on L2(R) and converge strongly to HR. It is trivial then that λ(fǫ) ⊗ 1 converges strongly to HR ⊗ 1 on L2(R) ⊗alg AZ. Since these operators are all uniformly bounded, adjointable Z-module maps by [L], we see by the usual δ/3-argument, that their extensions to the completion, L2(R) ⊗Z AZ satisfy: λ(fǫ) ⊗ 1 → HR ⊗ 1 = H strongly on L2(R) ⊗Z AZ. It now follows from item (3) of Lemma 5.1 (with L2(R)⊗Z AZ in place of A) and Key Problem 8 that λ(fǫ) ⊗ 1 → H ultraweakly on HL2(R)⊗ZAZ = HA. It remains to see that λ(fǫ) ⊗ 1 = Ind(fǫ) on HA. Now the former is initially defined on L2(R)⊗alg AZ while the latter is initially defined on V (A) = A. Since they are both defined on the common dense domain Cc(R)⊗ AZ, it suffices to check equality there. This is a trivial calculation. Remark 8.10. It follows from the previous lemma that for ξ ∈ A (cid:3) H(ξ) = norm lim ǫ→0 V π(fǫ)V ∗ξ. And since V π(fǫ)V ∗ξ(s) =Z fǫ(t)ξ(s − t)dt =Zt≥0 1 πit ξ(s − t)dt f or s ∈ R, we can formally write: (Hξ)(s) =Z 1 πit ξ(s − t)dt f or ξ ∈ A and s ∈ R where we understand the integral to be the principal-value integral converging in the norm of HA. 9. The INDEX THEOREM We quickly recap for the benefit of the reader what we've done so far. We begin with a unital C ∗-algebra A and a unital C ∗-subalgebra, Z of the centre of A. We assume that we have a faithful, unital Z-trace τ and a continuous action α : R → Aut(A) leaving τ and hence Z invariant. In short, the 4-tuple (A, Z, τ, α) is our object of study. As Standing Assumptions, we will assume that we have a concrete ∗-representation of A on a Hilbert space H which carries a faithful, unital u.w.-continuous Z-trace ¯τ : A → Z extending τ where as before A and Z denote respectively, the ultraweak closures of A and Z on H. Since A is concretely represented on this Hilbert space, we do not carry a special notation for this representation. Moreover there is an ultraweakly continuous action ¯α : R → Aut(A) If Z has a faithful state, ω then the GNS extending α and leaving ¯τ and Z invariant. representation of the state ¯ω = ω ◦ τ gives us a representation of A satisfying the Standing Assumptions by Proposition 2.1. 43 We defined AZ to be the C ∗-subalgebra of A generated by A and Z, so that ¯α restricts to a norm-continuous action of R on AZ and ¯τ restricts to a faithful, unital Z-trace on AZ . We defined A = Cc(R, AZ) to be a ∗-algebra with the usual ¯α-twisted convolution multiplication. There is a natural (right) pre-Hilbert Z-module structure on A making it into a Z-Hilbert algebra as defined in section 3. We defined HA to be the Paschke dual of all bounded Z-module mappings from A to Z (i.e., all Z-linear "Z-valued functionals" on A). Then L(HA) is a type I von Neumann algebra with centre Z. The point of this set-up is that the von Neumann subalgebra U(A) of L(HA) generated by the left multiplications π(x) of A on HA contains Z in its centre and has a faithful, normal semifinite Z-trace σ, defined on the two-sided ideal, U(A)σ = π(A2 b) via: σ(π(ξη)) = hξ∗, ηi, for ξ, η ∈ Ab the (full) Z-Hilbert algebra of (left) bounded elements in HA. At this point we look at a von Neumann algebra M = V U(A)V ∗ in L(HA) which also contains Z in its centre. M is generated by representations, π(·) := V πA(·)V ∗ of AZ and λ(·) := V U(·)V ∗ of R. The integrated representation π × λ is denoted by Ind. The canonical trace on M is denoted by τ and has domain of definition: Mτ = {S ∈ MS = V π(ξη)V ∗ some ξ, η ∈ Ab}. And for S = V π(ξη)V ∗, In particular, if x, y ∈ L2(R, AZ) with π(x) and π(y) bounded, then the operator Ind(x)∗Ind(y) is in the ideal of definition of the Z-trace, τ on M, and τ (S) = hξ∗, ηi. τ [Ind(x)∗Ind(y)] =Z ¯τ (x(t)∗y(t))dt. Definition 9.1. We consider the semifinite von Neumann algebra, with the faithful, normal, semifinite Z-trace obtained by restricting τ . For a ∈ A we define the Toeplitz operator N := PMP Ta := P π(a)P ∈ N . We recall from Section 1 that δ is the infinitesimal generator of α on A and that a 7→ 1 2πi τ (δ(a)a−1) : dom(δ)−1 → Zsa 44 is a group homomorphism which is constant on connected components and so extends uniquely to a group homomorphism A−1 → Zsa which is constant on connected compo- nents and is 0 on Z −1. With this convention and all the above notation, we state our index theorem. Much of the work that we have done so far is to make sense of the the statement of the following theorem and to make sense of the index calculations of [CMX] and [PhR] in this generality. It is interesting that the conclusions of the theorem are insensitive to the choice of a suitable representation of A satisfying the standing assumptions. In particular, if the representation is chosen using Proposition 2.1, the conclusions of the theorem are insensitive to the choice of a faithful state on Z. Theorem 9.2. Let A be a unital C ∗-algebra and let Z ⊆ Z(A) be a unital C ∗-subalgebra of the centre of A. Let τ : A → Z be a faithful, unital Z-trace which is invariant under a continuous action α of R. Then for any a ∈ A−1 ∩ dom(δ), the Toeplitz operator Ta is Fredholm relative to the trace τ on N = P (Ind(A ⋊ R)′′)P , and τ -Index(Ta) = −1 2πi τ (δ(a)a−1). We follow the second proof of [CMX], Section 25.2 (cf section 3 of [PhR]. Now relative to the decomposition 1 = P + (1 − P ) we see that π(a) =(cid:20) Ta B C D (cid:21) , where and similarly, B = P π(a)(1 − P ) = P [P, π(a)] = 1 2 P [H, π(a)], C = 1 2 [H, π(a)]P. Thus, we are led to calculate the general commutator [H, π(a)] for a ∈ dom(δ). Lemma 9.3. For any a ∈ dom(δ), [H, π(a)] belongs to Mτ where x ∈ C0(R, AZ) ∩ L2(R, AZ) is given by x(t) = αt(a) − a . πit 2. In fact, [H, π(a)] = Ind(x), Proof. Now, Ind(fǫ) converges strongly on A to H, so we easily compute for ξ ∈ A: where [Ind(fǫ), π(a)]ξ = Ind(xǫ)ξ xǫ(t) =(cid:26) αt(a)−a πit 0 t ≥ ǫ else . So, the Ind(xǫ) are uniformly bounded operators that converge pointwise on A to [H, π(a)]. Now, since x(t) → (πi)−1δ(a) as t → 0 and kx(t)k2 ≤ 4kak2 π2t2 , 45 we see that x ∈ C0(R, AZ) ∩ L2(R, AZ). One easily calculates that for ξ ∈ A kInd(x)ξ − Ind(xǫ)ξkZ ≤ kInd(x)ξ − Ind(xǫ)ξk2 → 0, and so Ind(x) and [H, π(a)] agree on A. That is, by the discussion in 8.6, π(x) = V ∗Ind(x)V is left bounded and Ind(x) = [H, π(a)] in L(HA). (cid:3) We want to use the Z-trace version of Hormander's formula (Theorem A3 and Corollary A4 in the Appendix) to calculate the τ -index of the Toeplitz operator Ta as τ ([Ta, Ta−1]). So we are led to examine such commutators in the hopes that they are in fact trace-class (they are). Corollary 9.4. If a, b ∈ dom(δ) we have TaTb−Tab ∈ Mτ∩N = N τ . In particular, if b = a−1 then Ta and Tb are τ -Fredholm operators in N . In general, if ab = ba, then [Ta, Tb] ∈ N τ . Proof. We easily calculate (see cor.3.3 of [PhR]): (1) (2) TaTb − Tab = P π(a)(P − 1)π(b)P = · · · = 1 4 P [H, π(a)][H, π(b)]P which is in Mτ ∩ PMP = N τ . If ab = ba, then [Ta, Tb] = (TaTb − Tab) + (Tba − TbTa) ∈ N τ . (cid:3) Discussion. In the case that a, b ∈ dom(δ) commute we have by equation (1) and a small calculation: (3) (4) [Ta, Tb] = P π(a)(P − 1)π(b)P − P π(b)(P − 1)π(a)P = · · · = P (π(a)H π(b) − π(b)H π(a))P, 1 2 and both of these terms are trace-class. Applying the trace to equation (4) we get: (5) τ ([Ta, Tb]) = 1 2 τ (P (π(a)H π(b) − π(b)H π(a))P ). 46 On the other hand, applying the trace to equation (3) , using the cyclic property of the trace and a little calculation (see [PhR]) we get: (6) τ ([Ta, Tb]) = 1 2 τ ((1 − P )(π(a)H π(b) − π(b)H π(a))(1 − P )). Defining and averaging equations (4) and (6) we get: T := π(a)H π(b) − π(b)H π(a), (7) τ ([Ta, Tb]) = 1 4 τ (P T P + (1 − P )T (1 − P )), and both of these terms are trace-class. Unfortunately, T itself is not usually trace-class. However, T is in Mτ Lemma 9.5. (cf lemma 3.4 of [PhR]) Suppose a, b ∈ dom(δ) and ab = ba. Then 2 by the following lemma. T = π(a)H π(b) − π(b)H π(a) belongs to Mτ given by y(t) = (πit)−1(aαt(b) − bαt(a)). 2; in fact it has the form Ind(y) where y is the function in C0(R, AZ)∩L2(R, AZ) Proof. It is straightforward to verify that we can also write: T = [H, π(b)]π(a) − [H, π(a)]π(b). Then by Lemma 9.3 we see that T = Ind(y) where y(t) = (αt(b) − b)αt(a) πit − (αt(a) − a)αt(b) πit = aαt(b) − bαt(a) πit . Since y(t) → (πi)−1(δ(b)a − δ(a)b) in the norm of A as t → 0, y is a continuous A-valued function. As ky(t)k ≤ 2kakkbk/πt for t 6= 0, we also see that y ∈ L2(R, AZ). Remark. In the previous lemma y(0) = (πi)−1(δ(b)a − δ(a)b) = −2(πi)−1δ(a)b. Combining this with equation (7) of the previous discussion would yield the desired formula: (cid:3) τ ([Ta, Tb]) = −1 2πi τ (δ(a)b), assuming that the operator T is trace-class. Since T is generally not trace-class, we need an approximate identity argument. Lemma 9.6. If S ∈ Mτ and {fn} is a sequence of functions in Cc(R)+ ⊂ Cc(R, AZ) each having integral 1 and symmetric supports about 0 shrinking to 0 then τ (S) = uw lim n→∞ τ (Ind(fn)S). 47 Proof. As in the proof of Lemma 8.8, we see that the operators, Ind(fn) = V π(fn)V ∗ are uniformly bounded on HA by 1 and converge strongly to 1 on L2(R)⊗ ¯AZ. In particular, for all x, y ∈ A we have by Paschke's Cauchy-Schwarz inequality (Propn. 2.3 of [Pa]): τ [Ind(x)Ind(y)] = hx∗, yi = hy∗, xi = norm lim n→∞hy∗, π(fn)xi = norm lim = norm lim n→∞ n→∞h(fnx)∗, yi = norm lim n→∞ τ [Ind(fn)Ind(x)Ind(y)]. τ [Ind(fnx)Ind(y)] Now, by item (3) of Lemma 5.1 we see that for all ξ, η ∈ Ab: τ [Ind(ξ)Ind(η)] = uw lim n→∞ τ [Ind(fn)Ind(ξ)Ind(η)]. Since every S ∈ Mτ has the form S = Ind(ξ)Ind(η) for some ξ, η ∈ Ab, we are done. Proposition 9.7. If a, b ∈ dom(δ) and ab = ba, then [Ta, Tb] ∈ N τ and τ [Ta, Tb] = −1 2πi τ (δ(a)b). (cid:3) Proof. Let {fn} be as in the previous lemma. Then, by equation (7) of the Discussion, the previous two lemmas, and the fact that Ind(fn)P = P Ind(fn) we get: τ ([Ta, Tb]) = 1 4 τ (P T P + (1 − P )T (1 − P )) τ (Ind(fn)(PTP + (1 − P)T(1 − P))) τ (Ind(fn)PTP + Ind(fn)(1 − P)T(1 − P)) τ (PInd(fn)TP + (1 − P)Ind(fn)T(1 − P)) τ (PInd(fn)T + (1 − P)Ind(fn)T) τ (Ind(fn)T) = uw lim = uw lim = uw lim = uw lim = uw lim = uw lim 1 4 1 4 1 4 1 4 1 4 1 4 1 τ (Ind(fn)Ind(y)) a − In fact, this last limit is easily seen to converge in norm, so that = uw lim 4πiZ fn(t)τ(cid:18) αt(b) − b t αt(a) − a t b(cid:19) dt. τ ([Ta, Tb]) = 1 4πi = −1 2πi τ (δ(b)a − δ(a)b) τ (δ(a)b). where and, B = P π(a)(1 − P ) = P [P, π(a)] = 1 2 P [H, π(a)] ∈ Mτ 2, π(a) =(cid:20) Ta B C D (cid:21) , C = 1 2 [H, π(a)]P ∈ Mτ 2. 48 (cid:3) Proof of Theorem 9.2. Recall that relative to the decomposition 1 = P +(1−P ) we have: By Corollary A4 of the Appendix and the previous proposition we have: τ -Index(Ta) = τ ([Ta, Ta−1]) = −1 2πi τ (δ(a)a−1). This completes the proof of Theorem 9.2. Corollary 9.8. If ϕ : A1 → A2 defines a morphism from (A1, Z1, τ1, α1) to (A2, Z2, τ2, α2) and if a ∈ A−1 1 ∩ (dom(δ1)) then ϕ(a) ∈ A−1 τ1-Index(Ta) ∈ (Z1)sa while τ2-Index(Tϕ(a)) ∈ (Z2)sa and also 2 ∩ (dom(δ2)) and (cid:3) ϕ(τ1-Index(Ta)) = τ2-Index(Tϕ(a)). Proof. This follows immediately from Proposition 1.3 and Theorem 9.2. (cid:3) 10. EXAMPLES 1. Kronecker (scalar trace) Example. Recall: A = C(T2), the C ∗-algebra of con- tinuous functions on the 2-torus, with the usual scalar trace τ given by the Haar measure on T2, and α : R → Aut(A) is the Kronecker flow on A determined by the real number, µ. That is, for s ∈ R, f ∈ A, and (z, w) ∈ T2 we have: (αs f )(z, w) = f(cid:0)e−2πis z, e−2πiµs w(cid:1) . In this case, Z = Z = C and so AZ = A. Hence our Z-Hilbert algebra A = Cc(R, A) is just a Hilbert algebra in the ordinary sense and HA = L2(R, L2(T2)). Now, denoting H = L2(T2), we have that the C ∗-crossed product A ⋊α R is represented on L2(R,H) by the induced representation of Definition 8.7 as follows: for s, t ∈ R, ξ ∈ Cc(R, A) ⊆ L2(R,H) and f ∈ A we define (π(f ) ξ) (s) = α−1 s (f ) · ξ(s) and (λt ξ) (s) = ξ(s − t). 49 Thus, π × λ is a faithful representation of A ⋊α R on L2(R,H). It is well-known that for µ irrational, M = (π × λ(A ⋊α R))′′ is a II∞ factor, [CMX]. In general M is a semifinite von Neumann algebra and π : A → M. Now, if δ is the densely defined derivation on A generating the representation α : R → Aut(A) and we let u ∈ U(A) be the function u(z, w) = w then u is a smooth element for δ and δ(u) = −(2πiµ)u. Thus by Theorem 9.2, the Toeplitz operator Tu := P π(u)P is Fredholm relative to the trace τ in the semifinite von Neumann algebra, N = PMP and its index is given by: τ -Index(Tu) = −1 2πi τ (δ(u)u∗) = µ. 2. General Kronecker Examples. Recall Z = C(X) is any commutative unital C ∗-algebra with a faithful state ω and θ ∈ Zsa is any self-adjoint element in Z. Recall A = C(T2, Z) = C(X) ⊗ C(T2), and τ : A → Z is given by the "slice-map" τ = idZ ⊗ ϕ where ϕ is the trace on C(T2) given by Haar measure. That is, for f ∈ A = C(T2, Z) we have τ (f ) =ZT2 f (z, w)d(z, w) ∈ Z, and τ is a faithful, tracial conditional expectation of A onto Z. Recall that ¯ω := ω◦τ = ω⊗ϕ is a faithful (tracial) state ¯ω on A. We use the element θ ∈ Zsa to define a τ -invariant action {αt} of R on A: αt(f )(x, z, w) = f (x, e−2πitz, e−2πiθ(x)tw), for f ∈ A, t ∈ R, x ∈ X, and z, w ∈ T. Let (π,H) be the GNS representation of A induced by ¯ω then there is a continuous unitary representation {Ut} of R on H so that (π, U) is covariant for α on A. Also, {Ut} implements an uw-continuous "extension" of α to ¯α acting on A := π(A)′′. Morover, letting Z := π(Z)′′, there exists a unique faithful unital, uw-continuous Z-trace ¯τ : A → Z "extending" τ, and ¯α leaves ¯τ invariant. That is, in this representation on H, we have that Standing Assump- tions are also satisfied. We simplify our notation and write L2(X), L2(T2), L∞(X), and L∞(T2) for L2(X, ω), L2(T2, ϕ), L∞(X, ω), and L∞(T2, ϕ), respectively. Then, in this representation, one easily verifies that: H = L2(X) ⊗ L2(T2), as Hilbert spaces, and Z = L∞(X) ⊗ 1, and AZ = L∞(X) ⊗ C(T2) as C ∗ − algebras, and A = L∞(X) ¯⊗L∞(T2) as von Neumann algebras. Identifying Z = L∞(X), our L∞(X)-Hilbert algebra is A = Cc(R, L∞(X) ⊗ C(T2)) with the ¯α-twisted convolution multiplication and L∞(X)-valued inner product for f, g ∈ A given 50 by: ¯τ ((f (t))∗g(t))dt τ (Ind(f )∗Ind(g)) = hf, gi =ZR = ZR(cid:18)ZT2 (f (t)[(z, w)])∗g(t)[(z, w)]d(z, w)(cid:19) dt. αt(v)(x, z, w) = e−2πiθ(x)tw and so δ(v)(x, z, w) = −2πiθ(x)w. Now, consider the following unitary v in A: v(x, z, w) = w. Then Hence, (δ(v)v∗)(x, z, w) = −2πi· θ(x). Since the trace τ on A is just the slice map idZ ⊗ ϕ we see that τ (δ(v)v∗) = −2πi · θ. Hence, by Theorem 9.2, the Toeplitz operator Tv is Fredholm relative to the trace τ on N = P (Ind(A ⋊ R)′′)P , and τ -Index(Tv) = −1 2πi τ (δ(v)v∗) = θ ∈ C(X) = Z ֒→ Z ⊗ C(T2) = A. 3. Fiberings of Toeplitz operators. Recall that for any fixed x ∈ X (where Z = C(X)) the evaluation map at x yields a homomorphism from A = Z ⊗ C(T2) to C(T2) which defines a morphism from Example 2 to Example 1 which carries θ to µ := θ(x). Moreover this morphism carries v to u = v(x). So that Index(Tu) = µ = θ(x) = (Index(Tv))(x). That is, the Toeplitz operator Tv fibers over X as the Toeplitz operators Tθ(x) and moreover for each x ∈ X: Index(Tv(x)) = (Index(Tv))(x). so the Index fibers accordingly. Similarly, for any fixed x ∈ X (where Z = C(X)) the evaluation map at x yields a homo- morphism from A = Z ⊗ Aθ to Aθ which defines a morphism from (Z ⊗ Aθ, Z, id ⊗ τθ, αη) to (Aθ, C, τθ, αη(x)). This morphism carries 1 ⊗ V to V . Since Index(T1⊗V ) = η and Index(TV ) = η(x) we see that: Index(T1⊗V )(x) = Index(TV ) = Index(T1⊗V (x)). 4. C ∗-algebra of the Integer Heisenberg group. Recall that A = C ∗(H) is the C ∗-algebra of the Integer Heisenberg group viewed as the universal C ∗-algebra generated by three unitaries U, V, W satisfying: W U = UW, W V = V W, and UV = W V U. In this case Z = C ∗(W ) is the centre of A and also equals C ∗(C) the C ∗-algebra generated by C = hWi the centre of H. The trace τ : C ∗(H) → C ∗(C) on functions in l1(H) ⊂ C ∗(H) is just given by restriction to C. Our Hilbert space H = l2(H) acted on by the left regular representation of C ∗(H). The restriction of this action to Z = C ∗(C) on l2(H) = M(n,m)∈Z2 l2(C · (V nU m)) 51 is unitarily equivalent to 1Z2 ⊗ πC(C) on L(n,m)∈Z2 l2(C). In this labelling of the cosets, multiplication by W acts the same on each coset: it increases the power of W by one. Multiplication by V acts as the identification of l2(C · (V nU m)) with l2(C · (V n+1U m)): that is, it acts as a permutation of the copies of l2(C) while acting on the basis elements as the identity on l2(C). However, multiplication by U not only maps l2(C · (V nU m)) to l2(C · (V nU m+1)), but it also acts on the basis elements of l2(C) by sending W k to W k+1. In this representation we recall that the map τ : C ∗(H) → C ∗(C) is given by τ (x) = 1Z2⊗ExE, where E is the projection of l2(H) onto l2(C). Thus we have an action α : R → Aut(A), that fixes Z = C ∗(W ) and leaves the Z-valued trace τ invariant. A short calculation using Theorem 9.2 then gives us the nontrivial index: τ -Index(TV nU mW p) = (nθ + m) ∈ Z = C ∗(W ). APPENDIX: FREDHOLM THEORY RELATIVE to a Z-VALUED TRACE on a von NEUMANN ALGEBRA We let N denote a semifinite von Neumann algebra and let Z denote a unital von Neumann subalgebra of the centre of N . We suppose that we have a faithful, normal, semifinite Z-trace φ defined on N+ as in Definition 6.1. We will show that using φ as a dimension function we can adapt M. Breuer's arguments in [Br1], [Br2] to obtain a Fredholm theory involving a Z-valued index with the usual algebraic and topological stability properties, and in which the role of the compact operators is replaced by the norm-closed ideal, Kφ N generated by the projections of φ-finite trace. A projection E in N will be called φ-f inite if φ(E) ∈ Z+. Since φ is faithful, it is clear that any φ-finite projection is also finite in the Murray-von Neumann sense. An operator T ∈ N is called φ-F redholm if the projection NT on ker(T ) is φ-finite and there is a φ-finite projection E with range(1 − E) ⊆ range(T ). Since φ-finite projections are finite, every φ- Fredholm operator is Fredholm in Breuer's sense. If T is φ-Fredholm, the φ-index of T is by definition φ-Index(T ) := φ(NT ) − φ(NT ∗) : we shall see below that T ∗ is also φ-Fredholm so that φ-Index(T ) is a well-defined self-adjoint element of Z. We observe as we did in [PhR] that the ideal Kφ N can also be described as the closure of any of: (1) the span of the φ-finite projections in N , (2) the span of the φ-finite elements in N , (3) the algebra of elements T ∈ N whose range projection RT is φ-finite. This ideal is clearly contained in Breuer's ideal K generated by all the finite projections in N . 52 Now the further remarks and proofs concerning how Breuer's arguments carry over to this situation follow verbatim from Appendix B of [PhR]. So, we obtain the analogues of Breuer's theorems exactly as we did in [PhR]. Theorem (A1). Let φ be a faithful, normal, semifinite Z-trace on the von Neumann algebra N , and let Kφ N be the norm-closed ideal in N generated by the φ-finite projections. (1) (The Fredholm alternative) If T ∈ Kφ N , then (1 − T ) is φ-Fredholm and φ-Index(1 − T ) = 0. (2) (Atkinson's Theorem) An operator T ∈ N is φ-Fredholm if and only if T + Kφ N is invertible in N /Kφ N . (3) If S and T are φ-Fredholm, then so are S∗ and ST , and φ-Index(S∗) = −(φ-Index(S)), φ-Index(ST ) = φ-Index(S) + φ-Index(T ). The following corollary is proved exactly as Corollary B2 of [PhR]. Corollary (A2). The set Fφ(N ) of φ-Fredholm operators is open in the norm topology of N , and the index map T 7→ φ-Index(T ) is locally constant on Fφ(N ). The following trace formula for the index goes back to Calder´on for pseudodifferential operators. The general type I case is due to Hormander [H] but Connes also has an elegant proof [Co]. One of the authors generalised Hormander's proof to the case of a factor of type II∞ in [Ph], Theorem A7. It is this latter proof that goes through essentially verbatim to our present setting, so we refer the reader to Appendix A of [Ph] for the proof. Theorem (A3). Let φ be a faithful, normal, semifinite Z-trace on the von Neumann algebra N , and let S, T ∈ N so that R1 = 1 − ST and R2 = 1 − T S are both n-summable for some integer n > 0. Then, T is a φ-Fredholm operator and φ-Index(T ) = φ(Rn 1 ) − φ(Rn 2 ). Corollary (A4). Let A be a unital C ∗-algebra and let Z ⊆ Z(A) be a unital C ∗-subalgebra of the centre of A. Let τ : A → Z be a faithful, unital Z-trace which is invariant under a continuous action α of R. Then for any a ∈ A−1 ∩ dom(δ), the Toeplitz operator Ta is Fredholm relative to the trace τ on N = P (Ind(A ⋊ R)′′)P , and τ -Index(Ta) = τ ([Ta, Ta−1]). Proof. We let T = Ta and S = Ta−1 and φ = τ in the statement of the previous theorem. Then, R1 = 1 − Ta−1Ta = Ta−1a − Ta−1Ta ∈ N τ by Corollary 9.4 and similarly, R2 ∈ N τ . Then, by the previous theorem, Ta is τ -Fredholm and τ -Index(Ta) = τ (R1) − τ (R2) = τ ([Ta, Ta−1]). (cid:3) References 53 [AP] J. Anderson and W. Paschke, The Rotation Algebra, Houston J. Math. 15 (1989), 1-26. [Arv] W. Arveson, Subalgebras of C ∗-algebras, Acta Math., 123 (1969), 141-224. [Br1] M. Breuer, Fredholm Theories in von Neumann Algebras, I, Math. Ann., 178 (1968), 243-254. [Br2] M. Breuer, Fredholm Theories in von Neumann Algebras, II, Math. Ann., 180 (1969), 313-325. [Co] A. Connes, Noncommutative Differential Geometry, Publ. Math. Inst. Hautes Etudes Sci., 62 (1985), 41-144. [CMX] R. Curto, P. S. Muhly, and J. Xia, Toeplitz operators on flows, J. Functional Analysis, 93 (1990), 391-450. [Dix] J. Dixmier, Les alg`ebres d'op´erateurs dans l'espace Hilbertien (Alg`ebres de von Neumann), Gauthier- Villars, Paris, 1969. [DM] H. Dym and H.P. McKean, Fourier Series and Integrals, Academic Press, New York, London, 1972. [H] L. Hormander, The Weyl Calculus of Pseudodifferential Operators, Comm. Pure Appl. Math., 32 (1979), 359-443. [Ji] R. Ji,Toeplitz Operators on Noncommutative Tori and Their Real-valued Index, Proc. Symp. Pure [K] [L] Math. (Amer. Math. Soc.), vol. 51, Part 2 (1990), pages 153-158. I. Kaplansky, Modules over operator algebras, Amer. J. Math., 75 (1953), 839-858. E. C. Lance, Hilbert C ∗-modules, London Math. Soc. Lecture Notes Series 210, Cambridge University Press, Cambridge, 1995. [Le] M. Lesch, On the Index of the Infinitesimal Generator of a Flow, J. Operator Theory, 26 (1991), 73-92. [PR] J.A. Packer and Iain Raeburn, On the Structure of Twisted Group C ∗-algebras, Trans. Amer. Math. Soc. 334 (1992), 685-718. [Pa] W. Paschke, Inner Product Modules Over B ∗-algebras, Trans. Amer. Math. Soc., 182 (1973), 443-468. [Ped] G.K. Pedersen, C ∗-Algebras and their Automorphism Groups, Academic Press, London, 1979. [Ph] John Phillips, Spectral Flow in Type I and II Factors -- A New Approach, Fields Inst. Comm., 17 (1997), 137-153. [PhR] John Phillips and Iain Raeburn, An Index Theorem for Toeplitz Operators with Noncommutative Symbol Space, J. Functional Analysis, 120 (1994), 239-263. [R] M. Rieffel, Morita Equivalence for C ∗-algebras and W ∗-algebras, J. Pure and Applied Algebra, 5 (1974), 51-96. J. Tomiyama, On the projection of norm one in W ∗-algebras, Proc. Japan Acad., 33 (1957), 608-612. [T] [U] H. Umegaki, Conditional expectation in an operator algebra I, Tohoku Math. J., 6 (1954), 358-362.
1107.5153
1
1107
2011-07-26T09:59:17
Topologies on Central Extensions of Von Neumann Algebras
[ "math.OA" ]
Given a von Neumann algebra $M$ we consider the central extension $E(M)$ of $M.$ We introduce the topology $t_c(M)$ on $E(M)$ generated by a center-valued norm and prove that it coincides with the topology of convergence locally in measure on $E(M)$ if and only if $M$ does not have direct summands of type II. We also show that $t_c(M)$ restricted on the set $E(M)_h$ of self-adjoint elements of $E(M)$ coincides with the order topology on $E(M)_h$ if and only if $M$ is a $\sigma$-finite type I$_{fin}$ von Neumann algebra.
math.OA
math
United Nations Educational, Scientific and Cultural Organization and International Atomic Energy Agency THE ABDUS SALAM INTERNATIONAL CENTRE FOR THEORETICAL PHYSICS TOPOLOGIES ON CENTRAL EXTENSIONS OF VON NEUMANN ALGEBRAS Sh. A. Ayupov 1 Institute of Mathematics and Information Technologies, Uzbekistan Academy of Sciences Dormon yoli str., 29, 100125, Tashkent, Uzbekistan and The Abdus Salam International Centre for Theoretical Physics, Trieste, Italy Department of Mathematics, Karakalpak state university Ch. Abdirov str.,1, K. K. Kudaybergenov 2 230113, Nukus, Uzbekistan and R. T. Djumamuratov 3 Department of Mathematics, Karakalpak state university Ch. Abdirov str.,1, 230113, Nukus, Uzbekistan Abstract Given a von Neumann algebra M we consider the central extension E(M ) of M. We introduce the topology tc(M ) on E(M ) generated by a center-valued norm and prove that it coincides with the topology of convergence locally in measure on E(M ) if and only if M does not have direct summands of type II. We also show that tc(M ) restricted on the set E(M )h of self-adjoint elements of E(M ) coincides with the order topology on E(M )h if and only if M is a σ-finite type If in von Neumann algebra. MIRAMARE -- TRIESTE 1Senior Associate of ICTP. Corresponding author. sh [email protected] [email protected] [email protected] 1 1 1 0 2 l u J 6 2 ] . A O h t a m [ 1 v 3 5 1 5 . 7 0 1 1 : v i X r a 1 Introduction In the series of paper [1]-[5] we have considered derivations on the algebra LS(M) of locally measurable operators affiliated with a von Neumann algebra M, and on various subalgebras of LS(M). A complete description of derivations has been ob- tained in the case of von Neumann algebras of type I and III. A comprehensive survey of recent results concerning derivations on various algebras of unbounded operators affiliated with von Neumann algebras is presented in [4]. A general form of automorphisms on the algebra LS(M) in the case of von Neumann algebras of type I has been obtained in [5]. In proof of the main results of the above papers the crucial role is played by the co-called central extensions of von Neumann algebras and also by various topologies considered in [3]. Let M be an arbitrary von Neumann algebra with the center Z(M) and let LS(M) denote the algebra of all locally measurable operators with respect M. We consider the set E(M) of all elements x from LS(M) for which there exists a sequence of mutually orthogonal central projections {zi}i∈I in M with W zi = 1, such that zix ∈ M for all i ∈ I. It is known [3] that E(M) is a *-subalgebra in LS(M) with the center S(Z(M)), where S(Z(M)) is the algebra of all measurable operators with respect to Z(M), moreover, LS(M) = E(M) if and only if M does not have direct summands of type II. i∈I A similar notion (i.e. the algebra E(A)) for arbitrary *-subalgebras A ⊂ LS(M) was independently introduced by M.A. Muratov and V.I. Chilin [7]. The algebra E(M) is called the central extension of M. It is known ([3], [7]) that an element x ∈ LS(M) belongs to E(M) if and only if there exists f ∈ S(Z(M)) such that x ≤ f. Therefore for each x ∈ E(M) one can define the following vector-valued norm x = inf{f ∈ S(Z(M)) : x ≤ f }. This center-valued norm naturally generates a topology on E(M) which denoted by tc(M). In this paper we study the relationship between the topology tc(M) on E(M) generated by the above center-valued norm, the topology t(M) -- of convergence locally in measure, and the order topology to(M) on E(M)h. We prove that tc(M) coincides with the topology t(M) on E(M) if and only if M does not have direct summands of type II. We show that tc(M) coincides with the order topology on E(M)h if and only if M is a σ-finite type If in algebra. 2 Central extensions of von Neumann algebras Let H be a complex Hilbert space and let B(H) be the algebra of all bounded linear operators on H. Consider a von Neumann algebra M in B(H) with the operator norm k · kM . Denote by P (M) the lattice of projections in M. A linear subspace D in H is said to be affiliated with M (denoted as DηM), if 2 u(D) ⊂ D for every unitary u from the commutant M ′ = {y ∈ B(H) : xy = yx, ∀x ∈ M} of the von Neumann algebra M. A linear operator x on H with the domain D(x) is said to be affiliated with M (denoted as xηM) if D(x)ηM and u(x(ξ)) = x(u(ξ)) for all ξ ∈ D(x). A linear subspace D in H is said to be strongly dense in H with respect to the von Neumann algebra M, if 1) DηM; 2) there exists a sequence of projections {pn}∞ n=1 in P (M) such that pn ↑ 1, n = 1 − pn is finite in M for all n ∈ N, where 1 is the identity in pn(H) ⊂ D and p⊥ M. A closed linear operator x acting in the Hilbert space H is said to be measurable with respect to the von Neumann algebra M, if xηM and D(x) is strongly dense in H. Denote by S(M) the set of all measurable operators with respect to M (see [10]). A closed linear operator x in H is said to be locally measurable with respect to n=1 of central the von Neumann algebra M, if xηM and there exists a sequence {zn}∞ projections in M such that zn ↑ 1 and znx ∈ S(M) for all n ∈ N (see [11]). It is well-known [6], [11] that the set LS(M) of all locally measurable operators with respect to M is a unital *-algebra when equipped with the algebraic operations of strong addition and multiplication and taking the adjoint of an operator, and contains S(M) as a solid *-subalgebra. Let (Ω, Σ, µ) be a measure space and from now on suppose that the measure µ has the direct sum property, i. e. there is a family {Ωi}i∈J ⊂ Σ, 0 < µ(Ωi) < ∞, i ∈ J, such that for any A ∈ Σ, µ(A) < ∞, there exist a countable subset J0 ⊂ J and a set B with zero measure such that A = S (A ∩ Ωi) ∪ B. i∈J0 We denote by L0(Ω, Σ, µ) the algebra of all (equivalence classes of) complex measurable functions on (Ω, Σ, µ) equipped with the topology of convergence in measure. Consider the algebra S(Z(M)) of operators which are measurable with respect to the center Z(M) of the von Neumann algebra M. Since Z(M) is an abelian von Neumann algebra it is *-isomorphic to L∞(Ω, Σ, µ) for an appropriate measure space (Ω, Σ, µ). Therefore the algebra S(Z(M)) coincides with Z(LS(M)) and can be identified with the algebra L0(Ω, Σ, µ) of all measurable functions on (Ω, Σ, µ). The basis of neighborhoods of zero in the topology of convergence locally in measure on L0(Ω, Σ, µ) consists of the sets W (A, ε, δ) = {f ∈ L0(Ω, Σ, µ) : ∃B ∈ Σ, B ⊆ A, µ(A \ B) ≤ δ, f · χB ∈ L∞(Ω, Σ, µ), f · χBL∞(Ω,Σ,µ) ≤ ε}, 3 where ε, δ > 0, A ∈ Σ, µ(A) < +∞, and χB is the characteric function of the set B ∈ Σ. Recall the definition of the dimension functions on the lattice P (M) of projection from M (see [6], [10]). By L+ we denote the set of all measurable functions f : (Ω, Σ, µ) → [0, ∞] (modulo functions equal to zero µ-almost everywhere ). Let M be an arbitrary von Neumann algebra with the center Z(M) ≡ L∞(Ω, Σ, µ). Then there exists a map D : P (M) → L+ with the following properties: (i) d(e) is a finite function if only if the projection e is finite; (ii) d(e + q) = d(e) + d(q) for p, q ∈ P (M), eq = 0; (iii) d(uu∗) = d(u∗u) for every partial isometry u ∈ M; (iv) d(ze) = zd(e) for all z ∈ P (Z(M)), e ∈ P (M); (v) if {eα}α∈J , e ∈ P (M) and eα ↑ e, then d(e) = sup α∈J d(eα). This map d : P (M) → L+, is a called the dimension functions on P (M). Recall that for an element x ∈ LS(M) the projection defined as c(x) = inf{z ∈ P (Z(M)) : zx = x} is called the central cover of x. Remark 2.1. Let M be a type I von Neumann algebra. If p, q ∈ P (M) abelian projections are faithful (i.e. with c(p) = c(q) = 1,) then the property (iii) implies that 0 < d(p)(ω) = d(q)(ω) < ∞ for µ-almost every ω ∈ Ω. Therefore replacing d by d(p)−1d we can assume that d(p) = c(p) for every faithful abelian projection p ∈ P (M). Thus for all e ∈ P (M) we have that d(e) ≥ c(e). The basis of neighborhoods of zero in the topology t(M) of convergence locally in measure on LS(M) consists (in the above notations) of the following sets V (A, ε, δ) = {x ∈ LS(M) : ∃p ∈ P (M), ∃z ∈ P (Z(M)), xp ∈ M, xpM ≤ ε, z⊥ ∈ W (A, ε, δ), d(zp⊥) ≤ εz}, where ε, δ > 0, A ∈ Σ, µ(A) < +∞. The topology t(M) is metrizable if and only if the center Z(M) is σ-finite (see [6]). Given an arbitrary family {zi}i∈I of mutually orthogonal central projections in zi = 1 and a family of elements {xi}i∈I in LS(M) there exists a unique M with W element x ∈ LS(M) such that zix = zixi for all i ∈ I. This element is denoted by x = P i∈I zixi. i∈I 4 We denote by E(M) the set of all elements x from LS(M) for which there exists zi = 1, a sequence of mutually orthogonal central projections {zi}i∈I in M with W such that zix ∈ M for all i ∈ I, i.e. i∈I E(M) = {x ∈ LS(M) : ∃zi ∈ P (Z(M)), zizj = 0, i 6= j, _ zi = 1, zix ∈ M, i ∈ I}, i∈I where Z(M) is the center of M. It is known [3] that E(M) is *-subalgebras in LS(M) with the center S(Z(M)), where S(Z(M)) is the algebra of all measurable operators with respect to Z(M), moreover, LS(M) = E(M) if and only if M does not have direct summands of type II. A similar notion (i.e. the algebra E(A)) for arbitrary *-subalgebras A ⊂ LS(M) was independently introduced recently by M.A. Muratov and V.I. Chilin [7]. The algebra E(M) is called the central extension of M. It is known ([3], [7]) that an element x ∈ LS(M) belongs to E(M) if and only if there exists f ∈ S(Z(M)) such that x ≤ f. Therefore for each x ∈ E(M) one can define the following vector-valued norm x = inf{f ∈ S(Z(M)) : x ≤ f } (2.1) and this norm satisfies the following conditions: 1)kxk ≥ 0; kxk = 0 ⇐⇒ x = 0; 2)kf xk = f kxk; 3)kx + yk ≤ kxk + kyk; 4)xy ≤ xy; 5)xx∗ = x2 for all x, y ∈ E(M), f ∈ S(Z(M)). 3 Topologies on the central extensions of von Neumann al- gebras Let M be an arbitrary von Neumann algebra with the center Z(M) ≡ L∞(Ω, Σ, µ). On the space E(M) we consider the following sets: O(A, ε, δ) = {x ∈ E(M) : x ∈ W (A, ε, δ)} , (3.1) where ε, δ > 0, A ∈ P, µ (A) < +∞. The following proposition gives elementary properties of the sets O(A, ε, δ), which immediately follow from the corresponding properties of the sets V (A, ε, δ) (see [6, Proposition 3.5.1]). Proposition 3.1. Let ε, εj > 0 and δ, δj > 0, j = 1, 2, A ∈ Σ, µ(A) < ∞. Then 5 i) λO(A, ε, δ) = O(A, λε, δ) λ ∈ C, λ 6= 0; ii) O(A, ε1, δ1) ⊆ O(A, ε2, δ2) ε1 ≤ ε2, δ1 ≤ δ2; iii) O(A, ε1, δ1) + O(A, ε2, δ2) ⊆ O(A, ε1 + ε2, δ1 + δ2); iv) O(A, ε1, δ1)O(A, ε2, δ2) ⊆ O(A, ε1ε2, δ1 + δ2); v) O∗(A, ε, δ) = O(A, ε, δ), where O∗(A, ε, δ) = {x∗ : x ∈ O(A, ε, δ)}; vi) T{O(A, ε, δ) : ε > 0, δ > 0, A ∈ Σ, µ(A) < ∞} = {0}. From Proposition 3.1 it follows that the system of sets {x + O(A, ε, δ)}, (3.2) where x ∈ E(M), ε > 0, δ > 0, A ∈ Σ, µ(A) < ∞, defines on E(M), a Hausdorff vector topology tc(M), for which the sets (3.2) form the base of neighborhoods of the element x ∈ E(M). Moreover in this topology the involution is continuous and the multiplication is jointly continuous, i.e. (E(M), tc(M)) is a topological *-algebra. From [4, Proposition 5.3] it follows that (E(M), tc(M)) is complete. Thus we obtain the following result. Proposition 3.2. i) (E(M), tc(M)) is a complete topological *-algebra; ii) M is a tc(M)-dense in E(M). Proof. i) is proved above. ii). Let x ∈ E(M) and A ∈ Σ, µ(A) < ∞, ε, δ > 0. For n ∈ N put Bn = {ω ∈ A : x(ω) ≤ n}. Since µ(A \ Bn) → 0 as n → ∞ there is k ∈ N such that µ(A \ Bk) ≤ ε. Put xk = χBkx. Then xk ≤ k1 and x − xkχBk = xχBk − xkχBk = xχBk − xχBk = 0. Thus xk ∈ x + O(A, ε, δ). This means that M tc(M ) = E(M). The proof is complete. Remark 3.3. Note that if M is a commutative von Neumann algebra then x = x for each x ∈ E(M), and therefore O(A, ε, δ) = W (A, ε, δ) for all ε, δ > 0, A ∈ P, µ (A) < +∞. Hence the topology tc(M) on E(M) coincides with the topology of convergence locally in measure t(M). If M is a factor, then E(M) = M and tc(M) = tk·kM , where tk·kM uniform topology on M. Proposition 3.4. i) A net {pα} ⊂ P (M) converges to zero with respect to the topol- ogy tc(M) if and only if c(pα) locally in measure on Z(M). t(Z(M )) −→ 0, where t(Z(M)) is the topology of convergence 6 ii) A net {xα} ⊂ E(M) converges to zero with respect to the topology tc(M) if and only if e⊥ family for the operator xα. λ (xα) tc(M ) −→ 0 for any λ > 0, where {eλ(xα)} is a spectral projections Proof. i) The proof immediately follows from the definition of the topology tc(M) and the equality p = c(p), p ∈ P (M). ii) Let xα tc(M ) −→ 0 and λ > 0. Take any A ∈ Σ, µ(A) < ∞, 0 < ε < λ/2, δ > 0. t(M ) −→ 0, then there exists α0 such that xα ∈ W (A, ε, δ) for each α ≥ α0. Since xα Therefore there exists Bα ∈ Σ, Bα ⊆ A such that µ(A \ Bα) ≤ δ, xαχBαM ≤ ε. Thus xαχBαM ≤ ε, i.e. 2 then from the last inequality we have that c(e⊥ λ (xα))χBα = 0. The inequality µ(A \ Bα) ≤ δ implies that c(e⊥ xαχBα ≤ εχBα. Since ε < λ λ (xα)) ∈ W (A, ε, δ), i.e. c(e⊥ t(Z(M )) −→ 0. Thus e⊥ tc(M ) −→ 0. λ (xα)) λ (xα) Now let e⊥ ε (xα) tc(M ) −→ 0 and 0 < ε < 1, δ > 0. Then c(e⊥ t(M ) −→ 0. Therefore ε (xα)) ∈ W (A, ε, δ) for all α ≥ α0. Hence there exists ε (xα))χBαM ≤ ε < 1. Thus there exists α0 such that c(e⊥ Bα ∈ Σ, Bα ⊆ A such that µ(A \ Bα) ≤ δ, c(e⊥ c(e⊥ ε (xα))χBα = 0, i.e. xαχBα ≤ εχBα. Therefore ε (xα)) and xαχBαM ≤ ε µ(A \ Bα) ≤ δ. Thus xα ∈ W (A, ε, δ), i.e. complete. xα t(Z(M )) −→ 0. Therefore xα tc(M ) −→ 0. The proof is Let t(M) denote the topology on E(M) induced by the topology t(M) from LS(M). Proposition 3.5. The topology tc(M) is stronger than the topology t(M) of conver- gence locally in measure. Proof. It is sufficient to show that O(A, ε, δ) ⊂ V (A, ε, δ). (3.3) Let x ∈ O(A, ε, δ), i.e. x ∈ W (A, ε, δ). Then there exists B ∈ Σ such that B ⊆ A, µ(A \ B) ≤ δ, and xχB ∈ L∞(Ω, Σ, µ), kxkχBM ≤ ε. Put z = p = χB. Then xp = xχB = xχB ∈ L∞(Ω, Σ, µ), i.e. xp ∈ M BχB = 0, one has and moreover xpM ≤ ε. Since µ(A \ B) ≤ δ and z⊥χB = χ⊥ z⊥ ∈ W (A, ε, δ). Therefore xpM ≤ ε, z⊥ ∈ W (A, ε, δ), zp⊥ = χBχ⊥ B = 0 7 and hence x ∈ V (A, ε, δ), i.e. O(A, ε, δ) ⊂ V (A, ε, δ). The proof is complete. Proposition 3.6. If M is a type I or III von Neumann algebra and 0 < ε < 1, then O(A, ε, δ) = V (A, ε, δ). Proof. From above (3.3) we have that O(A, ε, δ) ⊂ V (A, ε, δ). Therefore it is suffi- cient to show that V (A, ε, δ) ⊂ O(A, ε, δ). Let x ∈ V (A, ε, δ). Then there exist p ∈ P (M) and z ∈ P (Z(M)) such that xp ∈ M, xpM ≤ ε, z⊥ ∈ W (A, ε, δ), d(zp⊥) ≤ εz. If M is of type I then Remark 2.1 implies that d(zp⊥) ≥ c(zp⊥). Now from d(zp⊥) ≤ εz it follows that c(zp⊥) ≤ εz. From 0 < ε < 1 we obtain that zp⊥ = 0. If M is of type III then the finiteness of the projection zp⊥ implies that zp⊥ = 0. Thus z = zp. Put z = χE for an appropriate E ∈ Σ. Since z⊥ ∈ W (A, ε, δ) one has that χΩ\E ∈ W (A, ε, δ). Thus there exists B ∈ Σ such that B ⊆ A, µ(A \ B) ≤ δ, χΩ\EχB ≤ ε < 1. Hence χB ≤ χE. So we obtain xχB ≤ xχE = xz = xz = xzp = xp ≤ ε. This means that x ∈ O(A, ε, δ). The proof is complete. Proposition 3.6 implies that following Theorem 3.7. If M is a type I or III von Neumann algebra then the topologies t(M) and tc(M) coincide. Proposition 3.8. If M is of type II then t(M) < tc(M). Proof. Since M is a type II then there exists a decreasing sequence of projections {pn} in M such that c(pn) = 1 and d(pn) = 1 2n for all n ∈ N. Then {pn} converges to zero with respect to the topology locally in measure. Indeed take any neighborhood of zero V (A, ε, δ) in the topology t(M). Put z = 1, p = p⊥ k , where the number k is such that 1 2k < ε. For n ≥ k we have that pnp = pnp⊥ k = (pnpk)p⊥ k = 0, and z⊥ ∈ W (A, ε, δ) d(zp⊥) = d(pk) = 1 2k 1 ≤ εz. This means that pn ∈ V (A, ε, δ) for all n ≥ k, i.e. {pn} converges to zero with respect to the topology locally in measure. On the other hand the equality c(pn) = 1 implies that pn = 1. Thus a sequence {pn} does not converges to zero in the topology tc(M). Hence t(M) < tc(M). The proof is complete. 8 Theorem 3.7 and Proposition 3.8 imply the following result which describes the class of von Neumann algebras M for which the topologies t(M) and tc(M) coincide. Theorem 3.9. The following conditions on a given von Neumann algebra M are equivalent: i) t(M) = tc(M); ii) M does not have direct summands of type II. By E(M)h we denote the set of all selfadjoint elements in E(M). A net {xα}α∈I ⊂ (o) −→ x), if there exist nets E(M)h is called (o)-convergent to x ∈ E(M)h (denoted xα {aα}α∈I and {bα}α∈I in E(M)h, such that aα ≤ xα ≤ bα for each α ∈ I and aα ↑ x, bα ↓ x. The strongest topology on E(M)h for which (o)-convergence of nets implies their convergence in the topology is called the order topology, or the (o)-topology, and is denoted by to(M). Let tch(M) (respectively th(M)) denote the topology on E(M)h induced by the topology tc(M) (respectively t(M)) from E(M). We now describe class of von Neumann algebras M for which the topologies tc(M) and to(M) coincide. Theorem 3.10. (i) tch(M) ≤ to(M) if and only if M is of type If in; (ii) tch(M) = to(M) if and only if M is a σ-finite type If in algebra. Proof. (i) Let tch(M) ≤ to(M). If the algebra M does not has type If in then there exists a nonzero projection z ∈ P (Z(M)) and a sequence of mutually orthogonal projections {pn}∞ pn is a contradiction with z 6= 0. Hence M is a type If in algebra. t(Z(M )) −→ 0. Since pn = c(pn) = z it follows that z = 0, this tch(M ) −→ 0. Hence pn n=1 in M with c(pn) = z, n ∈ N. Then pn (o) −→ 0, and therefore Conversely let M be a type If in algebra. Then by [3, Proposition 1.1] we have that LS(M) = E(M). Thus theorem 3.9 implies that tch(M) = th(M). Since th(M) ≤ to(M) (see [8, Theorem 1 (i)]) then tch(M) ≤ to(M). (ii) If tch(M) = to(M) then M is a type If in algebra (see (i)). Again using the theorem 3.9 we have that tch(M) = th(M). Thus th(M) = to(M). Now by [8, Theorem 1 (ii)] follows that M is a σ-finite algebra. Conversely let M be a σ-finite type If in algebra. Then by theorem 3.9 we have that tch(M) = th(M) and by [8, Theorem 1 (ii)] we obtain that th(M) = to(M). Hence tch(M) = th(M) = to(M). The proof is complete. Theorem 3.10 yields the following corollary. Corollary 3.11. The following assertions are true: 9 (i) If M is a σ-finite von Neumann algebra but is not type If in, then to(M) < th(M); (ii) If M is not a σ-finite von Neumann algebra but is type If in, then th(M) < to(M). Proposition 3.12. The topology tc(M) is locally convex if and only if M is *- isomorphic to the C ∗-product L Mj, where Mj are factors. j∈J Proof. Let tc(M) be a locally convex topology on E(M). Since tc(M) induces the topology t(Z(M)) on Z(E(M)) = S(Z(M)), we have that (S(Z(M)), t(Z(M))) is a locally convex space. It follows from [9, 12, Ch. V, §3] that Z(M) is an atomic von Neumann algebra. Hence, the algebra M is *-isomorphic to the C ∗-product L Mj, where Mj are factors for all j ∈ J. j∈J Conversely, let M = L j∈J Mj, where Mj are factors. Then E(Mj) = Mj, tc(M) = tk·kMj , E(M) = Y Mj j∈J and, hence the topology tc(M) is a Tychonoff product of the normed topologies tk·kMj , that is, tc(M) is a locally convex topology. The proof is complete. Similarly, we obtain the following Proposition 3.13. The topology tc(M) can be normed if and only if M = n L j=1 Mj, where Mj are factors, j = 1, n, n ∈ N. Acknowledgments Part of this work was done within the framework of the Associateship Scheme of the Abdus Salam International Centre for Theoretical Physics (ICTP), Trieste, Italy. The first author thank ICTP for providing financial support and all facilities (July- August, 2011). This work is supported in part by the DFG AL 214/36-1 project (Germany). References [1] Albeverio S., Ayupov Sh. A., Kudaybergenov K. K., Derivations on the algebra of measurable operators affiliated with a type I von Neumann algebra, Siberian Adv. Math. 18 (2008) 86 -- 94. [2] Albeverio S., Ayupov Sh. A., Kudaybergenov K. K., Structure of derivations on various algebras of measurable operators for type I von Neumann algebras, J. Func. Anal. 256 (2009) 2917 -- 2943. 10 [3] Ayupov Sh. A., Kudaybergenov K. K., Additive derivations on algebras of mea- surable operators, ICTP, Preprint, No IC/2009/059, -- Trieste, 2009. -- 16 p. (accepted in Journal of operator theory). [4] Ayupov Sh. A., Kudaybergenov K. K., Derivations on algebras of measurable operators, Infin. Dimens. Anal. Quantum Probab. Relat. Top. 13 (2010) 305 -- 337. [5] Albeverio S., Ayupov Sh. A., Kudaybergenov K. K., Djumamuratov R. T., Automorphisms of central extensions of type I von Neumann algebras, arXiv: 1104.4698. (2011) 16 p. [6] Muratov M.A., Chilin V.I., Algebras of measurable and locally measurable op- erators, Institute of Mathematics Ukrainian Academy of Sciences, Kiev, 2007. [7] Muratov M.A., Chilin V.I., Central extensions of *-algebras of measurable op- erators, Doklady AN Ukraine, no 2 (2009) 24-28. [8] Muratov M.A., Chilin V.I., (o)-topologies on *-algebras of locally measurable operators, Ukrainan Jour. Math., 61 (2009) 1798-1808. [9] Sarymsakov T.A., Ayupov Sh.A, Khadzhyev D., Chilin V.I., Ordered algebras, FAN, Tashkent, 1983 (in Russian). [10] Segal I., A non-commutative extension of abstract integration, Ann. Math. 57 (1953) 401 -- 457. [11] Yeadon F.J., Convergence of measurable operators. Proc. Camb. Phil. Soc. 74 (1973) 257-268. 11
1210.0336
1
1210
2012-10-01T10:10:08
W*-superrigidity for group von Neumann algebras of left-right wreath products
[ "math.OA", "math.GR" ]
We prove that for many nonamenable groups \Gamma, including all hyperbolic groups and all nontrivial free products, the left-right wreath product group G := (Z/2Z)^(\Gamma) \rtimes (\Gamma \times \Gamma) is W*-superrigid. This means that the group von Neumann algebra LG entirely remembers G. More precisely, if LG is isomorphic with L\Lambda for an arbitrary countable group \Lambda, then \Lambda must be isomorphic with G.
math.OA
math
W∗-superrigidity for group von Neumann algebras of left-right wreath products by Mihaita Berbec1 and Stefaan Vaes2 Abstract We prove that for many nonamenable groups Γ, including all hyperbolic groups and all nontrivial free products, the left-right wreath product group G := (Z/2Z)(Γ) ⋊ (Γ × Γ) is W∗-superrigid. This means that the group von Neumann algebra LG entirely remembers G. More precisely, if LG is isomorphic with LΛ for an arbitrary countable group Λ, then Λ must be isomorphic with G. 1 Introduction and statements of the main results Over the last years, Popa's deformation/rigidity theory lead to a lot of progress in the classifi- cation of group measure space II1 factors L∞(X) ⋊ G associated with free, ergodic, probability measure preserving actions of countable groups (cf. the surveys in [Po06a, Va10a, Io12a]). In comparison, our understanding of group von Neumann algebras LG is much more limited. Connes' theorem of [Co76] implies that all II1 factors LG coming from amenable groups G with infinite conjugacy classes (icc) are isomorphic. Although nonamenable groups with nonisomor- phic group II1 factors were already discovered in [MvN43, Sc63, McD69], the general question on how LG depends on G remains largely unanswered, especially when G is a "classical group" like SL(n, Z) or a free group Fn. The first rigidity phenomena for group von Neumann algebras emerged in [Co80a], and in [Co80b], Connes asked whether icc property (T) groups G and Λ with isomorphic group von Neumann algebras, LG ∼= LΛ, must necessarily be isomorphic groups. Although this rigidity conjecture remains wide open, deformation/rigidity theory has provided large classes C of icc groups such that two groups G and Λ in the class C must be isomorphic whenever they have isomorphic group II1 factors, see e.g. [Po01, Po04, IPP05, PV06]. This is for instance the case for the class C of all wreath product groups Z/2Z ≀ Γ with Γ an icc property (T) group, see [Po04]. Note however that both G and Λ are assumed to belong to the class C, so that it is not excluded that LG ∼= LH for a group H that is nonisomorphic with G and that lies outside the class C. Even more so, in the case where G = Z/2Z ≀ Γ and Γ is torsion-free, a nonisomorphic H 6∼= G with LH ∼= LG always exists by [IPV10, Theorem 1.2]. Only in [IPV10], the first W∗-superrigidity theorem for group von Neumann algebras was established: for a large class of generalized wreath product groups G = (Z/2Z)(I) ⋊ Γ, it was shown that if LG ∼= LΛ for an arbitrary group Λ, then Λ must be isomorphic with G. Such a group G is called W∗-superrigid (see Definition A for the precise terminology). So G is W∗-superrigid if the group von Neumann algebra LG "remembers" G. The class of groups covered by [IPV10] contains all (Z/2Z)(I) ⋊ (Γ ≀ Z), where Γ is an arbitrary nonamenable group and I = (Γ ≀ Z)/Z. In this paper, we extend the results of [IPV10] and prove W∗-superrigidity for the more natural left-right wreath products G = (Z/2Z)(Γ) ⋊ (Γ × Γ), 1KU Leuven, Department of Mathematics, [email protected] Supported by Research Programme G.0639.11 of the Research Foundation -- Flanders (FWO) 2KU Leuven, Department of Mathematics, [email protected] Partially supported by ERC Starting Grant VNALG-200749, Research Programme G.0639.11 of the Research Foundation -- Flanders (FWO) and KU Leuven BOF research grant OT/08/032. 1 where the direct product Γ × Γ acts on Γ by left-right multiplication, and where Γ is either the free group Fn with n ≥ 2, or any icc hyperbolic group, or any nontrivial free product Γ1 ∗ Γ2. The precise statement is given in Theorem B below. We expect that for most nonamenable icc groups Γ, the left-right wreath product group (Z/2Z)(Γ) ⋊ (Γ × Γ) is W∗-superrigid. As we explain in Remark 6.2, this is however not true for arbitrary nonamenable icc groups Γ. To prove our W∗-superrigidity theorem, we follow the approach of [IPV10], by considering the comultiplication ∆ : LΛ → LΛ ⊗ LΛ that is induced by another group von Neumann algebra decomposition LG = LΛ and carefully analyzing how ∆ relates to the initial von Neumann algebra structure of LG. The following are the two major steps in the proof. We first use Popa's malleable deformation for Bernoulli actions (see [Po03]) and his spectral gap rigidity (see [Po06b]) to prove that the subalgebra L(Γ × Γ) ⊂ LG is invariant under ∆, up to unitary conjugacy. We next use the recent results on normalizers of amenable subalgebras in crossed products by hyperbolic groups (see [PV12]), and in crossed products by arbitrary free product groups (see [Io12b]), to prove that also the subalgebra L(cid:0)(Z/2Z)(Γ)(cid:1) ⊂ LG is invariant under ∆, up to unitary conjugacy. Both steps together bring us to a point where the general results of [IPV10] can be applied. Contrary to the approach of [IPV10], our proof does not use the clustering techniques of [Po04], but uses the recent results of [PV12, Io12b] instead. As a consequence, we can also prove W∗- superrigidity for a number of subgroups of generalized wreath product groups. In particular, we let H be any nontrivial torsion-free abelian group and let Γ, as above, be either the free group Fn with n ≥ 2, or any icc hyperbolic group, or any nonamenable free product Γ1 ∗ Γ2. We define H0 as the subgroup of H (Γ) consisting of those elements x with Pg xg = 0. Then we prove that H0 ⋊ (Γ × Γ) is always W∗-superrigid (see Theorem B). Definition A. A countable group G is called W∗-superrigid if the following holds: if Λ is any countable group and if π : LΛ → (LG)r is a ∗-isomorphism for some r > 0, then r = 1 and there exist an isomorphism of groups δ : Λ → G, a character ω : Λ → T and a unitary w ∈ LG such that π(vs) = ω(s) w uδ(s) w∗ for all s ∈ Λ . Here (vs)s∈Λ and (ug)g∈G denote the canonical generating unitaries of LΛ, resp. LG. The following is our main result. The proof is given at the end of Section 8, as a consequence of the more general Theorem 8.1. Theorem B. Assume that Γ is one of the following groups: • an icc hyperbolic group, • a finitely generated, icc, nonamenable, discrete subgroup of a connected noncompact rank one simple Lie group with finite center, • a free product Γ1 ∗ Γ2 with Γ1 ≥ 2 and Γ2 ≥ 3. All of the following generalized wreath product groups G are W∗-superrigid in the sense of Definition A : 1. the group (Z/nZ)(Γ) ⋊ (Γ × Γ) where n ∈ {2, 3}, 2 2. the kernel of the homomorphism H (Γ) ⋊ (Γ × Γ) → H : xg 7→ Pk∈Γ xk, where H is an arbitrary nontrivial torsion-free abelian group. Remark C. Let Γ be a group as in Theorem B. Assume moreover that Γ has no nontrivial characters. Let H be an an arbitrary nontrivial torsion-free abelian group and denote by G0 the kernel of the homomorphism H (Γ) ⋊ (Γ × Γ) → H given in Theorem B. At the end of section 8, we prove that G0 has no characters either. So the conclusion of Theorem B becomes stronger: whenever Λ is a countable group and π : LΛ → (LG0)r is a ∗-isomorphism, we have r = 1 and there exist an isomorphism of groups δ : Λ → G0 and a unitary w ∈ L(G0) such that π(vs) = w uδ(s) w∗ for all s ∈ Λ. 2 Preliminaries 2.1 Popa's intertwining-by-bimodules We recall Popa's intertwining-by-bimodules theorem. In the formulation of the theorem, we also introduce the notations P ≺ Q and P ≺f Q that are used throughout this article. Theorem 2.1 ([Po03, Theorem 2.1 and Corollary 2.3]). Let (M, τ ) be a tracial von Neumann algebra. Assume that p, q ∈ M are projections and that P ⊂ pM p and Q ⊂ qM q are von Neumann subalgebras with P being generated by a group of unitaries G ⊂ U (P ). Then the following three statements are equivalent. • There exist a nonzero partial isometry v ∈ M1,n(C) ⊗ pM q, a projection q0 ∈ Mn(C) ⊗ Q and a normal ∗-homomorphism θ : P → q0(Mn(C) ⊗ Q)q0 such that xv = vθ(x) for all x ∈ P . • There is no sequence of unitaries (wn) in G satisfying kEQ(x∗wny)k2 → 0 for all x, y ∈ pM q . • There exists a nonzero P -Q-subbimodule of pL2(M )q that has finite right Q-dimension. We write P ≺ Q if these equivalent conditions hold. We write P ≺f Q if P p0 ≺ Q for all nonzero projections p0 ∈ P ′ ∩ pM p. Sometimes we write P ≺M Q to stress the ambient von Neumann algebra M . Note that when the von Neumann algebra M has a nonseparable predual, then sequences have to be replaced by nets in the formulation of Theorem 2.1. Lemma 2.2 ([Va10b, Section 2]). Let Γ be a countable group and Γ y (B, τ ) a trace preserving action. Put M = B ⋊ Γ. Let p ∈ M be a projection and P ⊂ pM p a von Neumann subalgebra. (a) Assume that Λ < Γ is a subgroup. The set of projections p0 ∈ P ′ ∩ pM p satisfying P p0 ≺f B ⋊ Λ attains its maximum in a projection p1 that belongs to the center of the normalizer of P inside pM p. Moreover P (p − p1) 6≺ B ⋊ Λ. (b) Assume that Λ1, Λ2 < Γ are subgroups with Λ2 ⊳ Γ being normal. If P ≺f B ⋊ Λj for all j ∈ {1, 2}, then P ≺f B ⋊ (Λ1 ∩ Λ2). 3 Proof. The first statement follows from [Va10b, Proposition 2.6 and Lemma 2.5], while the second statement follows from [Va10b, Lemmas 2.7 and 2.5]. We also need the following lemma. Lemma 2.3. Let Γ be a countable group and Γ y (B, τ ) a trace preserving action. Put M = B ⋊ Γ and let p ∈ M be a projection. Assume that Q ⊂ pM p is a von Neumann subalgebra that is normalized by a group of unitaries G ⊂ U (pM p). Let Λ < Γ be a subgroup. If Q ≺f B and G′′ ≺ B ⋊ Λ, then (Q ∪ G)′′ ≺ B ⋊ Λ. Proof. For every subset F ⊂ Γ, we denote by PF the orthogonal projection of L2(M ) onto the closed linear span of {bug b ∈ B, g ∈ F}. We say that a subset F ⊂ Γ is small relative to Λ if F is contained in a finite union of subsets of the form gΛh with g, h ∈ Γ. Assume that (Q ∪ G)′′ 6≺ B ⋊ Λ. Since U (Q)G is a group of unitaries generating (Q ∪ G)′′, we get from [Va10b, Lemma 2.4] sequences of unitaries an ∈ U (Q) and wn ∈ G such that kPF (anwn)k2 → 0 for every subset F ⊂ Γ that is small relative to Λ. Since G′′ ≺ B ⋊ Λ, Theorem 2.1 provides a nonzero partial isometry v ∈ M1,n(C) ⊗ pM , a projection q ∈ Mn(C) ⊗ (B ⋊ Λ) and a normal ∗-homomorphism θ : G′′ → q(Mn(C) ⊗ (B ⋊ Λ))q such that xv = vθ(x) for all x ∈ G′′. Denote p1 := vv∗ and fix 0 < ε < kp1k2/3. By the Kaplansky density theorem, we can take a finite subset F1 ⊂ Γ and an element v1 in the linear span of {bug b ∈ M1,n(C) ⊗ B, g ∈ F1} such that kv1k ≤ 1 and kv − v1k2 < ε. Denote F2 := F1ΛF −1 1. By construction, every xn lies in the image of PF2 and we have that kxnk ≤ 1, kwnp1 − xnk2 < 2ε for all n. Since Q ≺f B, we obtain from [Va10b, Lemma 2.5] a finite subset F3 ⊂ Γ such that kan − PF3 (an)k2 < ε for all n. In combination with the previous paragraph, we get that kanwnp1 − PF3 (an)xnk2 < 3ε for all n. Denote F4 := F3F2 and observe that F4 is still small relative to Λ. By construction, PF3(an)xn lies in the image of PF4 and we have thus shown that kanwnp1 − PF4(anwnp1)k2 < 3ε for all n. Since kPF (anwn)k2 → 0 for every subset F ⊂ Γ that is small relative to Λ, it follows from [Va10b, Lemma 2.3] that kPF4 (anwnp1)k2 → 0. Hence lim supn kanwnp1k2 ≤ 3ε. Since an and wn are unitaries, we arrive at the contradiction that kp1k2 ≤ 3ε < kp1k2. 1 . Observe that F2 is small relative to Λ. Write xn := v1θ(wn)v∗ 2.2 Bimodules and weak containment Let (M, τ ) be a tracial von Neumann algebra and Q ⊂ M a von Neumann subalgebra. The basic construction hM, eQi is defined as the von Neumann algebra acting on L2(M ) generated by M and the orthogonal projection eQ of L2(M ) onto L2(Q). Recall that hM, eQi equals the commutant of the right Q-action on L2(M ), i.e. hM, eQi = B(L2(M )) ∩ (Qop)′. Let M, N be tracial von Neumann algebras. An M -N -bimodule MHN is a Hilbert space H equipped with two commuting normal unital ∗-homomorphisms M → B(H) and N op → B(H). Any M -N -bimodule MHN gives rise to a ∗-homomorphism πH : M ⊗alg N op → B(H) given by πH(x ⊗ yop)ξ = xξy, for all x ∈ M , y ∈ N and ξ ∈ H. If (ρ, K) and (π, H) are unitary representations of a countable group Γ, we say that ρ is weakly contained in π if kρ(a)k ≤ kπ(a)k for all a ∈ CΓ. Similarly, if MKN and MHN are M -N -bimodules, we say that MKN is weakly contained in MHN if kπK(x)k ≤ kπH(x)k for all x ∈ M ⊗alg N op. 4 2.3 Relative amenability A tracial von Neumann algebra (M, τ ) is called amenable if there exists an M -central state on B(L2(M )) whose restriction to M equals τ . Also M is amenable if and only if the trivial M -M - bimodule ML2(M )M is weakly contained in the coarse M -M -bimodule M(L2(M ) ⊗ L2(M ))M. Definition 2.4 ([OP07, Section 2.2]). Let (M, τ ) be a tracial von Neumann algebra and let P ⊂ pM p and Q ⊂ M be von Neumann subalgebras. We say that P is amenable relative to Q, if there exists a P -central positive functional on the von Neumann algebra phM, eQip whose restriction to pM p equals τ . Similarly, if Γ is a countable group with subgroups Λ1, Λ2 < Γ, we say that Λ1 is amenable relative to Λ2 if the action of Λ1 on Γ/Λ2 by left translations admits an invariant mean. The following lemma is essentially contained in [MP03, Proposition 6]. For completeness, we provide a full proof. Lemma 2.5. Let Γ be a countable group and Γ y (B, τ ) a trace preserving action. Put M = B ⋊ Γ and let Λ1, Λ2 < Γ be subgroups. Then the following statements are equivalent. (a) B ⋊ Λ1 is amenable relative to B ⋊ Λ2 inside M . (b) LΛ1 is amenable relative to B ⋊ Λ2 inside M . (c) Λ1 is amenable relative to Λ2 inside Γ. Proof. (a) ⇒ (b) is trivial. (b) ⇒ (c). For every g ∈ Γ, we denote by δgΛ2 ∈ ℓ∞(Γ/Λ2) the function that is equal to 1 in gΛ2 and that is equal to 0 elsewhere. There is a unique unital normal ∗-homomorphism π : ℓ∞(Γ/Λ2) → hM, eB⋊Λ2 i satisfying π(δgΛ2) = ug eB⋊Λ2 u∗ g for all g ∈ Γ . By construction, π conjugates the left translation action of Γ on ℓ∞(Γ/Λ2) with the action (Ad ug)g∈Γ. Since LΛ1 is amenable relative to B ⋊ Λ2 inside M , we can take an LΛ1-central state Ω on hM, eB⋊Λ2 i. Then Ω ◦ π is a Λ1-invariant state on ℓ∞(Γ/Λ2). Hence (c) holds. (c) ⇒ (a). We denote by η : Γ → U (ℓ2(Γ/Λ2)) the unitary representation of Γ given by left translation operators. We then turn the Hilbert space L2(M )⊗ℓ2(Γ/Λ2) into an M -M -bimodule with the bimodule action given by (bug) · (x ⊗ ξ) · y := bugxy ⊗ ηgξ for all b ∈ B, g ∈ Γ, x, y ∈ M, ξ ∈ ℓ2(Γ/Λ2) . Since (c) holds, take a sequence of unit vectors ξn ∈ ℓ2(Γ/Λ2) satisfying limn kηgξn − ξnk2 = 0 for all g ∈ Λ1. Then the sequence of vectors 1 ⊗ ξn ∈ L2(M ) ⊗ ℓ2(Γ/Λ2) satisfies hx · (1 ⊗ ξn), 1 ⊗ ξni = τ (x) for all x ∈ M and lim n kbug · (1 ⊗ ξn) − (1 ⊗ ξn) · bugk2 = 0 for all b ∈ U (B), g ∈ Λ1 . Observe that there is a unique unitary operator θ : L2(hM, eB⋊Λ2 i) → L2(M ) ⊗ ℓ2(Γ/Λ2) satisfying θ(bug eB⋊Λ2 x) = bugx ⊗ δgΛ2 for all b ∈ B, g ∈ Γ, x ∈ M . This unitary θ is M -M -bimodular. Define Sn ∈ L2(hM, eB⋊Λ2 i) given by Sn := θ−1(1 ⊗ ξn). Choose a state Ω on hM, eB⋊Λ2 i as a weak∗-limit point of the sequence of states T 7→ hT Sn, Sni. By construction, Ω(x) = τ (x) for all x ∈ M and Ω is G- central, where G = {bug b ∈ U (B), g ∈ Λ1}. Using the Cauchy-Schwarz inequality, it follows that Ω is (B ⋊ Λ1)-central. So (a) holds. 5 We need two elementary lemmas. Lemma 2.6. Let (M, τ ) be a tracial von Neumann algebra and let P ⊂ pM p and Q ⊂ M be von Neumann subalgebras. The set of projections p0 ∈ P ′ ∩ pM p with the property that P p0 is amenable relative to Q, attains its maximum in a projection p1 that belongs to the center of the normalizer of P inside pM p. Proof. Denote by P the set of projections p0 ∈ P ′∩pM p with the property that P p0 is amenable relative to Q. If p0 ∈ P and u ∈ NpM p(P ), it is easy to check that up0u∗ ∈ P. It therefore suffices to prove the following two statements. 1. If p0, p1 ∈ P, then q := p0 ∨ p1 belongs to P. For all j ∈ {0, 1}, choose P pj-central positive functionals Ωj on pjhM, eQipj with the property that Ωj(x) = τ (x) for all x ∈ pjM pj. Define the positive functional Ω on qhM, eQiq by the formula Ω(T ) := Ω0(p0T p0) + Ω1(p1T p1). It is easy to check that Ω is P q-central and that the restriction of Ω to qM q is normal and faithful. By [OP07, Theorem 2.1], we get that P q is amenable relative to Q. 2. If pn is an increasing sequence in P that converges strongly to q, then also q ∈ P. Take P pn-central positive functionals Ωn on pnhM, eQipn with the property that Ωn(x) = τ (x) for all n ∈ N and all x ∈ pnM pn. Choose a positive functional Ω on qhM, eQiq as a weak∗ limit point of the sequence of functionals T 7→ Ωn(pnT pn). By construction, Ω is P q-central and Ω(x) = τ (x) for all x ∈ qM q. So q ∈ P. We also need the following special case of [PV11, Proposition 2.7]. Lemma 2.7 ([PV11, Proposition 2.7]). Let Γ be a countable group and Γ y (B, τ ) a trace preserving action. Put M = B ⋊ Γ. Let p ∈ M be a projection and P ⊂ pM p a von Neumann subalgebra. Assume that Λ1, Λ2 < Γ are subgroups with Λ2 ⊳ Γ being normal. If P is amenable relative to B ⋊ Λj for all j ∈ {1, 2}, then P is amenable relative to B ⋊ (Λ1 ∩ Λ2). We finally need the concept of a left amenable bimodule, see [Si10, Theorem 2.2] and [PV11, Definition 2.3]. Definition 2.8. Let (M, τ ) and (N, τ ) be tracial von Neumann algebras. Let P ⊂ M be a von Neumann subalgebra. An M -N -bimodule MKN is said to be left P -amenable if B(K) ∩ (N op)′ admits a P -central state whose restriction to M equals τ . If (M, τ ) is a tracial von Neumann algebra and if P ⊂ pM p, Q ⊂ M are von Neumann subalgebras, then by definition, P is amenable relative to Q if and only if the pM p-Q-bimodule pL2(M ) is left P -amenable. The following easy lemmas are essentially contained in [OP07, Section 2.2]. For completeness, we provide full proofs. Lemma 2.9. Let (M, τ ) and (N, τ ) be tracial von Neumann algebras. Let P ⊂ M be a von Neumann subalgebra and MKN an M -N -bimodule. The following two statements are equivalent. (a) There exists a nonzero P -central positive functional on B(K) ∩ (N op)′ whose restriction to M is normal. (b) There exists a nonzero projection p ∈ P ′ ∩ M such that the pM p-N -bimodule pM p(pK)N is left P p-amenable. 6 Proof. (a) ⇒ (b). Let Ω be a nonzero P -central positive functional on N := B(K) ∩ (N op)′ whose restriction to M , denoted by ω is normal. Take T ∈ L1(M )+ such that ω(x) = τ (xT ) for all x ∈ M . Note that T 6= 0. Since ω is P -central, we have that T ∈ L1(P ′ ∩ M ). Take ε > 0 small enough such that the spectral projection p := χ(ε,+∞)(T ) is nonzero. Note that p ∈ P ′ ∩ M and that we can take S ∈ p(P ′ ∩ M )+p such that T S = ST = p. The formula y 7→ Ω(S1/2yS1/2) defines P p-central positive functional on B(pK) ∩ (N op)′ whose restriction to pM p equals τ . So pM p(pK)N is left P p-amenable. (b) ⇒ (a). Assume that p ∈ P ′ ∩ M is a nonzero projection and that Ω is a P p-central positive functional on B(pK) ∩ (N op)′ whose restriction to pM p equals τ . Then the formula y 7→ Ω(pyp) defines a nonzero P -central positive functional on B(K) ∩ (N op)′ whose restriction to M is normal. Lemma 2.10. Let (M, τ ) be a tracial von Neumann algebra with von Neumann subalgebra P ⊂ M . Let K be an M -M -bimodule. Assume that ξn ∈ K is a sequence of vectors and ε > 0 such that • kxξnk ≤ kxk2 for all x ∈ M and n ∈ N, • kξnk ≥ ε for all n ∈ N, • for all x ∈ P , we have that limn kxξn − ξnxk = 0. Then there exists a nonzero projection p ∈ P ′ ∩ M such that the pM p-M -bimodule pK is left P p-amenable. Proof. Choose a positive functional Ω on B(K)∩(M op)′ as a weak∗ limit point of the sequence of positive functionals y 7→ hyξn, ξni. The conditions on ξn imply that Ω(x) ≤ τ (x) for all x ∈ M +, that Ω(1) ≥ ε2 and that Ω is P -central. In particular, Ω is nonzero and the restriction of Ω to M is normal. The conclusion now follows from Lemma 2.9. Lemma 2.11. Let (M, τ ) and (N, τ ) be tracial von Neumann algebras. Let P ⊂ M be a von Neumann subalgebra. Assume that for all j ∈ {1, . . . , ℓ}, we are given an M -N -bimodule Kj. If j=1 Kj is a left P -amenable M -N -bimodule, then there exists a j ∈ {1, . . . , ℓ} and a nonzero projection p ∈ P ′ ∩ M such that pKj is a left P p-amenable pM p-N -bimodule. Lℓ Proof. Put K :=Lℓ j=1 Kj and denote by pj the orthogonal projection of K onto Kj. Let Ω be a P -central state on B(K) ∩ (N op)′ whose restriction to M equals τ . Take j ∈ {1, . . . , ℓ} such that Ω(pj) 6= 0. Then the formula y 7→ Ω(pjypj) defines a nonzero P -central positive functional on B(Kj) ∩ (N op)′ whose restriction to M is smaller or equal than τ and hence normal. So the conclusion follows from Lemma 2.9. 2.4 Weak amenability and class S We very briefly introduce weak amenability and bi-exactness (class S) for countable groups. We only use these concepts in the following way: the first two families of groups in Theorem B are weakly amenable and in class S, so that we can apply the results of [PV12] to them. Recall from [CH88] that a countable group Γ is called weakly amenable if Γ admits a sequence of finitely supported functions fn : Γ → C tending to 1 pointwise and satisfying supn kfnkcb < ∞. Here kf kcb is the Herz-Schur norm, i.e. the cb-norm of the linear map LΓ → LΓ : ug 7→ f (g)ug. 7 Following [Oz03] (see also [BO08, Chapter 15]), a group Γ is said to be in class S (or bi-exact) if Γ is an exact group and if there exists a map µ : Γ → Prob Γ from Γ to the probability measures on Γ satisfying lim k→∞ kµ(gkh) − g · µ(k)k1 = 0 for all g, h ∈ Γ . It immediately follows that if Γ belongs to class S and if Λ < Γ is an infinite subgroup, then the centralizer of Λ inside Γ is amenable. Ozawa's theorem in [Oz03] says that much more is true: if Q ⊂ LΓ is any diffuse von Neumann subalgebra, then the relative commutant Q′ ∩ LΓ is amenable. 2.5 Property Gamma, inner amenability and McDuff II1 factors Recall that a II1 factor M is said to have property Gamma, if M admits a sequence of unitaries un ∈ M such that τ (un) = 0 for all n and limn kunx − xunk2 = 0 for all x ∈ M . Let G be an icc group and denote M := LG. By [Ef73], if M has property Gamma, then G must be inner amenable, meaning that the unitary representation (Ad g)g∈G on ℓ2(G − {e}) has almost invariant vectors: there exists a sequence of unit vectors ξn ∈ ℓ2(G − {e}) such that limn k(Ad g)(ξn) − ξnk2 = 0 for every g ∈ G. The converse can however fail, as was shown in [Va09]. Denote by R the unique hyperfinite II1 factor. A II1 factor M is said to be McDuff if M is isomorphic with M ⊗ R. Every McDuff II1 factor has property Gamma. By [McD69], a II1 factor M is McDuff if and only if M admits two central sequences of unitaries un, vn ∈ M such that τ (un) = τ (vn) = τ (unvnu∗ For every II1 factor M , we denote by Aut(M ) the group of automorphisms of M , which naturally is a Polish group. We denote by Inn(M ) := {Ad u u ∈ U (M )} the normal subgroup of inner automorphisms and by Out(M ) := Aut(M )/ Inn(M ) the quotient group. Then M is non-Gamma if and only if Inn(M ) is closed in Aut(M ). In that case, Out(M ) naturally becomes a Polish group as well. n) = 0 for all n. nv∗ 2.6 Weakly mixing actions and weakly mixing representations Recall that a unitary representation π : Γ → U (H) is called weakly mixing if π has no nonzero finite-dimensional globally (π(g))g∈Γ-invariant subspaces. Similarly, a probability measure preserving (pmp) action Γ y (X, µ) is called weakly mixing if the associated unitary representation Γ y L2(X) ⊖ C1 is weakly mixing. If Γ y (X, µ) is a pmp action, then the following conditions are equivalent: • Γ y (X, µ) is weakly mixing, • the diagonal action Γ y X × X : g · (x, y) = (g · x, g · y) is ergodic, • whenever Γ y (Y, η) is a pmp action and F : X × Y → C is a measurable function that is invariant under the diagonal action Γ y X × Y : g · (x, y) = (g · x, g · y), we have that F is a.e. equal to a function that only depends on the Y -variable. The following lemma is classical (see e.g. [PV06, Proposition 2.3 and Lemma 2.4] for a simple proof). 8 Lemma 2.12. Assume that the countable group Γ acts on the countable set I. Let (X0, µ0) be an arbitrary nontrivial standard probability space. Then the following conditions are equivalent. • For every i ∈ I, the orbit Γ · i is infinite. • For every finite subset F ⊂ I, there exists a g ∈ Γ such that g · F ∩ F = ∅. • The unitary representation Γ y ℓ2(I) is weakly mixing. • The generalized Bernoulli action Γ y (X0, µ0)I is weakly mixing. 3 Spectral gap rigidity for generalized Bernoulli actions Let G be a countable discrete group acting on a countable set I. Assume that (A0, τ ) is an arbitrary tracial von Neumann algebra. We denote by AI 0 the tensor product, with respect to τ , of copies of A0 indexed by I. We let G act on AI 0 by the generalized Bernoulli action: denoting by πi : A0 → AI 0 the embedding of A0 as the i-th tensor factor, this generalized Bernoulli action (σg)g∈G is given by σg ◦ πi = πg·i for all g ∈ G and i ∈ I. We consider the crossed product von Neumann algebra M := AI ⋊G. Whenever F ⊂ I, we write Stab F := {g ∈ G g·i = i, ∀i ∈ F}. 0 In [Po03, Po04], Popa discovered his fundamental malleable deformation for Bernoulli crossed ⋊ G and used it to establish the first W∗-rigidity theorems in the case products M = AG 0 where G has property (T). In [Po06b], Popa introduced his spectral gap methods to prove W∗-rigidity theorems for AG ⋊ G in the case where G is a direct product of nonamenable 0 groups. These methods and results have been generalized in many subsequent works (see e.g. [PV06, Va07, Io10, IPV10]) and were in particular extended to cover certain generalized Bernoulli actions, associated with general group actions G y I. So far, the spectral gap methods could only be employed under the assumption that Stab i is amenable for all i ∈ I (see e.g. [IPV10, Corollary 4.3]). In this section, we show that it is actually sufficient to have a constant κ > 0 such that Stab F is amenable for all subsets F ⊂ I with F ≥ κ. We use the following variant, due to [Io06], of Popa's malleable deformation for Bernoulli crossed products. Consider the free product A0 ∗ LZ with respect to the natural traces. Denote by fM := (A0 ∗ LZ)I ⋊ G the corresponding generalized Bernoulli crossed product. Define the self-adjoint h ∈ LZ with spectrum [−π, π] such that exp(ih) equals the canonical generating unitary u1 ∈ LZ. Put ut := exp(ith) and note that ut is a one-parameter group of unitaries with τ (ut) < 1 for all t 6= 0. As above we denote by πi : A0 ∗ LZ → (A0 ∗ LZ)I the embedding as the i-th tensor factor. We can then define the malleable deformation (αt)t∈R by t ) for all g ∈ G, t ∈ R, automorphisms of fM given by αt(ug) = ug and αt(πi(x)) = πi(utxu∗ i ∈ I and x ∈ A0 ∗ LZ. Denote ρt := τ (ut)2 and observe that 0 ≤ ρt < 1 for all t 6= 0. For every finite subset F ⊂ I, we denote by πF : AF 0 the natural embedding. Define the unital completely positive maps ψt : M → M given by ψt(x) = EM (αt(x)) for all x ∈ M . Whenever a ∈ AF 0 is the elementary tensor given by a = ⊗ i∈F ai with ai ∈ A0 ⊖ C1, we have 0 → AI ψt(πF (a)ug) = ρ F t πF (a)ug for all t ∈ R, g ∈ G . Therefore we consider the malleable deformation (αt)t∈R, and the corresponding completely positive maps (ψt)t∈R, as the tensor length deformation of the generalized Bernoulli crossed product M = AI 0 ⋊ G. 9 Theorem 3.1. Let G y I be an action of a countable group on a countable set. Assume that κ, ℓ > 0 are integers and that G1, . . . , Gℓ < G are subgroups with the following property: for every finite subset F ⊂ I with F ≥ κ, there exists an i ∈ {1, . . . , ℓ} such that Stab F is amenable relative to Gi. Assume that (A0, τ ) and (N, τ ) are arbitrary tracial von Neumann algebras. Consider as above the generalized Bernoulli crossed product M = AI ⋊ G with its tensor length deformation 0 αt ∈ Aut(fM ). Assume that p ∈ N ⊗ M is a nonzero projection and that P ⊂ p(N ⊗ M )p is a von Neumann subalgebra such that for all nonzero projections q ∈ P ′ ∩ p(N ⊗ M )p and all i = 1, . . . , ℓ, we have that P q is nonamenable relative to N ⊗ (AI 0 ⋊ Gi). Then sup k(id ⊗ αt)(b) − bk2 converges to 0 as t → 0. b∈U (P ′∩p(N ⊗M )p) [Po06b, Lemma 5.1] and [IPV10, Corollary 4.3]. The essential difference is that we replace the Put M := N ⊗ M and fM := N ⊗ fM . The proof of Theorem 3.1 follows closely the proofs of bimodule ML2(fM ⊖ M)M by the following M-M-submodule Kκ := span  x ∈ N , g ∈ G, F ⊂ I with κ ≤ F < ∞, a = ⊗ i∈F x ⊗ πF (a)ug (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ai ∈ A0 ∗ LZ ⊖ A0 for at least κ elements i ∈ F   ai with ai ∈ A0 ∗ LZ for all i and with . (3.1) Before proving Theorem 3.1, we need the following lemma. Lemma 3.2. Under the assumptions of Theorem 3.1, put Mi := N ⊗ (AI ⋊ Gi). Then there 0 exist Mi-M-bimodules Hi such that the M-M-bimodule Kκ is weakly contained in the M-M- bimodule Lℓ i=1(cid:0)L2(M) ⊗Mi Hi(cid:1). Proof. Let u ∈ LZ be the canonical generating unitary. Let A ⊂ A0 ⊖ C1 be an orthonormal basis of L2(A0) ⊖ C1. Define B ⊂ A0 ∗ LZ given by B := {un1a1un2a2 · · · unk−1ak−1unk k ≥ 1, and for all j, nj ∈ Z − {0}, aj ∈ A} . By construction, we have the following orthogonal decomposition of L2(A0 ∗ LZ) into A0-A0- subbimodules: L2(A0 ∗ LZ) = L2(A0) ⊕Mb∈B A0bA0 . Fix F ⊂ I finite, with F ≥ κ, and fix for all i ∈ F, ci ∈ B. Denote c := 1 ⊗ πF(cid:16) ⊗ i∈F ci(cid:17) ∈ N ⊗ (A0 ∗ LZ)I . Define the M-M-subbimodule of Kκ given by Kc := McM. Define the subgroup Λ < G given by Λ := {g ∈ G g · F = F, cg·i = ci for all i ∈ F} . (3.2) The formula x ⊗ y 7→ xcy defines an M-M-bimodular unitary between L2(M) ⊗Q L2(M) and Kc with Q := N ⊗ (AI−F ⋊ Λ). The different Kc span a dense subspace of Kκ. Also, if F, c and F ′, c′ are chosen as above, there are two possibilities: either there exists a g ∈ G such that 0 10 . 0 , while in the second case, we have Kc ⊥ Kc′ g·i = ci for all i ∈ F, or such a g ∈ G does not exist. In the first case, we have F ′ = g · F and c′ Kc = Kc′ Altogether we can choose a sequence of c's as above, denoted cn, such that Kκ is the orthogonal direct sum of its subbimodules Kcn. To each cn corresponds a finite subset Fn ⊂ I satisfying Fn ≥ κ, and a subgroup Λn < G given by (3.2). Note that by (3.2), we get that Stab Fn is a finite index subgroup of Λn. Writing Qn = N ⊗ (AI−Fn ⋊ Λn), we conclude that Kκ is isomorphic to the direct sum of the sequence of M-M-bimodules L2(M) ⊗Qn L2(M). By the assumptions of the lemma, for every n, there exists an i(n) ∈ {1, . . . , ℓ} such that Stab Fn is amenable relative to Gi(n) inside G. Since Stab Fn < Λn has finite index, also Λn is amenable relative to Gi(n) inside G. It then follows from Lemma 2.5 that N ⊗ (AI ⋊ Λn) is amenable 0 relative to Mi(n). A fortiori, Qn is amenable relative to Mi(n). By [PV11, Proposition 2.4.3], as the direct sum of all L2(M) ⊗Qn L2(M) with i(n) = i, it follows that Kκ is weakly contained i=1(cid:0)L2(M) ⊗Mi Hi(cid:1) as an M-M-bimodule. this means that ML2(M)Qn is weakly contained in M(cid:0)L2(M) ⊗Mi(n) L2(M)(cid:1)Qn. Defining Hi in Lℓ Proof of Theorem 3.1. Denote by PKκ the orthogonal projection of L2(fM) onto the closed subspace Kκ that we defined in (3.1). Denote U := U (P ′ ∩ p(N ⊗ M )p). We start by proving the following claim that is a variant of Popa's fundamental transversality property in [Po06b, Lemma 2.1]. Claim. If sup b∈U when t → 0. kPKκ((id ⊗ αt)(b))k2 → 0 when t → 0, then also sup b∈U k(id ⊗ αt)(b) − bk2 → 0 To prove the claim, we first determine a formula for kPKκ(id ⊗ αt)(y)k2 when y ∈ M. For every n ≥ 0, define the closed subspace Hn ⊂ L2(M) as Hn := span(cid:26) x ⊗ πF (a)ug (cid:12)(cid:12)(cid:12)(cid:12) x ∈ N , g ∈ G, F ⊂ I finite, F = n, a = ⊗ i∈F with ai ∈ A0 ⊖ C1 for all i ∈ F ai (cid:27) . Observe that L2(M) is the orthogonal direct sum of the Hn. Denote by Pn the orthogonal projection of L2(M) onto Hn. Fix a finite subset F ⊂ I with F ≥ κ and fix, for all i ∈ F, elements ai ∈ A0 ⊖ C1. Put a = ⊗ i∈F ai. For all x ∈ N and all g ∈ G, we have x ⊗ αt(πF (a)ug) = x ⊗ πF(cid:16) ⊗ i∈F utaiu∗ t(cid:17) ug (utaiu∗ = XG⊂F x ⊗ πG(cid:16) ⊗ i∈G t − ρtai)(cid:17) πF −G(cid:16) ⊗ i∈F −G ρtai(cid:17) ug . In this last sum, the term corresponding to G ⊂ F belongs to Kκ if G ≥ κ, and is orthogonal to Kκ if G < κ. Therefore, we have for all x ∈ N and all g ∈ G that (1 − PKκ)(x ⊗ αt(πF (a)ug)) = XG ⊂ F , x ⊗ πG(cid:16) ⊗ i∈G G < κ (utaiu∗ t − ρtai)(cid:17) πF −G(cid:16) ⊗ i∈F −G ρtai(cid:17) ug . Put y = x ⊗ πF (a)ug and assume that y′ = x′ ⊗ πF ′(a′)ug′ is of a similar form. Since hutau∗ t − ρta, utbu∗ t − ρtbi = (1 − ρ2 t ) τ (b∗a) for all a, b ∈ A0 ⊖ C1 , 11 we get that h(1 − PKκ)(id ⊗ αt)(y), (1 − PKκ)(id ⊗ αt)(y′)i = hy, y′i κ−1Xj=0(cid:0) F j (cid:1) (1 − ρ2 t )j ρ2(F −j) t , with both sides being zero if F 6= F ′. We conclude that for all y ∈ M, we have k(1 − PKκ)(id ⊗ αt)(y)k2 2 = ∞Xn=0 cκ(t, n) kPn(y)k2 2 where cκ(t, n) = min(κ−1,n)Xj=0 (cid:0) n j(cid:1) (1 − ρ2 t )j ρ2(n−j) t . Note that cκ(t, n) = 1 if n < κ. It follows that kPKκ(αt(y))k2 2 = ∞Xn=0 (1 − cκ(t, n)) kPn(y)k2 2 for all y ∈ M . (3.3) To prove the claim, assume that kPKκ(id ⊗ αt)(b)k2 → 0 when t → 0 . sup b∈U Choose ε > 0. Take t > 0 such that kPKκ (id ⊗ αt)(b)k2 < ε for all b ∈ U . Since cκ(t, n) → 0 when n → ∞ and t is fixed, we can take n0 such that cκ(t, n) < 1/2 for all n ≥ n0. It then follows from (3.3) that for all b ∈ U , we have ε2 > kPKκ (id ⊗ αt)(b)k2 2 ≥ 1 2 ∞Xn=n0 kPn(b)k2 2 . (3.4) We finally take s0 > 0 such that 1 − ρn follows that for all b ∈ U and all s < s0, we have s < ε2 for all s < s0 and all 0 ≤ n < n0. Using (3.4), it k(id ⊗ αs)(b) − bk2 2 = ≤ ∞Xn=0 n0−1Xn=0 2(1 − ρn s ) kPn(b)k2 2 2ε2 kPn(b)k2 2 + 2 ∞Xn=n0 kPn(b)k2 2 ≤ 2ε2 + 4ε2 . So, k(id ⊗ αs)(b) − bk2 ≤ 3ε for all s < s0 and all b ∈ U . This proves the claim. To prove the theorem, assume that sup{k(id ⊗ αt)(b) − bk2 b ∈ U} does not tend to 0 as t → 0. We will produce a nonzero projection q ∈ P ′ ∩ pMp and a j ∈ {1, . . . , ℓ} such that P q is amenable relative to Mj. This will conclude the proof of the theorem. By the claim above, we find an ε > 0, a t0 > 0, and for every 0 < t < t0, a unitary bt ∈ U such that kPKκ (id ⊗ αt)(bt)k2 ≥ ε. Define ξt := PKκ(id ⊗ αt)(bt). We have kξtk2 ≥ ε for all 0 < t < t0. For every fixed x ∈ P , we have that kxξt − ξtxk2 → 0 as t → 0. We finally have kxξtk2 ≤ kxk2 for all x ∈ M. So Lemma 2.10 provides a nonzero projection q ∈ P ′ ∩ pMp such that the qMq-M-bimodule qKκ is left P q-amenable. Using [PV11, Corollary 2.5] and Lemma 12 3.2, we find Mj-M-bimodules Hj such thatLℓ j=1 qL2(M)⊗Mj Hj is left P q-amenable. Making q ∈ P ′ ∩ pMp smaller, Lemma 2.11 yields a j ∈ {1, . . . , ℓ} such that qL2(M) ⊗Mj Hj is a left P q-amenable bimodule. By [PV11, Proposition 2.4.4], the qMq-Mj-bimodule qL2(M) is left P q-amenable. This precisely means that P q is amenable relative to Mj. We also need the following variant of [Po03, Theorem 4.1] and its subsequent generalizations in [Io10, Theorem 2.1] and [IPV10, Theorem 4.2]. Since our proof is almost identical, we are rather brief. Theorem 3.3. Let G y I be an action of a countable group on a countable set. Assume that (A0, τ ) and (N, τ ) are arbitrary tracial von Neumann algebra. Consider as above the generalized Bernoulli crossed product M = AI 0 ⋊ G with its tensor length deformation αt ∈ Aut(fM ). Assume that p ∈ N ⊗ M is a nonzero projection and that Q ⊂ p(N ⊗ M )p is a von Neumann subalgebra generated by a group of unitaries G ⊂ U (Q) with the property that k(id ⊗ αt)(b) − bk2 converges to 0 as t → 0. sup b∈G If G is icc, if N is a factor and if for all i ∈ I, we have that Q 6≺ N ⊗ (AI 0 exists a partial isometry v ∈ N ⊗ M with vv∗ = p and v∗Qv ⊂ N ⊗ LG. ⋊ Stab i), then there Proof. As above, we put M = N ⊗M and fM = N ⊗fM . We first prove the existence of a nonzero partial isometry v ∈ M with the properties that vv∗ ∈ Q′ ∩ pMp and that v∗Qv ⊂ N ⊗ LG. We reason exactly as in the proofs of [Po03, Theorem 4.1], [Io10, Theorem 2.1] and [IPV10, Theorem 4.2]. For completeness, we nevertheless provide some details. By the uniform convergence of id ⊗ αt on G, we find a t > 0 and a nonzero partial isometry is of the form t = 2−n. Since for all i ∈ I, we have that Q 6≺ N ⊗ (AI 0 from [IPV10, Lemma 4.1.1] that w0w∗ w0 ∈ pfM(id ⊗ αt)(p) such that xw0 = w0(id ⊗ αt)(x) for all x ∈ Q. We may assume that t automorphism β ∈ Aut(fM ) given by β(x) = x for all x ∈ M and β(πi(u1)) = u∗ ⋊ Stab i), it follows 0w0 ∈ (id ⊗ αt)(M). Define the period two 1 for all i ∈ I. By construction, β ◦ αt = α−t ◦ β. 0 ∈ M and w∗ We can now define w1 := (id ⊗ αt)((id ⊗ β)(v∗)v) and check that w1 is a nonzero partial isometry in pfM(id ⊗ α2t)(p) satisfying xw1 = w1(id ⊗ α2t)(x) for all x ∈ Q. Continuing inductively, we find a nonzero partial isometry w ∈ pfM(id ⊗ α1)(p) satisfying xw = w(id ⊗ α1)(x) for all x ∈ Q. Literally repeating a part of the proof of [IPV10, Theorem 4.2], we find a finite, possibly empty, subset F ⊂ I such that Q ≺ N ⊗ (AF ⋊ Stab i) for all i ∈ I, ensures that 0 F = ∅. So, Q ≺ N ⊗ LG. ⋊ Stab F). Our assumption that Q 6≺ N ⊗ (AI 0 Take n ∈ N, a nonzero partial isometry v ∈ M1,n(C) ⊗ p(N ⊗ M ), a projection q in Mn(C) ⊗ N ⊗ LG and a ∗-homomorphism θ : Q → q(Mn(C) ⊗ N ⊗ LG)q such that xv = vθ(x) for all x ∈ Q. Since Q 6≺ N ⊗ (AI ⋊ Stab i) for all i ∈ I, by [Va07, Remark 3.8], we may assume 0 that for all i ∈ I, we have θ(Q) 6≺ N ⊗ L(Stab i). By [IPV10, Lemma 4.1.1], we then get that θ(Q)′ ∩ q(Mn(C) ⊗ N ⊗ M )q ⊂ Mn(C) ⊗ N ⊗ LG . In particular, v∗v is a projection in Mn(C) ⊗ N ⊗ LG of trace at most 1. Since N ⊗ LG is a II1 factor, we may then assume that n = 1. So, we have found a nonzero partial isometry v ∈ M with the properties that vv∗ ∈ Q′ ∩ pMp and that v∗Qv ⊂ N ⊗ LG. 13 n are orthogonal projections in Q′∩pMp such that v∗ Let vn be a maximal sequence of nonzero partial isometries vn ∈ M with the property that the vnv∗ n. Since we can apply the previous paragraph to Qp0 ⊂ p0Mp0, the maximality of the sequence (vn) ensures us that p0 = 0. Since N ⊗ LG is a II1 factor and since the v∗ nQvn ⊂ N ⊗LG. Put p0 := p−Pn vnv∗ n = p, we can take partial isometries wn ∈ N ⊗ LG such that wnw∗ nvn form a sequence of projections in N ⊗ LG with nvn for all n nwn are orthogonal. Then v :=Pn vnwn is a partial isometry n = v∗ Pn vnv∗ and such that the projections w∗ in M with vv∗ = p and v∗Qv ⊂ N ⊗ LG. 4 Properties of amplified comultiplications Throughout this section, assume that M0 is a II1 factor and r > 0 such that M r 0 = LΛ for some countable group Λ. We denote by (vs)s∈Λ the canonical generating unitaries of LΛ and define the comultiplication ∆ : LΛ → LΛ ⊗ LΛ given by ∆(vs) = vs ⊗ vs for all s ∈ Λ. Up to unitary conjugacy, we have a uniquely defined amplified comultiplication ∆ : M0 → (M0 ⊗ M0)r that we continue to denote by ∆. At a certain point, we will need the explicit relation between the original comultiplication on LΛ and the amplified comultiplication on M0. This is spelt out in Remark 4.2. Whenever (M, τ ) is a tracial von Neumann algebra and M0 ⊂ M , we define as follows the 0 ⊂ M r. Choose a projection p ∈ Mn(C) ⊗ M0 with (Tr ⊗τ )(p) = r and define inclusion M r 0 ⊂ M r is M r defined up to conjugacy by a partial isometry in Mn(C) ⊗ M0. 0 := p(Mn(C) ⊗ M0)p and M r := p(Mn(C) ⊗ M )p. As such, the inclusion M r Apart from statement (b), the following result is essentially contained in [IPV10, Proposition 7.2]. For completeness, we nevertheless give a full proof. At a first reading of Proposition 4.1, one may very well assume that M0 = M , which is sufficient to prove Theorem B.1. The most general setup is only needed to prove Theorem B.2. Proposition 4.1. Let M0 be a II1 factor and r > 0 such that M r 0 = LΛ for some countable group Λ. As above, denote by ∆ : M0 → (M0 ⊗ M0)r the amplified comultiplication. Assume that M and fM are tracial von Neumann algebras such that M0 ⊂ M and M0 ⊂ fM . (a) If P ⊂ M is a von Neumann subalgebra and M0 6≺M P , then ∆(M0) 6≺M ⊗M M ⊗ P . (b) If P ⊂ fM is a von Neumann subalgebra and ∆(M0) is amenable relative to M r ⊗P inside M r ⊗ fM , then M0 is amenable relative to P inside fM . (c) If P ⊂ M0 is a von Neumann subalgebra that has no amenable direct summand, then for every nonzero projection q ∈ ∆(P )′ ∩ (M ⊗ M )r, we have that ∆(P )q is nonamenable relative to M r ⊗ 1. Proof. Throughout the proof, we fix a projection p ∈ Mn(C) ⊗ M0 with (Tr ⊗τ )(p) = r. We identify p(Mn(C) ⊗ M0)p = LΛ. (a) Let ∆ : LΛ → LΛ ⊗ LΛ : ∆(vs) = vs ⊗ vs be the original comultiplication. Since M0 6≺M P , also M r 0 6≺M P . By Theorem 2.1, we can take a sequence sn ∈ Λ such that kEP (x∗vsny)k2 → 0 for all x, y ∈ p(Cn ⊗ M ) . 14 We claim that kEM ⊗P (x∗∆(vsn)y)k2 → 0 for all x, y ∈ p(Cn ⊗ M ) ⊗ p(Cn ⊗ M ) . (4.1) Indeed, (4.1) is obvious when x = x1 ⊗ x2 and y = y1 ⊗ y2 are elementary tensors. Then (4.1) follows easily for general x, y as well. By (4.1) and Theorem 2.1, we have ∆(LΛ) 6≺M ⊗M M ⊗P . Then also the conclusion ∆(M0) 6≺M ⊗M M ⊗ P follows. (b) We first state two preliminary observations. (∗) Assume that Q and S are tracial von Neumann algebras and that M rHQ and fM rKS are left LΛ-amenable. To prove (∗), assume that Ω is a ∆(LΛ)-central state on B(H ⊗ K) ∩ (Qop ⊗ Sop)′ whose bimodules. If the (M r ⊗ fM r)-(Q ⊗ S)-bimodule H ⊗ K is left ∆(LΛ)-amenable, then fM rKS is restriction to M r ⊗ fM r equals the trace. Then the formula Ω0(T ) := Ω(1 ⊗ T ) defines a state on B(K) ∩ (Sop)′ that is (vs)s∈Λ-central and whose restriction to fM r equals the trace. In combination with the Cauchy-Schwarz inequality, it follows that Ω0 is actually LΛ-central. This concludes the proof of (∗). (∗∗) Assume that S is a tracial von Neumann algebra and that fMKS is a bimodule. We leave it to the reader to check that fMKS is left M0-amenable if and only if the bimodule fM r(p(Cn ⊗ K))S is left M r 0 -amenable. We are now ready to prove (b). By our assumptions, the bimodule M r ⊗ fML2(M r ⊗ fM )M r ⊗ P is left ∆(M0)-amenable. From (∗∗), we get that M r ⊗ fM rL2(M r ⊗ p(Cn ⊗ fM ))M r ⊗ P 0 -amenable. is left ∆(M r 0 )-amenable. It then follows from (∗) that fM r(p(Cn ⊗ L2(fM )))P is left M r Again using (∗∗), we get that fML2(fM )P is left M0-amenable, i.e. that M0 is amenable relative to P inside fM . (c) Assume that LΛKLΛ is an arbitrary bimodule. Denote by λ : L(Λ) → B(K) and ρ : (LΛ)op → B(K) the normal ∗-homomorphisms given by the left, resp. right bimodule action. It is easy to check that there is a unique normal ∗-homomorphism Ψ : LΛ⊗ (LΛ)op → B(K ⊗ K ⊗ K) : Ψ(vs ⊗ vop t ) = λ(vs)ρ(vop t )⊗ λ(vs)⊗ ρ(vop t ) for all s, t ∈ Λ . It follows in particular that the LΛ-LΛ-bimodule K ⊗ K ⊗ K given by vs · (ξ1 ⊗ ξ2 ⊗ ξ3) · vt = (vsξ1vt) ⊗ (vsξ2) ⊗ (ξ3vt) is contained in a multiple of the coarse LΛ-LΛ-bimodule. Applying this statement to the bimodule LΛL2(M r)LΛ, it follows that the ∆(LΛ)-∆(LΛ)-bimodule is contained in a multiple of the coarse ∆(LΛ)-∆(LΛ)-bimodule. Then also ∆(LΛ)(cid:0)L2(M r ⊗ M r) ⊗M r⊗1 L2(M r ⊗ M r)(cid:1)∆(LΛ) ∆(M0)(cid:0)L2(M r ⊗ M ) ⊗M r⊗1 L2(M r ⊗ M )(cid:1)∆(M0) is contained in a multiple of the coarse ∆(M0)-∆(M0)-bimodule. Assume now that q ∈ ∆(P )′ ∩ (M ⊗ M )r is a nonzero projection such that ∆(P )q is amenable relative to M r ⊗1. We must prove that P has an amenable direct summand. By our assumption (4.2) 15 and [PV11, Proposition 2.4.3], the bimodule M r ⊗ M(L2(M r ⊗ M )q)∆(P )q is weakly contained in the bimodule M r ⊗ M(cid:0)L2(M r ⊗ M ) ⊗M r⊗1 L2(M r ⊗ M )q(cid:1)∆(P )q . Viewing L2(∆(P )q) as a subspace of L2(M r ⊗ M )q, it follows that ∆(P )qL2(∆(P )q)∆(P )q is weakly contained in the bimodule ∆(P )q(cid:0)qL2(M r ⊗ M ) ⊗M r⊗1 L2(M r ⊗ M )q(cid:1)∆(P )q . Since the bimodule in (4.2) is contained in a multiple of the coarse ∆(M0)-∆(M0)-bimodule, we conclude that the trivial ∆(P )q-∆(P )q-bimodule is weakly contained in the coarse ∆(P )q-∆(P )q-bimodule. Hence ∆(P )q has an amenable direct summand. Then also P has an amenable direct summand. Remark 4.2. Assume that M0 is a II1 factor and r > 0 such that M r group Λ. Consider the comultiplication 0 = LΛ for some countable ∆ : LΛ → LΛ ⊗ LΛ : ∆(vs) = vs ⊗ vs for all s ∈ Λ . Take a projection p ∈ Mn(C) ⊗ M0 with (Tr ⊗τ )(p) = r and realize M r Realize (M0 ⊗ M0)r as M r ∆0 : M0 → M r Denote by ζ : Mn(C) ⊗ M0 → M0 ⊗ Mn(C) the flip isomorphism. Put 0 = p(Mn(C) ⊗ M0)p. 0 ⊗ M0. The relation between ∆ and the amplified comultiplication 0 ⊗ M0 can be concretized in the following slightly painful way. ∆1 := (id ⊗ id ⊗ ζ −1) ◦ (∆0 ⊗ id) ◦ ζ , which is a unital ∗-homomorphism from Mn(C) ⊗ M0 to M r element Z ∈ M r for all x ∈ M r 0 . 0 ⊗ Mn(C) ⊗ M0. We then find an 0 ⊗ Mn(C) ⊗ M0 such that Z ∗Z = ∆1(p), ZZ ∗ = p ⊗ p and ∆(x) = Z∆1(x)Z ∗ 5 Normalizers of relatively amenable subalgebras Throughout this section, we work in the following setup and under the following assumptions. We refer to Sections 2.4 and 2.5 for the definitions of weak amenability, class S and property Gamma. Setup. We are given a II1 factor M0, a countable group Λ and a number r > 0 such that M r 0 = LΛ. We assume that M0 ⊂ M , where M is of the form M = B ⋊ Γ for a given trace preserving action Γ y (B, τ ) of a countable group Γ. We denote by ∆ : M0 → M r 0 ⊗ M0 the amplified comultiplication, as in Section 4. Assumptions. 1. The group Γ satisfies one of the following conditions. (a) Γ is nonamenable, weakly amenable and in class S. (b) Γ = Γ1 ∗ Γ2 with Γ1 ≥ 2 and Γ2 ≥ 3 and M ′ 0 ∩ M ω = C1. 2. We have ∆(M0)′ ∩ M r ⊗ M = C1. 3. If Γ0 < Γ is a subgroup of infinite index, we have that M0 6≺M B ⋊ Γ0. 16 4. We have that M0 is nonamenable relative to B inside M . At a first reading, one may very well assume that M0 = M . In that case, assumption 2 follows because Λ is an icc group, while assumptions 3 and 4 are trivially satisfied. This will be enough to prove Theorem B.1. The general situation is only needed to prove Theorem B.2. The following theorem is a direct consequence of the main results in [PV12] and [Io12b]. Theorem 5.1. Assume that we are in the setup and under the assumptions described above. If P ⊂ M r ⊗ M is a von Neumann subalgebra such that ∆(M0) ⊂ NM r⊗M (P )′′ and such that P is amenable relative to M r ⊗ B, then P ≺f M r ⊗ B. Proof. Throughout the proof, we view M r ⊗ M as the crossed product (M r ⊗ B) ⋊ Γ. By assumption 2, we have that ∆(M0)′ ∩ (M ⊗ M )r = C1. Since ∆(M0) ⊂ NM r⊗M (P )′′, by Lemma 2.2.(a), it suffices to prove that P ≺ M r ⊗ B. First assume that Γ satisfies assumption 1.(a). By [PV12, Theorem 1.4], we have that either ∆(M0) is amenable relative to M r ⊗ B, or that P ≺ M r ⊗ B. Using Proposition 4.1.(b) and assumption 4, we see that the first option is impossible. So we indeed get that P ≺ M r ⊗ B. Next assume Γ satisfies assumption 1.(b). We apply the main results of [Io12b] and need to introduce some of the corresponding notations. We extend the action Γ y B to an action Γ∗ F2 y B by letting F2 act trivially. Denote fM := B ⋊ (Γ∗ F2). View F2 as the free product of t := exp(ithj ) and define the one parameter group of automorphisms θt of fM two copies of Z that we denote as (Z)j, j = 1, 2. Choose self-adjoint elements hj in L(Z)j with spectrum [−π, π] and with the property that exp(ihj ) is the canonical unitary that generates L(Z)j. Denote uj given by θt(x) = uj t )∗ for all x ∈ B ⋊ (Γj ∗ (Z)j). Denote by L the kernel of the natural surjective homomorphism of Γ ∗ F2 onto F2 and put N := B ⋊ L. Observe that we can view t x(uj Fix t ∈ (0, 1). Since P is amenable relative to M r ⊗ B inside M r ⊗ M , we have that (id ⊗ θt)(P ) M r ⊗ fM as the crossed product M r ⊗ fM = (M r ⊗ N ) ⋊ F2. is amenable relative to M r ⊗ θt(B) inside M r ⊗ fM . Since θt(B) = B ⊂ N , we get a fortiori that (id ⊗ θt)(P ) is amenable relative to M r ⊗ N . By [PV11, Theorem 1.6], we get that either (id ⊗ θt)(P ) ≺ M r ⊗ N , or that (id ⊗ θt)∆(M0) is amenable relative to M r ⊗ N . So we are in one of the following situations. Case 1. There exists a t ∈ (0, 1) such that (id ⊗ θt)(P ) ≺ M r ⊗ N . Using Proposition 4.1.(a) and assumption 3, we know that for all j ∈ {1, 2}, we have ∆(M0) 6≺ M r ⊗ (B ⋊ Γj). Since M r ⊗ M is the amalgamated free product of M r ⊗ (B ⋊ Γ1) and M r ⊗ (B ⋊ Γ2), amalgamated over M r ⊗ B, it follows from [Io12b, Theorem 3.2] that P ≺ M r ⊗ B. Case 2. For all t ∈ (0, 1), we have that (id ⊗ θt)∆(M0) is amenable relative to M r ⊗ N . Fix t ∈ (0, 1). We get that ∆(M0) is amenable relative to M r ⊗ θ−t(N ) inside M r ⊗ fM . By Proposition 4.1.(b), we conclude that M0 is amenable relative to θ−t(N ) inside fM . Hence θt(M0) is amenable relative to N inside fM . So θt(M0) is amenable relative to N inside fM 0 ∩ M ω = C1. Viewing M as the for all t ∈ (0, 1). Also it is part of assumption 1.(b) that M ′ amalgamated free product of B ⋊ Γ1 and B ⋊ Γ2, amalgamated over B, it follows from [Io12b, Theorem 5.1] that either M0 ≺ B ⋊ Γj for some j ∈ {1, 2}, or that M0 is amenable relative to B inside M . The first option contradicts assumption 3, while the second option contradicts assumption 4. Hence case 2 is ruled out. 17 6 Left-right wreath products and inner amenability We need the following elementary results on left-right wreath products H (Γ) ⋊ (Γ × Γ), where the direct product group Γ × Γ acts on the set Γ by left-right multiplication: (g, h) · k = gkh−1. We refer to Section 2.5 for the definition of inner amenability. Proposition 6.1. Let H and Γ be arbitrary countable groups with H 6= {e}. Write H := H (Γ) and consider the left-right wreath product G := H ⋊ (Γ × Γ). Denote by H1 the abelianization of H with quotient map p1 : H → H1. Define the homomorphism p : H → H1 : p(x) =Xg∈Γ p1(xg) . Denote by H0 the kernel of p and define G0 := H0 ⋊ (Γ × Γ). (a) If Γ is not inner amenable, also G is not inner amenable. Even more so, the unitary representation (Ad g)g∈Γ×Γ on ℓ2(G − {e}) does not have almost invariant vectors. So any subgroup of G that contains Γ × Γ is not inner amenable. (b) If Γ is nonamenable and finitely generated and if Γ has trivial center, then G is not inner amenable. (c) If Γ is infinite and has trivial center, then G0 and G are icc groups and (LH0)′ ∩ LG ⊂ LH and (LG0)′ ∩ LG = C1 . Statement (b) in the above proposition is not used in the paper. We added it in order to put it in contrast with Remark 6.2, where we show that there are nonamenable icc groups Γ such that LG is a McDuff II1 factor, and in particular such that G is not W∗-superrigid. Proof. Throughout the proof, we write G := Γ× Γ. We denote by PG the orthogonal projection of ℓ2(G) onto ℓ2(G). The action (Ad g)g∈Γ×{e} on G − G has finite stabilizers. Therefore, the restriction of the representation (Ad g)g∈Γ×{e} to the invariant subspace ℓ2(G − G) is weakly contained in the regular representation of Γ. (a) Assume that ξn ∈ ℓ2(G − {e}) is a sequence of vectors that is almost invariant under (Ad g)g∈G. By the remark in the first paragraph and because Γ is nonamenable, it follows that kξn − PG(ξn)k2 → 0. Note that (PG(ξn)) is a sequence of vectors in ℓ2(G − {e}) that is almost invariant under (Adg)g∈G. Since Γ is not inner amenable, also G is not inner amenable. Hence kPG(ξn)k2 → 0. So also kξnk2 → 0. (b) Assume that ξn ∈ ℓ2(G − {e}) is a sequence of vectors that is almost invariant under (Ad g)g∈G . By the remark in the first paragraph and because Γ is nonamenable, it follows that kξn − PG(ξn)k2 → 0. Fix an element s ∈ H − {e}. For every k ∈ Γ, denote by sk ∈ H (Γ) the element s viewed in position k. It is easy to check that PG ◦ (Ad sk) ◦ PG = PStab k. Since kξn −PG(ξn)k → 0 and since the sequence (ξn) is almost invariant under (Ad g)g∈G , we conclude that kξn − PStab F (ξn)k → 0 for every finite subset F ⊂ Γ. If {k1, . . . , kr} is a finite generating set for Γ, one checks that Stab{e, k1, . . . , kr} = {(g, g) g ∈ Center Γ}. Since Γ has trivial center, we get that kξnk → 0. (c) We start by proving the following claim: for every g ∈ G − {e}, there exist infinitely many k ∈ Γ such that g · k 6= k. To prove this claim, denote δ : Γ → G : δ(h) = (h, h). If g ∈ G is not conjugate with an element in δ(Γ), we have g · k 6= k for all k ∈ Γ and the claim is trivial. If 18 g ∈ G − {e} is conjugate with the element δ(h) ∈ δ(Γ), we may actually assume that g = δ(h). Given k ∈ Γ, we have g · k = k if and only if h commutes with k. So if g · k = k for all but finitely many k ∈ Γ, it follows that the centralizer Γ0 := CentrΓ(h) of h inside Γ has a finite complement. Since Γ0 < Γ is a subgroup and Γ is infinite, this implies that Γ0 = Γ. So h lies in the center of Γ. This is impossible, because we assumed that Γ has trivial center and that δ(h) = g 6= e. Having proven the claim above, we show that for every x ∈ G−H, we have that {zxz−1 z ∈ H0} is infinite. We write x = yg with y ∈ H and g ∈ G − {e}. Define F0 := {e} ∪ {g · e} ∪ {k ∈ Γ yk 6= e} . By the claim in the previous paragraph, we can inductively choose elements kn ∈ Γ such that g · kn 6= kn for all n and such that the sets F0, {k1, g · k1}, {k2, g · k2}, ... are all disjoint. Fix an element s ∈ H − {e}. For every k ∈ Γ, denote by sk ∈ H (Γ) the element s viewed in position k. Define the sequence of elements zn ∈ H0 given by zn := s−1 e skn. Since znxz−1 n = s−1 e skn y s−1 g·kn sg·e g , we get that all elements znxz−1 x ∈ G − H. This means that (LH0)′ ∩ LG ⊂ LH. It remains to prove that (LG0)′ ∩ LG = C1. Because of the previous paragraph, it suffices to observe that elements in H − {e} have an infinite conjugacy class under (Ad g)g∈Γ×{e}. n are distinct. So the set {zxz−1 z ∈ H0} is infinite for every Remark 6.2. There are nonamenable icc groups Γ such that G := H (Γ) ⋊ (Γ × Γ) is inner amenable, and even such that LG is a McDuff II1 factor (see Section 2.5 for terminology). Indeed, it suffices that Γ admits two sequences of elements (gn), (hn) with the property that gn and hn do not commute, but eventually commute with any fixed element of Γ. In that case, u(gn,gn) and u(hn,hn) form two noncommuting central sequences in LG, forcing LG to be McDuff. Such sequences can be easily found in the icc group S∞ of finite permutations of N, and hence also in the nonamenable icc group F2 × S∞. Because of the previous paragraph, not all nonamenable left-right wreath product groups are W∗-superrigid. 7 Comultiplications and relative commutants Lemma 7.1. Let G and G be countable groups and γi : G → G group homomorphisms, with i = 1, 2. Assume that for every h ∈ G − {e}, the set {γ1(g)hγ1(g)−1 g ∈ G} is infinite. Then the following statements are equivalent. (a) There exists an h ∈ G such that γ1(g) = hγ2(g)h−1 for all g ∈ G. (b) There exists a finite subset F ⊂ G such that F ∩ γ1(g)Fγ2(g)−1 6= ∅ for all g ∈ G. (c) The unitary representation π : G → U (ℓ2G) : π(g)ξ = uγ1(g)ξu∗ γ2(g) is not weakly mixing. 19 Proof. The equivalence of (b) and (c) follows from Lemma 2.12. The implication (a) ⇒ (b) is trivial by taking F = {h}. Conversely assume that (b) holds. By Lemma 2.12, we can take an h ∈ G such that F1 := {γ1(g)hγ2(g)−1 g ∈ G} is a finite set. It follows that F1F −1 is a finite subset of G that is globally invariant under (Ad γ1(g))g∈G. By our assumptions, it follows that F1F −1 1 = {e}. This means that F1 is a singleton. So F1 = {h} and we conclude that γ1(g) = hγ2(g)h−1 for all g ∈ G. 1 Lemma 7.2. Let Λ be an icc group and α, β ∈ Aut(LΛ). Denote by (vs)s∈Λ the canonical group of unitaries generating LΛ. Let ∆ : LΛ → LΛ ⊗ LΛ : ∆(vs) = vs ⊗ vs be the comultiplication. If (α ⊗ β)∆(LΛ) ≺ ∆(LΛ), there exist unitaries V, W ∈ LΛ, characters ω, µ : Λ → T and an automorphism δ ∈ Aut(Λ) such that α(vs) = ω(s) V vδ(s)V ∗ and β(vs) = µ(s) W vδ(s)W ∗ for all s ∈ Λ. Proof. We start by proving the following claim: if Λ admits a sequence of elements sn ∈ Λ such that kE∆(LΛ)(v∗ xα(vsn) ⊗ β(vsn)v∗ y)k2 = 0 lim n for all x, y ∈ Λ , (7.1) then (α ⊗ β)∆(LΛ) 6≺ ∆(LΛ). Indeed, if (7.1) holds, we multiply left and right by elements of the form ∆(va), ∆(vb) and conclude that lim n kE∆(LΛ)(cid:0)(vx ⊗ vy) (α ⊗ β)∆(vsn) (va ⊗ vb)(cid:1)k2 = 0 for all x, y, a, b, ∈ Λ . Using k · k2-approximations, it follows that the same holds when we replace vx ⊗ vy and va ⊗ vb by arbitrary elements of LΛ ⊗ LΛ. This then means that (α ⊗ β)∆(LΛ) 6≺ ∆(LΛ) and hence this proves the claim. Our assumption is that (α ⊗ β)∆(LΛ) ≺ ∆(LΛ). So by the claim above, there is no sequence of elements sn ∈ Λ satisfying (7.1). This means that there are finitely many xi, yi ∈ Λ, with i = 1, . . . , k, and a δ > 0 such that kXi=1 kE∆(LΛ)(v∗ xiα(vs) ⊗ β(vs)v∗ yi)k2 2 ≥ δ The left hand side can be computed and we conclude that kXi=1Xt∈Λ τ (v∗ xitα(vs))2 τ (v∗ tyiβ(vs))2 ≥ δ for all s ∈ Λ . for all s ∈ Λ . (7.2) As in [IPV10, Formula (3.1)], we define the height of an element a ∈ LΛ as Using (7.2), we find that for all s ∈ Λ, we have hΛ(a) := max{τ (v∗ t a) t ∈ Λ} . δ ≤ kXi=1Xt∈Λ τ (v∗ xitα(vs))2 τ (v∗ tyi β(vs))2 ≤ hΛ(α(vs))2 kXi=1Xt∈Λ τ (v∗ tyiβ(vs))2 = k hΛ(α(vs))2 . 20 So we get that hΛ(α(vs)) ≥ pδ/k for all s ∈ Λ. It then follows from [IPV10, Theorem 3.1] that there exist a unitary V ∈ LΛ, a character ω : Λ → T and an automorphism δ1 ∈ Aut(Λ) such that α(vs) = ω(s) V vδ1(s)V ∗ for all s ∈ Λ. By symmetry, we find the same description of the automorphism β, yielding a unitary W ∈ LΛ, a character µ : Λ → T and an automorphism δ2 ∈ Aut(Λ) such that β(vs) = µ(s) W vδ2(s)W ∗ for all s ∈ Λ. It remains to prove that up to an inner conjugacy, δ1 = δ2. Replacing α by (Ad V ∗) ◦ α and replacing β by (Ad W ∗) ◦ β, we still have that (α ⊗ β)∆(LΛ) ≺ ∆(LΛ). So there exist finitely many xi, yi ∈ Λ, with i = 1, . . . , k, and a δ > 0 such that (7.2) holds. Since now α(vs) = ω(s) vδ1(s) and β(vs) = µ(s) vδ2(s), the left hand side of (7.2) is zero, unless there exists an i ∈ {1, . . . , k} and a t ∈ Λ satisfying δ1(s) = xit and δ2(s) = tyi. This means that for every s ∈ Λ, there exists an i ∈ {1, . . . , k} such that δ1(s)yiδ2(s)−1 = xi. Since Λ is icc, it then follows from Lemma 7.1 that δ1 and δ2 are equal up to inner conjugacy. Let Λ be an icc group and assume that LΛ does not have property Gamma, so that Out(LΛ) we denote by αω the automorphism of LΛ given by αω(vs) = ω(s) vs for all s ∈ Λ. Using the is a Polish group (see Section 2.5 for notations and terminology). For every character ω ∈ bΛ, icc property, one checks that the map ω 7→ αω embeds bΛ continuously into Out(LΛ). Since bΛ is compact, we can thus view bΛ as a compact subgroup of Out(LΛ). A countable subgroup A of a Polish group B is said to be discrete if there exists a neighborhood U of the identity e in B such that U ∩ A = {e}. Lemma 7.3. Let M0 be a II1 factor without property Gamma. Let r > 0 and M r 0 = LΛ for some countable group Λ. Denote by ∆ : M0 → (M0 ⊗ M0)r the amplified comultiplication as in Section 4. Assume that M is a tracial von Neumann algebra with M0 ⊂ M and M ′ 0 ∩ M = C1. Let L ⊂ NM (M0) be a subgroup such that M = (M0 ∪ L)′′. Finally assume that the image of L in Out(M0) is a discrete torsion-free subgroup. Then the following holds. (a) If H ⊂ L2((M ⊗ M )r) ⊖ L2(∆(M0)) is a nonzero ∆(M0)-∆(M0)-subbimodule of finite left ∆(M0)-dimension, then there exist automorphisms β1, . . . , βk ∈ Aut(M0) and a unitary ψ : H → L2(M0)⊕k : ξ 7→ (ψ1(ξ), . . . , ψk(x)) such that ψi(∆(x) ξ ∆(y)) = x ψi(ξ) βi(y) for all x, y ∈ M0, ξ ∈ H, i = 1, . . . , k, and such that every βi generates a discrete infinite subgroup of Out(M0). (b) We have ∆(M0)′ ∩ (M ⊗ M )r = C1. Note that the setup of Lemma 7.3 would allow to write M as the cocycle crossed product of M0 and an outer cocycle action of L/(L ∩ Inn M0) on M0, but we do not need that formalism here. Proof. First note that statement (b) is a consequence of statement (a). Take an element T in ∆(M0)′ ∩ (M ⊗ M )r and write S := T − E∆(M0)(T ). Since M0 is a factor, it suffices to prove that S = 0. So assume that S 6= 0. Denote by H the closure of ∆(M0)S. Then H is a ∆(M0)-∆(M0)-subbimodule of L2((M ⊗ M )r) ⊖ L2(∆(M0)) that has finite left dimension. By construction H contains the nonzero vector S satisfying ∆(x)S = S∆(x) for all x ∈ M0. Write H as in (a). Since all automorphisms βi are outer, we have that ψi(S) = 0 for all i ∈ {1, . . . , k}. So S = 0, contradicting our assumption. 21 0 := p(Mn(C) ⊗ M0)p and (M0 ⊗ M0)r = M r We now start proving statement (a). Take a projection p ∈ Mn(C) ⊗ M0 with (Tr ⊗τ )(p) = r. Realize M r 0 ⊗ M0. Denote by ∆ : LΛ → LΛ ⊗ LΛ : ∆(vs) = vs ⊗ vs the original comultiplication. During the proof, to improve the clarity of the exposition, we denote the amplified comultiplication by ∆0 : M0 → M r 0 ⊗ M0. The relation between ∆0 and ∆ has been concretized in Remark 4.2. Put M r := p(Mn(C) ⊗ M )p so that literally M r 0 ⊂ M r. Let H ⊂ L2(M r ⊗ M ) ⊖ L2(∆0(M0)) be a ∆0(M0)-∆0(M0)-subbimodule of finite left ∆0(M0)-dimension. Using the notation of Remark 4.2, we put H′ := Z (id ⊗ id ⊗ ζ −1)(H ⊗ Mn(C)) Z ∗ and notice that H′ ⊂ L2(M r ⊗ M r) ⊖ L2(∆(M r finite left ∆(M r phisms β1, . . . , βk ∈ Aut(M r that 0 )-subbimodule of 0 )-dimension. To conclude the proof of the lemma, we have to find automor- 0 )⊕k : ξ 7→ (ψ1(ξ), . . . , ψk(x)) such 0 ) and a unitary ψ : H′ → L2(M r 0 )) is a nonzero ∆(M r 0 )-∆(M r ψi(∆(x) ξ ∆(y)) = x ψi(ξ) βi(y) for all x, y ∈ M r 0 , ξ ∈ H′, i = 1, . . . , k, and such that every βi generates an infinite discrete subgroup of Out(M r By our assumptions on M0 ⊂ M , we can choose a subset L0 ⊂ NM (M0) such that 0 ). L2(M ) = L2(M0) ⊕ MV ∈L0 L2(M0)V and such that for every V ∈ L0, the automorphism Ad V of M0 generates a discrete infinite subgroup of Out(M0). Fix V ∈ L0. Take a partial isometry v ∈ Mn(C) ⊗ M0 such that vv∗ = p and v∗v = (id ⊗ Ad V )(p). Write V ′ := v(1 ⊗ V ) and note that V ′ ∈ NM r (M r 0 ). As such, we find a subset L1 ⊂ NM r (M r 0 ) such that L2(M r) = L2(M r 0 ) ⊕ MV ∈L1 L2(M r 0 )V and such that for every V ∈ L1, the automorphism Ad V of M r subgroup of Out(M r Define the subset L2 ⊂ NM r⊗M r (M r 0 ) given by 0 ⊗ M r 0 ). 0 generates a discrete infinite L2 := {1 ⊗ V V ∈ L1} ∪ {V ⊗ 1 V ∈ L1} ∪ {V1 ⊗ V2 V1, V2 ∈ L1} . We get that L2(M r ⊗ M r) = L2(M r 0 ⊗ M r 0 ) ⊕ MW ∈L2 L2(M r 0 ⊗ M r 0 )W . (7.3) 0 ⊗ M r 0 is of the form αW ⊗ βW , where 0 ⊗M r 0 ) 0 )) is a ∆(M r 0 )- 0 )-subbimodule of finite left dimension. By [IPV10, Proposition 7.2.3], we get that H0 = Also for every W ∈ L2, the automorphism Ad W of M r at least one of the αW , βW generates a discrete infinite subgroup of Out(M r Denote by P0 the orthogonal projection of L2(M r ⊗M r) onto the closed subspace L2(M r and define H0 as the closure of P0(H′). Then H0 ⊂ L2(M r ∆(M r {0}. For every W ∈ L2, denote by PW the orthogonal projection of L2(M r ⊗ M r) onto the closed subspace L2(M r 0 ) ⊖ L2(∆(M r 0 ⊗ M r 0 ). 0 )W and define 0 ⊗ M r ϕW : L2(M r ⊗ M r) → L2(M r 0 ⊗ M r 0 ) : ϕW (ξ) = PW (ξ)W ∗ . 22 Since W normalizes M r 0 ⊗ M r 0 and since (7.3) is an orthogonal decomposition, we get that ϕW (∆(x) ξ ∆(y)) = ∆(x) ϕW (ξ) (αW ⊗ βW )∆(y) 0 and all ξ ∈ L2(M r ⊗ M r). Denote by HW the closure of ϕW (H′). Below we for all x, y ∈ M r prove the following statement: if HW 6= {0}, then there exists a unitary ψW : HW → L2(M r 0 ) and an automorphism γW ∈ Aut(M r 0 ) such that ψW (∆(x) ξ (αW ⊗ βW )∆(y)) = x ψW (ξ) γW (y) (7.4) for all x, y ∈ M r 0 and all ξ ∈ HW , and such that γW generates a discrete infinite subgroup of Out(M r 0 ). For the moment, we assume that the statement is proven and deduce the lemma from it. Whenever HW 6= {0}, we denote by KW the M r 0 ) with bimodule action x · ξ · y = xξγW (y). Then ψW ◦ ϕW : H′ → KW is a bimodular map with dense range. So, KW is isomorphic with a subbimodule of H′. Since H′ has finite left dimension and since H0 = {0}, it follows that H′ is isomorphic with the direct sum of finitely many KW 's. This proves the lemma. 0 -bimodule L2(M r 0 -M r 0 ) ≺ ∆(M r 0 )-(αW ⊗ βW )∆(M r 0 )-subbimodule of L2(M r 0 ⊗ M r 0 ). We get in particular that HW is a nonzero ∆(M r 0 ) that has finite left dimension. So it remains to prove the statement above. Assume that HW 6= {0}. By construction, HW is a ∆(M r 0 ) of finite left dimension. By Theorem 2.1, this means that (αW ⊗ βW )∆(M r 0 ). By Lemma 7.2, there exist characters ω, µ : Λ → T and an automorphism δ ∈ Aut(Λ) such that, after unitarily conjugating αW and βW , we have that αW (vs) = ω(s) vδ(s) and βW (vs) = µ(s) vδ(s) for all s ∈ Λ. Note that (αW ⊗ βW )∆(vs) = ∆(γW (vs)), where the automorphism γW ∈ Aut(M r 0 ) is defined by the formula γW (vs) = ω(s)µ(s) vδ(s). So (αW ⊗ βW )∆(M r 0 ) = ∆(M r 0 )- 0 ⊗ M r subbimodule of L2(M r It then follows from [IPV10, Proposition 7.2.3] that HW ⊂ L2(∆(M r 0 is a factor and HW 6= {0}, we get that 0 ) as being ∆−1. By construction, HW = L2(∆(M r (7.4) holds. It remains to prove that γW generates an infinite discrete subgroup of Out(M r 0 ). We know that at least one of the αW , βW generates an infinite discrete subgroup of Out(M r 0 ). Assume that this is the case for αW . View αW as an element of Out(M r compact subgroup of Out(M r 0 ) and view bΛ as a normalizes bΛ. Since bΛ is compact and since αW generates an infinite discrete subgroup, it follows that bΛ and αW together generate a copy of bΛ ⋊ Z as a closed subgroup of Out(M r Since γW ∈ αWbΛ, it then follows that also γW generates an infinite discrete subgroup of 0 ). Since αW (vs) = ω(s) vδ(s) for all s ∈ Λ, we have that αW 0 )). We can thus define ψ : HW → L2(M r 0 )). Since M r 0 )-∆(M r 0 ). Out(M r 0 ). For later use, we end this section with yet another elementary lemma. Lemma 7.4. Let Λ be a countable group and ∆ : LΛ → LΛ ⊗ LΛ the comultiplication given by ∆(vs) = vs ⊗ vs for all s ∈ Λ. If α, β ∈ Aut(LΛ) are automorphisms that satisfy (α ⊗ id) ◦ ∆ = ∆ ◦ β, then there exists a character ω : Λ → T such that α = β = αω, where αω(vs) = ω(s)vs for all s ∈ Λ. Proof. Since ∆(β(vs)) = α(vs) ⊗ vs, we see that α(vs) ⊗ vs ∈ ∆(LΛ). This implies that α(vs) must be a multiple of vs, for all s ∈ Λ. So we find a character ω : Λ → T such that α = αω. But then also β = αω. 23 8 Proof of Theorem B Theorem B will be a direct consequence of the following general result. Recall from Section 2.4 the notions of weak amenability and class S. Theorem 8.1. Let Γ be a countable group satisfying one of the following conditions. 1. Γ is nonamenable, icc, weakly amenable, belongs to class S and admits a bound on the orders of its finite subgroups. 2. Γ = Γ1 ∗ Γ2 with Γ1 ≥ 2 and Γ2 ≥ 3. Let H be a nontrivial abelian group with subgroup H0 < H. Assume that H/H0 is either trivial or torsion-free. Define H := H (Γ) and consider the homomorphism pH : H (Γ) → H : pH(x) =Xg∈Γ xg . H (H0) and G0 := H0 ⋊ (Γ × Γ). Denote H0 := p−1 If Λ is any countable group and π : LΛ → (LG0)r is a ∗-isomorphism for some r > 0, then r = 1 and Λ ∼= p−1 0 < H ′ such that H0 = H ′ More precisely, there exist group isomorphisms δ : Λ → p−1 0) ⋊ (Γ × Γ) for some abelian group H ′ with subgroup H ′ 0 and H/H0 ∼= H ′/H ′ 0. 0) ⋊ (Γ × Γ) and γ : H ′/H ′ H ′ (H ′ 0 → H ′ (H ′ θ(η) for all k ∈ \H ′/H ′ that π = Ad w ◦ αω ◦ πθ ◦ πδ where H/H0, a probability measure preserving isomorphism θ : cH ′ → bH satisfying θ(k + η) =bγ(k) + 0 and a.e. η ∈ cH ′, a character ω : G0 → T and a unitary w ∈ LG0 such 0) ⋊ (Γ × Γ)(cid:1) is the ∗-isomorphism given by πδ(vs) = uδ(s) for all • πδ : LΛ → L(cid:0)p−1 H (H0)⋊(Γ×Γ)(cid:1) is the natural ∗-isomorphism associated • πθ : L(cid:0)p−1 0)⋊(Γ×Γ)(cid:1) → L(cid:0)p−1 with an infinite tensor product of copies of θ ; H ′ (H ′ H ′ (H ′ s ∈ Λ ; • αω is the automorphism of LG0 given by αω(ug) = ω(g)ug for all g ∈ G0. This whole section is devoted to the proof of Theorem 8.1, following closely the strategy of [IPV10] and using many results of [IPV10]. At the end, we will deduce Theorem B, with case B.1 corresponding to the special case where H0 = H, and case B.2 corresponding to H0 = {e}. Throughout this section, we fix a countable icc group Γ that satisfies either condition 1 or condition 2 in Theorem 8.1. We also fix a nontrivial abelian group H with subgroup H0 < H such that H/H0 is either trivial or torsion-free. We denote H := H (Γ) and H0 := p−1 H (H0). We write G := Γ × Γ and we consider the left-right wreath product G := H ⋊ G, with its subgroup G0 := H0 ⋊ G. Put M := LG and M0 := LG0. We finally fix a countable group Λ, a positive number r > 0 and a ∗-isomorphism π : LΛ → M r 0 . To simplify notations, we do not explicitly write π and identify M r As in Section 4, we consider the amplified comultiplication ∆ : M0 → (M0 ⊗ M0)r. Note that the amplified homomorphism ∆ is only defined up to unitary conjugacy (see Remark 4.2 for details). 0 = LΛ. 24 Both condition 1 and condition 2 in Theorem 8.1 guarantee that Γ is not inner amenable. So by Proposition 6.1.(c), G0 and G are icc groups, M0 and M are II1 factors and M ′ 0 ∩ M ω = C1. Also M0 does not have property Gamma and Out(M0) is a Polish group. We write A := LH so that M = A ⋊ G. We also write A0 := LH0 so that M0 = A0 ⋊ G. Recall that a countable subgroup A of a Polish group B is said to be discrete if there exists a neighborhood U of the identity e in B such that U ∩ A = {e}. We start by two general lemmas on the structure of M0 and M . The first one is an immediate consequence of Popa's cocycle superrigidity theorem [Po06b, Theorem 1.1]. Lemma 8.2. Let β ∈ Aut(M0) and assume that there exists a nonzero vector ξ ∈ L2(M0) such that ξ0β(a) = aξ0 for all a ∈ A0. Then there exists a character ω : G → T and a unitary v ∈ NM (M0) such that β = (Ad v) ◦ αω, where the automorphism αω is defined as αω(aug) = ω(g) aug for all a ∈ A0, g ∈ G. If moreover β generates a discrete infinite subgroup of Out(M0), we have that EM0(v) = 0. 0 ∈ A0 and v∗ Proof. Taking the polar decomposition of ξ0, we find a nonzero partial isometry v0 ∈ M0 such that v0β(a) = av0 for all a ∈ A0. By Proposition 6.1.(c), we have that A′ 0 ∩ M0 = A0 and hence, v0v∗ 0v0 ∈ β(A0). Since G y A0 is ergodic, we can extend v0 to a unitary v1 ∈ U (M0) such that v1β(a)v∗ 1 = a for all a ∈ A0. Put β1 = (Ad v1) ◦ β. Since β1(a) = a for all a ∈ A0, we have β1(ug) = µg ug for all g ∈ G, where µg ∈ U (A0) and (µg)g∈G defines a 1-cocycle for the action G y A0. By Popa's cocycle superrigidity theorem [Po06b, Theorem 1.1] for the action G a unitary v2 ∈ U (A) and a character ω : G → T such that µg = ω(g) v∗ follows that v2β1(x)v∗ we have that v ∈ NM (M0) and β = (Ad v) ◦ αω. Finally, if β generates a discrete infinite subgroup of Out(M0), we know that as an element of σ y A, we find 2 σg(v2) for all g ∈ G. It 1v∗ 2, 2 = αω(x) for all x ∈ M0. In particular, v2 ∈ NM (M0). Putting v := v∗ Out(M0), β does not belong to the compact subgroup bG ⊂ Out(M0). So, v 6∈ U (M0). Since 0 ∩ M = C1, it follows that EM0(v) = 0. v ∈ NM (M0) and M ′ Lemma 8.3. Denote by He the copy of H inside H in position e ∈ Γ. Then M is generated by M0 and the group of unitaries L := {us s ∈ He} that normalize M0. The image of L in Out(M0) is a discrete subgroup of Out(M0) that is isomorphic with H/H0. 0 ∩ M ω = C1. So whenever (an) is a sequence Proof. By Proposition 6.1.(a), we know that M ′ of unitaries in U (M ) satisfying kxan − anxk2 → 0 for all x ∈ M0, there exists a sequence λn ∈ T such that kan − λn1k2 → 0. Assume that we have a sequence sn ∈ He such that Ad(usn), viewed as a sequence in Out(M0), converges to the identity. We must prove that sn belongs to (H0)e eventually. Since Ad(usn) converges to the identity in Out(M0), we find a sequence of unitaries wn ∈ U (M0) such that Ad(wnusn) → id in Aut(M0). This means that kxwnusn −wnusnxk2 → 0 for all x ∈ M0. It follows that we can take a sequence λn ∈ T such that kwnusn −λn1k2 → 0. So kusn −λnw∗ nk2 → 0. In particular, we get that kusn −EM0(usn)k2 → 0. Since kusn − EM0(usn)k2 = 1 whenever sn 6∈ (H0)e, we conclude that sn ∈ (H0)e eventually. We now start a systematic study of the amplified comultiplication ∆ : M0 → (M0 ⊗ M0)r. Lemma 8.4. We have that ∆(M0)′ ∩ (M ⊗ M )r = C1. Proof. This is an immediate consequence of Lemma 8.3, the assumption that H/H0 is torsion- free and part (b) of Lemma 7.3. 25 In what follows, we apply twice Theorem 5.1. So we need to check that the assumptions stated in the beginning of Section 5 are satisfied. Lemma 8.5. Both when viewing M as the crossed product M = B ⋊ ({e} × Γ) with B = A ⋊ (Γ × {e}), or as the crossed product M = B ⋊ (Γ × {e}) with B = A ⋊ ({e} × Γ), all assumptions in the beginning of Section 5 are satisfied. More concretely, we have (a) M ′ 0 ∩ M ω = C1, (b) ∆(M0)′ ∩ (M ⊗ M )r = C1, (c) if Γ0 < Γ is a subgroup of infinite index, M0 6≺ A ⋊ (Γ × Γ0) and M0 6≺ A ⋊ (Γ0 × Γ) , (d) M0 is nonamenable relative to A ⋊ (Γ × {e}), and nonamenable relative to A ⋊ ({e} × Γ), inside M . Proof. We already observed above that (a) follows from Proposition 6.1.(a). Statement (b) is given by Lemma 8.4. Statements (c) is straightforward and statement (d) follows from Lemma 2.5. Lemma 8.6. We have ∆(A0) ≺f A ⊗ A. Proof. Because of Lemma 8.5, we can apply Theorem 5.1 to the crossed product decompositions M = (A ⋊ (Γ × {e})) ⋊ Γ and M = (A ⋊ ({e} × Γ)) ⋊ Γ, and the abelian (hence amenable) von Neumann subalgebra ∆(A0) ⊂ M r ⊗ M . We conclude that ∆(A0) ≺f M r ⊗(cid:0)A ⋊ (Γ × {e})(cid:1) and ∆(A0) ≺f M r ⊗(cid:0)A ⋊ ({e} × Γ)(cid:1) . So by Lemma 2.2.(b), we get that ∆(A0) ≺f M r ⊗ A. By symmetry, we also have that ∆(A0) ≺f A ⊗ M r. Again by Lemma 2.2.(b), we conclude that ∆(A0) ≺f A ⊗ A. Lemma 8.7. Let G1 < G be a subgroup of infinite index. Then ∆(LG) 6≺ M r ⊗ (A ⋊ G1) and ∆(LG) 6≺ (A ⋊ G1) ⊗ M r. Proof. By symmetry, it suffices to prove that ∆(LG) 6≺ M r ⊗(A⋊G1). Assume the contrary. A combination of Lemma 8.6 and Lemma 2.3 then gives that ∆(M0) ≺ M r⊗(A⋊G1). Proposition 4.1.(a) now implies that M0 ≺ A ⋊ G1, contradicting the assumption that G1 < G has infinite index. We can view M as the generalized Bernoulli crossed product M = (LH)Γ ⋊ G. As in Section 3, we have the tensor length deformation by automorphisms αt of the tracial von Neumann algebra fM := (LH ∗ LZ)Γ ⋊ G. Lemma 8.8. Let P ⊂ (M ⊗ M )r be a von Neumann subalgebra such that for all nonzero projections p ∈ P ′ ∩ (M ⊗ M )r, we have that P p is nonamenable relative to M r ⊗ 1. Assume that ∆(LG) ⊂ N(M ⊗M )r (P )′′. Then sup k(id ⊗ αt)(b) − bk2 → 0 as t → 0. b∈U (P ′∩(M ⊗M )r) 26 Proof. We concretely realize the amplification (M ⊗ M )r as M r ⊗ M . Since A is abelian and hence amenable, we have that P p is nonamenable relative to M r ⊗A, for all nonzero projections p ∈ P ′ ∩ M r ⊗ M . Case 1: Γ is a nonamenable group in class S with the property that all finite subgroups of Γ have order at most κ − 1, for some fixed κ ∈ N. We consider the left-right action G y Γ. We claim that Stab F is amenable whenever F ⊂ Γ satisfies F ≥ κ. Indeed, every Stab F is isomorphic with a subgroup of Γ defined as the centralizer of κ distinct elements. These κ distinct elements necessarily generate an infinite subgroup of Γ. Since Γ belongs to class S, the centralizer of an infinite subgroup is amenable (see Section 2.4). This proves the claim. So the conclusion of the lemma follows immediately from Theorem 3.1, even without using the assumption that ∆(LG) ⊂ NM r⊗M (P )′′. Case 2: Γ = Γ1 ∗ Γ2 with Γ1 ≥ 2 and Γ2 ≥ 3. Denote by δ : Γ → Γ × Γ : δ(h) = (h, h) the diagonal embedding and consider the left-right action G y Γ. Whenever F ⊂ Γ and F ≥ 2, we have that Stab F is either cyclic, or conjugate to a subgroup of δ(Γ1), or conjugate to a subgroup of δ(Γ2). So the conclusion of the lemma follows from Theorem 3.1, once we have proven that P p is nonamenable relative to M r ⊗ (A ⋊ δ(Γj)) for all j ∈ {1, 2} and all nonzero projections p ∈ P ′ ∩ M r ⊗ M . By symmetry, it suffices to consider j = 1. Take a nonzero projection p ∈ P ′ ∩ M r ⊗ M and assume by contradiction that P p is amenable relative to M r ⊗ (A ⋊ δ(Γ1)). Denote by Q the normalizer of P inside M r ⊗ M . By assumption, we have that ∆(LG) ⊂ Q. Replacing p by the smallest projection in Z(Q) that dominates p and using Lemma 2.6, we still have that P p is amenable relative to M r ⊗ (A ⋊ δ(Γ1)). We now prove that P p is amenable relative to B := M r ⊗ (A ⋊ (Γ × {e})). We denote Mj := M r ⊗ (A ⋊ (Γ × Γj)). We can then view M r ⊗ M as the amalgamated free product of M1 and M2 over B. Since we assumed that P p is amenable relative to M r ⊗ (A ⋊ δ(Γ1)), we have a fortiori that P p is amenable relative to M1. Since p ∈ Z(Q), the normalizer of P p inside M r ⊗ M contains Qp. Since ∆(LG) ⊂ Q and since Γ × Γ1 has infinite index in G, it follows from Lemma 8.7 that Qp 6≺ M1. Then [Io12b, Corollary 2.12] implies that P p is amenable relative to B. By symmetry, we also get that P p is amenable relative to M r ⊗ (A ⋊ ({e} × Γ)). So Lemma 2.7 implies that P p is amenable relative to M r ⊗A, and hence also relative to M r ⊗1, contradicting our initial assumptions on P . Lemma 8.9. There exists a unitary Ω ∈ (M ⊗ M )r such that Ω ∆(LG) Ω∗ ⊂ (LG ⊗ LG)r . Proof. Also M ⊗ M can be viewed as a generalized Bernoulli crossed product M ⊗ M = (LH)I ⋊ (G × G), associated with G × G acting on the disjoint union I := Γ ⊔ Γ of two copies of Γ. The corresponding tensor length deformation precisely is αt ⊗ αt ∈ Aut(fM ⊗ fM ). Denote by δ : Γ → Γ × Γ : δ(h) = (h, h) the diagonal embedding. Observe that the stabilizer (in G × G) of an element i ∈ I is either of the form G × gδ(Γ)g−1 or gδ(Γ)g−1 × G, with g ∈ G. Since G is an icc group, the lemma will follow by applying Theorem 3.3 to the generalized Bernoulli action G × G y (LH)I , provided that we prove the following two statements. 1. sup g∈G k(αt ⊗ αt)∆(ug) − ∆(ug)k2 → 0 as t → 0. 2. ∆(LG) 6≺ M ⊗ (A ⋊ δ(Γ)) and ∆(LG) 6≺ (A ⋊ δ(Γ)) ⊗ M . 27 Proof of 1. By symmetry, it suffices to prove that k(id ⊗ αt)∆(ug) − ∆(ug)k2 → 0 as t → 0 . sup g∈G Since every g ∈ G is the product of an element in Γ × {e} and an element in {e} × Γ, again by symmetry, it suffices to prove that sup k(id ⊗ αt)∆(ug) − ∆(ug)k2 → 0 as t → 0 . (8.1) g∈{e}×Γ Denote P := ∆(L(Γ × {e})). By Proposition 4.1.(c), we have that P p is nonamenable relative to M r ⊗ 1 for all nonzero projections p ∈ P ′ ∩ (M ⊗ M )r. The unitaries ∆(ug), g ∈ {e} × Γ, all commute with P and the normalizer of P contains ∆(LG). So (8.1) follows from Lemma 8.8. Proof of 2. Since δ(Γ) has infinite index in G, statement 2 follows immediately from Lemma 8.7. Lemma 8.10. Write C := ∆(A0)′ ∩ (M ⊗ M )r. Then C ≺f A ⊗ A. Proof. We start by proving the existence of a nonzero projection p ∈ C ′ ∩ (M ⊗ M )r such that Cp is amenable relative to M r ⊗ 1. Assume the contrary. Since the normalizer of C contains ∆(M0) and since all unitaries in ∆(A0) commute with C, it follows from Lemma 8.8 that sup k(id ⊗ αt)∆(a) − ∆(a)k2 → 0 as t → 0. a∈U (A0) Lemma 8.9 implies in particular that k(id ⊗ αt)∆(ug) − ∆(ug)k2 → 0 as t → 0. sup g∈G Note that W := {aug a ∈ U (A0), g ∈ G} is a group of unitaries generating M0. The two formulae above imply that k(id ⊗ αt)∆(b) − ∆(b)k2 → 0 as t → 0. (8.2) sup b∈W We now apply Theorem 3.3. Denote as above δ : Γ → Γ × Γ : δ(h) = (h, h). The stabilizer of an element g ∈ Γ under the left-right action G y Γ can be conjugated into δ(Γ). From Proposition 4.1.(a), we know that ∆(M0) 6≺ M r ⊗ (A ⋊ δ(Γ)). So (8.2) and Theorem 3.3 imply that ∆(M0) can be unitarily conjugated into M r ⊗ LG. This is in contradiction with Proposition 4.1.(a). So we indeed find a nonzero projection p ∈ C ′ ∩ (M ⊗ M )r such that Cp is amenable relative to M r ⊗ 1. The normalizer of C contains ∆(M0) and by Lemma 8.4, we know that ∆(M0)′ ∩ (M ⊗ M )r = C1 . So by Lemma 2.6, we conclude that C is amenable relative to M r ⊗ 1. Applying twice Theorem 5.1, which is possible thanks to Lemma 8.5, it follows that C ≺f M r ⊗ (A ⋊ (Γ × {e})) and that C ≺f M r ⊗ (A ⋊ ({e} × Γ)). It then follows from Lemma 2.2.(b) that C ≺f M r ⊗ A. By symmetry, we also have that C ≺f A ⊗ M r. Again using Lemma 2.2.(b), we reach the desired conclusion that C ≺f A ⊗ A. Lemma 8.11. If H ⊂ ∆(A0)′ ∩ (M ⊗ M )r is a finite-dimensional, globally (Ad ∆(ug))g∈G- invariant subspace, then H ⊂ C1. 28 Proof. Put H′ := {x− E∆(M0)(x) x ∈ H}. The main part of the proof consists in showing that H′ = {0}. Assume on the contrary that H′ 6= {0}. Note that H′ is a finite-dimensional, globally (Ad ∆(ug))g∈G-invariant subspace of ∆(A0)′ ∩ (M ⊗ M )r and that H′ ⊂ (M ⊗ M )r ⊖ ∆(M0). Denote by K the closed linear span of ∆(M0)H′ inside L2((M ⊗ M )r). Observe that K is a ∆(M0)-∆(M0)-subbimodule of L2((M ⊗ M )r) ⊖ L2(∆(M0)) that has finite left dimension. By Lemma 8.3, the assumption that H/H0 is torsion-free, and Lemma 7.3, there exist automor- phisms β1, . . . , βk ∈ Aut(M0) and a unitary ψ : K → L2(M0)⊕k : ξ 7→ (ψ1(ξ), . . . , ψk(ξ)) such that ψi(∆(x) ξ ∆(y)) = x ψi(ξ) βi(y) for all x, y ∈ M0, ξ ∈ K, i = 1, . . . , k, i := ψi(H′)v and note that H′ and such that every βi generates a discrete infinite subgroup of Out(M0). Fix an i ∈ {1, . . . , k} and note that ψi(H′) 6= {0}. Take a nonzero vector ξ0 ∈ ψi(H′). Since the elements of H′ commute with ∆(A0), it follows that that ξ0βi(a) = aξ0 for all a ∈ A0. By Lemma 8.2, we then find a unitary v ∈ NM (M0) and a character ω : G → T such that βi = (Ad v) ◦ αω and such that EM0(v) = 0. Recall that αω(aug) = ω(g) aug for all a ∈ A0 and all g ∈ G. i is a finite-dimensional subspace of L2(M ) such that Put H′ ξa = aξ for all ξ ∈ H′ i, a ∈ A0 and such that H′ i is globally invariant under ξ 7→ ω(g) ugξu∗ g 0 ∩ M = A. So H′ for all g ∈ G. By Proposition 6.1.(c), we have A′ i ⊂ L2(A). It follows that H′ i is a finite-dimensional subspace of L2(A) that is globally invariant under the generalized Bernoulli action G y A. By Lemma 2.12, the latter is weakly mixing. It follows that H′ i ⊂ C1. So, ψi(H′) ⊂ Cv∗. Since ψi(H′) ⊂ L2(M0), while v∗ is orthogonal to L2(M0), we find that ψi(H′) = {0}, which is absurd. So we have proven that H′ = {0}, meaning that H ⊂ ∆(M0). So H = ∆(H0) where H0 ⊂ A′ 0 ∩ M0 = A0 and since the action G y A0 is weakly mixing, it follows that H0 ⊂ C1. Then also H ⊂ C1. 0 ∩ M0 is a finite-dimensional, globally (Ad ug)g∈G-invariant subspace. Since A′ Lemma 8.12. We have that r = 1 and that there exist a unitary v ∈ M0, a character ω : G → T and an injective group homomorphism ρ : G → Λ such that ω(g) vugv∗ = vρ(g) for all g ∈ G and ∆(vA0v∗) ⊂ vA0v∗ ⊗ vA0v∗ . Proof. We view M ⊗ M as the crossed product M ⊗ M = (A ⊗ A) ⋊ (G × G). By Proposition 6.1.(c), we have that (A ⊗ A)′ ∩ (M ⊗ M ) = A ⊗ A, meaning that the generalized Bernoulli action G × G y A ⊗ A is essentially free. By Lemma 8.9 and after a unitary conjugacy of ∆, we have ∆(LG) ⊂ (LG ⊗ LG)r. Put C := ∆(A0)′ ∩ (M ⊗ M )r. From Lemma 8.10, we know that C ≺f A ⊗ A. By construction, the unitaries ∆(ug), g ∈ G, normalize C. By Lemma 8.11, the action (Ad ∆(ug))g∈G on the center Z(C) of C is weakly mixing. Actually, Lemma 8.11 says that even the action (Ad ∆(ug))g∈G on C has no nontrivial finite-dimensional invariant subspaces. This means that all the assumptions of [IPV10, Theorem 6.1] are satisfied. Denote by N the von Neumann algebra generated by C and the unitaries (∆(ug))g∈G. Then ∆(M0) ⊂ N and it follows from Proposition 4.1.(a) that N 6≺ M ⊗ (A ⋊ G1) and N 6≺ (A ⋊ G1) ⊗ M whenever G1 < G has infinite index. So also all the assumptions of [IPV10, Corollary 6.2] are satisfied. From [IPV10, Theorem 6.1 and Corollary 6.2], it then follows that r = 1 and that there exist a unitary Ω1 ∈ M ⊗ M , a character ω : G → T and group homomorphisms γ1, γ2 : G → G such that Ω1∆(A0)Ω∗ 1 ⊂ A ⊗ A and Ω1∆(ug)Ω∗ 1 = ω(g) uγ1(g) ⊗ uγ2(g) . (8.3) 29 Since r = 1, we may from now on assume that M0 = LΛ and that ∆ : M0 → M0 ⊗ M0 is the original comultiplication given by ∆(vs) = vs ⊗ vs for all s ∈ Λ. By (8.3) and Lemma 8.7, the ranges of γ1 and γ2 are finite index subgroups of G. Denote by ζ : M ⊗ M → M ⊗ M : ζ(x ⊗ y) = y ⊗ x the flip automorphism. Since ζ ◦ ∆ = ∆, it follows from (8.3) that (uγ2(g) ⊗ uγ1(g)) ζ(Ω1) Ω∗ 1 (u∗ γ1(g) ⊗ u∗ γ2(g)) = ζ(Ω1) Ω∗ 1 for all g ∈ G . Because G is icc and because the subgroups γ1(G) < G and γ2(G) < G have finite index, we get that {(γ2(g)xγ2(g)−1, γ1(g)yγ1(g)−1) g ∈ G} is an infinite set for all (x, y) ∈ (G × G) − {e}. By Lemma 7.1, we then find an h ∈ G such that γ1(g) = hγ2(g)h−1 for all g ∈ G. This means that after replacing Ω1 by (1 ⊗ uh)Ω1, we may assume that γ1 = γ2. We denote this homomorphism as γ. It then also follows that ζ(Ω1) is a multiple of Ω1. Since ∆(ug) and uγ(g) ⊗ uγ(g) are unitarily conjugate, the homomorphism γ is injective. Define K0 := \H/H0 and identify K0 with the group of characters on H that are equal to 1 on H0. Whenever η ∈ K0, the formula eη : xg 7→ η(cid:16)Xh∈Γ xh(cid:17) for all x ∈ H (Γ), g ∈ G , defines a character on G and hence an automorphism αη ∈ Aut(M ) by the formula αη(uz) = eη(z)uz for all z ∈ G. Since η equals 1 on H0, we get that αη(a) = a for all a ∈ M0. More precisely, (αη)η∈K0 is a continuous action of K0 on M and the fixed point algebra of this action equals M0. Let η, η′ ∈ K0. Applying αη ⊗ αη′ to (8.3), it follows that Ω∗ 1(αη ⊗ αη′ )(Ω1) commutes with uγ(g) ⊗ uγ(g) for all g ∈ G. Since G is icc and γ(G) < G has finite index, we have that {(γ(g)xγ(g)−1, γ(g)yγ(g)−1) g ∈ G} is an infinite set for all (x, y) ∈ (G × G) − {e}. Using 1(αη ⊗ αη′)(Ω1) must be a multiple of 1 and we find Ψ(η, η′) ∈ T Lemma 2.12, it follows that Ω∗ such that (αη ⊗ αη′)(Ω1) = Ψ(η, η′) Ω1 for all (η, η′) ∈ K0 × K0 . (8.4) It follows that Ψ is a continuous character on K0 × K0. Since ζ(Ω1) is a multiple of Ω1, we also get that Ψ(η, η′) = Ψ(η′, η) for all (η, η′) ∈ K0 × K0. Since cK0 = H/H0, we find an x ∈ H such that Ψ(η, η′) = η(x)η′(x) for all (η, η′) ∈ K0 × K0. For every g ∈ Γ, denote by πg : LH → (LH)Γ the embedding of LH as the g-th tensor factor. Write Vx := πe(ux) and put Ω2 := (V ∗ x )Ω1. From (8.4), it follows that Ω2 ∈ M0 ⊗ M0. Denote by xe ∈ H (Γ) the element x ∈ H viewed in position e. Define the injective group homomorphism γ′ : G → G0 : γ′(g) = x−1 e γ(g)xe. It follows from (8.3) that x ⊗ V ∗ Ω∗ 2(uγ ′(g) ⊗ uγ ′(g))Ω2 = ∆(ω(g)ug) for all g ∈ G . (8.5) Since γ(G) has finite index in G and since G is icc, we have that {γ′(g)xγ′(g)−1 g ∈ G} is an infinite set for all x ∈ G0 − {e}. By Lemma 2.12, we get that the representation (Ad(uγ ′(g)))g∈G on L2(M0) ⊖ C1 is weakly mixing. It then follows from (8.5) and [IPV10, Lemma 3.4] that there exist unitaries w, v ∈ M0, a character ω′ : G → T and an injective group homomorphism ρ : G → Λ such that wuγ ′(g)w∗ = ω′(g) vρ(g) for all g ∈ G and Ω2 = (w∗ ⊗ w∗)∆(v) . In combination with (8.5), we get that ω′(g)2 ∆(vρ(g)) = ω′(g) vρ(g) ⊗ ω′(g) vρ(g) = wuγ ′(g)w∗ ⊗ wuγ ′(g)w∗ = ∆(ω(g) vugv∗) 30 for all g ∈ G. So also vugv∗ = ω(g)ω′(g)2 vρ(g). This implies that u∗ γ ′(g) w∗v ug = ω(g)ω′(g) w∗v for all g ∈ G . Lemma 7.1 then provides an element k ∈ G0 such that γ′(g) = kgk−1 for all g ∈ G. It follows that u∗ kw∗v ∈ C1 and that ω′ = ω. So, w is a multiple of vu∗ k and ω(g) vugv∗ = vρ(g) for all g ∈ G . From (8.3), we know that Ω2∆(A0)Ω∗ multiple of vu∗ k, we conclude that ∆(vA0v∗) ⊂ vA0v∗ ⊗ vA0v∗. 2 ⊂ A0 ⊗ A0. Since Ω2 = (w∗ ⊗ w∗)∆(v) and since w is a Proof of Theorem 8.1. The proof consists of three different parts. Writing Λ as a semidirect product Σ ⋊ G We do not explicitly write the isomorphism π : LΛ → (LG0)r, but directly identify LΛ = L(G0)r. We denote by ∆ : LΛ → LΛ ⊗ LΛ the comultiplication given by ∆(vs) = vs ⊗ vs for all s ∈ Λ. Recall from [IPV10, Lemma 7.1] that a von Neumann subalgebra P ⊂ LΛ satisfies ∆(P ) ⊂ P ⊗ P if and only if P = LS for a subgroup S < Λ. As above, we denote H := H (Γ) and H0 := p−1 H (H0). We write G := Γ × Γ and A := LH = (LH)Γ, with its subalgebra A0 := LH0. Finally, M := A ⋊ G = L(H ⋊ G) and M0 := A0 ⋊ G = LG0. For every character ω : G0 → T, we denote by αω the induced automorphism of M0 given by αω(ux) = ω(x)ux for all x ∈ G0. By Lemma 8.12, we get that r = 1 and that we can compose the identification LΛ = LG0 with an inner automorphism of LG0 and an automorphism of the form αω for a character ω : G → T such that after these compositions, we have ∆(A0) ⊂ A0 ⊗ A0 and ug = vρ(g) for all g ∈ G , (8.6) where ρ : G → Λ is an injective group homomorphism. It follows that A0 = LΣ for an abelian subgroup Σ < Λ and that we have written Λ as a semidirect product Λ = Σ ⋊ G, where G acts on Σ by group automorphisms. So from now on, we may assume that Λ = Σ ⋊ G in such a way that LΣ = A0 and vg = ug for all g ∈ G (denoting as above by (vs)s∈Λ the canonical unitaries for LΛ, and by (ua)a∈G0 the canonical unitaries for LG0). Proving that Σ is of the form p−1 H ′ (H ′ 0) Whenever we view Γ as the index set of the infinite tensor product A = (LH)Γ, we denote the elements of Γ by the letters i, j. We denote by g · i the left-right action of g ∈ G on i ∈ Γ. We denote by πi : LH → (LH)Γ the embedding of LH into (LH)Γ as the i-th tensor factor. We denote by (σg)g∈G the generalized Bernoulli action given by σg ◦ πi = πg·i. We finally denote by δ : Γ → G : δ(g) = (g, g) the diagonal embedding. Since Γ is icc, we have that δ(Γ) · i is infinite for all i ∈ Γ − {e}. By Lemma 2.12, the action (σδ(g))g∈Γ on (LH)Γ−{e} is weakly mixing and we have that πe(LH0) = {a ∈ A0 σδ(g)(a) = a for all g ∈ Γ} , πe(LH0) ⊗ πe(LH0) = {a ∈ A0 ⊗ A0 (σδ(g) ⊗ σδ(g))(a) = a for all g ∈ Γ} , πe(LH) ⊗ πe(LH) = {a ∈ A ⊗ A (σδ(g) ⊗ σδ(g))(a) = a for all g ∈ Γ} . (8.7) (8.8) (8.9) 31 1 with corresponding comultiplication ∆1 : LH ′ For the rest of the proof, we only consider the comultiplication ∆ restricted to LΣ. Since vg = ug for all g ∈ G, we have that ∆ ◦ σg = (σg ⊗ σg) ◦ ∆ for all g ∈ G. Using (8.7) and (8.8), it then follows that ∆(πe(LH0)) ⊂ πe(LH0) ⊗ πe(LH0). This means that we find an abelian group H ′ 1, and an identification LH ′ 1 = LH0 such that ∆ ◦ πe = (πe ⊗ πe) ◦ ∆1. Composing with (σg ⊗ σg)g∈G, it follows that ∆ ◦ πi = (πi ⊗ πi) ◦ ∆1 for all i ∈ Γ. So we can view πi as well as an injective group homomorphism of H ′ 1 into Σ. Since the von Neumann algebras πi(LH0), i ∈ Γ, are in tensor product position inside LΣ, it follows that the subgroups πi(H ′ 1) < Σ, i ∈ Γ, are in direct sum position inside Σ. Fix an element x ∈ H. The formula Ωx(g) := πe(ux)πg·e(u∗ x) defines a 1-cocycle for the action (σg)g∈G on A0. Hence g 7→ ∆(Ωx(g)) is a 1-cocycle for the generalized Bernoulli action (σg ⊗ σg)g∈G on (LH)Γ ⊗ (LH)Γ. By Popa's cocycle superrigidity theorem [Po06b, Theorem 1.1], we find a unitary Vx ∈ (LH)Γ ⊗ (LH)Γ such that 1 → LH ′ 1 ⊗ LH ′ ∆(Ωx(g)) = Vx (σg ⊗ σg)(V ∗ x) for all g ∈ G . By construction, Ωx(δ(g)) = 1 for all g ∈ Γ. From (8.9), it then follows that Vx = (πe ⊗ πe)(Ux) for a unitary Ux ∈ LH ⊗ LH. So we get that ∆(πe(ux)πg·e(u∗ x)) = (πe ⊗ πe)(Ux) (πg·e ⊗ πg·e)(U ∗ x ) for all x ∈ H, g ∈ G . Applying σh ⊗ σh for an arbitrary h ∈ G, and combining with the earlier definition of ∆1, we find that ∆((πi ⊗ πj)(ux ⊗ u∗ ∆ ◦ πi = (πi ⊗ πi) ◦ ∆1 x)) = (πi ⊗ πi)(Ux) (πj ⊗ πj)(U ∗ x ) for all i ∈ Γ . for all x ∈ H, i, j ∈ Γ , (8.10) Define H2 := {(x, y) ∈ H × H x + y ∈ H0}. Then H2 is generated by the subgroups H0 × H0 and {(x, −x) x ∈ H}. Since LH ′ 1 = LH0, the von Neumann algebra generated by the elements {(πi ⊗ πj)(ux ⊗ u∗ 1), i ∈ Γ, equals the von Neumann algebra generated by all the (πi ⊗ πj)(LH2), which is the whole of A0 = LΣ. x) i, j ∈ Γ, x ∈ H}, together with the algebras πi(LH ′ So the formulae in (8.10) entirely determine ∆. Also note that for a given x ∈ H, the unitary Ux is uniquely determined up to multiplication by a scalar in T. Finally observe that for x ∈ H0, we have Ux = ∆1(ux), up to multiplication by a scalar in T. In particular, Ux ∈ LH0 ⊗ LH0 whenever x ∈ H0. For all distinct i, j ∈ Γ, denote by πij : LH2 → A0 the embedding into the i'th and j'th coordinate. It follows from (8.10) that we can identify LH2 = LH ′ 2 for some abelian group 2 with the corresponding comultiplication ∆2 : LH ′ H ′ 2 ⊗ LH ′ 2 given by the following formulae that use the tensor leg numbering notation. 2 → LH ′ ∆2(ux ⊗ u∗ ∆2(a ⊗ b) = (∆1(a))13 (∆1(b))24 x) = (Ux)13 (U ∗ x )24 for all x ∈ H , for all a, b ∈ LH ′ 1 . (8.11) By construction, we have ∆ ◦ πij = (πij ⊗ πij) ◦ ∆2. So we can view πij as an injective group 1 as a subgroup of H ′ homomorphism πij : H ′ 2 and that under this identification πij(a, b) = πi(a) + πj(b) for all (a, b) ∈ H ′ 2 → Σ. Note that we can naturally view H ′ 1 × H ′ 1 × H ′ 1. We denote by K := bH the group of characters on H and by K0 < K the closed subgroup of characters that are identically 1 on H0. We identify K0 = \H/H0. Whenever ω ∈ K, we denote by αω ∈ Aut(LH) the induced automorphism given αω(ux) = ω(x)ux for all x ∈ H. 32 Applying αω in the i-th coordinate yields the automorphism αi αω in all coordinates yields the automorphism αΓ that αΓ ω ∈ K0. ω ∈ Aut((LH)Γ), while applying ω ∈ Aut((LH)Γ). By construction, we have ω(a) = a for all ω ◦ πi = πi ◦ αω. A given a ∈ A = (LH)Γ belongs to A0 if and only if αΓ Fix x ∈ H. Since ∆(A0) ⊂ A0 ⊗ A0, the left hand side of the formulae in (8.10) is invariant under αΓ ω ⊗ id for all ω ∈ K0. Since Ux is uniquely determined up to a scalar, it follows that (αω ⊗ id)(Ux) is a multiple of Ux for every ω ∈ K0. So we find an element γ(x) ∈ H/H0 such that (αω ⊗ id)(Ux) = ω(γ(x)) Ux for all ω ∈ K0 . When x ∈ H0, we have that Ux ∈ LH0 ⊗ LH0 and hence γ(x) = 0. well-defined group homomorphism from H/H0 to H/H0. The formulae in (8.10) entirely determine ∆ so that it follows that (αi ω◦γ for all i ∈ Γ and all ω ∈ K0. Using Lemma 7.4, we conclude that γ = id and that every automorphism αi ω is induced by a character of Σ. It follows that there are group homomorphisms ψi : Σ → H/H0 such that ω ⊗id)◦∆ = ∆◦αi It follows that γ is a αi ω(vs) = ω(ψi(s)) vs for all s ∈ Σ, i ∈ Γ, ω ∈ K0 . ω ◦ πij = πij ◦ (αω ⊗ id), we have that ψi ◦ πij = ψ. Since (αω ⊗ id)(x) = (id ⊗ αω)(x) ω ◦ πij = πij ◦ (αω ⊗ id). Hence ψj ◦ πij = −ψ. We further have that A similar reasoning, using (8.11) instead of (8.10), provides a homomorphism ψ : H ′ such that (αω ⊗ id)(vs) = ω(ψ(s)) vs for all s ∈ H ′ Since αi for all x ∈ LH ′ ψk ◦ πij = 0 if k 6∈ {i, j}. We already observed above that the subgroups πi(H ′ Denote by Σ1 < Σ the subgroup generated by the πi(H ′ that LΣ1 = (LH0)Γ. It follows that 1) < Σ, i ∈ Γ, are in a direct sum position. 1 = LH0, we have 1), i ∈ Γ. Since LH ′ 2, we have αj 2 → H/H0 2, ω ∈ K0. LΣ1 = {x ∈ A0 αi ω(x) = x for all i ∈ Γ, ω ∈ K0} and hence Σ1 =\i∈I Ker ψi . Every permutation β ∈ Perm Γ defines an automorphism γβ of (LH)Γ by permuting the tensor It follows from (8.10) that (γβ ⊗ γβ) ◦ ∆ = ∆ ◦ γβ, so that γβ induces a group factors. automorphism of Σ. By construction, we have γβ ◦ πi = πβ(i) and γβ ◦ πij = πβ(i),β(j). It is now easy to check that all assumptions of Lemma 8.13 are satisfied. We conclude from 0 < H ′ and a G-equivariant Lemma 8.13 that there exists an abelian group H ′ with subgroup H ′ group isomorphism p−1 H ′ (H ′ 0) → Σ. Proving that the isomorphism π is of the required form We put H′ := H ′(Γ) and H′ LH0, with the above identification of LΣ and LH′ Λ = H′ 0 ⋊ G and where the isomorphism 0 := p−1 H ′ (H ′ 0). Precomposing the original identification of LΣ and 0, we have brought us to the point where π : L(H′ 0 ⋊ G) → L(H0 ⋊ G) 0) = LH0 and π(ug) = ug for all g ∈ G. satisfies π(LH′ Denote by ϕ : LH′ 0. Note that ϕ is a G-equivariant ∗-isomorphism. To conclude the proof of Theorem 8.1, it remains to prove that ϕ must be 0 → LH0 the restriction of π to LH′ 33 of the following special form: there exist a group isomorphism γ : H ′/H ′ 0 → H/H0, a G- invariant character µ : H0 → T and a trace preserving ∗-isomorphism ϕ0 : LH ′ → LH such that ϕ0 ◦ αω◦γ = αω ◦ ϕ0 for all ω ∈ \H/H0 and such that ϕ = αµ ◦ ϕΓ 0 . Here the ∗-isomorphism ϕΓ 0 : (LH ′)Γ → (LH)Γ is defined as the infinite tensor product of copies of ϕ0. Denote K = bH, K ′ = cH ′, K0 = \H/H0 and K ′ embed K0 as a subgroup of K Γ diagonally. We similarly consider K ′ 0 = \H ′/H ′ 0. Consider the compact group K Γ and 0 < (K ′)Γ. We identify LH′ 0 = L∞(cid:16) (K ′)Γ 0 (cid:17) and LH0 = L∞(cid:16) K Γ K0(cid:17) . K ′ We can then view ϕ = θ∗ where θ is a probability measure preserving (pmp), G-equivariant isomorphism θ : (K ′)Γ K ′ 0 → K Γ K0 . y (K ′)Γ and G × K0 y K Γ. By Popa's cocycle super- Consider the natural actions G × K ′ 0 rigidity theorem [Po06b, Theorem 1.1] and [PV06, Lemma 5.2], there exist a pmp isomorphism eθ : (K ′)Γ → K Γ, a group homomorphism β : G → K0 : g 7→ βg and a continuous group isomorphism bγ : K ′ 0 → K ′ such that (8.12) eθ((g, k) · ω) = (g, βgbγ(k)) ·eθ(ω) 0 and a.e. ω ∈ (K ′)Γ. and eθ(ω) + K0 = θ(ω + K ′ 0) , for all (g, k) ∈ G × K ′ Fix x ∈ H and denote Fx : K Γ → T : Fx(ω) = ωe(x). As before, denote by δ : Γ → G : δ(g) = (g, g) the diagonal embedding. One checks that (Fx ◦eθ)(δ(g) · ω) = βg(x) (Fx ◦eθ)(ω) for all g ∈ Γ and a.e. ω ∈ (K ′)Γ . Since Γ is icc, it follows from Lemma 2.12 that the action of δ(Γ) on (K ′)Γ−{e} is weakly mixing, so that the function ω 7→ (Fx ◦ eθ)(ω) only depends on the coordinate ωe. Since this holds for all x ∈ H, we find a pmp isomorphism θ0 : K ′ → K such that (eθ(ω))e = θ0(ωe) for a.e. ω. By construction, we have θ0(k + ω) = bγ(k) + θ0(ω) for all k ∈ K ′ 0 and a.e. ω ∈ K ′. Writing ϕ0 := (θ0)∗, we obtain the trace preserving ∗-isomorphism ϕ0 : LH ′ → LH satisfying ϕ0 ◦ αω◦γ = αω ◦ ϕ0 for all ω ∈ \H/H0. Evaluating (8.12) in the coordinate e, we find that βδ(g) = 0 for all g ∈ Γ, so that β(g,h) = ρg −ρh for a group homomorphism ρ : Γ → K0 : g 7→ ρg. We also find that eθ(ω)g = θ0(ωg) + ρg for all g ∈ Γ and a.e. ω ∈ (K ′)Γ. Define µ ∈ K Γ/K0 as µ := (ρg)g∈Γ + K0. Then µ is a G- invariant element of K Γ/K0, i.e. a G-invariant character on H0. By construction, we have that ϕ = αµ ◦ ϕΓ 0 . A combinatorial lemma Whenever I is a countable set and H is a countable abelian group with subgroup H0 < H, we consider the direct sum H (I), the group homomorphism pH : H (I) → H : pH(x) =Xg∈I xg and the subgroup p−1 group automorphisms that leave the subgroup p−1 H (H0) of H (I). The group Perm I of all permutations of I acts on H (I) by H (H0) globally invariant. 34 For every i ∈ I, we have a natural embedding µi : H0 → p−1 H (H0) of H0 into the i-th coordinate. Writing (H × H)H0 := {(x, y) ∈ H × H x + y ∈ H0}, we also have natural embeddings µij : (H × H)H0 → p−1 H (H0) into the i-th and j-th coordinate, whenever i and j are distinct elements of I. The subgroups µij((H × H)H0) generate p−1 The following elementary lemma abstractly characterizes this whole setup. The lemma is actually much more awkward to state than to prove. H (H0). Lemma 8.13. Let Σ be a countable abelian group and I a countably infinite set. Assume that we are given the following data: • countable abelian groups H1 and H2 such that H1 × H1 < H2, • for all i ∈ I, an injective homomorphism πi : H1 → Σ, • for all distinct i, j ∈ I, an injective homomorphism πij : H2 → Σ, • an abelian group L and, for all i ∈ I, a group homomorphism ψi : Σ → L, • a group homomorphism ψ : H2 → L, • an action of the group of all permutations β ∈ Perm I by group automorphisms γβ of Σ, such that the following conditions hold: • the subgroups πij(H2) generate Σ, • the subgroups πi(H1) are in a direct sum position inside Σ and generate a subgroup of Σ denoted by Σ1, • we have πij(a, b) = πi(a) + πj(b) for all (a, b) ∈ H1 × H1 ⊂ H2, • we have ψi ◦ πij = ψ = −ψj ◦ πij, • we have ψk ◦ πij = 0 if k 6∈ {i, j}, • we have Σ1 =Ti∈I Ker ψi, • for every β ∈ Perm I, we have γβ ◦ πi = πβ(i) and γβ ◦ πij = πβ(i),β(j). Then there exist a countable abelian group H with subgroup H0 < H and group isomorphisms δ1 : H0 → H1 , δ2 : (H × H)H0 → H2 and δ : p−1 H (H0) → Σ such that, using the notations µi and µij introduced before the lemma, we have • δ conjugates the actions of Perm I, • δ ◦ µi = πi ◦ δ1, • δ ◦ µij = πij ◦ δ2. 35 Proof. We may assume that I = N. Since the subgroups πi(H1) < Σ are in a direct sum position, we can assemble the πi into an isomorphism π : H (N) 1 → Σ1. Note that π conjugates the natural actions of Perm N. 1 Fix x ∈ H2. Observe that y := π12(x) + π23(x) + π31(x) belongs to the kernel of all ψi, i ∈ N. Hence, y = π(z) for some element z ∈ H (N) . It follows that z is invariant under cyclic permutations of (1, 2, 3). It also follows that z is invariant under all permutations that fix 1, 2 and 3. Since there are only finitely many k ∈ N with zk 6= 0, we conclude that y must be of the form y = π1(ρ(x)) + π2(ρ(x)) + π3(ρ(x)), where ρ : H2 → H1 is a group homomorphism. Also note that ρ(a, b) = a + b for all (a, b) ∈ H1 × H1 ⊂ H2. We define H := Ker ρ. We define the subgroup H0 < H given by H0 := {(a, −a) a ∈ H1}. We denote δ1 : H0 → H1 : δ1(a, −a) := a. By construction, we have that π12(x) + π23(x) + π31(x) = 0 for all x ∈ H. Applying γβ for an arbitrary permutation β of N, it follows that πij(x) + πjk(x) + πki(x) = 0 (8.13) for all x ∈ H and all distinct i, j, k ∈ N. Fix x ∈ H2. Observe that y := π12(x) + π21(x) belongs to the kernel of all ψi, i ∈ N. We also have that γβ(y) = y when β is the permutation of N that flips 1 and 2, as well as when β is a permutation that fixes 1 and 2. Reasoning as above, it follows that π12(x) + π21(x) = −π1(η(x)) − π2(η(x)), where η : H2 → H1 is a group homomorphism. We only introduced the minus sign to make the following computation easier. Applying γβ for an arbitrary permutation β of N, we get that πji(x) = −πij(x) + πi(η(x)) + πj(η(x)) for all x ∈ H2 and all distinct i, j ∈ N. We prove that η(x) = 0 for all x ∈ H. Fix x ∈ H and consider the element y := π12(x) + π23(x) + π34(x) + π41(x) . A first computation, using (8.13), yields y = −π31(x) + π34(x) + π41(x) = π1(η(x)) + π3(η(x)) + π13(x) + π34(x) + π41(x) = π1(η(x)) + π3(η(x)) . An analogous second computation gives y = π12(x) − π42(x) + π41(x) = π2(η(x)) + π4(η(x)) + π12(x) + π24(x) + π41(x) = π2(η(x)) + π4(η(x)) . Since the groups πi(H1) are in a direct sum position inside Σ, both computations together imply that η(x) = 0 for all x ∈ H. It follows that πij(x) = −πji(x) for all x ∈ H and all distinct i, j ∈ N. In combination with (8.13), we get that πij(x) + πjk(x) = πik(x) (8.14) for all x ∈ H and all distinct i, j, k ∈ N. We claim that the homomorphism δ2 : (H × H)H0 → H2 : δ2(x, y) = x + (0, δ1(x + y)) 36 is an isomorphism of groups satisfying δ2(x, y) = (δ1(x), δ1(y)) for all (x, y) ∈ H0 × H0. This last formula is immediate. It already implies that the image of δ2 contains both H and H1 ×H1. Since for every x ∈ H2, we have that x − (0, ρ(x)) ∈ H, the surjectivity of δ2 follows. Since ρ(δ2(x, y)) = δ1(x + y), the injectivity of δ2 follows as well. Using (8.14), it follows that the formula δ : p−1 H (H0) → Σ : δ(x) = πn+1(δ1(pH(x))) + nXi=1 πi,n+1(xi) whenever xk = 0 for all k > n is independent of the choice of n and hence a well-defined homomorphism satisfying δ ◦ µij = πij ◦ δ2 and δ ◦ µi = πi ◦ δ1. It immediately follows that δ conjugates the respective actions of Perm N and that δ is surjective. To prove the injectivity of δ, we first claim that H0 = H ∩ Ker ψ. The inclusion ⊂ is obvious. Conversely, assume that y ∈ H and ψ(y) = 0. Put z = π12(y). We get that z ∈ Ker ψk for all k ∈ N. So z ∈ Σ1. Since γβ(z) = z for every permutation β that fixes 1 and 2, we find that y ∈ H1 × H1. Since y ∈ H, we obtain the claim that y ∈ H0. If now δ(x) = 0, we get that ψ(xi) = ψi(δ(x)) = 0 for all i ∈ N. So x belongs to H (I) 0 . Since δ ◦ µi = πi ◦ δ1, the restriction of δ to H (I) is injective. 0 Proofs of Theorem B and Remark C Proof of Theorem B. A hyperbolic group Γ has only finitely many conjugacy classes of finite subgroups (see e.g. [Br99]). By Selberg's lemma [Se60], a finitely generated linear group Γ (over a field of characteristic zero) has a finite index subgroup that is torsion-free. In both cases, Γ admits a bound on the possible orders of its finite subgroups. By the work of [CH88, Sk88, Oz03, Oz07] (see [PV12, Lemma 2.4] for a more detailed explanation), we also have in both cases that Γ is weakly amenable and that Γ belongs to class S. So every group Γ that appears in Theorem B satisfies the conditions of Theorem 8.1. We will apply Theorem 8.1. The conclusion of Theorem 8.1 describes the given ∗-isomorphism π : LΛ → (LG0)r as a composition of an inner automorphism, "group like" isomorphisms implemented by group isomorphisms and characters, and the ∗-isomorphism πθ that need not be group like in general. We now prove that in the situation of Theorem B, also πθ is group like. 1. Assume that H = Z/nZ with n ∈ {2, 3} and put G = H (Γ) ⋊ (Γ × Γ). We apply Theorem 8.1 with H0 = H. This provides an abelian group H ′ with H ′ = H. So, H ′ ∼= H and we may assume that H ′ = H. It only remains to prove that the automorphism πθ : LG → LG is group like. But since LH has dimension 2 or 3, it is not hard to check that every automorphism δ : H → H. Then πθ is group like as well. 2. We apply Theorem 8.1 with H0 = {0}. Since H ′ ∼= H, we may assume that H ′ = H. Then θ : LH → LH is of the form θ = αω ◦ πδ for some character ω ∈ bH and group automorphism θ : bH → bH is a pmp isomorphism satisfying θ(k + ω) = k + θ(ω) for a.e. k, ω ∈ bH. So we find a fixed ω0 ∈ bH such that θ(ω) = ω + ω0 for a.e. ω ∈ bH. But then πθ is the identity map. H ({0}) and G0 = H0 ⋊ G. Put K = bH. Since G has no nontrivial characters, we only need Proof of Remark C. Assume that Γ has no nontrivial characters. Put G = Γ × Γ, H0 = p−1 to prove that H0 has no nontrivial G-invariant characters. This means that we have to prove that the action of G on the compact space K Γ/K only has 0 as a fixed point. One checks 37 that the G-fixed points in K Γ/K are precisely the points (αg)g∈Γ + K where α : Γ → K is a homomorphism. Since Γ has no nontrivial characters and K is abelian, such a homomorphism is constantly equal to 0. References [Br99] [BO08] N. Brady, Finite subgroups of hyperbolic groups. Int. J. Algebra Comput. 10 (2000), 399-406. N.P. Brown and N. Ozawa, C∗-algebras and finite-dimensional approximations. Graduate Studies in Mathematics 88. American Mathematical Society, Providence, 2008. A. Connes, Classification of injective factors. Ann. of Math. (2) 104 (1976), 73-115. [Co76] [Co80a] A. Connes, A factor of type II1 with countable fundamental group. J. Operator Theory 4 (1980), 151-153. [Co80b] A. Connes, Classification des facteurs. In Operator algebras and applications, Part 2 (Kingston, 1980), Proc. Sympos. Pure Math. 38, Amer. Math. Soc., Providence, 1982, pp. 43-109. [CH88] M. Cowling and U. Haagerup, Completely bounded multipliers of the Fourier algebra of a [Ef73] [Io06] [Io10] [Io12a] [Io12b] simple Lie group of real rank one. Invent. Math. 96 (1989), 507-549. E.G. Effros, Property Γ and inner amenability. Proc. Amer. Math. Soc. 47 (1975), 483-486. A. Ioana, Rigidity results for wreath product II1 factors. J. Funct. Anal. 252 (2007), 763-791. A. Ioana, W∗-superrigidity for Bernoulli actions of property (T) groups. J. Amer. Math. Soc. 24 (2011), 1175-1226. A. Ioana, Classiffication and rigidity for von Neumann algebras. In Proceedings of the 6th Eu- ropean Congress of Mathematics (Krakow, 2012), European Mathematical Society Publishing House, to appear. A. Ioana, Cartan subalgebras of amalgamated free product II1 factors. Preprint. arXiv:1207.0054 [IPP05] A. Ioana, J. Peterson and S. Popa, Amalgamated free products of weakly rigid factors and calculation of their symmetry groups. Acta Math. 200 (2008), 85-153. [IPV10] A. Ioana, S. Popa and S. Vaes, A class of superrigid group von Neumann algebras. Ann. Math., to appear. arXiv:1007.1412 [McD69] D. McDuff, Central sequences and the hyperfinite factor. Proc. London Math. Soc. (3) 21 (1970), 443-461. [MP03] N. Monod and S. Popa, On co-amenability for groups and von Neumann algebras. C. R. Math. Acad. Sci. Soc. R. Can. 25 (2003), 82-87. [MvN43] F.J. Murray and J. von Neumann, Rings of operators IV, Ann. Math. 44 (1943), 716-808. [Oz03] [Oz07] [OP07] N. Ozawa, Solid von Neumann algebras. Acta Math. 192 (2004), 111-117. N. Ozawa, Weak amenability of hyperbolic groups. Groups Geom. Dyn. 2 (2008), 271-280. N. Ozawa and S. Popa, On a class of II1 factors with at most one Cartan subalgebra. Ann. Math. 172 (2010), 713-749. S. Popa, On a class of type II1 factors with Betti numbers invariants. Ann. of Math. 163 (2006), 809-899. S. Popa, Strong rigidity of II1 factors arising from malleable actions of w-rigid groups, I. Invent. Math. 165 (2006), 369-408. S. Popa, Strong rigidity of II1 factors arising from malleable actions of w-rigid groups, II. Invent. Math. 165 (2006), 409-452. S. Popa, Deformation and rigidity for group actions and von Neumann algebras. In Pro- ceedings of the International Congress of Mathematicians (Madrid, 2006), Vol. I, European Mathematical Society Publishing House, 2007, p. 445-477. S. Popa, On the superrigidity of malleable actions with spectral gap. J. Amer. Math. Soc. 21 (2008), 981-1000. [Po01] [Po03] [Po04] [Po06a] [Po06b] 38 [PV06] [PV11] [PV12] [Sc63] [Se60] [Si10] [Sk88] [Va07] [Va09] [Va10a] [Va10b] S. Popa and S. Vaes, Strong rigidity of generalized Bernoulli actions and computations of their symmetry groups. Adv. Math. 217 (2008), 833-872. S. Popa and S. Vaes, Unique Cartan decomposition for II1 factors arising from arbitrary actions of free groups. Preprint. arXiv:1111.6951 S. Popa and S. Vaes, Unique Cartan decomposition for II1 factors arising from arbitrary actions of hyperbolic groups. Preprint. arXiv:1201.2824 J. Schwartz, Two finite, non-hyperfinite, non-isomorphic factors. Comm. Pure Appl. Math. 16 (1963), 19-26. A. Selberg, On discontinuous groups in higher-dimensional symmetric spaces. In Contribu- tions to function theory, Tata Institute of Fundamental Research, Bombay, pp. 147-164. T. Sinclair, Strong solidity of group factors from lattices in SO(n, 1) and SU(n, 1). J. Funct. Anal. 260 (2011), 3209-3221. G. Skandalis, Une notion de nucl´earit´e en K-th´eorie (d'apr`es J. Cuntz). K-Theory 1 (1988), 549-573. S. Vaes, Explicit computations of all finite index bimodules for a family of II1 factors. Ann. Sci. ´Ecole Norm. Sup. 41 (2008), 743-788. S. Vaes, An inner amenable group whose von Neumann algebra does not have property Gamma. Acta Math. 208 (2012), 389-394. S. Vaes, Rigidity for von Neumann algebras and their invariants. In Proceedings of the Inter- national Congress of Mathematicians (Hyderabad, 2010), Vol. III, Hindustan Book Agency, 2010, p. 1624-1650. S. Vaes, One-cohomology and the uniqueness of the group measure space decomposition of a II1 factor. Math. Ann., to appear. arXiv:1012.5377 39
1006.4044
1
1006
2010-06-21T11:55:31
Property (T) and exotic quantum group norms
[ "math.OA" ]
Utilizing the notion of property (T) we construct new examples of quantum group norms on the polynomial algebra of a compact quantum group, and provide criteria ensuring that these are not equal to neither the minimal nor the maximal norm. Along the way we generalize several classical operator algebraic characterizations of property (T) to the quantum group setting which unify recent approaches to property (T) for quantum groups with previous ones. The techniques developed furthermore provide tools to answer two open problems; firstly a question by B\'edos, Murphy and Tuset about automatic continuity of the comultiplication and secondly a problem left open by Woronowicz regarding the structure of elements whose coproduct is a finite sum of simple tensors.
math.OA
math
PROPERTY (T) AND EXOTIC QUANTUM GROUP NORMS DAVID KYED AND PIOTR M. SO LTAN Dedicated to Ryszard Nest on the occasion of his 60th birthday Abstract. Utilizing the notion of property (T) we construct new examples of quantum group norms on the polynomial algebra of a compact quantum group, and provide criteria ensuring that these are not equal to neither the minimal nor the maximal norm. Along the way we generalize several classical operator algebraic characterizations of property (T) to the quantum group setting which unify recent approaches to property (T) for quantum groups with previous ones. The techniques developed furthermore provide tools to answer two open problems; firstly a question by B´edos, Murphy and Tuset about automatic continuity of the comultiplication and secondly a problem left open by Woronowicz regarding the structure of elements whose coproduct is a finite sum of simple tensors. 0 1 0 2 n u J 1 2 ] . A O h t a m [ 1 v 4 4 0 4 . 6 0 0 1 : v i X r a 1. Introduction This paper is devoted to the theory of compact and discrete quantum groups. Both of these classes of quantum groups have been studied in detail by many authors and suffer from no shortage of exciting examples ([29, 21, 20, 27, 28]). It is known that a given compact quantum group G can be described by more than one C∗-algebra (see e.g. [30, 4]); the most useful choices being the "maximal" and the "minimal" (also called reduced ) completions of the algebra Pol(G) of polynomial functions on G. It often happens that the canonical quotient map from the maximal completion C(Gmax) to the minimal one C(Gmin) is an isomorphism (in other words G is co-amenable, [4]). However, many interesting situations can arise when G is described by a C∗-algebra sitting "in between" the maximal and minimal one (cf. [23]), but unfortunately there are not many examples of such compact quantum groups (apart from obvious direct product constructions). Unlike compact quantum groups, the discrete quantum groups (i.e. the duals of compact quan- tum groups) are all co-amenable -- there is just one C∗-algebra for each discrete quantum group. However, within this class of quantum groups one can find very interesting examples. In particular, there are discrete quantum groups with property (T) which we will study in this paper. We will use property (T) to construct special C∗-norms on the algebras of polynomials on (compact) dual quantum groups of property (T) discrete quantum groups. The completions of these polynomial algebras will be "exotic" in the sense that they will sit in between the maximal and minimal com- pletions. The canonical bijection between corepresentations of a discrete quantum group bG and ∗-representations of the C∗-algebra C(Gmax) will play a very important part in our investigation. Let us briefly discuss the content of the paper. In Subsections 1.1 and 1.2 we introduce the notation and list certain preliminary results from the theory of compact and discrete quantum groups. Section 2 provides necessary definitions and facts from the theory of corepresentations of quantum groups; we describe the standard operations of forming tensor products and contragre- dient corepresentations, emphasizing the link with representations of the dual object. The regular corepresentation of a discrete quantum group is introduced in Subsection 2.4 and in Subsection 2.5 we prove a quantum group version of a classical theorem from representation theory of locally compact groups. This theorem will be useful in the following sections. Date: June 7, 2018. 2010 Mathematics Subject Classification. Primary: 20G42, 46L89, secondary: 22D25. Key words and phrases. compact quantum group, discrete quantum group, property (T). P.S. partially supported by Polish government grant no. N201 1770 33, European Union grant PIRSES-GA- 2008-230836 and Polish government matching grant no. 1261/7.PR UE/2009/7. D.K. supported by The Danish Council for Independent Research Natural Sciences. 1 2 DAVID KYED AND PIOTR M. SO LTAN In Section 3 property (T) for discrete quantum groups is recalled and several known facts about quantum groups with property (T) are listed. Then in Section 4 the classical characterization of property (T) in terms of isolated points in the space of irreducible representations is extended to the quantum group setting. This result provides a direct link between property (T) of Fima ([14]) and earlier definitions in [5, 19]. As a consequence, in Section 5 we are able to show that a discrete quantum group bG has property (T) if and only if the C∗-algebra C(Gmax) has property (T) in the sense of Bekka ([6]). Finally in Section 6 we extend the characterization of property (T) by existence of a minimal projection in the full group C∗-algebra ([25]) to the setting of discrete quantum groups. The notion of a quantum group norm on the algebra of polynomials on a compact quantum group is defined in Section 7, where we also recall some basic facts about such norms. Then in Section 8 we give the first construction of a quantum group norm making the counit continu- ous. We call this procedure "adjoining the neutral element to a compact quantum group." This construction provides examples of quantum group norms which differ from the reduced one as well as the maximal one in the absence of both amenability and property (T). More complicated examples are collected in Section 9, where, starting from a property (T) discrete quantum group bG, we construct a certain quantum group norm k ·kΠ on Pol(G). The completion of Pol(G) in this norm provides many examples of interesting (exotic) compact quantum groups. These examples lead us to a (negative) answer to a question of B´edos, Murphy and Tuset whether any C∗-norm on Pol(G) arising from a representation weakly containing the regular one is necessarily a quantum group norm (cf. [4] and Section 9). Along the way we also give an example which shows that a very useful theorem of S.L. Woronowicz about compact quantum groups with faithful Haar mea- sure ([32, Theorem 2.6(2)]) cannot be generalized to all compact quantum groups. Several of our examples are co-commutative and we use some well known results from harmonic analysis (for which we refer e.g. to [7] and geometric group theory ([12]) to analyze them. The paper uses the standard language of quantum group theory on operator algebra level ([31, 32, 16]). In particular for a C∗-algebra A the symbol M(A) will denote the multiplier algebra of A. All Hilbert spaces we will consider will be separable and the inner products will be linear in the second variable. Similarly all C∗-algebras, except multiplier algebras, will be assumed to be separable and the tensor product of C∗-algebras will always be the spatial one. The term "representation" will always mean a ∗-representation. 1.1. Notation. We shall adopt the convention of e.g. [21, 14, 18] and always look at discrete quantum groups as duals of compact quantum groups. Thus any discrete quantum group will be denoted by bG. The C∗-algebra of "continuous functions on bG vanishing at infinity" will be denoted by c0(bG) and its comultiplication by b∆. Thus The compact quantum group G dual to bG can be described via many different objects. The polynomial algebra of G, i.e. the Hopf ∗-algebra spanned by matrix elements of finite dimensional corepresentations of G, will be denoted by Pol(G). The universal enveloping C∗-algebra of Pol(G), i.e. its completion with respect to the maximal C∗-norm will be denoted by C(Gmax). The Hilbert space obtained via GNS construction from the Haar measure h of G will be denoted by L2(G). The completion of Pol(G) in the norm coming from representing Pol(G) on L2(G) will be denoted by C(Gmin). Each of the algebras Pol(G), C(Gmax) and C(Gmin) has its own comultiplication, but we will use the same symbol ∆ for all of them. bG =(cid:0)c0(bG),b∆(cid:1). The possible other completions of Pol(G) will be denoted by C(G) or C(G(cid:3)), where in the space reserved by "(cid:3)" a symbol indicating the nature of the completion will be placed. For example, if we choose a faithful representation π of Pol(G) on some Hilbert space then the resulting C∗- In case the C∗-norm used to complete Pol(G) completion of Pol(G) will be written as C(Gπ). is a quantum group norm (see Section 7) the C∗-algebra C(G(cid:3)) will carry a comultiplication extending that of Pol(G) and we will continue to denote it by the symbol ∆. The only exception to this will come in parts of Section 9, where the distinction between comultiplications on different PROPERTY (T) AND EXOTIC QUANTUM GROUP NORMS 3 completions of Pol(G) will be necessary. The von Neumann algebra obtained as the bicommutant of C(Gmin) in B(cid:0)L2(G)(cid:1) will be denoted by L∞(G). The set of equivalence classes of irreducible corepresentations of G will be denoted by Irr(G). For α ∈ Irr(G) we will choose (and fix throughout the paper) a corepresentation uα in the class α. The dimension of uα will be denoted by nα. Thus uα is a unitary element of Mnα(C) ⊗ Pol(G). As Pol(G) naturally embeds into C(Gmax) and C(Gmin) (or any C(G(cid:3)) for that matter), we can regard uα as element of Mnα(C) ⊗ C(Gmax) or Mnα(C) ⊗ C(Gmin) etc. Let us also recall that a discrete quantum group bG is unimodular if its left and right Haar measures coincide (cf. [21, Section 3]). This is equivalent to many different conditions (cf. [32, Theorem 2.5]). The one we will use is that of G being a compact quantum group of Kac type which manifests itself in the fact that the antipode of G is a ∗-anti-automorphism. 1.2. Some preliminary results. Recall from [21, Section 3], [32, Section 4] that Mnα(C) c0(bG) = Mα∈Irr(G) L2(G) = Mα∈Irr(G) H α and it naturally acts on L2(G) which is the GNS Hilbert space for the Haar measure h of G. We have the decomposition with H α the subspace of L2(G) spanned by matrix elements of uα. The set i,j α ∈ Irr(G), i, j ∈ {1, . . . , nα}(cid:9) (cid:8)uα (1.1) duced. It is the element is not an orthonormal basis of L2(G) in general (cf. the Peter-Weyl-Woronowicz relations in [32, Section 7]), but if necessary the representatives (uα) of classes in Irr(G) can be chosen so that it is an orthogonal system ([11, Proposition 2.1]). Also let us note that if G is of Kac type, then the i,j α ∈ Irr(G), i, j = 1, . . . , nα(cid:9) is an orthonormal basis of L2(G). system (cid:8)√nαuα In [21, Section 2] the universal bicharacter describing the duality between G and bG was intro- of M(cid:0)c0(bG) ⊗ C(Gmax)(cid:1). It is of great importance and we will use it throughout the paper. The action of c0(bG) on L2(G) is described in detail e.g. in [32]. Interpreting [32, Formula 5.3] in accordance with our notation we obtain for a ∈ Pol(G) and ξ ∈ C(Gmax)∗ the formula where we view Pol(G) as a dense subspace of L2(G). Let us fix α and i, j ∈ {1, . . . , nα} and take for ξ the functional satisfying (cid:0)(id ⊗ ξ)w(cid:1)a = (id ⊗ ξ)∆(a), w = Mα∈Irr(G) (1.2) uα ξ(uβ k,l) = δα,βδi,kδj,l (for existence of such a ξ cf. [20, Section 1]) and put a = uα (id ⊗ ξ)w = eα i,j ∈ Mnα(C) ⊂ Mα∈Irr(G) and r,l. Then we have Mnα(C) = c0(bG) r,l) = (id ⊗ ξ) (id ⊗ ξ)∆(uα i,j acts on a basic element uα nαXk=1 Thus eα k,l = r,k ⊗ uα uα r,l of H α as eα i,j : uα r,l 7−→ δj,luα r,i. nαXk=1 ξ(uα k,l)uα r,k = nαXk=1 δi,kδj,luα r,k = δj,luα r,i. (1.3) 4 DAVID KYED AND PIOTR M. SO LTAN The lesson from this is that if H is a Hilbert space and m ∈ Mnα(C)⊗ B(H) ⊂ M(cid:0)c0(bG)⊗ K (H)(cid:1) is a matrix of operators m1,1 ... m =  . . . m1,nα . . . . . . mnα,nα ...   mnα,1 then for r, l = 1, . . . , nα and any ξ ∈ H we have r,l ⊗ mi,jξ = eα i,juα m(uα r,l ⊗ ξ) = r,i ⊗ mi,jξ = In particular, if m(η ⊗ ξ) = η ⊗ ξ for any η ∈ H α then taking η = uα nαXi,j=1 nαXi,j=1 δj,luα nαXi=1 uα r,i ⊗ mi,lξ r,l yields uα r,l ⊗ ξ = uα r,i ⊗ mi,lξ nαXi=1 so that mi,lξ = δi,lξ. 2. Corepresentations of discrete quantum groups In this section we collect the standard facts about corepresentations of discrete quantum groups. Most of what we write here applies to all locally compact quantum groups and possibly more general quantum groups (cf. [24]), but in what follows we will stick with discrete quantum groups. Let bG be a discrete quantum group. A unitary corepresentation of bG on a Hilbert space HU is a unitary U ∈ M(cid:0)c0(bG) ⊗ K (H)(cid:1) such that (One usually expects the right hand side of the above equation to read U13U23, but this is really not so much different because U ∗ satisfies such an equation.) Let w be the universal bicharacter describing the duality between bG and G (defined by (1.2)). Then it is known ([24, Section 5.1]) that any corepresentation U ∈ M(cid:0)c0(bG) ⊗ K (HU )(cid:1) is of the form U = (id ⊗ πU )w, where πU is a (uniquely determined) representation of C(Gmax) on the Hilbert space HU . We will not consider non-unitary corepresentations. Let U be a corepresentation of bG. Then U = (id ⊗ π)w for some representation π of C(Gmax). Now for any α ∈ Irr(G) we define This is sometimes called the α-component of U , but note that U α it is nothing like a sub- corepresentation. U α = (id ⊗ π)uα. Let us now describe some operations on corepresentations. (b∆ ⊗ id)U = U23U13. 2.1. Tensor product. Take two corepresentations U = (id ⊗ πU )w ∈ M(cid:0)c0(bG) ⊗ K (HU )(cid:1), V = (id ⊗ πV )w ∈ M(cid:0)c0(bG) ⊗ K (HV )(cid:1) ⊤ V of U and V is defined as U12V13 ∈ M(cid:0)c0(bG) ⊗ K (HU ⊗ HV )(cid:1). ⊤ V =(cid:0)id ⊗ [(πU ⊗ πV )◦∆](cid:1)w. U  of bG. The tensor product U  Another way to view the tensor product is Indeed, (id ⊗ ∆)w = w12w13. PROPERTY (T) AND EXOTIC QUANTUM GROUP NORMS 5 2.2. Contragredient representation. If H is a Hilbert space and H the complex conjugate Hilbert space then we have the anti-isomorphism given by ⊤ : B(H) ∋ m 7−→ ⊤(m) = m⊤ ∈ B(H) m⊤x = m∗x. tation V c of V is defined as V Let V = (id⊗ πV )w ∈ M(cid:0)c0(bG)⊗ K (HV )(cid:1) be a corepresentation. The contragredient represen- bR⊗⊤ = (bR ⊗ ⊤)V ∈ M(cid:0)c0(bG) ⊗ K (HV )(cid:1) (cf. [24, Section 3]), where bR is the unitary antipode ([33, Theorem 1.5(4)]). Again there is another way to view V c: V )w, where V c = (id ⊗ πc πc V = ⊤◦πV ◦R and R is the unitary antipode of G. This can be seen from (cid:0)id ⊗ [⊤◦πV ◦R](cid:1)w = (bR ⊗ ⊤)(id ⊗ πV )(bR ⊗ R)w = (bR ⊗ ⊤)(id ⊗ πV )w = V c 2.3. Containment, weak containment, equivalence, etc. Since there is a one to one corre- because (bR ⊗ R)w = w ([24, Formula 5.34]). spondence between corepresentations of bG and representations of the C∗-algebra C(Gmax) we can define the notions of containment, weak containment, equivalence and weak equivalence of corep- resentations by the corresponding notions from representation theory of C∗-algebras (see e.g. [13] or Section 4). We will write U ≤ V if U is contained in (i.e. is a sub-corepresentation of) V in the sense that πU is a subrepresentation of πV . Similarly we will write U 4 V if πU 4 πV (weak containment). Two corepresentations U and V are equivalent if πU and πV are unitarily equivalent, while U and V are weakly equivalent if U 4 V and V 4 U . We have the following simple lemma: Lemma 2.1. Let U, U1, V and V1 be corepresentations of a discrete quantum group bG. Then (1) if U ≤ U1 and V ≤ V1 then U  (2) if U ≤ V then U c ≤ V c. Remark 2.2. Let U be a finite dimensional corepresentation of a discrete quantum group bG, i.e. U ∈ M(cid:0)c0(bG) ⊗ K (HU )(cid:1) and dim HU = n < ∞. Then, upon choosing an orthonormal basis in HU , we can identify U with an n× n unitary matrix of elements of M(cid:0)c0(bG)(cid:1) which satisfy ⊤ V ≤ U1  ⊤ V1, for i, j = 1, . . . , n. If we put ui,j = U ∗ b∆(Ui,j) = j,i then b∆(ui,j) = nXk=1 nXk=1 Uk,j ⊗ Ui,k ui,k ⊗ uk,j. matrix group (to see that condition 3. of that definition is satisfied, consider the restriction of the Let B be the C∗-subalgebra of M(cid:0)c0(bG)(cid:1) generated by {ui,j}i,j=1,...,n. Then B is unital (because U is unitary) and b∆ restricts to a comultiplication B → B⊗B. Then(cid:0)B,b∆(cid:12)(cid:12)B(cid:1) is a compact quantum antipode of bG to the ∗-algebra generated by matrix elements of U ∗, cf. [33, Theorem 1.6(4)]). Furthermore, U is a unitary corepresentation of the opposite quantum group ([17, Section 4]). Using the results of [30, Section 3], [17, Section 4] and [22, Subsection 4.6] one can show that ⊤ U c contains the trivial representation. Note, however, that the definition of contragredient U  corepresentation in [30] is different from the one we have adopted and one is forced to use modular properties of the Haar measure of (cid:0)B,b∆(cid:12)(cid:12)B(cid:1). 6 DAVID KYED AND PIOTR M. SO LTAN 2.4. The regular corepresentation. The regular corepresentation of a discrete quantum group bG is W = (id ⊗ λ)w, where λ is the quotient map C(Gmax) → C(Gmin) ⊂ B(cid:0)L2(G)(cid:1). Proposition 2.3. The regular corepresentation is equivalent to its contragredient W Proof. Let us first define a unitary map Z : L2(G) → L2(G). We put c. Zuα k,l = R(uα k,l ∗), where R is the unitary antipode of G. The unitarity of Z follows from the calculation: i,jE . Let us examine the operator Zλ(a)Z ∗ for a ∈ Pol(G). On a vector R(uα k,l uβ k,l k,l ∗) ∈ L2(G) we have: Zλ(a)Z ∗R(uα k,l DZuα k,l Zuβ ∗ ∗ i,j i,j k,l ∗uβ ) R(uα k,l )∗R(uα k,l ∗uβ i,jE =DR(uβ ∗)E = h(cid:0)R(uβ ∗)(cid:1) = h(cid:0)R(uα i,j)(cid:1) i,j(cid:1) =Duα = h(cid:0)uα ∗) = Z(cid:0)λ(a)uα k,l(cid:1) = Z(a · uα k,l) k,l)∗(cid:1) = R(cid:0)(auα = R(cid:0)uα ∗a∗(cid:1) ∗(cid:1) = R(a∗) · R(cid:0)uα ∗(cid:1) R(cid:0)uα = λ(cid:0)R(a)(cid:1)∗ R(cid:0)uα ∗(cid:1). = λ(cid:0)R(a)(cid:1)⊤ k,l k,l k,l k,l Thus Z establishes unitary equivalence between λ and ⊤◦ λ◦ R, which is the same as unitary equivalence between W and W c. (cid:3) Remark 2.4. It is a well known fact that the tensor product W (cf. [24, Corollary 20]). In view of Proposition 2.3, we see that W ⊤ W is weakly contained in W ⊤ W c 4 W. 2.5. A theorem about corepresentations. We end this section with a quantum group gener- alization of [7, Proposition A.1.12] which gives a necessary and sufficient condition for a tensor product of two representations of a discrete group to have an invariant vector. Let H and K be Hilbert spaces and denote by HS(H, K) the space of Hilbert-Schmidt operators from H to K. There is a canonical unitary mapping Ψ : H ⊗ K −→ HS(cid:0)K, H(cid:1) given by x ⊗ y 7−→ xihy , where we use the Dirac notation: xihy is the operator K ∋ z 7−→ hy zi x ∈ H. This yields an isomorphism AdΨ : B(H ⊗ K) → B(cid:0)HS(cid:0)K, H(cid:1)(cid:1) Lemma 2.5. For S ∈ B(H), R ∈ B(K) and T ∈ HS(cid:0)K, H(cid:1) we have Proof. Calculate for T = Ψ(x ⊗ y) = xihy and extend the result by linearity and continuity. (cid:3) Theorem 2.6. Let U and V be corepresentations of a discrete quantum group bG. Then (cid:0)AdΨ(S ⊗ R)(cid:1)(T ) = S◦T ◦R⊤. AdΨ(x) = ΨxΨ∗. PROPERTY (T) AND EXOTIC QUANTUM GROUP NORMS 7 U  ⊤ V contains the trivial corepresentation. ⊤ V contains the trivial corepresentation then there exists a Proof. Ad (1). This follows directly from Lemma 2.1 and Remark 2.2. finite-dimensional corepresentation W contained both in U and in V c. (1) if W is a finite dimensional corepresentation of bG such that W ≤ U and W ≤ V c then (2) if bG is unimodular and U  Ad (2). As in the remarks preceding Lemma 2.5 we write Ψ for the canonical unitary HU⊗HV → HS(cid:0)HV , HU(cid:1) and AdΨ for the isomorphism B(HU ⊗ HV ) → B(cid:0)HS(cid:0)HV , HU(cid:1)(cid:1) Let us form the tensor product U  ⊤ V ) ∈ M(cid:0)c0(bG) ⊗ K(cid:0)HS(cid:0)HV , HU(cid:1)(cid:1)(cid:1). ⊤ V , so that X contains the trivial corepresen- X = (id ⊗ AdΨ)(U  ⊤ V . Let Since tation. This means that X has a non-zero invariant vector. Then X is a corepresentation of bG equivalent to U  the component X α ∈ Mnα(C) ⊗ B(cid:0)HS(cid:0)HV , HU(cid:1)(cid:1) is X = (id ⊗ AdΨ)(cid:0)id ⊗(cid:2)(πU ⊗ πV )◦∆(cid:3)(cid:1)w, X α = (id ⊗ AdΨ)(cid:0)id ⊗(cid:2)(πU ⊗ πV )◦∆(cid:3)(cid:1)uα = (id ⊗ AdΨ)(cid:0)id ⊗(cid:2)(πU ⊗ πV )◦∆(cid:3)(cid:1) nαXi,j=1 eα i,j ⊗ uα = (id ⊗ AdΨ)(id ⊗ πU ⊗ πV ) nαXi,j,k=1 so that i,j ⊗(cid:18) nαXk=1 πU (uα = (id ⊗ AdΨ) eα nαXi,j i,j = AdΨ(cid:18) nαXk=1 i,k) ⊗ πV (uα By Lemma 2.5, for T ∈ HS(cid:0)HV , HU(cid:1) and η ∈ H α we have πU (uα X α X α(η ⊗ T ) = nαXi,j eα i,jη ⊗(cid:18) nαXk=1 πU (uα i,k)◦T ◦πV (uα k,j)⊤(cid:19) eα i,j ⊗ uα i,j k,j i,k ⊗ uα k,j)(cid:19) i,k) ⊗ πV (uα k,j)(cid:19). Now let T be an invariant vector for X. In view of the discussion at the end of Subsection 1.2 X α i,j(T ) = δi,jT which reads nαXk=1 πU (uα i,k)◦T ◦πV (uα k,j)⊤ = δi,jT. We have assumed that bG is unimodular, i.e. that G is of Kac type. In particular, if κ is the antipode of G then κ = R is a ∗-anti-automorphism and κ2 = id. Moreover κ(uα ∗. Therefore k,j) = uα j,k Multiplying both sides of this equation by πc j,p) and summing over p we obtain nαXk=1 πU (uα V (uα j,k ∗) = δi,j T i,k)◦T ◦πc V (uα j,p(cid:19) = ∗uα uα j,k V(cid:18) nαXp=1 i,p)◦T = T ◦πc πU (uα V (uα i,p) πU (uα i,k)◦T ◦πc nαXp=1 δi,jT ◦πc V (uα j,p) nαXk=1 or equivalently because uα is unitary. (2.1) 8 DAVID KYED AND PIOTR M. SO LTAN Since (2.1) is true for all α ∈ Irr(G) and all i, p ∈ {1, . . . , nα}, we have U (1 ⊗ T ) = (1 ⊗ T )V c i.e. T intertwines V c and U . It follows that T T ∗ ∈ K (HU ) intertwines U with itself. Note that T T ∗ is a non-zero com- pact, positive operator. Therefore it has an eigenvalue λ > 0 with finite multiplicity. Moreover the corresponding eigenprojection also intertwines U with itself. This clearly leads to a finite dimensional sub-corepresentation W of U . Similarly T ∗T is a self-intertwiner of V c and there is a sub-corepresentation W ′ of V c corresponding to λ (the non-zero parts of spectra of T T ∗ and T ∗T coincide). Moreover, it is easy to see that the partial isometric part of the polar decomposition of T ∗ establishes an equivalence between W and W ′. (cid:3) Remark 2.7. Let us remark that the first part of Theorem 2.6 in the Kac case can be established in a simple calculation without resorting to the techniques described in Remark 2.2. Indeed, using ⊤ W c. The the notation of Theorem 2.6 (and its proof), we first note that U  corepresentation W  ⊤ W c is equivalent to ⊤ V contains W  (id ⊗eπ)w ∈ M(cid:0)c0(bG) ⊗ K (HS(HW , HW ))(cid:1) where the representation eπ when restricted to Pol(Gmax) is In view of Lemma 2.5, this means that for T ∈ HS(HW , HW ) and a ∈ Pol(Gmax) we have eπ : Pol(Gmax) ∋ a 7−→ AdΨ(cid:0)(πW ⊗ πW )(id ⊗ κ)∆(a)(cid:1). (cid:0)eπ(a)(cid:1)(T ) =X πW (a(1))◦T ◦πW(cid:0)κ(a(2))(cid:1). Since HW is finite-dimensional, we can take T = 1 to obtain (cid:0)eπ(a)(cid:1)(1) = (πW ⊗ πW )(cid:18)X(a(1))(cid:0)κ(a(2))(cid:1)(cid:19)1 = (πW ⊗ πW )(cid:0)m(id ⊗ κ)∆(a))1 = ε(a)1 for all a ∈ Pol(Gmax). It follows that the trivial corepresentation is contained in W  ⊤ W c. 3. Property (T) for discrete quantum groups In a recent paper by P. Fima [14], Kazhdan's property (T) is studied in the setting of discrete quantum groups. The definition is analogous to the classical definition for discrete groups and goes as follows. (cid:13)(cid:13)V α(η ⊗ ξ) − η ⊗ ξ(cid:13)(cid:13) < δkηkkξk Definition 3.1 ([14]). Let bG be a discrete quantum group and let V ∈ M(cid:0)c0(bG) ⊗ K (HV )(cid:1) be a unitary corepresentation of bG on the Hilbert space HV . For E ⊂ Irr(G) a finite subset and δ > 0 a vector ξ ∈ HV is said to be (E, δ)-invariant with respect to V if for all α ∈ E and all η ∈ H α. The corepresentation V has almost invariant vectors if such a ξ ∈ HV exists for all finite subsets E ⊆ Irr(G) and all δ > 0, and the discrete quantum group bG is said to have property (T) if every corepresentation with almost invariant vectors has a non-zero invariant vector. Remark 3.2. It was shown in [18] how property (T) for bG can be interpreted using the corre- spondence between corepresentations of bG and representations of C(Gmax). More precisely, bG has property (T) if and only if the following holds: if π : C(Gmax) → B(H) is a representation and n→∞(cid:13)(cid:13)π(a)ξn − ε(a)ξn(cid:13)(cid:13) = 0 for all there exists a sequence (ξn)n∈ of unit vectors H such that lim a ∈ C(Gmax) then there exists a unit vector ξ ∈ H with π(a)ξ = ε(a)ξ for all a ∈ C(Gmax). Remark 3.3. Actually, the study of property (T) for quantum groups began before the paper [14]. In [19] property (T) was studied in the setting of Kac algebras and in [4] it was introduced for the class of algebraic quantum groups. However, we will use Fima's approach in the following, since it fits our purposes best. Using the results obtained in the present paper we will see later that the different approaches are equivalent in the case of discrete quantum groups. PROPERTY (T) AND EXOTIC QUANTUM GROUP NORMS 9 In the following theorem we summarize some of the results obtained in [14]. Theorem 3.4. If bG is a discrete quantum group with property (T) then the following holds: (1) bG is finitely generated, i.e. the compact dual is a matrix quantum group. (2) There exists a finite subset E0 ⊆ Irr(G) and a δ0 > 0 such that every corepresentation with (E0, δ0)-invariant vectors has a non-zero invariant vector. Such a pair (E0, δ0) is called a Kazhdan pair for bG. (3) bG is unimodular. Furthermore, Fima links property (T) of bG with property (T) of L∞(G) (in the sense of Connes and Jones [10]) in the case whenbG is i.c.c. Property (T) forbG can also be described by means of the "positive definite functions" on bG as well as by a vanishing of cohomology result analogues to the classical Delorme-Guichardet theorem. We shall not elaborate further on these characterizations and refer the reader to [18] for details. Property (T) turns out to be essential in our search for exotic quantum group norms and in the following section we develop the results needed to construct these norms. The results obtained are of independent interests and parallel nicely classical results for discrete groups. 4. Property (T) and the Jacobson topology Let again G be a compact quantum group with C∗-algebra C(G) and Hopf ∗-algebra Pol(G). In this section we investigate the connection between property (T) forbG and the topology on the spec- trum Spec(cid:0)C(Gmax)(cid:1) consisting of equivalence classes of irreducible representations of C(Gmax). Recall from [13] that Spec(cid:0)C(Gmax)(cid:1) has a natural topology, called the Jacobson topology, which is intimately linked with the notion of weak containment. For the convenience of the reader we briefly recall this notion. Definition 4.1 ([13]). Let S be a set of representations of C(Gmax) and π some given representa- tion. Then π is said to be weakly contained in S, written π 4 S, if every vector state associated with π is a weak∗ limit (i.e. pointwise limit) of states which are linear combinations of vector functionals associated with the representations in S. Proposition 4.2 ([13, Theorem 3.4.10]). If S ⊂ Spec(cid:0)C(Gmax)(cid:1) and π ∈ Spec(cid:0)C(Gmax)(cid:1) then the following are equivalent: (1) π is in the closure of S (with respect to the Jacobson topology) in Spec(cid:0)C(Gmax)(cid:1), (2) π is weakly contained in S, (3) every vector state associated with π is the weak∗ limit of vector states associated with S. Recall that a functional ϕ : C(Gmax) → C is said to be a vector functional associated with as set of representations S if there exists ρ ∈ S and ξ ∈ Hρ such that ϕ(a) = hξ ρ(a)ξi for all a ∈ C(Gmax). Lemma 4.3. Let π : C(Gmax) → B(H) be a representation. Then π has almost invariant vectors if and only if the counit ε is weakly contained in π. Proof. If π has almost invariant vectors there exists a sequence (ξn)n∈ of unit vectors in H such that for all a ∈ C(G). Defining ϕn(a) = hξn π(a)ξni we have (cid:12)(cid:12)ϕn(a) − ε(a)(cid:12)(cid:12)2 0 (cid:13)(cid:13)π(a)ξn − ε(a)ξn(cid:13)(cid:13) −−−−→n→∞ =(cid:13)(cid:13)π(a)ξn − ε(a)ξn(cid:13)(cid:13)2 lim n→∞(cid:12)(cid:12)ϕn(a) − ε(a)(cid:12)(cid:12) = 0 and ϕn(a∗a) − ϕ(a∗)ϕ(a) ≥ 0 by the Cauchy-Schwarz inequality. Thus −(cid:0)ϕn(a∗a) − ϕn(a∗)ϕn(a)(cid:1). (4.1) and we conclude that ε 4 π. 10 DAVID KYED AND PIOTR M. SO LTAN Conversely, if ε 4 π we get a net (ξι) of unit vectors in H such that the net of vectors states (ϕι), ϕι(a) = hξι π(a)ξιi converges pointwise to ε on C(Gmax). But then for each ι we have (cid:12)(cid:12)ϕι(a) − ε(a)(cid:12)(cid:12)2 =(cid:13)(cid:13)π(a)ξι − ε(a)ξι(cid:13)(cid:13)2 −(cid:0)ϕι(a∗a) − ϕι(a∗)ϕι(a)(cid:1) ι (cid:13)(cid:13)π(a)ξι− ε(a)ξι(cid:13)(cid:13) = 0 for each a ∈ C(Gmax). This shows that π has almost (cid:3) as in (4.1) and hence lim invariant vectors. Definition 4.4. Let Ω be a set of positive functionals on C(Gmax) and let ϕ be another positive functional. Then ϕ is said to be approximated on finite sets by elements in Ω if the following holds: for all finite E ⊂ Irr(G) and all δ > 0 there exists ω ∈ Ω such that for all α ∈ E and all i, j ∈ {1, . . . , nα} Lemma 4.5. A positive functional ϕ on C(Gmax) is approximated on finite sets by elements from Ω ⊂ C(Gmax)∗ + if and only if there exists a sequence (ωn)n∈ of elements of Ω such that ϕ in the weak∗ topology. ωn(a) −−−−→n→∞ Proof. If ϕ is approximated by functionals from Ω on finite sets just pick an increasing sequence (En)n∈ of finite subsets of Irr(G) with Irr(G) as its union and choose ωn ∈ Ω such that ϕ(a) for every a ∈ Pol(G). Moreover, in this case ωn −−−−→n→∞ ∀ α ∈ En ∀ i, j ∈ {1, . . . , nα} : (cid:12)(cid:12)ϕ(uα i,j) − ωn(uα i,j)(cid:12)(cid:12) < 1 n . Since each matrix coefficient is contained in En from a certain point on, we get the desired pointwise convergence on the set of matrix coefficients, and since these span Pol(G) linearly, the pointwise convergence holds on all of Pol(G). If, conversely, we have a sequence (ωn)n∈ of elements of Ω converging pointwise to ϕ on Pol(G) then clearly ϕ is approximated by functionals in Ω on finite subsets. That the convergence holds on all of C(Gmax) is seen by a standard "epsilon over three" argument. (cid:3) Remark 4.6. Lemma 4.5 shows that ϕ is approximated on finite sets by elements of Ω if and only if ϕ lies in the weak∗ closure of Ω. We will also use the following functional analytic version of a Lemma in [7]. Lemma 4.7. If ϕ1 and ϕ are non-zero, positive linear functionals on C(Gmax) and ϕ ≥ ϕ1 then the GNS representation π1 associated with ϕ1 is contained in the GNS representation π associated with ϕ. If ϕ is already a vector functional associated to some representation ρ then π ≤ ρ. Proof. Let H and H1 be the GNS Hilbert spaces associated to ϕ1 and ϕ respectively, with cyclic vectors Ξ1 and Ξ. Since ϕ − ϕ1 ≥ 0 we get (cid:12)(cid:12)ϕ(uα i,j ) − ω(uα i,j)(cid:12)(cid:12) < δ. In particular (cid:13)(cid:13)π1(a)Ξ1(cid:13)(cid:13)2 = hΞ1 π1(a∗a)Ξ1i = ϕ1(a∗a) ≤ ϕ(a∗a) =(cid:13)(cid:13)π(a)Ξ(cid:13)(cid:13)2 (cid:0)π(a)Ξ = 0(cid:1) =⇒ (cid:0)π1(a)Ξ1 = 0(cid:1). . 1 2 . Thus T : H → H1 defined by π(a)Ξ 7→ π1(a)Ξ1 is well defined and bounded. Moreover, T is trivially seen to intertwine π and π1. Put K = (ker T )⊥ and note that K is π-invariant. Since T ∗ intertwines π1 and π the operator T ∗T is a self-intertwiner of π and by functional calculus the same is true for T = (T ∗T ) Consider now the polar decomposition T = UT, where U is an isometry from K onto ran T = If we can prove that U is also an intertwiner then U ∗ provides us with an equivariant H1. embedding of H1 into H proving that π1 ≤ π. To see that U intertwines we calculate for any ξ ∈ H which shows that U restricted to ran(T) intertwines. But since ran(T) = ran(T ∗) = (ker T )⊥ = K, we are done. If furthermore ϕ(a) = hη ρ(a)ηi for some representation ρ on a Hilbert space L then V : H → L given by π(a)Ξ 7→ ρ(a)η is a well defined isometry intertwining the GNS representation π with ρ. π1(a)UTξ = π1(a)T ξ = T π(a)ξ = UTπ(a)ξ = U π(a)Tξ, (cid:3) PROPERTY (T) AND EXOTIC QUANTUM GROUP NORMS 11 Remark 4.8. Note that if (in the above proof) η is cyclic for ρ then V is an equivalence between π and ρ. This, for instance, is always the case if ρ is irreducible. With the aid of the above lemmas we are now able to prove the following quantum group generalization of the classical characterization of property (T) in terms of Fell's topology. Theorem 4.9. A discrete quantum group bG has property (T) if and only if the trivial representa- tion ε is an isolated point in Spec(cid:0)C(Gmax)(cid:1). Proof. Assume first that bG has property (T). Since ε is finite dimensional and irreducible {ε} is automatically closed in Spec(cid:0)C(Gmax)(cid:1), so we need to show that {ε} is also open. Now the complement {ε}∁ is closed if and only if ε 6∈ {ε}∁ which happens if and only if ε is not weakly contained in {ε}∁. If this were the case then by [13, Theorem 3.4.10] we find a net (πι) of elements of Spec(cid:0)C(Gmax)(cid:1)\{ε} and for each ι a unit vector ξι in the representation space Hι of πι such that the vector functionals ϕι : a 7→ hξι πι(a)ξιi converge pointwise to ε on C(Gmax). The functionals ϕι satisfy (cid:12)(cid:12)ϕι(a) − ε(a)(cid:12)(cid:12)2 (cf. (4.1)) and hence (cid:13)(cid:13)πι(a)ξι − ε(a)ξι(cid:13)(cid:13) −→ι Define now −(cid:0)ϕι(a∗a) − ϕι(a∗)ϕι(a)(cid:1) =(cid:13)(cid:13)πι(a)ξι − ε(a)ξι(cid:13)(cid:13)2 π =Mι 0 for all a ∈ C(Gmax). πι : C(Gmax) −→ B(cid:16)Mι Hι(cid:17). Then, by construction, π has almost invariant vectors and by property (T) it must have a non-zero invariant unit vector η = (ηι). Then at least one ηι0 is non-zero and hence invariant for πι0. Thus ε ≤ πι0 contradicting the choice of πι0 in Spec(cid:0)C(Gmax)(cid:1) \ {ε}. Assume now that bG does not have property (T) and pick a representation π : C(Gmax) → B(H) with almost invariant vectors but without non-zero invariant ones. We may therefore choose a sequence (ξn)n∈ of unit vectors in H such that (cid:13)(cid:13)π(a)ξn − ε(a)ξn(cid:13)(cid:13) −−−−→n→∞ 0 for every a ∈ Pol(G). Putting ϕn(a) = hξn π(a)ξni we obtain a sequence of positive functionals satisfying relation (4.1). ε(a) for all a ∈ Pol(G) and Lemma 4.5 assures that ϕn → ε in the weak∗ Hence ϕn(a) −−−−→n→∞ topology. Our aim is to show that ε is not an isolated point in the spectrum, i.e. that ε is weakly contained in Spec(cid:0)C(Gmax)(cid:1) \ {ε}. in Spec(cid:0)C(Gmax)(cid:1) \ {ε}. Denote this set by Ω. Let E be a finite subset of Irr(G) and let δ > 0 be given. Choose an n0 ∈  such that (cid:12)(cid:12)ϕn0 (uα Hence, by Lemma 4.5 we have to show that ε can be approximated on finite sets by elements from the set consisting of linear combinations of positive functionals associated with the representations for all α ∈ E and i, j ∈ {1, . . . , nα}. Recall also that ϕn0 (a) = hξn0 π(a)ξn0i, with ε (cid:2) π. Since the state space S(cid:0)C(Gmax)(cid:1) is the weak∗-closed convex hull of the the set of pure states of C(Gmax), there exists a net (ϕι) of elements of S(cid:0)C(Gmax)(cid:1) converging pointwise to ϕn0 and with where tι ∈ [0, 1] and ψι is a linear combination of pure states different from ε, i.e. ψι ∈ Ω. By compactness of [0, 1] and weak∗ compactness of S(cid:0)C(Gmax)(cid:1) we may, upon passing to subnets, assume that tι −→ι t and (ψι) converges pointwise to a state ψ. Then ϕι = tιψι + (1 − tι)ε, i,j) − ε(uα i,j)(cid:12)(cid:12) < δ 2 the property that If t 6= 1 then Lemma 4.7 implies that ε is contained in the GNS representation associated to ϕn0 which, in turn, is contained in π -- contradiction with the choice of π. Hence t = 1 and thus ϕn0 is the pointwise limit of the net (ψι). Hence there exists an index ι0 such that ϕn0 = tψ + (1 − t)ε. (cid:12)(cid:12)ϕn0 (uα i,j) − ψι0(uα i,j)(cid:12)(cid:12) < δ 2 12 DAVID KYED AND PIOTR M. SO LTAN i,j) − ψι0(uα Remark 4.10. Equipped with Theorem 4.9 one can easily prove that a discrete quantum group for all α ∈ E and all i, j ∈ {1, . . . , nα}. Thus for all α ∈ E and i, j ∈ {1, . . . , nα} we have functionals in Ω as desired. i,j)(cid:12)(cid:12) < δ and since ψι0 is in the set Ω, we have shown that ε is approximated by (cid:12)(cid:12)ε(uα bG has property (T) in the sense of Definition 3.1 if and only if bG has property (T) as defined by E. B´edos, R. Conti and L. Tuset in [5, Definition 7.15] and if and only if the associated Kac algebra (in von Neumann algebraic formulation) has property (T) as defined by Petrescu and Joita in [19, Definition 3.1] (cf. [19, Theorem 3.3]). (cid:3) 5. Connection with property (T) for C∗-algebras In the paper [6] B. Bekka introduced property (T) for unital C∗-algebra endowed with a tracial state. His definition is a C∗-analogue of the corresponding definition for II1-factors due to Connes and Jones ([10]) and goes as follows: Definition 5.1. A unital C∗-algebra A admitting a tracial state is said to have property (T) if there exists a finite F ⊂ A and a constant c > 0 such that if a Hilbert A-bimodule H has a unit vector ξ such that kaξ − ξak < c (T) in the sense of Bekka. for all a ∈ F then there exists a non-zero vector ξ′ ∈ H such that aξ = ξa for all a ∈ A. Theorem 5.2. The discrete quantum group bG has property (T) if and only if C(Gmax) has property Note that the counit ε : C(Gmax) → C is a tracial state so that Bekka's definition, which only covers C∗-algebras admitting tracial states, can be applied. The proof is greatly inspired by the the proof of [14, Theorem 3]. Proof of Theorem 5.2. Assume that bG has property (T) and let (E, δ) be a Kazhdan pair. We now prove that E ′ =(cid:8)uα i,j α ∈ E, i, j ∈ {1, . . . , nα}(cid:9) and δ′ = δ max{nα√nα α ∈ E} constitute a Kazhdan pair for the C∗-algebra C(Gmax). Assume therefore that H is a Hilbert space which is also a C(Gmax)-bimodule and assume furthermore that ξ is an (E ′, δ′)-central vector; i.e. that kuα i,jξ − ξuα i,jk < δ′ for all α ∈ E and i, j ∈ {1, . . . , nα}. Denoting the left action C(Gmax) → B(H) by π and the right action C(Gmax)op → B(H) by ρ we obtain a new representation σ : C(Gmax) → B(H) by setting σ = m◦(cid:0)[ρ◦ R] ⊗ π(cid:1)◦ ∆ which corresponds to the corepresentation V of bG given by V α = (id ⊗ ρ)(uα ∗)(id ⊗ π)(uα). For α ∈ E we now obtain, using (1.3) and the fact that G is Kac so that (5.1) (cid:8)√nαuα i,j i, j = 1, . . . , nα(cid:9) is an orthonormal basis of H α, that (cid:13)(cid:13)V α(uα r,l ⊗ ξ) − uα r,l ⊗ ξ(cid:13)(cid:13) =(cid:13)(cid:13)(cid:0)(id ⊗ π)(uα)(cid:1)(uα i,j ⊗ π(uα =(cid:13)(cid:13)(cid:13)(cid:13) nαXi,j=1(cid:0)eα =(cid:13)(cid:13)(cid:13)(cid:13) nαXi=1 =(cid:13)(cid:13)(cid:13)(cid:13) nαXi=1 uα r,i ⊗ π(uα uα r,i ⊗ (uα r,l ⊗ ξ)(cid:13)(cid:13) r,l ⊗ ξ) −(cid:0)(id ⊗ ρ)(uα)(cid:1)(uα nαXi,j=1(cid:0)eα i,j)(cid:1)(uα i,j)(cid:1)(uα i,j ⊗ ρ(uα r,l ⊗ ξ) − i,l)ξ(cid:13)(cid:13)(cid:13)(cid:13) nαXi=1 i,l)ξ − i,l)(cid:13)(cid:13)(cid:13)(cid:13) i,lξ − ξuα uα r,i ⊗ ρ(uα r,l ⊗ ξ)(cid:13)(cid:13)(cid:13)(cid:13) PROPERTY (T) AND EXOTIC QUANTUM GROUP NORMS 13 =vuut =vuut nαXi=1 nαXi=1 kuα r,ik2kuα i,lξ − ξuα i,lk2 1 nαkuα i,lξ − ξuα i,lk2 < δ′. Now we take η ∈ H α and expand it in the orthonormal basis (5.1): η = nαXr,i=1 nα(cid:10)uα r,i η(cid:11) uα r,i. Then so that ≤ r,l ⊗ ξ) − uα r,l ⊗ ξ) − uα (cid:13)(cid:13)V α(η ⊗ ξ) − η ⊗ ξ(cid:13)(cid:13) =(cid:13)(cid:13)(cid:13)(cid:13) nαXr,i=1 nα(cid:10)uα nαXr,i=1 nα(cid:12)(cid:12)(cid:10)uα nαXr,i=1 nα(cid:12)(cid:12)(cid:10)uα r,i η(cid:11)(cid:12)(cid:12) ≤vuut nαXr,i=1(cid:12)(cid:12)(cid:10)√nαuα nαXr,i=1(cid:12)(cid:12)(cid:10)√nαuα (cid:13)(cid:13)V α(η ⊗ ξ) − η ⊗ ξ(cid:13)(cid:13) < δ′√nα r,l ⊗ ξ(cid:1)(cid:13)(cid:13)(cid:13)(cid:13) r,i η(cid:11)(cid:0)V α(uα r,l ⊗ ξ(cid:13)(cid:13) r,i η(cid:11)(cid:12)(cid:12)(cid:13)(cid:13)V α(uα nαXr,i=1(cid:12)(cid:12)(cid:10)√nαuα r,i η(cid:11)(cid:12)(cid:12)δ′ = δ′√nα r,i η(cid:11)(cid:12)(cid:12)2vuut nαXr,i=1 r,i η(cid:11)(cid:12)(cid:12) ≤ δ′√nαkηknα = δ′kηkn nαXr,i=1(cid:12)(cid:12)(cid:10)√nαuα 1 = kηknα, < r,i η(cid:11)(cid:12)(cid:12) By the Cauchy-Schwarz inequality 3 2 α ≤ δkηk It is easily seen that ξ′ is a central vector and we conclude that C(Gmax) has Remark 5.3. Let us emphasize that Theorem 5.2 together with [8, Proposition 3.2] shows that, If, conversely, C(Gmax) has property (T) then from [8, Proposition 3.2] it follows that every for all η ∈ H α. Therefore, since (E, δ) is a Kazhdan pair for bG, there exists a V -invariant unit vector ξ′ ∈ H. property (T). finite dimensional, irreducible representation is an isolated point in Spec(cid:0)C(Gmax)(cid:1). In particular ε is an isolated point and therefore bG has property (T) by Theorem 4.9. as in the classical case, bG has property (T) if and only if all finite dimensional representations of C(Gmax) are isolated in Spec(cid:0)C(Gmax)(cid:1). Corollary 5.4. Let bG be an infinite discrete quantum group. i.e. one with dim c0(bG) = ∞. Assume that bG has property (T). Then the regular corepresentation W of bG does not weakly contain any Proof. If U is a finite dimensional corepresentation of bG and U 4 W then by Remark 5.3 we have U ≤ W. By Lemma 2.1 and Remark 2.4 we have finite dimensional corepresentation. (cid:3) ⊤ U c ≤ W U  ⊤ W c 4 W. ⊤ U c contains the trivial representation (Remark 2.2), so W must weakly contain the triv- But U  ial representation which is impossible for property (T) infinite discrete quantum group (cf. [5, Theorem 7.17], [14, Remark 4]). (cid:3) 14 DAVID KYED AND PIOTR M. SO LTAN 6. Minimal projections and property (T) We recall that a projection p in a unital C∗-algebra A is called minimal if pAp = Cp. We prove here the following quantum group version of the classical characterization of property (T) in terms of minimal projections in the maximal group C∗-algebra (see [1, 25]). The proof follows the lines of the corresponding proof in [25]. As usual bG denotes a discrete quantum group. Proposition 6.1. The following are equivalent: (1) bG has property (T), (2) there exists a unique minimal projection in the center of C(Gmax) with ε(p) = 1, (3) there exists a minimal projection p ∈ C(Gmax) with ε(p) = 1. Proof. We first prove (1)⇒(2). If bG has property (T) then ε is isolated in Spec(cid:0)C(Gmax)(cid:1). Hence the spectrum splits into a disjoint union of open subsets as Spec(cid:0)C(Gmax)(cid:1) = {ε} ∪ {ε}∁ and thus C(Gmax) splits accordingly (as a C∗-algebra!) into the direct sum of two closed, two-sided ideals I and J defined, implicitly, as {ε} =(cid:8)π ∈ Spec(cid:0)C(Gmax)(cid:1) π(I) 6= {0}(cid:9), {ε}∁ =(cid:8)π ∈ Spec(cid:0)C(Gmax)(cid:1) π(J) 6= {0}(cid:9). Clearly we have J = ker ε and thus I is one-dimensional. The unit now splits as 1 = (e, f ) and p = (e, 0) clearly does the job. If another minimal, central projection p′ with ε(p′) = 1 existed then we would have pp′ = pp′p = λp for some λ ∈ C and since ε(p) = ε(p′) = 1 we have λ = 1. Thus p ≤ p′ and by minimality p = p′. The implication (2)⇒(3) is obvious. Lastly we prove (3)⇒(1). Let therefore p ∈ C(Gmax) be minimal with ε(p) = 1. Then by [25, Lemma 1] there exists a unique π ∈ Spec(cid:0)C(Gmax)(cid:1) such that π(p) 6= 0. Since ε clearly is such a representation we have {ε} =(cid:8)ρ ∈ Spec(cid:0)C(Gmax)(cid:1) ρ(p) 6= 0(cid:9) =(cid:8)ρ ∈ Spec(cid:0)C(Gmax)(cid:1) ρ(cid:0)C(Gmax)p C(Gmax)(cid:1) 6= {0}(cid:9) which by definition is open in Spec(cid:0)C(Gmax)(cid:1). Since {ε} is always closed, this proves that ε is an isolated point in Spec(cid:0)C(Gmax)(cid:1) and hence that bG has property (T). 7. Quantum group norms (cid:3) In [4, Section 3] the question of completing the polynomial algebra Pol(G) under different C∗- norms was addressed. In particular, the authors considered C∗-norms on Pol(G) for which the comultiplication Pol(G) → Pol(G) ⊗alg Pol(G) extends to a ∗-homomorphism of the completions. Such norms were called regular. We feel that the term "regular" is already overused in the literature on quantum groups (let us mention e.g. the regularity condition for multiplicative unitaries of [2, 3] or the regular corepresentation of Subsection 2.4). Therefore we would like to propose the following terminology: Definition 7.1. Let G be a compact quantum group and let k · k∼ be a C∗-norm on Pol(G). Let C(G∼) be the completion of Pol(G) in the norm k · k∼. The C∗-norm k · k∼ is called a quantum group norm if the comultiplication ∆ : Pol(G) → Pol(G)⊗alg Pol(G) extends to a ∗-homomorphism C(G∼) → C(G∼) ⊗ C(G∼). B´edos, Murphy and Tuset proved, among other things, that the norm coming from the repre- sentation of Pol(G) on L2(G) is the smallest quantum group norm on Pol(G) (cf. Remark 2.4 for an argument that it is a quantum group norm). Also the universal or maximal C∗-norm on Pol(G), i.e. the supremum of all C∗-norms on Pol(G), was proved in [4] to be a quantum group norm. In the next sections we will construct examples of quantum group norms with various interesting properties. In particular we will obtain examples of compact quantum groups G sitting strictly "between" their minimal and maximal versions. We will provide such examples both admitting a continuous counit and without this property. 0 0 eπ : Pol(G) ∋ a 7−→(cid:20)π(a) ε(a)(cid:21) ∈ B(H ⊕ C) kakeπ =(cid:13)(cid:13)eπ(a)(cid:13)(cid:13) = max(cid:8)(cid:13)(cid:13)π(a)(cid:13)(cid:13),(cid:12)(cid:12)ǫ(a)(cid:12)(cid:12)(cid:9) and let k · keπ be the norm defined by eπ: for all a ∈ Pol(G). Proposition 8.1. The C∗-norm k · keπ on Pol(G) is a quantum group norm. Proof. Take a ∈ Pol(G). We have (eπ⊗eπ)∆(a) =Xeπ(a(1)) ⊗eπ(a(2)) π(a(1)) ⊗ π(a(2)) π(a(1))ε(a(2))   =X 0 0 0  (π ⊗ π)∆(a)  0 0 0 = 0 0 0 0 0 0 ε(a)   . 0 π(a) 0 0 0 0 π(a) 0 0 0 ε(a(1))π(a(2)) 0 0 0 0 ε(a(1))ε(a(2))   PROPERTY (T) AND EXOTIC QUANTUM GROUP NORMS 15 8. Adjoining the neutral element to a compact quantum group Let G be a compact quantum group. We may view C(G) as embedded into B(H) for some Hilbert space H, so that the inclusion Pol(G) ֒→ C(G) becomes a representation of the ∗-algebra Pol(G), say π : Pol(G) → B(H). Consider now the representation (8.1) (8.2) Therefore Since the norm defined by π is a quantum group norm, we have (cid:13)(cid:13)(π ⊗ π)∆(a)(cid:13)(cid:13) ≤ (cid:13)(cid:13)π(a)(cid:13)(cid:13) for all a ∈ Pol(G). It follows that (cid:13)(cid:13)(eπ ⊗eπ)∆(a)(cid:13)(cid:13) = max(cid:8)(cid:13)(cid:13)(π ⊗ π)∆(a)(cid:13)(cid:13),(cid:13)(cid:13)π(a)(cid:13)(cid:13),(cid:12)(cid:12)ε(a)(cid:12)(cid:12)(cid:9). (cid:13)(cid:13)(eπ ⊗eπ)∆(a)(cid:13)(cid:13) = max(cid:8)(cid:13)(cid:13)π(a)(cid:13)(cid:13),(cid:12)(cid:12)ε(a)(cid:12)(cid:12)(cid:9) =(cid:13)(cid:13)eπ(a)(cid:13)(cid:13). (8.3) Let C(Geπ) be the completion of Pol(G) with respect to the norm k · keπ. Then (8.3) shows that ∆ : Pol(G) → Pol(G) ⊗alg Pol(G) extends to an isometry C(Geπ) → C(Geπ) ⊗ C(Geπ) (minimal -- spatial -- tensor product). (cid:3) Definition 8.2. Let G be a compact quantum group. The compact quantum group obtained by (8.1). Proposition 8.3. Assume that G does not admit a continuous co-unit. Then there exists a central the construction described in the above proposition will be denoted by eG and called the quantum group G with neutral element adjoined. Thus, by definition C(eG) = C(Geπ), where eπ is defined by projection p in C(eG) such that Proof. As before we write π for the representation Pol(G) ֒→ C(G) ⊂ B(H) for some Hilbert space H and eπ for the direct sum of π and ε. Denote by k · kπ and k · keπ the associated C∗-norms on Pol(G). Since ε is unbounded on (Pol(G),k·kπ), for each n ∈  there exists an ∈ Pol(G) such that kankπ = 1 and (cid:12)(cid:12)ε(an)(cid:12)(cid:12) > n. Let bn = 1 0, while ε(bn) = 1 The completion C(eG) of Pol(G) in k·keπ is isomorphic to the closure ofeπ(cid:0)Pol(G)(cid:1) inside B(H⊕C). Note that the sequence (cid:0)eπ(bn)(cid:1)n∈ C(eG) ∼= Cp ⊕ C(G). ε(an) an. Clearly (cid:13)(cid:13)π(bn)(cid:13)(cid:13) = kbnkπ −−−−→n→∞ eπ(bn) =(cid:20)π(bn) converges in B(H ⊕ C), since ε(bn)(cid:21) for all n. 0 0 16 DAVID KYED AND PIOTR M. SO LTAN and clearly It is now clear that p commutes with all elements of the form p = lim 0 1(cid:21) . 0 n→∞eπ(bn) =(cid:20)0 (cid:20)π(a) ε(a)(cid:21) 0 0 (a ∈ Pol(G)) and that we have the isomorphism C(eG) ∼= Cp ⊕ C(G). Remark 8.4. (cid:3) (1) If G is a compact quantum group with continuous counit (e.g. G might be co-amenable, [23]). cf. [4]) then we have G = eG because k · keπ is equal to the original norm on C(G). (2) The quantum group eG has, by construction, continuous counit. Moreover the comultipli- cation on C(eG) obtained by extending that on Pol(G) is injective (cf. the discussion in (3) Suppose that G 6= eG = Gmax. Note that it is obvious from the proof of Proposition 8.3 that the value of the counit (extended from Pol(G) to C(eG)) on the projection p is 1. Therefore by Propositions 8.3 and 6.1 we have that bG has property (T). We may therefore take for G the reduced (minimal) version of a non-co-amenable com- pact quantum group whose dual does not have property (T). Then in the sense that the canonical morphisms C(Gmax) → C(eG) → C(G) are not isomorphisms. (4) The situation when G 6= eG = Gmax is also very interesting. We give and example of this phenomenon in Section 9. G 6= eG 6= Gmax 9. Exotic quantum group norms the representation of C(Gmax) defined as the direct sum of all its infinite dimensional irreducible As before we consider a discrete quantum group bG. In this section we will assume that bG is infinite (dim c0(bG) = ∞) and that bG has property (T). Throughout this section we let Π be representations. The corresponding corepresentation of bG will be denoted by V: V = (id ⊗ Π)w. Proposition 9.1. We have λ 4 Π. Proof. By Corollary 5.4 λ does not weakly contain any finite dimensional representation. This means that the support of λ in Spec(cid:0)C(Gmax)(cid:1) does not have any finite dimensional represen- tations in its closure. Therefore the support of λ is contained in the set of infinite dimensional representations, i.e. the support of Π. Hence λ 4 Π. (cid:3) The above result yields the following corollary: Corollary 9.2. The seminorm k·kΠ defined on Pol(G) by Π is a norm and C(Gmin) is a quotient of the completion C(GΠ) of Pol(G) in the norm k · kΠ. Theorem 9.3. k · kΠ is a quantum group norm on Pol(G). Proof. We will show that (Π⊗ Π)◦∆ 4 Π. We will do this using the language of corepresentations of bG instead of that of representations of C(Gmax). ⊤ V does not weakly contain a finite dimensional corepre- sentation. Assume the contrary and let U be a finite dimensional corepresentation of bG such that U 4 V direct sums of irreducible ones) we have U ≤ V ⊤ U c ≤ (V U  ⊤ V. By Remark 5.3 (and the fact that finite dimensional corepresentations decompose into Clearly it is enough to show that V ⊤ V. Therefore, by Lemma 2.1 ⊤ V)c. ⊤ V) ⊤ (V PROPERTY (T) AND EXOTIC QUANTUM GROUP NORMS 17 But U  ⊤ U c contains the trivial corepresentation of bG (Remark 2.2), so ⊤ V) ⊤ (V ⊤ (V (V ⊤ V)c ≈ V ⊤ V c  ⊤ V c) 3.4(3)) we can use Theorem 2.6(2) to see that V must then contain a finite dimensional corepre- sentation. This is a contradiction with the construction of V. (cid:3) (cf. [24, Formula (3.7)]) contains the trivial corepresentation. Since bG is unimodular (Theorem For our infinite, discrete property (T) quantum groupbG we now obtain the following two results Corollary 9.4. The compact quantum group GΠ obtained via completion of Pol(G) in the norm k · kΠ does not admit a continuous counit. This is evident, because our assumptions on bG imply that ε 64 Π. The same technique as the one used in the proof of Theorem 9.3 gives an answer to a question asked in [4, End of Section 3], namely if every C∗-norm on Pol(G) defined by a representation which weakly contains the regular one is a quantum group norm. Corollary 9.5. Assume that bG has a non-trivial finite-dimensional corepresentation U . Let U0 be an irreducible subrepresentation of U and let π0 be the corresponding representation of C(Gmax): U0 = (id ⊗ π0)w. Then the representation defined as the direct sum of all irreducible representations of C(Gmax) except the trivial one weakly contains the regular representation and the associated norm is not a quantum group norm. Let us now discuss one special case when the compact quantum group GΠ has quite unexpected properties. Let us consider the cocommutative example with bG = Γ, a discrete Kazhdan group which is minimally almost periodic, i.e. it has no non-trivial finite dimensional irreducible repre- sentations.1 Then it is easily seen that Π ⊕ ε is weakly equivalent to the universal representation of Pol(G) = C[Γ]. In other words, (9.1) Remark 9.6. fGΠ = Gmax. (1) Let us note that if bG is an infinite discrete property (T) group with only one irreducible finite dimensional corepresentation, namely the trivial one, then the minimal projection p ∈ C(Gmax) associated to this representation has a very peculiar property. Let ∆Π be the comultiplication on C(GΠ) and let ρ : C(Gmax) → C(GΠ) be the quotient map. Of course we have ρ(p) = 0. Note further that ρ is faithful on Pol(G) ⊂ C(Gmax) (e.g. because the regular representation λ factors through ρ). If we denote by ∆max the comultiplication on C(Gmax) then we have so that (ρ ⊗ ρ)∆max(p) = 0. However, due to the decomposition (ρ ⊗ ρ)◦∆max = ∆Π◦ρ C(Gmax) ∼= C(GΠ) ⊕ Cp we clearly have ker ρ ⊗ ρ = (Cp ⊗ p) ⊕(cid:0)p ⊗ C(GΠ)(cid:1) ⊕(cid:0)C(GΠ) ⊗ p(cid:1) ⊂ C(Gmax) ⊗alg C(Gmax). This means that ∆max(p) ∈ C(Gmax) ⊗alg C(Gmax), but p 6∈ Pol(G), as ρ(p) = 0. This example provides an answer to the famous question whether any element of a C∗-algebra C(G) whose image under the coproduct is a finite sum of simple tensors must belong to Pol(G). The affirmative answer for compact quantum groups with faithful Haar measure 1 Examples of discrete property (T) groups which are minimally almost periodic have been constructed by Gromov in [15] (cf. [26, Theorem 3.4], more explicit examples have been constructed in [9]). The result of Gromov provides (uncountably many) pairwise non-isomorphic infinite discrete property (T) torsion groups. In particular, they cannot contain a non-abelian free subgroup, so by Tits' alternative ([12, Section 42]) they cannot be linear, i.e. subgroups of GL(N, ) for a field  of characteristic 0. Moreover, by [26, Lemma 3.5], these groups are simple. Since they are not linear, they must be minimally almost periodic. 18 DAVID KYED AND PIOTR M. SO LTAN was given by S.L. Woronowicz in [32, Theorem 2.6(2)]. Our example shows that this is not the case if the dual of G is minimally almost periodic with property (T). A crucial fact here is that there actually exists a comultiplication on C(GΠ) or, in other words, that finitely many irreducible finite dimensional corepresentations. k · kΠ is a quantum group norm. Note also, that a similar argument applies if bG has only (2) The reader will have noticed that in fact the situation that fGΠ = Gmax is equivalent to C(Gmax) having no irreducible finite dimensional representations except ε. In other words (9.1) holds if and only if bG is minimally almost periodic. The above example leads to an important question, namely whether we have Gmin = GΠ. It seems that this could actually be the case in some examples, but we have not been able to produce one (nor find it in literature). However, as the next proposition says, at least for the cocommutative examples, the case that Gmin 6= GΠ is rather common. Proposition 9.7. Let Γ be an infinite discrete group with Kazhdan's property (T) such that the regular representation of Γ is weakly equivalent to the sum of all infinite dimensional irreducible representations of Γ. Then any non-amenable subgroup of Γ must have finite index. In particular, Γ cannot be linear. Before proving this proposition let us state one lemma. Lemma 9.8. Let L be a subgroup of a discrete countable group G such that the permutation representation λG/L of G on ℓ2(G/L) is weakly contained the regular representation λG of G. Then L is amenable. Proof. The characteristic function of L is a positive definite function associated with the permuta- tion representation λG/L (consider the coefficient of λG/L arising from the vector in ℓ2(G/L) which is the delta-function in the point L of G/L). Therefore, since λG/L is weakly contained in λG, the characteristic function of L is a pointwise limit of positive definite functions with finite support. Restricting these functions to L yields a net of finitely supported positive definite functions on L approximating pointwise the constant function 1. This proves that L is amenable. (cid:3) Proof of Proposition 9.7. Let Λ be a non-amenable subgroup of Γ. Then, by Lemma 9.8, the permutation representation λΓ/Λ cannot be weakly contained in λΓ. By assumptions on Γ there must be a finite dimensional representation σ of Γ weakly contained in λΓ/Λ (there must be an irreducible representation σ weakly contained in λΓ/Λ and not in λΓ, but all infinite dimensional ones are weakly contained in λΓ). Consider now the tensor product of σ and its contragredient representation. This is weakly contained in the tensor product of λΓ/Λ with its contragredient which is equivalent to the tensor square of λΓ/Λ. Of course σ⊗σc contains the trivial representation, so the square of λΓ/Λ contains (strongly - by property (T)) the trivial representation. Now the tensor square of λΓ/Λ is equivalent to the permutation representation of Γ on ℓ2((Γ/Λ) × (Γ/Λ)) with diagonal action. If this representation has a fixed vector, then Γ must have a finite orbit for the diagonal action on (Γ/Λ)×(Γ/Λ). (If a group Γ acts on a set S and the associated permutation representation in ℓ2(S) has a non-zero fixed vector ξ, then expanding this vector in the canonical orthonormal basis and acting on it shows that the coefficients of ξ are constant along orbits - therefore there must be a finite orbit.) This means that Λ has finite index in Γ because if (γΛ, γ ′Λ) is an element of (Γ/Λ) × (Γ/Λ) which has finite orbit, then there are γ1, . . . , γn, γ ′ n ∈ Γ such that 1, . . . , γ ′ for all x ∈ Γ. But {xγΛ x ∈ Γ} is all of Γ/Λ, so Γ/Λ is contained in the union (xγΛ, xγ ′Λ) ∈(cid:8)(γ1Λ, γ ′ 1Λ), . . . , (γnΛ, γ ′ nΛ)(cid:9) γ1Λ ∪ ··· ∪ γnΛ. This establishes that any non-amenable subgroup of Γ has finite index. Since Γ has property (T) it is finitely generated, so if Γ furthermore were linear the Tits alternative ([12, Section 42]) implies that it is either virtually solvable (which is impossible because it is non-amenable) or contains a non-abelian free subgroup. It is easy to see that then Γ must also contain non-amenable subgroups of infinite index. (cid:3) PROPERTY (T) AND EXOTIC QUANTUM GROUP NORMS 19 It follows from Proposition 9.7 that if we take bG = Γ to be a linear infinite Kazhdan group, say Γ = SL(3, Z), admitting non-trivial finite-dimensional irreducible representations, then we have Acknowledgments. The authors are grateful to Vadim Alekseev, Narutaka Ozawa, Andrzej Zuk and, in particular, Alain Valette for helpful comments regarding the results in Section 9. The proof of Proposition 9.7 is due to Alain Valette. Gmin 6= GΠ 6= fGΠ 6= Gmax. References [1] C.A. Akemann & M.E. Walter: Unbounded negative definite functions. Canad. J. Math. 33 no. 4 (1981), 862 -- 871. [2] S. Baaj & G. Skandalis: Unitaries muliplicatifs et dualit´e pour les poiduits crois´es de C∗-alg´ebres. Ann. Sci- ent. ´Ec. Norm. Sup. 4e s´erie, t. 26 (1993), 425 -- 488. [3] S. Baaj, G. Skandalis & S. Vaes: Non-semi-regular quantum groups coming from number theory. Com- mun. Math. Phys. 235 no. 1 (2003), 139 -- 167. [4] E. B´edos, G.J. Murphy & L. Tuset: Co-amenability of compact quantum groups. J. Geom. Phys. 40 (2001), 129 -- 153. [5] E. B´edos, R. Conti, L. Tuset: On amenability and co-amenability of algebraic quantum groups and their corepresentations. Canad. J. Math. 57 no. 1 (2005), 17 -- 60. [6] B. Bekka: A Property (T) for C∗-algebras. Bull. London Math. Soc. 38 (2006), 857-867. [7] B. Bekka, P. de la Harpe & A. Valette: Kazhdan's property (T). New Mathematical Monographs 11. Cambridge University Press, Cambridge, 2008. [8] N.P. Brown: Kazhdan's property T and C∗-algebras. J. Funct. Anal. 240 no. 1 (2006), 290 -- 296. [9] P.-E. Caprace & B. R´emy: Simplicit´e abstraite des groupes de Kac-Moody non affines. C. R. Math. Acad. Sci. Paris 342 no. 8 (2006), 539 -- 544. [10] A. Connes & V.F.R. Jones: Property T for von Neumann algebras. Bull. London Math. Soc. 17 no. 1 (1985), 57 -- 62. [11] M. Daws: Operator biprojectivity of compact quantum groups. Proc. Amer. Math. Soc. 138 (2010), 1349 -- 1359. [12] P. de la Harpe: Topics in geometric group theory. University of Chicago Press 2000. [13] J. Dixmier: C∗-algebras. North-Holland Publishing Co., Amsterdam, 1977. Translated from the French by Francis Jellett, North-Holland Mathematical Library, Vol. 15. [14] P. Fima: Kazhdan's property (T) for discrete quantum groups. arXiv:0812.0665v1 [math.OA]. [15] M. Gromov: Hyperbolic groups. In Essays in group theory, Math. Sci. Res. Inst. Publ. 8, 1987, pp. 75 -- 263. [16] J. Kustermans & S. Vaes: The operator algebra approach to quantum groups. Proc. Nat. Acad. Sci. USA 97 no. 2 (2000), 547 -- 552. [17] J. Kustermans & S. Vaes: Locally compact quantum groups in the von Neumann algebraic setting. Math. Scand. 92 no. 1 (2003), 68 -- 92. [18] D. Kyed: A Delorme-Guichardet theorem for quantum groups. arXiv:1003.5181v1 [math.OA]. [19] S. Petrescu & M. Joita: Property (T) for Kac algebras. Rev. Roumaine Math. Pures Appl. 37 no. 2 (1992), 163 -- 178. [20] P. Podle´s: Symmetries of quantum spaces. Subgroups and quotient spaces of quantum SU (2) and SO(3) groups. Commun. Math. Phys. 170 (1995), 1 -- 20. [21] P. Podle´s & S.L. Woronowicz: Quantum deformation of Lorentz group. Comm. Math. Phys. 130 no. 2 (1990), 381 -- 431. [22] P.M. So ltan: Quantum Bohr compactification. Ill. J. Math. 49 no. 4 (2005), 1245 -- 1270. [23] P.M. So ltan: On actions of compact quantum groups. arXiv:1003.5526v2 [math.OA]. To appear in Illinois Journal of Mathematics. [24] P.M. So ltan & S.L. Woronowicz: From multiplicative unitaries to quantum groups II. J. Funct. Anal. 252 (2007), 42 -- 67. [25] A. Valette: Minimal projections, integrable representations and property (T). Arch. Math. 43 (1984), 397 -- 406. [26] A. Valette: Old and new about Kazhdan's property (T). In Representations of Lie groups and quantum groups, V. Baldoni & M.A. Picardello editors. Pitman Research notes in Mathematics Series 311 1993, pp. 271 -- 333. [27] A. Van Daele: Discrete quantum groups. J. Algebra 180 no. 2 (1996), 431 -- 444. [28] A. Van Daele & S. Wang: Universal quantum groups. Int. J. Math. 7 no. 2 (1996), 747 -- 764. [29] S.L. Woronowicz: Twisted SU(2) group. An example of noncommutative differential calculus. Publ. RIMS, Kyoto University 23 (1987), 117 -- 181. [30] S.L. Woronowicz: Compact matrix pseudogroups. Comm. Math. Phys. 111 (1987), 613 -- 665. [31] S.L. Woronowicz: Unbounded elements affiliated with C∗-algebras and non-compact quantum groups. Com- mun. Math. Phys. 136 (1991), 399 -- 432. 20 DAVID KYED AND PIOTR M. SO LTAN [32] S.L. Woronowicz: Compact quantum groups. In Sym´etries Quantiques, les Houches, Session LXIV 1995, Elsevier 1998, pp. 845 -- 884. [33] S.L. Woronowicz: From multiplicative unitaries to quantum groups. Int. J. Math. 7 no. 1 (1996), 127 -- 149. Mathematisches Institut, Georg-August-Universitat Gottingen E-mail address: [email protected] URL: http://www.uni-math.gwdg.de/kyed/ Institute of Mathematics, Polish Academy of Sciences and Department of Mathematical Methods in Physics, Faculty of Physics, University of Warsaw E-mail address: [email protected] URL: http://www.fuw.edu.pl/~psoltan/en/
1006.2979
2
1006
2011-05-18T12:28:17
Isomorphisms and Fusion Rules of Orthogonal Free Quantum Groups and their Complexifications
[ "math.OA", "math.QA" ]
We show that all orthogonal free quantum groups are isomorphic to variants of the free orthogonal Wang algebra, the hyperoctahedral quantum group or the quantum permutation group. We also obtain a description of their free complexification. In particular we complete the calculation of fusion rules of all orthogonal free quantum groups and their free complexifications.
math.OA
math
Isomorphisms and Fusion Rules of Orthogonal Free Quantum Groups and their Free Complexifications by Sven Raum(1,2) Abstract We show that all orthogonal free quantum groups are isomorphic to variants of the free orthogonal Wang algebra, the hyperoctahedral quantum group or the quantum permutation group. We also obtain a description of their free complexification. In particular we complete the calculation of fusion rules of all orthogonal free quantum groups and their free complexifications. 1 Introduction One problem in the theory of compact quantum groups is to find examples whose invariants can be calculated. The fusion rules of a compact quantum group are one of these invariants. Fusion rules give a complete description of equivalence classes of irreducible corepresentations and a de- composition of the tensor product of two of them into irreducible corepresentations. One approach to this problem is given by 'free quantum groups' as defined in [5]. These are orthogonal quantum groups, i.e. subgroups of the free orthogonal Wang algebra, whose intertwiners can be described by non-crossing partitions. Given natural numbers k and l the set Part(k, l) denotes the set of all partitions on two rows with k and l points, respectively. That is, an element P ∈ Part(k, l) is a partition of the disjoint union {1, ..., k}⊔{1, ..., l}. Alternatively it can be described by a diagram connecting the k points in the upper row and the l points in the lower row according to the partition ⋅ ... ⋅ ⋅ ⎧⎪⎪⎨⎪⎪⎩ ⎫⎪⎪⎬⎪⎪⎭ P⋅ ⋅ ... ⋅ of {1, ..., k}⊔{1, ..., l}. P is called non-crossing if it can be represented by a diagram with no lines crossing. The set of all non-crossing partitions on k and l points is denoted by NC(k, l). Let n, k, l ∈  and let (ei) be the standard basis of C n. Let i = (i1, ..., ik) ∈ {1, ..., n}k and j = (j1, ..., jl) ∈ {1, ..., n}l be multi indices and P ∈ Part(k, l). We set P(i, j) = 1 if and only if the diagram and those of j in the lower row. If P connects different numbers set P(i, j) = 0. n)⊗l by n)⊗k to (C Using this notation, a partition P ∈ Part(k, l) defines a linear map TP from (C P(i1, ..., ik ; j1, ..., jl)⋅ ej1 ⊗ ...⊗ ejl . diagram P joins only equal numbers after writing the entries of i in the upper row of the above TP(ei1 ⊗ ...⊗ eik) = Q j1,...,jl 1Research partially supported by Marie Curie Research Training Network Non-Commutative Geometry MRTN- CT-2006-031962 and by K.U.Leuven BOF research grant OT/08/032 2Department of Mathematics; K.U.Leuven; Celestijnenlaan 200B; B -- 3001 Leuven (Belgium). E-mail: [email protected] 1 A subspace of Hom((C by a family (TP) where P runs through some subset of Part(k, l). n)⊗l) is by definition spanned by partitions if it is linearly generated n)⊗k,(C In [7] the free unitary Wang algebra Au(n)∶= C ∗(uij , 1 ≤ i, j ≤ n(uij)ij ,(u∗ ij)ij are unitary) and the free orthogonal Wang algebra Ao(n)∶= C ∗(uij , 1 ≤ i, j ≤ n(uij)ij =(u∗ ij)ij is unitary) were introduced. Moreover in [8] the quantum permutation group As(n)∶= C ∗⎛⎜⎝uij , 1 ≤ i, j ≤ n RRRRRRRRRRRRRR (uij) =(u∗ ij) is unitary and uij are partial isometries summing up to one in every row and every column ⎞⎟⎠ was defined. Note that "are partial isometries" can be replaced by "are projections". The three last named algebras are compact matrix quantum groups in the sense of Woronowicz [9]. The following class of quantum groups will be of interest in this paper. Definition 1.1 Let (A, U) be a compact matrix quantum group. Then it is called free if • The morphism (Au(n), Uu) →(As(n), Us) mapping the entries of Uu to those of Us factorizes through (A, U). • The intertwiner spaces Hom(U i1 ⊠ ⋯ ⊠ U ik , U j1 ⊠ ⋯ ⊠ U jl), iα, jβ ∈ {1, } are spanned by ij) is the conjugate corepresentation of U and ⊠ denotes the tensor partitions, where U = (u∗ If the first condition is strengthened by requiring that the morphism (Ao(n), Uo) → (As(n), Us) factors through (A, U), then A it is called orthogonal free. product of corepresentations. In [5] the following classification was achieved. Theorem 1.2 There are exactly six orthogonal free quantum groups. Namely (i) The free orthogonal Wang algebra. (ii) The quantum permutation group. (iii) The hyperoctahedral quantum group Ah(n)∶= C ∗uij, 1 ≤ i, j ≤ n (uij) =(u∗ ij) is unitary and uij are partial isometries  . (iv) The bistochastic quantum group Ab(n)∶= C ∗⎛⎜⎝uij, 1 ≤ i, j ≤ n RRRRRRRRRRRRRR 2 (uij) =(u∗ ij) is unitary and uij sum up to one in every row and every column ⎞⎟⎠ . (v) The symmetrized bistochastic quantum group (vi) The symmetrized quantum permutation group RRRRRRRRRRRRRR Ab'(n)∶= C ∗⎛⎜⎝uij, 1 ≤ i, j ≤ n RRRRRRRRRRRRRRRRRRR As'(n)∶= C ∗⎛⎜⎜⎜⎝ uij, 1 ≤ i, j ≤ n (uij) =(u∗ ij) is unitary and uij sum up to the same element in every row and every column (uij) =(u∗ ij) is unitary and uij are partial isometries summing up to the same element in every row and every column ⎞⎟⎠ . ⎞⎟⎟⎟⎠ . and Vergnioux given in [6]. The fusion rules of (1) were calculated in [1], those of (2) in [3] and those of (3) in [6]. We show that the remaining examples are slight modifications of Ao(n) and As(n). In particular we can derive their fusion rules and find that Ab'(n) and As'(n) are counterexamples to a conjecture by Banica In [4] the free complexification of orthogonal free quantum groups was considered. If (A, U) is a orthogonal free quantum group, then its free complexification (A, U) is by definition the sub-C*- algebra of the free product A∗ C(S 1) generated by the entries of U ∶= U ⋅ idS1 = (uij ⋅ idS1). Here idS1 denotes the canonical generator of C(S1). As Banica shows in [4] the intertwiners between tensor products of the fundamental corepresentation and its conjugate can be described by the intertwiners of the orthogonal free quantum group it comes from. With additional requirements we can calculate the fusion rules of the free complexification from the fusion rules of the original orthogonal free quantum group. These additional requirements are fulfilled by Ao(n) and Ah(n), which gives the fusion rules of Ak(n) = Ah(n). Those of Au(n) = Ao(n) are known from [2]. From [4] we know that Ab(n) = Ab'(n) and As(n) = As'(n). We denote Ab(n) =∶ Ac(n) and As(n) =∶ Ap(n). They can be decomposed and described in terms of Ao(n) and As(n) again. Acknowledgment: I want to thank both Thomas Timmermann for suggesting to work on fusion rules of free quantum groups and Stefaan Vaes for helpful discussions about this article, especially on the last section. Moreover, I want to thank the referee for helpful comments. 2 Preliminaries We will mainly work with compact matrix quantum groups as defined by Worono-wicz in [9]. If A is a *-algebra and U ∈ Mn(A) we denote by U the matrix whose entries are conjugated, i.e. U ij =(Uij)∗. A pair (A, U) of a C*-algebra A and a unitary U ∈ Mn(A) is called a compact matrix quantum group if • A is generated by the entries of U , • there is a *-homomorphism ∆∶ A → A⊗min A mapping uij to ∑k uik ⊗ ukj, • the matrix U is invertible. A morphism of compact matrix quantum groups (A, U) φ → (B, V) is a *-homo-morphism A → B such that φ(uij) = vij where U and V must have the same size. There is at most one morphism from 3 one quantum group to another. If there is a morphism (A, U) →(B, V) then we say that (B, V) is a quantum subgroup of (A, U). a *-homomorphism ∆∶ A → A⊗min A such that Every compact matrix quantum group is also a compact quantum group, i.e. a C*-algebra A with morphism of compact quantum groups. • (∆⊗ id)○ ∆ =(id⊗ ∆)○ ∆, • span(A⊗ 1)∆(A) = span(1⊗ A)∆(A) = A⊗ A. A morphism of compact quantum groups (A, ∆A) φ → (B, ∆B) is a *-homomorphism from A to B such that ∆B ○ φ = (φ⊗ φ)○ ∆A. Every morphism of compact matrix quantum groups is also a We will also refer to a quantum group (A, U) or (A, ∆) as A. If (A, ∆A) and (B, ∆B) are quantum groups, then we denote by (A, ∆A)⊗(B, ∆B) the direct sum of quantum groups and by (A, ∆A)∗(B, ∆B) their free product. We will also write A⊗ B and A∗ B. A unitary corepresentation matrix of (A, ∆) is a unitary matrix V ∈ Mm(A) such that ∆(vij) = ∑k vik ⊗ vkj. In particular a one dimensional corepresentation matrix is just a unitary group-like element of A. 3 Free fusion rings In this section we will introduce free fusion rings and prove that they are free unital rings. We will use the following notation for words in free monoids. Let M = mon(S) be a free monoid over a set S. If w ∈ M is a word of length k, then we write wi for the i-th letter of w, 1 ≤ i ≤ k. Hence w = w1w2w3 . . . wk−1wk. Definition 3.1 A free fusion monoid is a free monoid M = mon(S) over a set S with a fusion ⋅ ∶ S × S → S ∪{∅} and a conjugation (i) The fusion ⋅ is associative, where we make the convention that s⋅ s′ is the empty set if one of ∶ S Ð→ S. They must satisfy the following conditions. s, s′ is the empty set. (ii) The conjugation is involutive, i.e. s = s for all s ∈ S. (iii) Fusion and conjugation are compatible in the following sense. For all s1, s2, s3 ∈ S we have s1⋅ s2 = s3 ⇔ s2⋅ s3 = s1 A set S equipped with fusion and conjugation is called a fusion set. The fusion and conjugation of S induce a fusion and a conjugation on M via • w⋅w′ = w1 . . . wk−1(wk⋅w′ ∅. 1)w′ 2 . . . w′ l where this fusion is the empty set by convention if wk⋅w′ 1 = • w = wk . . . w1 4 If M = mon(S) is a free fusion monoid, we can turn ZM into an associative ring by aw ⋅ aw′ = Q (axz + ax⋅z). w=xy w′=yz Here w, w′ are words in M , aw and aw′ are the corresponding elements in ZM , xy, yz and xz denote the concatenation of words and the second term in the sum is by convention always ignored if the fusion x⋅ z is empty. Actually condition (3) of the previous definition is a necessary condition for making ZM associative, as it can be seen by considering (as1 ⋅ as2)⋅ as3 = as1 ⋅(as2 ⋅ as3) for s1, s2, s3 ∈ S. A *-ring isomorphic to ZM for some fusion monoid M is called a free fusion ring. From the point of view of rings, free fusion rings are very easy. Actually they are free. The proof of the following lemma was already given in [6] in some special cases. Lemma 3.2 A free fusion ring over a fusion set S is the free unital ring over as, s ∈ S. There are coefficients C w Proof Let ZM be the fusion ring over a fusion set S. It suffices to show that ZM is a free Z-module with the basis as1⋯ask with k ∈  and s1, . . . , sk ∈ S. So it suffices to express the elements of the Z-basis aw, w ∈ M as Z-linear combinations of the elements as1⋯ask with k ∈  and s1, . . . , sk ∈ S and to show that {as1⋯askk ∈ , s1, . . . , sk ∈ S} is Z-linearly independent. aw, where w is s1...sk ∈ Z such that as1⋯ask = as1...sk + ∑w<k C w the length of the word w ∈ M . This shows that {as1⋯askk ∈ , s1, . . . , sk ∈ S} is linearly in- s1...sl ∈ Z such that as1...sk = aw1⋯aww. This shows that all aw, w ∈ M are linear combinations of as1⋯ask as1⋯ask+∑w<k Dw ◻ dependent. Moreover, by induction on k there are coefficients Dw with k ∈  and s1, . . . , sk ∈ S. s1...sk s1...sk Remark 3.3 Free fusion rings can be used to describe fusion rules very shortly and there is hope to use free fusion rings as a starting point for proofs of several properties of quantum groups. See section 10 of [6] for a comment on these possibilities. However in order to justify the concept of free fusion rings intrinsically it would be good to answer the following question affirmatively. Is every fusion ring of a compact quantum group that is free as a unital ring a free fusion ring? 4 Some isomorphisms of combinatorial quantum groups In this section we will consider combinatorial quantum groups A∗(n) for ∗ ∈ {b, b′, s′, c, p}. They are free products or direct sums of known quantum groups. For ∗ ∈ {b′, s′, c, p} it turns out that their fusion rings are not free. Theorem 4.1 We have the following isomorphisms of compact quantum groups (not necessarily preserving the fundamental corepresentation). (i) Ab(n) is isomorphic to Ao(n− 1). (ii) As'(n) is isomorphic to the direct sum As(n)⊗ C ∗(Z~2Z). (iii) Ab'(n) is isomorphic to the free product Ab(n)∗ C ∗(Z~2Z). (iv) Ap(n) is isomorphic to the free product As(n)∗ C(S1). (v) Ac(n) is isomorphic to the free product Ab(n)∗ C(S1). 5 The key observation for the rest of 4.1 is the following lemma. is bistochastic if and only if T tU T is of block form with 1 in the upper left corner and an orthogonal U = U unitary, where A is any unital C ∗-algebra. Then U is bistochastic if and only if the vector Remark 4.2 Note that in the case n ≤ 3 we have the isomorphisms As(n) ≅ C(Sn) and Ao(1) ≅ C({−1, 1}). So the given descriptions can be further simplified. Theorem 4.1(1) is proven by the following remark. Let U ∈ Mn(A) be an orthogonal matrix, i.e. (1, 1, . . . , 1)t is a right eigenvector and (1, 1, . . . , 1) is a left eigenvector of U . If T ∈ Mn(C) denotes any orthogonal matrix such that T(1, 0, . . . , 0)t =(1~√n, . . . , 1~√n)t , then an orthogonal matrix U (n− 1)×(n− 1) matrix in the lower right corner. Lemma 4.3 Let ∗ ∈ {b′, s′, c, p}. The fundamental corepresentation of A∗(n) contains a one di- mensional non-trivial corepresentation Uz which fulfils Uz ⊠ Uz ≃ 1. If ∗ ∈ {b′, s′} then Uz ≃ Uz. Proof Consider ∗ = b′, s′ first. The element z = ∑i uij is easily seen to be a unitary group-like element, so it corresponds to a one dimensional unitary corepresentation of A∗(n). Consider the group Sn ⊕ Z~2Z ⊂ Un as permutation matrices with entries +1 and −1. Let USn⊕Z~2Z be the canonical fundamental corepresentation of C(Sn ⊕ Z~2Z). Then the image of z under the map (A∗(n), U∗) →(C(Sn⊕ Z~2Z), USn⊕Z~2Z) is −1, so z is non-trivial. For ∗ = p, c consider z ∶= idS1 as coming from the copy of C(S1). This copy is contained in A∗(n), since the trivial corepresentation is contained in the fundamental corepresentation of Ab(n) and As(n). Using the relations of A∗(n) we can check the rest of the claim by simple calculations. ◻ Remark 4.4 The last lemma shows, that the fusion rules of neither of the quantum groups A∗(n) for ∈ {b′, s′, c, p} can be described by a free fusion ring. Actually in a free fusion ring any element a ≠ 1 satisfies a⋅ a∗ ≠ 1. This gives two counterexamples to the conjecture that for n ≥ 4 the fusion rules of all free orthogonal quantum groups can be described by a free fusion ring, which was stated in [6]. as a sub quantum group cannot be the sum of more than two irreducible corepresentations. Remark 4.5 The fundamental corepresentation of any matrix quantum group that has (As(n), Us) particular the last lemma already gives a decomposition U ≃ Uz ⊞ V with Uz non-trivial and one dimensional and V irreducible, where U is the fundamental corepresentation of A∗(n). Proof (Proof of Theorem 4.1) The isomorphism of (2) is given by As(n)⊗C ∗(Z~2Z) → As'(n)∶ ↦ z. This map exists since z is central in As'(n) as an easy calculation shows. ij⊗ 1 ↦ us′ ij⋅ z, 1⊗ u1 us The inverse map is given by In As'(n) → As(n)⊗ C ∗(Z~2Z)∶ us' ij → us ij ⊗ u1. if and only if T tU T is a block matrix with a self-adjoint unitary in the upper left corner and an In order to prove (3) we use again an orthogonal matrix T ∈ Mn(C) such that T(1, 0, ..., 0)t = (1~√n, ..., 1~√n)t. Then a matrix U ∈ Mn(A) for some C ∗-algebra A satisfies the relations of Ub' orthogonal (n − 1) × (n − 1) matrix in the lower right corner. This proves Ab'(n) ≅ Ao(n − 1) ∗ C ∗(Z~2Z) ≅ Ab(n)∗ C ∗(Z~2Z). The isomorphism of (4) is given by As(n)∗ C(S1) → Ap(n)∶ us ij ↦ up ij ⋅ z∗, idS1 ↦ z. 6 The isomorphism of (5) is given by Ab(n)∗ C(S1) → Ac(n)∶ ub ij ↦ uc ij ⋅ z∗, idS1 ↦ z. All the isomorphisms respect the comultiplication, since z is group-like. Hence, they are isomor- phisms of quantum groups. ◻ 5 Fusion rules for free products and the quantum group Ak(n) In this section we describe the fusion rules of the free complexification Ak(n) ≅ Ah(n). Instead of referring to Ak(n) explicitly, we will work in a more general setting and deduce its fusion rules as a corollary. Roughly the main statement of this section is given by the following theorem. See theorem 5.5 for a precise statement. its fusion rules are free. Assume further that 1 ∉ U ⊠2k+1 for any k ∈ . Then the fusion rules of Theorem 5.1 Let (A, U) be an orthogonal compact matrix quantum group, i.e. U = U , such that (A, U) are free and can be described in terms of the fusion rules of (A, U). The following theorem is due to Wang [8]. be complete sets of representatives of irreducible corepresentations of A and B, respectively. Then Theorem 5.2 Let (A, ∆A) and (B, ∆B) be compact quantum groups. Let (U α)α∈A and (U β)β∈B the corepresentations (W γ1 ⊠ ⋯ ⊠ W γn) with n ∈ , all W γi in {U α α ∈ A} and {U β β ∈ B} product (A, ∆A)∗(B, ∆B). and neighbours not from the same set, form a complete set of irreducible representations of the free The following observation will be useful when studying the fusion rules of a free complexification. Remark 5.3 Let A∗ B be a free product of compact quantum groups with irreducible corepresenta- tions W γ1 ⊠⋯⊠ W γn and W δ1 ⊠⋯⊠ W δm as in the last theorem. Then (i) If W γn and W δ1 are not corepresentations of the same factor of the free product, then W γ1 ⊠ i=1 W ǫi+δW γn ,W δ1⋅ (ii) If W γn and W δ1 are corepresentations of the same factor and W γn⊠W δ1 = ∑k ⋯⊠ W γn ⊠ W δ1 ⊠⋯⊠ W δm is an irreducible corepresentation of A∗ B. 1 is the decomposition into irreducible corepresentations, then W γ1 ⊠⋯⊠ W γn ⊠ W δ1 ⊠⋯⊠ W δm = k Q i=1(W γ1 ⊠⋯⊠ W γn−1 ⊠ W ǫi ⊠ W δ2 ⊠⋯⊠ W δm) + δW γn ,W δ1 ⋅ W γ1 ⊠⋯⊠ W γn−1 ⊠ W δ2 ⊠⋯⊠ W δm and the first k summands of this decomposition are irreducible. fusion rules are described by a free fusion ring over the fusion set S. Assume further that 1 ∉ U ⊠2k+1 for any k ∈ . For the rest of this section fix an orthogonal compact matrix quantum group (A, U) such that its Note that the fusion ring of A is the fusion subring of Rep(A∗ C(S1)) that is generated by U ⊠ z, 7 where z denotes the identity on the circle. even (respectively Repirr We will construct the free complexification S of S and prove that the fusion rules of (A, U) are described by S. We begin by constructing S. Let Repirr odd) be the set of classes of irreducible corepresentations of A that appear as subrepresentations of an even (respectively odd) tensor power of U . We have Repirr Seven ⊂ S (resp. Sodd ⊂ S) be the set of elements corresponding to corepresentations from Repirr (resp. Repirr the first copy of Seven (resp. Sodd) by S(1) odd = ∅ due to Frobenius duality and the requirement 1 ∉ U 2k+1 for all k ∈ . Let odd). The set S is then by definition the disjoint union Seven⊔ Seven⊔ Sodd⊔ Sodd. Denote odd) and the second one by S(2) even ∩ Repirr even (resp. S(2) even (resp. S(1) odd). even What follows is motivated by the following point of view: odd as z∗⋅ s for s ∈ Sodd. even as a plain copy of those in Seven. The elements of S(2) odd as s⋅ z and Remark 5.4 We consider element of S(1) are of the form z∗ ⋅ s ⋅ z for some s ∈ Seven. Similarly we consider elements of S(1) elements of S(2) Define a conjugation on S by the conjugation on S leaving S(1) even and S(2) exchanging S(1) well defined. A fusion on S can be defined according to the following table. odd. Note that Seven = Seven and Sodd = Sodd, i.e. the conjugation on S is even globally invariant and odd and S(2) even ⋅ S(1) even S(2) even S(1) odd S(2) odd S(1) even even∪{∅} S(1) odd∪{∅} S(2) ∅ ∅ S(2) even ∅ ∅ S(2) S(2) even∪{∅} odd∪{∅} S(1) odd odd∪{∅} S(1) even ∪{∅} S(2) ∅ ∅ S(2) odd ∅ ∅ S(2) S(1) odd∪{∅} even∪{∅} Now we can state a precise version of 5.1. The row gives the element which is fused from the right with an element coming from the set indicated by the column. The fusion is empty if this is indicated by the table and is otherwise the usual fusion of two elements of S lying in the part of S indicated by the table. Note that this definition makes sense, since Seven ⋅ Seven, Sodd ⋅ Sodd ⊂ Seven ∪ {∅} and Seven ⋅ Sodd, Sodd ⋅ Seven ⊂ Sodd∪{∅}. It is easy to see that S with this structure is a fusion set. Theorem 5.5 Let (A, U) be an orthogonal compact matrix quantum group such that its fusion rules k ∈ . Then the fusion rules of (A, U) are given by the free complexification S of S. We construct a complete set of corepresentations of A. In order to do so we associate an irreducible corepresentations of (A, U) to any element of R∶= Repirr are described by a free fusion ring over the fusion set S. Assume further that 1 ∉ U ⊠2k+1 for any odd. We denote odd ). Let V be a irreducible corepresentation even. Then V and z∗ ⋅ V ⋅ z are corepresentations of A. Actually, if V is an irreducible subrepresentation of U ⊠2k then V is an irreducible subrepresentation of (U ⊠ U)⊠k and z∗ ⋅ V ⋅ z is an irreducible subrepresentation of (U ⊠ U)⊠k. We consider V as an element of Repirr,(1) z∗ ⋅ V ⋅ z as an element of Repirr,(2) with it corepresentations V ⋅ z ∈ Repirr,(1) . Note that elements s from S give and odd then we can associate and z∗⋅ V ∈ Repirr,(2) even . Similarly we see that if V ∈ Repirr the i-th copy of Repirr in Repirr even⊔ Repirr even⊔ Repirr odd⊔ Repirr odd) by Repirr,(i) even (Repirr (Repirr,(i) even even odd odd 8 corepresentations Us by this identification. Consider a word w = w1 . . . wk with letters in R. We say that w is reduced if in the sequence Uw1 , . . . , Uwn a z is never followed by z∗ and Ux is always followed by z or z∗. In formal terms: ∀1 ≤ i ≤ k − 1 ∶(wi ∈ Repirr,(1) (wi ∈ Repirr,(2) even ∪ Repirr,(2) even ∪ Repirr,(1) odd ⇒ wi+1 ∈ Repirr,(2) odd ⇒ wi+1 ∈ Repirr,(1) odd )∧ even ∪ Repirr,(2) odd ) even ∪ Repirr,(1) Any such reduced word w = w1 . . . wk gives rise to an irreducible corepresentation of A by Uw ∶= Uw1 ⊠ . . . ⊠ Uwk and different reduced words give rise to inequivalent corepresentations by 5.2. Since any iterated tensor product of U and U decomposes as a sum of irreducible corepresentations of the type Uw, where w is a reduced word with letters in R, any irreducible corepresentation of A is equivalent to some Uw. Definition 5.6 Consider now a word w = w1 . . . wk with letters in S. It is called connected if every z is followed by a z∗. Formally: ∀1 ≤ i ≤ k − 1 ∶(wi ∈ S(1) (wi ∈ S(2) even ∪ S(2) even ∪ S(1) odd ⇒ wi+1 ∈ S(1) odd ⇒ wi+1 ∈ S(2) odd)∧ even ∪ S(1) odd) even ∪ S(2) The following definition says how we can associate irreducible corepresentations of A to words with letters in S. Definition 5.7 If w is an arbitrary word with letters in S then it has a unique decomposition in R. We set Uw ∶= Uw′ w = x1 . . . xl into maximal connected words. This gives rise to a unique reduced word w′ with letters Next we have to do some preparations in order to prove theorem 5.5. Definition 5.8 Let x = x1 . . . xm be a word in S. Then xi is the letter in S corresponding to xi and x ∶= x1 x2 . . . xm. xy is a connected word. Remark 5.9 Note that if x is a connected word with letters in S then according to remark 5.4 it can be written as zi0 ⋅ x⋅ zi1, i0, i1 ∈{0, 1,−1} and we have Ux = zi0 ⊠ Ux⊠ zi1. Definition 5.10 Let x, y be connected words with letters in S. We say that (x, y) fits together if Lemma 5.11 Let x = x1 . . . xm and y = y1 . . . yn be connected words with letters in S such that (xm, y1) fits together. Write Ux = zi0 ⊠ Ux⊠ zi1 and Uy = zj0 ⊠ U y ⊠ zj1. Then Uab⊞ Ua⋅b. Ux⊠ Uy = zi0 ⊠⎛ Uab⊞ Ua⋅b⎞ ⎠⊠ zj1 = Q x=ac,y=cb ⎝ Q x=ac,y=cb We have to prove that for all x = ac, y = cb Proof Since (x, y) fits together, we have zi1 ⊠ zj0 = 1. So by remark 5.3 the first equation follows. (i) zi0 ⊠ Uab⊠ zj1 = Uab (ii) zi0 ⊠ Ua⋅b⊠ zj1 = Ua⋅b. 9 (1) follows from the way irreducible corepresentations are associated to connected words remarked in 5.9. In order to prove (1), note that ab is connected, since a, b are connected and (a, b) fits together. So For (2) note that, since (a, b) fits together, a⋅ b = ∅ if and only if a⋅ b = ∅. If a⋅ b ≠ ∅ then it is ◻ connected and (2) follows by remark 5.9 again. Now we can give the proof of Theorem 5.5 Proof (Proof of Theorem 5.5) Let x = x1 . . . xk and y = y1 . . . yl be words with letters in S. We have to show that Ux⊠ Uy = Q x=ac,y=cb Uab⊞ Ua⋅b Let x = u1 . . . um and y = v1 . . . vn be the decomposition in maximal connected words. We identify them with letters in R. Then Ux = zi0 ⊠ U u1 ⊠ zi1 ⊠ U u2 ⊠ zi2 ⊠⋯⊠ U um−1 ⊠ zim−1 ⊠ zim ⊠ U um ⊠ zim+1 ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ Uy = zj0 ⊠ U v1 ⊠ zj1 ⊠zj2 ⊠ U v2 ⊠ zj3 ⊠⋯⊠ zjn−1 ⊠ U vn−1 ⊠ zjn ⊠ U vn ⊠ zjn+1 ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ = Uum , = Uv1 with i1, ..., im−2, j3, ..., jn ∈{1,∗}, i0, im, j0, j2 ∈{0,∗} and im−1, im+1, j1, jn+1 ∈{0, 1}. We are going to consider the two cases (xk, y1) do or do not fit together. Assume that (xk, y1) do not fit together. This means zim+1 ⋅ zj0 ≠ 1. Then Ux⊠ Uy is irreducible by Theorem 5.2. Moreover, xy = u1 . . . umv1 . . . vn is a decomposition in maximal connected words. So Ux ⊠ Uy = Uxy. On the other hand (xk, y1) not fitting together implies xk ≠ y1 and xk⋅y1 =∅. So ∑x=ac,y=cb Uab⊞ Ua⋅b = Uxy. Assume now that (xk, y1) fits together. This means zim+1 ⋅ zj0 = 1. By Lemma 5.11 Uab⊞ Ua⋅b)⊠ zj2 ⊠ U v2 ⊠ zj3 ⊠⋯⊠ U vn ⊠ zjn+1 Ux⊠ Uy = zi0 ⊠ U u1 ⊠ zi1 ⊠⋯⊠ U um−1 ⊠ zim−1⊠ This completes the proof for the first case. um=ac,v1=cb ( Q = zi0 ⊠ U u1 ⊠ zi1 ⊠⋯⊠ U um−1 ⊠ zim−1⊠ (( zj2 ⊠ U v2 ⊠ zj3 ⊠⋯⊠ U vn ⊠ zjn+1 . um=ac,v1=cb,a≥1 or b≥1 Q Uab⊞ Ua⋅b)⊞ δum,v1 ⋅ 1)⊠ By applying the induction hypothesis to the term zi0 ⊠ U u1 ⊠⋯⊠ zim−1 ⊠ δum,v1 ⋅ 1⊠ zj2 ⊠ U v2 ⊠⋯⊠ zjn+1 = δum,v1 ⋅ Uu1u2...um−1 ⊠ Uv2v3...vn Uab⊞ Ua⋅b. Ux⊠ Uy = Q we obtain ◻ We are now going to deduce the fusion rules of Ak(n). The following result is proven in [6] and describes the fusion rules of Ah(n). x=ac,y=cb 10 Theorem 5.12 Let Sh ∶= {u, p} with fusion u⋅ u = p⋅ p = p, u⋅ p = p⋅ u = u and trivial conjugation. The fusion rules of (Ah(n), Uh) are given by the free fusion ring over Sh in such a way that Uu ≃ Uh and Up⊞ 1 ≃(u2 ij). Using this theorem we obtain the following corollary in the case A = Ak(n). Corollary 5.13 The irreducible corepresentations of Ak(n) are described by the fusion set Sk ∶= {u, v, p, q} with fusion given by v p u ⋅ u ∅ q p ∅ v u ∅ p ∅ ∅ v p ∅ u ∅ ∅ q q v q and conjugation u = v, p = p, q = q. The elements of Sk correspond to the following corepresentations. • The class of the fundamental corepresentation U is Uu. • The class of U is Uv. • The class of the corepresentation (u∗ • The class of the corepresentation (uij ⋅ u∗ ij ⋅ uij) is Up⊞ 1 ij) is Uq ⊞ 1 Proof We only have to prove the part about the concrete description of Uu, Uv, Up and Uq. The fact that Uu is the class of the fundamental corepresentation is obvious from the construction. Uv ≃ U follows directly. ij ⋅ uij) and (uij ⋅ u∗ It is easy to check that (u∗ ij) are corepresentation of Ak(n). We have the decomposition U ⊠ U ≃ Uuv ⊞ Up ⊞ 1. Moreover the construction in this section shows that Uuv is n2 − n dimensional and Up is n− 1 dimensional. Since (uij ⋅ u∗ ij) is non trivial, it suffices to give at least two linearly independent intertwiners from the n dimensional corepresentation (uij ⋅ u∗ ij) to n)⊗2 ∶ ei ↦ ∑j ej ⊗ ej. n)⊗2 ∶ ei ↦ ei⊗ ei and C U ⊠ U . Two such intertwiners are C ◻ The proof for (uij ⋅ u∗ ij) works similarly. n →(C n →(C References [1] T. Banica, The representation theory of the free O(n) compact quantum group. (Théorie des représentations du groupe quantique compact libre O(n).). C. R. Acad. Sci., (3) 322 (1996), [2] T. Banica, The free compact quantum group U(n). (Le groupe quantique compact libre U(n).). 241-244. Commun. Math. Phys., (1) 190 (1997), 143-172. [3] T. Banica, Symmetries of a generic coaction. Math. Ann., (4) 314 (1999), 763-780. [4] T. Banica, A note on free quantum groups. (Une note sur les groupes quantiques libres.). Ann. Math. Blaise Pascal, (2) 15 (2008), 135-146. 11 [5] T. Banica and R. Speicher, Liberation of orthogonal Lie groups. Adv. Math., (4) 222 (2009), 1461-1501. [6] T. Banica and R. Vergnioux, Fusion rules for quantum reflection groups. J. Noncommut. Geom., (3) 3 (2009), 327-359. [7] S. Wang, Free products of compact quantum groups. Commun. Math. Phys., (3) 167 (1995), 671-692. [8] S. Wang, Quantum symmetry groups of finite spaces. Commun. Math. Phys., (1) 195 (1998), 195-211. [9] S.L. Woronowicz, A remark on compact matrix quantum groups. Lett. Math. Phys., (1) 21 (1991), 35-39. 12
0805.1566
5
0805
2010-05-03T09:18:28
Bass-Serre rigidity results in von Neumann algebras
[ "math.OA" ]
We obtain new Bass-Serre type rigidity results for ${\rm II_1}$ equivalence relations and their von Neumann algebras, coming from free ergodic actions of free products of groups on the standard probability space. As an application, we show that any non-amenable factor arising as an amalgamated free product of von Neumann algebras $\mathcal{M}_1 \ast_B \mathcal{M}_2$ over an abelian von Neumann algebra $B$, is prime, i.e. cannot be written as a tensor product of diffuse factors. This gives, both in the type ${\rm II_1}$ and in the type ${\rm III}$ case, new examples of prime factors.
math.OA
math
BASS-SERRE RIGIDITY RESULTS IN VON NEUMANN ALGEBRAS IONUT CHIFAN AND CYRIL HOUDAYER Abstract. We obtain new Bass-Serre type rigidity results for II1 equiv- alence relations and their von Neumann algebras, coming from free er- godic actions of free products of groups on the standard probability space. As an application, we show that any non-amenable factor arising as an amalgamated free product of von Neumann algebras M1 ∗B M2 over an abelian von Neumann algebra B, is prime, i.e. cannot be written as a tensor product of diffuse factors. This gives, both in the type II1 and in the type III case, new examples of prime factors. 1. Introduction We prove in this paper new rigidity results for amalgamated free products (hereafter abbreviated AFP) M = M1 ∗B M2 of semifinite von Neumann algebras over a common amenable von Neumann subalgebra. In the spirit of [21], these results can be viewed as von Neumann algebras analogs of "subgroup theorems" and "isomorphism theorems" for AFP of groups in Bass-Serre theory. Our main "subalgebra theorem" (see Theorem 1.1 below) shows that for any subalgebra Q ⊂ M with no amenable direct summand (e.g. Q non-amenable subfactor), the relative commutant P = Q′ ∩ M can be conjugated by an inner automorphism of M into either M1 or M2. This "subalgebra theorem" allows us to classify large classes of AFP fac- tors as well as to prove structural properties for these algebras. For instance, we prove that any non-amenable factor M = M1 ∗B M2 arising as an AFP over an abelian subalgebra cannot be decomposed into a tensor product of diffuse factors, i.e. M is a prime factor. This gives many new examples of prime factors, of type II1 and of type III. The typical "isomorphism theorem" we derive (see Theorem 1.5 below) shows that if θ : M ≃ N t is a ∗-isomorphism from an AFP II1 factor M = M1 ∗A ··· ∗A Mm onto the amplification by some t of an AFP II1 factor N = N1 ∗B ··· ∗B Nn, A, B abelian, 1 ≤ m, n ≤ ∞, with each Mi and each Nj containing large com- muting subalgebras, then m = n and θ(A ⊂ Mi) is unitarily conjugate to (B ⊂ Nj)t, for all i, after permutation of indices. Results of this type have been first obtained by Ozawa in his pioneering paper [24] for plain free products M = M1 ∗ M2 of semi-exact II1 factors. 2000 Mathematics Subject Classification. 46L10; 46L54; 46L55. Key words and phrases. Amalgamated free products; Deformation/rigidity; Spectral gap; Intertwining techniques. 1 2 I. CHIFAN AND C. HOUDAYER He showed that if Q ⊂ M is a non-amenable subfactor with P = Q′ ∩ M hyperfinite II1 factor, then P can be conjugated into M1 or M2 by a unitary in M . For the proof, he first used his C∗-algebraic techniques to get a "finite dimensional" P -Mi bimodule, and then Popa's intertwining subalgebras techniques to conclude. In their breakthrough paper [21], Ioana, Peterson & Popa showed that for B, M1, M2 arbitrary finite von Neumann algebras, any rigid subalgebra P (i.e. having relative property (T)) of M = M1 ∗B M2 can be conjugated into M1 or M2 by a unitary in M . Unlike Ozawa, this result used Popa's "deformation/rigidity" techniques in [32, 33] (with the relative property (T) being used for the rigidity) to first get a finite dimensional P -Mi bimodule, then the intertwining subalgebras techniques to conclude. Peterson, using his L2-derivations techniques [28], obtained similar Kurosh-type results for plain free products M = M1 ∗ M2, where M1, M2 are "L2-rigid" II1 factors. In [30], Popa showed that in many previous "deformation/rigidity" argu- ments (e.g. the W∗ strong rigidity of factors arising from Bernoulli actions of property (T) groups), the property (T) condition can be completely re- moved, using instead a "spectral gap" rigidity. This allowed treating many groups which do not have property (T), such as products of an arbitrary non-amenable and an arbitrary infinite group. Ozawa & Popa used this "spectral gap" rigidity arguments [26] to prove that the normalizer of any diffuse amenable subalgebra of the free group factors generates an amenable von Neumann algebra. We combine the deformation techniques of [21], and the spectral gap rigid- ity techniques of [30, 31] in order to prove our key technical theorem. This is the above mentioned "subalgebra theorem", of Bass-Serre type, which is behind all the results of this paper. For A, B ⊂ M finite von Neumann algebras, we recall from [32, 33] that the symbol A (cid:22)M B roughly means that a corner of A can be unitarily conjugated into a corner of B inside M . Theorem 1.1 (Key intertwining result). For i = 1, 2, let (Mi, τi) be a finite von Neumann algebra with a common amenable von Neumann subalgebra B ⊂ Mi, such that τ1B = τ2B. Let M = M1 ∗B M2 be the amalgamated free product. Let Q ⊂ M be a von Neumann subalgebra with no amenable direct summand (e.g. Q a non-amenable subfactor). Then there exists i = 1, 2 such that Q′ ∩ M (cid:22)M Mi. We briefly recall below the concepts that we will play against each other to prove Theorem 1.1: (1) The first ingredient we will use is the "malleable deformation" by automorphisms (αt, β) defined on M = M∗B (B⊗L(F2)), introduced in [21]. It represented one of the key tools that lead to the computa- tion of the symmetry groups of AFP of weakly rigid factors. It was shown in [30] that this deformation automatically features a certain "transversality property" (see Lemma 2.1 in [30]) which will be of essential use in our proof. BASS-SERRE RIGIDITY RESULTS IN VON NEUMANN ALGEBRAS 3 (2) The second ingredient we will use is the spectral gap rigidity principle discovered by Popa in [30, 31]. We prove that for any von Neumann subalgebra Q ⊂ M with no amenable direct summand, the action by conjugation Ad(U (Q)) y M has "spectral gap" relative to M : for any ε > 0, there exist δ > 0 and a finite "critical" subset F ⊂ U (Q) such that for any x ∈ ( M )1 (the unit ball of M ), if kuxu∗ − xk2 ≤ δ, ∀u ∈ F , then kx − EM (x)k2 ≤ ε. (3) A combination of (1) and (2) yields that for any Q ⊂ M with no amenable direct summand, the malleable deformation (αt) necessar- ily converges uniformly in k · k2 on the unit ball of Q′ ∩ M . Then using the intertwining techniques from [21], one can embed Q′ ∩ M into M1 or M2 inside M . We prove in fact a more general version of Theorem 1.1 (see Theorem 4.2), involving semifinite AFP. Indeed, given an AFP type III factor M, these techniques allow us to work with its core M ⋊σ R, which is of type II∞ [10, 39], rather than M itself. We then obtain the following theorem that generalizes many previous results on the indecomposability of factors as tensor products, i.e. primeness (see [17, 18, 24, 28, 37]), and moreover gives new examples of prime factors, of type II1 and of type III: Theorem 1.2 (Primeness for AFP over abelian). For i = 1, 2, let Mi be a von Neumann algebra. Let B ⊂ Mi be a common abelian von Neumann subalgebra, with B 6= Mi, such that there exists a faithful normal conditional expectation Ei : Mi → B. Denote by M = M1 ∗B M2 the amalgamated free product. If M is a non-amenable factor, then M is prime. Using some of Ueda's results on factoriality and non-amenability of plain free products and of amalgamated free products over a common Cartan subalgebra (see [41, 42, 43, 44]), we obtain the following corollaries: Corollary 1.3. For i = 1, 2, let (Mi, ϕi) be any von Neumann algebra endowed with a faithful normal state. Assume that the centralizer Mϕ1 is diffuse and M2 6= C. Then the free product (M, ϕ) = (M1, ϕ1) ∗ (M2, ϕ2) is a prime factor. Corollary 1.4. For i = 1, 2, let Mi be a non-type I factor, and B ⊂ Mi be a common Cartan subalgebra. Then the amalgamated free product M = M1 ∗B M2 is a prime factor. 1 In particular, let Γ = Γ1 ∗ Γ2 be a free product of countable infinite groups. Let σ : Γ y (X, µ) be a free action such that the measure µ is quasi-invariant under σ, and such that the restricted action σΓi is ergodic and non-transitive for i = 1, 2. Then the crossed product L∞(X, µ) ⋊ Γ is a prime factor. Theorem 1.1 allows us to obtain new W∗/OE Bass-Serre type rigidity results for actions of free products of groups. In order to state the main result, we first introduce the following notation. Fix integers m, n ≥ 1. For 4 I. CHIFAN AND C. HOUDAYER each i ∈ {1, . . . , m}, and j ∈ {1, . . . , n} let Γi = Gi × Hi Λj = G′ j × H ′ j j are not amenable be ICC (infinite conjugacy class) groups, such that Gi, G′ and Hi, H ′ j are infinite. Denote Γ = Γ1 ∗ ··· ∗ Γm and Λ = Λ1 ∗ ··· ∗ Λn. Let σ : Γ y (X, µ) be a free m.p. action of Γ on the probability space (X, µ) such that σi := σΓi is ergodic. Write A = L∞(X, µ), Mi = A ⋊ Γi, M = A ⋊ Γ and R(Γi y X), R(Γ y X) for the associated equivalence relations. Likewise, denote by ρ : Λ y (Y, ν) a free m.p. action of Λ on the prob- ability space (Y, ν) such that ρj := ρΛj is ergodic. Write B = L∞(Y, ν), Nj = B ⋊ Λj, N = B ⋊ Λ, and R(Λj y Y ),R(Λ y Y ) for the associated equivalence relations. We obtain the following analogs of Theorem 7.7 and Corollary 7.8 of [21]. Theorem 1.5 (W∗ Bass-Serre rigidity). If θ : M → N t is a ∗-isomorphism, then m = n, t = 1, and after permutation of indices there exist unitaries uj ∈ N such that for all j Ad(uj)θ(Mj) = Nj Ad(uj)θ(A) = B. In particular R(Γ y X) ≃ R(Λ y Y ) and R(Γj y X) ≃ R(Λj y Y ), for any j. Corollary 1.6 (OE Bass-Serre rigidity). If R(Γ y X) ≃ R(Λ y Y )t, then m = n, t = 1, and after permutation of indices, we have R(Γj y X) ≃ R(Λj y Y ), for any j. Conventions and notations. Throughout this paper, we write M for an arbitrary von Neumann algebra and M for a semifinite von Neumann algebra. Usually a state is denoted by ϕ or ψ and a trace is denoted by τ if it is finite and Tr if it is semifinite. States, traces and conditional expectations are always assumed to be faithful and normal. We shall denote by Mn := Mn(C) ⊗ M and M∞ := B(ℓ2)⊗M. Every von Neumann algebra is assumed to have separable predual. Also, (M)1 is the unit ball of M w.r.t. the operator norm. In Section 2, we extend some of Popa's intertwining techniques from finite to semifinite von Neumann algebras as well as some results from [21]. In Section 3, we give a generalization of Popa's spectral gap property in the context of semifinite amalgamated free products over an amenable von Neu- mann subalgebra. In Section 4, we prove Theorem 1.1. In Section 5, we prove Theorem 1.2 and deduce several corollaries. Finally, we give further rigidity results for finite amalgamated free products. Acknowledgements. The authors would like to express their warmest thanks to Stefaan Vaes for illuminating discussions on this paper. They are BASS-SERRE RIGIDITY RESULTS IN VON NEUMANN ALGEBRAS 5 also very grateful to the stimulating environment at University of California, Los Angeles, where this work was done. 2. Intertwining techniques for semifinite von Neumann algebras 2.1. Right Hilbert modules. Let (B, τ ) be a finite von Neumann algebra with a distinguished trace. Since τ is fixed, we simply denote L2(B, τ ) by L2(B). Let H be a right Hilbert B-module, i.e. H is a (separable) Hilbert space together with a normal ∗-representation π : Bop → B(H). For any b ∈ B, and ξ ∈ H, we shall simply denote π(bop)ξ = ξb. By the general theory, we know that there exists an isometry v : H → ℓ2⊗L2(B) such that v(ξb) = v(ξ)b, for any ξ ∈ H, b ∈ B. Since p = vv∗ commutes with the right B-action on ℓ2⊗L2(B), it follows that p ∈ B(ℓ2)⊗B. Thus, as right B-modules, we have H ≃ p(ℓ2⊗L2(B)). On B(ℓ2)⊗B, we define the following semifinite trace Tr (which depends on τ ): for any x = [xij]i,j ∈ (B(ℓ2)⊗B)+, Tr ([xij]i,j) =Xi τ (xii). We set dim(HB) = Tr(vv∗). Note that the dimension of H depends on τ but does not depend on the isometry v. Indeed take another isometry w : H → ℓ2⊗L2(B), satisfying w(ξb) = w(ξ)b, for any ξ ∈ H, b ∈ B. Note that vw∗ ∈ B(ℓ2)⊗B and w∗w = v∗v = 1. Thus, we have Tr(vv∗) = Tr(vw∗wv∗) = Tr(wv∗vw∗) = Tr(ww∗). Assume that dim(HB) < ∞. Then for any ε > 0, there exists a central projection z ∈ Z(B), with τ (z) ≥ 1 − ε, such that the right B-module Hz is finitely generated, i.e. of the form pL2(B)⊕n for some projection p ∈ Mn(C) ⊗ B. 2.2. Intertwining-by-bimodules device in the semifinite setting. In [32, 33], Popa introduced a very powerful tool to prove the unitary conjugacy of two von Neumann subalgebras of a tracial von Neumann algebra (M, τ ). If A, B ⊂ (M, τ ) are two possibly non-unital von Neumann subalgebras, denote by 1A, 1B the units of A and B. Note that we endow the finite von Neumann algebra B with the trace τ (1B · 1B)/τ (1B). Theorem 2.1 (Popa, [32, 33]). Let A, B ⊂ (M, τ ) be two possibly non-unital embeddings. The following are equivalent: (1) There exist n ≥ 1, a possibly non-unital ∗-homomorphism ψ : A → Bn and a non-zero partial isometry v ∈ M1,n(C)⊗1AM 1B such that xv = vψ(x), for any x ∈ A. (2) The A-B bimodule 1AL2(M )1B contains a non-zero A-B subbimod- ule H such that dim(HB) < ∞. 6 I. CHIFAN AND C. HOUDAYER (3) There is no sequence of unitaries (uk) in A such that kEB(a∗ukb)k2 → 0,∀a, b ∈ 1AM 1B. If one of the previous equivalent conditions is satisfied, we shall say that A embeds into B inside M and denote A (cid:22)M B. For our purpose, we need to extend Popa's intertwining techniques for semifinite von Neumann algebras. Namely, let (M, Tr) be a von Neumann algebra endowed with a faithful normal semifinite trace. We shall simply denote by L2(M ) the M -M bimodule L2(M, Tr), and by k·k2,Tr the L2-norm associated with Tr. We will use the following well-known inequality (k · k∞ is the operator norm): kxξyk2,Tr ≤ kξk2,Trkxk∞kyk∞,∀ξ ∈ L2(M ),∀x, y ∈ M. We shall say that a projection p ∈ M is Tr-finite if Tr(p) < ∞. Then p is necessarily finite. Moreover, pM p is a finite von Neumann algebra and τ := Tr(p · p)/ Tr(p) is a faithful normal tracial state on pM p. Remind that for any projections p, q ∈ M , we have p ∨ q − p ∼ q − p ∧ q. Then it follows that for any Tr-finite projections p, q ∈ M , p ∨ q is still Tr-finite and Tr(p ∨ q) = Tr(p) + Tr(q) − Tr(p ∧ q). Note that if a sequence (xk) in M converges to 0 strongly, then for any non-zero Tr-finite projection q ∈ M , kxkqk2,Tr → 0. Indeed, xk → 0 strongly in M ⇐⇒ x∗ =⇒ qx∗ =⇒ Tr(qx∗ kxkq) → 0 =⇒ kxkqk2,Tr → 0. kxk → 0 weakly in M kxkq → 0 weakly in qM q Moreover, there always exists an increasing sequence of Tr-finite projections (pk) in M such that pk → 1 strongly. Theorem 2.2. Let (M, Tr) be a semifinite von Neumann algebra. Let B ⊂ M be a von Neumann subalgebra such that TrB is still semifinite. Denote by EB : M → B the unique Tr-preserving conditional expectation. Let p ∈ M be a projection such that Tr(p) < ∞. Let A ⊂ pM p be a von Neumann subalgebra. The following conditions are equivalent: (1) There exists a non-zero Tr-finite projection q ∈ B such that the A- qBq bimodule L2(pM q) contains a non-zero A-qBq subbimodule H such that dim(HqBq) < ∞, where qBq is endowed with the trace Tr(q · q)/ Tr(q). (2) There is no sequence of unitaries (uk) in A such that EB(x∗uky) → 0 strongly, for any x, y ∈ pM . Definition 2.3. Under the assumptions of Theorem 2.2, if one of the equiv- alent conditions is satisfied, we shall still say that A embeds into B inside M and still denote A (cid:22)M B. BASS-SERRE RIGIDITY RESULTS IN VON NEUMANN ALGEBRAS 7 Proof of Theorem 2.2. (1) =⇒ (2). Write e = p ∨ q which is a Tr-finite projection in M . Thus Condition (1) tells exactly that A (cid:22)eM e qBq in the sense of Theorem 2.1. Hence there exist n ≥ 1, a possibly non-unital ∗-homomorphism ψ : A → (qBq)n and a non-zero partial isometry v ∈ M1,n(C) ⊗ pM q such that xv = vψ(x), for any x ∈ A. Automatically, v∗v ≤ ψ(p). Assume that there exists a sequence of unitaries (uk) in A such that EB(x∗uky) → 0 strongly, for any x, y ∈ pM . In particular, we get Since ψ(uk) are unitaries in ψ(p)Bnψ(p) and v∗ukv = ψ(uk)v∗v, we have kEBn(v∗ukv)k2,Trn → 0. kEBn (v∗v)k2,Trn = kEBn (v∗v)ψ(uk)k2,Trn = kEBn (v∗vψ(uk))k2,Trn = kEBn (v∗ukv)k2,Trn → 0. This implies that EBn(v∗v) = 0 and thus v∗v = 0. Contradiction. (2) =⇒ (1). From (2), there exist ε > 0, a finite set F ⊂ L2(B) and a finite set K ⊂ (pM )1 such that max x,y∈K,ξ∈F kEB(x∗uy)ξk ≥ ε,∀u ∈ U (A). Fix an increasing sequence of Tr-finite projections (pk) in B such that pk → 1 strongly. Since pk(ξ) → ξ for any ξ ∈ F , and B ∩ L2(B) is dense in L2(B), we obtain that there exist ε′ > 0 and k0 ∈ N large enough, such that x,y∈K kEB(x∗uy)pkk2,Tr ≥ ε′,∀u ∈ U (A),∀k ≥ k0. max Moreover we have k(1 − pk)EB(x∗uy)pkk2,Tr ≤ k(1 − pk)x∗uypkk2,Tr ≤ k(1 − pk)x∗k2,Tr. Since K is finite and x ∈ pM with p a Tr-finite projection, it follows that k(1 − pk)EB(x∗uy)pkk2,Tr → 0, uniformly for any x, y ∈ K, and u ∈ U (A). Thus there exist ε′′ > 0, a Tr-finite projection q = pk in B for k large enough, such that x,y∈K kqEB(x∗uy)qk2,Tr ≥ ε′′,∀u ∈ U (A). max The rest of the proof is now exactly the same as the one of [33], because if we denote by e = p ∨ q, we are working in the finite von Neumann algebra eM e. (cid:3) 8 I. CHIFAN AND C. HOUDAYER 2.3. Controlling quasi-normalizers in a semifinite AFP. We first fix some notation. Let P ⊂ Q be an inclusion of von Neumann algebras. We denote by NQ(P ) := {u ∈ U (Q) : uP u∗ = P} P ai and P x ⊂Xi the group of all unitaries in Q that normalize P inside Q. The normalizer of P inside Q is the von Neumann algebra NQ(P )′′. Since every unitary in P ′ ∩ Q normalizes P , we have P ′ ∩ Q ⊂ NQ(P )′′. We say that the inclusion P ⊂ Q is regular if NQ(P )′′ = Q. More generally, we denote by QN Q(P ) := {x ∈ Q : ∃a1, . . . , an ∈ Q, xP ⊂Xi aiP} the set of all elements in Q that quasi-normalize P inside Q. Note that QN Q(P ) is a unital ∗-algebra. The quasi-normalizer of P inside Q is the von Neumann algebra QN Q(P )′′. Moreover, we have P ′ ∩ Q ⊂ QN Q(P ). We say that the inclusion P ⊂ Q is quasi-regular if QN Q(P )′′ = Q. Important convention. For i = 1, 2, let Mi be a von Neumann algebra and B ⊂ Mi be a common von Neumann subalgebra. Assume that there exist faithful normal conditional expectations Ei : M → B. Write (M, E) := (M1, E1)∗B (M2, E2) the amalgamated free product. We will simply denote M = M1∗B M2 if no confusion is possible. We shall say that M is a semifinite amalgamated free product, if there exists a semifinite faithful normal trace Tr on M such that: • TrMi and TrB are still semifinite; • Tr◦E = Tr and Tr◦Ei = Tr. Whenever we refer to a trace Tr on a semifinite amalgamated free product, we always mean a trace Tr that satisfies the previous conditions. We prove the following analog of Theorem 1.2.1 of [21]. Nevertheless, the proof follows the same strategy as the one of Theorem 4.6 of [20]. Theorem 2.4. Let M = M1∗B M2 be a semifinite amalgamated free product. Denote by Tr the semifinite trace on M . Let p ∈ M1 be a projection such that Tr(p) < ∞. Let Q ⊂ pM1p be a von Neumann subalgebra such that Q (cid:14)M1 B. Then, any Q-pM1p subbimodule H of L2(pM p) such that dim(HpM1p) < ∞, is contained in L2(pM1p). In particular, Q′ ∩ pM p, NpM p(Q)′′ and QN pM p(Q)′′ are contained in pM1p. Proof. Since Q (cid:14)M1 B, we know there exists a sequence of unitaries (un) in Q such that EB(a∗unb) → 0 strongly, for any a, b ∈ pM1. We prove the following claim: Claim 2.5. ∀x, y ∈ pM p ⊖ pM1p,kEpM1p(xuny)k2,Tr → 0. Proof of Claim 2.5. Let x and y be reduced words in (M )1 with letters al- ternatingly from M1⊖ B and M2⊖ B. We assume that both x and y contain at least a letter from M2 ⊖ B. We set x = x′a with a = 1 if x ends with a letter from M2 ⊖ B and a equal to the last letter of x otherwise. Note that x′ is a reduced word ending with a letter from M2 ⊖ B. In the same way, BASS-SERRE RIGIDITY RESULTS IN VON NEUMANN ALGEBRAS 9 we set y = by′ with b = 1 if y begins with a letter from M2 ⊖ B and b equal to the first letter of y otherwise. Note that y′ is a reduced word beginning with a letter from M2 ⊖ B. Then for z ∈ Q, we have EM1(xzy) = EM1(x′EB(azb)y′). Apply this equality to z = un. Since EB(aunb) → 0 strongly, it follows from the equation above that EM1(xuny) → 0 strongly. Consequently, kEpM1p(pxunyp)k2,Tr = kpEM1(xuny)pk2,Tr → 0. Using Kaplansky density theorem together with the fact that un ∈ Q ⊂ pM1p, we are done. (cid:3) The rest of the proof is now exactly the same as the one of Theorem 4.6 (cid:3) in [20]. 3. Spectral gap property for semifinite AFP In this section, we will freely use the language of Hilbert bimodules over von Neumann algebras (see [3, 7]). We collect here a few properties we will be using throughout. Let M, N, P be any von Neumann algebras. For M -N bimodules H, K, denote by πH (resp. πK) the associated ∗- representation of the binormal tensor product M ⊗bin N op on H (resp. on K). We refer to [14] for the definition of ⊗bin. We say that H is weakly contained in K and denote it by H ≺ K if the representation πH is weakly contained in the representation πK, that is if ker(πH) ⊃ ker(πK). For a von Neumann algebra M , denote by L2(M ) the standard representation of M that gives the identity bimodule. Let H, K be M -N bimodules. The following are true: (1) Assume that H ≺ K. Then, for any N -P bimodule L, we have H⊗N L ≺ K⊗N L, as M -P bimodules. Exactly in the same way, for any P -M bimodule L, we have L⊗M H ≺ L⊗M K, as P -N bimodules (see Lemma 1.7 in [3]). (2) A von Neumann algebra B is amenable iff L2(B) ≺ L2(B)⊗L2(B), as B-B bimodules. Let B, M, N be von Neumann algebras such that B is amenable. Let H be any M -B bimodule and let K be any B-N bimodule. Then, as M -N bimodules, we have H⊗BK ≺ H⊗K (straightforward consequence of (1) and (2)). We prove the following analog of Lemma 2 in [31]: Proposition 3.1. Let M = M1 ∗B M2 be a semifinite amalgamated free product, and denote by Tr the semifinite trace. Assume that B is amenable. Let p ∈ M1 be a projection such that Tr(p) < ∞. Let Q ⊂ pM1p be a von Neumann subalgebra with no amenable direct summand. Then, for any free ultrafilter ω on N, we have Q′ ∩ (pM p)ω ⊂ (pM1p)ω. 10 I. CHIFAN AND C. HOUDAYER Proof. Denote Ki = L2(Mi) ⊖ L2(B), for i = 1, 2. Since there exists a Tr- preserving normal conditional expectation F1 : M → M1, it follows that L2(M1) ⊖ L2(M1) is a M1-M1 subbimodule of L2(M ) and more precisely, we have the following isomorphism as M1-M1 bimodules (see [41, 47]): L2(M ) ⊖ L2(M1) ∼=Mn≥1 Hn, where Hn = L2(M1)⊗B 2n−1 z K2⊗BK1⊗B ··· ⊗BK1⊗BK2 ⊗BL2(M1). } { Let p ∈ M1 be a non-zero Tr-finite projection. Cutting down with p, as pM1p-pM1p bimodules, we have L2(pM p) ⊖ L2(pM1p) ∼=Mn≥1 pHnp. Since B is amenable, from the standard properties of composition and weak containment of correspondences recalled at the beginning of this sec- tion, it follows that as pM1p-pM1p bimodules Consequently, we obtain 2n−1 z K2⊗K1⊗··· ⊗K1⊗K2 ⊗L2(M1p). pHnp ≺ L2(pM1)⊗ L2(pM p) ⊖ L2(pM1p) ≺M L2(pM1)⊗L2(M1p). } { Note now that as a left pM1p-module, L2(pM1) is always a submodule of L L2(pM1p), and exactly the same thing for the right pM1p module L2(M1p). Thus, we finally have L2(pM p) ⊖ L2(pM1p) ≺M L2(pM1p)⊗L2(pM1p). Since pM p is a finite von Neumann algebra, the proof is then exactly the same as the one of Lemma 2 of [31]. (cid:3) If Q ⊂ pM1p has no amenable direct summand, then for any ε > 0, there exist δ > 0 and a finite subset F ⊂ U (Q) such that for any x ∈ (pM p)1, (1) kux − xuk2,Tr < δ,∀u ∈ F =⇒ kx − EpM1p(x)k2,Tr < ε. Remind that a II1 factor N is said to be full if any central sequence is trivial, i.e. for any bounded sequence (xn) in N satisfying kxny− yxnk2 → 0 for any y ∈ N , then kxn − τ (xn)1k2 → 0. In the case of amalgamated free products of finite von Neumann algebras, we obtain the following corollary: Corollary 3.2. Let (Ni, τi) be a finite von Neumann algebra endowed with a distinguished trace, for i = 1, 2. Let B ⊂ Ni be a common amenable von Neumann subalgebra such that τ1B = τ2B. Denote by N = N1 ∗B N2 the amalgamated free product. If one of the Ni's is a full II1 factor, then N is a full II1 factor. BASS-SERRE RIGIDITY RESULTS IN VON NEUMANN ALGEBRAS 11 Proof. Assume that N1 is a full II1 factor. Fix ω a free ultrafilter on N. We have N ′ 1 = C. Since N1 is a non-amenable II1 factor, Proposition 3.1 yields N ′ 1 . Then, we obtain 1 ∩ N ω 1 ∩ N ω ⊂ N ω N ′ ∩ N ω ⊂ N ′ ⊂ N ′ = C. 1 ∩ N ω 1 ∩ N ω 1 Thus, N is a full II1 factor. (cid:3) 4. Key intertwining theorem for semifinite AFP 4.1. Notation. We fix some notation that we will be using throughout this section. For i = 1, 2, let Ni be a von Neumann algebra and let B ⊂ Ni be a common von Neumann subalgebra such that there exist normal faithful conditional expectations Ei : Ni → B. Write N = N1 ∗B N2 the amalgamated free product. We shall always assume that N is semifinite and denote by Tr its semifinite trace (see Section 2). Write Mi = Ni∗B (B⊗L(Z)) and M = M1 ∗B M2. We still denote by Ei : Mi → B the conditional expectation, and Tr the semifinite trace on M . Note that M = N ∗B (B⊗L(F2)). In Mi, denote by ui the generating Haar unitary of L(Z). Let f : T1 → ] − 1, 1] be the Borel function satisfying exp(π√−1f (z)) = z, ∀z ∈ T1. Define hi = f (ui) a selfadjoint element in Mi such that exp(π√−1hi) = ui. i = exp(tπ√−1hi) ∈ U (Mi). Following [21], define the deformation Write ut (αt) on M = M1 ∗B M2 by: αt = (Ad ut 1) ∗B (Ad ut 2),∀t ∈ R. Moreover define the period-2 automorphism β on M = N ∗B (B⊗L(F2)) by: β(x) = x,∀x ∈ N, β(ui) = u∗ i ,∀i ∈ {1, 2}. It was proven in [21] that the deformation (αt) satisfies a certain malleability type condition: αtβ = βα−t,∀t ∈ R. Note that αt and β are Tr-preserving. Hence, we shall still denote by αt and β the actions on L2(M ) and note that β(x) = x, for any x ∈ L2(N ). Recall that the s-malleable deformation (αt, β) automatically features a certain tranversality property. Proposition 4.1 (Popa, Lemma 2.1 in [30]). We keep the same notation as before. We have the following: (2) kx − α2t(x)k2,Tr ≤ 2kαt(x) − EN (αt(x))k2,Tr, ∀x ∈ L2(N ),∀t > 0. 12 I. CHIFAN AND C. HOUDAYER 4.2. Key intertwining theorem. All the theorems mentioned in the in- troduction will be consequences of the following general intertwining result: Theorem 4.2. We keep the same notation as before. Assume that B is amenable and N = N1 ∗B N2 is a semifinite amalgamated free product, where Tr denotes the semifinite trace. Let q ∈ N be a non-zero Tr-finite projection and let Q ⊂ qN q be a von Neumann subalgebra with no amenable direct summand. Then there exists i ∈ {1, 2} such that Q′ ∩ qN q (cid:22)N Ni. Proof. For i ∈ {1, 2}, write Mi = Ni ∗B (B⊗L(Z)), and define M = M1 ∗B M2. Note that M = N ∗B (B⊗L(F2)). We will be using the notation introduced in subsection 4.1. Let q ∈ N be a non-zero projection such that Tr(q) < ∞. Let Q ⊂ qN q be a von Neumann subalgebra with no amenable direct summand. Assume that Q′ ∩ qN q (cid:14)N Ni, for all i ∈ {1, 2}. We shall obtain a contradiction. Denote Q0 = Q′ ∩ qN q. Step (1) : Using the spectral gap condition and the transversal- ity property to find t > 0 and a nonzero intertwiner v between Id and αt. The first step of the proof uses a well-known argument due to Popa which appeared in Theorem 4.1 and Lemma 5.2 in [30]. For the sake of complete- ness, we will reproduce the argument from there. Let ε = 1 8kqk2,Tr. We know that there exist δ > 0, a finite subset F ⊂ U (Q), with q ∈ F , such that for every x ∈ (qM q)1, ku − αt(u)k2 Since αt → Id pointwise ∗-strongly as t → 0, and since for any u ∈ F k[x, u]k2,Tr ≤ δ,∀u ∈ F =⇒ kx − EqN q(x)k2,Tr ≤ ε. 2,Tr − ℜ Tr(u∗αt(u))(cid:1) 2,Tr − ℜ Tr(qu∗αt(u)q)(cid:1) , 48kqk2,Tr(cid:27) . max{ku − αt(u)k2,Tr : u ∈ F} ≤ min(cid:26) δ 2,Tr = 2(cid:0)kuk2 = 2(cid:0)kuk2 we may choose t = 1/2k small enough (k ≥ 1) such that 1 , 4 For every x ∈ Q0 and every u ∈ F ⊂ Q, writing q = (q − αt(q)) + αt(q), we have [qαt(x)q, u] = qαt(x)qu − uqαt(x)q = (q − αt(q))αt(x)qu + [αt(x), u] − uqαt(x)(q − αt(q)). For every x ∈ (Q0)1 and every u ∈ F ⊂ Q, since [u, x] = 0, we have k[qαt(x)q, u]k2,Tr ≤ k(q − αt(q))αt(x)quk2,Tr + k[αt(x), u]k2,Tr +kuqαt(x)(q − αt(q))k2,Tr ≤ 2kq − αt(q)k2,Tr + k[αt(x), u − αt(u)]k2,Tr ≤ 2kq − αt(q)k2,Tr + 2ku − αt(u)k2,Tr ≤ δ. BASS-SERRE RIGIDITY RESULTS IN VON NEUMANN ALGEBRAS 13 Thus, for every x ∈ (Q0)1, kqαt(x)q − EqN q(qαt(x)q)k2,Tr ≤ ε = 1 Now, for every x ∈ Q0, writing αt(q) = q + (αt(q) − q), we have 8kqk2,Tr. αt(x) = qαt(x)q + (αt(q) − q)αt(x)q + qαt(x)(αt(q) − q) +(αt(q) − q)αt(x)(αt(q) − q). Consequently, we get for every x ∈ (Q0)1, kαt(x) − EN (αt(x))k2,Tr ≤ 6kq − αt(q)k2,Tr +kqαt(x)q − EqN q(qαt(x)q)k2,Tr ≤ 6kq − αt(q)k2,Tr + ε ≤ 1 4kqk2,Tr. Using Proposition 4.1, we get for every x ∈ (Q0)1 1 2kqk2,Tr, where s = 2t. Thus, for every u ∈ U (Q0), we have kx − αs(x)k2,Tr ≤ ku∗αs(u) − qk2,Tr = ku∗(αs(u) − u)k2,Tr ≤ ku − αs(u)k2,Tr ≤ 1 2kqk2,Tr. Denote by C = cow{u∗αs(u) : u ∈ U (Q0)} ⊂ qM αs(q) the ultraweak closure of the convex hull of all u∗αs(u), where u ∈ U (Q0). Denote by a the unique element in C of minimal k · k2,Tr-norm. Since ka − qk2,Tr ≤ 1/2kqk2,Tr, necessarily a 6= 0. Fix u ∈ U (Q0). Since u∗aαs(u) ∈ C and ku∗aαs(u)k2,Tr = kak2,Tr, necessarily u∗aαs(u) = a. Taking v = pol(a) the polar part of a, we have found a non-zero partial isometry v ∈ qM αs(q) such that (3) xv = vαs(x),∀x ∈ Q0. The rest of the proof, namely Steps (2) and (3), are very similar to the reasoning in Lemma 4.8 and Theorem 6.1 in [34], Theorems 4.1 in [32] and Theorem 4.3 in [21] (see also Theorem 5.6 in [20]). For the sake of completeness, we will nevertheless give a detailed proof. Step (2) : Using the malleability of (αt, β) to lift Equation (3) till s = 1. Note that it is enough to find a non-zero partial isometry w ∈ qM α2s(q) such that xw = wα2s(x),∀x ∈ Q0. Indeed, by induction we can go till s = 1. Remind that β(z) = z, for every z ∈ N . Write w = αs(β(v∗)v). Since v ∈ qM αs(q), we see that w ∈ αsβαs(q)M α2s(q). But αsβαs = β. Hence, w ∈ qM α2s(q). Note that 14 I. CHIFAN AND C. HOUDAYER vv∗ ∈ Q′ Theorem 2.4 that Q′ 0 ∩ qM q. Since Q0 (cid:14)N Ni, it follows that Q0 (cid:14)N B. We know from 0 ∩ qM q ⊂ qN q. In particular, vv∗ ∈ qN q. Then, ww∗ = αs(β(v∗)vv∗β(v)) = αs(β(v∗)β(vv∗)β(v)) = αsβ(v∗v) 6= 0. Hence, w is a non-zero partial isometry in qM α2s(q). Moreover, for every x ∈ Q0 wα2s(x) = αs(β(v∗)vαs(x)) = αs(β(v∗)xv) = αs(β(v∗x)v) = αs(β(αs(x)v∗)v) = αsβαs(x)αs(β(v∗)v) = β(x)w = xw. Step (3) : Using the intertwining-by-bimodules technique to conclude. Thus, we have found a non-zero partial isometry v ∈ qM α1(q) such that xv = vα1(x),∀x ∈ Q0. (4) Note that v∗v ∈ α1(Q0)′ ∩ α1(q)M α1(q). Since α1 : qM q → α1(q)M α1(q) is a ∗-automorphism, and Q0 (cid:14)N B, Theorem 2.4 gives α1(Q0)′ ∩ α1(q)M α1(q) = α1(cid:0)Q′ 0 ∩ qM q(cid:1) ⊂ α1 (qN q) . Hence v∗v ∈ α1(qN q). Set A = L(F2). Denote Ki = L2(Ni) ⊖ L2(B). Denote by P1 the or- thogonal projection from L2(N ) on L2(B) ⊕ K1 ⊕ K2. Define the sub- space Halt ⊂ L2(M ) as the L2-closed linear span of B and the words in N1 ∗B N2 ∗B (B⊗A) with letters alternatingly from N1 ⊖ B, N2 ⊖ B, B⊗(A⊖ C1) and such that two consecutive letters never come from N1⊖ B, N2 ⊖ B. This means that letters from N1 ⊖ B and N2 ⊖ B are always separated by a letter from B⊗(A ⊖ C1). By the definition of α1, it follows that α1(L2(N )) ⊂ Halt. Denote by Palt the orthogonal projection of L2(M ) onto Halt. Since Q0 (cid:14)N Ni, for any i ∈ {1, 2}, there exists a sequence of unitaries (un) in Q0 such that ENi(a∗unb) → 0 strongly, ∀i ∈ {1, 2},∀a, b ∈ qN . Moreover, we have the following: Claim 4.3. ∀c, d ∈ qM α1(q), kEα1(N )(c∗und)k2,Tr → 0. Proof of Claim 4.3. Let c, d ∈ (M )1 (with M = N ∗B (B⊗A)) be either in B or reduced words with letters alternatingly from N⊖B and B⊗(A⊖C1). Set c = c′a, with a = c if c ∈ N , a = 1 if c ends with a letter from B⊗(A ⊖ C1) BASS-SERRE RIGIDITY RESULTS IN VON NEUMANN ALGEBRAS 15 and a equal to the last letter of c otherwise. Note that c′ is either equal to 1 or a reduced word ending with a letter from B⊗(A ⊖ C1). Exactly in the same way, set d = bd′, with b = d if d ∈ N , b = 1 if d begins with a letter from B⊗(A ⊖ C1) and b equal to the first letter of d otherwise. Note that d′ is either equal to 1 or a reduced word beginning with a letter from B⊗(A ⊖ C1). N1 ∗B N2 and by definition of the projection Palt, it is clear that For x ∈ N , write cxd = c′(axb)d′, and note that axb ∈ N . Since N = Palt(c′zd′) = 0,∀z ∈ L2(N ) ⊖ (L2(B) ⊕ K1 ⊕ K2). By definition of the conditional expectations EMi, i = 1, 2, and EB, it is easy to see that P1(aunb) = EM1(aunb) + EM2(aunb) − EB(aunb),∀n. (Note that aunb ∈ N ∩ L2(N ).) Then P1(aunb) → 0 strongly. Note that by construction of c′ and d′, we have Palt(cund) = Palt(c′P1(aunb)d′). In particular, since α1(L2(N )) ⊂ Halt, we get Eα1(N )(cund) = Eα1(N )(c′P1(aunb)d′). Thus, Eα1(N )(cund) → 0 strongly and kα1(q)Eα1(N )(cund)α1(q)k2,Tr → 0. Using Kaplansky density theorem together with the fact that un ∈ Q0 ⊂ qM q, we get the claim. (cid:3) Let's come back to Equation (4). Recall that for the unitaries (un) in Q0, we have unv = vα1(un). Since α1(un) are unitaries in α1(qN q), we get kv∗vk2,Tr = kv∗vα1(un)k2,Tr = kEα1(N )(v∗vα1(un))k2,Tr = kEα1(N )(v∗unv)k2,Tr → 0. Hence, v∗v = 0. Contradiction. (cid:3) 5. Applications to prime factors 5.1. Preliminaries. Let M be a von Neumann algebra. Let ϕ be a state on M. Denote by Mϕ the centralizer and M = M ⋊σϕ R the core of M, where σϕ is the modular group associated with the state ϕ. Denote by πσϕ = M → M , the representation of M in its core M , and denote by λs the unitaries in L(R) implementing the action σϕ. Consider the dual weight bϕ on M (see [39]) which satisfies the following: σ bϕ t (λs) = λs,∀s ∈ R. σ bϕ t (πσϕ (x)) = πσϕ(σϕ t (x)),∀x ∈ M Note that bϕ is a semifinite weight on M . Write θϕ the dual action of σϕ on M , where we identify R with its Pontryagin dual. Take now hϕ a non- singular positive self-adjoint operator affiliated with L(R) such that his ϕ = 16 I. CHIFAN AND C. HOUDAYER λs, for any s ∈ R. Define Trϕ := bϕ(h−1 trace on M and the dual action θϕ scales the trace Trϕ: ϕ ·). We get that Trϕ is a semifinite Trϕ ◦θϕ s (x) = e−s Trϕ(x),∀x ∈ M+,∀s ∈ R. There is also a functorial construction of the core of the von Neumann algebra M which does not rely on the choice of a particular state or weight ϕ (see [10, 12, 15]). This is called the non-commutaive flow of weights. Take two states ϕ, ψ on M. It follows from [15] and Theorem XII.6.10 in [38] that there exists a natural ∗-isomorphism Πϕ,ψ : M ⋊σϕ R → M ⋊σψ R such that Πϕ,ψ ◦ θϕ = θψ ◦ Πϕ,ψ and Trϕ = Trψ ◦Πϕ,ψ. In the rest of this section, we will simply denote by (M, θ, Tr) the non-commutative flow of weights, where θ is the dual action of R on the core M and Tr is the trace on M such that Tr◦θs = e−s Tr, for any s ∈ R. This construction does not depend on the choice of a state on M. Remind that if M is a factor, then the dual action θ is ergodic on the center Z(M ). We prove the following: Proposition 5.1. Let M be a non-amenable factor. Denote by (M, θ, Tr) its non commutative flow of weights. Then for any non-zero central projection z ∈ Z(M ), M z is not amenable. Moreover for any projection p ∈ M such that Tr(p) < ∞, pM p is a non-amenable finite von Neumann algebra. Proof. Assume that there is a non-zero central projection z ∈ Z(M ) such that M z is amenable. Thus M θt(z) = θt(M z) is still amenable. Define e =Wt∈R θt(z). It is clear that e ∈ Z(M ) and θt(e) = e, for any t ∈ R. By ergodicity of the action θ on Z(M ), we get e = 1. Now write zi = θti(z) for i = 1, 2, such that M zi is amenable. Note that z1 ∨ z2 = z1 + z2 − z1z2 = z1 + z2(1 − z1), so that M (z1 ∨ z2) = M z1 + M z2(1 − z1) is still amenable. Since amenability is stable under direct limits, and since Wt∈R θt(z) = 1, it follows that M is amenable. But by duality, M⊗B(L2(R)) ≃ M ⋊θ R. Consequently, M would be amenable. Contradiction. We may assume that M is properly infinite, so that M itself is properly infinite. Let p ∈ M be non-zero projection such that Tr(p) < ∞. Denote z = z(p) the central support of p in M . Since M is properly infinite, M z is still properly infinite and M z ≃ pM p⊗B(ℓ2). Since M z is not amenable, pM p is not amenable. (cid:3) 5.2. Main result. We first introduce some notation. Let (B, τ ) be a finite von Neumann algebra of type I: for example B = C, B is finite dimensional or B = L∞[0, 1]. For i = 1, 2, let Mi be a von Neumann algebra endowed with a conditional expectation Ei : Mi → B. We shall always assume that B 6= Mi. Denote by (M, E) = (M1, E1)∗B (M2, E2) the amalgamated free product. Write • ϕi = τ ◦ Ei, Mi = Mi ⋊σϕi R; • ϕ = τ ◦ E, M = M ⋊σϕ R. BASS-SERRE RIGIDITY RESULTS IN VON NEUMANN ALGEBRAS 17 Note that the modular groups satisfy the following equation: σϕ σϕ2 t (σϕ t ∗B , for any t ∈ R. Set λs the unitaries implementing the modular action t ). Define the canonical conditional expectations bEi : Mi → B⊗L(R) satisfying bEi(xλs) = Ei(x)λs, for any x ∈ Mi, for any s ∈ R. Exactly in the same way we can define bE : M → B⊗L(R). results (see Theorem 5.1 in [41]) that M , the core of M, is given by It follows from Ueda's t = σϕ1 (M, bE) = (M1, bE1) ∗(B⊗L(R)) (M2, bE2). Denote by Trϕi the semifinite trace coming from the dual weight bϕi on Mi. Exactly in the same way, denote by Trϕ the semifinite trace on M . It is straightforward to check: Trϕi ◦bEi = Trϕi Trϕ ◦bE = Trϕ . , TrϕB⊗L(R) are still semifinite. Then M is a semifinite Moreover TrϕMi amalgamated free product in the sense of Section 2. We will simply denote the semifinite trace Trϕ by Tr in the rest of the section. Theorem 5.2. Let (B, τ ) be a finite von Neumann algebra of type I. Assume B ⊂ Mi but B 6= Mi. Let Ei : Mi → B be a conditional expectation, for i = 1, 2. Denote by M = M1 ∗B M2 the amalgamated free product. Assume that M is a non-amenable factor. Then M is prime. Proof. We will be using the notation introduced at the beginning of this subsection. We prove the result by contradiction and we assume that M is not prime, i.e. M = P1⊗P2, where Pi is a diffuse factor (i.e. not of type I). Since M is a non-amenable factor, we may assume that P1 is a non-amenable factor. Thanks to Corollary 8 of [11], we may choose a state ψi on Pi such that the centralizer P ψi is a von Neumann algebra of type II1. Denote ψ = ψ1 ⊗ ψ2. Note that we can write the core M in two different ways: i M = (M1, bE1) ∗(B⊗L(R)) (M2, bE2) M = (P1⊗P2) ⋊σψ R. We denote by Tr the canonical trace on M scaled by the dual action θ (see the previous subsection). Denote by Pi = Pi ⋊σψi R the core of Pi. Fix a non-zero projection p ∈ L(R) ⊂ P1 such that Tr(p) < ∞. Since P1 is a non-amenable factor, the finite von Neumann algebra pP1p has no amenable direct summand (see Proposition 5.1). From Theorem 4.2, we know that there exists i = 1, 2 such that (pP1p)′ ∩ pM p (cid:22)M Mi. Note that P ψ2 2 p ⊂ (pP1p)′ ∩ pM p. In particular there exists n ≥ 1, a non-zero partial isometry v ∈ M1,n(C)⊗ M , a projection q ∈ M n i such that Trn(q) < ∞ and a (unital) ∗-homomorphism ρ : P ψ2 2 p → qM n i q such that xv = vρ(x), for any x ∈ P ψ2 2 p. Denote by Q = hP1, vv∗i the von Neumann subalgebra of M generated by P1 and vv∗. Then Q is still semifinite. We have vv∗ ∈ Q and 18 I. CHIFAN AND C. HOUDAYER 2 p)′ ∩ qM nq. Since P ψ2 2 p (cid:14)M B⊗L(R). Consequently, Theorem 2.4 yields v∗v ∈ qM n v∗v ∈ ρ(P ψ2 2 p is a von Neumann algebra of type II1, then P ψ2 i q, so that we may assume v∗v = q. Moreover, Theorem 2.4 yields v∗Qv ⊂ qM n i q. With the finite projection vv∗ ∈ Q, we can find a sequence of partial isometries (ul) ∈ Q such that u∗ l ∈ Z(Q). De- fine w := [ulv]l ∈ M1,∞(C)⊗M n. Note that, ww∗ = z, and w∗P1w ⊂ w∗w(M n l ulv) is not of finite trace in M n i ⊗B(ℓ2). Note that now we are working in the semifinite amalgamated free product: l ul ≤ vv∗ and z := Pl ulu∗ i ⊗B(ℓ2))w∗w. But w∗w = Diag(v∗u∗ M ∞ = M ∞ 1 ∗(B⊗L(R))∞ M ∞ 2 . Fix now an increasing sequence of projections (pk) in L(R) ⊂ P1 such that p0 = p, Tr(pk) < ∞, for any k, and pk → 1 strongly. Denote wk := pkw ∈ M1,∞(C)⊗M n, and note that wk is still a partial isometry (since ww∗ ∈ Z(Q)) and Tr(wkw∗ k) < ∞. We apply now the same strategy as before. Since w∗P ψ1 1 pkw ⊂ w∗ kwk(M n kwk is a subalgebra of type II1, an application of Theorem 2.4 (for each k) yields w∗pkP2pkw ⊂ kwk(M n w∗ i ⊗B(ℓ2))w∗ i ⊗B(ℓ2))w∗ kwk. Thus, we get for any k, w∗pkP2pkw ⊂ w∗w(M n i ⊗B(ℓ2))w∗w. i ⊗B(ℓ2))w∗w. Since pk → 1 strongly, we get w∗P2w ⊂ w∗w(M n i ⊗B(ℓ2))w∗w, for any j = Consequently, we obtain w∗Pjw ⊂ w∗w(M n 1, 2. Since ww∗ commutes with P1, the von Neumann algebra generated by w∗P1w and w∗P2w is exactly w∗M w, and w∗M w ⊂ w∗w(M n i ⊗B(ℓ2))w∗w. Cutting down with the projection p0 = p, this implies in particular that 0w0(M n i ⊗B(ℓ2))w∗ w∗pM pw ⊂ w∗ 0w0(M n⊗B(ℓ2))w∗ w∗ 0w0. Therefore, 0w0 = w∗ 0w0(M n i ⊗B(ℓ2))w∗ 0w0. Since B⊗L(R) 6= Mi, by definition of the amalgamated free product M = M1 ∗(B⊗L(R)) M2, we get a contradiction. (cid:3) Theorem 5.2 is no longer true for non-amenable factors arising as amal- gamated free products over an amenable von Neumann algebra. Look at the following trivial counter-example: for i = 1, 2 take Ni a II1 factor, write Mi = R⊗Ni, where R is the hyperfinite II1 factor and Ei = Id⊗τi. Then M := (R⊗N1) ∗R (R⊗N2) = R⊗(N1 ∗ N2) is a Mc Duff II1 factor and hence not prime. 5.3. Examples of prime factors. We deduce now several corollaries of Theorem 5.2 and give new examples of prime factors. We first consider the case of plain free products. For i = 1, 2, let (Mi, ϕi) be any von Neu- mann algebra endowed with a faithful normal state. Denote by (M, ϕ) = (M1, ϕ1) ∗ (M2, ϕ2) the free product. The von Neumann algebra M is known to be a full factor (i.e. Inn(M) is closed in Aut(M) [9]) if one of the following conditions holds: BASS-SERRE RIGIDITY RESULTS IN VON NEUMANN ALGEBRAS 19 (1) (Barnett [4]): ∃u ∈ U (Mϕ1 1 ),∃v, w ∈ U (Mϕ2 2 ), ϕ1(u) = ϕ2(v) = ϕ2(w) = ϕ2(v∗w) = 0; (2) (Ueda [42, 43]): Mϕ1 1 is diffuse and M2 6= C. We thank Y. Ueda for pointing out to us (2). Consequently, we obtain Corollary 5.3. Assume that M satisfies (1) or (2) so that M is a full factor. Then M is a prime factor. Gao & Junge proved in [17] that any free product (M, ϕ) = ∗i∈I (Mi, ϕi) of amenable von Neumann algebras is solid in a general sense and hence prime. It was proven by Ricard & Xu in [36] that such a free product M always has the complete metric approximation property, (denoted c.m.a.p.) there exists a net (φn) of normal finite-rank maps on M such that i.e. lim supkφnkcb ≤ 1 and φn → idM in the pointwise-ultraweak topology (see [3, 13]). Take now any countable group Γ such that Λcb(Γ) > 1, e.g. Γ = Z2 ⋊ SL(2, Z). Then for any von Neumann algebra M 6= C, endowed with a faithful normal state ϕ, the free product L(Γ) ∗ (M, ϕ) is a non-amenable factor, thus prime by Theorem 5.2 and which does not have the c.m.a.p. Consequently, Theorem 5.2 gives many examples of prime factors that do not have the c.m.a.p. We consider now the case of amalgamated free products. Firstly, the finite case. For i = 1, 2, let Mi be a II1 factor and B ⊂ Mi be a common abelian von Neumann subalgebra, such that τ1B = τ2B. Write M = M1 ∗B M2. We thank S. Vaes for showing us the following claim. Claim 5.4. The amalgamated free product M is a non-amenable II1 factor. Proof of Claim 5.4. The fact that M is always a II1 factor follows from Theorem 1.1 of [21]. We consider the following alternative: Assume B is not diffuse. Let p ∈ B be a non-zero minimal projection. It is straightforward to check that pM1p ∗pB pM2p ⊂ pM p. Since pB = Cp, we get pM1p ∗ pM2p ⊂ pM p. It is obvious that a free product of II1 factors is never amenable. Thus, M itself is non-amenable. Assume B is diffuse. For n ≥ 3, since B is diffuse, we may choose orthogonal projections p1, . . . , pn ∈ B such that Pi pi = 1 and τ (pi) = 1/n. Since M1 and M2 are both II1 factors denote by (ei,j) (resp. (fi,j)) a system of matrix unit in M1 (resp. M2) such that ei,i = fi,i = pi,∀i ∈ {1, . . . , n}. Instead of writing {1, . . . , n} for the set of indices, we will be using the notation Zn := Z/nZ, which is more convenient. Write u = Xi∈Zn v = Xi∈Zn ei,i+1 ∈ M1 fi,i+1 ∈ M2. 20 I. CHIFAN AND C. HOUDAYER It is straightforward to check that u ∈ U (M1), v ∈ U (M2) and un = vn = 1. Moreover, we have upku∗ = vpkv∗ = pk−1,∀k ∈ Zn. Hence for any k ∈ Zn but k 6= 0 and any j ∈ Zn, we have ukpj = pj−kuk. Applying EB, we obtain EB(uk)pj = pj−kEB(uk) = EB(uk)pj−k, since B is assumed to be abelian. This implies EB(uk) = 0. Likewise, we get EB(vk) = 0. It follows in particular that u and v are ∗-free in M w.r.t. the trace τ . Since u and v generate two copies of L(Zn), we have shown that L(Zn) ∗ L(Zn) ⊂ M . Since L(Zn) ∗ L(Zn) is non-amenable for n ≥ 3, it follows that M is a non-amenable II1 factor. Corollary 5.5. For i = 1, 2, let Mi be a II1 factor and B ⊂ Mi be a common abelian von Neumann subalgebra. Then the amalgamated free product M1∗B M2 is a prime II1 factor. (cid:3) More generally, for i = 1, 2, let now Mi be a non-type I factor such that A ⊂ Mi is a common Cartan subalgebra, i.e. (necessarily unique). • There exists a faithful normal conditional expectation Ei : Mi → A • A ⊂ Mi is a MASA, i.e. A′ ∩ Mi = A. • A ⊂ Mi is regular, i.e. NMi(A)′′ = Mi. It follows from Ueda's results (see [41, 44]), that under these assumptions, the amalgamated free product M1 ∗A M2 is a non-amenable factor. It is even non-Mc Duff (see Theorem 8 in [43]). Thus, we get Corollary 5.6. Assume that Mi is a non-type I factor and A ⊂ Mi is a common Cartan subalgebra. Then the amalgamated free product M1 ∗A M2 is a prime factor. In particular, let Γ = Γ1 ∗ Γ2 be a free product of countable infinite groups. Let σ : Γ y (X, µ) be a free action which leaves the measure is ergodic µ quasi-invariant. Assume moreover that the restriction σΓi and non-transitive (see [40]). One can view the crossed product M := L∞(X, µ) ⋊ Γ as the amalgamated free product M = M1 ∗L∞(X,µ) M2, with Mi = L∞(X, µ) ⋊ Γi and Ei : Mi → L∞(X, µ) is the canonical condi- tional expectation. Consequently, we obtain, under those assumptions, that L∞(X, µ) ⋊ Γ is a prime factor. We point out that the assumption of ergodicity on σΓi cannot be removed in general. Indeed amenable factors may appear as amalgamated free prod- ucts over a Cartan subalgebra. It suffices to take an amenable free ergodic action F2 y (X, µ), leaving the measure µ quasi-invariant. It follows that M = L∞(X, µ) ⋊ F2 is an amenable factor, hence non-prime. Nevertheless, M is the amalgamated free product M = (L∞(X, µ) ⋊ Z) ∗L∞(X,µ) (L∞(X, µ) ⋊ Z) BASS-SERRE RIGIDITY RESULTS IN VON NEUMANN ALGEBRAS 21 We quickly remind such a construction and refer to Section 6 of [44] for further details. For the free group Fn = hg1, . . . , gni on n ≥ 2 generators, denote by ∂Fn its boundary: ∂Fn = {(ωi) ∈YN∗{g1, g−1 1 , . . . , gn, g−1 n } : ∀i ∈ N, ωi 6= ω−1 i+1}. The boundary ∂Fn is a compact space for the product topology and its topology is generated by the following clopen sets: Ω(γ) = {ω = (ωi) ∈ ∂Fn : ω1 = γ1, . . . , ωr = γr}, for a reduced word γ = γ1 ··· γr. It is easy to see that the action of Fn by left multiplication on ∂Fn is continuous. By [1], this action is known to be amenable. Consider now the probability measure µ defined on ∂Fn by: µ(Ω(γ)) = 1 2n(cid:18) 1 2n − 1(cid:19)l(γ)−1 with the word length function l(·). It follows from [22, 27, 35] that µ is quasi-invariant under Fn and moreover the action Fn y (∂Fn, µ) is free and ergodic. It follows that the associated crossed product von Neumann algebra L∞(∂Fn, µ) ⋊ Fn is an amenable factor. This factor is moreover of (see [35]). Consequently by Connes' result [8], it is the unique type III 1 . Write now Fn = Γ1 ∗ Γ2 as AFD factor of type III 1 a free product of infinite groups. Thus we have , denoted by R 1 2n−1 2n−1 2n−1 R 1 2n−1 = (L∞(∂Fn, µ) ⋊ Γ1) ∗L∞(∂Fn,µ) (L∞(∂Fn, µ) ⋊ Γ2). 6. Further rigidity results for finite AFP 6.1. A Kurosh type result and consequences. In the finite case, we can obtain some more precise results. We have the following analog of Theorem 5.1 in [21]: Theorem 6.1. Let (Mi, τi) be a finite von Neumann algebra endowed with a distinguished trace, for i = 1, 2. Let B ⊂ Mi be a common amenable von Neumann subalgebra such that τ1B = τ2B. Denote by M = M1 ∗B M2 the amalgamated free product. Let Q ⊂ M be a von Neumann subalgebra with no amenable direct summand. Denote Q0 = Q′∩ M the relative commutant. Then there exists i = 1, 2 such that Q0 (cid:22)M Mi. (1) Assume that Q0 (cid:14)M B. Then there exists i = 1, 2, n ≥ 1 and a non-zero partial isometry v ∈ M1,n(C) ⊗ M such that v∗Q0v ⊂ M n i (2) Assume that Q0 (cid:14)M B and M1, M2 are factors. Then, there exists a unique pair of projections q1, q2 ∈ Z(Q′ 0 ∩ M ), satisfying q1 + q2 = 1, and unitaries ui ∈ U (M ) such that ui(Q0qi)u∗ i ⊂ Mi, for i = 1, 2. Proof. Theorem 4.2 yields i = 1, 2 such that Q0 (cid:22)M Mi. Thus, there exists n ≥ 1, a projection q ∈ M n i , a non-zero partial isometry v ∈ M1,n(C) ⊗ M and a unital ∗-homomorphism θ : Q0 → qM n i q such that xv = vθ(x), for 22 I. CHIFAN AND C. HOUDAYER any x ∈ Q0. Assume now that Q0 (cid:14)M B and let us prove (1). Since v∗v ∈ θ(Q0)′ ∩ qM nq and θ(Q0) (cid:14)qM nq B, it follows from Theorem 1.1 in [21] that v∗v ∈ qM n i q so that we may assume v∗v = q. Consequently, v∗Q0v ⊂ M n i . For (2), the proof is now exactly the same as the one of Theorem 5.1 of [21]. (cid:3) Under additional assumptions, an amalgamated free product M1 ∗B M2, with B just amenable, might be prime. From a result of Hermann & Jones (see Lemma 1 in [19]), if a non-inner amenable group Γ acts on a finite von Neumann algebra (P, τ ) in a trace-preserving way, then the crossed product M = P ⋊ Γ satisfies M ′ ∩ M ω ⊂ P ω. If we moreover assume that the action is strongly ergodic, i.e. L(Γ)′ ∩ P ω = C, then M ′ ∩ M ω = C, and M is a full II1 factor. Combining this observation and Corollary 3.2, we can obtain the following result: Theorem 6.2. Let (B, τ ) be any finite amenable von Neumann algebra. (1) For i = 1, 2, let (Ni, τi) be finite von Neumann algebras such that τ1B = τ2B = τ . Write N = N1 ∗B N2 and assume that N1 is a full II1 factor. Then N is a full prime II1 factor. (2) Let Γ1, Γ2 be countable discrete groups such that Γ1 ≥ 2 and Γ2 ≥ 3. Denote Γ = Γ1 ∗ Γ2, which is automatically non inner-amenable. For any strongly ergodic trace-preserving action of Γ on B, B ⋊ Γ is a full prime II1 factor. Proof. We note that for a non-prime II1 factor N = N1⊗N2, N1, N2 are necessarily non-amenable. Then the proof is very similar to the one of Theorem 6.1. (cid:3) 6.2. Solidity and semisolidity. Following [23, 24], a von Neumann alge- bra M is said to be solid if for any diffuse von Neumann subalgebra A ⊂ M , the relative commutant A′ ∩ M is amenable. It is said to be semisolid if for any type II1 von Neumann subalgebra A ⊂ M , the relative commutant A′ ∩ M is amenable. Ozawa proved that L(Γ) is solid for any countable group Γ in the class S (see [23]), and L∞(X, µ) ⋊ Γ is semisolid for any free ergodic m.p. action of a class S group Γ on the non-atomic probability space (X, µ) (see [24]). Moreover, he showed that the following countable groups are in the class S: word-hyperbolic groups [23], the wreath products Λ ≀ Γ with Λ amenable and Γ ∈ S [24] and Z2 ⋊ SL(2, Z) [25]. We prove the following stability properties: Theorem 6.3. For i = 1, 2, let (Mi, τi) be a finite diffuse von Neumann algebra with a distinguished trace. (1) M1 and M2 are solid iff the free product M1 ∗ M2 is solid. (2) Take B ⊂ Mi a common von Neumann subalgebra of type I such that τ1B = τ2B. Assume that M1 and M2 are II1 factors. Then M1 and M2 are semisolid iff the amalgamated free product M1 ∗B M2 is semisolid. BASS-SERRE RIGIDITY RESULTS IN VON NEUMANN ALGEBRAS 23 Proof. Since proofs of (1) and (2) are similar, and since moreover (1) can be deduced from Theorem 1.4 of [28], we only prove (2). Since M1 is a II1 factor and B is of type I, it follows from Theorem 1.1 in [21] that M is a II1 factor. We prove the result by contradiction. Assume that there exists a von Neumann subalgebra Q ⊂ M with a non-amenable direct summand such that Q′ ∩ M is of type II1. Since M is a factor, by looking at an amplification over a corner of Q, we may assume that Q has no amenable direct summand and Q0 = Q′ ∩ M is still of type II1. From Theorem 6.1, there exist a unitary u ∈ M , a non-zero projection q ∈ Z(Q′ 0 ∩ M ), and i = 1, 2 such that u(Q0q)u∗ ⊂ Mi. Denote p = uqu∗. Theorem 1.1 in [21] implies that u((Q′ 0 ∩ M )q ∨ Q0q)u∗ ⊂ pMip. Note that Q0q is of type II1. Moreover since Q ⊂ Q′ no amenable direct summand, it follows that Q′ direct summand either. Thus, with q ∈ Z(Q′ algebra (Q′ semisolid. 0 ∩ M and Q has 0 ∩ M has no amenable 0 ∩ M ), the von Neumann 0 ∩ M )q is not amenable. This contradicts the fact that pMip is (cid:3) We cannot obtain the same statement as (2) for solidity: namely, even if M1, M2 are solid and B is diffuse abelian, M = M1 ∗B M2 is not solid in general. For example take the following inclusion of free groups Λ = ha, b2i ⊂ ha, bi = Γ. Note that [Γ : Λ] = ∞. Look at the following generalized Bernoulli shift Γ y [0, 1]Γ/Λ = Γ y Yg∈Γ/Λ [0, 1]g. It is a free ergodic m.p. action. Write M = L∞(cid:0)[0, 1]Γ/Λ(cid:1) ⋊ Γ. Since Λ acts trivially on L∞([0, 1]eΛ) and since Λ is not amenable, the relative commutant L∞([0, 1]eΛ)′ ∩ M is not amenable. If we want to get solidity of such an amalgamated free product, we need additional assumptions. Let Γ be a countable group, and σ : Γ y (X, µ) be a free ergodic m.p. action. We shall say that this action is solid if for any diffuse von Neumann subalgebra A ⊂ L∞(X, µ), the relative commutant A′∩(L∞(X)⋊Γ) is amenable. We motivate this definition with the following result: Theorem 6.4. Let Γ = Γ1 ∗ Γ2 be a free product of countable groups and consider Γ y (X, µ) a free ergodic m.p. action on the probability space. Denote by M = L∞(X, µ) ⋊ Γ, Mi = L∞(X, µ) ⋊ Γi and note that M = M1 ∗L∞(X,µ) M2. Then M is solid iff M1, M2 are solid and the action Γ y (X, µ) is solid. Proof. We only need to prove the "if" part. We prove the result by con- tradiction. Since M is II1 factor, there exists a von Neumann subalgebra A ⊂ M with no amenable direct summand such that A0 = A′ ∩ M is dif- fuse. Thus, we know there exists i = 1, 2 such that A0 (cid:22)M Mi. There exist n ≥ 1, a projection p ∈ M n i p and i , a unital ∗-homomorphism ψ : A0 → pM n 24 I. CHIFAN AND C. HOUDAYER a non-zero partial isometry v ∈ M1,n(C)⊗ M such that xv = vψ(x), for any (v∗v). x ∈ A0. We may assume that p equals the support projection of EM n L∞(X)n. The same proof as (2) of Theorem 6.3 will lead to a contradiction, namely it will contradict the fact that Mi is solid. First case: ψ(A0) (cid:14)M n i i i Second case: ψ(A0) (cid:22)M n L∞(X)n. Using Remark 3.8 in [46], it follows that A0 (cid:22)M L∞(X). Then there exists m ≥ 1, a projection q ∈ L∞(X)m, a non-zero partial isometry w ∈ M1,m(C)⊗ M and a unital ∗-homomorphism θ : A0 → qL∞(X)mq such that xw = wθ(x), for any x ∈ A0. Since θ(A0) is diffuse, by solidity of the action, θ(A0)′ ∩ qM mq is amenable. Consequently, 0 ∩ M )ww∗ is amenable w∗w(θ(A0)′ ∩ qM mq)w∗w is amenable and ww∗(A′ as well. Since A has no amenable direct summand, and A ⊂ A′ 0 ∩ M , it follows that A′ 0 ∩ M has no amenable direct summand either. We get a contradiction. (cid:3) We refer to [6] for some applications of the notion of solid action in ergodic theory. 6.3. W∗/OE Bass-Serre rigidity results. Let (X, µ) be the standard Borel non-atomic probability space. Let R be a countable Borel measure- preserving equivalence relation on (X, µ). Denote by [R], the full group of all Borel m.p. isomorphisms φ : X → X such that (x, φ(x)) ∈ R for almost every x ∈ X. Denote by [[R]], the set of all partial Borel m.p. isomorphisms φ : dom(φ) → rng(φ), such that (x, φ(x)) ∈ R for almost every x ∈ dom(φ). A partial Borel isomorphism φ ∈ [[R]] is said to be properly outer if φ(x) 6= x, for almost any x ∈ dom(φ). Remind the following notion of freeness for equivalence relations due to Gaboriau. Definition 6.5 (Gaboriau, [16]). Let (Rk)k∈N be a sequence of m.p. equiv- alence relations on the probability space (X, µ). The sequence (Rk) is said to be free if for any n ≥ 1, for any i1 6= ··· 6= in ∈ N, for any φj ∈ [[Rij ]], whenever φj is properly outer, the product φ1 ··· φn is still properly outer. In order to state the main result, we first introduce some notation. Fix integers m, n ≥ 1. For each i ∈ {1, . . . , m}, and j ∈ {1, . . . , n} let Γi = Gi × Hi Λj = G′ j × H ′ j j are not amenable j are infinite. Note that Γi and Λj have a vanishing first L2-Betti be ICC (infinite conjugacy class) groups, such that Gi, G′ and Hi, H ′ number (see [5, 29]). Denote Γ = Γ1 ∗ ··· ∗ Γm and Λ = Λ1 ∗ ··· ∗ Λn. Let σ : Γ y (X, µ) be a free m.p. action of Γ on the probability space (X, µ) such that σi := σΓi is ergodic. Write A = L∞(X, µ), Mi = A ⋊ Γi, M = A ⋊ Γ, and R(Γi y X),R(Γ y X) for the associated equivalence relations. Likewise, denote by ρ : Λ y (Y, ν) a free m.p. action of Λ on the prob- ability space (Y, ν) such that ρj := ρΛj is ergodic. Write B = L∞(Y, ν), BASS-SERRE RIGIDITY RESULTS IN VON NEUMANN ALGEBRAS 25 Nj = B ⋊ Λj, N = B ⋊ Λ, and R(Λj y Y ),R(Λ y Y ) for the associated equivalence relations. Then we have R(Γ y X) ≃ R(Γ1 y X) ∗ ··· ∗ R(Γm y X) R(Λ y Y ) ≃ R(Λ1 y Y ) ∗ ··· ∗ R(Λn y Y ). We obtain the following analogs of Theorem 7.7 and Corollary 7.8 of [21]. Using our Theorem 4.2, the proofs are then exactly the same. These results can be viewed as Bass-Serre type rigidity results. Theorem 6.6. If θ : M → N t is a ∗-isomorphism, then m = n, t = 1, and after permutation of indices there exist unitaries uj ∈ N such that for all j Ad(uj)θ(Mj) = Nj Ad(uj)θ(A) = B. In particular R(Γ y X) ≃ R(Λ y Y ) and R(Γj y X) ≃ R(Λj y Y ), for any j. Corollary 6.7. If R(Γ y X) ≃ R(Λ y Y )t, then m = n, t = 1, and after permutation of indices, we have R(Γj y X) ≃ R(Λj y Y ), for any j. Corollary 6.7 has been recently generalized by Alvarez & Gaboriau [2] to all non-amenable countable groups Γi, Λj with a vanishing first L2-Betti number. See [2] for a precise statement. References [1] S. Adams, Boundary amenability for word hyperbolic groups and an application to smooth dynamics of simple groups. Topology 33 (1994), 763 -- 783. [2] A. Alvarez & D. Gaboriau, Free products, orbit equivalence and measure equiva- lence rigidity. arXiv:0806.2788 [3] C. Anantharaman-Delaroche, Amenable correspondences and approximation properties for von Neumann algebras. Pacific J. Math. 171 (1995), 309 -- 341. [4] L. Barnett, Free product von Neumann algebras of type III. Proc. Amer. Math. Soc. 123 (1995), 543 -- 553. [5] B. Bekka & A. Valette, Group cohomology, harmonic functions and the first L2-Betti number. Potential Anal. 6 (1997), 313 -- 326. [6] I. Chifan & A. Ioana, Ergodic subequivalence relations induced by a Bernoulli action. arXiv:0802.2353 [7] A. Connes, Noncommutative geometry. Academic Press. San Diego, California, 1994. [8] A. Connes, Classification of injective factors. Ann. of Math. 104 (1976), 73 -- 115. [9] A. Connes, Almost periodic states and factors of type III1. J. Funct. Anal. 16 (1974), 415 -- 445. [10] A. Connes, Une classification des facteurs de type III. Ann. Sci. ´Ecole Norm. Sup. 6 (1973), 133 -- 252. [11] A. Connes & E. Størmer, Homogeneity of the state space of factors of type III1. J. Funct. Anal. 28 (1978), 187 -- 196. [12] A. Connes & M. Takesaki, The flow of weights on factors of type III. Tohoku Math. J. 29 (1977), 473 -- 575. [13] M. Cowling & U. Haagerup, Completely bounded multipliers of the Fourier alge- bra of a simple Lie group of real rank one. Invent. Math. 96 (1989), 507 -- 549. 26 I. CHIFAN AND C. HOUDAYER [14] E.G. Effros & E.C. Lance, Tensor products of operator algebras. Adv. in Math. 25 (1977), 1 -- 34. [15] A.J. Falcone & M. Takesaki, Non-commutative flow of weights on a von Neu- mann algebra. J. Funct. Anal. 182 (2001), 170 -- 206. [16] D. Gaboriau, Cout des relations d'´equivalence et des groupes. Invent. Math. 139 (2000), 41 -- 98. [17] M. Gao & M. Junge, Examples of prime von Neumann algebras. Int. Math. Res. Notices. Vol. 2007 : article ID rnm042, 34 pages. [18] L. Ge, Applications of free entropy to finite von Neumann algebras, II. Ann. of Math. 147 (1998), 143 -- 157. [19] R.H. Hermann & V.F.R. Jones, Central sequences in crossed products. Cont. Math. 62 (1987), 539 -- 544. [20] C. Houdayer, Construction of type II1 factors with prescribed countable fundamen- tal group. J. reine angew Math. 634 (2009), 169-207. [21] A. Ioana, J. Peterson & S. Popa, Amalgamated free products of w-rigid factors and calculation of their symmetry groups. Acta Math. 200 (2008), 85 -- 153. [22] G. Kuhn & T. Steger, More irreducible boundary representations of free groups. Duke Math. J. 82 (1996), 381 -- 436. [23] N. Ozawa, Solid von Neumann algebras. Acta Math. 192 (2004), 111 -- 117. [24] N. Ozawa A Kurosh-type theorem for type II1 factors. Int. Math. Res. Notices. Vol. 2006 : article ID 97560, 21 pages. [25] N. Ozawa, An example of a solid von Neumann algebra. Hokkaido Math. J. 38 (2009), 557 -- 561. [26] N. Ozawa & S. Popa, On a class of II1 factors with at most one Cartan subalgebra. Ann. of Math., to appear. arXiv:0706.3623 [27] C. Pensavalle & T. Steger, Tensor products with anisotropic principal series representations of free groups. Pacific J. Math. 173 (1996), 181 -- 202. [28] J. Peterson, L2-rigidity in von Neumann algebras. Invent. Math. 175 (2009), 417 -- 433. [29] J. Peterson & A. Thom, Group cocycles and the ring of affiliated operators. arXiv:0708.4327 [30] S. Popa, On the superrigidity of malleable actions with spectral gap. J. Amer. Math. Soc. 21 (2008), 981 -- 1000. [31] S. Popa, On Ozawa's property for free group factors. Int. Math. Res. Notices. Vol. 2007 : article ID rnm036, 10 pages. [32] S. Popa, Strong rigidity of II1 factors arising from malleable actions of w-rigid groups I. Invent. Math. 165 (2006), 369-408. [33] S. Popa, On a class of type II1 factors with Betti numbers invariants. Ann. of Math. 163 (2006), 809 -- 899. [34] S. Popa, Some rigidity results for non-commutative Bernoulli Shifts. J. Funct. Anal. 230 (2006), 273 -- 328. [35] J. Ramagge & G. Robertson, Factors from trees. Proc. Amer. Math. Soc. 125 (1997), 2051 -- 2055. [36] ´E. Ricard & Q. Xu, Khintchine type inequalities for reduced free products and applications. J. reine angew. Math. 599 (2006), 27 -- 59. [37] D. Shlyakhtenko, Prime type III factors. Proc. Nat. Acad. Sci., 97 (2000), 12439 -- 12441. [38] M. Takesaki, Theory of Operator Algebras II. EMS 125. Springer-Verlag, Berlin, Heidelberg, New-York, 2000. [39] M. Takesaki, Duality for crossed products and structure of von Neumann algebras of type III. Acta Math. 131 (1973), 249 -- 310. [40] A. Tornquist, Orbit equivalence and actions of Fn. J. Symbolic Logic 71 (2006), 265 -- 282. BASS-SERRE RIGIDITY RESULTS IN VON NEUMANN ALGEBRAS 27 [41] Y. Ueda, Amalgamated free products over Cartan subalgebra. Pacific J. Math. 191 (1999), 359 -- 392. [42] Y. Ueda, Remarks on free products with respect to non-tracial states. Math. Scand. 88 (2001), 111 -- 125. [43] Y. Ueda, Fullness, Connes' χ-groups, and ultra-products of amalgamated free prod- ucts over Cartan subalgebras. Trans. Amer. Math. Soc. 355 (2003), 349 -- 371. [44] Y. Ueda, Amalgamated free products over Cartan subalgebra, II. Supplementary results and examples. Advanced Studies in Pure Mathematics, Operator Algebras and Applications. 38 (2004), 239 -- 265. [45] S. Vaes, Rigidity results for Bernoulli actions and their von Neumann algebras (after S. Popa). S´eminaire Bourbaki, expos´e 961. Ast´erisque 311 (2007), 237-294. [46] S. Vaes, Explicit computations of all finite index bimodules for a family of II1 fac- tors. Ann. Sci. ´Ecole Norm. Sup. 41 (2008), 743 -- 788. [47] D.-V. Voiculescu, K.J. Dykema & A. Nica, Free random variables. CRM Mono- graph Series 1. American Mathematical Society, Providence, RI, 1992. Math Dept, Vanderbilt University, Nashville, TN 37240 and IMAR, Bucharest, Romania E-mail address: [email protected] CNRS ENS Lyon, UMPA UMR 5669, 69364 Lyon cedex 7, France E-mail address: [email protected]
1306.2163
1
1306
2013-06-10T10:54:25
Algebraic reformulation of Connes embedding problem and the free group algebra
[ "math.OA" ]
We give a modification of I. Klep and M. Schweighofer algebraic reformulation of Connes' embedding problem by considering *-algebra of the countably generated free group. This allows to consider only quadratic polynomials in unitary generators instead of arbitrary polynomials in self-adjoint generators.
math.OA
math
Algebraic reformulation of Connes embedding problem and the free group algebra. Kate Juschenko, Stanislav Popovych Abstract We give a modification of I. Klep and M. Schweighofer algebraic reformulation of Connes' embedding problem by considering ∗-algebra of the countably generated free group. This allows to consider only quadratic polynomials in unitary generators instead of arbitrary poly- nomials in self-adjoint generators. KEYWORDS: Connes' Embedding Problem, I I1-factor, sum of her- mitian squares, positivity. 1 Introduction. Let ω ∈ β(N) \ N be a free ultrafilter on N and R be the hyperfinite II1- factor with faithful tracial normal state τ . Then the subset Iω in l∞(N, R) consisting of (x1, x2, . . .) with limn→ω τ (x∗ nxn) = 0 is a closed ideal in l∞(N, R) and a quotient algebra Rω = l∞(N, R)/Iω is a von Neumann II1-factor called ultrapower of R. It is naturally endowed with a faithful tracial normal state τω((xn) + Iω) = lim n→ω τ (xn). A. Connes' embedding problem asks whether every finite von Neumann algebra with fixed normal faithful tracial state can be embedded into Rω in a trace-preserving way. It is well know that Connes' embedding problem is equivalent to the problem whether every finite set x1, . . . , xn of self-adjoint contractions in 0 2000 Mathematics Subject Classification: 46L07, 46K50 (Primary) 16S15, 46L09, 16W10 (Secondary) 1 arbitrary II1-factor (M, τ ) has matricial microstates, i.e whether for any ε > 0 and t ≥ 1 there is k ∈ N and self-adjoint contractive k×k-matrices A1, . . . , An such that tr(w(x1, . . . , xn))−τ (w(A1, . . . , An)) < ε for all words w of length at most t. In [3] D. Hadwin proved that solving Connes' embedding problem in affirmative is equivalent to proving that there is no polynomial p(x1, ..., xn) in non-commutative variables such that 1. trk(p(A1, . . . , An)) ≥ 0 for every k and self-adjoint contractions A1, ..., An ∈ Mk. 2. τ (p(T1, . . . , Tn)) < 0, where T1, . . . , Tn are self-adjoint contractive ele- ments in a finite factor with trace τ . Recently I. Klep and M. Schweighofer established that Connes' embed- ding problem has the following equivalent algebraic reformulation. Let f (X1, . . . , Xm) be a self-adjoint element in a free associative algebra KhXi with countable family of self-adjoint generators X = {X1, X2, . . .}, where K = R or K = C. If tr(f (A1, . . . , Am)) ≥ 0 for any n and family of self- adjoint contractive matrices A1, . . . , Am ∈ Mn(K) then f has the property that for every ε > 0 we have εe + f = g + c where c is a sum of commutators in KhXi, g belongs to quadratic module generated by 1 − X 2 i and e is the unit in KhXi. Recall that a quadratic module is the smallest subset of KhXi containing unit, closed under addition and conjugation x → g∗xg by arbitrary g ∈ KhXi. In the present paper we consider the group ∗-algebra F of the countably generated free group F∞ = hu1, u2, . . .i instead of KhXi. One reason is that we can use a more standard and well known set of hermitian squares {g∗gg ∈ F } instead of quadratic module M and the second that we can bound the degree of polynomials f in the above reformulation by 2. This modification provides the following. Theorem. Connes' embedding conjecture is true iff for any self-adjoint f ∈ F of the form f (u1, . . . , un) = αe +Pi6=j αiju∗ i uj condition T r(f (V1, . . . , Vn)) ≥ 0 (1) for every m ≥ 1 and every n-tuple of unitary matrices V1, . . . , Vn ∈ U(m) implies that for every ε > 0, εe + f = g + c where c is a sum of commutators and g is a sum of Hermitian squares. 2 We will call f satisfying (1) a trace-positive quadratic polynomial. Ele- ments of the form g+c with c being a sum of commutators are called cyclically equivalent to g (see Section 2). In Section 3 we study a subset of correlation matrices of the form [tr(U ∗ i Uj)]ij where U1, . . . , Un runs over n-tuple of unitary matrices and tr(U) denotes nor- malized trace of U. Using Clifford algebra methods we show that this set contains all correlation matrices with real coefficients. This implies that all trace-positive quadratic polynomials f with real coefficients do satisfy the property from the above theorem. The description of the set {[tr(U ∗ i Uj)]ij U1, . . . , Un ∈ Um(C), m ≥ 1} seems In this case it is equivalent to the prob- to be unknown even for n = 3. lem of description of the set of triples (tr(U), tr(V ), tr(UV )) where U and V are unitary matrices. Note that the lists of possible eigenvalues of U, V and UV can be described by generalization of Horn's inequalities (see [2]) but little is know about possible traces (tr(U), tr(V ), tr(UV )). The only known connection between these traces seems to be the inequality q1 − tr(UV )2 ≤ q1 − tr(U)2 +q1 − tr(V )2 established in [13]. Acknowledgements The authors are indebted to Igor Klep and Markus Schweighofer for care- ful reading of the paper and a number of suggestions which helped to improve the paper a lot. 2 An algebraic reformulation of Connes' prob- lem. Let F be the ∗-algebra of the countably generated free group F∞. Let K denote the R-subspace in Fsa generated by the commutators f g − gf (f, g ∈ cyc F ). We will say that f and g in F are cyclically equivalent (denote f ∼ g) if f − g ∈ K. Let Σ2(F ) denote the set of positive elements of the ∗-algebra j fj with fj ∈ F . An element of the form f ∗f is called Hermitian square and therefore the cone Σ2(F ) is called the cone of Hermitian squares. F , i.e. elements of the form Pm j=1 f ∗ Definition 1. Let C be a subset of the vector space V . An element v ∈ C is called an algebraic interior point of C if for every u ∈ V there is ε > 0 in R s.t. v + λu ∈ C for all 0 ≤ λ ≤ ε. 3 Definition 2. Let A be a unital ∗-algebra with the unit e. Then 1. An element a ∈ Asa is called bounded if there is α ∈ R+ such that αe ± a ∈ Σ2(A). 2. An element x = a + ib with a, b ∈ Asa is bounded if the elements a and b are such. 3. The algebra A is bounded if all its elements are bounded. It is well known that the set of all bounded elements in A is a ∗-subalgebra in A and that an element x ∈ A is bounded if and only if xx∗ is such (see for example [9, 5]). In particular F is a bounded ∗-algebra. Obviously this implies that the unit of the algebra is an algebraic interior point of Σ2(F ). The following lemma is a modification of Theorem 3.12 in [7]. Lemma 3. Let f ∈ F be self-adjoint. If for any II1 factor M with faithful normal tracial state τ and separable predual and every n-tuple of unitary elements U1, . . . , Un in the unitary group U(M) of M we have that τ (f (U1, . . . , Un)) ≥ 0 then for every ε > 0, εe + f ∼ g for some g ∈ Σ2(F ). Proof. Clearly Σ2(F ) + K is a convex cone in R-space Fsa. Since e is an an algebraic internal point of Σ2(F ) it is also an algebraic internal point of Σ2(F ) + K. Assume that there is ε > 0 such that εe + f 6∼ g for any g ∈ Σ2(F ), i.e. εe + f 6∈ Σ2(F ) + K. By Eidelheit-Kakutani separation theorem there is R-linear unital functional L0 : Fsa → R s.t. L0(Σ2(F ) + K) ⊆ R≥0 and L0(εe + f ) ∈ R≤0. Since −K ⊂ Σ2(F ) + K we have that L0(K) = 0. In particular extending L0 to C-linear functional on F we get a tracial functional L. Since L maps Σ2(F ) into the non-negative reals it defines a pre-Hilbert space structure on F by means of sesquilinear for hp, qi = L(q∗p), p, q ∈ F . Let N = {p : hp, pi = 0}. By Cauchy-Schwarz inequality N = {p : L(q∗p) = 0 for all q ∈ F } and hence is a left ideal. Let H0 be the pre-Hilbert space F /N. Consider the left regular representation π : F → L(H0). Since π is a ∗-homomorphism for every f ∈ F operator π(f ) is bounded as a linear combination of unitary operators. Thus π(f ) can be extended to the bounded operator acting on the Hilbert space H which is the completion of H0. Thus we have a representation π : F → B(H) with a cyclic vector ξ = e + N and such that L(p) = hπ(p)ξ, ξi. In particular L is a contractive tracial state 4 on F and thus defines a tracial state of the universal enveloping C ∗-algebra C ∗(F ). By Banach-Alaoglu and Krein-Milman theorem we can assume that L is an extreme point in the set of all tracial states and thus π(F ) generates a factor von Neumann algebra M (see [3]). Clearly M is a finite factor. If it is type I then it should be C (since ξ is a trace vector) and thus can be embedded into any II1-factor in trace preserving way. Thus we can assume that M is a type II1-factor. But then condition L(f ) < 0 is impossible. Corollary 4. If self-adjoint f ∈ F has real coefficients and for any real type II1 von Neumann algebra (M, τ ) with normal faithful tracial state τ and every n-tuple of unitary elements U1, . . . , Un in M we have that τ (f (U1, . . . , Un)) ≥ 0 then the same holds for the complex II1 von Neumann algebras. Proof. Element f can be written as f = α +Pwj 2Pwj j ) with αwj ∈ R and for complex trace τ and U1, . . . , Un ∈ U(M) we will have τ (f ) = α + αwj Re τ (wj), i.e. τ (f ) = (Re τ )(f ). To finish the proof note that M can be regarded as a real finite von Neumann algebra with faithful trace Re τ . αwj (wj + w∗ Lemma 5. If f ∈ R[F∞], f = f ∗ and for every real type II1 von Neumann cyc algebra (M, τ ) we have that τ (f ) ≥ 0 then for every ε > 0, ε + f ∼ g for some g ∈ nPm j=1 g∗ j gj m ∈ N, gj ∈ RhF∞io. Proof. The proof of this statement can be obtained by obvious modification of the proof of lemma 3. The only nontrivial part is that the unit e is an algebraic internal point but this is equivalent to RhF∞i being bounded ∗- algebra. The proof of the last fact can be found in [12]. This lemma gives another proof of corollary 4. In sequel we will need the following lemma. Lemma 6. If (M, τ ) is a II1 factor which can be embedded into Rω and f ∈ F is self-adjoint then the condition tr(f (V1, . . . , Vn)) ≥ 0 for all m ≥ 0 and all unitary V1, . . . , Vn in Mm×m(C) implies that τ (f (U1, . . . , Un)) ≥ 0 for all unitary U1, . . . , Un in M. 5 trace on Rω we can find a representing sequences nu(k) Proof. Considering M as a subalgebra in Rω and τ as a restriction of the for Uk, k = 1, . . . , n in l∞(N, R) which are unitary elements in von Neumann algebra l∞(N, R). This can be done since every unitary in von Neumann algebra Rω can be lifted to a unitary in von Neumann algebra l∞(N, R) with respect to canonical morphism π : l∞(N, R) → Rω. Taking j sufficiently large we can approximate mixed moments of U1, . . . , Uk up to order m, i.e. τ (Us1 . . . Ust) with t ≤ m and s1, . . . , st ∈ {1, . . . , n}, by the mixed moments of unitary matrices u(k) j o∞ 1 , . . . , u(k) n . j=1 The following theorem is Proposition 4.6 in [6] Theorem 7. (E. Kirchberg) Let (M, τ ) be von Neumann algebra with separable predual and faithful normal tracial state τ . If for all n ≥ 1 and for all unitaries u1, . . . , un in M and for arbitrary ε > 0 there exists m ≥ 1 and unitary m × m matrices V1, . . . , Vn ∈ U(m) s.t. for all i, j: τ (u∗ i uj) − 1 m τ (uj) − Tr(V ∗ i Vj) < ε, 1 m Tr(Vj) < ε (2) (3) then M can be embedded into Rω. Remark 8. We may drop condition (3) since we may take u0 = 1, u1, . . . , un and by (2) find matrices W0, . . . , Wn such that τ (u∗ i Wj) < ε for all i and j. Thus (2) and (3) will be satisfied if we take Vj = W ∗ m Tr(W ∗ i uj) − 1 0 Wj. The proof of the following theorem is an adaptation of the proof of Propo- sition 3.17 from [7]. Theorem 9. Let (M, τ ) be II1-factor with separable predual. If for every i uj the condition self-adjoint element f ∈ F of the form f = α +Pi6=j αiju∗ Tr(f (V1, . . . , Vn)) ≥ 0 for all m ≥ 1 and every n-tuple of unitary matrices V1, . . . , Vn ∈ U(m) implies that τ (f (U1, . . . , Un)) ≥ 0 for all unitaries U1, . . . , Un in M then M can be embedded into Rω. 6 Pi6=j αiju∗ Proof. Take n ≥ 1. Consider the finite dimensional vector space W = {αe + i ujαij ∈ C}. Denote by C the convex hull of the set F of the functionals T ∈ W ∗ of the form T (p) = 1 m Tr(p(V1, . . . , Vn)) where m ≥ 1 and V1, . . . , Vn ∈ U(m). Take arbitrary n-tuple of unitary elements U1, . . . , Un in M and put L(p) = τ (p(U1, . . . , Un)) for p ∈ W . Assume that L 6∈ C. By Hahn-Banach theorem there is f ∈ W ∗∗ = W and c ∈ R s.t. Re(L(f )) < c < Re(T (f )) for all T ∈ C. Since e ∈ W we can substitute f − c instead of f and thus assume that c = 0. Since T (f ∗) = T (f ) for every T ∈ C and L(f ∗) = L(f ) we have that L(f + f ∗) = 2Re(L(f )) < 0 < 2Re(T (f )) = T (f + f ∗) which is a contradiction. Thus L ∈ C. Let T be a rational convex combination of elements T1, . . . , Ts from F and Tk corresponds to n-tuples Vj,k. Then T = 1 q (p1T1 + . . . + psTs) for some positive integers p1, . . . , ps, q. Taking block-diagonal Vj = (V ⊗p1 j,s ) we see that T ∈ F . Thus each element of C, in particular element L can be approximated by elements of F . By the Kirchberg's Theorem we have that M can be embedded into Rω. j,1 ⊕ . . . ⊕ V ⊗ps Theorem 10. Connes' embedding conjecture problem has affirmative solu- i uj condition tion iff for any self-adjoint f ∈ F of the form f = αe+Pi6=j αiju∗ Tr(f (V1, . . . , Vn)) ≥ 0 for every m ≥ 1 and every n-tuple of unitary matrices V1, . . . , Vn ∈ U(m) implies that for every ε > 0, εe + f ∼ g with g ∈ Σ2(F ). Proof. If Connes' embedding problem has affirmative solution and quadratic f ∈ Fsa is such that Tr(f (V1, . . . , Vn)) ≥ 0 for every m ≥ 1 and every n-tuple of unitary matrices V1, . . . , Vn ∈ U(m) then by lemma 6 we have τ (f (U1, . . . , Un)) ≥ 0 for any unitary U1, . . . , Un in M. Hence by lemma 3, εe + f is cyclically equivalent to a sum of Hermitian squares. This proves that the conditions of the theorem are necessary. If εe + f is cyclically equivalent to an element in Σ2(F ) for every ε > 0 then clearly τ (f (U1, . . . , Un)) ≥ 0 for any unitary U1, . . . , Un in M. Hence the sufficiency of the theorem conditions follows from Theorem 9. 7 3 The trace-positive quadratic polynomials. adjoint quadratic polynomials f = αe + Pi6=j αiju∗ The results of the preceding section motivate the study of trace-positive self- i uj in unitary generators u1, . . . , un, i.e. polynomials having the property that Tr(f (V1, . . . , Vn)) ≥ 0 for every m ≥ 1 and every n-tuple of unitary matrices V1, . . . , Vn ∈ U(m). If A denotes the matrix α/n α12 . . . α1n α12 α/n . . . α2n . . . . . . . . . . . . α1n α2n . . . α/n     then Tr f (U1, . . . , Un) ≥ 0 can be expressed as positivity of the sum of all entries of the Schur product A ◦ X where X = [tr(U ∗ i Uj)]ij. Thus the trace-positive polynomials f can be characterized as those for which the sum of all entries of A ◦ X for all X ∈ Kn := {[tr(U ∗ m ≥ 1, U1, . . . , Un ∈ U(m)}. Thus our primary objective is to describe the sets Kn ⊆ Mn(C). Note that in the case A is positive semidefinite we have f ∈ Σ2(F ). Indeed in this case A is a sum of rank one positive i Uj)]ij semidefinite matrices A = Ps(βs,1, . . . , βsn)T (βs,1, . . . , βsn) and hence f = Ps(Pj βs,juj)∗(Pj βs,juj). We will also be interested in real analog of the sets Kn, i.e. the sets Kn(R) = Kn ∩ Mn(R). Note that the sets of the traces of monomials of unitary operators and their asymptotic properties in the context of Connes' embedding problem also studied in [10] and [11]. A self-adjoint matrix A such that f = (u−1 n )A(u1, . . . , un)T is defined uniquely except for the diagonal entries. This motivates the following definition. We will call A and B diagonally equivalent and write A d∼ B if A − B is a diagonal matrix with vanishing trace. 1 , . . . , u−1 Definition 11. Let S ⊆ Mn(C) and A ∈ Mn(C) be self-adjoint. We say that A is S-positive and denote A ≥S 0 if there is self-adjoint B such that A d∼ B and bijsij ≥ 0 Xij for all s ∈ S. The three natural choices for S will be Fn = {(tij)tjj = 1 and tij ≤ 1 for all i, j} , 8 Pn ⊂ Fn consisting of positive matrices and the set Kn ⊂ Fn. Clearly, an i uj is self-adjoint matrix A = [aij] is Kn-positive iff f = Pi aiie +Pi6=j aiju∗ a trace positive quadratic polynomial. Note that if then A ≥Fn 0 Tr A ≥ Xi6=j aij 1 , . . . , u−1 and hence A d∼ B for some diagonally dominant matrix B. In this case polynomial f = (u−1 n )A(u1, . . . , un)T is a sum of hermitian squares. However if A ≥Pn 0 then A need not be diagonally equivalent to positive matrix. Note that for the three choices of S mentioned above one can use equality instead of diagonal equivalence since diagonal entries of elements in S equal to 1. The following lemma gives a description of cyclically equivalent quadratic polynomials. Lemma 12. For every matrix A the element (u−1 cyclically equivalent to 1 , . . . , u−1 n )A(u1, . . . , un)T is g−1 k (u−1 1 , . . . , u−1 n )Agk(u1, . . . , un)T gk Xk for any finite collection g1, . . . , gk ∈ F∞ and any matrices Ag such that Agk d∼ A. Xk (4) (5) Any element g ∈ F such that g cyc ∼ f is of the form (4) for some matrices satisfying (5). Moreover for self-adjoint g matrices Ag can also be chosen to be self-adjoint. Proof. The lemma follows from the following easy observation. For any w1 and w2 in F∞ the element w1 − w2 is a commutator ab − ba for some a, b ∈ F∞ if and only if w1 and w2 are conjugated. Hence K consists of finitely supported sums of the form Xj Xk αjkg−1 k wjgk where wj, gk belong to F∞ and Pk αjk = 0 for all j. 9 4 The Clifford Algebras and positive polyno- mials with real coefficients. For a real Hilbert space V there is a unique associative algebra C(V ) with a linear embedding J : V → C(V ) with generating range and such that for all x, y ∈ V J(x)J(y) + J(y)J(x) = 2 hx, yi . (6) The algebra C(V ) is called Clifford algebras associated to V . Clifford algebra can be realized on a Hilbert space such that for every x ∈ V with kxk = 1 operator J(x) is symmetry, i.e. J(x)∗ = J(x) and J(x)2 = I. To see this consider Pauli matrices U = (cid:18) 1 0 −1 (cid:19) , Q(cid:18) 0 1 1 0 (cid:19) . 0 Clearly U and Q are self-adjoint unitary matrices and U 2 = I, Q2 = I, QU + UQ = 0. Then matrices Qj = U ⊗ . . . ⊗ U ⊗ Q ⊗ I ⊗ I . . . are symmetries and {Qi, Qj} = 2δijI. Hence operator J(x) = Pj xjQj is also a symmetry for unit real vector x. For further properties of Clifford algebras we refer to the books [1] and [8]. Theorem 13. For every real correlation matrix P ∈ Mn(R) there is n- tuple of symmetries S1, . . . , Sn in finite dimensional real Hilbert space s.t. P = [tr(S∗ i Sj)]ij. Proof. Every correlation n × n-matrix P is a Gram matrix for a system of unit vectors x1, . . . , xn, i.e. P = [hxi, xji]ij. Taking Clifford symmetries Sj = J(xj) as in the paragraph preceding the theorem we see that P = [tr(S∗ i Sj)]ij. Proposition 14. For every n ≥ 1 the closure Tn(R) of the set of matrices {[τ (U ∗ i Uj)]ijU1, . . . , Un ∈ U(M)} does not depend on real type II1 von Neumann algebra (M, τ ). If self-adjoint f (u1, . . . , un) ∈ F has real coefficients and possess prop- erty that for every n-tuple of unitary matrices U1, . . . , Un ∈ U(m) we have cyc ∼ g for some g ∈ tr(f (U1, . . . , Un)) ≥ 0 then for every ε > 0, εe + f nPm j=1 g∗ j gjm ∈ N, gj ∈ R[F∞]o. 10 Proof. Since every II1 factor contains matrix algebras of arbitrary size we see that Tn(R) coincides with the set of correlation matrices. The last statement follows from Lemma 5. Corollary 15. If quadratic f ∈ F , f (u1, . . . , un) = α +Pi6=j αiju∗ that i uj is such for all unitary matrices U1, . . . , Un then f = 0. Tr(f (U1, . . . , Un)) = 0 Proof. For every k 6= j and t ∈ [0, 1] the matrix P1 = I + (Ekj + Ejk)t is a real correlation matrix. Hence by the theorem there are unitary matrices U1, . . . , Un such that P1 = [tr(U ∗ t Us)]ts. Then the matrix P2 = I + (iEkj − iEjk)t is equal to [tr(V ∗ t Vs)]ts where Vt = Ut for t 6= j and Vj = iUj are unitary matrices. Hence α + (αkj + αjk)t = 0 and α + (αkj − αjk)it = 0. From which follows that α = αkj = 0 and hence f = 0. References [1] O. Bratelli, D. W. Robinson, Operator algebras and quantum sta- tistical mechanics, Springer-Verlag, New York-Heidelberg-Berlin; Vol- ume II, 1981. [2] W. Fulton, Eigenvalues, invariant factors, highest weights and Schu- bert calculus, Bull. Amer. Math. Soc. 37, (2000), 209 -- 249. [3] D. Hadwin, Noncommutative moments problem, Proc. Amer. Math. Soc. 129 (2001), no. 6, 1785 -- 1791. [4] R. Horn, C. Johnson, Matrix Analysis, Cambridge University Press, 1985 [5] K. Juschenko, S. Popovych, Matrix ordered operator algebras, In- diana Univ. Math. J.. 58, (2009), 1203 -- 1218 [6] E. Kirchberg, On nonsemisplit extensions, tensor products and exact- ness of group C ∗-algebras, Invent. Math. 112 (1993), no. 3, 449-489 [7] I. Klep, M. Schweighofer, Connes' Embedding Conjecture and sums of Hermitian squares, Adv. Math., 217(2008), 1816 -- 1837. 11 [8] G. Pisier, Introduction to Operator Space Theory, Cambridge Univer- sity Press, 2003 [9] S. Popovych, Conditions for embedding a ∗-algebra into a C ∗-algebra, Methods of Funct. Analysis and topology. 5, No.3 (1999) 40-48. [10] F. Radulescu, Combinatorial aspects of Connes' Embedding Conjec- ture and asymptotic distribution of traces of products of unitaries, Op- erator theory 20, 197 -- 205, Theta Ser. Adv. Math., 6, Theta, Bucharest, 2006. [11] F. Radulescu, Convex sets associated with von Neumann algebras and Connes' Approximate Embedding problem, Math. Res. Lett. 6 (1999), no. 2, 229 -- 236. [12] I. Vidav, On some ∗-regular rings, Acad. Serbe Sci., Publ. Inst. Math. 13 (1959), 73 -- 80 [13] B. Wang, F. Zhang, A trace inequality for unitary matrices, Amer. Math. Monthly 101 (1994), no. 5, 453 -- 455. 12
1211.6552
3
1211
2013-02-21T15:19:04
Tannaka-Krein duality for compact quantum homogeneous spaces. I. General theory
[ "math.OA", "math.QA" ]
An ergodic action of a compact quantum group G on an operator algebra A can be interpreted as a quantum homogeneous space for G. Such an action gives rise to the category of finite equivariant Hilbert modules over A, which has a module structure over the tensor category Rep(G) of finite dimensional representations of G. We show that there is a one-to-one correspondence between the quantum G-homogeneous spaces up to equivariant Morita equivalence, and indecomposable module C*-categories over Rep(G) up to natural equivalence. This gives a global approach to the duality theory for ergodic actions as developed by C. Pinzari and J. Roberts.
math.OA
math
Tannaka -- Kreın duality for compact quantum homogeneous spaces. I. General theory K. De Commer Department of Mathematics, University of Cergy-Pontoise, UMR CNRS 8088, F-95000 Cergy-Pontoise, France e-mail: [email protected] M. Yamashita Department of Mathematics, Ochanomizu University Otsuka 2-1-1, Bunkyo, 112-8610, Tokyo, Japan e-mail: [email protected] Abstract An ergodic action of a compact quantum group G on an operator algebra A can be interpreted as a quantum homogeneous space for G. Such an action gives rise to the category of finite equivariant Hilbert modules over A, which has a module structure over the tensor category ReppGq of finite-dimensional representations of G. We show that there is a one-to-one correspondence between the quantum G-homogeneous spaces up to equivariant Morita equivalence, and indecomposable module C-categories over ReppGq up to natural equivalence. This gives a global approach to the duality theory for ergodic actions as developed by C. Pinzari and J. Roberts. Keywords: compact quantum groups; C-algebras; Hilbert modules; ergodic actions; module categories AMS 2010 Mathematics subject classification: 17B37; 20G42; 46L08 Introduction In the study of compact group actions on topological spaces, homogeneous spaces play a key role as fundamental building blocks. Ever since the foundational works of I. Gelfand and M. Neumark, the notion of unital C-algebras is known to be a rich generalization of compact topological spaces, and one frequently interprets them as function algebras on (compact) 'quan- tum spaces'. In this more general noncommutative framework, a generally accepted notion of 1 'compact quantum homogeneous space' for a compact group is that of a continuous ergodic action of the group on a unital C-algebra, that is, an action for which the scalars are the only invariant elements. In the same way as compact topological spaces are generalized to unital C-algebras, S.L. Woro- nowicz [37, 39] generalized the notion of compact topological groups to that of compact quantum groups. His axiom system for compact quantum groups is a very simple and natural one involv- ing the coproduct homomorphism dualizing the product map of groups. The resulting theory turns out to be strikingly rich, but at the same time as structured as the classical one. As in the classical case, we have the Haar measure, the Peter -- Weyl theory and the Tannaka -- Kreın duality ([39, 38, 18]). One may also formulate the notion of actions of compact quantum groups on quantum spaces, in a way which respects the Gelfand -- Neumark duality when applied to the continuous map G X Ñ X defining a classical group action. In this framework there is also a natural candidate for the 'quantum homogeneous spaces' over compact quantum groups, by using the formalism of ergodic (co)actions [29, 7]. In this paper, we aim to characterize such quantum homogeneous spaces in the spirit of the Tannaka -- Kreın duality. Such a duality theory for ergodic actions was already developed in [28], where the notion of quasi-tensor functor, a special kind of isometrically lax functor, was used. For practical purposes however, the lack of a strong tensor structure on such a functor makes it difficult to let algebra run its course in computations, due to the appearance of extraneous projections as stumbling blocks. Taking a cue from the theory of fusion categories, we rather formulate a duality theory in terms of module C-categories over the tensor C-category of finite-dimensional representations of G. Indeed, module categories over fusion categories are known to correspond to a good generalized notion of subgroup/homogeneous space (see A. Ocneanu's pioneering work in the subfactor context [25], and more recent developments in the purely algebraic framework [1, 26, 12]). Module C-categories can equivalently, and more concretely, be described in terms of tensor functors into a category of bi-graded Hilbert spaces. This formulation then makes at the same time the connection with the 'fiber functor theory' from [6], which corresponds to non-graded Hilbert spaces and ergodic actions of full quantum multiplicity, and with the theory of [28], which corresponds to considering one particular component of such a graded tensor functor. In the purely algebraic setting, such bi-graded tensor functors also lead to the construction of weak Hopf algebras, i.e. quantum groupoids [15, 16, 11], and Hopf -- Galois actions [32, 33, 30]. The relation with ergodic actions comes by means of a crossed product construction and a Morita theory for quantum groupoids, but we will not further go in to this in this paper. We also mention that a different kind of Tannaka-Kreın duality for classical homogeneous spaces was developed in [17], and for actions on finite quantum spaces in [5, 4] within the framework of planar algebras. Here is a short summary of the contents of the paper. The first two sections will cover prelim- inaries and fix notations. They are meant as an aid for readers who are not familiar with the 2 methodology. In the first section, we will recall the basic concepts concerning compact quantum groups and quantum homogeneous spaces. In the second section, we introduce the necessary prerequisites concerning C-categories, tensor C-categories and module C-categories. Then, in the next five sections, we prove our main results. In the third section, we explain how quan- tum homogeneous spaces lead to indecomposable module C-categories. In the fourth section, we briefly expand on the algebraic content of a general compact quantum group action, so that in the fifth section, we can concentrate on the essential part of the reconstruction of a quantum homogeneous space from an indecomposable module C-category. In the short sixth section we show that this establishes essentially an equivalence between the two notions. In the seventh section, we give further comments on the functoriality of this correspondence. In the appendix, we explain the link between module C-categories and bi-graded tensor functors. It is mainly meant to provide details for, as well as to generalize, the remark which appears in the proof of Theorem 2.5 of [11]. In the accompanying paper [8], we apply the results of the present paper to the case of the compact quantum group SU qp2q. Conventions To have consistency when working with Hilbert C-modules, we will always take the inner product xξ, ηy of a Hilbert space to be linear in η and antilinear in ξ. When ξ and η are vectors in a Hilbert space H , we write ωξ,η for the functional T ÞÑ xξ, T ηy on BpH q. When A and B are C-algebras, A b B denotes their minimal tensor product unless otherwise stated. 1 Compact quantum groups and related structures 1.1 Compact quantum groups Definition 1.1 ([39]). A compact quantum group G consists of a unital C-algebra CpGq and a faithful unital -homomorphism ∆ : CpGq Ñ CpGq b CpGq satisfying the coassociativity condition p∆ b idq ∆ " pid b∆q ∆ and the cancelation condition r∆pCpGqqp1 b CpGqqsn-cl " CpGq b CpGq " r∆pCpGqqpCpGq b 1qsn-cl, where n-cl means taking the norm-closed linear span. We recall from [39] that any compact quantum group admits a unique positive state ϕG which satisfies pid bϕGqp∆pxqq " ϕGpxq1 " pϕG b idqp∆pxqq, x P CpGq. (1.1) This state is called the invariant state (or the Haar state) of CpGq. Definition 1.2. The compact quantum group G is called reduced if the invariant state ϕG is faithful. 3 In the rest of the paper, we will always work with reduced compact quantum groups. This is no serious restriction, as to any G one can associate a reduced companion which has precisely the same representation theory as G. Definition 1.3. A unitary corepresentation u of CpGq on a Hilbert space Hu is given by a unitary element u of BpHuq b CpGq satisfying the multiplicativity condition pid b∆qpuq " u12u13 P BpHuq b CpGq b CpGq, where the leg numbering indicates at which slot in a multiple tensor product one places the element, filling the blank spots with units. A unitary corepresentation u is said to be finite- dimensional when Hu is so. When u and v are unitary corepresentations of CpGq, an operator T P BpHu, Hvq is said to be an intertwiner between u and v if it satisfies vpT b 1q " pT b 1qu. A unitary corepresentation u is called irreducible if the space of intertwiners from u to itself is one-dimensional. In what follows we will refer to unitary corepresentations of CpGq as unitary representations of G. 1.2 Quantum homogeneous spaces Definition 1.4 ([7, 29]). Let G be a compact quantum group. An action of G on a unital C-algebra A is a faithful unital -homomorphism α : A Ñ A b CpGq satisfying the coaction condition pid b∆q α " pα b idq α and the density condition rp1 b CpGqqαpAqsn-cl " A b CpGq. We call the action ergodic if the space AG " tx P A αpxq " x b 1u is equal to C1. If pA, αq is an ergodic action, we will use the notation A " CpXq, and refer to the symbol X as the quantum homogeneous space. If X is a quantum homogeneous space for G, then CpXq carries a canonical faithful positive state ϕX, determined by the identity pid bϕGqpαpxqq " ϕXpxq1 px P CpXqq. It is the unique state on CpXq which is α-invariant, pϕX b idqαpxq " ϕXpxq1 for all x P CpXq. 4 2 C-categories 2.1 Semi-simple C-categories Definition 2.1 ([13]). A C-category D is a C-linear category whose morphism spaces are Banach spaces satisfying the submultiplicativity condition }ST } ď }S}}T } for composition of morphisms S and T , and admitting antilinear 'involutions' : MorpX, Y q Ñ MorpY, Xq, T ÞÑ T , which behave contravariantly and satisfy the C-condition }T T } " }T }2 for each morphism T . A linear functor between two C-categories is called a C-functor if it preserves the -operation. Remark 2.2. Let D and D1 be C-categories. Let FunpD, D1q be the category • whose objects are the C-functors from D to D1, and • whose morphisms between two functors F, G : D Ñ D1 consist of the natural transforma- tions φ‚ " pφX : F X Ñ GXqXPD such that p}φX}qXPD is uniformly bounded. Then FunpD, D1q is a C-category with the norm }φ‚} " supXPD }φX} and the involution pφqX " pφXq. Definition 2.3 ([13]). We say that an object X in a C-category D is simple if MorpX, Xq is isomorphic to C. We call D semi-simple [23, Section 1.6] if D admits finite direct sums and if any of its objects is isomorphic to a finite direct sum of simple objects. Remark 2.4. A C-category D is semi-simple if and only if all morphism spaces are finite- dimensional and 'idempotents split'. The latter condition means that any self-adjoint projection p P MorpX, Xq is of the form vv for some isometry v P MorpY, Xq. Furthermore, a semi-simple C-category also has a zero object 0, i.e. an object which is both initial and terminal. Definition 2.5. Let J be a set, and D a semi-simple C-category. We say that D is based on J if we are given a bijection between J and a maximal family of mutually non-isomorphic simple objects in D. We then write Xr for the simple object associated with r P J. By definition, any object X in a semi-simple C-category D based on J is isomorphic to a direct sum 'rPJ mrXr. The integer mr is called the multiplicity of Xr in X, and is uniquely determined by mr " dimpMorpXr, Xqq. Then for any object X and any irreducible Xr, the complex vector space MorpXr, Xq admits a natural structure of Hilbert space by the inner product xS, T y " ST P MorpXr, Xrq " C. Examples of semi-simple C-categories will be presented in Section 3 and the appendix. They can be seen as categorified versions of Hilbert spaces, cf. the slightly different context of [3]. As with Hilbert spaces, there is essentially only one semi-simple C-category for each cardinal number, the cardinality of the set of isomorphism classes of irreducible objects in the given semi-simple C-category, cf. Lemma A.1.6. However, true to this analogy, they arise in various 5 presentations in practical situations, from concrete to abstract. For the moment, it will suffice to have the following characterization of equivalences between semi-simple C-categories. Lemma 2.6. Let D and D1 be semi-simple C-categories, with D based on an index set J. Let F be a C-functor from D to D1. Then F is an equivalence of categories if and only if the set tF pXrq r P Ju forms a maximal set of mutually non-isomorphic irreducible objects in D1. Proof. The necessity of the condition is obvious. Let us see that it is also sufficient. Let X be an irreducible object of D and let m be a nonnegative integer. Then the C-algebra EndpmXq is isomorphic to MmpCq, where the identity morphisms of the direct summands form a partition of unity by mutually equivalent minimal projections. Since F pXq is also an irreducible object, it follows that F induces a C-algebra isomorphism between EndpmXq and EndpF pmXqq -- EndpmF pXqq. More generally, given a finite direct sum X " 'rPJ mrXr, we can conclude that F provides an isomorphism between EndpXq and EndpF pXqq. Finally, by considering this argument for X ' Y , we conclude that F gives a bijection from MorpX, Y q to MorpF pXq, F pY qq for any objects X, Y , that is, F is a fully faithful functor. As the set tF pXrq r P Ju forms a maximal set of mutually non-isomorphic irreducible objects in D1, we also have that F is essentially surjective. From [21, Theorem IV.4.1], we conclude that F is an equivalence. 2.2 Tensor C-categories Definition 2.7. [10] A (strict) tensor C-category C " pC, b, 1q consists of a C-category C together with a bilinear C -functor b : C C Ñ C and an object 1 P C such that there are equalities of functors ´ b p´ b ´q " p´ b ´q b ´, 1 b ´ " idC " ´ b 1. The 'strictness' condition refers to the on the nose associativity of b. In most examples which arise in practice, the associativity only holds up to certain coherence isomorphisms [21, Chapter VII]. But for the cases we will encounter, the coherence isomorphisms will be obvious and one can safely ignore them. Also for abstract tensor categories, one can almost always restrict oneself to the setting of strict tensor categories by Mac Lane's coherence theorem [21, Section VII.2]. This coherence result holds as well on the C-level. Definition 2.8 ([10, 20]). Let C be a tensor C-category. An object U in C is said to admit a conjugate or dual if there exists a triple p ¯U , RU , ¯RU q with ¯U P C and pRU , ¯RU q a couple of morphisms RU : 1 Ñ ¯U b U, ¯RU : 1 Ñ U b ¯U satisfying the conjugate equations p ¯R U b idU qpidU bRU q " idU , pR U b id ¯U qpid ¯U b ¯RU q " id ¯U . (2.1) 6 The full subcategory of all objects in C admitting duals is denoted by Cf. A tensor C-category C is called rigid if C " Cf. Remarks 2.9. 1. [20, Theorem 2.4] When U and V are in Cf, the product ¯V b ¯U of their duals is in duality with U b V . Moreover, if p ¯U , RU , ¯RU q makes a dual for U, then pU, ¯RU , RU q makes a dual for ¯U. It follows that Cf is a rigid C-tensor subcategory of C. If pRU , ¯RU q satisfy the conjugate equations, then for any λ P C also pλRU , ¯λ´1 ¯RU q satisfy the same equations. When the unit of C is irreducible, then for U irreducible and ¯U a fixed dual, this is the only arbitrariness in the choice of pRU , ¯RU q. 2. For any U, the object ¯U , when it exists, is unique up to isomorphism. 3. When the unit of C is irreducible, then for any irreducible U with dual ¯U, one can always arrange for a solution pRU , ¯RU q of the conjugate equations which is normalized, i.e. such that R U RU is a strictly positive real number which is uniquely determined by U. It is called the quantum dimension of U. ¯RU . Then by the above scaling result, dimqpUq " R U RU " ¯R U Examples 2.10. 1. The category of all Hilbert spaces and bounded maps is a tensor C- category for the ordinary tensor product of Hilbert spaces. The maximal rigid subcategory consists of all finite-dimensional Hilbert spaces. If H is a finite-dimensional Hilbert space, the complex conjugate space H can be taken as its conjugate object, where the maps RH and ¯RH are given by R H : H b H Ñ C, ¯ξ b η Ñ xξ, ηy, R H : H b H Ñ C, ξ b ¯η Ñ xη, ξy. 2. For any compact quantum group G, the category ReppGq of its finite-dimensional unitary representations together with the intertwiners forms a rigid tensor C-category with irre- ducible unit object. The tensor product u lT v of two representations u and v is defined to be the representation on Hu b Hv given by the unitary u13v23 P BpHuq b BpHvq b CpGq. When u is an object of ReppGq, its dual can be given by a unitarization of pj b idqpu´1q P BpHuq b CpGq, where j : BpHuq Ñ BpHuq is the natural anti-isomorphism character- ized by jpT q ¯ξ " T ξ. Unlike the case of Hilbert spaces or compact groups, u lT v is not isomorphic to v lT u in general. 3. [10, 36] For a fixed C-category D, let EndpDq denote the category of C-endofunctors, cf. Remark 2.2. Then EndpDq is a tensor C-category, with the b-structure F b G " F G given by the composition of endofunctors, and with the identity functor providing the unit. The associated rigid category EndpDqf consists of adjointable functors whose unit and co-unit maps are uniformly bounded. We recall the notion of strong tensor functor and tensor equivalence. Definition 2.11. Let C1 and C2 be two tensor C-categories. A strong tensor C-functor from C1 to C2 consists of a C-functor F : C1 Ñ C2 together with natural unitary transformations ψU,V : F pUq b F pV q Ñ F pU b V q, c : 1C2 Ñ F p1C1q, 7 satisfying certain coherence conditions [24, Section 1.2]. It is called a tensor equivalence if the underlying functor F is an equivalence. Example 2.12. If G is a compact quantum group, there is a natural forgetful functor from ReppGq to Hf, sending each unitary representation u to the underlying Hilbert space Hu, and acting as the identity on intertwiners. The natural transformations ψ and c are identity maps. In general, there can exist other faithful strong tensor C-functors from ReppGq to Hf besides this canonical one, cf. [6], but each one of them determines a unique compact quantum group ([38]). The following lemma will be used at some point. Lemma 2.13 ([20]). Let C1 and C2 be tensor C-categories, and F : C1 Ñ C2 a strong tensor C-functor. If C1 is rigid, then the image of F is contained in pC2qf. Proof. If U P C1, then the compatibility of F with the tensor products can be used to construct a duality between F pUq and F p ¯U q. Hence the image of F is inside pC2qf. 2.3 Module C-categories Definition 2.14. Let C be a tensor C-category with unit object 1, and D a C-category. One says that D " pD, M, φ, eq is a C-module C-category if M : C D Ñ D is a bilinear -functor with natural unitary transformations φ : Mpp´ b ´q, ´q „Ñ Mp´, Mp´, ´qq, e : Mp1, ´q „Ñ id, satisfying certain obvious coherence conditions, cf. [27], which we will spell out below. We say that D is semi-simple if the underlying C-category is semi-simple. We say that D is indecomposable or connected if, for all non-zero X, Y P D, there exists an object U P C such that MorpMpU, Y q, Xq ‰ 0. In the following, we will use the more relaxed notation U b X for MpU, Xq, and similarly for morphisms. The coherence conditions can then be written in the following form, as the commutation of the diagrams pU b V b W q b X φU,V bW,X U b ppV b W q b Xq (2.2) φU bV,W,X idU bφV,W,X pU b V q b pW b Xq φU,V,W bX / U b pV b pW b Xqq, 8 / /     / and U b X φU,1,X 7♦♦♦♦♦♦♦♦♦♦♦ '❖❖❖❖❖❖❖❖❖❖❖ φ1,U,X U b p1 b Xq idU bX idU beX '❖❖❖❖❖❖❖❖❖❖❖ 7♦♦♦♦♦♦♦♦♦♦♦♦ eU bX / U b X. 1 b pU b Xq (2.3) Examples 2.15. 1. Let D be a C-category. Then D is a module C-category for EndpDq in an obvious way. 2. Let G be a compact (quantum) group and H be a closed (quantum) subgroup of G. Then ReppHq is a ReppGq-module C-category in a natural way: the action of π P ReppGq on θ P ReppHq is defined as πH b θ. In other words, this is induced by the restriction functor ReppGq Ñ ReppHq, which is a strong tensor C-functor. 3. More generally, if C1 and C2 are tensor C-categories, and F a strong tensor C-functor from C1 to C2, then C2 becomes a C1-module C-category by the association MpX, Y q " F pXq b Y . We will need the following interplay between dual objects and the module structure. Lemma 2.16. Let C be a rigid tensor C-category, and let D be a C-module C-category. For any U in C and any objects X, Y in D, we have an isomorphism MorpU bY, Xq -- MorpY, ¯UbXq, called the Frobenius isomorphism associated with pRU , ¯RU q. Proof. This can be proved by a standard argument involving the conjugate equations, cf. Propo- sition A.4.2. The appropriate notion of morphisms between module C-categories is the following. Definition 2.17. Let D and D1 be module C-categories over a fixed tensor C-category C. A C-module homomorphism from D to D1 is given by a pair pG, ψq, where G is a functor from D to D1 and ψ is a unitary natural equivalence Gp´ b ´q Ñ ´ b G´, such that the diagrams of the form Gp1 b Xq ψ1,X / 1 b GX w♣♣♣♣♣♣♣♣♣♣♣♣ e Gpeq GX (2.4) 9 ' 7 ' / 7   / w and (2.5) ψU,V bX 4✐✐✐✐✐✐✐✐✐✐✐✐✐✐✐✐✐ GpU b pV b Xqq GpφU,V,X q GppU b V q b Xq commute. U b GpV b Xq idU bψV,X *❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯❯ U b pV b GXq ψU bV,X φU,V,GX / pU b V q b GX An equivalence between D and D1 is a morphism pG, ψq for which G is an equivalence of categories. The following section is dedicated to the ReppGq-module C-categories which are the star actors of this paper. 3 Equivariant Hilbert modules Definition 3.1 ([2]). Let X be a quantum homogeneous space for a compact quantum group G. An equivariant Hilbert C-module E over X is a right Hilbert CpXq-module E, carrying a coaction αE : E Ñ E b CpGq, where the right hand side is the exterior product of E with the standard right Hilbert CpGq-module CpGq, satisfying the density condition rp1 b CpGqqαEpEqsn-cl " E b CpGq " rαEpEqp1 b CpGqqsn-cl and the compatibility conditions 1. @x P CpXq, @ξ P E : αEpξ xq " αEpξqαXpxq, 2. @ξ, η P E : xαEpξq, αEpηqyCpXqbCpGq " αXpxξ, ηyCpXqq. Remark 3.2. An equivariant Hilbert C-module is necessarily saturated, and in particular faithful as a right CpXq-module. Indeed, otherwise the closed linear span of txξ, ηyCpXq ξ, η P Eu would give a proper equivariant closed 2-sided ideal I in CpXq. But any invariant state on CpXq{I would induce a non-faithful invariant state over CpXq, which is a contradiction. To any equivariant Hilbert CpXq-module one can associate a special unitary which implements the coaction. Definition 3.3. Let X be a quantum homogeneous space for a compact quantum group G, and E an equivariant Hilbert C-module over X. One defines the associated unitary morphism XE P LCpXqbCpGq pE bαX pCpXq b CpGqq, E b CpGqq 10 * 4     / by the formula XEpξ b px b hqq " αEpξqpx b hq. Example 3.4. Consider a set ‚ with one element, and consider Cp‚q " C with the trivial right action αtriv : Cp‚q Ñ Cp‚q b CpGq, 1 Ñ 1 b 1. Then an equivariant Hilbert C-module over ‚ is nothing but a representation of G. Indeed, a right Hilbert Cp‚q-module is just a Hilbert space H . Then the receptacle of the unitary operator in Definition 3.3 can be identified with BpH q b CpGq. This gives the correspondence of the equivariant Hilbert C-modules over ‚ and the unitary representations of G. We will denote the equivariant Hilbert space associated to u as pHu, δuq. We will be particularly interested in a subcategory of equivariant Hilbert C-modules which admit a nice decomposition into irreducible objects. Definition 3.5. An equivariant Hilbert C-module E is called • finite if it is finitely generated projective as a right CpXq-module, and • irreducible if the space LGpEq " tT P LpEq αEpT ξq " pT b 1qαEpξq for all ξ P Eu is one-dimensional. Any irreducible equivariant Hilbert C-module is finite in the above sense, as seen in the next proposition. Proposition 3.6. An equivariant C-module is finite if and only if the C-algebra LGpEq is finite-dimensional. Proof. Let XE be the unitary morphism associated with αξ as in Definition 3.3. Then, the map x ÞÑ XEpx bαξ 1qX E defines a coaction of CpGq on LpEq, and the ideal of compact endomorphisms is a G-invariant subalgebra [2]. Moreover, LGpEq is precisely the G-fixed point subalgebra of LpEq. First, let us prove that an equivariant module over X is finitely generated projective over CpXq when LGpEq is finite-dimensional. We can reduce it to the case of LGpEq " C by taking a decomposition associated with a partition of unity by minimal projections in LGpEq. Then, taking any non-zero positive compact endomorphism x of E, we see that pid bϕGqpXαpx bαE 1qX αq is simultaneously compact and nonzero positive scalar in LpEq. Hence E is finitely generated projective over CpXq [19, Lemma 6.5]. Conversely, suppose that we are given a finitely generated projective CpXq-module E admitting a compatible corepresentation of CpGq. Then, the crossed product module E ¸ G, which is finitely generated projective over CpXq ¸ G, admits a natural faithful representation of LGpEq as CpXq ¸ G-module homomorphisms. By the ergodicity of G on X, we know that CpXq ¸ G is a direct sum of algebras of compact operators [7]. Hence, for any finitely generated projective module over CpXq ¸ G, the module 11 endomorphisms must form a finite-dimensional algebra. This implies that LGpEq is finite- dimensional. In particular, any irreducible equivariant Hilbert C-module E over CpXq gives another quantum homogeneous space LpEq " KpEq, by the action as given in the beginning of the above proof. Definition 3.7. A quantum homogeneous space Y is called equivariantly Morita equivalent to X if there exists an irreducible equivariant Hilbert C-module E over CpXq and an equivariant C-algebra isomorphism CpYq Ñ KpEq. We say that such an equivariant Hilbert module E and associated isomorphism implement the Morita equivalence. Note that the above terminology is justified by Remark 3.2. Notation 3.8. Let G be a compact quantum group, and X a quantum homogeneous space over G. We let DX denote the category of finite equivariant Hilbert C-modules over X, whose morphisms are the equivariant adjointable maps between Hilbert C-modules. Proposition 3.9. The category DX is a semi-simple C-category. Proof. By the above proposition, for any object E in DX, the algebra MorpE , Eq is a finite- dimensional C-algebra. Moreover, if p P MorpE , Eq is a projection, then pE is again an object of DX. Remark 2.4 then implies the assertion. In view of Example 3.4, it can be seen that finite (resp. modules play a similar role as the finite-dimensional (resp. irreducible) representations of G. irreducible) equivariant Hilbert C- Now let E be a finite equivariant Hilbert CpXq-module, and let u be a finite-dimensional unitary representation of G. Then we can amplify E with u to obtain the equivariant Hilbert module u lT E. As a Hilbert CpXq-module, u lT E is the amplification Hu b E of E with the Hilbert space Hu. The coaction of CpGq is given by the formula pu lT αE qpξ b ηq " u13pξ b αpηqq, Then obviously u lT E is still finite. We record the following facts for later reference. Lemma 3.10. For any E P DX, there exists a representation u of G for which there is an isometric morphism of E into u lT CpXq. Proof. This is a consequence of the equivariant stabilization, see Section 3.2 of [34]. Proposition 3.11. Let X be a quantum homogeneous space for a compact quantum group G. Denote by DX the C-category of finite equivariant Hilbert CpXq-modules. Then the operation ReppGq DX Ñ DX, pu, Eq ÞÑ u lT E defines a connected ReppGq-module C-category structure on DX. 12 Proof. The maps necessary to complete the ReppGq-module category structure are obvious, coming from the ordinary associativity maps for the concrete tensor products of the underlying Hilbert spaces and Hilbert C-modules. Let us prove that DX is connected over ReppGq. Let E and F be arbitrary objects in D. By Lemmas 3.10 and 2.16, we can find a representation u such that CpXq appears inside u lT E. Then, again by Lemma 3.10, we can a suitable representation v such that Morpv lT E , F q ‰ 0. Hence D is connected. Remark 3.12. The equivariant K-group K G 0 pCpXqq is a free abelian group generated by the ir- reducible classes of DX. Note that for compact groups, the above picture was already presented, modulo some of the terminology, in [35, section 9]. Its extension to the compact quantum group setting was treated in [31]. We aim to show in the next sections that the module C-category DX, together with the distinguished element corresponding to the standard Hilbert C-module CpXq, remembers the quantum homogeneous space X. 4 An algebraic approach to quantum group actions In this section, we will provide a characterization of quantum homogeneous spaces and equiv- ariant Hilbert modules with the analysis drained out of it. This intermediate step will make the Tannaka -- Kreın machine of the next section run more smoothly. The main argument provides an algebraic description of an arbitrary action of a compact quantum group G. It is based on results which appear already in [7, 29]. We first recall the notion of Hopf -algebra associated with a compact quantum group. Definition 4.1. [39] Let G be a compact quantum group. If u is a finite-dimensional unitary representation of G, the elements pid bωξ,ηqpuq P CpGq for ξ, η P Hu are called the matrix coefficients of u. The set of all such elements with the u ranging over the representations of G form a dense Hopf -subalgebra PpGq Ď CpGq. Definition 4.2. Let G be a compact quantum group. Let A be a unital -algebra. An algebraic action of G on A is defined to be a Hopf -algebra coaction αA : A Ñ A b PpGq, the tensor product on the right being the algebraic one, such that A G is a unital C-algebra, and such that the following positivity condition is satisfied: The map x ÞÑ EGpxq " pid bϕGqαpxq P A G is completely positive on A . (P) To be clear, the complete positivity means that for any n P N and any element a P A b MnpCq, the element pEG b idqpaaq is a positive element in the C-algebra A G b MnpCq. 13 Lemma 4.3. Let G be a compact quantum group with an action αA on a unital C-algebra A. Let A denote the linear span of pid bϕGqpαApxqp1 b gqq for x P A and g P PpGq. Then A is a dense unital -subalgebra of A on which αA restricts to an algebraic action. Proof. See [29, Theorem 1.5], and [7, Lemma 11 and Proposition 14], whose proofs do not depend on the ergodicity assumption made there. The complete positivity of EG follows from the way it is defined in (P); namely, -homomorphisms, states, their amplifications, and their compositions are completely positive. Proposition 4.4. Let G be a compact quantum group with an algebraic action αA on a unital -algebra A . Then there exists a unique C-completion A of A to which αA extends as a coaction of CpGq. Moreover, AG " A G. Proof. We denote by B the C-algebra A G. By the complete positivity assumption on EG, the B-valued inner product xa, byB " EGpabq on A gives a pre-Hilbert B-module structure. We want to show that the left representation of A on itself by left multiplication extends to the Hilbert module completion A. Let a be an arbitrary element of A . Since the image of αA ends up in the algebraic tensor product of A and PpGq, there is a finite-dimensional unitary representation u of G and an intertwiner from ¯u to A whose image contains a. Let us choose an orthonormal basis ei of Hu, and put uij " pωei,ej b idqpuq. Then, the above statement means that there are elements ai P A such that • a can be written as a linear combination ři λiai, and • the elements ai transform according to pu jiq, so αA paiq " řj aj b u ji. i ai P B. bining the inequalities a that The unitarity of u implies that ři a Since B is a C-algebra, one has the inequality ři a j aj ď ři a i ai›››B EG`ba j ajb ď ›››ÿi a i ai}B. Fix now some j. Com- i ai in A with the previous one, the positivity of EG implies i ai ď }ři a EGpbbq, @b P A . It follows that left multiplication with each aj is bounded, so that a extends as a left multipli- cation operator to A. We obtain in this way a faithful -representation A Ñ LBpAq. Define A to be the norm- completion of A in this representation. We claim that the coaction αA extends to A. Consider the transformation X on A b PpGq defined by Xpa b gq " αA paqp1 b gq. Then, the invariance of ϕG implies that X extends to a unitary morphism on the right Hilbert B-module Ab L 2pGq. By a routine computation we obtain that a ÞÑ Xpa b 1qX for a P A gives the extension αA of αA to A. From this formula for αA, it also follows that we have pid bϕGqαpaq " xa 1B, 1ByB for all a P A. It follows that the invariant elements of A lie in B. 14 It remains to prove the uniqueness of A. Let us assume that A is an arbitrary unital C-algebra satisfying the conclusion of the lemma. Then EG can, by the same formula, be extended to a conditional expectation from A to B. Since G is reduced, this conditional expectation is faithful. Now, if a P A Ď A and R ă }a}, the functional calculus shows that there is a positive element b P A such that pRbq2 ă baab. Thus, the norm of a can be characterized by }a} " sup bPA zt0u}EGpbaabq} }EGpbbq} 1 2 . Hence the C-norm on A is uniquely determined in terms of pA , αA q. Proposition 4.5. Let G be a compact quantum group. Then the correspondences A ÞÑ A and A ÞÑ A of Lemma 4.3 and Proposition 4.4 can be extended to respective functors Alg and Comp between the categories of actions of G and algebraic actions of G. Moreover, Comp Alg is naturally equivalent to the identity functor. Here, the morphisms on the respective categories are understood to be the equivariant unital -homomorphisms. Proof. Let A and B be unital C-algebras endowed with G-actions, and let f : A Ñ B be an equivariant unital -homomorphism. The equivariance implies that f restricts to an equivariant -homomorphism A Ñ B. This gives the functor Alg. Conversely, suppose that A and B are unital -algebras with algebraic G-actions, A and B their respective completions. Then the direct sum A'B admits a canonical G-action extending the ones on the direct summands. If f : A Ñ B is an equivariant unital -homomorphism, the map pid f qpaq " a ' f paq is a faithful G-equivariant homomorphism from A to A ' B. Proposition 4.4 implies that the C-norm on A induced by id f has to agree with the A-norm. Hence f extends to an equivariant -homomorphism A Ñ B. This way we obtain the functor Comp. Now, the natural equivalence between Comp Alg and the identity functor follows directly from density part in Lemma 4.3 and the uniqueness part in Proposition 4.4. Remark 4.6. The composition Alg Comp is not equivalent to the identity functor in gen- eral. For example, if A is given by the function algebra of closed disk Cp ¯Dq endowed with the rotation action of Up1q, the algebra A contains many Up1q-invariant norm dense subalgebras corresponding to the various decaying conditions around the origin. However, on the subcate- gory of the actions with finite-dimensional fixed point algebras, Alg Comp is indeed equivalent to the identity functor. 15 5 Tannaka -- Kreın construction Let G be a compact quantum group. We take a set I indexing the equivalence classes of irreducible objects in ReppGq, and a distinguished irreducible object ua for each a P I. When convenient, we will abbreviate ua by a. The index corresponding to the unit object of ReppGq will be written as o. We identify Ho with C (canonically) by means of the tensor structure. It will be handy to use the following Penrose -- Einstein-like notation. It concerns the natural map àaPI Morpua, uq b Ha Ñ Hu, ÿi xi b ξi ÞÑ ÿi xipξiq (5.1) for any representation u. This map is an isomorphism, see Lemma A.1.4. Notation 5.1. We will write the inverse of (5.1) as ξ ÞÑ ξa b ξa, so that ξ " ξaξa. For the rest of this section, we will fix a semi-simple ReppGq-module C-category D. Notation 5.2. For objects x, y in D, we denote by A y x the vector space A y x " àaPI Morpua b y, xq b Ha The direct sum on the right hand side is the algebraic one. We can endow A y PpGq-comodule structure αy x " 'apid bδaq, where δa is defined in Example 3.4. x with the Remark 5.3. The space 'a Morpua b y, xq b Ha may be seen as the coend of the functor Cop C Ñ Vect sending pu, vq to Morpu b y, xq b Hv, see for instance [23, Section 2], [21, Chapter IX]. Our goal is to make the A y y into algebraic actions for G, and the A y pre-Hilbert modules for A y y . x into equivariant right Notation 5.4. When f stands for an element in A y x , its leg in Morpua b y, xq (resp. in Ha) for a P I is denoted by f a (resp. fa). Thus, the expression of the form f a b fa is understood to represent f . We will combine this notation with Notation 5.1. This notation can be seen as analogous to the Sweedler notation for coproducts. As an example, consider fixed a, b P I, and elementary tensors f " x b ξ and g " y b η respectively in Morpua b y, xq b Ha and Morpub b y, xq b Hb. Choose a maximal family of mutually orthogonal isometric morphisms pιc ab,kqk from uc to ua lT ub. Then we have f c b gd b pfc b gdqe b pfc b gdqe " ÿc,k x b y b ιc ab,k b pιc ab,kqpξ b ηq inside Àc Morpua b y, xq b Morpub b y, xq b Morpuc, ua b ubq b Hc. As an exercise to get acquainted with the notation, the reader could try to prove the following interchange law rpξa b idvqpξa b ηqcs b pξa b ηqc " pξ b ηqc b pξ b ηqc -- ξ b η, 16 where ξ, η are arbitrary vectors respectively in Hu and Hv. Definition 5.5. Let x, y, z be objects in D. We define a multiplication map A y x A z y Ñ A z x (5.2) by the formula f g " pf gqc b pf gqc " rf apida bgbqφa,b,zppfa b gbqc b idzqs b pfa b gbqc. where φa,b,z is the associator from Definition 2.14. Proposition 5.6. The multiplication (5.2) is associative. Proof. Let pf, g, hq P A y x A z y A w z . First, the product pf gqh can be expressed as rrf apida bgbqφa,b,zppfa b gbqc b idzqspidc bhdqφc,d,wpppfa b gbqc b hdqe b idwqs b ppfa b gbqc b hdqe. Taking composition at c and using naturality of φ, the above is equal to rf apida bgbqpida b idb bhdqφa,b,dbwφa lT b,d,wppfa b gb b hdqe b idwqs b pfa b gb b hdqe. Similarly, the expression f pghq reduces to rf apida bgbqpida b idb bhdqpida bφb,d,wqφa,b lT d,wppfa b gb b hdqe b idwqs b pfa b gb b hdqe. The conclusion then follows from the associativity constraint on φ. Proposition 5.7. Let x and y be objects in D, and let ey P Morpuo b y, yq be the structure map of tensor unit included in the module package. Then the element 1y " ey b 1 P A y y is a right unit for the multiplication map A y x , and a left unit for the multiplication map A y y Ñ A y x A y y A x y Ñ A x y . Proof. Take f P A x y . Then the formula for the product f 1y reads rf apida beyqφa,o,yppfa b 1qc b idyqs b pfa b 1qc. Since Morpuc, uo lT uaq ‰ 0 if and only if a " c for a, c P I, the unit constraint on e reduces this expression to f apida b idyq b fa " f . This shows that 1y is a left unit. An analogous argument shows that 1y is also a left unit. It follows that we can make a category A having the same objects as D, and with morphism space from x to y the linear space A x y are unital algebras. It contains D as a faithful sub-category, as shown by the following lemma. y . In particular the 'endomorphism spaces' A y Lemma 5.8. There is a linear functor D Ñ A which is the identity on objects, and which sends f P Morpy, xq to f ey b 1 P A y x . 17 Proof. This is proven in the same way as Proposition 5.7. In the following, we will identify Morpy, xq with its image inside A y x . x and g P A z Proposition 5.9. Take x, y, z objects in D. Let f P A y y . Then αz xpf gq " αy xpf qαz ypgq. Proof. When pE, αq a right comodule over PpGq, let us write, for x P E, Then, resorting again to the notation of Example 3.4, one has αpxq " xp0q b xp1q P E b PpGq. δu lT vpξ b ηq " ξp0q b ηp0q b ξp1qηp1q. Using that ξc b δcpξcq " pξp0qqc b pξp0qqc b ξp1q, the element αz xpf gq can thus be computed as rf apida bgbqφa,b,zpfa b gbqcs b δcppfa b gbqcq " rf apida bgbqφa,b,zpfap0q b gbp0qqcs b pfap0q b gbp0qqc b fap1qgbp1q. On the other hand, the way the coaction αy x is defined implies that f a b fap0q b fap1q " pfp0qqa b pfp0qqa b fp1q. It follows that αz xpf gq can be expressed as rpfp0qqapida bpgp0qqbqφa,b,zppfp0qqa b pgp0qqbqcs b ppfp0qqa b pgp0qqbqc b fp1qgp1q, which is precisely αy xpf qαz ypgq. We will now define a -operation A y so we first fix our conventions concerning duals. x Ñ A x y . Here the rigidity of ReppGq will come into play, Notation 5.10. When f u P Morpu b y, xq, we write fu P Morpy, ¯u b xq for its image of the Frobenius isomorphism associated with pRu, ¯Ruq (see Lemma 2.16). So, fu " pid¯u bf uqφ¯u,u,ypRu b idyqe y . Similarly, when ξu P Hu, we define ξu P H¯u by the formula ξu " pξ u b iduq ¯Rup1q, where ξ for a vector ξ P H is the obvious map H Ñ C. Definition 5.11. We define the anti-linear conjugation map : A y x Ñ A x y by f " pf q¯a b pf q¯a " p fa q b fa . 18 Since the above formula involves both Ra and ¯R independent of the choice of the duality morphisms. a for each a P I, the definition of is actually Proposition 5.12. The operation is anti-multiplicative. Proof. Let f P A y x and g P A z y . Then by definition of the product, pgf q¯c b pgf q¯c " rpgq¯apid¯a bpf q ¯bqφ¯a,¯b,xpppgq¯a b pf q¯bq¯c b idxqs b ppgq¯a b pf q¯bq¯c " rp ga qpid¯a bp fb qqφ¯a,¯b,xpp ga b fb q¯c b idxqs b p ga b fb q¯c. Let us concentrate first on the part φ ¯a,¯b,xpid¯a b fb q ga . Choose as solution for the conjugate equations for b lT a the couple ppid¯a bRb b idaqRa, pida b ¯Rb b id¯aq ¯Raq. Then, using naturality and coherence for φ and e, we can write, after some diagram manipulations, φ ¯a,¯b,xpid¯a b fb q ga " pid¯a lT ¯b bpf bpidb bgaqqqφ¯a lT ¯b,b,abzφ¯a lT ¯b lT b,a,zpRb lT a b idzqe z . Substituting in the expression for gf and pulling through the factor ppgqa b pf qbqc b idx, we find that gf is equal to the expression rezpR b lT a b idzqpp ga b fb q¯c b idb lT a b idzqφ ¯cbb,a,zφ ¯c,b,abzpid¯c bppidb bgaqf bqqs b p ga b fb q¯c. Now for vectors ξ and η in representation spaces, we have rR b lT app ξa b ηb q¯c b idb b idaqs b p ξa b ηb q¯c " rR c pid¯c bppηb b ξaqcqqs b pc ηb b ξaq, which can be verified using the natural isomorphism and the conjugate equations for pR, ¯Rq. It follows that gf can be written as 'c Morp¯c lT b lT a, 1q b H¯c Ñ H b b H a rezpR c b idzqpid¯c bppfb b gaqcq b idzqφ ¯cbb,a,zφ ¯c,b,abzpid¯c bppidb bgaqf bqqs b pc fb b gaq. Using once more coherence and naturality for φ, this reduces to pf gq. Proposition 5.13. The operation is involutive. Proof. Let f P A y x . By the definition of the -operation, pf q can be written as rexpR ¯a b idxqφ a,¯a,xpida b id¯a bf aqpida bφ¯a,a,yqpida bRa b idyqpida be yqs b p ¯R a b idaqpfa b ¯R¯ap1qq. Using again naturality and coherence for φ and e, this can be rewritten pf q " rf apR ¯a b ida b idyqpida bRa b idyqs b p ¯R a b idaqpfa b ¯R¯ap1qq. But since we may replace the conjugate solution pR¯a, ¯R¯aq with p ¯Ra, Raq, the conjugate equations for pRa, ¯Raq show that the above expression reduces to f . 19 Proposition 5.14. For f P A y x , we have αy xpf q " αx ypf q. Proof. The coaction on f can be written as reypR a b idyqφ ¯a,a,ypid¯a bf aqs b pf a b u¯aqp ¯Rap1q b 1q. Since ¯Ra P Morpuo, ua lT u¯aq, one has pu¯aq23p ¯Raq12 " pu aq13puaq13pu¯aq23p ¯Raq12 " pu aq13p ¯Raq12. Thus, we obtain y pf q " reypR αx a b idyqφ ¯a,a,ypid¯a bf aqs b pupfa b 1qq 13p ¯Rap1q b 1q " αy xpf q, which proves the assertion. Lemma 5.15. There is a natural equivariant -isomorphism A x'y x'y -- A x x A x y A y x A y y . Proof. This follows from the natural decomposition Endpx ' yq -- Endpxq Morpy, xq Morpx, yq Endpyq , which passes through all further structure imposed on the A y x . Lemma 5.16. We have pA y x qG " Morpy, xq. Furthermore, for f P A y y , we have pid bϕGqpαy ypf qq " f ofoe y P Endpyq. Proof. These formulas follow from the definition of αy representations. x and the orthogonality of irreducible Theorem 5.17. For each object y of D, the coaction of PpGq on A y of G. y defines an algebraic action Proof. The only thing left to prove is the complete positivity (P) for the map EG " pid bϕGqαy y. By Lemma 5.15, it is enough to show that EG is positive on A y y for arbitrary y. Let f, g P A y y . Then we have f g " reypR a b idyqφ ¯a,a,ypid¯a bf agbqφ¯a,a,yppppf a b id¯aq ¯Rap1q b gbqqc b idyqs Applying EG to this means taking the value at c " o. 20 b ppf a b id¯aq ¯Rap1q b gbqc. Since ua and ub are irreducible, there exists an embedding of uo into u¯a lT ub if and only if b " a. In that case an isometric embedding is given by pdimq uaq´1{2Ra for the normalized choice of pRa, ¯Raq. Thus, we obtain, using the conjugate equations for pRa, ¯Raq in the last step, ppf a b id¯aq ¯Rap1q b gaqoppf a b id¯aq ¯Rap1q b gaqo " as a morphism from uo to u¯a lT ua. Hence, " 1 dimq ua xfa, gay dimq ua pf a b R aqp ¯Rap1q b gaqRa Ra EGpf gq " " xfa, gay dimq ua 1 dimq ua eypR a b idyqφ ¯a,a,ypid¯a bf agaqφ¯a,a,ypRa b idyqe y eypR a b idyqφ ¯a,a,ypid¯a bxf, gyMorpy,yqqφ¯a,a,ypRa b idyqe y , where xf, gyMorpy,yq " xfa, gayf aga is the standard Morpy, yq-valued inner product on A y this formula, it follows that EG is indeed completely positive. x . From Remark 5.18. In [26], the construction of an action from a module category is carried out internally within the tensor category. There are two obstacles for attempting such a construction in our setting. The first obstacle is a finiteness problem, in that the algebra underlying an ergodic action will in general live inside a completion of the tensor category. This could be taken care of by standard techniques. The second obstacle is that we want our algebras to be endowed with a good -structure. Now ergodic actions on finite-dimensional C-algebras can be characterized abstractly inside of ReppGq as (irreducible) abstract Q-systems ([20], [22]). However, the definition of Q-system is too restrictive if we want to allow non-finite quantum homogeneous spaces. So although it seems manageable to lift both of the above obstacles separately, we do not know how to tackle them in combination. At this stage, we can apply the material developed in the previous section. Notation 5.19. For each object y in D, we denote the G-C-algebraic completion of A y Proposition 4.4) by Ay of Lemma 5.15 as y (see y. We denote the block decomposition of Ax'y induced by the isomorphism Ax'y x'y " Ax Ax x Ay x y Ay y . In this way, for general x, y, the space Ax Hilbert Ay y-module, together with a unital -homomorphism from Ax y naturally has the structure of an equivariant right x into LAy ypAy xq. Lemma 5.20. When x and y are objects in D with y irreducible, then the action of G on Ay y is ergodic, and Ay x is a finite equivariant Hilbert Ay y-module. Proof. From the block decomposition as in Notation 5.19, we may as well suppose that also x is irreducible. Then by Lemma 5.16 and Proposition 4.4, we obtain that the actions on Ay y 21 x are ergodic. Since the image of Ax and Ax deduce from Remark 3.2 that either we have an identification Ax is in particular finitely generated projective, or else Ay x in LpAy x " 0. xq must by construction contain KAy x -- KAy ypAy yq, we xq, in which case Ay x y pAy x are Banach spaces with the -operations Ay The Ay It follows that we can make a C-category A having the same objects as D, and with morphism space from x to y given by the Banach space Ax y. By Lemma 5.8, it contains a faithful copy of the C-category D, which are precisely the fixed points under the G-action on the morphism spaces. y satisfying the C-condition. x Ñ Ax Proposition 5.21. Let y be a fixed irreducible object in D, and let Ay be the category with ‚ objects the Ay x, where x ranges over the objects in D, and ‚ with morphism space MorAypz, xq the space KAy z, Ay xq. Then we have a C-functor Fy : A Ñ Ay, sending x to Ay x and an element f P Az ypAy multiplication with this element. Moreover, the resulting maps MorApz, xq Ñ KAy G-equivariant. ypAy x to left z, Ay xq are Proof. Since the modules Ay elements in Az map is then a formality to check. The equivariance follows from Proposition 5.9. x indeed gives compact operators from Ay x are finitely generated projective over Ay y, left multiplication with x. The functoriality of the given z to Ay Example 5.22. Let H be a quantum subgroup of G. We have seen in Example 2.15 that ReppHq is a ReppGq-module category. When w is an irreducible unitary representation of H, we find that A w w -- MorppuaqH b w, wq b Ha -- p ¯w b p¯uaqH b wqH b Ha -- pBpHwq b P pGqqH, the fixed points being with respect to the w-induced left H-action on BpHwq b P pGq. It then follows that the action of G on CpXwq given by Lemma 5.20 is equal to the right translation action on the fixed point algebra pBpHwq b CpGqqH. 6 Correspondence between the constructions Let G be a compact quantum group, and let X be a quantum homogeneous space over G. It is known [28] that the G-algebra CpXq can be recovered from the associated 'spectral functor' u ÞÑ HomGpHu, CpXqq on ReppGq, where the right hand side simply means the space of G-equivariant linear maps. In general, if we ignore the problem of completion, any right comodule E over CpGq can be recovered from its spectral functor by the formula HomGpHa, Eq b Ha » E , (6.1) àaPI 22 up to completion. The algebra structure of CpXq was recovered from the usual tensor structure on the forgetful functor of ReppGq, and the 'quasi-tensor' structure on the spectral functor. The above general scheme and our construction of G-algebra in the previous section are related by the following simple translation. Lemma 6.1. Let u P ReppGq, and let pE , αEq be a G-equivariant Hilbert C-module over CpXq. Then one has a natural isomorphism HomGpHu, Eq » HomG,CpXqpHu b CpXq, Eq, (6.2) where the right hand side denotes the space of linear G-equivariant, right CpXq-linear maps. Proof. If T P HomGpHu, Eq, the map ξ b x ÞÑ T pξqx from Hu b CpXq to E is G-equivariant and right CpXq-linear. On the other hand, the inverse correspondence is given by pulling back with the embedding Hu Ñ Hu b CpXq, ξ ÞÑ ξ b 1. The above isomorphism can be regarded as an adjunction between the 'scalar extension by CpXq' functor and the 'scalar restriction' functor (forgetting the action of CpXq). Moreover, CpXq itself can be regarded as an irreducible object in the category DX by the ergodicity. Hence, if E is a finite equivariant Hilbert module over CpXq, we have for the right hand side of (6.2) that HomG,CpXqpHu b CpXq, Eq " Morpu b CpXq, Eq, the latter a morphism space in DX. We use here implicitly that adjointability is automatic for CpXq-module maps between finitely generated projective modules). In the following, we use Notation 5.19. Proposition 6.2. Let ‚ denote the object CpXq in DX. Then the G-C-algebra A‚ antly isomorphic to CpXq. This isomorphism is induced by the embedding ‚ is equivari- A ‚ ‚ Ñ CpXq, f ÞÑ f apfa b 1q. (6.3) Proof. By Lemma 6.1, A ‚ ‚ can be identified with 'a HomGpHa, CpXqq b Ha, and the map (6.3) is identified with the canonical embedding (6.1). We obtain the assertion by comparing our product structure on A ‚ ‚ with the one in [28, Theorem 8.1]. Proposition 6.3. Let D be a connected module C-category over ReppGq. Let y P D be an irreducible object, and write Ay y " CpXyq. Then there is an equivalence of ReppGq-module C-categories D -- DXy, by restricting the functor Fy from Proposition 5.21 to D. Proof. First of all, Lemma 5.20 ensures us that Fy has the proper range on objects. Since D is realized inside the category A by taking the G-invariants in morphism spaces, the equivariance part of Proposition 5.21 ensures that Fy also has the proper range on morphisms. In the following, we will mean by Fy its restriction to D. 23 We next show that Fy is a ReppGq-module homomorphism. Let u be a finite-dimensional representation of G, and let x be an object in D. Then, the spectral subspace functors associated with Ay ubx is, by definition, determined by the spaces pMorpua b y, u b xqqaPI , but the Frobenius isomorphism implies that these are equal to x are the same: the one for Ay ubx and u b Ay Morpp¯u lT uaq b y, xq » Morp¯u lT ua, Ay xq " Morpua, Hu b Ay xq for a P I. The resulting linear isomorphism A y x is by construction a G- homomorphism. y -linear and isometric by the same type of calculation as in the previous section. The coherence conditions for Fy follow from the naturality for scalar restriction/extension and from the fact that we can canonically take u lT v " ¯v lT ¯u using the chosen duality morphisms for u and v. ubx Ñ Hu b A y It is right A y It remains to show that the sets of irreducible classes are in bijection under the functor Fy. By the connectedness of D, for any object x, there exists an (irreducible) representation u of G such that Morpu b y, xq ‰ 0. Hence Ay x is a non-zero Hilbert module. As in the proof of Lemma 5.20, it follows that Ay x is irreducible if x is irreducible. If further x and z are irreducible, we must have by the same reasoning that the map Ax Ax x Az x z Az z Ñ KpAy xq x, Ay zq KpAy KpAy z, Ay zq xq KpAy is an isomorphism. Using Lemma 5.16, we see that if x and z are non-isomorphic irreducible objects, Ay z are not equivalent in DXy. x and Ay Now, any object in DXy is a subobject of u lT CpXyq for some finite-dimensional representation u of G. As Fy preserves the module structure, and as CpXyq is the image of y by construction, we find that any object of DXy is isomorphic to an object in the image of Fy. By Lemma 2.6, we conclude that Fy is an equivalence of ReppGq-module C-categories. To conclude this section, we summarize our main result in the following theorem, which will also include the formalism on bi-graded Hilbert spaces developed in the Appendix. Indeed, in our setup, abstract module C-categories will arise naturally from the study of quantum homogeneous spaces, and one then passes to the bi-graded Hilbert space picture to reveal the combinatorial structure in a more tangible form, cf. the remark after Theorem 1.5 in [14]. This will be exploited in our forthcoming paper [8] to classify the ergodic actions of the quantum SU qp2q groups for 0 ă q ď 1. Theorem 6.4. Let G be a compact quantum group. There is a one-to-one correspondence between the following notions. 1. Ergodic actions of G (modulo equivariant Morita equivalence). 2. Connected module C-categories over ReppGq (modulo module equivalence). 3. Connected strong tensor functors from ReppGq into bi-graded Hilbert spaces (modulo nat- ural tensor equivalence). 24 The connectedness of a strong tensor functor F into J-bi-graded Hilbert spaces means that it can can not be decomposed as a direct sum F1 ' F2 with the Fi strong tensor functors into Ji-bi-graded Hilbert spaces, J " J1 Y J2 with J1 and J2 disjoint. Proof. The equivalence between the first two structures is a direct consequence of Proposi- tions 6.2 and 6.3, where the arbitrariness of the choice of irreducible object corresponds pre- cisely to equivariant Morita equivalence, cf. the remark above Notation 3.8. The equivalence between the last two is a consequence of Proposition A.4.2, under which the connectedness can be easily seen to be preserved. Let us give a little more detail on the direct correspondence between tensor functors and ergodic actions. Let J be a set, and pFrsqr,sPJ be a connected strong tensor functor from ReppGq into column-finite J-bi-graded Hilbert spaces. Then by Proposition A.4.2, HJ f has a structure of ReppGq-module C-category, in such a way that Frspuq -- Morpxr, u b xsq. Hence for r, s elements of J, the spaces A xs xr which were constructed in Section 5 can be explicitly expressed as A xs xr " àaPI Frspaq b Ha, since we can identify Morpu b xs, xrq with the conjugate Hilbert space of Morpxr, u b xsq by means of the adjoint map. 7 Categorical description of equivariant maps In this last section, we investigate the relationship between equivariant maps between quantum homogeneous spaces and equivariant functors between module C-categories. Let X and Y be quantum homogeneous spaces over G, respectively given by the coactions α : CpXq Ñ CpXq b CpGq and β : CpYq Ñ CpYq b CpGq. A G-morphism from Y to X is repre- sented by a unital -algebra homomorphism θ from CpXq to CpYq satisfying the G-equivariance condition pθ b idq α " β θ. Given such a homomorphism θ, we obtain a -preserving functor θ# : DX Ñ DY defined as the extension of scalars E ÞÑ E bCpXq CpYq . We may assume that this functor maps the distinguished object CpXq of DX to the one of DY, namely CpYq. When u P ReppGq and E P DX, let ψθ denote the isomorphism θ pHu b Eq bCpXq CpYq Ñ Hu b pE bCpXq CpYqq, pξ b xq b y ÞÑ ξ b px b yq. Then ψθ can be considered as a natural unitary transformation ψθ : θ#p´ b ´q Ñ ´ b pθ#´q between functors from ReppGq DX to DY. This ψθ enables one to complete θ# to a module C-category homomorphism between DX and DY, cf. Definition 2.17. We aim to characterize the G-equivariant morphisms of quantum homogeneous spaces in terms of their associated categories and functors between them. 25 Theorem 7.1. Let X and Y be quantum homogeneous spaces over G. Let pG, ψq be a ReppGq- module homomorphism from DX to DY satisfying GpCpXqq " CpYq. Then there exists a G- equivariant -homomorphism θ from CpXq to CpYq such that θ# is naturally isomorphic to G. Furthermore, two ReppGq-module homomorphisms pG, ψq and pG, ψ1q with the same underlying functor give rise to the same homomorphism θ if and only if ψ and ψ1 are conjugate by a unitary self-equivalence of G. Proof. By Proposition 6.2, we know that CpXq can be identified with a completion of the space A " 'aPI MorpHa b CpXq, CpXqq b Ha, and similarly for CpYq as a completion of the space B " 'aPI MorpHa b CpYq, CpYqq. For any u P ReppGq, the action of G and ψ u,CpXq induces a linear map ΨE u : Mor pHu b CpXq, Eq Ñ Mor pHu b CpYq, GEq , sending f to Gpf qψ as a map from A to B. u,CpXq. When E " CpXq, we write ΨCpXq u " Ψu, and we put θ " 'aPIΨa b ida We first want to show that this is an algebra homomorphism. Let f and g be elements of A . The effect of θ on f g can be expressed, using the notation from Definition 5.5, as pθpf gqqc b pθpf gqqc " rG`f apida bgbqppfa b gbqc b idCpXqq ψ c,CpXqs b pfa b gbqc (7.1) where we have dropped the associativity constraint for the module category since the latter is concrete. By functoriality of G, naturality of ψ and coherence of ψ, the morphism part in the left leg of the above formula can be written as Gpf aqψ a,CpXqpida bGpgbqqpida bψ b,CpXqqppfa b gbqc b idCpYqq, which can be simplified to Ψapf aqpida bΨbpgbqqppfa b gbqc b idCpYqq. Since we can write θpf q " Ψapf aq b fa, we conclude that indeed θpf gq " θpf qθpgq. In the same way, the unitality of θ is proven. Next, let us observe that θ is compatible with the involution on both algebras. This is a consequence of the facts that G 'commutes' with the morphisms in R and intertwines the - operations on DX and DY, and of naturality of ψ. Since θ is equivariant by construction, it then follows from Proposition 4.5 that θ can be extended uniquely to an equivariant -homomorphism from CpXq to CpYq, which we denote by the same symbol. Finally, we have to prove that θ# and G are equivalent. Let E be an object of DX, and write AE " 'aPI Morpua b CpXq, Eq b Ha, which we know can be identified with a dense subset of E. Similar notation will be used for B. Then for f P AE and g P B, we can define an element nE pf, gq in BGE by nEpf, gq " ΨE " rΨE apf aqpfa b gq apf aqpida bgbqppfa b gbqc b idCpYqqs b pfa b gbqc. 26 This will give a linear map nE from the algebraic tensor product AE b B to BGE . By construc- tion, it extends to the canonical isomorphism θ#CpXq » CpYq " GCpXq at the object CpXq. Using ReppGq-equivariance, it then follows that nE also extends to a unitary from θ#pEq to GpEq for E of the form u lT CpXq for some representation u of G. By the connectedness of DX and linearity, we deduce that this holds for arbitrary E. Hence nE induces a natural unitary transformation n : θ#E Ñ GE. The way in which n is constructed shows that the canonical ψθ is interchanged with ψ, i.e. Indeed, taking ξ P Hu, f P AE and g P CpYq, we have that pid bηEq pψθqu,E " ψu,E ηubE . ηubEppξ b f q b gq " Gppξ b f qaqψ a,CpXqppξ b f qa b gq. On the other hand, ψ u,Epid bηEqpξ b pf b gqq " ψ u,Erξ b Gpf cqψ c,CpXqpfc b gqs c,CpXqqpξ b fc b gq " Gpidu bf cqψ " Gpidu bf cqψ " Gpidu bf cqψ " Gppidu bf cqppξ b fcqa b idCpXqqqψ u,cbCpXqp1 b ψ ubc,CpXqpξ b fc b gq ubc,CpXqppξ b fcqapξ b fcqa b gq a,CpXqppξ b fcqa b gq, which then reduces to the expression above. It follows that if we have a different ψ1 which leads to the same θ, we can construct by means of the two n-maps for ψ and ψ1 a unitary self-equivalence of G which conjugates ψ and ψ1. Conversely, if µ is a natural unitary equivalence from G to itself, the µ-conjugated natural transformation ψµ " pidu bµCpXqqψµ ubCpXq : Gpu b CpXqq Ñ u b CpYq gives the same map Morpu b CpXq, CpXqq Ñ Morpu b CpYq, CpYqq as the one induced by ψ. Example 7.2. Let K ă H be an inclusion of quantum subgroups of G. Then, the restriction functor ReppHq Ñ ReppKq is a ReppGq-module homomorphism, and maps the trivial represen- tation of H to the one of K. The induced G-equivariant homomorphism CpHzGq Ñ CpKzGq is the canonical inclusion of fixed point subalgebras for the respective left translation actions. We now want to interpret Theorem 7.1 in the context of bi-graded Hilbert spaces. We keep X and Y fixed quantum homogeneous spaces for G. In the following, we let J (resp. J 1) be an index set of the irreducible objects in DX (resp. DY). We denote the index corresponding to CpXq (resp. CpYqq by ‚ (resp. ). The J J-graded (resp. J 1 J 1-graded) Hilbert space associated with the action of u P C on DX (resp. DY) is denoted by pF X pqpvqqp,qPJ 1), and the corresponding unitaries by rspuqqr,sPJ (resp. pF Y φ rs,u,v : F rspu b vq Ñ 'tF rtpuq b F tspvq. 27 Then if θ : CpXq Ñ CpYq is an equivariant -homomorphism, we have the J 1 J-graded Hilbert space 'p,rFpr associated with θ#, where Fpr " Morpxp, θ#xrqq for p P J 1, r P J, cf. Section A.2. From Theorem 7.1, we then obtain the following corollary. Corollary 7.3. Let X and Y be quantum homogeneous spaces for G. The equivariant homo- morphism from CpXq to CpYq are in one-to-one correspondence with the classes of families of Hilbert spaces Fpr, p P J 1 and r P J, and unitary maps ψu pr : àsPJ Fps b F X srpuq Ñ àqPJ 1 F Y pqpuq b Fqr for u P ReppGq, r P J, and p P J 1, such that ψo diagrams p,r is δp,r times the identity, Fp,‚ " δp,, the 'sFps b F X srpuq 's id bF X srpT q 'sFps b F X srpvq ψu pr ψv pr 'qF Y pqpuq b Fqr 'qF Y pqpT qbid / 'qF Y pqpvq b Fqr are commutative for any T P Morpu, vq, and 's,tFps b F X stpuq b F X trpvq 'q idpq bψv qr +❱❱❱❱❱❱❱❱❱❱❱❱❱❱❱❱❱❱❱ 'q,wF Y pqpuq b F Y qwpvq b Fwr 'q,tF Y pqpuq b Fqt b F X trpvq 'tψu ptbidtr 4❤❤❤❤❤❤❤❤❤❤❤❤❤❤❤❤❤❤ 's idps bφX sr,u,v 'sFps b F X srpu b vq is commutative. ψubv pr 'wφY pw,u,v / 'wF Y pwpu b vq b Fwr Here two families pFpr, ψu unitaries Urs : Fpr Ñ Gpr such that qtq and pGpr, µu qtq belong to the same class if and only if there are p'wpidqw bUwsqqψu qt " µu qtp'spUqs b idstqq for all q P J 1, t P J and u P ReppGq. In practice, one only needs to verify the above assumptions for all irreducible u (in which case the naturality condition simplifies), or for tensor products of a b-generating object (in which case the constraint condition simplifies). Moreover, the fact that the above Hilbert spaces are often one-dimensional in special cases makes the problem of determining the possible ψ more tractable. rspuaqqrs and pqpuaqqpq for a P I, and pdimpFprqqpr. These are multiplicity matrices as considered in [35] Important invariants of pX, Y, θq are the families of integer-valued matrices pdim F X pdim F Y and [31] (see Remark 3.12). 28 / /     / + 4     / Let us examine them more closely in the particular case when the larger algebra CpYq is of full quantum multiplicity [6, 9]. This is the case if and only if DY is based on a singleton tyu. Thus, the functor θ# : DX Ñ DY itself can be classified among the C-functors by the dimension of the vector spaces Fr " Morpθ#pxrq, yq for r P J. The next result is useful in determining the coideals inside the full quantum multiplicity ones even when there is no trace, c.f. [31, Corollary 4.21]. Proposition 7.4. Let X and Y be quantum homogeneous spaces over G. Assume that CpYq is of full quantum multiplicity, and that there is a G-equivariant homomorphism θ from CpXq to CpYq. Then, for any u P ReppGq, the matrix pdim F X rspuqqr,sPJ has an integer-valued eigenvector for the eigenvalue dim F Ypuq. Proof. The vector pdim FrqrPJ satisfies dim Fr dim F X rspuq " dim Mor pθ#pu b xsq, CpYqq ÿrPJ for any s P J (the above sum makes sense because pF X eigenvector of the eigenvalue dim F Ypuq. rspuqqr,sPJ is banded). Hence it is an " dim Morpu b θ#xs, CpYqq " dim F Ypuq dimpFsq Appendix. Concrete C-categories In this appendix, we pick up the discussion which we started in Section 2. It is, essentially, an elaborate write-out of the remark appearing in the proof of Theorem 2.5 of [11]. A.1 Concrete semi-simple C-categories As we will show in Lemma A.1.6, there is essentially only one semi-simple C-category based on a given set J. This can easily be shown by using Lemma 2.6, but we would like to have a more concrete formula for the inverse of such an equivalence functor. To accomplish this, we first establish some preliminaries results. The first goal is to generalize the direct sum construction in the setting of C-categories. Definition A.1.1. Let D be a C-category. Let X be an object of D, and H a finite- dimensional Hilbert space. An H -amplification of X is an object H b X together with a linear map θH X : H Ñ MorpX, H b Xq such that 1. For all ξ, η P H , we have θH X pξqθH 2. If ξi is an orthonormal basis of H , then ři θH X pηq " xξ, ηy idX. X pξiqθH X pξiq " idH bX. 29 Note that, if H " 0, the second condition above implies that the H -amplification is a zero object. Similarly, if H " C, the H -amplification is equivalent to the identity functor. Lemma A.1.2. Let D be a C-category admitting finite direct sums, and H a Hilbert space of finite dimension. Then any object of D admits an H -amplification. The ensuing operation Hf D Ñ D can be extended to an Hf-module C-category structure on D. We recall that Hf is the category of finite-dimensional Hilbert spaces. i"1 for each H . For an object X P D, define i"1X of n copies of X. With vi denoting the i-th isometric Proof. Choose a fixed orthonormal basis peiqn H b X as the direct sum 'n injection X Ñ 'iX, the θH X pξq " řixei, ξyvi are easily seen to satisfy the conditions for an H -amplification. The resulting construction is obviously functorial in X. If x is an operator H Ñ K , we choose an orthonormal basis pfiqi for H and define x b idX to be the operator X pfiq from H b X to K b X. Again, this is clearly independent of the chosen basis for H , and will give functoriality on the H -component. Finally, the associator for the module structure can be made as follows: given Hilbert spaces H and K with respective bases pfiq and pgjq, we define ři θK X pxfiqθH φH ,K ,X " ÿi,j θH K bXpfiqθK X pgjqθH bK X pfi b gjq as a morphism pH b K q b X Ñ H b pK b Xq. As a consequence of the Hf-module structure, we obtain a natural isomorphism MorpH b X, K b Y q » K b H b MorpX, Y q. In the presentation of the right hand side, composition of morphisms involves the concatenation of the form H b ¯H Ñ C 'in the middle' by means of the inner product. Notation A.1.3. Let D be a semi-simple C-category based on the set J. Let r P J and X P D. We denote by Xprq the Hilbert space MorpXr, Xq. Lemma A.1.4. Let D be a semi-simple C-category based on an index set J. Then there is a natural unitary equivalence X Ñ 'rPJ Xprq b Xr for X P D. Proof. Let Y be another object of D. Considering the central support of range projections for morphisms in MorpY, Xq, we see that the map MorpXr, Xq b MorpY, Xrq Ñ MorpY, Xq àrPJ induced by composition of morphisms is an isomorphism. The left hand side of the above is, by definition of the amplification, canonically isomorphic to Mor pY, 'rPJ Xprq b Xrq. By the Yoneda lemma, we obtain the assertion. 30 The next definition provides the canonical semi-simple C-category with which we will want to compare an arbitrary one. Definition A.1.5. Let J be a set. A J-graded Hilbert space is a Hilbert space H endowed with a direct sum decomposition H " 'rPJ Hr (the right hand side should be understood as the Hilbert space direct sum). They form a C-category HJ by considering as morphisms the grading-preserving operators, MorpH , K q " tT P BpH , K q @r P J : T pHrq Ď Kru " !pTrqrPJ P źrPJ BpHr, Krq sup rPJ }Tr} ă 8). The full subcategory of J-graded finite-dimensional Hilbert spaces is denoted HJ f . f then forms a semi-simple C-category, based on the set J in a natural way. The category HJ Namely, an irreducible object for the label r P J is given by the graded Hilbert space Cr which has C as component at place r and 0 at the other places. Lemma A.1.6. Let D be a semi-simple C-category based on a set J. Then the categories D and HJ f are unitarily equivalent, an adjoint pair of equivalences being given by X ÞÑ àrPJ Xprq, H ÞÑ àrPJ Hr b Xr, where Hr denotes the r-th component of H . Proof. An equivalence between D and HJ nition A.1.1 to define invertible unit and co-unit maps for the stated functors. f can be established by using Lemma A.1.4 and Defi- A.2 Functors and natural transformations The goal of this section is to give an equally concrete description of functors between semi-simple C-categories, and natural transformations between them. Let J and J 1 be index sets. Let H " 'pPJ 1,rPJ Hpr be a Hilbert space endowed with a direct sum decomposition over the set J 1 J. We also assume that H is column-finite in the sense that řp dimpHprq is finite for all r. In particular all Hpr are finite-dimensional. Then one has f given by pF H K qp " 'rHpr b Kr on objects, and pF H pT qqp " f to HJ 1 a functor F H from HJ 'r idpr bTr on morphisms. If J 2 is another index set and H 1 is a column-finite J 2J 1-graded Hilbert space, the composition of functors F H 1 and F H is given by F K , where the J 2 J-graded Hilbert space K is given by the l8pJ 1q-balanced tensor product pH 1 âl8pJ 1q H qvr " àpPJ 1 pH 1 vp b Hprq pv P J 2, r P Jq. 31 Let D (resp. D1) be a semi-simple C-category based on an index set J (reps. J 1), with a system of irreducible objects pXrqrPJ (resp. pYpqpPJ 1). The next proposition shows that any functor between abstract semi-simple C-categories is induced by a column-finite J 1 J-graded Hilbert space as above. Proposition A.2.1. Let F be a C-functor from D to D1. Up to the unitary equivalence of Lemma A.1.6, F is naturally equivalent to the functor induced by the J 1 J-graded Hilbert space H F whose pp, rq-th component is MorpYp, F pXrqq. Proof. First of all, the graded Hilbert space 'p,r MorpYp, F pXrqq is indeed column-finite, as the F pXrq splits into a finite number of irreducible objects. A natural equivalence as in the statement of the proposition must then be given by unitary maps for p P J 1. On the direct summand at r, we define φp as the map H F pr b Xprq Ñ pF Xqppq φp : àrPJ MorpYp, F pXrqq b MorpXr, Xq Q f b g ÞÑ F pgq f P MorpYp, F Xq. Then the resulting map is indeed unitary by the semi-simplicity of D. The compatibility with the morphisms in D is apparent from the above definition of φp. Suppose we are given two J 1 J-graded Hilbert spaces H and K , and an operator T P BpH , K q which respects the grading. Then, we obtain a natural transformation ηT of F H into F K by the formula pηT M qp " àrPJ Tpr b idMr : F H pM qp Ñ F K pM qp, because the norm of this operator is uniformly bounded by }T }. Thus, we obtain a morphism from F H to F K in the category FunpHJ f q (see Remark 2.2). f , HJ 1 Conversely, uniformly bounded norm from F to G. Then the induced maps let F and G be functors from D to D1, and η be a natural transformation of T η pr : MorpYp, F pXrqq Ñ MorpYp, GpXrqq, f ÞÑ ηXr f has a norm bounded from above by }η}. Now, from the way F and H F is identified in Proposition A.2.1, one sees that the above correspondences T ÞÑ ηT and η ÞÑ T η are inverse to each other. We record this for reference in the following proposition. Proposition A.2.2. Let F and G be functors from D to D1. Then morphisms from F to G in FunpD, D1q can be naturally identified with grading preserving bounded operators from H F to H G. 32 A.3 Concrete semi-simple tensor C-categories We next apply the above constructions to the endomorphism tensor category EndpDqf associ- ated with a semi-simple C-category D. Notation A.3.1. Let J be an index set, and denote by E J the C-category of column-finite J J-graded Hilbert spaces H " 'r,sPJHrs. As morphisms, we take the bounded operators v : H Ñ K which preserve the grading. By the results of Section A.2, we can identify E J with the tensor C-category of C-endofunctors f . Thus, the tensor product, is given by the l8pJq-balanced tensor product, and the unit on HJ object 1J is given by l2pJq with the diagonal J J-grading p1J qst " δs,tC. Lemma A.3.2. The maximal rigid subcategory E J satisfy the condition f of E J has as its objects those H which sup r ÿs pdimpHrsq ` dimpHsrqq ă 8. In particular, all Hrs are finite-dimensional, and only a finite number of Hrs are non-zero on each 'row' and 'column', i.e. the grading is banded. The dual dpH q of H can then be given by dpH qrs " Hsr " H sr with duality morphisms RH : l2pJq Ñ àr,sPJ ¯RH : l2pJq Ñ àr,sPJ Hrs b Hrs, Hrs b Hrs, δs ÞÑ ÿr,i δr ÞÑ ÿs,i ξpr,sq i b ξpr,sq i i b ξpr,sq ξpr,sq i , where the ξpr,sq i form an orthogonal basis of Hrs. Proof. The restriction on the dimensions of the Hrs ensures that both operators RH and ¯RH are bounded. It is then straightforward to check that they satisfy the snake identities for a duality. Conversely, suppose that 'rsHrs admits a dual 'rsGrs by means of duality morphisms pR, ¯Rq. Then the latter decompose into maps Rrs : C Ñ Grs b Hsr, ¯Rrs : C Ñ Hrs b Gsr. Let us write Jrspξq " pξ b idqpRrsp1qq P Hsr, Irspηq " pη b idqp ¯Rrsp1qq P Gsr for ξ P Grs and η P Hrs. Then Jrs gives an anti-linear map from Grs to Hsr, and Irs from Hrs to Gsr. The snake identities (2.1) imply that Irs is the inverse of Jsr. By the boundedness of R and ¯R, we obtain that suprřs TrpJ suprřs TrpI equality as supsřr TrppJ rsIrsq " } ¯R}2. Since Irs " J ´1 rsJrsq´1q " } ¯R}2. rsJrsq " }R}2, and similarly sr , the trace property allows us to rewrite the latter 33 Suppose now that that the condition suprřspdimpHrsq ` dimpHsrqq ă 8 is not satisfied. Then by symmetry we may assume that there exists a sequence rn such that řs dimpHrn,sq ě n. This implies that we can also find sn and a strictly positive eigenvalue λ of J λ ď }R}2 n . But as λ´1 ď } ¯R}2, this gives a contradiction. Jrn,sn such that rn,sn We now show that if D is a semi-simple C-category based on an index set J, then EndpDqf is tensor equivalent with E J f . Proposition A.3.3. Let D be a semi-simple C-category, based on an index set J. Then the categories EndpDqf and E J f are tensor equivalent, by means of the associations F ÞÑ àpr,sqPJJ MorpXr, F pXsqq, H ÞÑ "X ÞÑ àr,sPJ Hrs b Xpsq b Xrı. Proof. We have already remarked that there are mutually inverse tensor equivalences EndpDq Ø E J . Since equivalences preserve duality, they restrict to equivalences between EndpDqf and E J f . A.4 Module C-categories and bi-graded tensor functors This section essentially establishes that also in the categorical set-up, there is an equivalence between modules and representations. Combined with the material of the previous sections, it allows one to present a concrete and workable version of a semi-simple module C-category. Lemma A.4.1. Let C be a tensor C-category, and D a C-category. Then there is an equivalence between C-module C-category structures M on D and strong tensor C-functors F : C Ñ EndpDq. Proof. For module structures M and tensor functors F , we have the associations M ÞÑ rFM : U ÞÑ MpU, ´qs, F ÞÑ rMF : pU, Xq ÞÑ F pUqpXqs, mapping all other structural morphisms in the obvious ways. These maps are clearly inverses to each other. We can now state the following useful result. Proposition A.4.2. Let C be a tensor C-category, and let J be a set. Then there is an equivalence between 1. module C-structures on J-based semi-simple C-categories, and 2. strong tensor C-functors C Ñ E J f . 34 Given a module C-category pD, M, φ, eq, the corresponding tensor functor C Ñ E J f is given by F : U Ñ àr,s MorpXr, MpU, Xsqq. Writing the right hand side above as 'r,sFrspUq, the coherence maps for tensoriality are encoded as isometries FrspUq b FstpV q Ñ FrtpU b V q, f b g ÞÑ φ U,V,Xt pidU bgq f, (A.4.1) Proof. By Lemma A.4.1 and Lemma 2.13, a C-module C-category structure on a semi-simple C-category D based on J is equivalent to giving a strong tensor C-functor from C to EndpDqf. Composing with the tensor equivalence from Proposition A.3.3, we obtain the correspondence stated in the proposition. References [1] N. Andruskiewitsch and J. M. Mombelli, On module categories over finite-dimensional Hopf algebras, J. Algebra 314 (2007), 383 -- 418. [2] S. Baaj and G. Skandalis, C -alg`ebres de Hopf et th´eorie de Kasparov ´equivariante, K- Theory, 2 (1989),683 -- 721. [3] J.C. Baez, Higher-Dimensional Algebra II: 2-Hilbert Spaces, preprint (1996) arXiv:q-alg/9609018 [math.QA] [4] T. Banica, The planar algebra of a coaction, J. Operator Theory 53 (2005), 119 -- 158. [5] T. Banica, Quantum automorphism groups of homogeneous graphs, J. Funct. Anal. 224 (2005), 243 -- 280. [6] J. Bichon, A. De Rijdt and S. Vaes, Ergodic coactions with large quantum multiplicity and monoidal equivalence of quantum groups, Comm. Math. Phys. 262 (2006), 703 -- 728. [7] F. Boca, Ergodic actions of compact matrix pseudogroups on C -algebras, In: Recent advances in operator algebras (Orl´eans, 1992). Ast´erisque 232 (1995), 93 -- 109. [8] K. De Commer and M. Yamashita, Tannaka-Kreın duality for compact quantum homo- geneous spaces. II. Classification of quantum homogeneous spaces for quantum SU p2q, preprint (2012) arXiv:1212.3413 [math.OA]. [9] A. De Rijdt and N. Vander Vennet, Actions of monoidally equivalent compact quantum groups and applications to probabilistic boundaries, Ann. Inst. Fourier (Grenoble) 60 (1) (2010), 169 -- 216. [10] S. Doplicher J.E. Roberts, A new duality theory for compact groups, Invent. Math. 98 (1989), 157 -- 218. 35 [11] P. Etingof and V. Ostrik, Module categories over representations of SLqp2q and graphs, Math. Res. Lett. 11 (1) (2004), 103 -- 114. [12] S. Gelaki, Module categories over affine group schemes, preprint (2012) arXiv:1209.1155 [math.QA]. [13] P. Ghez, R. Lima and J.E. Roberts, W-categories, Pac. J. Math. 120 (1985), 79 -- 109. [14] P. Grossman and N. Snyder, Quantum subgroups of the Haagerup fusion categories, preprint (2011) arXiv:1102:2631 [math.OA]. [15] R. Haring-Oldenburg, Reconstruction of weak quasi-Hopf algebras, J. Algebra 194 (1) (1997), 14 -- 35. [16] T. Hayashi, A canonical Tannaka duality for finite semisimple tensor categories, preprint (1999) arXiv:math/9904073 [math.QA] [17] N. Iwahori and M. Sugiura, A duality theorem for homogeneous compact manifolds of compact Lie groups, Osaka J. Math. 3 (1966), 139 -- 153. [18] A. Joyal and R. Street, An introduction to Tannaka duality and quantum groups, Lecture Notes in Math. 1488, Springer-Verlag (1991), 413 -- 492. [19] G.G. Kasparov, The operator K-functor and extensions of C-algebra, Math. USSR Izvestija 16 (3) (1981), 513 -- 572. [20] R. Longo and J.E. Roberts, A theory of dimension, K-theory 11 (1997), 103 -- 159. [21] S. Mac Lane, Categories for the working mathematician, Springer-Verlag, Berlin, Heidel- berg, New York (1971). [22] M. Muger, From subfactors to categories and topology I. Frobenius algebras in and Morita equivalence of tensor categories, J. Pure Appl. Alg. 180 (2003), 81 -- 157. [23] M. Muger, Tensor categories: A selective guided tour, Revista de la Uni´on Matem´atica Argentina 51 (1) (2010), 95 -- 163. [24] M. Muger, J. E. Roberts and L. Tuset, Representations of algebraic quantum groups and reconstruction theorems for tensor categories, Alg. Repres. Theor. 7 (2004), 517 -- 573. [25] A. Ocneanu, The classification of subgroups of quantum SU pNq, In "Quantum symmetries in theoretical physics and mathematics (Bariloche, 2000)", Contemp. Math. 294, 133 -- 159, Amer. Math. Soc., Providence, RI, 2002. [26] V. Ostrik, Module categories, weak Hopf algebras and modular invariants, Transform. Groups 8 (2) (2003), 177 -- 206. [27] V. Ostrik, Module categories over the Drinfeld double of a finite group, Int. Math. Res. Not. 27 (2003), 1507 -- 1520. 36 [28] C. Pinzari and J. E. Roberts, A duality theorem for ergodic actions of compact quantum groups on C -algebras, Comm. Math. Phys. 277 (2) (2008), 385 -- 421. [29] P. Podle´s, Symmetries of Quantum Spaces. Subgroups and Quotient spaces of Quantum SU p2q and SOp3q Groups, Comm. Math. Phys. 170 (1995), 1 -- 20. [30] P. Schauenburg, Hopf-Galois and Bi-Galois extensions, Galois theory, Hopf algebras, and semiabelian categories, Fields Inst. Comm. 43, AMS (2004), 469-515. [31] R. Tomatsu, Compact quantum ergodic systems, J. Funct. Anal. 254 (2008), 1 -- 83. [32] K.-H. Ulbrich, Galois extensions as functors of comodules, Manuscripta Math. 59 (1987), 391 -- 397. [33] K.-H. Ulbrich, Fiber functors of finite dimensional comodules, Manuscripta Math. 65 (1989), 39 -- 46. [34] R. Vergnioux, KK -th´eorie ´equivariante et op´eratour de Julg-Valette pour les groupes quan- tiques, Ph.D. thesis, Universit´e Paris Diderot-Paris 7, 2002. [35] A. Wassermann, Ergodic actions of compact groups on operator algebras. I. General theory, Ann. of Math. 130 (2) (1989), 273 -- 319. [36] S.L. Woronowicz, Duality in the C -algebra theory, Proceedings of the International Congress of Mathematiciens, Warsaw (1983), 1347 -- 1356. [37] S.L. Woronowicz, Compact matrix pseudogroups, Comm. Math. Phys. 111 (1987), 613 -- 665. [38] S.L. Woronowicz, Tannaka-Kreın duality for compact matrix pseudogroups. Twisted SU pNq groups, Invent. Math. 93 (1988), 35 -- 76. [39] S.L. Woronowicz, Compact quantum groups, in "Sym´etries quantiques" (Les Houches, 1995), North Holland, Amsterdam (1998), 845 -- 884. 37
1709.03281
2
1709
2018-07-13T04:58:16
Reconstructing the Bost--Connes semigroup actions from K-theory
[ "math.OA", "math.NT" ]
We complete the classification of Bost--Connes systems. We show that two Bost--Connes C*-algebras for number fields are isomorphic if and only if the original semigroups actions are conjugate. Together with recent reconstruction results in number theory by Cornelissen--de Smit--Li--Marcolli--Smit, we conclude that two Bost--Connes C*-algebras are isomorphic if and only if the original number fields are isomorphic.
math.OA
math
RECONSTRUCTING THE BOST–CONNES SEMIGROUP ACTIONS FROM K-THEORY YOSUKE KUBOTA AND TAKUYA TAKEISHI Abstract. We complete the classification of Bost–Connes systems. We show that two Bost–Connes C*-algebras for number fields are isomorphic if and only if the original semigroups actions are conjugate. Together with recent reconstruction results in number theory by Cornelissen–de Smit–Li–Marcolli–Smit, we conclude that two Bost–Connes C*-algebras are isomorphic if and only if the original number fields are isomorphic. 1. Introduction The Bost–Connes system (AK , σK,t) is a C*-dynamical system attached to a number field K. A specific feature of this system is that its dynamics, in particular the behavior of its KMS-states, reflects the arithmetics of the original number field. For example, the set of extremal KMS-states at low temperature equips a free transitive action of the Galois group Gab K . At high temperature there is a unique KMS-state. The critical temperature β = 1 is nothing but the critical point of the Dedekind zeta function, which is the partition function of the Bost–Connes system. After the pioneering work by Bost–Connes [3] in the case of Q, the gen- eralization of this dynamical system to an arbitrary number field has been a leading problem in the study of Bost–Connes systems. It was completed after a 15-year effort by many mathematicians such as Ha–Paugam [9] (the definition of the Bost–Connes system), Laca–Larsen–Neshveyev [12] (the KMS-classification) and Yalkinoglu [19] (construction of the arithmetic sub- algebra). For the definition of the Bost–Connes system, one starts with an action of the semigroup IK of integral ideals of K on a compact space YK, which is defined by using the Artin reciprocity map in class field theory. It asso- ciates a groupoid GK and the Bost–Connes C*-algebra AK is the groupoid C*-algebra C ∗ r (GK ). The R-action σK is induced from the absolute norm function N : IK → R≥0. See Section 2.1 for more details. According to the philosophy of anabelian geometry, it is natural to expect that the Bost–Connes system remembers the original number field. More precisely, two Bost–Connes systems AK and AL have been conjectured to be isomorphic if and only if the number fields K and L are isomorphic. Indeed, as is mentioned above, the dynamics of the Bost–Connes system such as the KMS states recovers many of the arithmetics of the number field. For example, an isomorphism of Bost–Connes systems immediately induces a bijection between abelianized Galois groups. This problem was Date: July 16, 2018. 1 2 Y. KUBOTA AND T. TAKEISHI first considered in [7] and recently a remarkable partial solution is given by Cornelissen, de Smit, Li, Marcolli and Smit in [5, 6], where it is proved that the Bost–Connes semigroup actions YK x IK and YL x IL are conjugate if and only if K is isomorphic to L. Alongside these results, recent works by the second author [17,18] provide a new perspective on this problem: Even if we forget the R-action σK,t, the underlying Bost–Connes C*-algebra AK has rich information. A key observation is that AK has the canonical structure of a C*-algebra over 2PK (see Subsection 2.3 for more details). In particular, the main theorem of [18] clarifies that we can reconstruct the Dedekind zeta function of K from the ordered K0-group of irreducible sub-quotients without using the R-action and KMS-states. In this paper, we establish this idea in a complete way. Our main theorem is the following: Theorem 1.1. Let K and L be number fields. The following are equivalent: (1) The semigroup actions YK x IK and YL x IL are conjugate. (2) The Bost–Connes groupoids GK and GL are isomorphic. (3) The Bost–Connes systems (AK , σK,t) and (AL, σL,t) are R-equivariantly isomorphic. (4) The Bost–Connes C*-algebras AK and AL are isomorphic. (5) There is a bijection PK ∼= PL (which identifies 2PK ∼= 2PL ) and an ordered KK(2PK )-equivalence between AK and AL. (6) There is a bijection PK ∼= PL and a family of ordered isomorphisms K ) → K∗(BF L ) ϕF : K∗(BF K = ∂F,p such that ϕF ∪{p}◦∂F,p L ◦ϕF for any finite subset F ⊂ PK ∼= PL. Here we say that an isomorphism of K0-groups is ordered if it gives a bijection of the positive cones (see for example [2, Section 6]). The precise meaning of (5) is that there is an invertible element in KK(2PK ; AK , AL) ([11, Definition 4.1], see also [14, Definition 3.1]) which induces a family of ordered isomorphisms between filtered K0-groups [15, Definition 2.4]. The C*-algebra BF K are defined in Definition 2.11 and Definition 2.14 respectively. K and the homomorphism ∂F,p The essential step is (6)⇒(1). For the proof, we essentially give a recon- struction procedure of the topological space YK and the action of IK from the given K-theoretic data. Combining Theorem 1.1 with recent results [5, Theorem 3.1] and [6, The- orem 3.1] mentioned above, we complete the classification of Bost–Connes systems and the underlying C*-algebras. Corollary 1.2. Let two number fields K and L satisfy one of the equivalent conditions (1) - (6) in Theorem 1.1. Then, K is isomorphic to L. Theorem 1.1 can be viewed as not only a classification result of C*- algebras, but also a construction of an invariant of number fields. Actually, the implication (6)⇒(1) means that the family of ordered K∗-groups with boundary homomorphisms provides a complete invariant of number fields. It would be interesting to relate this invariant with other known invariants in number theory. RECONSTRUCTING THE BOST–CONNES SEMIGROUP ACTIONS 3 This paper is organized as follows. In Section 2, we revisit the global struc- ture of the Bost–Connes semigroup action and Bost–Connes C*-algebra from the viewpoint of the valuation map. In Section 3, we give a reconstruction procedure of profinite completions of free abelian groups from the K∗-group of the crossed product. Finally, the proof of Theorem 1.1 is given in Section 4. In Appendix A, we introduce a more direct reconstruction of the profinite completion provided for the authors by Xin Li. Acknowledgment. The authors are gratefully indebted to Xin Li for his interest and allowing them to expose his elegant alternative proof in this paper. They also would like to thank Yuki Arano, Kazuki Tokimoto and Makoto Yamashita for for helpful discussions. In addition, the second au- thor would like to thank Koji Fujiwara for the support when the second author belonged to the Department of Mathematics of Kyoto University and Research Institute for Mathematical Sciences, Kyoto University. The first author is partially supported by MEXT's Program for Leading Grad- uate Schools, the Research Fellowship of the JSPS (No. 26-7081) and JSPS KAKENHI Grant Number JP17H06461. The second author is supported by JSPS KAKENHI Grant Number JP17H06785. 2. Structure of the Bost–Connes semigroup actions 2.1. Preliminaries. First of all, we give a quick review of the definition of the Bost–Connes system and related objects. Throughout this paper, we write N for the additive semigroup {n ∈ Z n ≥ 0}. We start with some notational conventions in algebraic number theory. We basically follow the notations of [16]. For a commutative ring R, we use the symbol R∗ for the set of unit elements. For a family of locally compact spaces {Xi}i∈I and compact open subspaces Yi ⊂ Xi, the restricted direct product is defined to be Y′ i∈I (Xi, Yi) := {(xi) ∈Yi∈I Xi xi 6∈ Yi for only finitely many i's}. Let K be a number field with the integer ring OK. We write PK for the set of prime ideals of OK . Let IK denote the set of nonzero integral ideals of K, which forms a semigroup by the multiplication. By the unique prime factorization, it is isomorphic to the free abelian group with the basis PK, pN. Similarly, the group JK of fractional ideals of K is i.e., IK ∼= Lp∈PK isomorphic to the direct sum Lp∈PK Let OK denote the ring Qp∈PK Q′ For each p ∈ PK, the corresponding local field Kp has the integer ring p is isomorphic to Z through the valuation vp. Op, let AK,f denote the ring of finite adeles K,f ∩ OK. A finite idele a ∈ A∗ K,f generates a fractional ideal (a) ∈ JK and this correspondence induces an isomorphism K,f / O∗ A∗ The group A∗ ∼= JK . It restricts to an isomorphism O♮ Op and the quotient K∗ p /O∗ (Kp, Op) and set O♮ K,f of finite ideles acts on Gab K through the Artin reciprocity K / O∗ K ∼= IK . p∈PK K map pZ. K := A∗ φK : A∗ K,f → Gab K . 4 Y. KUBOTA AND T. TAKEISHI Let ¯φK(g) := φK(g)−1. The group A∗ cation, which is denoted by α. Now, the product action α × ¯φK of A∗ AK,f × Gab K,f also acts on AK,f by the multipli- K,f onto K induces a JK -action on the quotient space XK := AK,f × O∗ K Gab K = (AK,f × Gab K )/(α × ¯φK)( O∗ K ). This action is explicitly written as a·[b, γ] := [ba, φK (a)−1γ] for b ∈ AK,f and γ ∈ Gab K , where a is a finite idele with a = (a). It restricts to an IK-action on YK := OK × O∗ K Gab K ⊂ XK. The Bost–Connes groupoid attached to K is the semigroup transformation groupoid GK := YK ⋊ IK. More precisely, GK is the subgroupoid of transformation groupoid XK ⋊ JK associated to the group action XK x JK (in the sense of [4, Example 5.6.3]) defined by GK = {(x, a) ∈ XK ⋊ JK x and a · x are in YK}. Note that the target space G0 K = YK is compact. The Bost–Connes C*-algebra is the associated groupoid C*-algebra AK := r (GK ) (see for example [4, Section 5.6]). In other words, it is the corner C ∗ subalgebra AK = 1YK (C0(XK ) ⋊ JK )1YK of the crossed product C0(X)⋊JK , where 1YK ∈ C0(XK ) is the characteristic function on YK. The dual action of the Pontrjagin dual group JK ∼= (R/Z)∞ on C0(X) ⋊ JK has the property that its restriction to C0(X) is trivial. Hence it restricts to an action on AK . The R-action σK,t on AK is defined as the composition of this JK-action with the dual homomorphism R ∼= R → JK of where NK is the absolute norm function. That is, σK,t is determined by − log NK ( · ) : JK → R, for t ∈ R, f ∈ C(YK) and a ∈ IK. σK,t(f ua) = NK (a)itf ua 2.2. A decomposition of YK . In previous works [17, 18] of the second author, the structure of the closures of orbits of GK were studied in order to determine the primitive ideal space Prim(AK ). Here we revisit them from the viewpoint of the valuation maps. Let N and Z denote N ∪ {+∞} and Z ∪ {+∞} with the order topology pN acts by the product action of N x N. Similarly we define the JK -space (Z, N). Since the valuation vp : Kp → Z is invariant under p on the domain, the composition respectively. Let Yval := NPK , on which the semigroup IK ∼= Lp∈PK Xval := Q′ the action by O∗ p∈PK (cid:0)Yp∈PK vp(cid:1) ◦ pr1 : AK,f × Gab K → Xval, RECONSTRUCTING THE BOST–CONNES SEMIGROUP ACTIONS 5 where pr1 denotes the projection onto the first factor, induces a JK -equivariant proper continuous map valK : XK → Xval. It restricts to an IK-equivariant map YK → Yval. The orbit space Xval/JK with the quotient topology is isomorphic to the power set 2PK with the power-cofinite topology (that is, the family of subsets of the form UF := {C ∈ 2PK C ∩ F = ∅}, where F runs over all finite subsets of PK , forms an open basis of 2PK ) by the map JK · (np) 7→ {p ∈ PK np = ∞}. This is because the power-cofinite topology is nothing but the product topol- ogy of 2 = {0, 1} with the topology {∅, {0}, 2}. For a subset S ⊂ PK , we write X S val for the orbit corresponding to the complement Sc. That is, where X S val := JK · xS 0 ⊂ Xval, 0 )p =(cid:26) 0 (xS if p ∈ S, +∞ if p 6∈ S. We remark that the closure X S corresponds to the fact that val is equal to the union FS′⊂S X S′ val, which {Sc} = {C ∈ 2PK Sc ⊂ C} ⊂ 2PK . The stabilizer subgroup of xS 0 is and the fiber val−1 K (xS 0 ) is canonically isomorphic to J S K :=Mp6∈S pZ ⊂ JK GS K := Gab K /(cid:0)Yp6∈S O∗ p(cid:1), on which J S K acts by multiplication through the homomorphism ¯φS K : J S K → GS K induced from the restriction of ¯φK to the subgroup A∗ For example, G∅ (where P 1 quotient. K := {(k) ∈ JK k ∈ K∗ K is isomorphic to the narrow class group Cl 1 K : JK → Cl 1 +}) and φ∅ p6∈S(Kp, Op). K := JK /P 1 K K is equal to the K,f ∩Q′ Now, we use a canonical choice of a complement JK,S := J Sc K =Mp∈S pZ K . Then, JK ∼= J S K × JK,S and JK,S acts on X S val freely. Therefore, the of J S subspace X S K := val−1 K (X S val) =(cid:16)Yp∈S O♮ p ×Yp6∈S {0}(cid:17) × O∗ K Gab K ⊂ XK 6 Y. KUBOTA AND T. TAKEISHI K is often identified with the space(cid:0)Qp∈S O♮ (afterwards X S equivariantly isomorphic to JK,S × GS through K ) is JK- K on which JK acts by multiplication p(cid:1)× O∗ Gab K idJK,S × ¯φS Indeed, the identification ηS ηS K(a, γ) := [aS K : JK,S × J S K → JK,S × GS K . K : JK,S × GS K → X S 0 · a, φK (a)−1γ] ∈ A∗ (2.1) K is explicitly given by K,f × O∗ Gab K , K where a ∈ A∗ K,f with (a) = a and γ ∈ Gab 0 for the finite adele determined by (aS aS p 6∈ S. K is a lift of γ ∈ GS 0 )p = 1 for p ∈ S and (aS K . Here we write 0 )p = 0 for In summary, we get the following. Lemma 2.2. There is a bijection ηK :=(cid:16) GS⊂PK ηS K(cid:17) : GS⊂PK (JK,S × GS K ) → XK . Moreover, it restricts to a bijection betweenFS⊂PK IK,S :=Lp∈S pN. Next we describe the topology on YK and each Y S decomposition. For S ⊂ S′ ⊂ PK , let IK,S ×GS K and YK, where K in terms of the above qS′,S K : (cid:16) Yp∈S′ Op(cid:17) × O∗ K Gab K →(cid:16)Yp∈S Op(cid:17) × O∗ K Gab K denote the surjection induced from the projection Qp∈S′ Op → Qp∈S Op. Then, by (2.1), the composition K)−1 ◦ qS′,S θS′,S K := (ηS is written as K ◦ ηS′ K : IK,S′ × GS′ K → IK,S × GS K (2.3) θS′,S K (ab, γ) = (a, φS K (b)−1 · πS′,S for ab ∈ IK,S′, where a ∈ IK,S and b ∈ IK,S′ ∩ I S πS′,S K : GS′ K denotes the projection. K → GS K (γ)) For T ⊂ S ⊂ PK, we define the map ΘT from FS′⊂S IK,S′ × GS′ one-point compactification (IK,T × GT K ∪ {∗} by K to the K, and γ ∈ GS K . Here K )+ = IK,T × GT if T ⊂ S′, otherwise, =(cid:26) θS′,T K c∗ ΘT IK ×GS′ K where c∗ denote the constant map to ∗. Lemma 2.4. Let S ⊂ PK. The bijection ηK in Lemma 2.2 gives rise to a homeomorphism between Y S K with the weakest topology such that ΘT is continuous with respect to the usual topology of IK,T × GT K for each T ⊂ PK. K andFS′⊂S IK,S′ ×GS′ K for the setFS′⊂S IK,S′ × GS′ Proof. We write Z S the statement of the lemma. Note that Z S K with the topology given in K is Hausdorff because its topology K → K : Y S is induced from the inclusion QT ΘT . The composition ΘT ◦ η−1 RECONSTRUCTING THE BOST–CONNES SEMIGROUP ACTIONS 7 K )+ is continuous because it is the composition of qS,T (IK,T × GT with the collapsing map Y T bijection from Y S is compact and Z S K to Z S K is Hausdorff. K → (Y T K → Y T K K gives a continuous K, which is actually a homeomorphism because Y S K (cid:3) K )+. That is, η−1 K : Y S For a finite subset F ⊂ PK , the structure of the JK -space X F K , that is, K is well understood in class field theory. the homomorphism φF K : J F K → GF Lemma 2.5. Let F be a finite subset of PK. K is an extension of the narrow class group Cl 1 K by a (1) The group GF quotient of the group Qp∈F O∗ (2) The homomorphism φF p. K factors through an isomorphism lim←−m∈IK,F J F K /P m K → GF K, where IK,F ∼= NF is equipped with the product partial order and K := {(k) ∈ JK k ∈ K∗ P m + such that k ≡ 1 modulo m}. Proof (cf. [13, Proposition 1.1]). Let K∞ :=Qp∞ Kp be the completion at all infinite places and let K o the Artin reciprocity map gives an isomorphism A∗ A∗ ∞ be the connected component of K∗ ∞K∗ → Gab ∞. Then, K , where K /K o := O∗ p and O(n) and U m := U m p := (1 + pn) for n ≥ 1. For m ∈ IK, ∞. Then, since U (m) is an open f × K o K := A∗ K,f × K∗ ∞. We write as O(0) p let U (m) subgroup of A∗ := Qp O(mp) K, we get f p GF K ∼= A∗ K/K∗K o ∞(Yp6∈F p ) ∼= lim←−m O∗ A∗ K/K∗K o ∞U (m) f ∼= lim←−m A∗ K/K∗U (m). CK/C m The right hand side is by definition isomorphic to the projective limit lim←−m K J F K /P m (see [16, Definition VI.1.2, Definition VI.1.7]), which is isomorphic to lim←−m K by [16, Proposition VI.1.9]. These isomorphisms are all JK -equivariant by construction. Now, (1) follows from (2) because GF K /φK (Qp∈F O∗ p ) ∼= G∅ K is isomorphic (cid:3) to Cl 1 K. Finally we get the following reconstruction of the Bost–Connes semigroup action YK x IK . Proposition 2.6. Let K and L be number fields. Let us fix a bijection χ : PK → PL and write jχ : JK → JL and jF for the induced isomorphisms. Assume that there is a family of isomorphisms ΦF : GF K → Gχ(F ) for any finite subset F ⊂ PK such that the diagrams K → J χ(F ) χ : J F L L (2.7) J F K jF χ φF K / / GF K ΦF J χ(F ) L φχ(F ) L / Gχ(F ) L commute. Then, there is a homeomorphism Ψ : YK → YL such that (Ψ, jχ) gives rise to a conjugate of semigroup actions YK x IK and YL x IL.     / 8 Y. KUBOTA AND T. TAKEISHI Proof. In the proof, we omit χ and use the same symbol F for its image in PL for simplicity of notation. First, consider the diagram (2.8) J F K φF K / GF K πF,E K / GE K jF χ ΦF ΦE J F L φF L / / GF L πF,E L / GE L for finite subsets E ⊂ F of PK, where πF,E K → GE K. Then the left square commutes by assumption. The large outer square also commutes because it is a restriction of (2.7) to subgroups J F L . Since φF K has a dense image, the right square also commutes. Recall that θF,E K denotes the quotient GF K is written in (2.3) by using only φF by the commutativity of (2.7) and (2.8), the diagram K . Therefore, K and πF,E K and J F IK,F × GF K θF,E K IK,E × GE K jF χ ×ΦF jE χ ×ΦE IL,F × GF L θF,E L / IL,E × GE L commutes. Hence the map ΨF := GE⊂F jE χ × ΦE : G IK,E × GE K →G IL,E × GE L is a homeomorphism from Y F K to Y F and jχ gives rise to a conjugacy of Y F L by Lemma 2.4. Moreover, each ΨF K x IK and Y E L x IL. Now we get a homeomorphism Ψ : YK → YL as the projective limit of Y F K by the connecting maps , where F runs over all finite subsets of PK. Since K is IK-equivariant, the pair (Ψ, jχ) gives rise to a conjugate of (cid:3) ΨF 's. Indeed, YK is the projective limit lim←−F σF,E each σF,E YK x IK and YL x IL. K := FF ′⊂F θF ′,F ′∩E K 2.3. Subquotients of AK . Here we observe how the decomposition given in Subsection 2.2 is reflected to the structure of the Bost–Connes C*-algebra AK. Throughout this paper, we use the symbol ϕ⋊Γ : A⋊Γ → B ⋊Γ for the ∗-homomorphism between (reduced) crossed product C*-algebras induced from a Γ-equivariant ∗-homomorphism ϕ : A → B. As is proved in [18, Proposition 3.17], there is a continuous surjection ψK : Prim(AK ) → Xval/JK ∼= 2PK , that is, AK has a canonical structure of the C*-algebra over 2PK in the sense of [11] (see also [14, Definition 2.3]). This map is characterized by the property that the pull-back of a JK -invariant open subset U of Xval corresponds to the ideal AK (U ) := 1YK (C0(val−1 K (U )) ⋊ JK )1YK of AK . /   /     / / /     / RECONSTRUCTING THE BOST–CONNES SEMIGROUP ACTIONS 9 This ψK is an intrinsic structure of the C*-algebra AK in the following sense. Lemma 2.9. Let K and L be number fields. Assume that there is an iso- morphism ϕ : AK → AL. Then, there is a bijection χ : PK → PL such that the diagram (ϕ−1)∗ Prim(AK) Prim(AL) ψK 2PK χ ψL / 2PL commutes. That is, ϕ : AK → AL is a ∗-isomorphism over 2PK . Proof. Let Prim2(A) denote the set of second maximal primitive ideals in the sense of [18, Definition 3.9]. It is proved in [18, Proposition 3.11] that Prim2(AK ) ⊂ Prim(A) is a locally compact Hausdorff space and its con- nected components are in one-to-one correspondence with the elements of PK . More precisely, there is a locally constant map ψK : Prim2(AK ) → PK such that ψK (P ) = { ψK (P )}c ∈ 2PK . Therefore, we get a bijection χ : PK ψ−1 K−−→ π0(Prim2(AK ))) (ϕ−1)∗ −−−−→ π0(Prim2(AL)) ψL−−→ PL. Moreover, this choice of χ makes the above diagram commute because for any P ∈ Prim(A) we have ψK (P ) = { ψK (Q) Q ∈ Prim2(A), P ⊂ Q}c ∈ 2PK , which follows from [18, Proposition 3.6]. (cid:3) Similarly, the C*-algebra Aval := 1Yval(C0(Xval) ⋊ JK )1Yval also has the structure of a C*-algebra over 2PK which is characterized by Aval(U ) = 1Yval (C0(U )⋊JK)1Yval for any JK -invariant open subset U ⊂ Xval. Note that it is isomorphic to the tensor product of infinite copies of the Toeplitz algebra T := 1N(C0(Z) ⋊ Z)1N. Moreover, the ∗-homomorphism valK := val∗ K ⋊ JK : C0(Xval) ⋊ JK → C0(XK ) ⋊ JK gives rise to a ∗-homomorphism from Aval to AK , for which we use the same symbol valK . It maps Aval(U ) to AK (U ) for any open subset U ⊂ 2PK , that is, it is a ∗-homomorphism over 2PK . Next we relate the structure of a C*-algebra over 2PK on AK with the decomposition in Lemma 2.2. Following the terminology in [14], we say that a subset of 2PK of the form U \ V , where U, V are open subsets of 2PK , is locally closed. For a locally closed subset Z = U \ V of 2PK , we associate a subquotient AK(Z) := AK(U )/AK (U ∩ V ) / /     / 10 Y. KUBOTA AND T. TAKEISHI of AK, which is independent of the choice of such U and V and has the structure of a C*-algebra over Z. In particular, for a locally closed subset Z and an open subset U of 2PK , we get an exact sequence (2.10) 0 → AK (Z ∩ U ) → AK (Z) → AK(Z \ U ) → 0. Recall that {S} = {T ∈ 2PK S ⊂ T } for S ⊂ PK . For a finite subset F ⊂ PK , {F c} =(cid:16) \E(F {Ec} c(cid:17) \ {F c} c is a locally closed subset. Definition 2.11. We define the C*-algebras BF K := AK({F c}) = 1Y F val := Aval({F c}) = 1Y F BF K (C0(X F (C0(X F K ) ⋊ JK )1Y F val) ⋊ JK )1Y F K , val . val K and BF The C*-algebras BF of AK and Aval respectively. Note that BF the tensor product of C ∗ Moreover, since valK is a ∗-homomorphism over 2PK , we get a ∗-homomorphism val will play the role of composition factors val is canonically isomorphic to K with the compact operator algebra K(ℓ2(IK,F )). r J F Lemma 2.12. Let BF K . Then, there is an isomorphism val → BF K. valF K : BF K ) ⋊ J F K := C(GF K : BF ξF K ⋊ J F K → BF K ⊗ K(ℓ2(IK,F )), K )⊗ id, where πF : GF K is given by the restriction of K ◦ valF such that ξF K = (πF ∗ Proof. The isomorphism ξF ηF ∗ K ⋊JK : C0(X F to the subalgebra BF K ×GF The second claim follows from valK ◦ ηF ∗ K → C0(JK,F ×GF K, where ηF ∗ K )⋊J F K : J F K → pt is the projection. K)⋊JK ∼= (C(GF K → X F K = id ×πF . K )⋊JK )⊗K(ℓ2(JK,F )) K is the map given in (2.1). (cid:3) For a finite subset F ⊂ PK and p ∈ F c, let Fp := F ∪ {p}. Then, the two-point subset {F c, F c p } =(cid:16) \E(F {Ec p} c(cid:17) \ {F c p } c is locally closed. We apply (2.10) for Z = {F c, F c exact sequences p } and U := {F c} c to get (2.13) Note that AK({F c, F c 0 → BFp 0 → BFp K → 0, K → AK ({F c,F c val → Aval({F c,F c p }) and Aval({F c, F c p }) ∼= 1(C0(X F p }) ∼= 1(C0(X F p }) → BF p }) → BF p }) are explicitly written as K ∪ X Fp val ∪ X Fp K ) ⋊ JK )1, val) ⋊ JK )1, val → 0. AK({F c, F c Aval({F c, F c RECONSTRUCTING THE BOST–CONNES SEMIGROUP ACTIONS 11 K ∪ X Fp where X F open subset of X Fp Y Fp val ). K (resp. X F K (resp. X Fp val ∪ X Fp val) and 1 is the constant function on Y Fp val) is equipped with the topology as an K (resp. Definition 2.14. We write ∂F,p K : K∗(BF ∂F,p val : K∗(BF K ) → K∗+1(BFp K ), val) → K∗+1(BFp val), for the boundary homomorphism associated to the exact sequences (2.13). Remark 2.15. Since X F val ∪ X Fp val is identified with the subspace (cid:16)Yq∈F Z(cid:17) × Z+ ×(cid:16) Yq6∈Fp {∞}(cid:17) ⊂(cid:16)Yq∈F Z+(cid:17) × Z+ ×(cid:16) Yq6∈Fp Z+(cid:17) ∼= Xval, the second exact sequence in (2.13) is isomorphic to the tensor product of r J Fp C ∗ K ⊗ K(ℓ2(JK,F )) with the Toeplitz extension 0 → K(ℓ2(pN)) → T → C ∗ r (pZ) → 0. Therefore, ∂F,p is given by the Kasparov product with the KK1-class [T ] ∈ val KK1(C ∗(pZ), C) represented by the Toeplitz extension. (Recall that an ex- tenson of C*-algebras determines an element of the KK1-group. A basic reference is [2, Section 18].) Finally, we discuss the use of KK(X)-theory in the study of the Bost– Connes C*-algebra. Here we omit the detail of KK(X)-theory [11] (see also [14] and [1]) and only remark the following two points. First, for two C*-algebras A, B over a topological space X, a ∗-homomorphism over X from A to B gives an element of KK(X; A, B). Second, an element ϕ ∈ KK(X; AK , AL) induces a family of homomorphisms ϕZ∗ : K∗(A(Z)) → K∗(B(Z)) for any locally closed subsets Z ⊂ X such that the diagrams K∗(A(Z \ W )) K∗(A(Z)) K∗(A(W )) ϕ(Z\W )∗ ϕZ∗ ϕW ∗ K∗(B(Z \ W )) / K∗(B(Z)) / K∗(B(W )) ∂ ∂ K∗+1(A(Z \ W )) ϕZ∗ / K∗+1((B(Z \ W )) commute for any closed subset W ⊂ Z ([14, Definition 2.4], see also [1, Proposition 3.2.1]). Note that a KK(X)-equivalence given by a ∗-isomorphism over X is ordered, that is, each of the induced isomorphism ϕZ∗ gives a bi- jection between positive cones K0(A(Z))+ ∼= K0(B(Z))+. We apply this commutativity for the exact sequences (2.13). First, since valK is a ∗-homomorphism over 2PK , the induced homomorphisms valF K∗ : K∗(Bval) → K∗(BF K) / /   / /   / /     / / / 12 Y. KUBOTA AND T. TAKEISHI make the diagrams (2.16) K∗(BF val) ∂F,p K / K∗+1(BFp val) valF K∗ val Fp K∗ K∗(BF K ) ∂F,p val / / K∗+1(BFp K ) commute. Second, if AL is regarded as a C*-algebra over 2PK by a fixed iden- tification χ : PK → PL and there is an KK(2PK )-equivalence ϕ ∈ KK(2PK ; AK , AL), then it gives a family of isomorphisms ϕF ∗ : K∗(BF K) → K∗(Bχ(F ) L ) such that the diagrams (2.17) K∗(BF K ) ϕF ∂F,p K K∗+1(BFp K ) ϕFp K∗(Bχ(F ) L ) ∂χ(F ),χ(p) L / K∗+1(Bχ(Fp) L ) commute for each finite subset F ⊂ PK . Given the ∗-isomorphism ϕ : AK → AL and χ given in Lemma 2.9, we have such a KK(2PK )-equivalence. We remark that the discussion in this paragraph shows (4) ⇒ (5) ⇒ (6) of Theorem 1.1. 3. Reconstructing profinite completions from K-theory of the crossed product C*-algebra In this section we study C*-algebras of the form C(G) ⋊ Γ, where Γ is a countable free abelian group and G is its profinite completion. Our goal is to show that its K-group remembers the completion, that is, the homomorphism Γ → G. For simplicity of notation, we use the same symbol π for quotients of compact abelian groups when the domain and range are specified. Throughout this paper, for a C*-algebra A we use the symbol K∗(A) for the Z/2-graded group K0(A) ⊕ K1(A). The tensor product of K∗-groups is also taken in the category of Z/2-graded abelian groups. Let N be a set of rational prime numbers. A group G is a pro-N group if it is a projective limit of finite groups whose orders are factorized as a product of primes in N . A pro-N completion of a discrete group Γ is a pro- N group G equipped with a group homomorphism f : Γ → G whose image is dense in G. In other words, G is a projective limit of finite quotients Γ/Γn of Γ, where Γn is a decreasing sequence of normal subgroups of Γ such that [Γ : Γn] is factorized as a product of primes in N . We remark that we do not assume that the homomorphism f is injective. For example, the quotient f : Γ → Γ/Π is a pro-N completion if [Γ : Π] is factorized as the product of primes in N . Hereafter we deal with a finitely generated pro-N completion of a count- able free abelian group. We say that a profinite group G is finitely generated /     / /     / RECONSTRUCTING THE BOST–CONNES SEMIGROUP ACTIONS 13 if it is topologically finitely generated, that is, there is a finite family of el- ements of G spanning a dense subgroup. For a set N of rational prime numbers, Z[N −1] denotes the smallest subring of Q containing p−1 for all p ∈ N and M [N −1] := M ⊗Z Z[N −1] for an abelian group M . Lemma 3.1. Let N be a set of rational prime numbers and let G be a finitely generated pro-N group. Then, G is decomposed as the productQp∈N Gp such p × Fp for dp ∈ Z>0 and a finite p-group F . that each Gp is isomorphic to Zdp Proof. We use the notation N n for pn for all p ∈ N . The group G is canonically regarded as a finitely generated module over the ring Z/nZ ∼=Yp∈N lim←− N n Zp. Let 1Zp denote the unit of Zp, which is an idempotent in Q Zp, and set Gp := 1Zp · G. Then G ∼= Q Gp and each Gp is a finitely generated Zp- module. Hence we get the conclusion by the structure theorem for finitely generated modules over a PID (principal ideal domain). (cid:3) We represent the order of profinite groups by using supernatural numbers like G =Q plp, where lp = logp Fp if dp = 0 and lp = ∞ if dp ≥ 1. Lemma 3.2. Let F be a finite subset of PK. Then, there is a finite set NF of rational prime numbers such that φF K is a pro-NF completion. K → GF K : J F Proof. For a prime p ∈ PK over a rational prime number p, the group U (1) principal units of Kp is a pro-p group and the quotient O∗ is finite (see for example [16, Proposition II.5.3]). Let Np denote the union of {p} with the set of prime numbers dividing O∗ p is a pro-Np group. Let NF denote the union ofSp∈F Np with the set of prime numbers dividing K. Then GF K has dense image by Lemma 2.5 (2). K is a pro-NF group by Lemma 2.5 (1). Moreover, (cid:3) K := Cl 1 h1 the map φF . Then, O∗ p/U (1) p /U (1) of p p p 3.1. K-groups of C(G) ⋊ Γ. For the calculation of K∗(C(G) ⋊ Γ), we start with the case that G is finite, that is, G ∼= Γ/Π for a finite index subgroup Π. First of all, we review a special case of Green's imprimitivity theorem [8, Theorem 17]. Let σ denote the regular representation of Γ to ℓ2(Γ/Π). Recall that C(Γ/Π) ⋊ Γ ∼= span[(C(Γ/Π) ⊗ 1) · ((σ ⊗ id)(C ∗ r Γ))] ⊂ K(ℓ2(Γ/Π)) ⊗ C ∗ r Γ. We write κ for this inclusion. Let p ∈ C(Γ/Π) be the support function on 0 ∈ Γ/Π. Then the ∗-homomorphism j := p ⊗ idC∗ r Γ : C ∗ r Γ → K(ℓ2(Γ/Π)) ⊗ C ∗ r Γ bijects the subalgebra C ∗ quently, j0 := jC∗ r Π induces the isomorphism K∗(C ∗ r Π onto the full corner p(C(Γ/Π) ⋊ Γ)p. Conse- r Π) ∼= K∗(C(Γ/Π) ⋊ Γ). 14 Y. KUBOTA AND T. TAKEISHI Lemma 3.3. Let Γ and Π be as above. Let ι : C ∗ r Γ denote the inclusion, let π : Γ/Π → pt denote the quotient and let π∗ ⋊ Γ : C ⋊ Γ → C(Γ/Π) ⋊ Γ denote the induced ∗-homomorphism. Then, the composition r Π → C ∗ ι∗ ◦ (j0)−1 ∗ ◦ (π∗ ⋊ Γ)∗ : K∗(C ∗ r Γ) → K∗(C ∗ r Γ) is multiplication by [Γ : Π]. In particular, (π∗ ⋊ Γ)∗ is injective. Proof. By definition j∗ gives a canonical identification of K∗(C ∗ K∗(K(ℓ2(Γ/Π)) ⊗ C ∗ r Γ). Since j ◦ ι = κ ◦ j0, we get ∗ ◦ (π∗ ⋊ Γ)) = κ∗ ◦ (π∗ ⋊ Γ)∗ = (σ ⊗ idC∗ j∗ ◦ (ι∗ ◦ (j0)−1 r Γ)∗. r Γ) with Since σ is homotopic to the trivial representation onto the [Γ : Π]-dimensional vector space ℓ2(Γ/Π) (because Γ is connected), the right hand side is multi- plication by [Γ : Π]. (cid:3) Next, we give a more explicit calculation of the K-group. r Γ) ∼= K∗(Γ) and the exterior algebra V∗Γ (actually, this is It is well- known in topological K-theory that there is a canonical isomorphism be- tween K∗(C ∗ an isomorphism as Hopf algebras). Here, each element of Γ is of odd de- gree, that is, K0(C ∗ In the context of K-theory of C*-algebras, this isomorphism is understood in terms of the Kasparov product in the following way. First, the K-group of the C*-algebra of the free abelian group generated by a single element v is r Γ) ∼= VevenΓ and K1(C ∗ r Γ) ∼= VoddΓ. K∗(C ∗ r (Zv)) ∼= Z[1] ⊕ Zβv ∼=V∗(βv), where the Bott element βv ∈ K1(C ∗ r (Zv)) is represented by the unitary uv. For an independent family of elements v1, . . . , vk ∈ Γ, the Kasparov product determines the element βv1 ⊗ . . . ⊗ βvk ∈ K∗(Ni C ∗ We use the same letter for its image in K∗(C ∗ of Γ, we get the homomorphism r (Zvi)) ∼= K∗(C ∗ r (Li Zvi)). r Γ). By choosing a basis {vi}i (3.4) V∗(βv1 , βv2 , . . . ) ∋ βvi1 ∧ · · · ∧ βvik 7→ βvi1 ⊗ . . . ⊗ βvik ∈ K∗(C ∗ r Γ), which is well-defined by the graded commutativity of the Kasparov product [10, Theorem 5.6]. It is actually an isomorphism due to the Kunneth formula (recall that the Kunneth homomorphism K∗(A) ⊗ K∗(B) → K∗(A ⊗ B) is nothing but the Kasparov product as above). We canonically identify the left hand side with V∗Γ by the correspondence v 7→ βv. For a rank k free abelian group equipped with an orientation, the element βΣ := βv1 ⊗ . . . ⊗ βvk ∈ K∗(C ∗ r Σ) is independent of the choice of an oriented basis {vi}i and generates VkΣ ∼= Z. Hereafter, we use the same symbol βΣ for its image in K∗(C ∗ r Γ) if Σ is an oriented direct summand (that is, a direct summand with a fixed orientation) of Γ. Lemma 3.5. Let Π be a finite index subgroup of Γ. Through the isomor- r Γ) is phism K∗(C ∗Γ) ∼= V∗Γ, the homomorphism ι∗ : K∗(C ∗ identified with the inclusion V∗Π →V∗Γ. r Π) → K∗(C ∗ RECONSTRUCTING THE BOST–CONNES SEMIGROUP ACTIONS 15 Proof. When Γ = Zv and Π = nZv, the statement can be checked directly. Actually, ι∗([1]) = [1] and ι∗(βnv) = βnv = nβv. For general Γ and Π, let us choose a basis {vi}i of Γ such that Π =Li niZvi. Now the claim follows from the functoriality of the Kunneth isomorphism. (cid:3) In particular, ι∗ is an isomorphism after tensoring with Z[N −1] if [Γ : Π] is factorized as primes in N . Together with Lemma 3.3, we can see that so is (π∗ ⋊ Γ)∗ : K∗(C ∗ r Γ) → K∗(C(Γ/Π) ⋊ Γ). Now we go back to the study of the K-group of C(G) ⋊ Γ for general G. Lemma 3.6. Let Γ be a countable free abelian group, let ϕ : Γ → G be a pro-N completion and let π : G → pt denote the quotient. Then, (π∗ ⋊ Γ)∗ : K∗(C ∗ r Γ) → K∗(C(G) ⋊ Γ) is an isomorphism after tensoring with Z[N −1]. Proof. Since C(G)⋊ Γ is isomorphic to the inductive limit lim−→k it follows from the above observations. C(Γ/Γk)⋊ Γ, (cid:3) Lemma 3.6 means that the K-group of C(G) ⋊ Γ determines an interme- diate subgroup (3.7) V∗Γ ⊂ K∗(C(G) ⋊ Γ) ⊂V∗Γ[N −1]. Here we simply write V∗Γ[N −1] for (V∗Γ)[N −1] = V∗(Γ[N −1]). In the following subsections, we observe that this data has rich information, enough to reconstruct the pro-N completion f : Γ → G. Hereafter we often regard K∗(C(G) ⋊ Γ) as a subgroup of QΓ and omit the homomorphism (π∗ ⋊ Γ)∗ for simplicity of notations. Definition 3.8. Let Γ be a countable free abelian group and let f : Γ → G be a profinite completion of Γ. For k ∈ Z>0, we write Kk Γ(G) := K∗(C(G) ⋊ Γ) ∩VkQΓ. We remark that Lemma 3.3 means that 1 Kn (3.9) [Γ : Π] Γ(Γ/Π) = ·VnΠ as subgroups of VnQΓ because the composition restricts to the standard inclusion V∗Γ → V∗QΓ. Therefore, for a general −−−−→V∗QΓ ∼=−→V∗Π Γ(Γ/Π) 1 [Γ:Π] ι∗ K∗ profinite completion G = lim←− Γ/Γk, we get 1 (3.10) Kn Γ(G) = ∞[k=1(cid:16) [Γ : Γk] ·VnΓk(cid:17). In particular, we get a direct sum decomposition K∗(C(G) ⋊ Γ) =Mn Kn Γ(G). 16 Y. KUBOTA AND T. TAKEISHI 3.2. Reconstructing pro-p completions. We start with the case that N = {p}. Hereafter, we use the following symbol: for an element x of an abelian group M , we define the supernatural number δ(x, M ) =Qp plp(x,M ) to be lp(x, M ) := sup{l ∈ Z>0 x ∈ pkM } ∈ Z>0 ∪ {∞}. The following lemma is a direct consequence of Lemma 3.3. Lemma 3.11. Let f : Γ → G be a pro-p completion of a countable free abelian group. Let Σ be an oriented rank d direct summand of Γ. Then δ(βΣ, Kd Γ(G)) = G/f (Σ). Proof. Let {Γk}k be a decreasing sequence of subgroups of Γ such that f factors through an isomorphism lim←− Γ/Γk → G. By (3.10), we have δ(βΣ, Kd Γ(G)) = sup k δ(βΣ, Kd Γ(Γ/Γk)). On the other hand, G/f (Σ) is equal to the supremum of Γ/(Γk + Σ). Hence the proof of the lemma is reduced to the case that G = Γ/Π. Let us choose an oriented basis v1, . . . , vd of Σ such that Σ∩Π =Li niZvi. We remark that Σ ∩ Π is a direct summand of Π. By Lemma 3.3, ι∗ ◦(j0)−1 (π∗ ⋊ Γ)∗(βΣ) coincides with ∗ ◦ [Γ : Π]βv1 ∧ · · · ∧ βvd = = [Γ : Π] n1 . . . nd [Γ : Π] n1 . . . nd βn1v1 ∧ · · · ∧ βndvd ι∗(βΣ∩Π) and hence (j0)−1 ∗ ◦ (π∗ ⋊ Γ)(βΣ) = [Γ : Π] n1 . . . nd βΣ∩Π. Since δ(βΣ∩Π,VdΠ) = 1, we get δ(βΣ, Kd Γ(Γ/Π)) = δ((j0)−1 [Γ : Π] n1 . . . nd = ∗ ◦ (π∗ ⋊ Γ)(βΣ),V∗Π) [Γ : Π] = [Σ : Σ ∩ Π] = Γ/(Π + Σ). (cid:3) Lemma 3.12. Let Γ be a free abelian group and let f : Γ → G be a pro-p completion such that G ∼= Zp. Then, f factors through the isomorphism Γ/(Γ ∩ pkK1 Γ(G)) → G. lim←− k→∞ Proof. Since G ∼= lim←− it suffices to show that Γ ∩ K1 k G/pkG ∼= lim←− Γ(G) = f −1(pkG). k Γ/f −1(pkG), We show that x ∈ pkK1 out loss of generality we may assume δ(x, Γ) = 1 because δ(pkx, K1 pkδ(x, K1 Γ(G)). Lemma 3.11 implies that Γ(G) if and only if f (x) ∈ pkG for x ∈ Γ. With- Γ(G)) = δ(βx, K1 Γ(G)) = G/f (Zx). RECONSTRUCTING THE BOST–CONNES SEMIGROUP ACTIONS 17 Recall that a closed subgroup f (Zx) of G ∼= Zp is of the form plG for some l ≥ 0. Now we get the conclusion because pk divides G/f (Zx) if and only if f (x) ∈ pkG. (cid:3) For d ∈ Z>0, let Sd(Γ, G) denote the set of oriented rank d direct sum- mands Σ of Γ such that there exists x ∈ Γ \ Σ with δ(βx ∧ βΣ, Kd+1 Γ (G)) = 1. Note that such x satisfies δ(x, Γ) = 1 and hence Zx ⊕ Σ is also a direct summand of Γ. By the fundamental theorem of finitely generated modules over a PID and Lemma 3.11, an oriented direct summand Σ is in Sd(Γ, G) if and only if G/f (Σ) is a singly generated Zp-module. For an oriented rank d direct summand Σ of Γ, let βΣ ∧ · denote the endo- morphism on V∗Γ[p−1] taking the exterior product with βΣ. In particular, it induces a homomorphism from Γ[p−1] ∼=V1Γ[p−1] to Vd+1Γ[p−1]. Lemma 3.13. Let Γ be a free abelian group, G := Zd a pro-p completion. Set p and let f : Γ → G be Γk := Γ ∩ \Σ∈Sd−1(Γ,G) (βΣ ∧ · )−1(pkKd Γ(G)). Then, f factors through the isomorphism lim←− Γ/Γk Proof. For Σ ∈ Sd−1(Γ, G), let GΣ := G/f (Σ) and let fΣ denote the com- position of f with the quotient G → GΣ. Note that GΣ is isomorphic to Zp since the Zp-rank of GΣ is equal to 1 and GΣ is singly generated. ∼= G. We claim that (βΣ ∧ · )−1(Kd Γ(G)) = K1 Γ(GΣ). Indeed, Lemma 3.11 implies that δ(βx ∧ βΣ, Kd Γ(G)) = G/f (Σ ⊕ Zx) = GΣ/f (Zx)) = δ(βx, K1 Γ(GΣ)) = p∞ for x ∈ Σ. Γ(GΣ)) for x ∈ Γ \ Σ with δ(x, Γ) = 1 and δ(βx, K1 Now fΣ factors through the isomorphism Γ Γ ∩ (βΣ ∧ · )−1(pkKd Γ(G)) lim←− k → GΣ by Lemma 3.12. Hence the direct product GΣ For the proof of the lemma, it suffices to show that the product of pro- Y fΣ : Γ → YΣ∈Sd−1(Γ,G) factors through an injection lim←− Γ/Γk →Q GΣ. jections G → Q GΣ is injective. Since f has dense image, there is a finite basis of the Zp-module G. Now, Σi := Lj6=i Zvj for i = 1, . . . , d satisfies Σi ∈ Sd−1(Γ, G) and G →Qi GΣi is injective. family {v1, . . . , vd} of elements of Γ such that {f (v1), . . . , f (vd)} forms a free Lemma 3.14. Let Γ be a free abelian group and let f : Γ → G be a pro- p completion such that G is finitely generated. Let F denote the torsion subgroups of G. (cid:3) (1) Let d be the minimal integer such that Kd of VdΓ[p−1]. Then, it is equal to the rank of G. Γ(G) is a proper subgroup 18 Y. KUBOTA AND T. TAKEISHI (2) Let Λ be an oriented rank d direct summand of Γ minimizing δ(βΛ, Kd and set H := f (Λ). Then, G/H is isomorphic to F . In particular, N := δ(βΛ, Kd (3) The subgroup Γ(G)) is equal to F . Γ(G)) Π := {x ∈ Γ N −1βx ∧ βΛ ∈ Kd+1 Γ (G)} is equal to f −1(H). (4) Under the identification V∗QΠ ∼=V∗QΓ induced from the inclusion Π → Γ, we get K∗ Π(H) = N · K∗ Γ(G). Consequently, Sd(Π, H) consists of oriented direct summands Σ of Π such that there exists x ∈ Π with δ(ι∗(βx ∧ βΣ), Kd Γ(G)) = N . Proof. Since f has dense image, there is a finite family {vi}i∈I of elements of Γ such that {f (vi)}i∈I freely generates a submodule M of G such that G ∼= M × F . Now we get δ(βΛ, Kd Γ(G) ( Γ(G)) = Vd[p−1]. On the other hand, Lemma 3.11 also implies that δ(βΣ, K∗ Γ(G)) < ∞ by Lemma 3.11 and hence Kd p∞ for any direct summand Σ of the rank less than d. In fact, this choice of Λ actually minimizes δ(βΛ, Kd Γ(G)) as in the state- ment of (2). This follows from the fact that the order of the quotient of G by any free subgroup of the same rank divides F . Now (2) follows from Lemma 3.11. Next we show (3). Let π : G → G/H denote the quotient. It is obvious that both Π and f −1(H) contain Λ. On the other hand, for x ∈ Γ \ Λ with x = ply and δ(y, Γ) = 1, we have δ(βx ∧ βΛ, Kd+1 Γ (G)) = plδ(βy ∧ βΛ, Kd+1 Γ (G)) = plG/f (Zy ⊕ Λ) = plF/hπ ◦ f (y)i by Lemma 3.11. The right hand side is equal to F = N if and only if the order of f (y) divides pl, that is, f (x) = 0. Finally, (4) immediately follows from (3.10) as K∗ Γ(G) = = ∞[k=1(cid:16) 1 [Γ : Π] 1 [Γ : Γk] ·V∗Γk(cid:17) ∞[k=1(cid:16) 1 [Π : Γk] ·VnΓk(cid:17) = N −1K∗ Π(H), if we choose {Γk} as Γ1 = Π. (cid:3) Theorem 3.15. Let f : Γ → G, d, Λ, N and Π be as in Lemma 3.14. Set Πk := Π ∩ \Σ∈Sd−1(Π,H) (βΣ ∧ · )−1(pkN · Kd Γ(G)). Then, f factors through an isomorphism lim←− Γ/Πk → G. Proof. By construction, H is a free Zp-module. Therefore, f : Π → H factors Π/Πk → H by Lemma 3.13 and Lemma 3.14 through the isomorphism lim←−k (4). The result follows because G = Γ ×Π H. (cid:3) RECONSTRUCTING THE BOST–CONNES SEMIGROUP ACTIONS 19 Corollary 3.16. Let Γ be a free abelian group, let fi : Γ → Gi (i = 1, 2) be two pro-p completions such that Gi are finitely generated and let πi : Gi → pt denote the quotient. Suppose that we have the isomorphism ϕ : K∗(C(G1) ⋊ Γ) → K∗(C(G2) ⋊ Γ) such that ϕ ◦ (π∗ Φ : G1 → G2 such that Φ ◦ f1 = f2. 1 ⋊ Γ)∗ = (π∗ 2 ⋊ Γ)∗. Then, there is a group isomorphism Γ(G1) = K∗ Proof. The assumption means that K∗ Λ, N , Π and Sd−1(Π, H) used to define the decreasing sequence {Πk} de- groups of V∗Γ ⊂ V∗QΓ. It is checked in Lemma 3.14 that all the data d, Γ(G) ⊂VQΓ. Therefore pend only on the intermediate subgroup V∗Γ ⊂ K∗ we get isomorphisms fi : lim←−k isomorphism. Γ/Πk → Gi and Φ := f2 ◦ f −1 is the desired (cid:3) Γ(G2) as intermediate sub- 1 3.3. Reconstructing pro-N completions. Let N be a subset of prime numbers and let f : Γ → G be a pro-N completion such that G is finitely generated. By Lemma 3.1, the group G is decomposed as Gp1 × · · · × Gpk . For p ∈ N , let πp : G → Gp be the projection. Lemma 3.17. For p ∈ N , the subgroup K∗ to K∗ Γ(Gp). Γ(G) ∩V∗Γ[1/p] ofV∗QΓ is equal Proof. Let Np := N \ {p}. Since K∗ Γ(G) ∩ Γ[1/p], it suffices to show that the inclusion K∗ Γ(G) induces an isomorphism after tensoring with Z[N −1 ]. Moreover, by (3.10) it is enough to consider the case that G = Γ/Π. Let Πp be the subgroup of Γ such that Γ/Πp ∼= Gp, that is, [Γ : Πp] is a power of p and p does not divide [Πp : Π]. Then Πp[N −1 Γ(Gp) is included in K∗ Γ(Gp) ⊂ K∗ ] = Π[N −1 ] and hence p p p K∗ Γ(G)[N −1 p ] =(cid:16) = 1 [Γ : Π]V∗Π(cid:17)[N −1 [Γ : Πp]V∗Πp[N −1 1 p p ] = K∗ Γ(Gp)[N −1 p ]. ] (cid:3) Corollary 3.18. Let fi : Γi → Gi (i = 1, 2) be two pro-N completions of free abelian groups and let πi : Gi → pt denote the quotient. Suppose that we have an isomorphism F : Γ1 → Γ2 and ϕ : K∗(C(G1) ⋊ Γ) → K∗(C(G2) ⋊ Γ) such that the diagram K∗(C ∗ r Γ1) F∗ K∗(C ∗ r Γ2) (π∗ 1 ⋊Γ1)∗ K∗(C(G1) ⋊ Γ1) ϕ (π∗ 2 ⋊Γ2)∗ / K∗(C(G2) ⋊ Γ2) commutes. Then, there is a group isomorphism Φ : G1 → G2 such that Φ ◦ f1 = f2 ◦ F . / /     / 20 Y. KUBOTA AND T. TAKEISHI Proof. First, by replacing f2 with f2 ◦ F , we may assume that Γ1 = Γ2 and F = id without loss of generality. For p ∈ N and i = 1, 2, let πi,p : Gi → (Gi)p denote the quotient. By Lemma 3.17, ϕ induces an isomorphism ϕp : K∗ 2,p ⋊ Γ)∗. We apply Corollary 3.16 to get an isomorphism Φp : G1,p → G2,p. Finally, Φ := (cid:3) Γ(G2,p) such that (π∗ 1,p ⋊ Γ)∗ ◦ ϕp = (π∗ Γ(G1,p) ∼= K∗ Qp∈N Φp is the desired isomorphism. 4. Reconstructing the Bost–Connes semigroup action In this section we give a proof of Theorem 1.1. Throughout this section, we fix a total order on PK in order to fix the orientation on direct summands JK,F for finite subsets F ⊂ PK . For finite subsets F and F ′ of PK with F ∩ F ′ = ∅, we use the symbol βF K ). We write πF F ′ for the element βJK,F ′ ∈ K0(C ∗ K for the projection GF K → pt. r J F The essential step is (6)⇒(1). Here we reconstruct the semigroup action K) and homomorphisms K ). First of all, recall that we have an ordered YK x IK from the family of ordered groups K∗(BF ∂F,p K : K∗(BF isomorphism K) → K∗+1(BFp ξF K∗ : K∗(BF K ) → K∗(BF K ), where ξF profinite completions φF K is as in Lemma 2.12. We will apply Corollary 3.18 to reconstruct K . To this end, we need to reconstruct K ) from boundary homomorphisms. Recall K is called the narrow class K ) ⋊ JK and JK /P 1 the inclusion V∗J F K = C(JK/P 1 K → K∗(BF K → GF K =: h1 K : J F that B∅ number. Lemma 4.1. The boundary homomorphism ∂F,p the equalities K is uniquely determined by ∂F,p K (valF ∂F,p K (valF K∗(βF K∗(βF F ′ p F ′)) = 0, )) = (−1)N (F ′,p)+1val Fp K∗(βFp F ′ ), for any finite subset F ′ ⊂ F c {q ∈ F ′ p < q}. p . Here, N (F ′, p) denotes the inversion number Proof. By (3.7), a homomorphism from K∗(BF uniquely determined by the image of its subgroup (πF ∗ K ⋊ J F This shows the uniqueness of homomorphisms from K∗(BF with the above equalities since valF K ⋊ J F in Lemma 2.12. K ) to a torsion-free group is K)∗(K∗(C ∗J F K )). K ) to K∗+1(BFp K ) K )∗, as is shown K∗)−1 ◦ (πF ∗ K∗ = (ξF Since the diagram (2.16) commutes, it suffices to show ∂F,p val (βF F ′) = 0, ∂F,p val (βF F ′ p ) = (−1)N (F ′,p)+1βFp F ′ . To see this, recall that the identification K∗(C ∗ Kasparov product as in (3.4). In particular, we have r JK ) ∼=V∗JK is given by the F ′ = βFp βF F ′ ⊗[1C∗(pZ)], βF F ′ p = (−1)N (F ′,p)βFp F ′ ⊗ βp RECONSTRUCTING THE BOST–CONNES SEMIGROUP ACTIONS 21 as elements of K∗(C ∗ the associativity of the Kasparov product, we get K ) ∼= K∗(C ∗ K ⊗ C ∗ r J F r J Fp r (pZ)). Now, by Remark 2.15 and F ′ ⊗[1C∗ F ′) = (βFp ∂F,p val (βF F ′ ∧ βp) = (−1)N (F ′,p)(βFp ∂F,p val (βF r (pZ)] ⊗C∗ r (pZ)[T ] = 0 and βp ⊗C∗ r (pZ)]) ⊗C∗ since [1C∗ r (pZ)[T ] = 0, F ′ ⊗ βp) ⊗C∗ r (pZ)[T ] = (−1)N (F ′,p)+1βFp F ′ r (pZ)[T ] = −1 ∈ KK(C, C). (cid:3) Let F = {p1, . . . , pl} be a finite subset of PK with p1 ≤ · · · ≤ pl. We write as Fi := {pl−i+1, . . . , pl} ⊂ F and DF K := ξF K ◦ ∂Fl−1,p1 K ◦ · · · ◦ ∂F1,pl−1 K Similarly we define DF val : K∗(C ∗ ◦ ∂∅,pl K ◦ (ξ∅ r JK) → K∗+l(C ∗ K)−1 : K∗(B∅ r J F K ). K) → K∗+l(BF K ). Lemma 4.2. There is a unique ordered homomorphism such that the image (τ F the narrow class number h1 K ◦ DF K . K ) → R τ F K : K0(BF K )(Kl(B∅ K )) is Z. Moreover, it maps [1BF ] to K Proof. For a free abelian group Γ and its finite index subgroup Π, the K0- group of the C*-algebra C(Γ/Π) ⋊ Γ ∼= C ∗ r (Π) ⊗ MΓ/Π admits a unique ordered homomorphism to R up to scalar multiplication (see [2, Exercise 6.10.3]). Hence the inductive limit K0(BF K ) = lim−→ m∈IK,F K0(C(J F K /P m K ) ⋊ J F K ) (we remark that this equality is obtained from Lemma 2.5 (2)) also admits a unique ordered homomorphism to R up to scalar multiplications. We take the unique ordered homomorphism τ F ]) = K ([1BF h1 K . Then, the composition τ F K · Z. This is because it coincides with the map induced from the canonical trace on K∗(C ∗ K by the uniqueness of ordered homomorphisms from K∗(C ∗ K such that τ F K ⋊ JK )∗ has the image h1 r JK ) multiplied with h1 K ) to R mapping [1C∗ K . Hence we obtain K ◦ (πF ∗ ] to h1 r J F K r J F K h1 K(τ F K ◦ DF K )(Kl(B∅ K )) ⊂ (τ F = (τ F ⊂ (τ F K ◦ DF K ◦ (πF ∗ K ◦ (πF ∗ K ) ⊂ (π∅∗ KKl(B∅ K )((π∅∗ K ⋊ J F K ⋊ J F K ⋊ JK)∗(Kl(B∅ val(Kl(B∅ K )∗)(DF K )∗)(K0(BF val))) val))) val)) ⊂ h1 K · Z. K ⋊ JK)∗(Kl(C ∗ r JK)), which follows Here we use the fact h1 from (3.9). The remaining task is to show that the image of τ F K to Cl 1 Z. Recall that φ∅ K surjects J F each pi ∈ F there is ai ∈ J F K such that ai · pi ∈ P 1 ∗-homomorphisms as in Lemma 3.3 for the inclusion P 1 ι = val∅ K). Then the element K contains 1 ∈ K by Lemma 2.5 (2). Therefore, for K . Let ι and j0 be K ⊂ JK (note that K ◦ DF ζ := βa1·p1 ⊗ . . . ⊗ βal·pl = (βp1 + βa1 ) ⊗ . . . ⊗(βpl + βal) ∈ K∗(C ∗ r P 1 K ) 22 Y. KUBOTA AND T. TAKEISHI satisfies ι∗(ζ) = h1 β∅ (for i = 1, . . . , N ), Lemma 4.1 implies that K · j0(ζ) ∈ K∗(B∅ K ) by Lemma 3.3. Since ζ is written as by ci ∈ Z and finite subsets Gi of PK satisfying F 6⊂ Gi F +PN i=1 ciβ∅ Gi DF K (ι∗(ζ)) =DF K(cid:0)val∅ K∗(cid:0)β∅ F + NXi=1 ciβ∅ Gi(cid:1)(cid:1) =DF K (val∅ K∗(β∅ F )) = (−1)lvalF K∗(βF ∅ ) = ι∗[1C∗ r J F K Consequently we get (τ F K ◦ DF K)((−1)lj0(ζ)) = 1. ]. (cid:3) Now we assume that there is a family of isomorphisms ϕF : K∗(BF K) → ◦ ϕF ◦ K)−1. By the commutativity of the diagrams (2.16) and (2.17) and the ) as in the condition (6) of Theorem 1.1. Let ϕF := ξχ(F ) L L K∗(Bχ(F ) (ξF definition of DF K , the diagrams K∗(C ∗ r JK ) DF val / K∗+l(C ∗ r J F K) K∗(B∅ K ) DF K / K∗+l(BF K ) (4.3) π∅∗ K ⋊JK πF ∗ K ⋊J F K ϕ∅ ϕF K∗(B∅ K ) DF K / / K∗+l(BF K ), K∗(B∅ L) also commute. Dχ(F ) L / / K∗+l(Bχ(F ) ) L Proposition 4.4. Assume the condition (6) of Theorem 1.1. (1) The isomorphism jχ : JK → JL induced from χ restricts to an iso- morphism P 1 (2) The diagram K → P 1 L. K∗(C ∗ r JK ) (jχ)∗ K∗(C ∗ r JL) (π∅∗ K ⋊JK )∗ (π∅∗ L ⋊JL)∗ K∗(B∅ K ) ϕ∅ / K∗(B∅ L) commutes. Proof. By Lemma 4.1 and Lemma 4.2, we have (τ F K ◦ DF K )(β∅ F ′) =(cid:26) h1 0 K if F = F ′, otherwise. Therefore, the homomorphism ΨK from K∗(B∅ K ) to V∗JK determined by ΨK(x) :=XF (−1)F (τ F K ◦ DF K )(x) · β∅ F , where F runs over all finite subsets of PK , satisfies that the composition V∗JK ∼= K∗(C ∗ is the multiplication by h1 ι∗ ◦ (j0)−1, where ι∗ and j0 are as in Lemma 3.3 for the inclusion P 1 r JK ) K . Comparing it with Lemma 3.3 we get ΨK = K ⊂ JK. K) ΨK−−→V∗JK K ⋊JK)∗ (π∅∗ −−−−−−−→ K∗(B∅ /     /     / /     / RECONSTRUCTING THE BOST–CONNES SEMIGROUP ACTIONS 23 Therefore we get Im ΨK =V∗P 1 (4.5) K and in particular Im ΨK ∩V1JK = P 1 K . Consider the diagram (4.6) K∗(C ∗ r JK ) (jχ)∗ K∗(C ∗ r JL) (π∅∗ K ⋊JK )∗ (π∅∗ L ⋊JL)∗ K∗(B∅ K ) ϕ∅ / K∗(B∅ L) ΨK ΨL ∧∗jχ V∗JK /V∗JL The right square commutes by the commutativity of the right diagram in (4.3) and the uniqueness of τ F K in Lemma 4.2. Hence we get (1) by (4.5). In particular we get h1 K = h1 L, which shows that the outer square of (4.6) commutes. Since ΨK and ΨL are injective, a diagram chasing shows that the left square also commutes. (cid:3) Proof of Theorem 1.1. The steps (1)⇒(2)⇒(4), (1)⇒(3)⇒(4) are obvious. (4) ⇒ (5) ⇒ (6) is explained at the last paragraph of Subsection 2.3. We show (6)⇒(1). In the proof, for a finite subset F , we omit χ and use the same symbol F for its image in PL for simplicity of notation. Consider the diagram K∗(C ∗ r JK) DF val (PPPPPPPPPPPP π∅∗ K ⋊JK (1) πF ∗ K ⋊J F K / K∗(B∅ K ) w♣♣♣♣♣♣♣♣♣♣♣ DF K K∗+l(BF K ) (jχ)∗ (2) (jχ)∗ (5) ϕF (4) ϕ∅ DF val 6♥♥♥♥♥♥♥♥♥♥♥♥♥ K∗(C ∗ r JL) πF ∗ L ⋊J F (3) L π∅∗ L ⋊JL / K∗+l(BF L ) g◆◆◆◆◆◆◆◆◆◆◆ DF L K∗(B∅ L). K∗+l(C ∗ r J F K ) K∗+l(C ∗ r J F L ) We have already proved the commutativity of the large outer square and diagrams (1), (3), (4) at Proposition 4.4 (2) and (4.3) respectively. The diagram (2) also commutes by definition. Since DF val is surjective, a diagram chasing shows that the diagram (5) also commutes. Finally, with the help of Lemma 3.2 we can apply Corollary 3.18 to get L such that the diagram (2.7) commutes, which con- (cid:3) isomorphisms GF K cludes the theorem by Proposition 2.6. ∼= GF Lastly, we give two remarks. First, the proof of Theorem 1.1 actually gives a procedure for reconstruction of the semigroup action from K-theoretic data. This is a stronger result than a mere classification. Second, Theorem 1.1 does not mean that the isomorphism is reconstructed from K-theoretic data. Indeed, if we have an automorphism ϕ on AK as a C*-algebra over 2PK (e.g., the action of Gab K induced from its action by multiplication on the sec- ond factor of YK), then we can apply the commutativity of the diagram (5) / /   / /     / / / /   (   w /     / / / 6 g 24 Y. KUBOTA AND T. TAKEISHI in the proof of Theorem 1.1 above for the family {ϕF to see that ϕF ∗ are identity maps. ∗ : K∗(BF K) → K∗(BF K )} Appendix A. Another direct reconstruction of profinite completions from K-theory by Xin Li Let us present an alternative approach to Corollary 3.18. Let Γ → G be a profinite completion, where Γ is a countable free abelian group and G is a profinite completion of Γ which is topologically finitely generated, as Γ/Γn with finite index subgroups Γ1 ⊇ Γ2 ⊇ in Section 3. Let G = lim←−n Γ3 ⊇ . . . of Γ. Let ι be the map K∗(C ∗Γ) → K∗(C(G) ⋊ Γ) induced by the canonical homomorphism C ∗Γ → C(G) ⋊ Γ. In the following, we give a concrete recipe allowing us to reconstruct Γ → G from ι. i=1 Zei, Σs = Ls Write Γ = L∞ i=1 Zei. The basis {ei} gives rise to a canonical isomorphism Zs ∼= Σs, which is the only one we use to identify Σs with Zs. Let Gs = lim←−n Σs/(Σs∩Γn), and ps : Σs → Gs is the canonical map. We have a canonical isomorphism Ls r=0Vr Σs ∼= K∗(C ∗Σs) as in Section 3, so that we can consider the composition ιs : Vs−1 Σs → K∗(C ∗Σs) → ι−→ K∗(C(G) ⋊ Γ). Define K∗(C ∗Γ) Ts := {x ∈ K∗(C(G) ⋊ Γ) : ∃ n ∈ Z, n > 0 with n · x ∈ Im(ιs)} . We have a canonical identification (A.1) Σs ∼=^s−1 Σs, isomorphism Hom(Σs, Z) ∼= Σs via the pairing Σs × Σs → Z, (x, y) 7→ hx, yi, which is given as follows: The pairing Σs ×Vs−1 Σs →Vs Σs ∼= Z, (x, y) 7→ x ∧ y gives an isomorphism Vs−1 Σs ∼= Hom(Σs, Z), and we obtain a second where h·, ·i is the standard Euclidean inner product h(xi), (yi)i =Pi xi · yi. Let is be the composite of ιs with the isomorphism Σs ∼= Vs−1 Σs from (A.1), is : Σs ∼=Vs−1 Σs ֒→ Ts. Theorem A.2. There is an isomorphism ϕs of Ts with a subgroup Qs of Qs uniquely determined by requiring that ϕs ◦ is is the canonical isomor- phism Σs ∼= Zs. Then ϕs induces an isomorphism Ts/is(Σs) ∼= Qs/Zs, ϕs−→ Qs/Zs ֒→ again denoted by ϕs. Let ψs be the composite ψs : Ts/is(Σs) Rs/Zs ∼= cZs ∼= cΣs. Here cΣs stands for Pontrjagin dual, and the isomor- phism Rs/Zs ∼= cZs sends x ∈ Rs/Zs to χ ∈ cZs, where χ(z) = e2πihx,zi. Let cψs : Σs → \Ts/i(Σs) be the dual map of ψs. Then there is a (unique) isomor- phism ωs : \Ts/i(Σs) ∼= Gs such that ωs ◦cψs = ps. The canonical inclusions Σs ֒→ Σ, Gs ֒→ G give rise to isomorphisms Γ ∼= lim−→s Gs ∼= G is the original profinite completion such that Γ ∼= lim−→s Γ → G, where the connecting maps Σs → Σs+1 and Gs → Gs+1 are the ones induced by the canonical inclusion Σs ֒→ Σs+1. lim−→s −→ lim−→s Σs and lim−→s Gs ∼= G Σs ps Proof. We first explain the last claim. By assumption, G is topologically finitely generated. Hence for s big enough, the canonical map Gs → G is an isomorphism, i.e., Σs + Γn = Γ for all n, or equivalently, the canonical RECONSTRUCTING THE BOST–CONNES SEMIGROUP ACTIONS 25 map Σs/(Σs ∩ Γn) → Γ/Γn is surjective. So the sequence Gs → Gs+1 → . . . becomes stationary. Thus, it suffices to show that for s big enough (i.e., for s such that Σs + Γn = Γ for all n), there is ωs such that ωs ◦cψs = ps. Let us fix s with Σs + Γn = Γ for all n. To simplify notation, we set Σ := Σs, ι := ιs, T := Ts and so on, i.e., we drop the index s. We have an isomorphism (A.3) K∗(C(G) ⋊ Γ) ∼= lim−→ {K∗(C(Γ/Γ) ⋊ Γ) → K∗(C(Γ/Γ1) ⋊ Γ) → . . . } ∼= lim−→ {K∗(C ∗Γ) → K∗(C ∗Γ1) → . . . } . Here we identify K∗(C(Γ/Γn) ⋊ Γ) with K∗(C ∗Γn) by embedding C ∗Γn as a full corner into C(Γ/Γn) ⋊ Γ as in the paragraph before Lemma 3.3. For every n, we have by Lemma 3.3 a commutative diagram (A.4) K∗(C ∗Γ) dn· K∗(C ∗Γ) K∗(C ∗Γn) Here the horizontal map is the composite of the first n structure maps in the inductive limit (A.3), the vertical map is induced by the canonical inclusion Γn ֒→ Γ, and the map dn· : K∗(C ∗Γ) → K∗(C ∗Γ) is multiplication with dn = [Γ : Γn]. If we now choose an isomorphism µn : Γ ∼= Γn, then we can expand (A.4) to (A.5) K∗(C ∗Γ) dn· K∗(C ∗Γ) K∗(C ∗Γn) (µn)∗ K∗(C ∗Γ) This gives maps K∗(C ∗Γ) → K∗(C ∗Γ) (the arrow in (A.5) from the lower left to the lower right copy of K∗(C ∗Γ)) such that K∗(C(G) ⋊ Γ) can be identified with the inductive limit of K∗(C ∗(Γ)) → K∗(C ∗Γ) → . . . . Now we choose µn : Γ ∼= Γn such that µn(Σ) = Σ ∩ Γn. This is possible µnΣ−→ as Σ ∩ Γn is a direct summand of Γn. As Σ ∼= Zs, the composition Σ Σ ∩ Γn ֒→ Σ is given by a matrix Mn with integer entries. Restricting the left copy of K∗(C ∗Γ) to Vs−1 Σ in (A.5), we obtain (A.6) dn· Vs−1 Σ Vs−1 Σ Vs−1(Σ ∩ Γn) Vs−1 Mn Vs−1(µnΣ) Vs−1 Σ 26 Y. KUBOTA AND T. TAKEISHI After deletingVs−1(Σ∩Γn) and identifyingVs−1 Σ with Σ as in (A.1), (A.6) becomes (A.7) Σ dn· Σ (M adg n )t Σ n is the adjugate matrix of Mn, uniquely determined by M adg Here M adg · Mn = det(Mn) · I (I being the identity matrix). (M adg n )t is the transpose of M adg n . As dn = [Γ : Γn] = [Σ : Σ ∩ Γn] = det(Mn), the missing map Σ → Σ (from the lower left to the lower right copy of Σ in (A.7)) must be given by M t n. Hence we can complete (A.7) to n (A.8) Σ M t n dn· Σ (M adg n )t Σ Thus T is the inductive limit of the stationary inductive system Σ → Σ → . . . where the composition of the first n structure maps is given by M t n : Σ → Σ (and this determines the inductive limit). It is easy to see that T contains this inductive limit. Conversely, that T is contained in this inductive limit follows from the fact that Vs−1 Σ is a direct summand in K∗(C ∗Γ), so that if n · x lies in Vs−1 Σ ⊆ K∗(C ∗Γ) for some n ∈ Z, n > 0 and x ∈ K∗(C ∗Γ), then x itself must lie inVs−1 Σ. Now, there is only one way to complete the diagram (A.9) Σ . . . Σ Qs if we start with Σ ∼= Zs ֒→ Qs as our first vertical map and want that the diagram commutes (the first row in (A.9) is the inductive system from above giving rise to T ). The completed diagram is given by (A.10) Σ Qs M −t 1 Σ . . . Σ . . . M −t n Here M −t n is the inverse of M t isomorphism ϕ : T ∼= Q =Sn M −t (A.10). ψ becomes T /i(Σ) ∼= Q/Zs =Sn(cid:0)M −t For fixed n, the image of M −t because n (as a matrix over Q). Now the desired n Zs ⊆ Qs arises as the inductive limit in n Zs/Zs(cid:1) ֒→ Rs/Zs ∼= cZs ∼= bΣ. n Zs/Zs ֒→ Rs/Zs ∼= cZs is \Zs/MnZs ⊆ cZs n Zs = {x ∈ Rs : hx, zi ∈ Z ∀ z ∈ MnZs} . M −t Hence ψ is given by T /i(Σ) ∼= lim−→ n \Zs/MnZs = lim−→ n \Σ/(Σ ∩ Γn) ֒→ bΣ. RECONSTRUCTING THE BOST–CONNES SEMIGROUP ACTIONS 27 Therefore, up to composing with an isomorphism called ω, bψ is given by Σ/(Σ ∩ Γn), as claimed. Σ → lim←−n (cid:3) Remark A.11. If Γ itself is finitely generated, the proof becomes even easier, as we can take Σ = Γ. References [1] Rasmus Bentmann, Filtrated K-theory and classification of C*-algebras, Diplom The- sis., 2010. [2] Bruce Blackadar. K-theory for operator algebras, Second, Mathematical Sciences Re- search Institute Publications, vol. 5, Cambridge University Press, Cambridge, ISBN 0-521-63532-2, (1998). [3] J.-B. Bost and A. Connes, Hecke algebras, type III factors and phase transitions with spontaneous symmetry breaking in number theory, Selecta Math. (N.S.) 1 (1995), no. 3, 411–457. [4] Nathanial P. Brown and Narutaka Ozawa. C ∗-algebras and finite-dimensional approx- imations, Graduate Studies in Mathematics, vol. 88, American Mathematical Society, Providence, RI, ISBN 978-0-8218-4381-9; 0-8218-4381-8, (2008). [5] Gunther Cornelissen, Bart de Smit, Xin Li, Matilde Marcolli, and Harry Smit, Recon- structing global fields from Dirichlet L-series, preprint, arXiv:1706.04515[math.NT], 2017. [6] Gunther Cornelissen, Xin Li, Matilde Marcolli, and Harry Smit, Reconstruct- in the abelianized Galois group, preprint, ing global fields from dynamics arXiv:1706.04517[math.NT], 2017. [7] Gunther Cornelissen and Matilde Marcolli, Quantum Statistical Mechanics, L-series and Anabelian Geometry, preprint, arXiv:1009.0736[math.NT], 2010. [8] Philip Green, The local structure of twisted covariance algebras, Acta Math. 140 (1978), no. 3-4, 191–250. [9] Eugene Ha and Fr´ed´eric Paugam, Bost-Connes-Marcolli systems for Shimura vari- eties. I. Definitions and formal analytic properties, IMRP Int. Math. Res. Pap. 5 (2005), 237–286. [10] Gennadi G. Kasparov, The operator K-functor and extensions of C ∗-algebras, Izv. Akad. Nauk SSSR Ser. Mat. 44 (1980), no. 3, 571–636, 719. [11] Eberhard Kirchberg, Das nicht-kommutative Michael-Auswahlprinzip und die Klas- sifikation nicht-einfacher Algebren, C ∗-algebras (Munster, 1999), 2000, pp. 92–141, Springer, Berlin. [12] Marcelo Laca, Nadia S. Larsen, and Sergey Neshveyev, On Bost-Connes types systems for number fields, J. Number Theory 129 (2009), no. 2, 325–338. [13] Marcelo Laca, Sergey Neshveyev, and Mak Trifkovi´c, Bost-Connes systems, Hecke algebras, and induction, J. Noncommut. Geom. 7 (2013), no. 2, 525–546. [14] Ralf Meyer and Ryszard Nest, C ∗-algebras over topological spaces: the bootstrap class, Munster J. Math. 2 (2009), 215–252. [15] , C∗-algebras over topological spaces: filtrated K-theory, Canad. J. Math. 64 (2012), no. 2, 368–408. [16] Jurgen Neukirch. Algebraic number theory, Grundlehren der Mathematischen Wis- senschaften [Fundamental Principles of Mathematical Sciences], vol. 322, Springer- Verlag, Berlin, ISBN 3-540-65399-6, (1999). Translated from the 1992 German original and with a note by Norbert Schappacher, With a foreword by G. Harder. [17] Takuya Takeishi, Irreducible representations of Bost-Connes systems, J. Noncommut. Geom. 10 (2016), no. 3, 889–906. [18] , Primitive ideals and K-theoretic approach to Bost-Connes systems, Adv. Math. 302 (2016), 1069–1079. [19] Bora Yalkinoglu, On arithmetic models and functoriality of Bost-Connes systems. With an appendix by Sergey Neshveyev, Invent. Math. 191 (2013), no. 2, 383–425. 28 Y. KUBOTA AND T. TAKEISHI iTHEMS Research Group, RIKEN, 2-1 Hirosawa, Wako, Saitama 351-0198, Japan E-mail address: [email protected] Kyoto Institute of Technology, 606-8585, Japan E-mail address: [email protected]
1004.0262
1
1004
2010-04-02T00:22:01
On inductive limits of type I C*-algebras with one-dimensional spectrum
[ "math.OA" ]
The class of separable C*-algebras which can be written as inductive limits of continuous-trace C*-algebras with spectrum homeomorphic to a disjoint union of trees and trees with a point removed is classified by the Cuntz semigroup.
math.OA
math
ON INDUCTIVE LIMITS OF TYPE I C*-ALGEBRAS WITH ONE-DIMENSIONAL SPECTRUM ALIN CIUPERCA, GEORGE A. ELLIOTT, AND LUIS SANTIAGO Abstract. The class of separable C*-algebras which can be written as inductive limits of continuous-trace C*-algebras with spectrum homeomorphic to a disjoint union of trees and trees with a point removed is classified by the Cuntz semigroup. 1. Introduction In recent years the Cuntz semigroup has become an important tool of investigation in the theory of C*-algebras, particularly in the work related to the Elliott classification program. As an invariant, it plays a significant role in the theory of both simple and non-simple C*- algebras (see [25], [3], [2], [22]). In this paper, we shall show that the Cuntz semigroup is effective as an invariant for the class of C*-algebras that can be expressed as the inductive limit of a sequence A1 → A2 → A3 → · · · where each building block Ai is a separable continuous-trace C*-algebra with spectrum homeomorphic to a disjoint union of trees and trees with a point removed. The term tree will refer to a contractible one-dimensional finite CW complex, in other words, a contractible space obtained from a finite discrete space V whose elements we shall call vertices, by attaching a finite collection E of 1-cells, which we shall call edges. Without loss of generality we may assume that a tree is a subspace of the plane. sequential inductive limits of building blocks of the form LN Our work can be viewed as a continuation of a number of previous investigations in the classification of C*-algebras. The problem of classifying inductive limits of continuous trace C*-algebras with spectrum homeomorphic to the closed interval [0, 1] was addressed, and results were obtained, in [11] in the simple case, and in [3] in the not necessarily simple case. A particular class of C*-algebras classified in our paper are the C*-algebras obtained as k=1 Mmk (C0(Xk)), where the spaces Xk are trees or trees with a point removed. The case that the spaces Xk are compact received considerable attention in the more classical framework of classification of simple approximately homogeneous (AH) algebras, where usage of the classical Elliott invariant led to remarkable results. The first such result, the classification of approximate interval (AI) algebras, was obtained by one of the present authors in [7], where the situation Xk = [0, 1] was treated. Important generalizations of this result were obtained, in two different directions. On the one hand, the requirement of simplicity for the inductive limit was kept in [14], where Li extended the classification to the case where the spaces Xk are trees. On the other hand, classification results for certain classes of non-simple AI algebras were obtained by Stevens in [23] and by Robert in [18]. In [3], two of the present authors showed that, for C*-algebras of stable rank one, the Cuntz invariant and another C*-algebra invariant -- the Thomsen semigroup -- determine each other, 1 in a natural way. Using Thomsen's classification result [24], it was inferred that the Cuntz semigroup is a complete invariant for the class of separable approximate interval algebras. It became apparent that the Cuntz semigroup is a good candidate to be considered in the classification of not necessarily simple C*-algebras. The main result of the present paper states that the Cuntz semigroup functor classifies the ∗-homomorphisms from a sequential inductive limit of separable continuous-trace C*- algebras with spectrum homeomorphic to a disjoint union of trees and trees with a point removed to a C*-algebra of stable rank one: Theorem 1. Let A be a sequential inductive limit of separable continuous-trace C*-algebras with spectrum homeomorphic to a disjoint union of trees and trees with a point removed. Let B be a C*-algebra of stable rank one. Suppose that there is a Cuntz morphism α : Cu(A) → Cu(B) such that α[sA] ≤ [sB], where sA and sB are a strictly positive element of A and any positive element of B, respectively. It follows that there exists a ∗-homomorphism φ : A → B, unique up to approximate unitary equivalence, such that Cu(φ) = α. It follows from this theorem that the invariant consisting of the Cuntz semigroup to- gether with a distinguished element of it classifies sequential inductive limits of separable continuous-trace C*-algebras with spectrum homeomorphic to a disjoint union of trees and trees with a point removed: Corollary 1. Let A and B be sequential inductive limits of separable continuous-trace C*- algebras with spectrum homeomorphic to a disjoint union of trees and trees with a point removed. Let sA and sB be strictly positive elements of A and B, respectively. Suppose that there is a Cuntz semigroup isomorphism α : Cu(A) → Cu(B) such that α[sA] = [sB]. It follows that there exists a ∗-isomorphism φ : A → B, unique up to approximate unitary equivalence, such that Cu(φ) = α. Note that, as a consequence of Corollary 1 and Li's result [14], the Cuntz semigroup and the Elliott invariant both classify the simple sequential inductive limits of building blocks of k=1 Mmk (C0(Xk)), where the spaces Xk are trees. This is not surprising, since for a large variety of simple C*-algebras, the Cuntz semigroup can be recovered functorially from the Elliott invariant, as proved in [1] and, in the non-unital case, [12]. the form LN 2. Preliminary definitions and results 2.1. The Cuntz semigroup. For a C*-algebra A, let us denote by A+ the set of positive elements of A, and by eA the unitization of A. The following definition of the Cuntz semigroup is different from the original definition given in [5], in that, in addition to positive elements in matrix algebras over A, also positive elements in A ⊗ K are considered (K denotes the algebra of compact operators on a separable Hilbert space). As shown in [4], this form of the definition is very useful. (The relation between the two definitions for a non-stable algebra is not known, except -- see [4] -- in the case of stable rank one.) Definition 1. Let a and b be positive elements of A⊗K. Let us say that a is Cuntz smaller than b, denoted by a 4 b, if there exists a sequence (dn)n∈N in A ⊗ K such that dnbd∗ n → a. The elements a and b are called Cuntz equivalent, written a ∼ b, if a 4 b and b 4 a. It is immediate that 4 is a pre-order on the set of positive elements of A ⊗ K, so that ∼ is an equivalence relation. Given a ∈ (A ⊗ K)+ let us denote by [a] the Cuntz equivalence 2 class of a. The Cuntz semigroup of A, denoted by Cu(A), is defined as the set of equivalence classes of positive elements of A endowed with the order derived from the pre-order relation 4 (so that [a] ≤ [b] if a 4 b), and the addition operation [a] + [b] :=(cid:20)(cid:18) a 0 0 b (cid:19)(cid:21) , where the positive element inside the brackets in the right side of the equation above is identified with its image in A⊗K by any isomorphism of M2(A⊗K) with A⊗K induced by an isomorphism of K and M2(K). If φ : A → B is a ∗-homomorphism from the C*-algebra A to the C*-algebra B, then it induces an ordered semigroup morphism Cu(φ) : Cu(A) → Cu(B) defined on the Cuntz equivalence class of a positive element a ∈ A ⊗ K by Cu(φ)[a] = [(φ ⊗ id)(a)], where id : K → K denotes the identity operator on K. It was shown in [4] that Cu(A) is an object in the the category Cu of ordered abelian semigroups with zero with the following additional properties: (i) every increasing sequence has a supremum; (ii) each element is the supremum of a rapidly increasing sequence, by which is meant a sequence such that each term is compactly contained in the next, where we say that an element x is compactly contained in an element y, and write x ≪ y, if whenever y ≤ sup yn for an increasing sequence (yn)n∈N, then eventually x ≤ yn; (iii) the operation of passing to the supremum of an increasing sequence and the relation ≪ of compact containment are compatible with addition. The maps in the category Cu are ordered semigroup maps preserving the zero element, suprema of increasing sequences, and the relation of compact containment. In Theorem 2 of [4], the authors prove that the Cuntz semigroup constitutes a functor from the category of C*-algebras to the category Cu, and that this functor preserves inductive limits of sequences. We will also make use of another, stronger, equivalence relation among positive elements of a C*-algebra. Definition 2. Let a and b be positive elements of a C*-algebra A. We will say that a is Murray-von Neumann equivalent to b if there exists x ∈ A such that a = x∗x and b = xx∗. The following result (the proof of which is included for the convenience of the reader) will be useful on several occasions. Lemma 1. Let A be a C*-algebra. Let a and b be positive elements of A such that ka−bk ≤ ǫ. Then ka1/2 − b1/2k ≤ √ǫ. Proof. Since ka − bk ≤ ǫ, we have a ≤ b + ǫ1 eA, where 1 eA denotes the unit of the unitization of A. By Proposition 1.3.8 of [17] we have By symmetry, Therefore, a1/2 ≤ (b + ǫ1 eA)1/2 ≤ b1/2 + 1 eA√ǫ. b1/2 ≤ a1/2 + 1 eA√ǫ. ka1/2 − b1/2k ≤ √ǫ. 3 (cid:3) The following proposition is a restatement in terms of positive elements of Theorem 3 of [4]. For the convenience of the reader we include a proof of this statement. Lemma 2. Let A be C*-algebra, and let B be a hereditary subalgebra of A of stable rank one. Let δ > 0. If x, y ∈ A are such that xx∗, yy∗ ∈ B, x∗x ∈ y∗Ay, and (2.1) then there exists a unitary U in the unitization of B such that kx∗x − y∗yk < δ, kx − Uyk < √δ. Proof. Let x = V x and y = Wy denote the polar decompositions of x and y in the bidual of A (we use the notation x for the element (x∗x) 2 ). Set V xW ∗ = z. It is clear that z ∈ xA. Also, z ∈ Ay∗ since by assumption x∗x ∈ y∗Ay. Therefore, z ∈ B. Let ǫ > 0 (to be specified later). Since B has stable rank one there exists an invertible element z′ in the unitization of B such that 1 kz − z′k ≤ ǫ 2 , kz − z′k ≤ ǫ 2 . Denote by U ∈ eB the unitary in the polar decomposition of the invertible element z′. It follows that kz − Uzk ≤ kz − z′k + kUz′ − Uzk ≤ + = ǫ. ǫ 2 ǫ 2 Hence, (2.2) We have the following estimation: kz − Uzk < ǫ. kx − Uyk = kV x − UWyk ≤ kV x − UWxk + kUWx − UWyk ≤ kV xW ∗ − UWxW ∗k + kx − yk = kz − Uzk + kx − yk. From Equation (2.1) we have kx − yk < √δ. Taking ǫ = √δ − kx − yk in Equation (2.2) and using the preceding estimation we conclude that kx − Uyk ≤ kz − Uzk + kx − yk = √δ. (cid:3) Proposition 1. Let A be C*-algebra. If a, b ∈ A+ are such that a 4 b, and the hereditary subalgebra bAb has stable rank one, then a is Murray-von Neumann equivalent to an element of bAb. Proof. Since a 4 b, it follows by Lemma 2.2 of [13] that there are elements xn ∈ A, n = 1, 2,· · · , such that (cid:18)a − 1 22n(cid:19)+ = x∗ nxn, xnx∗ n ∈ bAb. (Given a positive element a and a real number t, we denote by (a − t)+ the evaluation of a in the function ct(x) = max(x − t, 0), for x ≥ 0.) 4 For each n ≥ 1, let us apply Lemma 2 to the elements xn and xn+1. Then there exists a n. We unitary Un such that kxn − Unxn+1k < 1 have n =(cid:18)a − n)∗x′ (x′ 22n(cid:19)+ , and kx′ it has a limit. Let us denote by x the limit of (x′ 2n . It follows that the sequence (x′ n+1k < 1 n − x′ 2n . For each n ≥ 1 set U1U2 · · · Un−1xn = x′ 1 n(x′ x′ n)∗ ∈ bAb, n)n∈N is a Cauchy sequence and hence x∗x = lim n (x′ n)∗x′ n = lim n)n∈N. Then n (cid:18)a − 1 22n(cid:19)+ = a. (cid:3) Also, xx∗ = limn x′ n(x′ n)∗ ∈ bAb. 2.2. Generators and relations. Throughout this paper we will only consider trees that are realized as subsets of C, with edges line segments of length 1. Note that any tree is homeomorphic to one in this class and so we do not lose any generality with this assumption. Let (X, v) be a rooted tree, that is to say, a tree X with a specified vertex, or root, v and with edges oriented with the natural orientation, away from the root. Let us denote by E(X, v) the set of edges of X with their respective orientations. Given two edges e1 of e2 of (X, v), we will say that e2 is next to e1 if the terminal vertex of e1 is the same as the initial vertex of e2. We will say that e1 is beside e2 if the initial vertex of e1 is the same as the initial vertex of e2. The orientation of the edges of X induces an order in the set of edges E(X, v). Given two edges e and e′, let us write e ≤ e′ if there is a sequence of edges e1, e2,· · · , en with e1 = e and en = e′ such that ei+1 is next to ei for all i. Let us consider the C*-algebra C0(X \ v) of continuous functions on X ⊆ C that vanish at the point v. To each edge e of (X, v) we associate the positive element ge of C0(X \ v) given by (2.3) ge(x) = x, x ∈ e = [0, 1], 1, x ∈ Xe, 0, x ∈ X \ (Xe ∪ e), where Xe denotes the subtree of X consisting of the edges e′ of (X, v) that are less than the edge e. In the equation (2.3) we are identifying the edge e with the interval [0, 1] in such a way that the initial and terminal points of e are identified with the points 0 and 1, respectively. Let us denote by G(X, v) the set of elements ge with e ∈ E(X, v). In the proposition below it is shown that G(X, v) generates the C*-algebra C0(X \ v). Given a rooted tree (X, v), let us denote by C∗hX, vi the universal C*-algebra on generators (he)e∈E(X,v) -- one generator he for each edge e -- subject to the relations (2.4) khek ≤ 1, he ≥ 0, he1he2 = 0, he1he2 = he2, if e1 is beside e2, if e2 is next to e1. Sometimes, in order to avoid confusion, we will write h(X,v) instead of he for the generators of the C*-algebra C∗hX, vi. The same notation will be used when referring to the elements ge of C0(X \ v). e 5 Proposition 2. Let (X, v) be a rooted tree. Then the C*-algebra C∗hX, vi is isomorphic to the C*-algebra C0(X \ v), by means of (2.5) he 7→ ge ∈ C0(X \ v). Proof. The set of elements (ge)e∈E(X,v), with ge defined in Equation (2.3), is a represention of the relations (2.4) in the C*-algebra C0(X \ v). It follows by Lemma 3.2.2 of [15] that there exists an isomorphism from C∗hX, vi to C0(X \ v) such that (2.5) holds, if and only if: (1) the elements ge, e ∈ E(X, v), generate the C*-algebra C0(X \ v); and (2) for any ∗-homomorphism φ : C∗hX, vi → C there exists a ∗-homomorphism ψ : C0(X \ v) → C such that φ(he) = ψ(ge). (Here we are also using that the C*-algebra C∗hX, vi is commutative, since by the relations (2.4) the elements he, e ∈ E(X, v), commute with each other.) Let e′ be an edge of (X, v). The sub-C*-algebra of C0(X \ v) generated by the element ge′ consists of the continuous functions on X that are constant on the setSe>e′ e, and zero on the set X \ (Se≥e′ e). These functions, when e′ varies through the set E(X, v), clearly generate the C*-algebra C0(X \ v). This shows that Condition (1) is satisfied. Let φ : C∗hX, vi → C be a ∗-homomorphism. Then the numbers φ(he), e ∈ E(X, v), satisfy the relations (2.4). It follows from these relations that the set of edges of (X, v) such that φ(he) 6= 0 consists of a sequence of edges e1, e2,· · · , ek, such that the initial vertex of e1 is v, and ei+1 is next to ei for all i. Moreover, we have φ(hek) > 0, and φ(hei) = 1 for 1 ≤ i < k. Let x be the point in ek such that gek(x) = φ(hek). Then the ∗-homomorphism ψ : C0(X \ v) → C such that ψ(f ) = f (x) satisfies ψ(ge) = φ(he) for all e ∈ E(X, v). This shows that Condition (2) is satisfied. (cid:3) 2.3. Continuous-trace C*-algebras. Let A be a C*-algebra and let A denote the spec- trum of A. A continuous-trace C*-algebra is a C*-algebra A which is generated as a closed two-sided ideal by the elements x ∈ A+ for which the function π → Tr(π(x)) is finite and continuous on A. In this paper we make use of the following fact about continuous-trace C*-algebras (see [6]): Proposition 3. Let A be a continuous-trace C*-algebra such that H3( A, Z) = 0. Then A is stably isomorphic to C0( A). In particular, the proposition above can be applied to the case that the spectrum of A is a finite disjoint union of trees and trees with a point removed. 3. The pseudometrics d(X,v) U and d(X,v) W Given a C*-algebra A and a rooted tree (X, v), let us denote by Hom(C0(X \ v), A) the set of ∗-homomorphisms from C0(X \ v) to A, and by Mor(Cu(C0(X \ v)), Cu(A)) the set of Cuntz semigroup morphisms from the Cuntz semigroup of C0(X \ v) to the Cuntz semigroup of A. on these sets, and in W Theorem 3 we will prove that these pseudometrics are equivalent. In this section we will define pseudometrics d(X,v) and d(X,v) U Given φ, ψ ∈ Hom(C0(X \ v), A) we define d(X,v) U (φ, ψ) by the formula d(X,v) U (φ, ψ) := inf U ∈ eA g∈G(X,v) kφ(g) − U ∗ψ(g)Uk, sup 6 In order to define the pseudometric d(X,v) where eA denotes the unitization of A, and G(X, v) denotes the set of generators of the C*-algebra C0(X \ v) corresponding to the edges of X as in (2.3). let us consider first the special case (X, v) = ([0, 1], 0). The pseudometric d([0,1],0) W -- or, for short, dW -- was defined in [3] by Ciuperca and Elliott. Given Cuntz semigroup morphisms α, β : Cu(C0(0, 1]) → Cu(A) the distance between α and β is defined by W (3.1) dW (α, β) := inf(cid:26)r ∈ R+(cid:12)(cid:12)(cid:12)(cid:12) α[(id − t − r)+] ≤ β[(id − t)+], β[(id − t − r)+] ≤ α[(id − t)+], for all t ∈ R+(cid:27) , where id denotes the identity function on (0, 1], and (id − t)+ denotes the positive part of the function id − t, for t ∈ R. Now let us consider a general rooted connected tree (X, v). For each generator g ∈ G(X, v) let χg : C0(0, 1] → C0(X \ v) denote the ∗-homomorphism such that χg(id) = g. Let us define d(X,v) W on the set Mor(Cu(C0(X \ v)), Cu(A)) by (3.2) d(X,v) W (α, β) := sup dW (α ◦ Cu(χg), β ◦ Cu(χg)). g∈G(X,v) Since dW is a pseudometric (see [3]), d(X,v) we show that when the C*-algebra A has stable rank one d(X,v) W is also a pseudometric. In the following proposition is actually a metric. W Proposition 4. Let A be a C*-algebra and let (X, v) be a rooted tree. The following state- ments hold: (i) If the C*-algebra A has stable rank one, then d(X,v) (ii) The space Hom(C0(X \ v), A) is complete with respect to the pseudometric d(X,v) is a metric. W U . Proof. (i) Let (X, v) be a rooted tree, and let α, β : Cu(C0(X \ v)) → Cu(A) be Cuntz semigroup morphisms such that d(X,v) By Theorem 1 of [19], Cu(C0(X \ v)) is naturally isomorphic (by a Cuntz semigroup isomorphism!) to the semigroup of lower semicontinuous functions from X \ v to N ∪ {∞} with the pointwise addition and order. The isomorphism is given by the rank map: W (α, β) = 0. We need to show that α = β. i=0 Cu(C0(X \ v)) ∋ [a] 7→ {(X \ v) ∋ x 7→ rank(a(x))}. (3.3) Let us denote the set of lower semicontinuous functions from X \ v to N ∪ {∞} by Lsc(X \ 1Ui, where the sets Ui are defined as Ui = {x f (x) > i}, for all i. Moreover, the sets Ui, i = 0, 1,· · · , are open since f is lower semicontinuous. For each open set Ui, i = 1, 2,· · · , there exists a sequence (Ui,j)∞ j=1 of v, N∪{∞}). Let f ∈ Lsc(X \ v, N∪{∞}). Then f =P∞ pairwise disjoint open connected subsets of X \ v such that Ui =S∞ it follows that f = supnP1≤i,j≤n Let us show that α(1U ) = β(1U ) for any open connected subset U of X \ v. It will follow from this that α = β, since the Cuntz semigroup morphisms α and β are additive and preserve suprema of increasing sequences, and any function in Lsc(X \ v, N ∪ {∞}) is the supremum of an increasing sequence of finite sums of characteristic functions of open subsets j=1 Ui,j. Since ∞Xi=0 ∞Xj=0 ∞Xi=0 1Ui = 1Ui,j . 1Ui,j , f = 7 of X \ v, as shown above. By assumption, d(X,v) pseudometric d(X,v) (see Equation (3.2)) W W (α, β) = 0. Hence, from the definition of the dW (α ◦ Cu(χg), β ◦ Cu(χg)) = 0, It follows that α(1X ǫ for all g ∈ G(X, v), where χg : C0(0, 1] → C0(X \ v) is the ∗-homomorphism defined by χ(id) = g. Since the C*-algebra A has stable rank one dW is a metric, as it was shown in [3] (or in Proposition 2 of [20]). Therefore, α ◦ Cu(χg) = β ◦ Cu(χg) for all g ∈ G(X, v). In particular, α[(ge − ǫ)+] = β[(ge − ǫ)+] for each e ∈ E(X, v), and each ǫ > 0. Let us denote by X ǫ e the subset of X \ v consisting of the edges of (X, v) that are less than e, and the points of e that are at a distance strictly larger than ǫ from the initial vertex of e. Then rank((ge − ǫ)+) = 1X ǫ e) for each e ∈ E(X, v), and each e. ǫ > 0. Note that each set X ǫ e and y < x (a point y is less than a point x if the non-overlapping path from y to the root v contains x), then y ∈ X ǫ e. Also note that every hereditary open subset of X \ v is the union of a finite number of pairwise disjoint sets that have the form X ǫ e, for some e ∈ E(X, v) and some ǫ > 0. Therefore, α(1U ) = β(1U ) for every hereditary open subset U of X \ v. Let U be a connected open subset of X \ v. We can choose a sequence of open subsets (Ui)∞ i=1 Ui, and such that for each i = 1, 2,· · · , there is a compact set Ki such that Ui ⊂ Ki ⊂ Ui+1. It follows that 1U = supi 1Ui, and 1Ui ≪ 1Ui+1 for all i ≥ 1. In addition, for each i = 1, 2,· · · , we can chose a hereditary open subset Vi ⊆ X \ v such that U ∪ Vi is hereditary, Vi ∪ Ki is compact, and Vi ∩ Ui = ∅. Hence, Vi is such that e is hereditary in the sense that, if x ∈ X ǫ i=1 such that U =S∞ e) = β(1X ǫ We have α(1Vi) = β(1Vi), α(1U ∪Vi) = β(1U ∪Vi), 1Ui∪Vi ≪ 1Ui+1∪Vi, 1Ui + 1Vi = 1Ui∪Vi. α(1Ui) + α(1Vi) = α(1Ui∪Vi) ≪ α(1Ui+1∪Vi) ≤ α(1U ∪Vi) = β(1U ∪Vi) ≤ β(1U ) + β(1Vi) = β(1U ) + α(1Vi). More briefly, Hence by Theorem 4.3 of [21], α(1Ui) + α(1Vi) ≪ β(1U ) + α(1Vi). α(1Ui) ≤ β(1U ). 1Ui we conclude that α(1U ) ≤ Taking the supremum over i ≥ 1 and using that 1U = supi β(1U ). By symmetry, β(1U ) ≤ α(1U ). Therefore, we have α(1U ) = β(1U ) for every open connected subset of X \ v. This concludes the proof of the statement (i). . Then there exists a subsequence (φni)i∈N of (φn)n∈N such that (ii) Let (φn)n∈N be a Cauchy sequence with respect to the pseudometric d(X,v) U (3.4) d(X,v) U (φni, φni+1) < 1 2i+1 , 8 d(X,v) U (φni, φn) < 1 2i+1 , for each i ∈ N, and each n > ni. It follows from the definition of the metric d(X,v) the first inequality above, that there exist unitaries Ui ∈ A, i = 1, 2,· · · , such that U , and from kφni(g) − U ∗ i φni+1(g)Uik < 1 2i+1 , for all g ∈ G(X, v). For each i ≥ 1, set Ad(Ui−1Ui−2 · · · U1) ◦ φni = φ′ ni. Then for each g ∈ G(X, v), the sequence (φ′ ni(g))i∈N is Cauchy in the norm topology. Hence, it converges. For each g ∈ G(X, v), let us denote by g the element limi φni(g). Then the set {g g ∈ G(X, v)} is a representation of the relations (2.4) in the C*-algebra A. Therefore, by Proposition 2, there exists a ∗-homomorphism φ : C0(X \ v) → A such that φ(g) = g, for all g ∈ G(X, v). Using the triangle inequality and the second inequality in Equation (3.4) we have d(X,v) U (φn, φ) ≤ d(X,v) (φn, φni) + d(X,v) U U 1 2i+1 + 0 + d(X,v) U < ni) + d(X,v) (φni, φ′ 1 2i+1 + d(X,v) U U (φ′ ni, φ) (φ′ ni, φ), (φ′ ni, φ) = for each i ∈ N, and for each n > ni. Therefore, the pseudometric d(X,v) is complete. U It follows that d(X,v) U (φn, φ) → 0 when n → ∞. (cid:3) Lemma 3. Let A be a unital C*-algebra of stable rank one, and let X be a tree. Consider two Cuntz semigroup morphisms α, β : Cu(C(X)) → Cu(A) such that α([1X]) = β([1X]), where 1X denotes the unit of C(X). Then dW (α ◦ Cu(χg), β ◦ Cu(χg)) = dW (α ◦ Cu(χ1−g), β ◦ Cu(χ1−g)), (3.5) for any element 0 ≤ g ≤ 1 of C(X), where χg as above takes id ∈ C0(0, 1] to g. W W Proof. Let 0 ≤ g ≤ 1 in C(X) be given. Let us denote by eχg the ∗-homomorphism from C[0, 1] to C(X) such that eχg(id) = g and eχg(1[0,1]) = 1X. Note that eχgC0(0,1] = χg and that, in the obvious sense, eχgC0[0,1) = χ1−g. Using the definition of the pseudometrics dW = d([0,1],0) and d([0,1],1) we have W dW (α ◦ Cu(χg), β ◦ Cu(χg)) = dW (α ◦ Cu(eχg), β ◦ Cu(eχg)), dW (α ◦ Cu(χ1−g), β ◦ Cu(χ1−g)) = d([0,1],1) (α ◦ Cu(eχg), β ◦ Cu(eχg)). Therefore, in order to show that equality (3.5) holds it is enough to show that dW (α, β) = d([0,1],1) (α, β) for all Cuntz semigroup morphisms α, β : Cu(C[0, 1]) → Cu(A) such that W α([1[0,1]]) = β([1[0,1]]). By Theorem 1 of [19] the Cuntz semigroup of C[0, 1] is isomorphic to the set of lower semicontinuous functions from [0, 1] to N ∪ {∞} with pointwise addition and order. Under this identification the element [(id − t)+] ∈ Cu(C[0, 1]) corresponds to the characteristic function of the set (t, 1]. We will denote this function by 1(t,1]. Let r > 0. It follows from the definition of dW and d([0,1],1) that W (3.6) (3.7) dW (α, β) < r ⇔ d([0,1],1) W (α, β) < r ⇔ α(1(t+r,1]) ≤ β(1(t,1]), β(1(t+r,1]) ≤ α(1(t,1]), α(1[0,t)) ≤ β(1[0,t+r)), β(1[0,t)) ≤ α(1[0,t+r)), 9 for all t ∈ R+, for all t ∈ R+. In order to show that dW (α, β) = d([0,1],1) implies d([0,1],1) (α, β) < r for all r ∈ R+, and vice versa. W Let us suppose that dW (α, β) < r. Then for all ǫ > 0 we have (somewhat as in [3]) W (α, β) it is enough to prove that dW (α, β) < r α(1[0,t−ǫ)) + α(1(t−ǫ,1]) ≤ α(1[0,1]) ≪ α(1[0,1]) = β(1[0,1]) ≤ β(1[0,t+r)) + β(1(t+r−ǫ,1]) ≤ β(1[0,t+r)) + α(1(t−ǫ,1]) (we are using relation (3.6) in order to obtain the last inequality above). By Theorem 1 of [9] (or by Theorem 4.3 of [21]) we can cancel α(1(t−ǫ,1]) from both sides of the preceding inequality. Thus, we obtain α(1[0,t−ǫ)) ≤ β(1[0,t+r)). Since ǫ is arbitrary and α preserves suprema of increasing sequences, α(1[0,t)) = α(sup ǫ 1[0,t−ǫ)) = sup ǫ α(1[0,t−ǫ)) ≤ β(1[0,t+r)). So, α(1[0,t)) ≤ β(1[0,t+r)). Interchanging α and β, as we may, we have β(1[0,t)) ≤ α(1[0,t+r)). Hence, d([0,1],1) W (α, β) < r by relation (3.7). This shows that (α, β) ≤ dW (α, β). d([0,1],1) W The opposite inequality follows by symmetry. (cid:3) Theorem 2. Let A be a C*-algebra of stable rank one and let (X, v) be a rooted tree. Let such that φ, ψ : C0(X \ v) → A be ∗-homomorphisms. Then for all ǫ > 0 there exists a unitary U in eA kφ(g) − U ∗ψ(g)Uk < (2N + 2)dW (Cu(φ ◦ χg), Cu(ψ ◦ χg)) + ǫ (3.8) for all g ∈ G(X, v), where N denotes the number of edges of X. Proof. We will use mathematical induction on the number of edges N of the tree X. When N = 1 we can identify the rooted tree (X, v) with ([0, 1], 0). In this case the inequality (3.8) was shown in the proof of Theorem 4.1 of [3] (for an explicit proof see Theorem 1 of [20]). Now let N > 1 and let us assume that the theorem holds for all rooted trees with number of edges strictly less than N, and in each case for an arbitrary C*-algebra of stable rank one in place of A. Let us show that the theorem holds for all rooted trees with N edges. Let (X, v) be a rooted tree with N edges and let φ, ψ : C0(X \ v) → A be ∗-homomorphisms. Let us denote by eφ,eψ : C(X) → eA the unitizations of φ and ψ. Then φ ◦ χg = eφ ◦ χg and ψ ◦ χg = eψ ◦ χg. It follows from these identities that the inequality (3.8) holds for the ∗- homomorphisms φ and ψ if and only if it holds for the ∗-homomorphismseφ and eψ. Therefore, we may assume that the ∗-homomorphisms φ and ψ have domain C(X), codomain eA, and Let us show that there is no loss of generality to assume that the number of edges of X having v as a vertex is strictly larger than one. Assume that v is a vertex of only one edge of X, say e. Denote by v′ the other vertex of the edge e. Since the number of edges of the tree X is strictly larger than one, v′ is a vertex of more than one edge of X. From the definition of the set of generators G(X, v′) we see that it is obtained from the set G(X, v) by replacing the generator ge associated to the edge e by the function 1 − ge ∈ C0(X \ v′). That is to say, (3.9) that they are unital. G(X, v′) = (G(X, v) \ {ge}) ∪ {1 − ge}. 10 Now if we apply Lemma 3 to the C*-algebra eA and to the Cuntz semigroup morphisms Cu(φ) and Cu(ψ) we see that the inequality (3.8) remains unchanged if a generator g is replaced by 1−g. This observation together with the equation (3.9) implies that the ∗-homomorphisms φ and ψ satisfy the inequality (3.8) for all g in G(X, v) if and only if they satisfy this condition with v′ in place of v. Thus, we may assume that v is a vertex of more than one edge of X. Let us denote by e1, e2,· · · , ek the edges of (X, v) with v as a vertex, and by v1, v2,· · · , vk their second vertices, respectively. Denote by ge1, ge2,· · · , gek ∈ G(X, v) the generators of C0(X \ v) associated to e1, e2,· · · , ek. Suppose that the number of indices i such that dW (Cu(φ ◦ χgei ), Cu(ψ ◦ χgei )) = 1, is strictly larger than one (note that by the definition of the metric dW we have dW (α, β) ≤ 1 for all Cuntz semigroup morphisms α and β). Changing the numbering of the edges e1, e2,· · · , ek if necessary, we may assume that (3.10) dW (Cu(φ ◦ χgei dW (Cu(φ ◦ χgei ), Cu(ψ ◦ χgei ), Cu(ψ ◦ χgei )) = 1, for 1 ≤ i ≤ k′, )) < 1, for k′ < i ≤ k. Denote by Y the subgraph of X consisting of the edges e of (X, v) such that either e < ei for some 1 ≤ i ≤ k, or e = ei for some k′ < i ≤ k. Let us define an equivalence relation ∼ on Y by taking vi ∼ vj for every 1 ≤ i, j ≤ k′, vi ∼ v for every 1 ≤ i ≤ k′, and x ∼ x for every x ∈ Y . Then the set (Y ′, [v1]), where Y ′ = Y /∼ and [v1] denotes the equivalence class of v1, has the structure of a rooted tree. The edges and the vertices of (Y ′, [v1]) are defined to be the images by the quotient map of the edges and the vertices of Y . In addition, the C*-algebra C0(Y ′ \ [v1]) is isomorphic to the subalgebra of C0(X \ v) generated by the the ∗-homomorphisms induced by the restrictions of the ∗-homomorphisms φ and ψ to the subalgebra of C0(X \ v) generated by the elements ge ∈ G(X, v), with e an edge of Y . By the inductive hypothesis and using that the number of edges of (Y ′, [v1]) is strictly less than elements ge ∈ G(X, v) such that e is an edge of Y . Let φ′, ψ′ : C0(Y ′ \ [v1]) → eA denote N we have that given ǫ > 0 there exists a unitary U ∈ eA such that kφ′(g) − U ∗ψ′(g)Uk < (2N + 2)dW (Cu(φ′ ◦ χg), Cu(ψ′ ◦ χg)) + ǫ, for all g ∈ G(Y ′, [v1]). It follows that (3.11) kφ(ge) − U ∗ψ(ge)Uk < (2N + 2)dW (Cu(φ ◦ χge), Cu(ψ ◦ χge)) + ǫ, for every edge e of Y . By Equation (3.10) the inequality above also holds for the edges e1, e2,· · · , ek′ since the left side of the inequality (3.11) is less than or equal to two, and N > 1. Using that the edges of (X, v) consist of the edges of Y and the edges e1, e2,· · · , ek′ we conclude that the inequality above holds for all e ∈ E(X, v). It follows that the statement of the theorem holds for the rooted tree (X, v). Therefore, we may assume that dW (Cu(φ ◦ χgei ), Cu(ψ ◦ χgei )) < ri < 1, for i = 1, 2,· · · , k and some positive numbers ri. By the definition of the pseudometric dW , the preceding inequality implies that φ((gei − ri)+) = (φ(gei) − ri)+ 4 ψ(gei), 11 (3.12) i xi, xix∗ j = x∗ φ((gei − ri)+) = x∗ i xj = 0 for i 6= i=1 xi = x, and for each i = 1, 2,· · · , k. Applying (ii) of Proposition 1 to the elements φ((gei − ri)+) and i ∈ ψ(gei)eAψ(gei). Note that the elements xi satisfy the orthogonality relations xix∗ ψ(gei), we obtain elements xi ∈ eA, i = 1, 2,· · · , k, such that j (this holds because the elements gei are pairwise orthogonal). Set Pk consider the polar decomposition x = V x of x in the bidual of eA. From the orthogonality homomorphism with image contained in the hereditary subalgebra ψ(gei)eAψ(gei) of eA, for relations satisfied by the elements xi it follows that xi = V xi. This last identity implies that the restriction of the map Ad(V ∗) ◦ φ to the C*-algebra (gei − ri)+C0(X \ v) is a ∗- each i = 1, 2,· · · , k. For each i ∈ {1, 2,· · · , k} let us denote by Xi the closure (in C) of the spectrum of the algebra (gei − ri)+C0(X \ v), and by wi ∈ Xi the point on the edge ei that is at distance ri from v. The set Xi ⊂ X can be given the structure of a tree by defining its vertices to be the vertices of X that belong to Xi, together with the point wi. The edges of Xi will simply be the edges of X that are subsets of Xi, together with the part of the edge ei that belongs to Xi. (We will refrain from insisting here that an edge has length one.) It follows from the fact that k is at least two that the number of edges of Xi is less than or equal to N − 1. Note that the C*-algebra (gei − ri)+C0(X \ v) is just C0(Xi \ wi). The restrictions of the maps Ad(V ∗) ◦ φ and ψ to the C*-algebra C0(Xi \ wi) are ∗- fore, by the inductive hypothesis, for each fixed ǫ > 0 there exists a unitary Ui, in the homomorphisms with images contained in the hereditary subalgebra ψ(gei)eAψ(gei). There- C*-algebra generated by the hereditary subalgebra ψ(gei)eAψ(gei) and the unit of eA, such kV φ(g)V ∗ − U ∗ i ψ(g)Uik < 2NdW (Cu(Ad(V ∗) ◦ φ ◦ χg), Cu(ψ ◦ χg)) + ǫ that = 2NdW (Cu(φ ◦ χg), Cu(ψ ◦ χg)) + ǫ, i=1 G(Xi, wi). for all g ∈ G(Xi, wi). Replacing Ui by a scalar multiple if necessary, we may assume that kV φ(g)V ∗ − U ∗ψ(g)Uk < 2NdW (Cu(φ ◦ χg), Cu(ψ ◦ χg)) + ǫ, i=1 xi, where the elements xi are given in the equation (3.12). It follows by (i) of Proposition 1 that i=1(Ui − 1) = U. Then U is a unitary element of eA since the elements Ui − 1 are pairwise orthogonal. Furthermore, for each i = 1, 2,· · · , k, we have i ψ(g)Ui = U ∗ψ(g)U for all g ∈ G(Xi, wi). Thus, U ∗ (3.13) Ui − 1 ∈ ψ(gei)eAψ(gei). Set 1 +Pk for all g ∈Sk Recall that V is the partial isometry in the polar decomposition of the element x =Pk for any δ > 0 there exists a unitary W ∈ eA such that Hence for any δ > 0 and any finite subset F of the hereditary subalgebra xeAx there exists a unitary W ∈ eA such that kV yV ∗ − W yW ∗k < δ, kV x − Wxk < δ. 12 for all y ∈ F . In particular, if we take F =Sk i=1 G(Xi, wi), and δ small enough, we find that the inequality (3.13) still holds if the partial isometry V is replaced by a suitable unitary W . Set UW = U ′. Then (3.14) kφ(g) − (U ′)∗ψ(g)U ′k < 2NdW (Cu(φ ◦ χg), Cu(ψ ◦ χg)) + ǫ, for all g ∈ Sk i=1 G(Xi, wi). Let us show that the unitary U ′ satisfies the conditions of the theorem. According to the tree structure given to the set Xi for each i = 1, 2,· · · , k, we have that all the elements of G(Xi, wi) belong to G(X, v) except for fi = 1 (gei − ri)+ (this 1−ri function, restricted to Xi, is the generator of C(Xi, wi) that corresponds to the (short) edge Xi ∩ ei; in other words, it is the generator gXi∩ei). In fact, (3.15) G(X, v) = (G(Xi, vi) ∪ {gei})\ {fi}. k[i=1 So, in order to show that the inequality (3.8) holds for the ∗-homomorphisms φ and ψ, and the unitary U ′, it is enough to check that it holds for the elements gei. (For the rest of the elements of G(X, v) it holds by (3.14) and (3.15).) By the definition of the metric dW , dW (Cu(φ ◦ χfi), Cu(ψ ◦ χfi)) = −ri)+), Cu(ψ ◦ χ(gei ), Cu(ψ ◦ χgei ≤ )) < ri (by hypothesis), it follows that dW (Cu(φ ◦ χ(gei dW (Cu(φ ◦ χgei )). 1 1 − ri 1 − ri Since dW (Cu(φ ◦ χgei ), Cu(ψ ◦ χgei −ri)+)) By the inequality (3.14) with g = fi, it follows that dW (Cu(φ ◦ χfi), Cu(ψ ◦ χfi)) < ri 1 − ri . k[i=1 1 kφ(fi) − (U ′)∗ψ(fi)U ′k < 2Nri 1 − ri + ǫ, for i = 1, 2,· · · , k. Hence, (3.16) for i = 1, 2,· · · , k. By the triangle inequality, kφ((gei − ri)+) − (U ′)∗ψ((gei − ri)+)U ′k < 2Nri + (1 − ri)ǫ ≤ 2Nri + ǫ, kφ(gei) − (U ′)∗ψ(gei)U ′k ≤ kφ(gei) − φ((gei − ri)+)k+ + kφ((gei − ri)+) − (U ′)∗ψ((gei − ri)+)U ′k + kψ((gei − ri)+) − ψ(gei)k. By the inequality (3.16) and the identity k(gei − ri)+ − geik = ri it follows that kφ(gei) − (U ′)∗ψ(gei)U ′k < ri + (2Nri + ǫ) + ri = (2N + 2)ri + ǫ, for i = 1, 2,· · · , k. Since this inequality holds for all numbers ri such that dW (Cu(φ ◦ χgei ), Cu(ψ ◦ χgei )) < ri < 1 we conclude that kφ(gei) − (U ′)∗ψ(gei)U ′k < (2N + 2)dW (Cu(φ ◦ χgei 13 ), Cu(ψ ◦ χgei )) + ǫ, for i = 1, 2,· · · , k. This shows that (3.8) holds for the elements gei, i = 1, 2,· · · , k. This concludes the proof of the theorem. (cid:3) Theorem 3. Let A be a C*-algebra of stable rank one and let (X, v) be a rooted tree. Let φ, ψ : C0(X \ v) → A be ∗-homomorphisms. Then (3.17) W (Cu(φ), Cu(ψ)), d(X,v) W (Cu(φ), Cu(ψ)) ≤ d(X,v) (φ, ψ) ≤ (2N + 2)d(X,v) U where N denotes the number of edges of X. Proof. Let us start by proving the first inequality of (3.17). By Corollary 9.1 of [3] (see also Lemma 1 of [20]) we have, for each g ∈ G(X, v), dW (Cu(φ ◦ χg), Cu(ψ ◦ χg)) ≤ d([0,1],0) U = inf (φ ◦ χg, ψ ◦ χg) u∈Akφ(g) − u∗ψ(g)uk u∈A ≤ inf = d(X,v) g∈G(X,v) kφ(g) − u∗ψ(g)uk sup (φ, ψ). U It follows that d(X,v) W (Cu(φ), Cu(ψ)) = sup g∈G(X,v) dW (Cu(φ ◦ χg), Cu(ψ ◦ χg)) ≤ d(X,v) U (φ, ψ). Now let us prove the second inequality of (3.17). Applying Theorem 2 to the ∗-homomorphisms kφ(g) − U ∗ψ(g)Uk < (2N + 2)dW (Cu(φ ◦ χg), Cu(ψ ◦ χg)) + ǫ, φ and ψ we obtain a unitary U ∈ eA such that to g ∈ G(X, v) and then taking the infimum with respect to U ∈ eA we obtain for all g ∈ G(X, v). Taking the suprema on both sides of the inequality above with respect (φ, ψ) < (2N + 2)d(X,v) W (Cu(φ), Cu(ψ)) + ǫ. d(X,v) U Since ǫ is arbitrary, the desired inequality follows. (cid:3) 4. Approximate lifting This section is devoted to the proof of Theorem 4 below. This theorem states that every Cuntz semigroup morphism between the Cuntz semigroups of the C*-algebra of continuous functions over a rooted connected tree and a general C*-algebra can be lifted approximately to a ∗-homomorphism between these C*-algebras. Before we proceed to prove Theorem 4 we need some preliminary results. The following result is Lemma 4 of [20]. (Lemma 4 is used in the proof of Theorem 4 of [20] which is the case of Theorem 4 below that the tree is an interval. The proof in the present more general setting is somewhat more subtle than the proof on the case of the interval.) Lemma 4. Let A be a C*-algebra and let {xk}n k=0 be elements of Cu(A) such that xk+1 ≪ xk for k = 0, 1,· · · , n − 1. Then there exists a ∈ (A ⊗ K)+ with kak ≤ 1 such that [a] = x0 and xk+1 ≪ [(a − k/n)+] ≪ xk for k = 1, 2,· · · , n − 1. 14 Lemma 5. Let A be a C*-algebra. Let a, b ∈ A+ and ǫ > 0. The following statements hold. (i) If a is Murray-von Neumann equivalent (and hence Cuntz equivalent) to b, then (a−ǫ)+ is Murray-von Neumann equivalent to (b − ǫ)+. (ii) If kak,kbk ≤ 1, ǫ ≤ 1, and a ∈ (b − ǫ)+A(b − ǫ)+, then kba − ak ≤ 2(1 − ǫ). Proof. (i) We must show that (x∗x − ǫ)+ is Murray-von Neumann equivalent to (x∗x − ǫ)+. Consider the polar decomposition x = vx of x in the bidual of A. The element y = v(x∗x − ǫ)1/2 (ii) If kbk ≤ ǫ then the statement is trivial. Let us suppose that kbk > ǫ > 0. With bǫ = fǫ(b), where fǫ(t) = min(kbk, (tkbk)/ǫ), we have bǫ(b − ǫ)+ = kbk(b − ǫ)+ and kb − bǫk = kbk − ǫ. By the first equation, bǫa = kbka, and hence by the second equation, + belongs to A and satisfies y∗y = (x∗x − ǫ)+ and yy∗ = (xx∗ − ǫ)+. kba − ak ≤ kba − kbkak + k(1 − kbk)ak ≤ 2(1 − ǫ). (cid:3) Theorem 4. Let A be a C*-algebra and let (X, v) be a rooted tree. Consider a Cuntz semigroup morphism α : Cu(C0(X \ v)) → Cu(A) satisfying α[sX] ≤ [sA], where sX and sA are a strictly positive element of C0(X\v) and a positive element of A, respectively. Then for every ǫ > 0 there exists a ∗-homomorphism φ : C0(X \ v) → A such that d(X,v) W (α, Cu(φ)) < ǫ. Proof. Let N be a positive integer (to be specified later). Consider the set G(X, v) of generators of the C*-algebra C0(X \ v) defined in Subsection 2.2. Recall that the elements of G(X, v) are indexed by the edges of (X, v). Let e be a fixed but arbitrary such edge. By Lemma 4, with n = 2N and xk = α[(ge − k/n)] for k = 0, 1,· · · , n− 1, there exists a positive element ae ∈ A ⊗ K of norm at most 1 such that α[ge] = [ae], and (4.1) for k = 1, 2,· · · , 2N − 1. Note that for each edge e, [(ae − (k + 1)/2N )+] ≪ α[(ge − k/2N )+] ≪ [(ae − k/2N )+], Xe′ next to e for all 0 ≤ t < 1 (as this holds before applying α), and therefore α[ge′] ≪ [(ae − (2N − 1)/2N )+]. (4.2) α[ge′] ≤ α[(ge − t)+] Xe′ next to e aeae′ = 0 if e and e′ are beside each other. Using the stability of A ⊗ K, we may choose the elements (ae)e∈E(X,v) in such a way that Let 0 < δ < 1/2N be such that (4.1) and (4.2) still hold when the elements ae are replaced with (ae − δ)+ (the existence of the number δ follows from (4.1) and (4.2), the definition of the relation ≪, and from the fact that for b = (ae − k/2N ), k = 0, 1,· · · , 2N − 1 (in fact for any positive element b!), [b] = supδ[(b − δ)+], and (b − δ′)+ ≪ (b − δ′′)+ if δ′ > δ′′). Let us construct as follows a family of positive elements (be)e∈E(X,v) in A such that be is Murray-von Neumann equivalent to (ae − δ)+, such that bebe′ = 0 if e and e′ are beside each other, and such that (4.3) be′ ∈ (be − (2N − 1)/2N )+A(be − (2N − 1)/2N )+ if be′ is next to be. Note that by (ii) of Lemma 5 this last relation implies that kbebe′ − be′k < 1/2N −1 15 for all edges e and e′ such that e′ is next to e. Let us carry out this construction inductively. Let us start by constructing the positive elements be associated to the edges e that have v as a vertex. Denote by ei, i = 1, 2,· · · , k, the edges of (X, v) with one vertex v. Using that the elements aei, i = 1, 2,· · · , k, are pairwise orthogonal we have [aei] = kXi=1 " kXi=1 aei# = More briefly,Pk (aei − δ)+ = kXi=1 kXi=1 α[gei] = α" kXi=1 kXi=1 aei − δ!+ = x∗x, i=1 aei 4 sA. By Lemma 2.2 of [13], there is x ∈ A such that gi# ≤ α[sX] ≤ [sA]. xx∗ ∈ sAAsA = A. Let x = V x be the polar decomposition of the element x in the bidual of A. Set V (aei − δ)+V ∗ = bei, for i = 1, 2,· · · , k. Then the elements bei, i = 1, 2,· · · , k, belong to A and are pairwise orthogonal. Now let us suppose that we have constructed the positive element be associated to the edge e and let us construct the elements be′ associated to the edges e′ that are next to the edge e. By the choice of δ, and the fact that α[ge] = [ae] for every edge e ∈ E(X, v), we obtain from relation (4.2) that Xe′ next to e [ae′] ≪ [(ae − δ − (2N − 1)/2N )+]. By (i) of Lemma 5 applied to the elements (ae − δ)+ and be, and ǫ = (2N − 1)/2N , we have [(ae − δ − (2N − 1)/2N )+] = [(be − (2N − 1)/2N )+]. Therefore, [ae′] ≤ [(be − (2N − 1)/2N )+]. Hence, since the terms of (ae′)e′ next to e are pairwise orthogonal, ae′ 4 (be − (2N − 1)/2N )+. Xe′ next to e Xe′ next to e Therefore, by Lemma 2.2 of [13] there is y ∈ A ⊗ K such that Xe′ next to e (ae′ − δ)+ = y∗y, and yy∗ ∈ (be − (2N − 1)/2N )+A(be − (2N − 1)/2N )+. be′. It is clear that the positive elements be′, e′ next to e, satisfy the required conditions. Let y = Wy be the polar decomposition of y in the bidual of A⊗K. Set W (ae′−δ)+W ∗ = Following this procedure we construct positive contractions (be)e∈E(X,v) -- one positive ele- ment be for each generator ge of C0(X \ v) -- such that bebe′ = 0 if e and e′ are beside each other, and kbebe′ − be′k < 1/2N −1 if e′ is next to e. It follows that these elements form a 1/2N −1 representation in A of the relations (2.4). 16 Fix ǫ > 0. Choose n ≥ 1 such that 1/2n−1 < ǫ. Using the weak stability of the relations (2.4) -- which follows from Theorem 5.1 of [16] and Theorem 14.1.4 of [15] -- we can choose N > n + 1 such that there are positive contractions ce ∈ A satisfying cece′ = 0 if e and e′ are beside each other, cece′ = ce′ if e′ is next to e, and kbe − cek < 1/2n+1. (4.4) The elements ce, e ∈ E(X, v), form a representation of the relations (2.4) in the C*-algebra A. Therefore, they induce a ∗-homomorphism φ : C0(X \ v) → A, which is defined on the generators of C0(X \ v) by φ(ge) = ce. Let us prove that φ is the desired homomorphism. Fix an edge e of (X, v). By Corollary 9.1 of [3] applied to the elements be − k/2n+1 and ce − k/2n+1, k = 0, 1,· · · , 2n+1, (or by Lemma 1 of [20] applied to the elements be and ce) we have (4.5) [(be − (k + 1)/2n+1)+] ≤ [(ce − k/2n+1)+], [(ce − (k + 1)/2n+1)+] ≤ [(be − k/2n+1)+], for k = 0, 1,· · · , 2n+1 − 1. Since be is Murray-von Neumann equivalent to (ae − ǫ)+, by (i) of Lemma 5 the relation (4.1) holds for k = 1, 2,· · · , 2N − 1 when ae is replaced with be. Therefore, [(be − (k + 1)/2N )+] ≤ α[(ge − k/2N )+] ≤ [(be − k/2N )+], for k = 1, 2,· · · , 2N − 1. Since N > n + 1 the preceding equation implies that [(be − (k + 1)/2n+1)+] ≤ α[(ge − k/2n+1)+] ≤ [(be − k/2n+1)+], (4.6) for k = 1, 2,· · · , 2n+1 − 1. It follows from the inequalities (4.5) and (4.6) that [(ce − (k + 1)/2n)+] = [(ce − (2k + 2)/2n+1)+] ≤ [(be − (2k + 1)/2n+1)+] ≤ α[(ge − 2k/2n+1)+] = α[(ge − k/2n)+], and α[(ge − (k + 1)/2n)+] = [(be − (k + 1)/2n)+] = [(be − (2k + 2)/2n+1)+] ≤ [(ce − 2k/2n+1)+] = [(ce − k/2n)+], for k = 1, 2,· · · , 2n − 1. These inequalities may be rewritten as Cu(φ)[(ge − (k + 1)/2n)+)] ≤ α[(ge − k/2n)+], α[(ge − (k + 1)/2n)+] ≤ Cu(φ)[(ge − k/2n)+], for k = 1, 2,· · · , 2n − 1. Any interval of length 1/2n−1 contains an interval of the form (k/2n, (k + 1)/2n) for some integer k. Thus, for every t ∈ [0, 1] there exists k such that (k/2n, (k+1)/2n) ⊆ (t, t+1/2n−1). It follows from the preceding inequalities that Cu(φ)[(ge − t − 1/2n−1)+] ≤ Cu(φ)[(ge − (k + 1)/2n)+] ≤ α[(ge − k/2n)+] ≤ α[(ge − t)+], 17 for t ∈ [0, 1]. Interchanging the roles of Cu(φ) and α (noting that they are symmetric) we also have α[(ge − t − 1/2n−1)+] ≤ Cu(φ)[(ge − t)+], for t ∈ [0, 1]. These inequalities can be restated as dW (Cu(φ) ◦ Cu(χge)), α ◦ Cu(χge)) ≤ 1/2n−1. (Note that in the definition of the pseudometric dW it is enough to take t in [0, 1].) Since the inequality above holds for all e ∈ E(X, v) we conclude that d(X,v) W (Cu(φ), α) ≤ 1/2n−1 < ǫ. (cid:3) 5. Proof of Theorem 1 and Corollary 1 Lemma 6. Let A and B be C*-algebras with B of stable rank one. Let φ, ψ : A → B be ∗-homomorphisms such that φ is approximately unitarily equivalent to ψ with the unitaries taken in the unitization of B ⊗ K. Then φ is approximately unitarily equivalent to ψ with the unitaries taken in the unitization of B. Proof. Let F be a finite subset of A, and let 0 < ǫ < 1. Let us choose a positive element a ∈ A such that kak < 1, and (5.1) ka2f − fk < ǫ, kaf a − fk < ǫ, for all f ∈ F (a can be chosen to be a suitable element of an approximate unit of the C*-algebra generated by the elements of F ). Since φ and ψ are approximately unitarily equivalent with unitaries taken in (B ⊗ K)e, there exists a unitary U ∈ (B ⊗ K)esuch that kUφ(a2)U ∗ − ψ(a2)k < ǫ, kUφ(f )U ∗ − ψ(f )k < ǫ, (5.2) for all f ∈ F . Set ψ(a)Uφ(a) = z. Then by the triangle inequality and the second inequality in (5.1) and (5.2) we have kzφ(f )z∗ − ψ(f )k = kψ(a)Uφ(af a)U ∗ψ(a) − ψ(f )k ≤ ≤ kψ(a)U(φ(af a) − φ(f ))U ∗ψ(a)k + kψ(a)(Uφ(f )U ∗ − ψ(f ))ψ(a)k + kψ(af a) − ψ(f )k ≤ ≤ ǫ + ǫ + ǫ = 3ǫ, for all f ∈ F . Also, using the first inequality in (5.2) we obtain that kz∗z − φ(a4)k = kφ(a)(Uψ(a2)U ∗ − φ(a2))φ(a)k < ǫ. It follows that kzk < 1 and (5.3) kzφ(f )z∗ − ψ(f )k < 3ǫ, kz − φ(a2)k < √ǫ, for all f ∈ F , with Lemma 1 being used to obtain the last inequality. Since the C*-algebra B has stable rank one we may assume that z is an invertible element of eB. Let us denote by 18 W the unitary in the polar decomposition of the element z. Then by the triangle inequality and the inequalities (5.1) and (5.3) we have kW φ(f )W ∗ − ψ(f )k ≤ kW (φ(f ) − φ(a2f ))W ∗k + kW φ(a2)φ(f )W ∗ − zφ(f )W ∗k+ + kzφ(f )W ∗ − zφ(f a2)W ∗k + kzφ(f )φ(a2)W ∗ − zφ(f )z∗k + kzφ(f )z∗ − ψ(f )k < < ǫ + √ǫ + ǫ + √ǫ + ǫ ≤ 5√ǫ, for all f ∈ F . ψ are approximately unitarily equivalent with unitaries taken in the unitization of B. Since the finite subset F and the positive number ǫ are arbitrary we conclude that φ and (cid:3) Let A and B be C*-algebras such that A has a strictly positive element -- say sA. Let us say that the ordered pair (A, B) has the property (P) if for any Cuntz semigroup morphism α : Cu(A) → Cu(B) such that α[sA] ≤ [sB], where sB is a positive element of B, there exists a ∗-homomorphism φ : A → B -- unique up to approximate unitary equivalence -- such that α = Cu(φ). Proposition 5. The following statements hold true: (i) Let B be a C*-algebra of stable rank one and let (X, v) be a rooted tree. If the pair (C0(X \ v), pBp) has the property (P) for every projection p of B, then the pair (C(X), B) has the property (P). (ii) If the pair of C*-algebras (A, B) has the property (P) and B has stable rank one, then (iii) Let C be a C*-algebra of stable rank one. the pair (Mn(A), B) has the property (P), for every n ∈ N. If the pairs of C*-algebras (A, D) and (B, D) have the property (P) for all hereditary subalgebras D of C, then the pair (A ⊕ B, C) has the property (P). (iv) If the pairs of C*-algebras (Ai, B) have the property (P) for a sequence A1 ρ1−→ A2 ρ2−→ · · · , (Ai, ρi), B) has the property (P). then the pair (lim −→ (v) Let A, B, and C be C*-algebras such that A is stably isomorphic to B, and C has stable rank one. If the pair (A, C ⊗ K) has the property (P), then the pair (B, C) has the property (P). Proof. (i) Let α : Cu(C(X)) → Cu(B) be a morphism in the Cuntz category satisfying α([1X]) ≤ [sB]. Let us show that it is induced by a ∗-homomorphism ϕ : C(X) → B. As B has stable rank one, the element α([1X]) appears as the Cuntz class of a projection of B, say p. We have the following identifications: Cu(pBp) ∼= {x ∈ Cu(B) x ≤ ∞[p]}, Cu(C0(X \ v)) ∼= {x ∈ Cu(C(X)) x ≤ ∞[sX]}, where sX is a strictly positive element of C(X, v). It follows that α[sX] ≤ [p], and α(Cu(C0(X \ v))) ⊆ Cu(pBp). By assumption the pair (Cu(C0(X \ v)), pBp) has the property (P). Therefore, there is a ∗-homomorphism φ : C0(X \ v) → pBp such that Cu(φ) is equal to the restriction of α to Cu(C0(X \ v)). Let us consider the restriction of α to Cu(C0(X \ v)). Denote by 19 φ : C(X) → B the ∗-homomorphism which agrees with φ on C0(X \ v) and satisfies φ(1) = p. Then Cu( φ) = α. (The proof of this fact is similar to the proof of (i) of Proposition 4.) Now let us show that if φ, ψ : C(X) → B are ∗-homomorphisms such that Cu(φ) = Cu(ψ), then they are approximately unitarily equivalent. Since Cu(φ)[1X] = Cu(ψ)[1X ] and B has stable rank one we may assume that φ(1X) = ψ(1X) = p, where p ∈ B is a projection. Let us denote by φ′, ψ′ : C0(X \ v) → pBp ⊆ B the restrictions of the ∗-homomorphisms φ and ψ to C0(X \ v). Since Cu(φ) = Cu(ψ), we have d(X,v) W (Cu(φ′), Cu(ψ)) = 0. Using the relation between d(X,v) (φ′, ψ′) = 0 (the unitaries taken in pBp). It follows now that φ and ψ are approximately unitarily equivalent, as desired. established in Theorem 3, we conclude that d(X,v) U and d(X,v) W U (ii) Let us start by showing that any Cuntz semigroup morphism α : Cu(Mn(A)) → B, satisfying α[sA ⊗ 1n] ≤ [sB], where 1n denotes the unit of Mn(C), can be lifted to a ∗- homomorphism φ : Mn(A) → B such that Cu(φ) = α. Let us consider the inclusion map iA : A → Mn(A) given by iA(a) = a ⊗ e1,1. The map Cu(iA) is an isomorphism. We have Cu(A) Cu(iA) / Cu(Mn(A)) α / / Cu(B). Since the pair (A, B) has the property (P), there exists a ∗-homomorphism ψ : A → B such that α ◦ Cu(iA) = Cu(ψ). It follows that α = Cu(ψ) ◦ Cu(iA)−1. Using the commutativity of the diagram ψ A iA B iB Mn(A) ψ⊗id / Mn(B) we obtain that Cu(ψ ⊗ id) = Cu(iB)◦ Cu(ψ)◦ Cu(iA)−1, and hence Cu(ψ ⊗ id) = Cu(iB)◦ α. Therefore, [(ψ ⊗ id)(sA ⊗ 1n)] = Cu(iB) ◦ α[sA ⊗ 1n] ≤ Cu(iB)[sB] = [sB ⊗ e1,1]. Since the C*-algebra A has stable rank one, Mn(A) has stable rank one. It follows by (ii) of Proposition 1 that there is an element x ∈ Mn(B) such that (ψ ⊗ id)(sA ⊗ 1n) = x∗x, xx∗ ∈ (sB ⊗ e1,1)Mn(B)(sB ⊗ e1,1). Let x = V x be the polar decomposition of the element x in the bidual of Mn(B). Then Ad(V ∗) ◦ (ψ ⊗ id) is a ∗-homomorphism with image contained in the hereditary subalgebra (sB ⊗ e1,1)Mn(B)(sB ⊗ e1,1). Also, Cu(Ad(V ∗) ◦ (ψ ⊗ id)) = Cu(ψ ⊗ id). Denote by γ : (sB ⊗ e1,1)Mn(B)(sB ⊗ e1,1) → B the isomorphism defined by γ(b ⊗ e1,1) = b for all b ∈ B. Then Cu(γ) = Cu(iB)−1, since Cu(B ⊗ e1,1) = Cu(B ⊗ K). Set γ ◦ Ad(V ∗) ◦ (ψ ⊗ id) = φ. It follows that Cu(φ) = α. Let φ, ψ : Mn(A) → B be ∗-homomorphisms such that Cu(φ) = Cu(ψ). Let us show that φ and ψ are approximately unitarily equivalent. Let us consider a finite subset of Mn(A) of the form {a ⊗ ek,l 1 ≤ k, l ≤ n, a ∈ F,kak < 1}, 20 / / /     / where F is a finite subset of A. Let 0 < ǫ < 1 and let b ∈ A be a positive contraction such that (5.4) for all a ∈ F . Since the pair (A, B) has the property (P) and Cu(φ◦ iA) = Cu(φ◦ iA), where kbab − ak < ǫ, kb2a − ak < ǫ, iA : A → Mn(A) is the inclusion map, there is a unitary U ∈ eB such that kUφ(a ⊗ e1,1)U ∗ − ψ(a ⊗ e1,1)k < ǫ, (5.5) for all a ∈ F ∪ {b2}. Let us set nXi=1 ψ(b ⊗ ei,1)Uφ(b ⊗ e1,i) = z. Using the inequalities (5.4) and (5.5) we get for all a ∈ F that kzφ(a ⊗ ek,l)z∗ − ψ(a ⊗ ek,l)k = kψ(b ⊗ ek,1)Uφ(bab ⊗ e1,1)U ∗ψ(b ⊗ e1,l) − ψ(a ⊗ ek,l)k ≤ ≤ kψ(b ⊗ ek,1)(Uφ(bab ⊗ e1,1)U ∗ − ψ(a ⊗ e1,1))ψ(b ⊗ e1,l)k + kψ(bab ⊗ ek,l) − ψ(a ⊗ ek,l)k < 2ǫ + ǫ = 3ǫ. φ(b ⊗ ei,1)(U ∗ψ(b2 ⊗ e1,1)U − φ(b2 ⊗ e1,1))φ(b ⊗ e1,i)k < ǫ. Also, using the inequality (5.5) we have kz∗z − φ(b4 ⊗ 1n)k = k nXi=1 Hence, we have found an element z ∈ B such that kz − φ(b2 ⊗ 1n)k < √ǫ, (5.6) kzφ(a ⊗ ek,l)z∗ − ψ(a ⊗ ek,l)k < 3ǫ, (5.7) for all a ∈ F and k, l = 1, 2,· · · , n. Since B has stable rank one we can approximate the hold when z is replaced by z′. Denote by W the unitary in the polar decomposition of the invertible element z′. Then, element z by an invertible element z′ ∈ eB in such a way that inequalities (5.6) and (5.7) still kW φ(a ⊗ ek,l)W ∗ − ψ(a ⊗ ek,l)k = kW (φ(a ⊗ ek,l) − φ(b2a ⊗ ek,l))W ∗k+ + kW (φ(b2 ⊗ ek,l) − z′)φ(a ⊗ ek,l)W ∗k + kWz′(φ(a ⊗ ek,l) − φ(ab2 ⊗ ek,l))W ∗k+ + kWz′φ(a ⊗ ek,l)(φ(b2 ⊗ ek,l) − z′)W ∗k + kz′φ(a ⊗ ek,l)(z′)∗ − ψ(a ⊗ ek,l)k < ǫ + √ǫ + 2ǫ + 2√ǫ + 3ǫ < 9√ǫ, for all a ∈ F and k, l = 1, 2,· · · , n. Therefore, the unitary W ∈ eB satisfies kW φ(a ⊗ ek,l)W ∗ − ψ(a ⊗ ek,l)k < 9√ǫ, for all a ∈ F and k, l = 1, 2,· · · , n. This proves that the ∗-homomorphisms φ and ψ are approximately unitarily equivalent. (iii) Let α : Cu(A⊕ B) → Cu(C) be such that α[sA ⊕ sB] ≤ [sC], where sA, sB are strictly positive elements of A and B, and sC is a positive element of C. Let us consider the positive elements a, b ∈ C ⊗K such that α[sA⊕ 0] = [a] and α[0⊕ sB] = [b]. By the stability of C ⊗K we may assume that a and b are orthogonal. We have α[sA ⊕ sB] = [a] + [b] = [a + b] ≤ [sC]. 21 Applying (ii) of Proposition 1 to the C*-algebra C and the positive elements a + b and sC, we can find an element x ∈ C ⊗ K such that a + b = x∗x, xx∗ ∈ C. Consider the polar decomposition x = V x of x in the bidual of C ⊗ K. Set V aV ∗ = a′ and V bV ∗ = b′. Then a′ and b′ are orthogonal elements of C + satisfying [a] = [a′] and [b] = [b′]. Also, we have the following natural isomorphisms: Cu(a′Aa′) ∼= {[z] ∈ Cu(C) [z] ≤ ∞[a′]}, Cu(b′Ab′) ∼= {[z] ∈ Cu(C) [z] ≤ ∞[b′]}. Using these identifications and the fact that Cu(A ⊕ B) is naturally isomorphic to Cu(A) ⊕ Cu(B), we can identify the morphism α with a pair of Cuntz semigroup morphisms (α1, α2), α1 : Cu(A) → Cu(a′Aa′), α1[sA] ≤ [a1], α2 : Cu(B) → Cu(b′Ab′), α1[sB] ≤ [b2]. Since by hypothesis the pairs (A, a′Aa′), and (B, b′Ab′) have the property (P) we can find ∗-homomorphisms φ1 : A → a′Aa′, φ2 : A → b′Ab′ such that Cu(φ1) = α1 and Cu(φ2) = α2. It follows that the ∗-homomorphism φ = (φ1, φ2) : A ⊕ B → C induces the morphism α at the level of Cuntz semigroups. Let φ, ψ : A ⊕ B → C be ∗-homomorphisms such that Cu(φ) = Cu(ψ). Let us show that Since Cu(φ) = Cu(ψ) we have φ(sA) ∼ ψ(sA), and φ(sB) ∼ ψ(sB) (we are using the identifications A ⊕ 0 ∼= A and 0 ⊕ B ∼= B). Again by (ii) of Proposition 1 we can find elements x1, x2 ∈ C such that φ and ψ are approximately unitarily equivalent. φ(sA) = x∗ φ(sB) = x∗ 1x1, 2x2, x1x∗ x2x∗ 1 ∈ ψ(sA)Cψ(sA), 2 ∈ ψ(sB)Cψ(sB). Set x1+x2 = x. Consider the partial isometry V in the polar decomposition of x in the bidual of C. Since the elements x1 and x2 satisfy the orthogonality relations x∗ 1 = 0, we have 1x2 = x2x∗ x1 = V x1, x2 = V x2. It follows from these identities that the map Ad(V ∗)◦ φ is a ∗-homomorphism. Also, we have Cu(Ad(V ∗)◦φ) = Cu(φ). Let us denote by φA and φB the restrictions of the ∗-homomorphism Ad(V ∗) ◦ φ to the C*-algebras A and B, respectively. Then φA(A) ⊆ ψ(sA)Cψ(sA), φB(B) ⊆ ψ(sB)Cψ(sB). Since by hypothesis the pair of C*-algebras (A, ψ(sA)Cψ(sA)) has the property (P), the ∗-homomorphism φA is approximately unitarily equivalent -- with the unitaries taken in the unitization of the C*-algebra ψ(sA)Cψ(sA) -- to the restriction of the *-homomorphism ψ to the C*-algebra A. In similar way the ∗-homomorphism φB and the restriction of the ∗-homomorphism ψ to the C*-algebra B are approximately unitarily equivalent in the uniti- zation of the C*-algebra ψ(sB)Cψ(sB). It follows that the ∗-homomorphisms Ad(V ∗)◦ φ and ψ are approximately unitarily equivalent in the unitization of the C*-algebra C. In order to 22 complete the proof let us show that the ∗-homomorphisms Ad(V ∗) ◦ φ and φ are approxi- mately unitarily equivalent. Recall that V is the partial isometry in the polar decomposition of the element x = x1 + x2. It follows by Proposition 1 applied to the C*-algebra C and the element x that for every ǫ > 0 and every finite subset F of the hereditary algebra x∗Cx there is a unitary U ∈ C ∼ such that kV zV ∗ − UzU ∗k < ǫ, (iv) Let A = lim −→ for all z ∈ F . This implies that Ad(V ∗) ◦ φ and φ are approximately unitarily equivalent, since φ(A ⊕ B) ⊆ x∗Cx. (Ai, ρi). For each 1 ≤ i < j let ρi,j : Ai → Aj denote the ∗-homomorphism ρj−1 ◦ ρj−2 ◦· · ·◦ ρi. Also, for each 1 ≤ i let ρi,∞ : Ai → A denote the ∗-homomorphism given by the inductive limit. Let α : A → B be a Cuntz semigroup morphism such that α[sA] ≤ [sB], where sA is a strictly positive elements of A, and sB is a positive element of B. For each i ≥ 1 set α ◦ Cu(ρi,∞) = αi. We have αi[sAi] = α[ρi,∞(sAi)] ≤ α[sA] ≤ [sB]. where sAi denotes a strictly positive element of Ai. Since the pairs (Ai, B) have the property (P) for all i ≥ 1, there exist ∗-homomorphisms φi : Ai → B, such that Cu(φi) = αi. For each i we have Cu(φi) = Cu(φi+1 ◦ ρi). Hence the ∗-homomorphisms φi and φi+1 ◦ ρi are approximately unitarily equivalent. By Subsection 2.3 of [8] there exists a ∗-homomorphism φ : A → B such that for every i ≥ 1 the ∗-homomorphisms φ◦ ρi,∞ and φi are approximately unitarily equivalent. Since the Cuntz functor is equal in ∗-homomorphisms that are approxi- mately unitarily equivalent we have Cu(φ◦ ρi,∞) = Cu(φi). Therefore, Cu(φ)◦ Cu(ρi,∞) = αi for all i ≥ 1. By the universal property of the inductive limit this implies that α = Cu(φ). To conclude the proof of (iv) let us show that if homomorphisms φ, ψ : A → B are such that Cu(φ) = Cu(ψ), then they are approximately unitarily equivalent. For each i ≥ 1 set φ◦ ρi,∞ = φi, and ψ◦ ρi,∞ = ψi. We have Cu(φi) = Cu(ψi) for i ≥ 1. Since for each i ≥ 1 the pair (Ai, B) has the property (P) the ∗-homomorphisms φi and ψi are approximately uni- tarily equivalent. It follows that the ∗-homomorphisms φ and ψ are approximately unitarily equivalent. (v) The pair (A, C ⊗ K) has the property (P) by assumption, and it follows by Lemma 6 that the pair (A, C) also has the property (P). It follows by (ii) and (iv) that also the pairs (A ⊗ K, C) and (A ⊗ K, C ⊗ K) have the property (P). Since A ⊗ K ∼= B ⊗ K, the pairs (B ⊗ K, C) and (B ⊗ K, C ⊗ K) have the property (P). Let α : Cu(B) → Cu(C) be a Cuntz semigroup morphism such that α[sB] ≤ [sC], where sB is a strictly positive element of B, and sC is a positive element of C. Let iB : B → B ⊗K and iC : C → C ⊗ K denote the inclusion maps iB(b) = b ⊗ e1,1 and iC(c) = c ⊗ e1,1. Then Cu(iB) and Cu(iC) are isomorphisms. Since the pair (B ⊗ K, C ⊗ K) has the property (P), there exists a ∗-homomorphism φ : B⊗K → C ⊗K such that Cu(φ) = Cu(iC)◦ α◦ Cu(iB)−1. We have Cu(φ ◦ iB)[sB] = (Cu(iC) ◦ α)[sB] ≤ [sC ⊗ e1,1]. By Proposition 1 there exists x ∈ C ⊗ K such that (φ ◦ iB)(sB) = x∗x, xx∗ ∈ (sC ⊗ e1,1)(C ⊗ K)(sC ⊗ e1,1). 23 Consider the polar decomposition x = V x of x in the bidual of C ⊗ K. Then (Ad(V ∗) ◦ φ ◦ iB)(B) ⊆ (sC ⊗ e1,1)(C ⊗ K)(sC ⊗ e1,1) = C ⊗ e1,1, and Cu(Ad(V ∗) ◦ φ ◦ iB) = Cu(φ ◦ iB). Denote by γ : C ⊗ e1,1 → C the ∗-isomorphism γ(c ⊗ e1,1) = c for all c ∈ C. Then Cu(γ) = Cu(iC)−1, since Cu(C ⊗ e1,1) = Cu(C ⊗ K). Set γ ◦ Ad(V ∗) ◦ φ ◦ iB = φ′. It follows that φ′ : B → C, and Cu(φ′) = Cu(γ ◦ Ad(V ∗) ◦ φ ◦ iB) = Cu(iC)−1 ◦ Cu(φ) ◦ Cu(iB) = α. Now let us show that if φ, ψ : B → C are ∗-homomorphisms such that Cu(φ) = Cu(ψ), then they are approximately unitarily equivalent, with the unitaries taken in the unitization of C. Since Cu(φ) = Cu(ψ) it follows that Cu(φ ⊗ id) = Cu(ψ ⊗ id), where id : K → K is the identity map. This implies that φ⊗id is approximately unitarily equivalent to ψ⊗id, with the equivalent. Since the images of both maps are contained in C, φ and ψ are approximately unitarily equivalent with the unitaries taken in the unitization of C, by Lemma 6. unitaries taken in (C ⊗K)e. Hence, (ψ⊗ id)◦ iB and (ψ⊗ id)◦ iB are approximately unitarily (cid:3) Proof of Theorem 1. By the statements (i), (iii), (iv), and (v) of Proposition 5, and by Proposition 3 it is enough to prove the theorem in the case A = C0(X \ v), where (X, v) is a rooted tree. The uniqueness of the homomorphism φ is a special case of Theorem 2. Let us prove its existence. By Theorem 4, for every n there exists φn : C0(X \ v) → B such that d(X,v) W (Cu(φn), α) < 1/2n+2. It follows by Theorem 2 that d(X,v) U (φn, φn+1) ≤ (2N + 2)d(X,v) W (Cu(φn), Cu(φn+1)) < 1 2n (2N + 2). This implies that (φn)n is a Cauchy sequence with respect to the pseudometric d(X,v) (ii) of Proposition 4, the space Hom(C0(X \ v), B) is complete with respect to d(X,v) there exists φ : C0(X\v) → B such that d(X,v) 3, . By U . Hence, (φ, φn) → 0. By the first inequality of Theorem U U d(X,v) W (Cu(φ), α) ≤ d(X,v) ≤ d(X,v) U W (Cu(φ), Cu(φn)) + d(X,v) W (Cu(φn), α), (φ, φn) + d(X,v) W (Cu(φn), α) → 0. In other words, d(X,v) metric, we have Cu(φ) = α, as desired. W (Cu(φ), α) = 0. Since, as shown in (i) of Proposition 4, d(X,v) W is a (cid:3) Proof of Corollary 1. Let A and B be sequential inductive limits of separable continuous- trace C*-algebras with spectrum homeomorphic to a disjoint union of trees and trees with a point removed. Let α : Cu(A) → Cu(B) be a Cuntz semigroup isomorphism such that α[sA] = [sB]. Then by Theorem 1, there are ∗-homomorphisms φ : A → B and ψ : B → A such that Cu(φ) = α, and Cu(ψ) = α−1. We have Cu(φ ◦ ψ) = Cu(idA) and Cu(ψ ◦ φ) = Cu(idB), where idA and idB denote the identity endomorphisms of A and B. Therefore by the uniqueness part of Theorem 1, φ ◦ ψ and idA, and ψ ◦ φ and idB are approximately unitarily equivalent. Hence, by Theorem 3 of [10] (cf. Section 4.3 of [10]), there exists an isomorphism ρ : A → B approximately unitarily equivalent to φ, and in particular by 24 definition of Cu(B) such that Cu(ρ) = Cu(φ) = α. The ∗-homomorphism ρ is unique -- up to approximate unitary equivalence -- by Theorem 1. (cid:3) 5.1. Remarks. It follows from (v) of Proposition 5 that Theorem 1 still holds if A is taken to be a full hereditary subalgebra of a C*-algebra that can be written as a sequential inductive limit of separable continuous-trace C*-algebras with spectrum homeomorphic to a disjoint union of trees and trees with a point removed. The same applies to Corollary 1 for the C*-algebras A and B. We don't know if in general every such full hereditary subalgebra can be written as a sequential inductive limit of continuous-trace C*-algebras with spectrum homeomorphic to a disjoint union of trees and trees with a point removed. We have the following partial result: Proposition 6. Let A be a full hereditary subalgebra of a C*-algebra that can be written as a sequential inductive limit of continuous-trace C*-algebras with spectrum homeomorphic to a compact tree. Then A can be written as a sequential inductive limit of continuous-trace C*-algebras with spectrum homeomorphic to a compact tree. Proof. Let B = lim Bi, where each Bi is a continuous-trace C*-algebras with spectrum −→ homeomorphic to a tree. Let A be a full hereditary subalgebra of B. Since B has stable rank one, by Corollary 4 of [4] the C*-algebra A can be written as a sequential inductive limit of hereditary subalgebras of the C*-algebras Bi, i = 1, 2,· · · . Let us denote these hereditary It follows that each C*-algebra Ai is a continuous-trace subalgebras by Ai, i = 1, 2,· · · . C*-algebra. Therefore, we have lim −→ (C0( Ai) ⊗ K) ∼= lim −→ (Ai ⊗ K) = A ⊗ K ∼= B ⊗ K. (C( Bi)⊗K), the C*-algebra B⊗K contains a nonzero projection. It follows Since B⊗K = lim −→ that for i large enough the C*-algebras Ai ⊗ K also contain nonzero projections. Therefore, Ai is compact for i large enough. Since for each i = 1, 2,· · · , the C*-algebra C( Ai) ⊗ K is a hereditary subalgebra of C( Bi) ⊗ K, and Bi is a compact tree, then for i large enough the set Ai is a compact tree. This concludes the proof of the proposition. (cid:3) Acknowledgments The research of the second author was supported by the Natural Sciences and Engineering Research Council of Canada. The third author was partially supported by DGI MICIIN-FEDER MTM2008-06201-C02-01, and by the Comissionat per Universitats i Recerca de la Generalitat de Catalunya. Part of this research was carried out in the 2008 -- 2009 academic year while the first and third authors were postdoctoral fellows at the Fields Institute for Research in Mathematical Sciences with the support of Dr. G. A. Elliott's research grant. The first author acknowledges the support of an AARMS Postdoctoral Fellowship, while the third author acknowledges the support of a Juan de la Cierva Postdoctoral Fellowship. References [1] N. P. Brown, F. Perera, and A. S. Toms, The Cuntz semigroup, the Elliott conjecture, and dimension functions on C ∗-algebras, J. Reine Angew. Math. 621 (2008), 191 -- 211. [2] A. Ciuperca, Some properties of the Cuntz semigroup and an isomorphism theorem for a certain class of non-simple C ∗-algebras, Thesis, University of Toronto, 2008. 25 [3] Alin Ciuperca and George A. Elliott, A remark on invariants for C ∗-algebras of stable rank one, Int. Math. Res. Not. IMRN 5 (2008), 33. [4] K. T. Coward, G. A. Elliott, and C. Ivanescu, The Cuntz semigroup as an invariant for C ∗-algebras, J. Reine Angew. Math. 623 (2008), 161 -- 193. [5] J. Cuntz, Dimension functions on simple C ∗-algebras, Math. Ann. 233 (1978), no. 2, 145 -- 153. [6] Jacques Dixmier, C ∗-algebras, North-Holland Publishing Co., Amsterdam, 1977. Translated from the French by Francis Jellett; North-Holland Mathematical Library, Vol. 15. [7] G. A. Elliott, A classification of certain simple C ∗-algebras, Quantum and non-commutative analysis (Kyoto, 1992), Math. Phys. Stud., Vol. 16, Kluwer Acad. Publ., Dordrecht, 1993, pp. 373 -- 385. [8] [9] , On the classification of C ∗-algebras of real rank zero, J. Reine Angew. Math. 443 (1993), 179 -- 219. , Hilbert modules over a C ∗-algebra of stable rank one, C. R. Math. Acad. Sci. Soc. R. Can. 29 (2007), no. 2, 48 -- 51. , Towards a theory of classification, Adv. Math. 223 (2010), no. 1, 30 -- 48. [10] [11] G. A. Elliott and C. Ivanescu, The classification of separable simple C*-algebras which are inductive limits of continuous-trace C*-algebras with spectrum homeomorphic to the closed interval [0, 1], J. Funct. Anal. 254 (2008), no. 4, 879 -- 903. [12] G. A. Elliott, L. Robert, and L. Santiago, The cone of lower semicontinuous traces on a C ∗-algebra (2008), preprint, arXiv:0805.3122v2. [13] E. Kirchberg and M. Rørdam, Infinite non-simple C ∗-algebras: absorbing the Cuntz algebras O∞, Adv. [14] L. Li, Classification of simple C ∗-algebras: inductive limits of matrix algebras over trees, Mem. Amer. Math. 167 (2002), no. 2, 195 -- 264. Math. Soc. 127 (1997), no. 605. [15] T. A. Loring, Lifting solutions to perturbing problems in C ∗-algebras, Fields Institute Monographs, Vol. 8, American Mathematical Society, Providence, RI, 1997. [16] , Stable relations. II. Corona semiprojectivity and dimension-drop C ∗-algebras, Pacific J. Math. 172 (1996), no. 2, 461 -- 475. [17] Gert K. Pedersen, C ∗-algebras and their automorphism groups, London Mathematical Society Mono- graphs, Vol. 14, Academic Press Inc. [Harcourt Brace Jovanovich Publishers], London, 1979. [18] L. Robert, Classification of non-simple approximate interval C ∗-algebras: the triangular case, Thesis, University of Toronto, 2006. [19] , The Cuntz semigroup of some spaces of dimension at most 2, C. R. Math. Acad. Sci. Soc. R. Can., to appear. [20] L. Robert and L. Santiago, On the classification of C ∗-homomorphisms from C0(0, 1] to a C ∗-algebra of stable rank greater than 1, J. Funct. Anal., to appear. [21] M. Rørdam and W. Winter, The Jiang-Su algebra revisited, J. Reine Angew. Math., to appear. [22] L. Santiago, Classification of non-simple C ∗-algebras: Inductive limits of splitting interval algebras, Thesis, University of Toronto, 2008. [23] K. H. Stevens, The classification of certain non-simple approximate interval algebras, Operator algebras and their applications, II (Waterloo, ON, 1994/1995), Fields Inst. Commun., Vol. 20, Amer. Math. Soc., Providence, RI, 1998, pp. 105 -- 148. [24] K. Thomsen, Inductive limits of interval algebras: unitary orbits of positive elements, Math. Ann. 293 (1992), no. 1, 47 -- 63. [25] A. S. Toms, An infinite family of non-isomorphic C ∗-algebras with identical K-theory, Trans. Amer. Math. Soc. 360 (2008), no. 10, 5343 -- 5354. Alin Ciuperca, Department of Mathematics and Statistics, University of New Brunswick, Fredericton, New Brunswick E3B 5A3, Canada E-mail address: [email protected] George A. Elliott, Department of Mathematics, University of Toronto, Toronto, On- tario M5S 2E4, Canada E-mail address: [email protected] 26 Luis Santiago, Departament de Matem`atiques, Edifici C, Universitat Aut`onoma de Barcelona, Bellaterra, Barcelona 08193, Spain E-mail address: [email protected] 27
1906.07553
1
1906
2019-06-18T13:23:41
Type-zero ternary corners
[ "math.OA" ]
In this paper we discuss the relationship between a TRO $\mathcal{T}$ and a sub-TRO $\mathcal{S}$ that is the range of a TRO-conditional expectation on $\mathcal{T}$, a \textit{ternary corner}, by investigating a special class $\mathcal{D}$ of bounded linear maps on~$\mathcal{T}$. We pay particular attention to the case when the TROs contain partial isometries.
math.OA
math
TYPE-ZERO TERNARY CORNERS Y. ESTAREMI AND M. MATHIEU Dedicated to the memory of Richard M. Timoney (1953 -- 2019). Abstract. In this paper we discuss the relationship between a TRO T and a sub- TRO S that is the range of a TRO-conditional expectation on T , a ternary corner, by investigating a special class D of bounded linear maps on T . We pay particular attention to the case when the TROs contain partial isometries. 1. Introduction A ternary ring of operators (TRO) between Hilbert spaces K and H is a norm-closed subspace X of B(K, H) which is algebraically closed under the ternary product [x, y, z] = xy∗z for all x, y, z ∈ X . A TRO X ⊆ B(K, H) is called a W ∗-TRO if it is weak∗ closed in B(K, H). TROs are widely studied by many authors; for instance, in [11], the authors proved that TROs form a special class of concrete operator spaces and characterized TROs in terms of the operator space theoretic properties. The interconnections between TROs and JC*-triples are studied in [4, 5, 6]; compare also [3]. It is well known that an operator space is injective if and only if it is completely isometric to a ternary corner of an injective C*-algebra (see, e.g., [1]). A fundamental tool to study TROs is the construction of the linking algebra, that is, a particular C*-algebra containing the related TRO as a corner. TROs and their associated linking algebras share many common properties wherefore the application of operator algebraic methods simplifies the study of TROs that are not algebras themselves. Basic properties of TROs are discussed in, e.g., [1, 8, 10, 12, 15] and references therein. In [7], the authors studied the relationship between unital C*-algebras and their unital C*-subalgebras that are the range of a C*-conditional expectation by defining a special class of bounded linear maps on the underlying C*-algebra. Inspired by this, in the present paper, we are going to give a similar characterization by introducing a new class of bounded linear maps on a TRO to study sub-TROs which are the range of a TRO- conditional expectation. The main tool that we use in the proofs are the contractive projections (equivalently, TRO-conditional expectations) on TROs and their properties. Projections and contractive projections on TROs are studied in [2, 8, 12, 13, 14], for example. Our results extend some results from [7] to the setting of TROs. 2010 Mathematics Subject Classification. 46K70, 46L89, 47B48. Key words and phrases. Ternary rings of operators, C*-algebras, conditional expectations. 1 2 Y. ESTAREMI AND M. MATHIEU In Section 2 we define a condition (called type-zero) for a sub-TRO S by using an approximate unit of the left linking algebra of S, which is similar to the one defined for C*-subalgebras in [7]. We also introduce an improvement of this condition (restricted type-zero). We prove that every TRO-conditional expectation is an extension of another TRO-conditional expectation that is defined on a further sub-TRO and is unique. The sub-TRO S of T is restricted type-zero if and only if S contains no non-zero TRO- left ideal of T . Some other related results are obtained. At the end of Section 3, using the operators associated to a TRO-conditional expectation, we show that, under a weak condition and without loss of generality, we can assume a sub-TRO be type-zero (Theorem 3.7). If a sub-TRO which is the range of a TRO-conditional expectation is type-zero, then the corresponding conditional expectation is faithful. In Section 4 we apply the results of Section 2 to the case when S contains a partial isometry e such that ee∗s = se∗e = s for all s ∈ S. 2. Type-zero TROs Let B(K, H) denote the space of bounded linear operators from the Hilbert space K into the Hilbert space H. Let T ⊆ B(K, H) be a TRO. Also, let LT and RT be the C*-algebras generated by T T ∗ and T ∗T , respectively (where T ∗ = {x∗ x ∈ T }). It is known that T T ∗T is norm dense in T . A norm-closed subspace J in a TRO T is called a TRO-left ideal if T T ∗J ⊆ J and TRO-right ideal if J T ∗T ⊆ J , and a TRO ideal if both conditions are satisfied. If J is a TRO ideal, then T J ∗T ⊆ J . Let T be a TRO. By a TRO-conditional expectation on T we mean a completely contractive projection on T (where T is equipped with the canonical operator space structure inherited from B(K, H)) or, equivalently, a continuous linear map E : T → T satisfying E ◦ E = E and, for all x, y, z ∈ T , (2.1) E(x)E(y)∗E(z) = E(xE(y)∗E(z)) = E(E(x)y∗E(z)) = E(E(x)E(y)∗z). It follows that the range of E is a closed sub-TRO of T (consisting of the fixed points of E) and T = S ⊕ K where S and K are the range and the kernel of E, respectively. Moreover, from (2.1), (2.2) SS ∗K ⊆ K, KS ∗S ⊆ K, so that the kernel K can be regarded as a bimodule over the left and right C*-algebras of S (see [8, 12]). Now we define the ternary corners in TROs which are our main object of investigation. Definition 2.1. Let T be a TRO and let S be a sub-TRO of T . Then S is called a ternary corner of T if there is a Banach space K ⊆ T such that T = S ⊕ K and condition (2.2) holds for K (cf. [12, Section 4.1]). It is not difficult to prove that a sub-TRO S of the TRO T is a ternary corner if and only if there is a TRO-conditional expectation E on T such that E(T ) = S. See, e.g., TERNARY CORNERS 3 [12, Theorem 4.1.3]. For a sub-TRO S ⊆ T we define S1 = {x ∈ T : T T ∗x ⊆ S}. In the next proposition we obtain some basic properties of S1. Proposition 2.2. For a ternary corner S in a TRO T , the following facts hold: (a) S1 is a TRO-left ideal in T and a TRO-right ideal in S. (b) S1 ⊆ S and S1 is norm closed. (c) If I is a TRO-left ideal in T which is contained in S, then I ⊆ S1. (d) S is a TRO-left ideal if and only S = S1. Proof. (a) For every y, z ∈ T and x ∈ S1 we have T T ∗yz∗x ⊆ T z∗x ⊆ S, and so yz∗x ∈ S1, i.e., T T ∗S1 ⊆ S1. Hence S1 is a TRO-left ideal in T . And clearly for each x ∈ S1 we have T T ∗xS ∗S ⊆ SS ∗S ⊆ S, which means that S1 is a TRO-right ideal in S. (b) Let {uα} be an approximate unit of LT . Then for every y, z ∈ T we have kyz∗ − uαyz∗k → 0, kyz∗ − yz∗uαk → 0 and so for x ∈ S1 we have kx − uαxk2 = k(x − uαx)(x − uαx)∗k = kxx∗ − uαxx∗ − xx∗uα + uαxx∗uαk ≤ 2kxx∗ − uαxx∗k → 0. Since uαx ∈ S and S is norm closed, it follows that x ∈ S. Thus S1 ⊆ S. Moreover, if {xα} ⊆ S1, x ∈ S and kxα − xk → 0, then we have kyz∗xα − yz∗xk ≤ kykkzkkxα − xk → 0. Hence S1 is norm closed. (c) If I is a TRO-left ideal in T which is contained in S, then T T ∗I ⊆ I ⊆ S, hence I ⊆ S1. (d) This is clear by definition. (cid:3) Remark 2.3. If T and S are W ∗-TROs, then S1 is also closed in the weak operator topology. Now we characterize when a ternary corner S is a TRO-ideal. Proposition 2.4. Let S be a ternary corner of T with T = S ⊕ K. Then S is a TRO-ideal if and only if KS ∗S = {0} and SS ∗K = {0}. Proof. Let S be a TRO-ideal. Then S is a TRO-left ideal and T T ∗S ⊆ S. So KS ∗S ⊆ S. Moreover, we have KS ∗S ⊆ K. Therefore KS ∗S ⊆ S ∩ K = {0}. 4 Y. ESTAREMI AND M. MATHIEU Since S is a TRO-right ideal, then ST ∗T ⊆ S and so SS ∗K ⊆ S. Moreover, by definition we also have SS ∗K ⊆ K. Therefore SS ∗K ⊆ S ∩ K = {0}. Conversely, let KS ∗S = {0} and SS ∗K = {0}. Since SS ∗S is norm dense in S and KK∗SS ∗S = K(SS ∗K)∗S, SS ∗SK∗K = S(KS ∗S)∗K, then we get that SK∗S = {0} = KK∗S and SS ∗K = {0}. Consequently, we have T T ∗S = SS ∗S + KS ∗S + SK∗S + KK∗S ⊆ S which tell us that S is a TRO-left ideal in T . And similarly we get that ST ∗T ⊆ S. Hence S is a TRO-right ideal in T . (cid:3) A similar argument as in the proof of Proposition 2.2, part (b) together with the continuity and module properties of E shows that, for every x ∈ LS, z ∈ RS and y ∈ T , the following identities hold: (2.3) E(xy) = xE(y), E(yz) = E(y)z. Let x ∈ T = S ⊕ K and write x = s + m for unique s ∈ S and m ∈ K. Let {cα} be an approximate unit in LS. Then limα cαs = s and therefore limα cαx = s + limα cαm provided the latter limit exists. Similarly, if limγ mdγ exists, then limγ xdγ = s + limγ mdγ where {dγ} is an approximate unit in RS. Under these assumptions, it follows that for all x ∈ T the nets {cαx} and {xdγ} converge. Now we define S0 = {x ∈ T : for all y, z ∈ T , lim α (yz∗cαx) exists and belongs to S}. In the sequel we will see that the definition of S0 is independent of the choice of ap- proximate unit. Let S be a ternary corner in T . By definition of S0 we have S1 ⊆ S0; in fact, since S1 ⊆ S and for every x ∈ S we have kx − cαxk → 0, lim(yz∗cαx) = yz∗x ∈ S, for all y, z ∈ T . Consequently, we get that S1 ⊆ E(S0). On the other hand, if x ∈ S0, then by definition, for every t ∈ LT , the net {tcαx} converges to an element of S; in particular, for every s ∈ LS, the net {scαx} converges to an element of S. Hence, if we choose a fixed α0 ∈ I, then limα(cα0cαx) ∈ S and so lim α = lim α (cα0cαx) = E(lim α cα0cαE(x) (cα0cαx)) = cα0E(x). Also, we have limα(cα0cα)x = cα0x; therefore cα0E(x) = cα0x. Since cα0 is arbitrary and limα cαE(x) = E(x), E(x) = limα cαx. Therefore yz∗E(x) = limα yz∗cαx ∈ S. Hence E(x) ∈ S1. From these observations we have the following proposition. TERNARY CORNERS 5 Proposition 2.5. Let T and S be TROs with S ⊆ T . conditional expectation corresponding to S, then we have If E : T → T is a TRO- (2.4) S0(T S) := S0 = {x ∈ T : E(x) ∈ S1} or, equivalently, E(S0) = S1. Moreover, ES0 : S0 → S0 is the unique TRO-conditional expectation onto S1. Note that one inclusion in (2.4) follows from E(S0) ⊆ S1 whereas the reverse inclusion from the fact, as shown above, that lim(yz∗cαE(x)) = yz∗E(x) ∈ S, for all y, z ∈ T and E(x) ∈ S1. Since S1 is norm closed, S0 is also norm closed. We next introduce the main concept of this paper. Definition 2.6. A sub-TRO S ⊆ T is called type-zero if S0 = {0}. Also, S is called restricted type-zero if S1 = {0}. Corollary 2.7. If S is a type-zero ternary corner, then its corresponding conditional expectation is faithful, i.e., K = {0}. Moreover, if the conditional expectation corre- sponding to S is faithful, then S1 = S0. Remark 2.8. By Proposition 2.2 and Definition 2.6 we obtain that the sub-TRO S of T is restricted type-zero if and only if S contains no non-zero TRO-left ideal of T . If S is restricted type-zero, then S contains no non-zero TRO ideal of T . Let us look at an example (cf. also [12, Section 4.5]). Let H be an arbitrary non-zero complex Hilbert space. For non-zero vectors ξ, ζ ∈ H, consider the rank-one operator ξ ⊗ ζ ∗ : H → H, ξ ⊗ ζ ∗(γ) = hγ, ζiξ, γ ∈ H. Let ξ ⊗ ξ ∗ be a rank-one self-adjoint projection. The column Hilbert space Hc is defined as follows Hc = B(H)(ξ ⊗ ξ ∗) = {η ⊗ ξ ∗ : η ∈ H}, with the inner product hη ⊗ ξ ∗, ζ ⊗ ξ ∗i = hη, ζi for ζ, η ∈ H. It is easy to see that Hc is a norm closed subspace of B(H) and closed under the ternary product; indeed we have (η1 ⊗ ξ ∗)(η2 ⊗ ξ ∗)∗(η3 ⊗ ξ ∗) = hη3, η2i(η1 ⊗ ξ ∗) where η1, η2, η3 ∈ H. Hence Hc is a TRO in B(H). Similarly, we consider the row Hilbert space Hr given by Hr = (ξ ⊗ ξ ∗)B(H) = {ξ ⊗ η∗ : η ∈ H} with the inner product hξ ⊗ η∗ 2i = hη1, η2i. Moreover, Hr is also a TRO in B(H). The ternary corners in Hc and in Hr, respectively are precisely the closed subspaces of Hc and of Hr, respectively. 1, ξ ⊗ η∗ Let S be a ternary corner of Hc, E be the corresponding TRO-conditional expectation and h1, h2, h0 ∈ Hc with hi = ηi ⊗ ξ ∗, i = 0, 1, 2. Since h1h∗ 2h0 = hη0, η2ih1, we get that h0 ∈ S1 if and only if h0 = 0, also h0 ∈ S0 if and only if E(h0) ∈ S1, and so if and only if E(h0) = 0. Therefore S1 = {0} and S0 = K. 6 Y. ESTAREMI AND M. MATHIEU Similarly, let S be a ternary corner in Hr, E be the corresponding TRO-conditional expectation and h1, h2, h0 ∈ Hr with hi = ξ ⊗ η∗ 2h0 = hη2, η1ih0, we get that h0 ∈ S1 if and only if h0 ∈ S and also, h0 ∈ S0 if and only if E(h0) ∈ S1, and so if and only if h0 ∈ Hr. i , i = 0, 1, 2. Since h1h∗ Moreover, let Ec : B(H) → B(H) be defined as: Ec(x) = x(ξ) ⊗ ξ ∗. Then Ec is a contractive projection and consequently a TRO-conditional expectation onto S = Hc. Thus Hc is a ternary corner of B(H) and Kc = ker(Ec) = {x ∈ B(H) : x(ξ) = 0}. Let x ∈ B(H). We have x ∈ S1 if and only if yz∗x ∈ Hc for all y, z ∈ B(H) which in turn is equivalent to yz∗x = yz∗x(ξ) ⊗ ξ ∗ for all y, z ∈ B(H). This entails that x ∈ S0 if and only if yz∗Ec(x) = yz∗Ec(x)(ξ) ⊗ ξ ∗ = yz∗x(ξ) ⊗ ξ ∗ = yz∗(x(ξ) ⊗ ξ ∗) for all y, z ∈ B(H). This implies that S0 = B(H) and so S1 = Hc. In a similar vein, by defining Er : B(H) → B(H) as Er(x) = ξ ⊗ x∗(ξ)∗ we obtain a contractive projection and consequently a TRO-conditional expectation Er onto S = Hr. So Hr is also a ternary corner in B(H) and Kr = ker(Er) = {x ∈ B(H) : x∗(ξ) = 0}. A similar argument as above shows that S1 = {0} and so S0 = Kr in this case. We can sum these observations up as follows. Remark 2.9. Let S ⊆ Hc be a ternary corner. Then S1 = {0} and S0 = K. Hence S is restricted type-zero. Moreover, if S ⊆ Hr is a ternary corner. Then S0 = Hr and S1 = S. So S is type-zero if and only if S = Hr = {0}. In addition, Hr is a restricted type-zero ternary corner of B(H). But Hc never can be type-zero in B(H). We shall prepare the introduction of the operators associated with TRO-conditional expectations in the subsequent section by some technical terminology and a lemma. As before, let {cα} be an approximate unit in LS and {dγ} an approximate unit in RS. Let Tl be the set of all x ∈ T such that {cαx} converges in T , Tr be the set of all x ∈ T such that {xdγ} converges in T , and Tl,r = Tl ∩Tr the set of all x ∈ T such that limα,γ(cαxdγ) exists and belongs to T . Direct computations show that Tl, Tr and Tl,r are sub-TROs of the TRO T . From now on we will assume that T = Tl,r. For every x ∈ Tl we have E(lim α (x − cαx)) = lim α (E(x) − cαE(x)) = 0. Hence Kl0 := {y ∈ T : y = lim α (x − cαx) for some x ∈ T } ⊆ K ⊆ S0. Since S ∩ K = {0}, it follows that Kl0 ∩ S1 = {0}. By these observations we get that S1 + Kl0 ⊆ S0. TERNARY CORNERS 7 On the other hand, we know that, for x ∈ S0, E(x) = limα cαx. So we can write x = limα(x − cαx) + E(x). By these observations we obtain the following proposition. Proposition 2.10. Let T and S be TROs with S ⊆ T . If S is a ternary corner of T , then we have S0 = S1 ⊕ Kl0. Corollary 2.11. If S ⊆ T is a type-zero ternary corner and x ∈ T , then limα cαx = x, and limγ xdγ = x as well. Let Kl = {y ∈ T : y = lim α Kr = {y ∈ T : y = lim γ Kl,r = {y ∈ T : y = lim α,γ cαx for some x ∈ K}, xdγ for some x ∈ K}, and cαxdγ for some x ∈ K}. In order to obtain a decomposition for K in the following lemma, we need a final auxiliary set T0 = {y ∈ T : y = lim λ lim β ((x − cλx) − (x − cλx)dβ) for some x ∈ T }. Lemma 2.12. Let S be a ternary corner in T with the TRO-conditional expectation E : T → S. Then we have the following: (a) K = (Kl + Kr) ⊕ T0; (b) Kl ∩ Kr = Kl,r, (c) K = Kl,r if and only if limα cαx = limγ xdγ = x for all x ∈ T . Proof. (a) Let a ∈ T . Then for α ∈ I and γ ∈ J we have a = cαa + adγ − cαadγ + a − cαa + cαadγ − adγ and consequently a − E(a) = cαa − E(a) + adγ − cαadγ + a − cαa + cαadγ − adγ. By taking limits with respect to α and γ we get that a − E(a) = lim α (cα(a − E(a))) + lim α,γ ((a − cαa)dγ) + lim α (a − cαa) + lim α,γ ((cαa − a)dγ). Therefore K ⊆ (Kl + Kr) ⊕ T0. Also, it is clear that (Kl + Kr) ⊕ T0 ⊆ K and so we have (b) For each x ∈ T we have K = (Kl + Kr) ⊕ T0. (cα(x − E(x))dγ) lim α,γ (cαxdγ − E(x)) = lim α,γ = lim α,γ = lim α,γ (cα(xdγ − E(xdγ)) ((cαx − E(cαx)dγ). This implies that Kl,r ⊆ Kl ∩ Kr. 8 Y. ESTAREMI AND M. MATHIEU Conversely, let α0 ∈ I and γ0 ∈ J. Then clearly we have limα cαcα0 = cα0 and limγ dγ0dγ = dγ0. Hence, if x0 ∈ Kl ∩ Kr, then x0 = lim α (cα(x − E(x))) = lim γ ((y − E(y))dγ) for some x, y ∈ T , hence x0 = lim α,γ (cα(x − E(x))dγ) = lim α,γ (cα(y − E(y))dγ). Thus x0 ∈ Kl,r. (c) If K = Kl,r, then for each x ∈ K we have limα cαx = limγ xdγ = x. Thus for every x ∈ T we have limα cαx − x = y ∈ K and so y = limα cαy = limα(cαx − cαx) = 0. Therefore limα cαx − x = y = 0. Similarly we have limγ xdγ − x = 0 and so x = limα cαx = limγ xdγ. Conversely, if limα cαx = limγ xdγ = x for all x ∈ T , then for a ∈ T , a − E(a) = lim α,γ cα(a − E(a))dγ ∈ Kl,r. Hence K = Kl,r. (cid:3) Corollary 2.13. Under the assumptions of Lemma 2.12 we get that Kl + Kr = {0} if and only if Kl = {0} if and only if Kr = {0}. 3. Operators induced by TRO-conditional expectations In [7], the authors studied a certain class of operators on a C*-algebra A associated to a conditional expectation Φ from A to a C*-subalgebra B. These operators can be regarded as a kind of generalized inner derivations with respect to the conditional expectation. Therefore properties of Φ reflect in properties of the algebra of all those operators and vice versa. In this section, we shall introduce a similar class of operators defined on a TRO with respect to a TRO-conditional expectation, and study their interrelations. Let a, b ∈ T and E be the TRO-conditional expectation corresponding to the sub- TRO S ⊆ T . The operator Da,b : T → T defined by Da,b(x) = E(a)E(b)∗x − aE(b)∗E(x) is linear, and, for a, b, a′, b′ ∈ T , we have Da,bDa′,b′ = DE(a)E(b)∗ a′,b′. Let D = D(T S) = lin {Da,b : a, b ∈ T }. Since E is contractive, we have kDa,bk ≤ 2 kakkbk. So the algebra D consists of bounded linear operators. Recall that T = ker E ⊕ S and that the elements of the kernel of E are of the form of a − E(a). Therefore for a, b, x ∈ T we have Da,b(x) = E(a)E(b)∗x − E(a)E(b)∗E(x) + E(a)E(b)∗E(x) − aE(b)∗E(x) = E(a)E(b)∗(x − E(x)) + (E(a) − a)E(b)∗E(x) = LE(a)E(b)∗ (x − E(x)) + L(E(a)−a)E(b)∗ E(x), TERNARY CORNERS 9 in which Ly is left multiplication by y. This implies that Da,b =(cid:18) LE(a)E(b)∗ L(E(a)−a)E(b)∗ 0 (cid:19) 0 with respect to the above decomposition of T . Setting δ : T ×T → D with δ(a, b) = Da,b we obtain mapping which is linear in the first variable and conjugate linear in the second variable. Furthermore, δ(K, T )D = {0} and δ(T , K)D = {0}. The conditional expectation corresponding to a ternary corner is not necessarily unique; nevertheless, in the next proposition, we will see that the definition of D(T S) is independent of the corresponding conditional expectation E. Proposition 3.1. Let T be a TRO and let S ⊆ T be a sub-TRO. Suppose that E1 and E2 are two TRO-conditional expectations corresponding to S. Let I be the identity map on T and G = E2 + I − E1. Then (i) G is invertible and G−1 = E1 + I − E2. Moreover we have E1G = E2, E2G−1 = E1, GE1 = E1, GE2 = E2 and (I − E2)(I − E1) = I − E2, (I − E1)(I − E2) = I − E1. (ii) Let D1 = D(T S; E1) and D2 = D(T S; E2). For a, b, x ∈ T , we set and Fa,b(x) = E1(a)E1(b)∗x − aE1(b)∗E1(x) Sa,b(x) = E2(a)E2(b)∗x − aE2(b)∗E2(x). Then Fa,b ∈ D1, Sa,b ∈ D2, G−1Fa,bG = SG−1(a),G−1(b) and the mapping θ : Fa,b → G−1Fa,bG is a bijection and consequently the map Θ : nXi=1 λiFai,bi → nXi=1 λiG−1Fai,biG is an algebra isomorphism from D1 onto D2. Proof. Part (i) is easy to prove just by direct calculations. So we proceed to (ii). Let a, b, x ∈ T . Then Fa,b(G(x)) = E1(a)E1(b)∗G(x) − aE1(b)∗E1(G(x)) = E1(a)E1(b)∗(E2(x) + x − E1(x)) − aE1(b)∗E2(x). Since E1Fa,b = 0 = E2Sa,b, we have G−1Fa,b(G(x)) = (Fa,b(G(x)) − E2Fa,b(G(x))) = E1(a)E1(b)∗(E2(x) + x − E1(x)) − aE1(b)∗E2(x) − E1(a)E1(b)∗(2E2(x) − E1(x)) + E2(a)E1(b)∗E2(x) = E1(a)E1(b)∗x − (E1(a) + a − E2(a))E1(b)∗E2(x) = E1(a)E1(b)∗x − G−1(a)E1(b)∗E2(x). 10 Y. ESTAREMI AND M. MATHIEU On the other hand, we have G−1E2 = E1. Therefore G−1Fa,b(G(x)) = E2(G−1(a))E2(G−1(b))∗x−G−1(a)E2(G−1(b))∗E2(x) = SG−1(a),G−1(b)(x). Since G is invertible and for a, b, c, d, Fa,bFc,d = FE1(a)E1(a)∗c,d, Θ is an algebraic isomor- phism. (cid:3) Now we can prove the following lemma. Lemma 3.2. With the above notation, Kl + Kr = lin {Da,b(x) : a, b, x ∈ T }. Proof. For a, b, x ∈ T we have Da,b(x) = E(a)E(b)∗x − aE(b)∗E(x) = E(a)E(b)∗x − E(E(a)E(b)∗x) + E(aE(b)∗E(x)) − aE(b)∗E(x) = lim α (cαE(a)E(b)∗x − E(E(a)E(b)∗x) + lim γ (E(aE(b)∗E(x)) − aE(b)∗E(x)dγ). Hence Da,b(x) ∈ Kl + Kr and so lin {Da,b(x) : a, b, x ∈ T } ⊆ Kl + Kr. Conversely, if y ∈ Kl, then First we notice that for each cα ∈ LS we can find xαi, yαi, aαi, bαi ∈ S such that y = lim α (cαx − E(x)) = lim α (cαx − cαE(x)). cα = lim (xαiy∗ αi + aαib∗ i αi) = lim i (E(xαi)E(yαi)∗ + E(aαi)E(bαi)∗). It follows that y = lim α lim (Dxαi ,yαi (x) + Daαi ,bαi (x)). i Thus Kl ⊆ lin {Da,b(x) : a, b, x ∈ T }, and similarly Kr ⊆ lin {Da,b(x) : a, b, x ∈ T }. (cid:3) We now define a unique TRO-conditional expectation on a special sub-TRO of T . Lemma 3.3. Let the map P : T → T be defined as P(x) = limα,γ cαxdγ for all x ∈ T . Then P is a TRO-conditional expectation onto B = {y : y = lim α,γ cαxdγ, for some x ∈ T }. Proof. Since all cα's are positive contractions, P is a contraction. Also, for each β ∈ I we have limα cαcβ = cβ and similarly for the approximate unit {dγ} of RS. So for x ∈ T and y, z ∈ B we get that P◦P(x) = P(x), P(xy∗z) = P(x)y∗z, P(zy∗x) = zy∗P(x), and P(yx∗z) = yP(x)∗z and thus P is a TRO-conditional expectation. (cid:3) In the next proposition we give a criterion entailing that there is a unique TRO- conditional expectation E onto S. Proposition 3.4. If D = {0}, then the form of the TRO-conditional expectation E necessarily is E(x) = limα,γ cαxdγ for all x ∈ T . That is, (3.1) S = {y : y = lim α,γ cαxdγ, for some x ∈ T }. TERNARY CORNERS 11 Proof. Let D = {0}. Then for all a, b, x ∈ T , E(a)E(b)∗x = aE(b)∗E(x). So, if we take a, b ∈ S, then we have ab∗x = ab∗E(x) and therefore we have cαx = cαE(x). Hence E(x) = limα cαx. On the other hand, if we take x, b ∈ S, then for all a ∈ T we have E(a)b∗x = ab∗x and therefore E(a) = limγ adγ. Consequently we get that E(x) = limα,γ cαxdγ for all x ∈ T . (cid:3) Using the results from the previous section, this situation can be characterized as follows. Theorem 3.5. Let S ⊆ T be a ternary corner and E be the corresponding TRO- conditional expectation. Then the following seven conditions are equivalent: (a) For every x ∈ T , limα cαx = limγ xdγ and S is of the form in (3.1); (b) D = {0}; (c) Da,bS = 0 for all a, b ∈ T ; (d) S is a TRO ideal in T ; (e) S = S1; (f) T = S0. Proof. By Proposition 3.4, (b) ⇒ (a). Conversely suppose that (a) holds. For y ∈ S written as y = limα,γ cαxdγ for some x ∈ T we have y = E(y) = E(lim α,γ cαxdγ) = lim α,γ E(cαxdγ) = lim α,γ cαE(x)dγ = E(x), where in the penultimate step we used the identities (2.3). Therefore, for each x ∈ T , E(x) = limα,γ cαxdγ. Using the second assumption in part (a), limα cαx = limγ xdγ for every x ∈ T , we obtain, for all x, y, z ∈ T , = lim = lim Dy,z(x) = lim α,γ cαDy,z(x)dγ α,γ(cid:0)cαE(y)E(z)∗xdγ − cαyE(z)∗E(x)dγ(cid:1) α,γ(cid:0)E(y)dγE(z)∗cαx − ydγE(z)∗cαE(x)(cid:1) α,γ(cid:0)E(y)E(z)∗cαxdγ − cαydγE(z)∗E(x)(cid:1) = E(y)E(z)∗E(x) − E(y)E(z)∗E(x) = 0. = lim Hence we get the implication (a) ⇒ (b). Let a, b ∈ T and s ∈ S. Then Da,b(s) = E(a)E(b)∗s − aE(b)∗s = (E(a) − a)E(b)∗s, hence {Da,b(s) : a, b ∈ T , s ∈ S} = KS ∗S. If x, y ∈ S and z ∈ K, then y = limγ ydγ and so zy∗x = limγ zdγy∗x. Thus KS ∗S = KrS ∗S and therefore Kr = {0} if and only if Da,bS = {0} for all a, b ∈ T . Finally by Corollary 2.13 and Lemma 3.2 we get that D = {0} if and only if Da,bS = {0} for all a, b ∈ T . So (b) and (c) are equivalent. 12 Y. ESTAREMI AND M. MATHIEU By Proposition 2.4 and the proof of the last part we get the implications (d) ⇒ (c) and (d) ⇒ (e). Also, we have SS ∗K = {0} if and only if Kl = {0}. Thus, by Proposition 2.4, we have S is right TRO ideal if and only if Kl = {0} and is a left TRO ideal if and only if Kr = {0}. By these observations we find that S is a TRO ideal if and only if D = {0}. So (d) and (c) are equivalent. By Proposition 2.5, we have E(S0) = S1 and also we have E(T ) = S. Therefore (e) and (f) are equivalent. Now we prove the implication (f) ⇒ (c). Let T = S0. Then, for every a, b, x ∈ T , we have ab∗E(x) ∈ S, i.e., E(ab∗E(x)) = ab∗E(x). Hence for each s ∈ S we conclude that Da,b(s) = E(a)E(b)∗s − aE(b)∗s = E(aE(b)∗s) − aE(b)∗s = aE(b)∗s − aE(b)∗s = 0. Hence (f) implies (c). Therefore the proof is complete. (cid:3) Our next result contains a necessary condition under which S0 is a TRO-ideal in T and is the analogue of Theorem 2.7 in [7]. Theorem 3.6. Let S ⊆ T be a ternary corner and E be the corresponding TRO- conditional expectation. Then the following hold: Therefore S0 ⊆Ta,b∈T ker Da,b. Conversely, let x ∈ T such that Da,b(x) = 0 for all a, b ∈ T . Then we have E(a)E(b)∗x = aE(b)∗E(x). From abx = abE(x) for all a, b ∈ S we have cαx = cαE(x) for each α ∈ I and thus E(x) = limα cαx and E(a)E(b)∗E(x) = E(a)E(b)∗x = aE(b)∗E(x). Now let y, z ∈ T and s, s′ ∈ S. Then yz∗s′s∗E(x) = E(yz∗s′)s∗E(x) ∈ S. Thus by taking limits on linear combinations of elements from SS ∗ converging to cα we get that yz∗E(x) ∈ S and so E(x) ∈ S1, equivalently x ∈ S0. (ii) We know that S0 is a TRO left ideal. To prove that it is a TRO right ideal let x ∈ S0, s, s′ ∈ S and y, z ∈ T . Then we have E(xs∗s′y∗z) = lim α cαE(xs∗s′y∗z) = E(E(x)s∗s′y∗z) = E(x)s∗E(s′y∗z). Since E(x) ∈ S1 and S1 is a TRO right ideal in S, we find that E(xs∗s′y∗z) = E(x)s∗E(s′y∗z) ∈ S1. (i) S0 =Ta,b∈T ker Da,b. (ii) If limα cαx = limγ xdγ, then S0 is a TRO ideal in T . Proof. (i) Let x ∈ S0 and a, b ∈ T . By definition we have ab∗E(x) ∈ S and therefore E(aE(b)∗E(x)) = aE(b)∗E(x). Moreover, E(x) = limα cαE(x) and E(b) = limα cαE(b). Hence, Da,b(x) = E(a)E(b)∗x − aE(b)∗E(x) = E(a)E(b)∗ lim α cαx − aE(b)∗E(x) = E(a)E(b)∗E(x)x − aE(b)∗E(x) = E(aE(b)∗E(x)) − aE(b)∗E(x) = 0. TERNARY CORNERS 13 By taking limits on linear combinations of elements of S ∗S that converge to dγ we get that E(xy∗z) ∈ S1 and consequently xy∗z ∈ S0. (cid:3) We recall the TRO quotient structure that we need in the sequel. Let I ⊆ T be a (closed) TRO ideal. Then the quotient operator space T /I = {x = x + I : x ∈ T } is a TRO with the ternary product xy∗z = [xy∗z and x∗ = x∗, for all x, y, z ∈ T . (Here we use that every operator space can be completely isometrically embedded into the bounded linear operators on some Hilbert space and that there is a unique TRO structure on the above quotient, by [1, Corollary 4.4.6].) Given TROs T1 and T2, a linear map Φ : T1 → T2 is called a TRO-homomorphism if Φ(xy∗z) = Φ(x)Φ(y)∗Φ(z) (x, y, z ∈ T1). Let T be another TRO and Φ : T → T1 be a TRO-homomorphism onto T1. If we set B = Φ(S), then B is sub-TRO of T1. Moreover, if E is the corresponding conditional expectation to S and E(ker Φ) ⊆ ker Φ, we define the map E1 : T1 → B as E1(Φ(x)) = Φ(E(x)) and obtain a well-defined TRO-conditional expectation onto B. It is clear that Φ(Da,b(x)) = DΦ(a),Φ(b)(Φ(x)). This guarantees that the map with Θ(Pn i=1 λiDai,bi) =Pn Θ : D(T S) → D(T1B) since for every a, b, c, d, x ∈ T , Da,bDc,d = DE(a)E(b)∗ c,d, we have i=1 λiDΦ(ai),Φ(bi), is well defined, linear and surjective. Also, Φ(Da,bDc,d(x)) = DΦ(a),Φ(b)DΦ(c),Φ(d)(Φ(x)). Θ(Da,bDc,d) = Θ(Da,b)Θ(Dc,d) Hence and therefore Θ(( nXi=1 λiDai,bi)( nXi=1 βjDcj ,dj )) = Θ( nXi=1 λiDai,bi)Θ( nXi=1 βjDcj ,dj ). Now let T1 = T /S0 and B = (S + S0)/S0. Moreover, suppose that Φ is the canonical quotient map, i.e., Φ(a) = a = a + S0 for every a ∈ T . If we define E : T /S0 → (S + S0)/S0 by E(a) = dE(a), then E is a TRO-conditional expectation onto (S + S0)/S0. In the next theorem we prove that (S + S0)/S0 is type-zero in T /S0. This, once again, is analogous to the C*-situation [7, Proposition 2.10]. Theorem 3.7. Let S ⊆ T be TROs and suppose that limα cαx = limγ xdγ for all x ∈ T . Also let E : T /S0 → (S + S0)/S0 be defined by E(a) = dE(a). Then the following statements hold: (i) E is a TRO-conditional expectation onto (S + S0)/S0. 14 Y. ESTAREMI AND M. MATHIEU (ii) (S + S0)/S0 is a type-zero sub-TRO in T /S0. (iii) The map Θ : D = D(T S) −→ D(T /S0 (S + S0)/S0) = bD with nXi=1 λiD ai, bi Θ( λiDai,bi) = nXi=1 is an algebraic isomorphism. Proof. Statements (i) and (ii) are taken care of by the above remarks. In order to prove (iii), let x ∈ B0. Then \Da,b(x) = Da,b(x) = 0 for all a, b ∈ T . This means that Da,b(x) ∈ S0 = S1 ⊕ Tl0. If Da,b(x) ∈ S1, then Da,b(x) = E(Da,b(x)) = 0. On the other hand, if Da,b(x) ∈ Tl0, then limα cαDa,b(x) = 0. So for s, s′ ∈ S, Ds′,s(x) = s′s∗(x − E(x)) ∈ Tl0 and therefore cα(x − E(x)) ∈ Tl0. Thus limα cα(x − E(x)) = 0, i.e., limα cαx = E(x). This implies that Da,b(x) = E(a)E(b)∗x − aE(b)∗(lim α cαx) = E(a)E(b)∗x − aE(b)∗x. Now, if a ∈ K, then lim α cαDa.b(x) = Da,b(x) = lim α (−cαE(b)∗x) and Da,b(x) = 0, because Da,b(x) ∈ Tl0. On the other hand, if a ∈ S, then Da,b(x) = aE(b)∗x − aE(b)∗x = 0. Consequently, Da,b(x) = 0 for all a, b ∈ T and by Theorem 3.6 we have x ∈ S0. Therefore x = 0. This means that B is type-zero. In addition, the map Θ is injective. To see this let a, b, c, d ∈ T and DΦ(a),Φ(b) = DΦ(c),Φ(d). Consequently, for all x ∈ T , we have Da,b(x) − Dc,d(x) ∈ S0 and so by our assumptions we get that Da,b(x) − Dc,d(x) = lim α cα(Da,b(x) − Dc,d(x)) = lim γ (Da,b(x) − Dc,d(x))dγ. By Proposition 2.10 we conclude that, if Da,b(x) − Dc,d(x) ∈ S1, then Da,b(x) − Dc,d(x) = E(Da,b(x) − Dc,d(x)) = 0. Also, if Da,b(x) − Dc,d(x) ∈ Kl0, then Da,b(x) − Dc,d(x) = lim α cα(Da,b(x) − Dc,d(x)) = 0. Therefore Da,b(x) = Dc,d(x) and Θ is injective. (cid:3) The map ¯Θ : D → bD with ¯Θ(T ) = Θ(T ) for T ∈ D and ¯Θ(T ) = limα Θ(Tα) for T ∈ D \ D and T = limα Tα, Tα ∈ D, is also an isomorphism between two Banach spaces. We conclude this section with two remarks on the invertibility of the operator Da,b. Lemma 3.8. Let λ 6= 0 and a, b ∈ T . Then Da,b − λI is injective on T if and only if LE(a)E(b)∗ − λI is injective on K. Proof. Let k ∈ K. Then Da,b(k) = E(a)E(b)∗k = LE(a)E(b)∗ k. This implies the necessity of the condition. Conversely, suppose that LE(a)E(b)∗ − λI is injective on K and Da,b(x) − TERNARY CORNERS 15 λx = 0. Then by applying E to this equation we get that λE(x) = 0, so x ∈ K. Therefore LE(a)E(b)∗ x − λx = 0 and so x = 0. (cid:3) Proposition 3.9. For λ 6= 0 and a, b ∈ T , the map Da,b − λI is invertible in B(T ) if and only if LE(a)E(b)∗ − λI is invertible in B(K). Proof. If LE(a)E(b)∗ −λI is invertible, then by Lemma 3.8 we get that Da,b −λI is injective in B(T ). Now we prove that Da,b−λI is surjective in B(T ). For x ∈ T we have x = s+k, with s ∈ S and k ∈ K. Since LE(a)E(b)∗ − λI is surjective, then Da,b = LE(a)E(b)∗ on K, so we have Da,bh − λh = k and Da,bh′ − λh′ = E(a)E(b)∗s − aE(b)∗s for some h, h′ ∈ K. This implies that (Da,b − λI)λ−1h′ = λ−1(E(a)E(b)∗s − aE(b)∗s). As a result (Da,b − λI)(λ−1(h′ − s)) = s and Da,b − λI is surjective. Conversely, suppose that Da,b − λI is invertible. Then LE(a)E(b)∗ − λI is injective. Since Da,b = LE(a)E(b)∗ on K, we get that for each k ∈ K, we can find x ∈ T such that Da,bx − λx = k. Upon applying E on this equation we find that x ∈ K and so LE(a)E(b)∗ − λI is surjective. (cid:3) The proof of the last result is almost identical to the one of the corresponding result in [7], Lemma 3.12. 4. Partial isometries and related ternary corners In this section we focus on those TROs that contain partial isometries (tripotents). Let e ∈ T be a partial isometry, that is ee∗e = e, such that ee∗x = x for all x ∈ T . Then ee∗xx∗ = xx∗ and thus, for every x ∈ LT , we have ee∗x = x. Similarly, xe∗e = x for all x ∈ T implies that xe∗e = x for all x ∈ RT . For the partial isometry e, we set Te = {x ∈ T : ee∗x = xe∗e = x} = {ee∗xe∗e : x ∈ T }. It is easy to see that Te is a norm closed subspace of T . Also, Te is a sub-TRO of T . Similarly, Te is a W*-sub-TRO of T provided T is a W*-TRO. There are lots of partial isometries in TROs, particularly in W*-TROs. To see this and for more details about partial isometries in TROs we refer to [8, 9, 15]. If S ⊆ T is a sub-TRO, a ∈ S is a partial isometry and {cα}α ⊆ S is an approximate unit such that aa∗s = sa∗a for s ∈ S, then limα cαx = aa∗x for all x ∈ T . This means that we can do the same for partial isometries that we did in the last section for approximate units. In the next lemma we obtain a unique TRO-conditional expectation corresponding to a partial isometry a ∈ T onto Ta. Lemma 4.1. Let a ∈ T be a partial isometry and Ψ : T → T be an map such that Ψ(x) = aa∗xa∗a, for every x ∈ T . Then Ψ is a TRO-conditional expectation onto the sub-TRO Ta ⊆ T . Proof. It is clear that Ψ is linear. Let x ∈ T . Since aa∗a = a, we obtain Ψ(Ψ(x)) = aa∗Ψ(x)a∗a = aa∗aa∗xa∗aa∗a = aa∗xa∗a = Ψ(x). 16 Y. ESTAREMI AND M. MATHIEU Hence Ψ ◦ Ψ = Ψ. Also, by a direct computation we get that for every x, y ∈ aa∗T a∗a and z ∈ T the following hold: Ψ(xy∗z) = xy∗Ψ(z), Ψ(xz∗y) = xΨ(z)∗y, Ψ(zy∗x) = Ψ(z)y∗x. So Ψ is a TRO-conditional expectation on T . (cid:3) Some sufficient and necessary conditions for uniqueness of a TRO-conditional expec- tation onto a sub-TRO involving the operators of the previous section are contained in the following result. Proposition 4.2. Let S ⊆ T be a ternary corner and let E be the corresponding TRO- conditional expectation. Let e ∈ S be a partial isometry such that ee∗x = x and xe∗e = x for all x ∈ S. If D = {0}, then S = Te and E(x) = ee∗xe∗e for all x ∈ T . In addition, if ee∗x = xe∗e and E(x) = ee∗xe∗e for all x ∈ T , then D = {0}. Proof. Let D = {0} and a, b, x ∈ T . Then E(a)E(b)∗x = aE(b)∗E(x) and so for a = b = e we have E(x) = ee∗x. Moreover, for b = x = e we get that E(a) = ae∗e. Consequently, for all x ∈ T , E(x) = ee∗xe∗e. Now let ee∗x = xe∗e and E(x) = ee∗xe∗e for all x ∈ T . For all a, b, x ∈ T , we have Da,b(x) = E(a)E(b)∗x − aE(b)∗E(x) = ee∗ae∗eE(b)∗x − aE(b)∗ee∗xe∗e = aE(b)∗x − aE(b)∗x = 0. Thus Da,b = 0 and therefore D = {0}. (cid:3) We record some further relations between K and D. Lemma 4.3. Let S ⊆ T be a ternary corner with the corresponding TRO-conditional expectation E . If there is a partial isometry e ∈ S such that ee∗x = x and xe∗e = x for all x ∈ S and also Tel = {x ∈ T : x = y − ee∗y for some y ∈ T } and Tel,er = {x ∈ T : x = y − ye∗e for some y ∈ Tel}, then (a) K = Tel,er ⊕ (Ke∗e + ee∗K), (b) Ke∗e ∩ ee∗K = ee∗Ke∗e, (c) K = ee∗Ke∗e if and only if ee∗x = x and xe∗e = x for all x ∈ T , (d) Ke∗e + ee∗K = alg {Da,b(x) : a, b, x ∈ T } = D. Proof. (a) It is clear that Tel,er ⊆ T , Ke∗e ⊆ T and ee∗K ⊆ T . Since for every x ∈ Tel,er, ee∗x = 0 = xe∗e, we get that Tel,er ∩ (Ke∗e + ee∗K) = {0}. Moreover, for each x ∈ T , we can write x = ee∗x + (x − ee∗x)e∗e + x − ee∗x − (x − ee∗x)e∗e TERNARY CORNERS 17 and so x − E(x) = x − ee∗x − (x − ee∗x)e∗e + ee∗(x − E(x)) + (x − ee∗x)e∗e. This implies that K ⊆ Tel,er ⊕ (Ke∗e + ee∗K) and so K = Tel,er ⊕ (Ke∗e + ee∗K). (b) Follows by direct computation. (c) If ee∗x = x and xe∗e = x for all x ∈ T , then Tel,er = {0} and ee∗K = Ke∗e = ee∗Ke∗e. Hence, by (a), we have K = ee∗Ke∗e. Conversely, if K = ee∗Ke∗e, then ee∗K = Ke∗e = K and so for each x ∈ T , 0 = ee∗(x−ee∗x) = x−ee∗x and 0 = (x−xe∗e)e∗e = x−xe∗e. As a result, we have ee∗x = xe∗e = x for all x ∈ T . (d) Let a, b ∈ T . Then De,e(x) = ee∗x − ee∗E(x) = ee∗(x − E(x)) and Dx,e(e) = E(x)e∗e − xe∗e = (E(x) − x)e∗e. Thus Ke∗e + ee∗K ⊆ alg {Da,b(x) : a, b, x ∈ T }. Conversely, let a, b, x ∈ T . Then Da,b(x) = E(a)E(b)∗x − E(a)E(b)∗E(x) + E(a)E(b)∗E(x) − aE(b)∗E(x) = ee∗E(a)E(b)∗(x − E(x)) + (E(a) − a)E(b)∗E(x)e∗e. This means that so the proof is complete. D ⊆ Ke∗e + ee∗K, (cid:3) Since ee∗Tel,er = Tel,ere∗e = {0}, by part (a) of Lemma 4.3, we deduce the next corollary. Corollary 4.4. Let S ⊆ T be a ternary corner with the corresponding TRO-conditional expectation E . The following conditions are equivalent: (i) Ke∗e + ee∗K = {0}; (ii) Ke∗e = {0}; (iii) ee∗K = {0}. In the case that there is an element e ∈ S such that ee∗x = xe∗e = x for all x ∈ S, we have since ee∗x = limα ee∗cαx = limα cαx. S0 = {x ∈ T : T T ∗ee∗x ⊆ S}, Proposition 4.5. Let S ⊆ T be a ternary corner with the corresponding TRO-conditional expectation E . Suppose there is a partial isometry e ∈ S such that ee∗x = x and xe∗e = x for all x ∈ S. Then the following conditions are equivalent: (i) S = ee∗T e∗e, E(x) = ee∗xe∗e and ee∗x = xe∗e for all x ∈ T ; (ii) D = {0}; (iii) Da,bS = 0, for all a, b ∈ T ; 18 Y. ESTAREMI AND M. MATHIEU (iv) S is a TRO-ideal of T ; (v) S = S1; (vi) T = S0. Proof. By Proposition 4.2 we have that (i) ⇐⇒ (ii). For every s ∈ S and a, b ∈ T we have Da,b(s) = E(a)E(b)∗s − aE(b)∗s = (E(a) − a)E(b)∗s. It follows that {Da,b(s) : a, b ∈ T , s ∈ S} = KS ∗S. This implies that Da,bS = 0, for all a, b ∈ T if and only if Ke∗e = {0}. Hence, by Corollary 4.4 and Lemma 4.3, we obtain that (ii) and (iii) are equivalent. Since limα cαx = ee∗x for all x ∈ T , by Proposition 2.10 we get that S0 = S1 ⊕ Tel. This implies that (v) and (vi) are equivalent. Also, we obtain that E(x) = ee∗x = ee∗xe∗e for all x ∈ S0. Hence if T = S0, then S = Te and it is a TRO-ideal. Moreover, if S is a TRO- ideal, then S = S1. These observations show that the implications (vi) ⇒ (iv) ⇒ (v) hold. This completes the proof. (cid:3) Theorem 4.6. Under the assumptions of Lemma 4.3 we have the following statements. (a) S0 =Ta,b∈T ker Da,b = S1 + Tel. (b) If ee∗x = xe∗e for all x ∈ T , then S0 is a TRO-ideal in T and so S + S0 is a sub-TRO of T and (S + S0)/S0 is a sub-TRO of T /S0. Proof. (a) Let x ∈ S0. Then E(x) = ee∗x and so for all a, b ∈ T we have Da,b(x) = E(a)E(b)∗x − aE(b)∗ee∗x = E(a)E(b)∗ee∗x − aE(b)∗x. Also, since x ∈ S0, for all y, z ∈ T , E(yz∗ee∗x) = yz∗ee∗x and in particular for s ∈ S, E(ys∗x) = E(ys∗ee∗x) = ys∗ee∗x = ys∗x. Therefore, Da,b(x) = E(a)E(b)∗ee∗x − aE(b)∗x = E(a)E(b)∗ee∗x − E(a)E(b)∗ee∗x = 0. This implies that S0 ⊆Ta,b∈T ker Da,b. Conversely, if x ∈ Ta,b∈T ker Da,b, then, for all a, b ∈ T , E(a)E(b)∗x = aE(b)∗ee∗x = aE(b)∗x and thus E(x) = ee∗x. Hence aE(b)∗x ∈ S. Therefore, ab∗ee∗x = (ab∗e)E(e)∗x ∈ S and so x ∈ S0. It is clear that S1 + Tel ⊆ S0. Moreover, for every x ∈ S0, we have x = ee∗x + x − ee∗x = E(x) + x − ee∗x ∈ S1 + Tel which proves the statement. (b) Let x ∈ S0 and y, z ∈ T . Since x ∈ S0, E(x) = ee∗x. Therefore E(xy∗z) = ee∗E(xy∗z) = E(ee∗xe∗ey∗z) = xe∗E(ey∗z). TERNARY CORNERS 19 Since e∗E(ey∗z) ∈ S ∗S and S1 is a TRO-right ideal in S, E(xy∗z) = ee∗xe∗E(ey∗z) ∈ S1. Thus xy∗z ∈ S0, that is, S0 is a TRO-right ideal in T . Clearly, S0 is also a TRO-left ideal in T . Consequently, S0 is a TRO-ideal in T . (cid:3) Let T1 be another TRO and Φ : T → T1 be a TRO-homomorphism onto T1. If e ∈ T is a partial isometry, then e1 = Φ(e) ∈ T1 is also a partial isometry. Moreover, if ee∗x = xe∗e = x for all x ∈ S, then e1e∗ 1e1 = b, for all b ∈ B = Φ(S). 1b = be∗ Let a, b, x ∈ T and e ∈ S be a partial isometry such that ee∗s = se∗e = s for all s ∈ S and ee∗x = xe∗e for all x ∈ T . Then, by Theorem 4.6 part (b), S0 is a TRO ideal in T . As we saw in the proof of Theorem 3.7, the map Θ : D(T S) −→ D(T1B) with Θ(Pn i=1 λiDai,bi) =Pn i=1 λiDΦ(ai),Φ(bi), is linear surjective and multiplicative. Let T1 = T /S0 and B = (S + S0)/S0. Moreover, suppose that Φ is the canonical quotient map, i.e., Φ(a) = a = a + S0 for every a ∈ T . If we define E : T /S0 −→ (S + S0)/S0 by E(a) = dE(a), then E is a TRO-conditional expectation onto (S + S0)/S0. Moreover, (S+S0)/S0 is a type-zero sub-TRO in T /S0. To see this, let x ∈ B0 such that Da,b(x) = 0 for all a, b ∈ T . Since \Da,b(x) = Da,b(x) = 0, we have Da,b(x) ∈ S0 for all a, b ∈ T . By Theorem 4.6, we get that Da,b(x) ∈ Tel and so ee∗Da,b(x) = 0. Now let a = b = e. Then De,e(x) = ee∗(x − E(x)) and so De,e(x) = ee∗De,e(x) = 0. Hence we have E(x) = ee∗x and so Da,b(x) = E(a)E(b)∗x − aE(b)∗x. If a ∈ K, then Da,b(x) = ee∗(−aE(b)∗x) and so Da,b(x) = ee∗Da,b(x) = 0, because Da,b(x) ∈ Tel. If a ∈ S, then Da,b(x) = E(a)E(b)∗x − aE(b)∗x = aE(b)∗x − aE(b)∗x = 0. Therefore Da,b(x) = 0 for all a, b ∈ T and so x ∈ S0, i.e., x = 0. Thus B0 = {0} and hence B is type-zero in T /S0. In addition, the map Θ is injective. To see this let a, b, c, d ∈ T and DΦ(a),Φ(b) = DΦ(c),Φ(d). Consequently, for all x ∈ T , we have Da,b(x) − Dc,d(x) ∈ S0 and so by our assumptions we get that Da,b(x) − Dc,d(x) = ee∗(Da,b(x) − Dc,d(x)). By Theorem 4.6 we conclude that if Da,b(x) − Dc,d(x) ∈ S1, then Da,b(x) − Dc,d(x) = E(Da,b(x) − Dc,d(x)) = 0. Also, if Da,b(x) − Dc,d(x) ∈ Tel, then Da,b(x) − Dc,d(x) = ee∗(Da,b(x) − Dc,d(x)) = 0. Therefore Da,b(x) = Dc,d(x) and so there is a one to one correspondence between D and bD. Therefore Θ is injective. By these observations we have the next theorem. 20 Y. ESTAREMI AND M. MATHIEU Theorem 4.7. Let S ⊆ T be TROs and e ∈ S be a partial isometry such that ee∗x = xe∗e and ee∗s = se∗e = s for all x ∈ T and s ∈ S. Let E : T /S0 −→ (S + S0)/S0 (i) E is a TRO-conditional expectation onto (S + S0)/S0. (ii) (S + S0)/S0 is a type-zero sub-TRO in T /S0. defined by E(a) = dE(a). Then the following statements hold: (iii) The map Θ : D = D(T S) → D(T /S0(S + S0)/S0) = bD, with Θ( λiDai,bi) = λiD ai, bi nXi=1 nXi=1 is an algebraic isomorphism. The map ¯Θ : D → bD with ¯Θ(T ) = Θ(T ) for T ∈ D and ¯Θ(T ) = limα Θ(Tα) for T ∈ D \ D and T = limα Tα, Tα ∈ D, is also an isomorphism between two Banach algebras. We finish our discussion with some pertinent examples. Example 4.8. Let M3(C) and ei,j be the matrix unit with 1 in the (i, j) position, 0 elsewhere with 1 ≤ i, j ≤ 3. If we set T = linear span{e1,1, e1,2, e1,3, e2,1, e2,2, e2,3} and then S is a sub-TRO of T and a b 0 0 0 0 0 0 0 S = linear span{e1,1, e1,2} =n    : a ∈ Co, SS ∗ =n  a 0 0 0 0 0 0 0 0   : a, b ∈ Co,   : a, b, c, d ∈ Co. a b 0 c d 0 0 0 0 SS ∗ =n  In this case we see that there are approximate units for LS and RS such that T 6= Tl,r. Moreover, S is a corner of T with K = linear span{e1,3, e2,1, e2,2, e2,3}. Note that there is no partial isometry in S. Example 4.9. Let T = M2,3(C); this is a TRO. If we put S =n(cid:18) a 0 b a′ 0 b′ (cid:19) : a, a′, b, b′ ∈ Co, then S is a sub-TRO of T . Moreover, the element e = (cid:18) 1 0 0 0 0 1 (cid:19) ∈ S is a partial isometry and ee∗x = xe∗e = x for all x ∈ S. Also, the corresponding TRO-conditional expectation is of the form of E(x) = ee∗xe∗e, for all x ∈ T . TERNARY CORNERS 21 Acknowledgement. This paper was written during the first-named author's sabbatical stay at Queen's University Belfast in 2018. References [1] D. P. Blecher and Ch. Le Merdy, Operator algebras and their modules. An operator space approach, London Math. Soc. Monographs 30, Claredon Press, Oxford, 2004. [2] D. P. Blecher and M. Neal, Completely contractive projections on operator algebras, Pacific J. Math. 2 (2016), 289 -- 324. [3] D. Bohle and W. Werner, The universal enveloping ternary ring of operators of a JB*-triple system, Proc. Edinb. Math. Soc. 57 (2014), 347 -- 366. [4] L. J. Bunce, B. Feely and R. M. Timoney, Operator space structure of JC*-triples and TROs, I, Math. Z. 270 (2012), 961 -- 982. [5] L. J. Bunce and R. M. Timoney, On the operator space structure of Hilbert spaces, Bull. London Math. Soc. 43 (2011), 1205 -- 1218. [6] L. J. Bunce and R. M. Timoney, On the universal TRO of a JC*-triple, ideals and tensor products, Q. J. Math. 64 (2013), 327 -- 340. [7] J. Daughtry, A. Lambert and B. Weinstock, Operators on C*-algebras induced by conditional expectations, Rocky Mountain J. Math. 25 (1995), 1243 -- 1275. [8] E. G. Effros, N. Ozawa and Z.-J. Ruan, On injectivity and nuclearity for operator spaces, Duke Math. J. 110 (2001), 489 -- 521. [9] L. A. Harris, A generalization of C*-algebras, Proc. London Math. Soc. 42 (1981), 331 -- 361. [10] M. Kaur and Z.-J. Ruan, Local properties of ternary rings of operators and their linking C*- algebras, J. Funct. Anal. 195 (2002), 262-305. [11] M. Neal and B. Russo, Operator space characterizations of C*-algebras and ternary rings, Pacific J. Math. 209 (2003), 339 -- 364. [12] R. Pluta, Ranges of bimodule projections and conditional expectations, Cambridge scholar pub- lishing, Newcastle upon Tyne, 2013. [13] P. Salmi and A. Skalski, Inclusions of ternary rings of operators and conditional expectations, Math. Proc. Cambridge Phil. Soc. 155 (2013), 475-482. [14] P. Salmi and A. Skalski, Actions of locally compact (quantum) groups on ternary rings of operators, their crossed products, and generalized Poisson boundaries, Kyoto J. Math. 57 (2017), 667-691. [15] H. Zettl, A characterization of ternary rings of operators, Adv. in Math. 48 (1983), 117 -- 143. Department of Mathematics, Payame Noor University (PNU), P. O. Box 19395-3697, Tehran, Iran E-mail address: [email protected] Mathematical Sciences Research Centre, Queen's University Belfast, Belfast BT7 1NN, Northern Ireland E-mail address: [email protected]
1511.02753
2
1511
2016-09-28T11:53:57
Hypercontractivity of heat semigroups on free quantum groups
[ "math.OA", "math.FA", "math.PR" ]
In this paper we study two semigroups of completely positive unital self-adjoint maps on the von Neumann algebras of the free orthogonal quantum group $O_N^+$ and the free permutation quantum group $S_N^+$. We show that these semigroups satisfy ultracontractivity and hypercontractivity estimates. We also give results regarding spectral gap and logarithmic Sobolev inequalities.
math.OA
math
HYPERCONTRACTIVITY OF HEAT SEMIGROUPS ON FREE QUANTUM GROUPS UWE FRANZ, GUIXIANG HONG, FRANC¸ OIS LEMEUX, MICHAEL ULRICH and HAONAN ZHANG Abstract. In this paper we study two semigroups of completely positive uni- tal self-adjoint maps on the von Neumann algebras of the free orthogonal quantum group O+ N . We show that these semigroups satisfy ultracontractivity and hypercontractivity esti- mates. We also give results regarding spectral gap and logarithmic Sobolev inequalities. N and the free permutation quantum group S+ . A O h t a m [ 2 v 3 5 7 2 0 . 1 1 5 1 : v i X r a Subject classification. 46L50, 47A30, 47D03 Keywords. Free quantum group, heat semigroup, hypercontractivity, logarithmic Sobolev inequality INTRODUCTION Since the 70s, when the word "hypercontractivity" was coined (see [18]), it has yielded a fruitful area of Mathematics. Stronger than the classical notion of con- tractivity, it has been shown that hypercontractivity is strongly linked to a class of inequalities called logarithmic Sobolev inequalities, which in turn have many appli- cations such as in statistical mechanics (see for instance [12] for the investigation of the Ising model based on log-Sobolev inequalities). With the rise of noncom- mutative mathematics, hypercontractivity has also been studied in the context of noncommutative Lp spaces, for instance in [15]. The hypercontractivity for semigroups on some cocommutative compact quan- tum groups such as von Neumann algebras of discrete groups, e.g. free products of Z2, etc., has been recently studied by Junge et al., see [13] and the references therein. The goal of this paper is to investigate hypercontractivity for semigroups on the free orthogonal quantum group and the free permutation quantum group. Differ- ent definitions for a Brownian motion (and hence for a heat semigroup) could be considered on these quantum groups; we will be interested in the ad-invariant gen- erating functionals in order to select semigroups that could pretend to the role of heat semigroups. This paper is only a short introduction to this topic and it is the authors' hope that much more work will be done in this direction. 1. COMPACT QUANTUM GROUPS AND HEAT SEMIGROUPS 1.1. Compact quantum group: definition. Compact quantum groups are a generalization of compact groups in the context of noncommutative mathematics. They are defined in the following way: 1 2 U. FRANZ, G. HONG, F. LEMEUX, M. ULRICH, H. ZHANG Definition 1.1. A compact quantum group is a pair G = (A, ∆) such that A is a unital C∗-algebra and ∆ : G → G ⊗ G is a comultiplication, i.e. it is a unital ∗-algebra homomorphism and it verifies: and, moreover, the quantum cancellation properties are verified, i.e. (∆ ⊗ id) ◦ ∆ = (id ⊗ ∆) ◦ ∆ Lin[(1 ⊗ G)∆(G)] = Lin[(G ⊗ 1)∆(G)] = G ⊗ G where Lin is the norm-closure of the linear span. The C∗-algebra A is also noted C(G). It is indeed a generalization, because for any compact group G, (C(G), ∆G) with the comultiplication arising from the group multiplication: ∆G : C(G) → C(G × G) ≃ C(G) ⊗ C(G) f 7→ ((x, y) 7→ f (x.y)) is a compact quantum group. The relevant examples for this article were defined by Wang, see [9, 22, 23]: Example 1.2 (Free Orthogonal Quantum Group, see [22]). Let N ≥ 2 and Cu(O+ N ) be the universal unital C∗-algebra generated by the N 2 self-adjoint elements uij, 1 ≤ i, j ≤ N verifying the relations: Xk ukiukj = δij =Xk We define a comultiplication ∆ by setting ∆(vij ) =Pk vik ⊗ vkj. Then (C(O+ is a compact quantum group called the Free Orthogonal Quantum Group. impose in addition commutativity, we recover the classical orthogonal group. Example 1.3 (Free Permutation Quantum Group, see [23]). Let N ≥ 2 and C(S+ N ) be the universal unital C∗-algebra generated by N 2 elements uij, 1 ≤ i, j ≤ N such that for all 1 ≤ i, j ≤ N : N ), ∆) If we uikujk u2 ij = uij = u∗ij Xk uik = 1 =Xk ukj We define a comultiplication ∆ by setting ∆(uij ) =Pk uik⊗ukj. Then, (C(S+ N ), ∆) is a compact quantum group called the Free Permutation Quantum Group. If we impose in addition commutativity, we find the classical permutation group. For G = O+ N , S+ uij, 1 ≤ i, j ≤ N and contained in C(G). It has a bialgebra structure by setting: N , we denote by Pol(G) the ∗-algebra generated by the generators ǫ(uij) = δij It is called the algebra of polynomials of G. Moreover, every compact quantum group is endowed with a Haar state, i.e. a normalized positive functional h : C(G) → C such that (h ⊗ id)∆(a) = h(a)1 = (id ⊗ h)∆(a) for each a ∈ G. The Haar state allows us to define the reduced C∗-algebra of a compact quantum group. If G is a compact quantum group, then we have the GNS representation of its Haar state h, i.e. a ∗-homomorphism π : Pol(G) → B(H) with H a Hilbert space and Ω ∈ H a unit vector, such that h(x) =< Ω, π(x)Ω > for all x ∈ Pol(G). The HYPERCONTRACTIVITY ON FREE QUANTUM GROUPS 3 reduced C∗-algebra Cr(G) is the norm completion of π(Pol(G)) in B(H). In this article, we will always consider the reduced C∗-algebra rather than the universal one. The reason for this is that the Haar state is faithful on the reduced C∗- algebra. The faithfulness of h is important to define the Lp spaces, which is done as follows. The space L∞(G) = Cr(G)′′ is the von Neumann algebra generated by Cr(G). We define Lp(G) for 1 ≤ p < ∞ as the completion of L∞(G) for the norm kxkp = [h((x∗x)p/2)]1/p. We recall here that the Haar state is a trace (i.e. h(ab) = h(ba)) whenever the compact quantum group is of Kac type, which is the case for the quantum groups O+ N treated in this paper. See [16] and the references therein for non-tracial Lp-spaces. N and S+ Let us now say a few words about corepresentations, for more details and nota- tions we refer to [8, 10]. A corepresentations of a compact quantum group G is a unitary matrix v ∈ Mk ⊗ G such that (id ⊗ ∆)(v) = v12v23, it is irreducible if the only scalar matrices that commute with v are multiples of the identity matrix. The set of all (equivalence classes of) irreducible corepresentations is denoted Irr(G). In the case of O+ N , the irreducible corepresentations can be indexed by N and we denote by (u(s) ij )1≤i,j≤dim Vs the coefficients of the sth irreducible corepre- sentation, Vs being their linear span. N and S+ 1.2. Markov semigroups. In order to investigate hypercontractivity of heat semi- groups, one must be able to define heat semigroups on the quantum groups at hand. We recall here for clarity's sake a certain number of important results, without proofs. More on this topic might be found in [8]. We can define L´evy processes on quantum groups (Definition 2.4 in [8]). If (jt)t≥0 is such a process, then we can associate to it a Markov semigroup Tt by putting Tt = (id ⊗ φt) ◦ ∆ where φt = Φ ◦ jt is the marginal distribution of jt. The L´evy process (jt)t is also associated to a generator L = dφt (actually, there is a one-to one correspondence between generators and L´evy processes, called the Schoenberg correspondence). dt (cid:12)(cid:12)(cid:12)t=0 It is important to mention the domain of the Markov semigroup. The operator Tt can either be seen as Tt : Cu(G) → Cu(G) or as Tt : Cr(G) → Cr(G). We will in the sequel take the second definition, due to our use of the reduced C∗-algebra. The semigroup is associated to a Markovian generator TL : Pol(G) → Pol(G) which is defined by TL = (id ⊗ L) ◦ ∆ = dTt The two semigroups treated in this paper are KMS-symmetric (even GNS- symmetric, which means that TL and Tt are self-adjoint on L2(G, h)), therefore they extend to σ-weakly continuous semigroups on the von Neumann algebra L∞(G) = Cr(G)′′, see, e.g., [7, Theorem 2.39]. dt (cid:12)(cid:12)t=0. Now, in the classical case, a heat semigroup is the Markov semigroup associated to a Brownian motion, which is a particular kind of L´evy process. So if we had a definition of such a Brownian motion on O+ N , we could define a heat semigroup and this semigroup should be naturally privileged in our study. Unfortunately, to define such an object is not an easy matter. In the classical case, Brownian motions are defined on Lie groups via the Laplace-Beltrami operator. On quantum groups, we do not have a differential structure which would allow us to define a quantum analogue to the Laplace-Beltrami operator. Alternative approaches must thus be found. N or S+ 4 U. FRANZ, G. HONG, F. LEMEUX, M. ULRICH, H. ZHANG One way to do so is to use the notion of gaussianity first introduced by Schurmann (as is done for instance in [19, see especially Section 5.3] to exhibit a Brownian motion on the unitary dual group). This approach nevertheless fails for S+ N , as indicated by [10, Proposition 8.6], since there are no gaussian generators on S+ N . As an alternative, we will be interested in the class of ad-invariant generating functionals (see Section 6 of [8]), i.e. the functionals invariant under the adjoint action. Linear functionals L : Pol(G) → C are ad-invariant iff there exist num- bers (cs)s such that L(u(s) ij ) = csδij for s ∈ Irr(G). They are classified for O+ N in [8, Section 10] and in [10, Section 10.4] for S+ In the classical case of Lie N . groups, [14, Propositions 4.4, 4.5] shows that ad-invariant processes (or, equiva- lently, conjuguate-invariant processes) on compact simple Lie groups have a gen- erator constituted of the Laplace-Beltrami operator plus a part due to the L´evy measure. It therefore seems reasonable to define a Brownian motion from within the class of ad-invariant functionals and this will be the approach which we will use in this paper. 1.3. Heat semigroup on the Free Orthogonal Quantum Group. We will need the definition of Chebyshev polynomials of the second kind. Definition 1.4. The Chebyshev polynomials of the second kind are the polynomials Us given by the relation Us(X) = ⌊s/2⌋ Xp=0 (−1)p(cid:18)s − p p (cid:19)X s−2p They are an orthonormal family for the scalar product defined via the semicircular measure. We recall the following proposition, found in [8, Proposition 10.3], showing that N are classified by pairs (b, ν) where b is a non-negative ad-invariant functionals on O+ real number and ν a finite measure with support on the interval [−N, N ]. Proposition 1.5. The ad-invariant generating functional on Pol(O+ N ) with charac- teristic pair (b, ν) (b ≥ 0 and ν a finite measure on [−N, N ]) acts on the coefficients of unitary irreducible representations of O+ N as: L(u(s) ij ) = δij Us(N ) −bU′s(N ) +Z N −N Us(x) − Us(N ) N − x ν(dx)! for s ∈ N, where Us denotes the sth Chebyshev polynomial of the second kind. The generator of the Markov semigroup, which is defined by: TL = (id⊗ L)◦ ∆, acts as: TL(u(s) ij ) = 1 Us(N ) −bU′s(N ) +Z N −N Us(x) − Us(N ) N − x ν(dx)! u(s) ij The Markov semigroup is given by Tt = exp(tTL). Here we will be interested in in the case b = 1 and ν = 0. Indeed, our formula is similar to Hunt's formula in the case of L´evy processes on Lie groups and it seems natural to take ν = 0, since it seems to play a role analogous to the L´evy measure in Hunt's formula. HYPERCONTRACTIVITY ON FREE QUANTUM GROUPS 5 Let us now investigate further this Markovian semigroup. We have: L(u(s) ij ) = − δij Us(N ) U′s(N ) Therefore, the eigenvalues of TL are given by: λs = − U′s(N ) Us(N ) with eigenspace Vs = span{u(s) ij , 1 ≤ i, j} and multiplicity ms = (dim u(s))2 = Us(N )2 (see [8], section 10). Now, since the leading coefficient of Us is equal to one, we can write these polynomials with the help of their zeros: Us(x) = (x − x1) . . . (x − xs) and therefore: U′s(x) Us(x) = for x ∈ R\{x1, . . . xs}. in this section and also in the next. s Xk=1 1 x − xk The following classical lemma about Chebyshev polynomials will be useful to us Lemma 1.6. The zeros of Us are comprised between −2 and 2. Proof. We will use the fact that the Chebyshev polynomials of the second kind √4 − x2 constitute an orthonormal family with regard to Wigner's semicircle law 1 π on [−2, 2]. Let ∈ N. Let us denote by S = {y1, . . . , yl} the set of all zeros of Us in (−2, 2) that have an odd multiplicity. We set Q =Ql k=1(X − xk). It is obvious that Q divides Us. Let us now assume that deg Q < s = deg Us. Therefore, we have: πp4 − x2dx = 0 But the very definition of Q means that the zeros of UsQ that are in (−2, 2) have an even multiplicity, i.e. UsQ has a constant sign on this interval. For the integral to be zero, we must have UsQ = 0, which is absurd. Therefore we must have Us = Q and this proves the lemma. (cid:3) Q(x)Us(x) Z 2 −2 1 We thus have the following lemma: Lemma 1.7. For N ≥ 2, s Xk=1 1 s s N ≤ −λs = U′s(N ) Us(N ) = N − xk ≤ where, for N = 2, we take the convention that 1/0 = ∞. Proof. The upper bound of −λs is a consequence of the previous lemma. To obtain the lower bound, note that N − xk > 0 for 1 ≤ k ≤ n. Since Pn k=1 xk = 0, we have N − 2 s2 1 s −λs = Xk=1 N − xk ≥ = s N . k=1 xk sN −Ps (cid:3) 6 U. FRANZ, G. HONG, F. LEMEUX, M. ULRICH, H. ZHANG 1.4. Heat semigroups on the Free Permutation Quantum Group. We rely on the results of [10] for S+ N . We consider semigroups with generating functionals defined by: L(u(s) ij ) = − δijU′2s(√N ) 2√N U2s(√N ) We follow the same reasoning as before. The eigenvalues are: U′2s(√N ) 2√N U2s(√N ) λs = − with eigenspace Vs = {u(s) We find the estimate: ij , 1 ≤ i, j ≤ dim Vs} and multiplicity ms = U2s(√N )2. Lemma 1.8. For N ≥ 4, s N ≤ −λs = 1 2√N 2s Xk=1 1 √N − xk ≤ s √N (√N − 2) where, for N = 4, we take the convention that 1/0 = ∞. 2. ULTRACONTRACTIVITY AND HYPERCONTRACTIVITY When we need to distinguish the semigroups, we will denote by T O t (resp. T S t ) the semigroup we introduced on O+ N (resp. S+ N ) 2.1. Ultracontractivity. We say that a semigroup Tt is ultracontractive if it is bounded from L2 into L∞ for all t > 0. In the sequel, we will denote by k.k∞ = k.k p = h((x∗x)p/2) the p-norm (h being the Haar state). the operator norm and by kxkp We will prove the following result: Theorem 2.1. Let Tt be a semigroup on a Kac-type compact quantum group, such that the following assumptions hold: • The subspaces Vs spanned by the coefficients of the irreducible corepresen- tations us are eigenspaces for the generator TL of the Markov semigroup, i.e. TLx = λsx for x ∈ Vs • We have an estimate of the form λs ≤ −αs for some α > 0. • We have an inequality of the form: kxk∞ ≤ (βs + γ)kxk2 for x ∈ Vs, with β, γ ≥ 0 and β, γ are independent of s. Then, Tt is ultracontractive: kTtxk∞ ≤pf (t)kxk2, where: f (t) = β2e−2αt(1 + e−2αt) + 2βγe−2αt(1 − e−2αt) + γ2(1 − e−2αt)2 . (1 − e−2αt)3 HYPERCONTRACTIVITY ON FREE QUANTUM GROUPS 7 Proof. We have for x =Ps xs with xs ∈ Vs: kTtxk∞ ≤ Xs∈N kTtxsk∞ =Xs e−αst(βs + γ)kxsk2 ≤ Xs eλstkxsk∞ ≤Xs ≤ Xs = pf (t)kxk2 e−αstkxsk∞ (βs + γ)2e−2αst!1/2 Xs 2!1/2 kxsk2 where we used the Cauchy-Bunyakovsky-Schwarz inequality. The computation of f (t) =Ps(β2s2 + 2βγs + γ2)e−2αst is done via the classical series: e−λk = 1 Xk∈N Xk∈N Xk∈N ke−λk = k2e−λk = 1 − e−λ e−λ (1 − e−λ)2 e−λ(1 + e−λ) (1 − e−λ)3 (cid:3) Let us mention the following nice consequence Corollary 2.2. We have for any heat semigroup satisfying the assumptions of Theorem 2.1: with f the same function as in Theorem 2.1. kTtxk∞ ≤ f (t/2)kxk1, Let us remark that, when t goes to zero, f (t) is equivalent to 1/t3. On Rd, the behavior when t goes to zero of a heat semigroup is in 1/td/2, as can be seen e.g. in [20, Property Rn, section II.1], so that we have here a behavior as if we were in "dimension" 6. Proof. We are follow the reasoning of [2, Corollary 3]. The semigroup is self-adjoint on L2(G, h), since L(u(s) ij ) = L(u(s) ji ) is real. So we can dualize the inequality of Theorem 2.1 to obtain kTtxk2 ≤pf (t)kxk1. We can then combine it to get: kTtxk∞ ≤pf (t/2)kTt/2xk2 ≤ f (t/2)kxk1 (cid:3) As a consequence of the Theorem, we deduce that the semigroup we considered on the Free Orthogonal Quantum Group is ultracontractive. Indeed, [5, Proof of Theorem 2.2] shows that there exists a constant D (depending on N ) such that: (2.1) when x ∈ Vs. Thus we can apply Theorem 2.1 with α = 1/N and β = γ = D. pending on N ) such that on S+ In the same way, [4, Theorem 4.10] shows that there exists a constant C (de- kxk∞ ≤ D(s + 1)kxk2 (2.2) N and for x ∈ Vs, we have: kxk∞ ≤ C(2s + 1)kxk2 8 U. FRANZ, G. HONG, F. LEMEUX, M. ULRICH, H. ZHANG This means that we can obtain ultracontractivity for our semigroup on S+ applying Theorem 2.1 with α = 1/N , β = 2C and γ = C. N by 2.2. Special cases O+ 2 . We have Us(2) = s + 1 and, differentiating the recurrence relation, we get U′s(2) = s(s + 1)(s + 2)/6. Therefore we know the exact value of the eigenvalues: 4 . We can say more in the case of O+ 2 and S+ If we take up the computations from Theorem 2.1, we get better estimates: λs = − s(s + 2) 6 kTtxk∞ ≤sD2Xs Xs e− s(s+2) 3 t(s + 1)2 ≤Xs e− s(s+2) 3 t(s + 1)2kxk2 e− s2 3 t(s + 1)2 Observe now that: and, moreover, ∞ Xs=0 e− s2 3 t ≤ 1 + This yields the inequality: se− s2 3 t ≤ 1 + ∞ Xs=1 s2e− s2 3 t ∞ Xs=1 3 t . For fixed t, we have: kTtxk∞ ≤pg(t)kxk2 with g(t) = 4D2 is decreasing on [q 3 3 ≤Z s0 3 ≤ The function s 7→ s2e− s2 t Let's set s0 =q 3 Z s0 Z ∞ Xs=1 Xs=s0 We do the change of variable u = spt/3: 3 ds ≤ 3 ds ≤ s2e− s2 t s2e− s2 t s2e− s2 t s2e− s2 t 3 et ∞ s0 s0 s0 0 0 s2e− s2 3 t + D2 ∞ Xs=1 t , +∞[ and increasing on [0,q 3 t ]. s2e− s2 t 3 ds + 3 et +Z ∞ s0 s2e− s2 t 3 ds 0 t(cid:19)3/2Z 1 (cid:18) 3 t(cid:19)3/2Z ∞ (cid:18) 3 1 u2e−u2 du ≤ u2e−u2 du ≤ u2e−u2 du + 3 et s2e− s2 t s2e− s2 t 3 ≤(cid:18) 3 3 ≤ 3 et 0 t(cid:19)3/2Z 1 +(cid:18) 3 t(cid:19)3/2Z ∞ 1 u2e−u2 du And by combining: t(cid:19)3/2Z ∞ (cid:18) 3 0 u2e−u2 du ≤ 3 et + s2e− s2 t 3 ≤ 2 3 et +(cid:18) 3 t(cid:19)3/2Z ∞ 0 u2e−u2 du In other words, when t goes to zero, g(t) behaves like t−3/2, and, in the spirit of the remark following Corollary 2.2, this yields a "dimension" 3 for the semigroup. and the "dimen- The same reasoning for S+ sion" of the semigroup on S+ 4 yields the eigenvalues λs = − s(s+1) 4 is also 3. 6 Xs=1 Xs=s0 ∞ ∞ Xs=0 HYPERCONTRACTIVITY ON FREE QUANTUM GROUPS 9 2.3. Hypercontractivity. Definition 2.3. We say that a semigroup Tt is hypercontractive if for each 2 < p < ∞, there exists a τp > 0 such that for all t ≥ τp we have: (2.3) kTtxkp ≤ kxk2 Let us remark that if the semigroup Tt is hypercontractive, then the inequality (2.3) is also true for 1 ≤ p ≤ 2 because for such a p and for any x ∈ C(G) we have kTtxkp ≤ kxkp ≤ kxk2. We can also notice that due to duality, we have: kTtxk2 ≤ kxkq for t ≥ τp and q such that 1/p + 1/q = 1. Therefore, for t big enough, Tt is also a contraction from Lq to L2 for any 1 < q < 2. Denote by DN and CN the constants from the inequalities (2.1) and (2.2), re- spectively. We know from [5, Proof of Theorem 2.2] and [4, Theorem 4.10] that DN ≥ 1, CN ≥ 1. Theorem 2.4. The semigroup T O t hypercontractive. t ) we consider on O+ N (resp. S+ (resp. T S N ) is Proof. We use the inequality p ≤ kh(x)1k2 kxk2 p + (p − 1)kx − h(x)1k2 p x ∈ L∞(G) for 2 < p < ∞, shown in [17, Theorem 1]. It can indeed be applied in our setting, with L∞(G) = Cr(G)′′ a von Neumann algebra and h a faithful, finite normal trace on it. We will write x = h(x)1 +Ps≥1 xs with xs ∈ Vs. We have h(Tt(x))1 = Tt(h(x)1) because the Vs are eigenspaces for Tt. Therefore, p + (p − 1)kTt(x − h(x)1)k2 kTt(x)k2 for t ≥ τp with τp such that: (p − 1)Xs≥1 (βs + γ)2e2λsτp ≤ 1. (cid:3) 2 2 p p ≤ kTt(h(x)1)k2 ≤ h(x)2 + (p − 1) kTt(xs)kp Xs≥1  eλstkxskp ≤ h(x)2 + (p − 1) Xs≥1  eλstkxsk∞ ≤ h(x)2 + (p − 1) Xs≥1  eλst(βs + γ)kxsk2 ≤ h(x)2 + (p − 1) Xs≥1  ≤ h(x)2 + (p − 1)Xs≥1(cid:0)(βs + γ)eλst(cid:1)2Xs≥1 kxsk2 2 2 2 ≤ kxk2 2 10 U. FRANZ, G. HONG, F. LEMEUX, M. ULRICH, H. ZHANG Proposition 2.5. Hypercontractivity is achieved for T O given by t at least from the time τ (O) p where X is the smallest real positive root of N 2 τ (O) log X, p = − X 3 − 3X 2 + 4X (1 − X)3 Hypercontractivity is achieved for T S t at least from the time τ (S) p where Y is the smallest real positive root of N 2 τ (S) log Y, p = − Y 3 − 2Y 2 + 9Y (1 − Y )3 = 1 (p − 1)D2 N . given by = 1 (p − 1)C2 N . Proof. We use the expression: (βs + γ)2e2λsτp = 1 (p − 1)Xs≥1 drawn from the proof of Theorem 2.4. The precise value of the eigenvalues is too cumbersome to compute, therefore we use a minoration of them: λs ≤ − λs ≤ − s N s N for O+ N , for S+ N . p 2τ (O) 2τ (S) By then setting X = exp(− N ) and using the classical series that were already used in the proof of Theorem 2.1, we obtain the desired equation for X and Y . The fact that the root must be the smallest real one (indeed one can easily check that there exists at least one root between 0 and 1) comes from the fact that we need to take the biggest time τp such that the inequalities N ) and Y = exp(− p X 3 − 3X 2 + 4X (1 − X)3 Y 3 − 2Y 2 + 9Y (1 − Y )3 ≤ ≤ 1 (p − 1)D2 N 1 (p − 1)C2 N , , are verified always for t ≥ τp. But X and Y diminish when the time increases. Therefore we need to choose the smallest positive root. (cid:3) For p ≥ 4 − ε0 with ε0 a nonnegative constant, we can obtain a better estimate of τ (O) p (resp. τ (N ) p ) as the following theorem shows: Theorem 2.6. There exists ε0 ≥ 0 such that for any p ≥ 4 − ε0, kT O t k2→p ≤ 1, for all t ≥ log(p − 1) + (1 − )N log DN , 2 p cN 2 dN 2 kT S t k2→p ≤ 1, with c = 2 log(√3+1) log 3 t ≥ for all log(p − 1) + (1 − ≈ 1.8297 . . ., and d = log(11+√105)−log 2 log 3 2 p )N log CN , ≈ 2.15096 . . .. HYPERCONTRACTIVITY ON FREE QUANTUM GROUPS 11 Proof. We only prove this theorem for O+ the inequality N , the case for S+ N is similar. We use again kxk2 p ≤ kh(x)1k2 p + (p − 1)kx − h(x)1k2 p x ∈ L∞(G) for 2 < p < ∞, shown in [17, Theorem 1]. By the Holder inequality, for p ≥ 1, kxskp ≤ (DN (s + 1))1− 2 p kxsk2, xs ∈ Vs. Therefore, kTt(x)k2 p ≤ h(x)2 + (p − 1) Xs≥1 ≤ h(x)2 + (p − 1) Xs≥1 ≤ h(x)2 + (p − 1)Xs≥1 2 eλstkxskp  eλst (DN (s + 1))1− 2 e2λst (DN (s + 1))2(1− 2 2 p kxsk2  p )Xs≥1 kxsk2 2. When t ≥ cN 2 log(p − 1) +(cid:16)1 − 2 p(cid:17) N log DN and s ≥ 1, we have p(cid:19) s log DN p(cid:19) log DN , 2λst ≤ −cs log(p − 1) − 2(cid:18)1 − ≤ −cs log(p − 1) − 2(cid:18)1 − 2 2 and e2λst ≤ (p − 1)−csD−2(1− 2 any p ≥ 4 − ε0, N p ) . So it suffices to show that for some ε0 ≥ 0, for Rp :=Xs≥1 φs(p) =Xs≥1 (p − 1)1−cs(s + 1)2(1− 2 p ) ≤ 1. An easy computation implies that φ′s(p) ≤ 0 if and only if 4(p − 1) p2 cs − 1 log(s + 1) . ≤ p2 Note that f1(p) = 4(p−1) for s ≥ 1, thus from c = 2 log(√3+1) c − 1 log 2 f1(p) ≤ f1(2) = 1 < log 3 is decreasing for p ≥ 2, and f2(s) = cs−1 log(s+1) is increasing ≈ 1.83 > 1.69 ≈ 1 + log 2 we deduce that = f2(1) ≤ f2(s), for all p ≥ 2, s ≥ 1. Hence each φs is decreasing for p ≥ 2, and Rp is also decreasing for p ≥ 2. Since we have Rp ≤ 1 for all p ≥ 4. So there exists ε0 ≥ 0 such that kT O Remark 2.7. We can see from [5, Proof of Theorem 2.2] that t k2→p ≤ 1. (cid:3) R4 =Xs≥1 s + 1 3cs−1 = 3(2 · 3c − 1) (3c − 1)2 = 1, DN ≤ (1 − q2)−1 r Ys=1(cid:0)1 − q2s(cid:1)−3 , 12 U. FRANZ, G. HONG, F. LEMEUX, M. ULRICH, H. ZHANG lim N log DN = 0, N→+∞ p )N log DN , disappears with r ≥ 1 and N = q + 1 which implies that the latter part of τ O as N → +∞. Indeed, it suffices to show that: log (1 − q2) Ys=1 So it is done when we prove for all s ≥ 1: q , 0 < q < 1. Thus we deduce that p in Theorem 2.6, (1− 2 (1 − q2s)3! = 0. lim q→0 1 q r 1 q lim q→0 log(1 − q2s) = 0. This is clear, since lim q→0 q2s log(1 − q2s) = −1 for all s ≥ 1. We have not been able to proof a similar result for S+ 1 N , since by [4, Theorem 4.10], CN → +∞ as N → +∞. 3. FURTHER PROPERTIES OF THE SEMIGROUPS We will note Pol(G)+ the subset of Pol(G) consisting of all such x such that x = x. 3.1. Spectral gap. Definition 3.1. We say that Tt verifies a spectral gap inequality with constant m > 0 if we have for all x ∈ Pol(G)+: mkx − h (x) k2 2 ≤ −h (xTLx) Proposition 3.2. Our semigroup T O with constant m = 1 N . t on O+ N verifies the spectral gap inequality Proof. The eigenvalues of the generator TL are of the form: s 1 N − xi , U′s (N ) Us (N ) = − Xi=1 λs = − and we have shown that λs ≥ s N for N ≥ 2. Let us now write x =Ps xs. We then get: − h (xTLx) =Xs U′s (N ) Us (N )kxsk2 2 1 Using the fact that the Vs are in orthogonal direct sum, we deduce that: −h (xTLx) ≥ N kxk2 2. But, we also see that kx − h (x) k2 ≤ kxk2 and thus we finally get: kx − h (x) k2 2 ≤ −N h (xTLx) . We can prove the following in the same way: (cid:3) Proposition 3.3. Our semigroup T S with constant m = 1 N . t on S+ N verifies the spectral gap inequality HYPERCONTRACTIVITY ON FREE QUANTUM GROUPS 13 3.2. Logarithmic Sobolev Inequalities. Hypercontractivity is equivalent to Log- arithmic Sobolev inequalities, or, shorter, log-Sobolev inequalities, see, e.g., [11] or [15, Theorem 3.8]. We derive here a log-Sobolev inequality for the generators of our heat semigroups. There is nothing new in this Section, we include it only for comparison. Proposition 3.4. There exists a constant t0 > 0, such that, if we denote q (t) = 4 2−t/t0 , we then have for 0 ≤ t ≤ t0: kT G t : L2 → Lq(t)k ≤ 1 N N or S+ where G = O+ Proof. We take for t0 the optimal time for hypercontractivity Tt : L2 → L4, then we have q(t0) = 4, and T0 = Id : L2 → L2 and Tt0 : L2 → L4 are contractions. The Proposition therefore follows by Stein interpolation. Theorem 3.5. For x ∈ L∞(G)+ ∩ D(T G Proposition 3.4, we have the following inequality: L ) and with the same assumptions as in (cid:3) h(cid:0)x2 log x(cid:1) − kxk2 2 log kxk2 ≤ − c 2 h (xTLx) where c = t0/2. Proof. We define : F (t) = kxtkq(t), where we note xt = Ttx. Because of Proposition 3.4, we know that log F (t) ≤ log F (0). Hence: d dt log F (t) t=0 ≤ 0 As in [15, Lemma 3.7] this term is given by: d dt log kxtkq = d dt(cid:18) 1 q q(cid:19) log kxtkq log kxtkq + qkxtkq qkxtkq log kxtkq + 1 1 q = − = − q q q q q d dtkxtkq q (cid:16)qh(cid:0)xq−1 t TL(xT )(cid:1) + qh(cid:0)xq t log(xt)(cid:1)(cid:17) . From this we obtain the desired inequality, because q (0) = 2, q (0) = 2/t0. (cid:3) 4. CONCLUSION We have studied in this paper two Markov semigroups, one on O+ N and the other on S+ N , which could be candidates for a Brownian motion on these quantum groups. We have shown that these semigroups are hypercontractive and satisfy log-Sobolev inequalities. Several natural questions are: What can be said about other semigroups on O+ N N ? What are the optimal times for hypercontractivity? What happens on or S+ other quantum groups, e.g. SUq(2), which are not Kac-type? Acknowledgements UF, MU and HZ are supported by MAEDI/MENESR and DAAD through the PROCOPE programme, and by MAEDI/MENESR and the Polish MNiSW through the POLONIUM programme. GH is supported by MINECO: ICMAT Severo Ochoa project SEV-2011-0087. 14 U. FRANZ, G. HONG, F. LEMEUX, M. ULRICH, H. ZHANG References [1] Keith Ball, Eric A. Carlen, and Elliott H. Lieb, Sharp uniforme convexity and smoothness inequalities for trace norms, Invent. Math., 115(1994), 463–482. [2] Philippe Biane, Free hypercontractivity, Comm. Math. Phys., 184(1997), 457–474. [3] Thierry Bodineau and Boguslaw Zegarlinski, Hypercontractivity via spectral theory, Infin. Dimens. Anal. Quantum. Probab. Rel. Top., 3(2000), 15–31. [4] Michael Brannan, Reduced operator algebras of trace-preserving quantum automorphism groups, Doc. math., 18(2013), 1349–1402. [5] Michael Brannan, Strong asymptotic freeness for free orthogonal quantum groups, Canadian Math. Bulletin, 57(2014), 708–720. [6] Raffaella Carbone and Emanuela Sasso, Hypercontractivity for a quantum Ornstein- Uhlenbeck semigroup, Probab. Theory Related Fields, 140(2000), 15–31. [7] Fabio Cipriani, Dirichlet forms on noncommutative spaces, In: "Quantum Potential Theory", Uwe Franz and Michael Schurmann (eds), Lecture Notes in Mathematics Vol. 1954, pp. 161– 276, Springer, 2008. [8] Fabio Cipriani, Uwe Franz, and Anna Kula, Symmetries of L´evy processes on compact quan- tum groups, their Markov semigroups and potential theory, J. Funct. Analysis, 266(2014), 2789–2844. [9] Alfons Van Daele and Shuzhou Wang, Universal quantum groups, Int. J. Math., 07(1996), 255–263. [10] Uwe Franz, Anna Kula, and Adam Skalski, L´evy Processes on Qantum Permutation Groups, In: "Noncommutative Analysis, Operator Theory and Applications", Fabio Cipriani, Fabrizio Colombo, Irene Sabadini (eds.), Birkhauser, 2016, see also arXiv:1510.08321. [11] Leonard Gross, Hypercontractivity and logarithmic Sobolev inequalities for the Clifford Dirichlet form, Duke Math. J., 42(1975), 383–396. [12] Richard Holley and Daniel Stroock, Logarithmic Sobolev inequalities and stochastic Ising models, J. Stat. phys., 46(1978), 1159–1194. [13] Marius Junge, Carlos Palazuelos, Javier Parcet, and Mathilde Perrin, Hypercontractivity in group von Neumann algebras, preprint arXiv:1304.5789, 2013. [14] Ming Liao, L´evy processes in Lie groups, volume 162 of Cambridge Tracts in Mathematics, Cambridge University Press, Cambridge 2004. [15] Robert Olkiewicz and Boguslaw Zegarlinski, Hypercontractivity in noncommutative Lp spaces, J. Funct. Analysis, 161(1999), 246–285. [16] Gilles Pisier and Quanhua Xu, Non-commutative Lp-spaces, Handbook of the geometry of Banach spaces, Vol. 2, 14591517, North-Holland, Amsterdam 2003. [17] Eric Ricard and Quanhua Xu, A noncommutative martingale convexity inequality, Ann. Prob. 44(2016), 867-882. [18] B. Simon and R. Hoegh-Krohn, Hypercontractive semi-groups and two dimensional self- coupled Bose fields, J. Funct. Analysis, 9(1972), 121–180. [19] Michael Ulrich, Construction of a free L´evy process as high-dimensional limit of a Brownian motion on the unitary group, Infin. Dim. Anal. Quant. Prob., 18(2015), 1550018. [20] N. Th. Varopoulos, L. Saloff-Coste, and T. Coulhon, Analysis and Geometry on Groups, Cambridge University Press, Cambridge 1992. [21] Roland Vergnioux, The property of rapid decay for discrete quantum groups, J. Operator Theory, 57(2007), 303–324. [22] Shuzhou Wang, Free products of compact quantum groups, Comm. Math. Phys., 167(1995), 671–692. [23] Shuzhou Wang, Quantum symmetry groups of finite spaces, Comm. Math. Phys., 195(1998), 195–211. [24] Haonan Zhang, A Noncommutative Martingale Convexity Inequality and Its Application to Hypercontractivity, Master's Thesis, University of Franche-Comt´e, Besan¸con, 2016. FRANZ, Laboratoire de Math´ematiques de Besanc¸on, Universit´e de Bourgogne-Franche- Comt´e, France E-mail address: [email protected] HYPERCONTRACTIVITY ON FREE QUANTUM GROUPS 15 HONG, School of Mathematics and Statistics, Wuhan University, Wuhan 430072, China E-mail address: [email protected] LEMEUX, Laboratoire de Math´ematiques de Besanc¸on, Universit´e de Bourgogne- Franche-Comt´e, France ULRICH, Laboratoire de Math´ematiques de Besanc¸on, Universit´e de Bourgogne- Franche-Comt´e, France E-mail address: [email protected] ZHANG, Laboratoire de Math´ematiques de Besanc¸on, Universit´e de Bourgogne-Franche- Comt´e, France E-mail address: [email protected]
1304.6540
3
1304
2015-06-01T17:03:17
Crossed products for actions of crossed modules on C*-algebras
[ "math.OA" ]
We decompose the crossed product functor for actions of crossed modules of locally compact groups on C*-algebras into more elementary constructions: taking crossed products by group actions and fibres in C*-algebras over topological spaces. For this, we extend the theory of partial crossed products from groups to crossed modules; extend Takesaki-Takai duality to Abelian crossed modules; show that equivalent crossed modules have equivalent categories of actions on C*-algebras; and show that certain crossed modules are automatically equivalent to Abelian crossed modules.
math.OA
math
CROSSED PRODUCTS FOR ACTIONS OF CROSSED MODULES ON C*-ALGEBRAS ALCIDES BUSS AND RALF MEYER Abstract. We decompose the crossed product functor for actions of crossed modules of locally compact groups on C ∗-algebras into more elementary con- structions: taking crossed products by group actions and fibres in C ∗-alge- bras over topological spaces. For this, we extend Takesaki -- Takai duality to Abelian crossed modules; describe the crossed product for an extension of crossed modules; show that equivalent crossed modules have equivalent cate- gories of actions on C ∗-algebras; and show that certain crossed modules are automatically equivalent to Abelian crossed modules. 1. Introduction A crossed module of locally compact groups C consists of two locally compact groups H and G with a continuous group homomorphism ∂ : H → G and a con- tinuous conjugation action c : G → Aut(H) such that ∂(cg(h)) = g∂(h)g−1, c∂(h)(k) = hkh−1 for all g ∈ G, h, k ∈ H. Strict actions of crossed modules on C∗-algebras and crossed products for such actions are defined in [1]. Here we are going to fac- torise this crossed product functor into more elementary operations, namely, tak- ing crossed products for actions of locally compact groups and taking fibres in C0(X)-C∗-algebras ([15]). We mostly work with the more flexible notion of action by correspondences in- troduced in [2]. By [2, Theorem 5.3], such actions are Morita -- Rieffel equivalent to ordinary strict actions, that is, actions by automorphisms. This requires, however, to stabilise the C∗-algebras involved; and certain induced actions that we need are easier to describe as actions by correspondences. Here we define the 2-category of crossed module actions by correspondences precisely, making explicit some hints in [2]. Then we translate the definition of this 2-category into the language of Fell bundles; this extends results in [2] from locally compact groups to crossed modules. We define saturated Fell bundles over crossed modules and correspon- dences between Fell bundles over crossed modules so that they are equivalent to actions by correspondences and transformations between such actions. Correspon- dences between Fell bundles contain Morita -- Rieffel equivalences of Fell bundles and representations of Fell bundles as special cases. The crossed product for a 2010 Mathematics Subject Classification. 46L55, 18D05. Key words and phrases. C∗-algebra, crossed module, Fell bundle, Takesaki -- Takai duality. Supported by the German Research Foundation (Deutsche Forschungsgemeinschaft (DFG)) grant ME 3248/1 and by CNPq (Ciências sem Fronteira) -- Brazil. 1 2 ALCIDES BUSS AND RALF MEYER crossed module action by correspondences is defined by a universal property for representations. Let Ci = (Gi, Hi, ∂i, ci) be crossed modules of locally compact groups. We call a diagram C1 → C2 → C3 of homomorphisms of crossed modules a (strict) extension if the resulting diagrams G1 → G2 → G3 and H1 → H2 → H3 are extensions of locally compact groups in the usual sense. If C2 acts on a C∗-algebra A, then C1 also acts on A by restriction of the C2-action. We show that A ⋊ C2 ∼= (A ⋊ C1) ⋊ C3 for a certain induced action of C3 by correspondences on A ⋊ C1. If Ci are ordinary groups Gi viewed as crossed modules, then G1 is a closed normal subgroup of G2 with quotient G3 = G2/G1. It is well-known that A ⋊ G1 carries a Green twisted action of (G2, G1) such that (A ⋊ G1) ⋊ (G2, G1) ∼= A ⋊ G2. Our theorem says that such a Green twisted action may be turned into an action of G3 by correspondences with the same crossed product. Our proof, in fact, generalises this idea to cover general extensions of crossed modules. Another special case is [1, Theorem 1], which says that (A⋊H)⋊C ∼= A⋊G if C = (G, H, ∂, c) is a crossed module and A is a G-C∗-algebra. Moreover, we generalise Takesaki -- Takai duality to Abelian crossed modules. We call a crossed module Abelian if the group G is Abelian and the action c is trivial (forcing H to be Abelian). Thus Abelian crossed modules are just contin- uous homomorphisms ∂ : H → G between Abelian locally compact groups. The Pontryagin dual C is the crossed module ∂ : G → H given by the transpose of ∂. Our duality theorem says that actions of C are equivalent to actions of the arrow groupoid H ⋊ G associated to C; this is the transformation groupoid for the action of G on H where g ∈ G acts by right translations by ∂(g). Our duality maps an action of C to its crossed product by G, equipped with the dual action of G and a canonical C0( H)-C∗-algebra structure that comes from the original action of H. The inverse equivalence takes the crossed product by G and extends the dual G-action on it to an action of C, using the original C0( H)- C∗-algebra structure. Thus our duality result merely enriches the usual Takesaki -- Takai duality by translating the action of H for a crossed module action into a C0( H)-C∗-algebra structure on the crossed product, and vice versa. In this setting, the crossed product for crossed module actions is equivalent to the functor of taking the fibre at 1 ∈ H for an action of H ⋊ G. Thus crossed products for Abelian crossed modules may be computed in two steps: first take a crossed product by an action of the Abelian group G, then take a fibre for a C0( H)-C∗-algebra structure. Now we describe our decomposition of crossed products for a general crossed module C = (G, H, ∂, c). Let G1 be the trivial group and H1 := ker ∂. Since ker ∂ is Abelian, the trivial maps ∂1 and c1 provide a crossed module C1 = (G1, H1, ∂1, c1). This fits into an extension C1 ֌ C ։ C2, C2 = (G2, H2, ∂2, c2), where G2 := G, H2 := H/H1 and ∂2 and c2 are the canonical induced maps. Next, let G3 := ∂2(H2) ⊆ G2, H3 := H2 and let ∂3 and c3 be the restrictions of ∂2 and c2. The crossed module C3 = (G3, H3, ∂3, c3) has the feature that ∂3 is injective with dense range; we call such crossed modules thin. CROSSED PRODUCTS FOR CROSSED MODULES 3 Since G3 is a closed normal subgroup of G, G4 := G/G3 is a locally compact group. Let H4 = 0, ∂4 and c4 be trivial. This gives another strict extension of crossed modules C3 → C2 → C4. Two applications of Theorem 5.2 give A ⋊ C ∼= (A ⋊ C1) ⋊ C2 ∼= ((A ⋊ C1) ⋊ C3) ⋊ C4. Thus it remains to study crossed products by crossed modules of the special forms C1, C3 and C4, where G1 = 0, C3 is thin, and H4 = 0. Since G1 = 0, C1 is a very particular Abelian crossed module. Here our duality says that actions of C1 on A are equivalent to C0(cH1)-C∗-algebra structures on A, where cH1 denotes the dual group of H1. The crossed product with C1 is the fibre at 1 ∈ cH1 for the corresponding C0(cH1)-C∗-algebra structure. Since H4 = 0, an action of C4 is equivalent to an action of the group G4; crossed products also have the usual meaning. To understand crossed products for the thin crossed module C3, we replace C3 by a simpler but equivalent crossed module. Equivalent crossed modules have equivalent categories of actions on C∗-algebras, and the equivalence preserves both the underlying C∗-algebra of the action and the crossed products. Many thin crossed modules are equivalent to Abelian crossed modules. We prove this for all thin crossed modules of Lie groups and provide both a sufficient criterion and a counterexample for thin crossed modules of locally compact groups. If the thin crossed module C3 is equivalent to an Abelian crossed module C5, then we may turn a C3-action on A into a C5-action on A with an isomorphic crossed product. By our duality theory, the crossed product by C5 is the fibre at 1 ∈ cH5 for a canonical C0(cH5)-C∗-algebra structure on A ⋊ G5. Assuming that C3 satisfies the mild condition to make it Abelian, we thus decom- pose the crossed product functor for our original crossed module C into four more elementary steps: taking fibres in C0(X)-C∗-algebras twice and taking crossed products by ordinary groups twice. 2. Crossed module actions by correspondences A correspondence between two C∗-algebras A and B is a Hilbert B-module E with a nondegenerate ∗-homomorphism A → B(E). Correspondences are the ar- rows of a weak 2-category Corr(2), with C∗-algebras as objects and isomorphisms of correspondences as 2-arrows (see [2]). Let C = (H, G, ∂, c) be a crossed module of locally compact groups. We may turn C into a strict 2-group with group of arrows G and 2-arrow space G × H, where (g, h) gives a 2-arrow g ⇒ g∂(h). Following [2, Section 4], we define an action of C by correspondences as a mor- phism C → Corr(2) in the sense of [2, Definition 4.1], with continuity conditions added as in [2, Section 4.1]; we define transformations between such actions as in [2, Section 4.2], again with extra continuity requirements; and we define modifica- tions between such transformations as in [2, Section 4.3]. This defines a 2-category Corr(C). The definitions in [2] for actions of general strict 2-categories may be simplified because C is a strict 2-group. Analogous simplifications are already discussed in detail in [2] for weak actions by automorphisms, that is, morphisms to the 4 ALCIDES BUSS AND RALF MEYER 2-category C∗(2) (see [2] for the definition of C∗(2)). For this reason, we merely state the simplified definitions without proving that they are equivalent to those in [2, Section 4]. Definition 2.1. Let C = (H, G, ∂, c) be a crossed module of locally compact groups. An action of C by correspondences consists of • a C∗-algebra A; • correspondences αg : A → A for all g ∈ G with α1 = A the identity correspondence; • isomorphisms of correspondences ωg1,g2 : αg2 ⊗Aαg1 → αg1g2 for g1, g2 ∈ G, where ω1,g and ωg,1 are the canonical isomorphisms αg ⊗A A ∼= αg and A ⊗A αg ∼= αg; • isomorphisms of correspondences ηh : A → α∂(h) for all h ∈ H, where η1 = IdA; • a C0(G)-linear correspondence α from C0(G, A) to itself with fibres αg; more explicitly, this means a space of continuous sections α ⊆ Qg∈G αg such that pointwise products of elements in α with elements of C0(G, A) on the left or right are again in α, pointwise inner products of elements in α are in C0(G, A), and the projections α → αg are all surjective; these must satisfy the following conditions: (1) the following diagram of isomorphisms commutes for all g1, g2, g3 ∈ G: αg3 ⊗A αg2 ⊗A αg1 Idαg3 ⊗A ωg1,g2 ωg2,g3 ⊗A Idαg1 αg2g3 ⊗A αg1 ωg1,g2g3 αg3 ⊗A αg1g2 ωg1g2,g3 αg1g2g3 (2) the following diagram of isomorphisms commutes for all h1, h2 ∈ H: A ⊗A A ηh2 ⊗A ηh1 α∂(h2) ⊗A α∂(h1) can ω∂(h1),∂(h2) A ηh1h2 α∂(h1h2) (3) the following diagram of isomorphisms commutes for all h ∈ H, g ∈ G: A ⊗A αg can αg can αg ⊗A A (2.2) ηh ⊗A Idαg Idαg ⊗A ηcg (h) α∂(h) ⊗A αg αg ⊗A α∂(cgh) ωg,∂(h) αg∂(h) ω∂(cgh),g α∂(cgh)g (4) fibrewise application of ωg2,g1 gives an isomorphism ω : π∗ 2α ⊗C0(G×G,A) π∗ 1α → µ∗α, CROSSED PRODUCTS FOR CROSSED MODULES 5 where π1 and π2 are the two coordinate projections G × G → G and µ is the multiplication map G × G → G; notice that the fibres of these two Hilbert modules over C0(G × G, A) at (g1, g2) ∈ G × G are αg2 ⊗A αg1 and αg1g2 , respectively; (5) fibrewise application of (ηh)h∈H gives an isomorphism η : C0(H, A) → ∂∗α. Since the maps ωg1,g2 and ηh are isomorphisms, the continuity conditions (4) 1α ⊆ Qg1,g2∈G αg2 ⊗A αg1 into µ∗α ⊆ and (5) hold if (ωg1,g2 )g1,g2∈G maps π∗ Qg1,g2∈G αg1g2 and (ηh)h∈H maps C0(H, A) ⊆ Qh∈H A into ∂∗α ⊆ Qh∈H α∂(h); 2 α⊗A π∗ these maps are automatically unitary (isometric and surjective). The trivial C-action on B is given by βg = B for g ∈ G, ωg2,g1 = IdB for g1, g2 ∈ G, ηh = IdB for h ∈ H, and β = C0(G, B). Definition 2.3. Let (A, α, ωA, ηA) and (B, β, ωB, ηB) be C-actions by correspon- dences. A C-equivariant correspondence or transformation between them consists of • a correspondence E from A to B, and • isomorphisms of correspondences χg : E ⊗B βg → αg ⊗A E with χ1 = 1, such that (1) for all h ∈ H, the following diagram commutes: A ⊗A E can E can E ⊗B B (2.4) ηA h ⊗A IdE IdE ⊗B ηB h α∂(h) ⊗A E E ⊗B β∂(h) χ∂(h) (2) for all g1, g2 ∈ G, the following diagram commutes: E ⊗B βg2 ⊗B βg1 IdE ⊗B ωB g1,g2 χg2 ⊗B Idβg1 αg2 ⊗A E ⊗B βg1 Idαg2 ⊗A χg1 E ⊗B βg1g2 αg2 ⊗A αg1 ⊗A E (2.5) χg1g2 αg1g2 ⊗A E ωA g1,g2 ⊗A IdE (3) pointwise application of χg gives a C0(G)-linear isomorphism χ : E ⊗B β → α ⊗A E. A transformation is called an (equivariant Morita -- Rieffel) equivalence if E is an equivalence, that is, the left A-action is given by an isomorphism A ∼= K(E). A transformation is called a (covariant) representation of (A, α) on B if the C-action on B is trivial and E is the correspondence associated to a nondegenerate ∗-homomorphism π : A → M(B). In this case, we also write π : (A, α) → B or simply π : A → B to denote the representation. The continuity condition (3) is equivalent to (χg)g∈G mapping E ⊗B β ⊆ Qg∈G E ⊗B βg into α ⊗A E ⊆ Qg∈G αg ⊗A E; this map is automatically unitary because it is fibrewise unitary. 6 ALCIDES BUSS AND RALF MEYER Definition 2.6. A modification between two transformations (E, χg) and (E ′, χ′ g) from A to B is a unitary W : E → E ′ such that for all g ∈ G the following diagram commutes: (2.7) E ⊗B βg W ⊗B Idβg E ′ ⊗B βg χg χ′ g αg ⊗A E Idαg ⊗A W αg ⊗A E ′ The notion of representation above leads to a definition of crossed products: Definition 2.8. A crossed product for a C-action (A, α, ω, η) by correspondences is a C∗-algebra B with a representation πu : A → B that is universal in the sense that any other representation A → C factors uniquely as f ◦ πu for a morphism (nondegenerate ∗-homomorphism) f : B → M(C). If a crossed product exists, then its universal property determines it uniquely up to canonical isomorphism because an isomorphism in the morphism category of C∗-algebras must be a ∗-isomorphism in the usual sense. We will construct crossed products later using cross-sectional C∗-algebras of Fell bundles. The actions, transformations, and modifications defined above are the objects, arrows and 2-arrows of a weak 2-category (that is, bicategory) Corr(C) with in- vertible 2-arrows. This statement contains the following assertions. Given two C-actions xi = (Ai, αi, ωi, ηi) for i = 1, 2, the transformations x1 → x2 (as objects) and the modifications (as morphisms) between such transformations form a group- oid CorrC(x1, x2); here the composition of modifications is just the composition of unitary operators. Given three C-actions x1, x2 and x3, there is a composition bifunctor CorrC(x2, x3) × CorrC(x1, x2) → CorrC(x1, x3); the composite of two transformations (E1, χ1) from x1 to x2 and (E2, χ2) from x2 to x3 is the transformation from x1 to x3 consisting of E1 ⊗A2 E2 and E1 ⊗A2 E2 ⊗A3 α3g IdE1 ⊗A2 χ2g −−−−−−−−→ E1 ⊗A2 α2g ⊗A1 E2 χ1g ⊗A1 IdE2 −−−−−−−−→ α1g ⊗A1 E1 ⊗A2 E2. This composition is associative and unital up to canonical isomorphisms with suitable coherence properties. To see the associators, notice that (E1 ⊗A2 E2)⊗A3 E3 and E1 ⊗A2 (E2 ⊗A3 E3) are not identical but merely canonically isomorphic. These canonical isomorphisms are the associators. They are canonical enough that being careful about them would lead to more confusion than leaving them out. The identity arrow on A is A with the canonical isomorphisms χg : A ⊗A αg ∼= αg ∼= αg ⊗AA. The composite of another arrow with such an identity arrow is canonically isomorphic to that arrow using the canonical isomorphisms A ⊗A E ∼= E and E ⊗B B ∼= E. These are the unit transformations. The coherence conditions for a weak 2-category listed in [2, Section 2.2.1] are trivially satisfied. 2.1. Translation to Fell bundles. For a locally compact group, it is shown in [2] that actions by correspondences are equivalent to saturated Fell bundles. When we reinterpret everything in terms of Fell bundles, transformations become CROSSED PRODUCTS FOR CROSSED MODULES 7 correspondences between Fell bundles, modifications become isomorphisms of such correspondences, and representations become representations of Fell bundles in the usual sense. Thus the cross-sectional C∗-algebra of the associated Fell bundle has the correct universal property for a crossed product. We want to extend all these results to crossed module actions. Most of the work is already done in [2]. Let C = (G, H, ∂, c) be a crossed module of locally compact groups and let (A, α, ω, η) be an action of C as in Definition 2.1. Forgetting η, the data (A, α, ω) is a continuous action by correspondences of the locally compact group G (as defined in [2] or as in our Definition 2.1 for G viewed as a crossed module). Results in [2] about locally compact groups show that the data (A, α, ω) is equivalent to a saturated Fell bundle over G. The map η gives us some extra data that describes how the group H acts. A transformation between actions of C is the same as a transformation between the resulting actions of G that satisfies an additional compatibility condition with the H-actions. And modifications for G- and C-actions are just the same. It remains to translate everything in Corr(C) related to the group H to the language of Fell bundles. For this we first need some notation. Let A be a Fell bundle over G with fibres Ag at g ∈ G. A multiplier of order g ∈ G of A is a pair µ = (L, R) (left and right multipliers) of maps L, R : A → A such that L(Af ) ⊆ Agf and R(Af ) ⊆ Af g for all f ∈ G, and aL(b) = R(a)b for all a, b ∈ A (see [3, VIII.2.14]). We write M(A)g for the set of multipliers of order g. We usually write µ · a = L(a) and a · µ = R(a). The maps L and R must be fibrewise linear and bounded. The adjoint of µ is defined by µ∗ · a = (a∗ · µ)∗ and a · µ∗ = (µ · a∗)∗, and µ is called unitary if µ∗µ = µµ∗ = 1 (the unit of M(A1)). The set M(A) = Sg∈G M(A)g of all multipliers of A is a Fell bundle over G viewed as a discrete group, called the multiplier Fell bundle of A. We endow M(A) with the strict topology: a net (µi) in M(A) converges strictly to µ ∈ M(A) if and only if µi ·a → µ·a and a·µi → a·µ in A for all a ∈ A. Let UM(A) be the group of unitary multipliers of A of arbitrary order. We may view the fibres Ag as Hilbert bimodules over A1 using the multiplica- tion in the Fell bundle and the inner products hx, yiA1 = x∗y on the right and A1 hx, yi = xy∗ on the left. Taking these operations fibrewise makes the space Γ0(A) of continuous sections of A vanishing at infinity a Hilbert bimodule over C0(G, A1) because the multiplication and involution in the Fell bundle are contin- uous. Lemma 2.9. Let A be a saturated Fell bundle. Then M(A)g is isomorphic to the space of adjointable operators A1 → Ag. The space of strictly continuous sections of M(A) is isomorphic to the space of adjointable operators C0(G, A1) → Γ0(A), that is, the multiplier Hilbert bimodule of Γ0(A) (as defined in [5, Chapter 1.2]). Proof. A multiplier µ of A of order g restricts to an adjointable map A1 → Ag, a 7→ µ · a, with adjoint b 7→ µ∗ · b. Furthermore, since A1 · Ag2 = Ag2 for all g2 ∈ G, an adjointable map A1 → Ag extends uniquely to a multiplier of A. For the last statement, we must show that a section (µg)g∈G of M(A) is strictly continuous if and only if pointwise application of µg and µ∗ g gives well-defined maps 8 ALCIDES BUSS AND RALF MEYER C0(G, A1) ↔ Γ0(A). The existence of a map C0(G, A1) → Γ0(A) is equivalent to the continuity of µg · a for all a ∈ A1, which is equivalent to the continuity of µg · a for all a ∈ A because A = A1 · A. Since the pointwise product maps for a saturated Fell bundle satisfy C0(G, A1) · Γ0(A) = Γ0(A) g := Ag−1 ), the continuity of µ∗ (with A∗ the continuity of µ∗ continuity of a · µg for all a ∈ A1 or to the continuity of a · µg for all a ∈ A. g · ag for all (ag) ∈ Γ0(A) is equivalent to g · ag for all (ag) ∈ C0(G, A1). This is in turn equivalent to the (cid:3) and Γ0(A) · Γ0(A∗) = C0(G, A1), Definition 2.10. Let C = (G, H, ∂, c) be a crossed module of locally compact groups. A Fell bundle over C is a Fell bundle A = (Ag)g∈G over G together with a strictly continuous group homomorphism υ : H → UM(A), such that (1) υh has order ∂(h) for all h ∈ H, and (2) a · υh = υcg(h) · a for all a ∈ Ag and h ∈ H. A Fell bundle over C is saturated if it is saturated as a Fell bundle over G. Theorem 2.11. Actions of C by correspondences are equivalent to saturated Fell bundles over C. Proof. Let (A, α, ω, η) be an action of C by correspondences as in Definition 2.1. The data (A, α, ω) subject to the conditions (1) and (4) in Definition 2.1 are an action of the locally compact group G by correspondences. We have simplified this compared to the definition in [2] by requiring the unit transformation A → α1 to be the identity. It is shown as in the proof of [2, Lemma 3.7] that any action of G by correspondences is equivalent to one with this extra property. Thus our data (A, α, ω) is equivalent to a saturated Fell bundle over G by [2, Theorem 3.17]. Its fibres are Ag = αg−1 ; the multiplication Ag1 ×Ag2 → Ag1g2 is a·b := ωg−1 (a⊗b); the involution is the unique one for which the inner product on Ag = αg−1 has the expected form: hx, yi = x∗ · y. 2 ,g−1 1 An isomorphism of right Hilbert A-modules ηh−1 : A → α∂(h−1) = A∂(h) is equivalent to a unitary multiplier υh of A of order ∂(h) by Lemma 2.9. We claim that the conditions in Definition 2.10 for this map υ : H → UM(A) are equivalent to the conditions (2), (3) and (5) in Definition 2.1. · a′ c−1 g (h) The condition ω∂(h1),∂(h2)(ηh2 ⊗A ηh1 ) = ηh1h2 is equivalent to υ being a group 1 = 1 ∈ A1, ag ∈ Ag, g ∈ G, h ∈ H; here we use 1 run through approximate for all h ∈ H, g ∈ G, ag ∈ Ag, which is equivalent homomorphism. The diagram (2.2) commutes if and only if υ∗ a1 · ag · υ∗ A1 · Ag · A1 = Ag and that υ∗ units, we get υ∗ h ·ag = ag ·υ∗ to condition (2) in Definition 2.10. h = υh−1. Letting a1 and a′ c−1 g (h) 1 for all a1, a′ h · a1 · ag · a′ The equivalence of the continuity conditions in Definitions 2.1 and 2.10 follows (cid:3) as in the proof of the second statement in Lemma 2.9. Since actions by automorphisms may be viewed as actions by correspondences, Theorem 2.11 implies that weak actions by automorphisms as considered in [2, Section 4.1.1] give rise to saturated Fell bundles. We make this more explicit for strict actions: CROSSED PRODUCTS FOR CROSSED MODULES 9 Example 2.12. Let (α, u) be an action of C = (G, H, ∂, c) by ∗-automorphisms on a C∗-algebra A (as defined in [1, Definition 3.1]). That is, α : G → Aut(A) is a (strongly continuous) action of G on A by ∗-automorphisms, u : H → UM(A) is a strictly continuous group homomorphism, and (1) α∂(h)(a) = uhau∗ (2) αg(uh) = ucg(h) for all g ∈ G and h ∈ H. h for all a ∈ A and h ∈ H; and Let A = A ×α G be the semidirect product Fell bundle over G for the action α (see [3, VIII.4]). Its operations are given by (a, f ) · (b, g) = (aαf (b), f g) and (a, g)∗ = (αg−1 (a∗), g−1) for all a, b ∈ A and f, g ∈ G. The multiplier Fell bundle M(A) is isomorphic to the semidirect product Fell bundle M(A) ×α G, where α is tacitly extended to a G-action on M(A). Therefore, the formulas υh · (a, g) := (au∗ h, ∂(h)g) and (a, g) · υh := (au∗ cg(h), g∂(h)) for h ∈ H, g ∈ G, a ∈ A define a unitary multiplier υh of order ∂(h) of A, where υh = (u∗ h, ∂(h)). The pair (A, υ) is a (saturated) Fell bundle over C in the above sense. Next we translate transformations between C-actions into correspondences be- tween Fell bundles over C. Let (A, υA) and (B, υB) be Fell bundles over C. The following definition is an extension of [2, Definition 3.21] from groups to 2-groups (that is, crossed modules). It should also be compared with the notion of equiva- lence between Fell bundles over groupoids appearing in [12, 14, 17]. Definition 2.13. A C-equivariant correspondence from (A, υA) and (B, υB) is a continuous Banach bundle E = (Eg)g∈G over G together with • a continuous multiplication A × E → E that maps Ag1 × Eg2 to Eg1g2; • a continuous multiplication E × B → E that maps Eg1 × Bg2 to Eg1g2; • a continuous inner product h␣, ␣i : E ×E → B that maps Eg1 ×Eg2 to Bg−1 1 g2 ; these must satisfy (1) associativity x · (y · z) = (x · y) · z for (x, y, z) in A × A × E, E × B × B, and A × E × B; (2) A1 · Eg = Eg = Eg · B1 for all g ∈ G; (3) ξ2 7→ hξ1, ξ2i is fibrewise linear for all ξ1 ∈ E and hξ1, ξ2i∗ = hξ2, ξ1i for all ξ1, ξ2 ∈ E; (4) hξ1, ξ2 · b2i = hξ1, ξ2ib2 for all b2 ∈ B, ξ1, ξ2 ∈ E; (5) hξ, ξi ≥ 0 in B1 for all ξ ∈ E, and kξk2 = khξ, ξik; (6) haξ1, ξ2i = hξ1, a∗ξ2i for all a ∈ A, ξ1, ξ2 ∈ E; (7) υA h for h ∈ H, g ∈ G, ξ ∈ Eg. cg(h) · ξ = ξ · υB An isomorphism of correspondences is a homeomorphism E → E ′ that is compati- ble with the left and right multiplication maps and the inner products. Theorem 2.14. Let (A, υA) and (B, υB) be saturated Fell bundles over C. The groupoid of Fell bundle correspondences from (A, υA) to (B, υB) and isomorphisms between such correspondences is equivalent to the groupoid of transformations and modifications between the C-actions associated to (A, υA) and (B, υB). 10 ALCIDES BUSS AND RALF MEYER Proof. Let (E1, χg) be a transformation from the C-action corresponding to A to the C-action corresponding to B. This contains isomorphisms Eg := Ag ⊗A E1 χg−1 −−−→ ∼= E1 ⊗B Bg for g ∈ G. The notation Eg leads to no serious confusion for g = 1 because A1 ⊗A ∼= E1 canonically. The spaces Eg are the fibres of a correspondence E1 = A ⊗A E1 Γ0(A) ⊗A E1 from C0(G, A) to C0(G, B); the continuity of χ gives Γ0(A) ⊗A E1 ∼= E1 ⊗B Γ0(B). We may topologise E := Sg∈G Eg so that Γ0(E) = Γ0(A) ⊗A E1 ∼= E1 ⊗B Γ0(B). The multiplications in A and B induce continuous multiplication maps A × E → E E × B → E, defining Eg as Ag ⊗A E1 and E1 ⊗B Bg, respectively. These multiplications satisfy (a1a2)ξ = a1(a2ξ) for all a1, a2 ∈ A, ξ ∈ E and ξ(b1b2) = (ξb1)b2 for all ξ ∈ E, b1, b2 ∈ B. The bimodule property (aξ)b = a(ξb) for all a ∈ A, ξ ∈ E, b ∈ B is equivalent to (2.5) by a routine computation. The nondegeneracy of correspondences implies Eg = A1 · Eg = Eg · B1. The B-valued inner product on E induces inner product maps h␣, ␣i : E g1 × Eg2 → Bg−1 1 g2 , where E g1 denotes the conjugate space of Eg1 and hξ1 ⊗ b1, ξ2 ⊗ b2i := b∗ 1hξ1, ξ2iBb2 for ξ1, ξ2 ∈ E, b1 ∈ Bg1, b2 ∈ Bg2 , and we identify B = B1. The required Hilbert module properties of this inner product are routine to check. A unitary multiplier u of A of degree g ∈ G gives an adjointable map Ag2 → Agg2 , which induces an adjointable map Eg2 → Egg2 , ξ 7→ u · ξ. Similarly, a unitary multiplier v of B of degree g ∈ G induces maps Eg2 → Eg2g, ξ 7→ ξ ·v, using the map Bg2 → Bg2g. In particular, υA h act on (Eg) by left and right multiplication maps. The condition (2.4) for a transformation is equivalent to υA h for all h ∈ H, ξ ∈ E. For ξ ∈ Eg and g ∈ G, this implies h · ξ = ξ · υB h and υB cg(h) · ξ = ξ · υB υA h . To see this, write ξ = aξ1 with a ∈ Ag, ξ1 ∈ E1, and compute aξ1 · υB h ξ1 = υA cg(h) · aξ1 using condition (2) in Definition 2.10. Thus a transformation between actions of C by correspondences yields a correspondence between the associated Fell bundles. h = aυA To show the converse, take a correspondence of Fell bundles E. Then Eg is a Hilbert module over B1, and the multiplication maps E1 ⊗B1 Bg → Eg must be unitary. These maps are the fibres of a continuous map, so that E ∼= E1 ⊗B1 B. The left multiplication gives unitaries Ag ⊗A1 E1 ∼= Eg ∼= E1 ⊗B1 Bg. These yield the maps χg−1 : Ag ⊗A1 E1 above, it is then routine to show that (E1, χg) is a transformation of C-actions. ∼−→ E1 ⊗B1 Bg. Reversing the computations A modification between transformations consists of an isomorphism of corre- ∼= E ′ ⊗B Bg = spondences E → E ′, which then induces isomorphisms Eg = E ⊗B Bg E ′ g. The compatibility with χg shows that these maps preserve also the left module structure, hence give an isomorphism of Fell bundle correspondences. Conversely, CROSSED PRODUCTS FOR CROSSED MODULES 11 an isomorphism of Fell bundle correspondences is determined by its action on the unit fibre E1, which is a unitary W that satisfies the compatibility condition (2.7). Thus our constructions of Fell bundle correspondences from C-actions and vice versa are fully faithful functors between the respective categories. These functors are inverse to each other up to natural isomorphisms. (cid:3) The invertible transformations between C-actions are those where E is an im- primitivity bimodule. In this case, E carries an A-valued left inner product as well such that Ahξ, ηi · ζ = ξ · hη, ζiB for all ξ, η, ζ ∈ E. This leads to an appropriate definition of a Morita -- Rieffel equivalence between Fell bundles over C. The func- toriality of crossed products implies that Morita -- Rieffel equivalent Fell bundles over C have Morita -- Rieffel equivalent crossed products. Now we specialise to a transformation (E1, χ) from A to B where the under- lying right Hilbert module of E1 is B itself. Thus the left A-action on E1 is a nondegenerate ∗-homomorphism A → M(B). Proposition 2.15. Transformations (E1, χ) between two C-algebras A and B with E1 = B as a Hilbert B-module correspond to morphisms between the Fell bundles A and B over C associated to A and B, that is, to maps f : A → M(B) • that are fibrewise linear; • satisfy f (a∗) = f (a)∗ and f (a1 · a2) = f (a1) · f (a2) for all a, a1, a2 ∈ A; • that are nondegenerate (f (A1) · B1 = B1); • satisfy f (υA h for all h ∈ H, where nondegeneracy has been used to h ) = υB extend f to multipliers; and • that are strictly continuous in the sense that the map A × B → B, (a, b) 7→ f (a) · b, is continuous. The joint continuity in the last condition is equivalent to the continuity of the maps a 7→ f (a) · b for all b ∈ B1. Proof. Let E be the Fell bundle correspondence associated to a transformation (E1, χ). We assume that E1 = B1 = B as a right Hilbert B1-module. This gives an isomorphism of Banach bundles E ∼= B that is right B-linear and unitary for the B-valued inner product because the multiplication map (ξ, b) 7→ ξ · b induces a unitary E1 ⊗B1 B ∼= E. Conversely, a Fell bundle correspondence with E = B as right Hilbert B-module has E1 = B1 as a right Hilbert B1-module. A Fell bundle correspondence with E = B as a right Hilbert B-module is the same as a continuous multiplication A × B → B with the properties in Defini- tion 2.13. These properties are equivalent to f being a morphism A → M(B) of Fell bundles over C as defined in the statement of the proposition. The equivalence of joint and separate continuity follows because kf (a)k ≤ kak for all a ∈ A. (cid:3) Finally, we specialise to representations of C-actions. These are, by definition, transformations to a C∗-algebra B equipped with the trivial C-action. The trivial C-action on B corresponds to the constant Fell bundle with fibre B and υB h = 1B for all h ∈ H. A morphism of Fell bundles from (A, υ) to such a constant Fell bundle is a map ρ : A → M(B) such that (1) ρ restricts to linear maps Ag → M(B) for all g ∈ G; (2) ρ(a∗) = ρ(a)∗ and ρ(a1 · a2) = ρ(a1) · ρ(a2) for all a, a1, a2 ∈ A; 12 ALCIDES BUSS AND RALF MEYER (3) for each b ∈ B, the map a 7→ ρ(a)b is continuous from A to B (continuity); (4) ρ(A1) · B = B (nondegeneracy); (5) ρ(υA h ) = 1 for all h ∈ H; the last condition involves the canonical extension of a representation of a Fell bundle to multipliers. Except for the last condition, this is the standard definition of a representation of a Fell bundle (see [3]). We are going to describe the crossed product of a C-action by correspondences as a quotient of the cross-sectional C∗-algebra of the associated Fell bundle, generalis- ing the description in [1] for strict C-actions. The cross-sectional C∗-algebra C∗(A) of a Fell bundle A over the locally compact group G is constructed in [3, Chap- ter VIII] as the C∗-completion of the space Cc(A) of compactly supported sections of A with a suitable convolution and involution. It comes with a canonical map ρu : A → M(C∗(A)), which is a representation of A viewed as a G-C∗-algebra (via Theorem 2.11). Definition 2.16. For a Fell bundle (A, υ) over C, let Iυ be the closed, two-sided ideal of C∗(A) generated by the multipliers {ρu(υh) − 1 : h ∈ H}, that is, Iυ := span{x(ρu(υh) − 1)y : x, y ∈ C∗(A), h ∈ H}. We call C∗(A, υ) := C∗(A)/Iυ the cross-sectional C∗-algebra of (A, υ). Since ρ(υh) = 1 is the only extra condition needed for a representation of A to be a representation of (A, υ), C∗(A, υ) is the largest quotient of C∗(A) on which ρu gives a representation of (A, υ) as a C-C∗-algebra. Proposition 2.17. The C∗-algebra C∗(A, υ) := C∗(A)/Iυ with the canonical representation of A is a crossed product (A, υ) ⋊ C. That is, representations of C∗(A, υ) correspond bijectively to representations of (A, υ). Proof. Morphisms C∗(A) → B are in bijection with representations of the Fell bundle A on B, mapping a representation ρ of A to its integrated form ∫ ρ (see [3, VIII.11]). Representations of C∗(A, υ) = C∗(A)/Iυ correspond bijectively to representations of C∗(A) that vanish on Iυ. Moreover, for a representation π of A, ∫ ρ(Iυ) = 0 ⇐⇒ ∫ ρ(ρu(υh)ξ − ξ) = 0 for all h ∈ H, ξ ∈ Cc(A) ⇐⇒ ρ(υh · a) = ρ(a) for all a ∈ A. Thus the bijection from representations of A on B to morphisms C∗(A) → B restricts to a bijection from representations of (A, υ) to morphisms C∗(A, υ) → B. (cid:3) 3. Equivalence of crossed modules One should expect equivalent crossed modules to have equivalent 2-categories of actions on C∗-algebras by correspondences. What does this mean? A functor Corr(C) → Corr(C′) for two crossed modules C and C′ consists of • a map F on objects, which maps each C-action x to a C′-action F (x); • for any two C-actions x1 and x2, a functor F : CorrC(x1, x2) → CorrC′ (F (x1), F (x2)); CROSSED PRODUCTS FOR CROSSED MODULES 13 • for any three C-actions x1, x2, x3, natural isomorphisms F (f ) ◦ F (g) ∼= F (f ◦ g) in CorrC′ (F (x1), F (x3)) for f ∈ CorrC(x2, x3), g ∈ CorrC(x1, x2); • natural isomorphisms F (1x) ∼= 1F (x); the natural transformations in the last two conditions must satisfy suitable coher- ence axioms (see [9]). Such a functor is an equivalence if the map F on objects is essentially surjective and the functors on the arrow groupoids are equivalences. An equivalence of 2-categories has a quasi-inverse that is again a functor, such that the compositions in either order are equivalent to the identity functor in a suitable sense. We will only consider functors constructed from homomorphisms of crossed modules. Let Ci = (Gi, Hi, ∂i, ci) for i = 1, 2 be crossed modules. A homomor- phism of crossed modules C1 → C2 is a pair of continuous group homomorphisms ϕ : G1 → G2, ψ : H1 → H2 such that ∂2 ◦ ψ = ϕ ◦ ∂1 and c2,ϕ(g)(ψ(h)) = ψ(c1,g(h)). Such a homomorphism induces a functor (ϕ, ψ)∗ : Corr(C2) → Corr(C1), (3.1) by sending a C2-action (A, α, ω, η) to (A, ϕ∗α, (ϕ × ϕ)∗ω, ψ∗η), a transformation (E, χ) to (E, ϕ∗χ), and a modification W again to W ; the natural transformations in the definition of a functor above are trivial and therefore coherent. When we translate to Fell bundles using Theorem 2.11, the functor (3.1) sends a Fell bundle (A, υ) over C1 to the Fell bundle (ϕ∗A, ψ∗υ) over C2, where ϕ∗A denotes the pull-back of A along ϕ and ψ∗υ = υ ◦ ψ. Definition 3.2. The arrow groupoid of C = (G, H, ∂, c) is the transformation groupoid G ⋊ H for the right action of the topological group H on the topological space G defined by g · h := g∂(h). A homomorphism (ϕ, ψ) of crossed modules induces a functor between the arrow groupoids. We call (ϕ, ψ) an equivalence if the induced functor on arrow groupoids gives an equivalence of topological groupoids. To make this precise, we turn a functor between the arrow groupoids into a Hilsum -- Skandalis morphism, that is, a topological space endowed with commuting actions of G1 ⋊ H1 and G2 ⋊ H2 (see [8]). Let X := G1 × H2 and define anchor maps πi : X → Gi for i = 1, 2 by π1(g1, h2) := g1 and π2(g1, h2) := ϕ(g1)∂2(h2). Define a left H1-action and a right H2-action on X by h · (g1, h2) := (g1∂1(h)−1, ψ(h)h2) (g1, h2) · h := (g1, h2h) for h ∈ H1, for h ∈ H2. These definitions turn X into a G1 ⋊ H1-G2 ⋊ H2-bispace. This bispace is always a Hilsum -- Skandalis morphism, that is, the action of G2 ⋊ H2 is free and proper and π1 induces a homeomorphism from the G2 ⋊ H2-orbit space to G1. The homo- morphism (ϕ, ψ) is an equivalence if the associated bispace is a Morita equivalence as in [13]. In our case, this happens if and only if the action of G1 ⋊ H1 on X is free and proper with orbit space projection π2. Equivalently, X is a Morita equivalence between G1 ⋊ H1 and G2 ⋊ H2 as in [13]. It is enough to check that H1 acts freely and properly on X with X/H1 ∼= G2 via π2. Lemma 3.3. The homomorphism (ϕ, ψ) is an equivalence if and only if 14 ALCIDES BUSS AND RALF MEYER (1) the map (∂1, ψ) : H1 → {(g1, h2) ∈ G1 × H2 : ϕ(g1) = ∂2(h2)} = G1 ×G2 H2 is a homeomorphism, where the codomain carries the subspace topology; (2) the map π2 : G1 ×H2 → G2, (g1, h2) 7→ ϕ(g1)·∂2(h2), is an open surjection. The first condition says that H1 is the pull-back of H2 along ϕ, the second condition is a transversality condition for this pull-back. Proof. The freeness of the H1-action on X means that no h 6= 1 in H1 has ψ(h) = 1 and ∂1(h) = 1, that is, the map in (1) is injective. 1)−1g1) = ∂2(h′ 2h−1 Let (g1, h2) ∈ G1 × H2 and (g′ 1, h′ 2) ∈ G1 × H2. They have the same π2-image if and only if ϕ(g1)∂2(h2) = ϕ(g′ 1)∂2(h′ 2), if and only if ((g′ 2 ) ∈ G1 × H2 satisfies ϕ((g′ 2 ); hence the surjectivity of the map in (1) means that the map π′ 2 is auto- matically continuous, and it is open if and only if π2 is because the projection X → X/H1 is open and continuous. Hence condition (2) means that π′ 2 is surjec- tive and open. Being injective and continuous as well, it is a homeomorphism. 2 : X/H1 → G2 induced by π2 is injective. The map π′ 1)−1g1, h′ 2h−1 Finally, the properness of the H1-action is equivalent to the following: for all compact subsets K ⊆ G1, L ⊆ H2, the set of h ∈ H1 with ∂1(h) ∈ K −1 · K and ψ(h) ∈ L · L−1 is compact. This is equivalent to the properness of the map in (1). We already know that this map is a continuous bijection. Such a map is proper if and only if it is a homeomorphism. (cid:3) For discrete crossed modules, [16, Proposition 6.3] says that the equivalences in the above sense are the acyclic cofibrations in a suitable model structure. The arrow groupoid does not yet encode the multiplication in G and the conju- gation action c. These are encoded in a continuous functor (3.4) M : (G ⋊ H) × (G ⋊ H) → (G ⋊ H), (g1, g2) 7→ g1 · g2, (g1, h1), (g2, h2) 7→ (g1 · g2, c−1 g2 (h1) · h2); here (g, h) denotes the 2-arrow h : g ⇒ g∂(h). The existence and associativity of this functor is equivalent to the axioms of a crossed module. The orbit space of the arrow groupoid is π1(C) := coker ∂; the group structure on the orbit space is induced by M ; the isotropy group of any g ∈ G is isomorphic to π2(C) := ker ∂. This is an Abelian group, and the action c induces a π1(C)-module structure on it; this module structure may also be expressed through M . Since we may express them through canonical extra structure on the arrow groupoid, the group π1(C) and the π1(C)-module π2(C) are invariant under equivalences; for crossed modules of locally compact groups, this invariance includes the induced topologies on them. Discrete crossed modules are classified up to equivalence by the group π1(C), the π1(C)-module π2(C), and a cohomology class in H3(π1(C), π2(C)) (see [10]). We shall not attempt such a complete classification of crossed modules of locally compact groups here. It is useful, however, to know that π1 and π2 are invariant under equivalence. Theorem 3.5. The functor Corr(C2) → Corr(C) induced by an equivalence of crossed modules (ϕ, ψ) : C → C2 is an equivalence of 2-categories. CROSSED PRODUCTS FOR CROSSED MODULES 15 This equivalence intertwines the crossed product functors on both categories. Proof. We must show that any C-action (A, α, ω, η) is isomorphic to (ϕ, ψ)∗(A2, α2, ω2, η2) for an action of C2 and that any transformation (E, χ) between C-actions is isomor- phic to (ϕ, ψ)∗(E2, χ2) for a transformation of C2-actions (E2, χ2) that is unique up to isomorphism. Since (ϕ, ψ)∗ does not change the underlying C∗-algebras, we may put A2 := A. Define isomorphisms of correspondences ηg,h : αg → αg∂(h) for g ∈ G, h ∈ H by ηg,h : αg ∼= A ⊗A αg ηh⊗AIdαg −−−−−−−→ α∂(h) ⊗A αg ω(g,∂(h)) −−−−−−→ αg∂(h). The space G× H is the space of 2-arrows in C, and an action in the sense of [2, Def- inition 4.1] provides ηg,h as above. We removed redundancy in Definition 2.1 and kept only η1,h = ηh. Now we need the whole family ηg,h. It depends continuously on (g, h) ∈ G × H. The naturality of the isomorphisms ωg1,g2 with respect to 2-arrows says that the diagrams αg2 ⊗A αg1 ωg1,g2 αg1g2 (3.6) ηg2,h2 ⊗A ηg1,h1 ηg1g2,c−1 g2 (h1)h2 αg2∂(h2) ⊗A αg1∂(h1) ωg1∂(h1),g2∂(h2) αg1∂(h1)·g2∂(h2) commute for all g1, g2 ∈ G, h1, h2 ∈ H. The isomorphisms ηg,h : αg → αg∂(h) for g ∈ G, h ∈ H turn α into a G ⋊ H- equivariant Hilbert bimodule over C0(G, A). Since G ⋊ H is Morita equivalent to G2 ⋊ H2 via the bispace X, we may transport this to an G2 ⋊ H2-equivariant Hilbert bimodule α2 over C0(G2, A). We describe α2 more explicitly. The pull-back π∗ 1 α ∼= α ⊗ C0(H2) along π1 : X → G is an H × H2-equivariant Hilbert module over C0(X, A). Let α2 ⊆ M(π∗ 1 α) be the space of all bounded continuous sections of π∗ 1α that are H-invariant and whose norm function is in C0(X/H). Since X/H ∼= G2 via π2 and the actions of H2 and H commute, we may view this as an H2-equivariant Hilbert C0(G2, A)-module. The left A-action survives these constructions because the action of H commutes with it. The fibre of α2 at g2 ∈ G2 is isomorphic to αg for any g ∈ G for which there is h2 ∈ H2 with π2(g, h2) = g2, that is, g2 = ϕ(g)∂2(h2). These isomorphisms on the fibres are continuous and give a canonical H-equivariant isomorphism ϕ∗α2 ∼= α. The pull-backs π∗ 1α, π∗ 2α and µ∗α in the definition of ω are H × H-equivariant Hilbert modules over C0(G × G, A). That is, they are representations of the groupoid (G ⋊ H)2 = (G × G) ⋊ (H × H). This groupoid is equivalent to (G2 ⋊ H2)2 via the bispace X 2. Now pull back π∗ 2α, µ∗α and the isomorphism 1α → µ∗α to X 2 and push all this down to the category of ω : π∗ 2, A)-modules by taking H 2-invariants. The resulting H 2 Hilbert C0(G2 2 α2, and µ∗α2, 2, A)-modules are canonically isomorphic to π∗ 2 -equivariant Hilbert C0(G2 2α ⊗C0(G×G,A) π∗ 1α, π∗ 1α2, π∗ 16 ALCIDES BUSS AND RALF MEYER respectively. Hence ω induces a C0(G2 × G2)-linear H2 × H2-equivariant unitary operator ω2 : π∗ 2α2 ⊗C0(G2×G2,A) π∗ 1α2 → µ∗α2. The H2 ×H2-equivariance shows that ω2 satisfies the analogue of (3.6), which gives conditions (2) and (3) in Definition 2.1. To prove that ω2 inherits condition (1) in Definition 2.1, we use that for each g1, g2, g3 ∈ G2 there are g′ 3 ∈ G such that 3 with ω2,g1,g2 = ωg′ . We suitable natural identifications of the fibres α2,gi ∼= ω. Hence (ϕ, ψ)∗(A, α2, ω2, η2) = also find a canonical isomorphism (ϕ × ϕ)∗ω2 (A, α, ω, η). 2, ω2,g1g2,g3 = ωg′ 3 , ω2,g1,g2g3 = ωg′ 3 , ω2,g2,g3 = ωg′ 2g′ 1,g′ ∼= αg′ ig′ and α2,gigj 2,g′ ∼= αg′ i 1,g′ 1g′ 2,g′ 1, g′ 2, g′ j In the same way, we may also lift a transformation (E, χ) of C-actions to one between the corresponding C2-actions. We may put E2 = E because our lifting does not change the underlying C∗-algebras. The assumptions in Definition 2.3 imply that the isomorphism χ is H-equivariant. Hence we may pull it back to an H × H2-equivariant isomorphism over X and then push down to G2 to get an H2-equivariant isomorphism χ2 : E ⊗B β2 → α2 ⊗A E over G2. The same arguments as above show that (E, χ2) is a transformation that is a (ϕ, ψ)∗-preimage of the transformation (E, χ), and the only one with this property up to isomorphism of transformations. It is clear that (ϕ, ψ)∗ maps a trivial action of C2 to a trivial action of C. Furthermore, on the level of transformations, E is a morphism (its underlying Hilbert module is B itself) if and only if (ϕ, ψ)∗E is a morphism. Therefore, (ϕ, ψ)∗ gives a bijection between C2-representations of (A, α2, ω2, η2) on B and C-representations of (A, α, ω, η) on B. By the universal property, there is a unique isomorphism A ⋊α,ω,η C ∼= A ⋊α2,ω2,η2 C2 that maps the universal C-representation to the universal C2-representation. (cid:3) Example 3.7. Let C = (G, H, ∂, c) be a crossed module and let G1 ⊆ G be a closed subgroup such that the map G1 × H → G, (g, h) 7→ g · ∂(h), is open and surjective. Let H1 := ∂−1(G1) and let ∂1 : H1 → G1 and c1 : G1 → Aut(H1) be the restrictions of ∂ and c. Then the embedding (ϕ, ψ) of C1 = (G1, H1, ∂1, c1) into C is an equivalence by Lemma 3.3. The second condition in Lemma 3.3 is our assumption. The first condition is that the group homomorphism H1 → {(g1, h) ∈ G1 × H : ϕ(g1) = ∂(h)}, h1 7→ (ψ(h1), ∂1(h1)), is a homeomorphism. Indeed, the map has the restriction of the first coordinate projection as a continuous inverse. Theorem 3.5 says that a Fell bundle over C1 extends to one over C in a natural and essentially unique way. Example 3.8. Let N ⊆ H be a closed c(G)-invariant subgroup such that ∂ re- stricts to a homeomorphism from N onto a closed subgroup of G; then ∂(N ) is normal in G. Let G2 := G/∂(N ), H2 := H/N , and let ∂2 : H2 → G2 and c2 : G2 → Aut(H2) be the induced maps. Then the projection map (ϕ, ψ) from C to C2 = (G2, H2, ∂2, c2) is an equivalence by Lemma 3.3. The second condition in Lemma 3.3 follows because already the projection G → G2 is open and surjective. CROSSED PRODUCTS FOR CROSSED MODULES 17 The first condition requires the map H → {(g, h2) ∈ G × H2 : ϕ(g) = ∂2(h2)}, h 7→ (ψ(h), ∂(h)), to be a homeomorphism. Injectivity and surjectivity of this map are clear, and its properness is not hard to check. This implies that the map is a homeomorphism. Theorem 3.5 says that any Fell bundle over C is the pull-back of a Fell bundle over C2, which is unique up to isomorphism and depends naturally on the original Fell bundle over C. The reason to expect this is that the unitary multipliers υh for h ∈ N trivialise our Fell bundle over N -cosets. Example 3.9. Even more specially, assume that C is group-like, that is, ∂ is a homeomorphism onto a closed (normal) subgroup of G. Theorem 3.5 says that C-actions are equivalent to actions of the locally compact group G/∂(H) viewed as a crossed module. Actions of the latter are equivalent to Fell bundles over the locally compact group G/∂(H) in the usual sense. Actions of C are a Fell bundle analogue of Green twisted actions. In particular, any Green twisted action of (G, ∂(H)) gives rise to a Fell bundle over G/∂(H). For discrete groups, this equivalence is already contained in [6]. We have defined when a crossed module homomorphism is an equivalence. Since its inverse is usually not described by a crossed module homomorphism, we are led to the following equivalence relation for crossed modules: Definition 3.10. Two crossed modules of locally compact groups C and C′ are equivalent if they are connected by a chain of crossed module homomorphisms C = C0 ← C1 → C2 ← C3 → · · · ← C′ where each arrow is a crossed module homomorphism that is an equivalence. Since equivalence of 2-categories formulated in terms of functors is a symmetric relation (quasi-inverses exist and are again functors), equivalent crossed modules have equivalent action 2-categories Corr(C) by Theorem 3.5. The topological group π1(C) and the topological π1(C)-module π2(C) are invariant under equivalence be- cause, as already mentioned, equivalences implemented by homomorphisms pre- serve π1 and π2. 3.1. Simplification of thin crossed modules. We call a crossed module C = (G, H, ∂, c) thin if ∂ is injective and has dense range. In the discrete case, this implies that C is equivalent to the trivial crossed module. An interesting example is the crossed module associated to a dense embedding Z → T, which acts on the corresponding noncommutative torus (see [1]). Being thin is an invariant of equivalence of crossed modules. In a thin crossed module, the action c is dictated by ∂(cg(h)) = g∂(h)g−1 because ∂ is injective. We want to simplify thin crossed modules up to equivalence. The best result is available if both G and H are Lie groups (we allow an arbitrary number of connected components). In that case, the following theorem gives a complete classification up to equivalence. Theorem 3.11. Any thin crossed module of Lie groups is equivalent to one where G = Rn for some n ∈ N and H is a dense subgroup of G with the discrete topology. 18 ALCIDES BUSS AND RALF MEYER And it is equivalent to one where G = Tn for some n ∈ N and H is a dense subgroup of G with the discrete topology. Two thin crossed modules Ci = (Gi, Hi, ∂i, ci) with Gi = Rni for i = 1, 2 and discrete Hi are equivalent if and only if there is an invertible linear map Rn1 → Rn2 mapping ∂1(H1) onto ∂2(H2). Proof. Let C = (G, H, ∂, c) be a thin crossed module of Lie groups. Let G1 ⊆ G be the connected component of the identity. Let H1 = ∂−1(G1) ⊆ H. Since ∂(H) is dense in G, it meets every connected component. Hence we are in the situation of Example 3.7, and C is equivalent to C1 = (G1, H1, ∂1, c1), where ∂1 and c1 are the restrictions of ∂ and c. The Lie group G1 is connected. Let ϕ : G2 → G1 be the universal covering of G1 and let H2 := {(h1, g2) ∈ H1 × G2 : ∂1(h1) = ϕ(g2)} be its pull-back to a covering of H1. Let ψ : H2 → H1 and ∂2 : H2 → G2 be the coordinate projections. Then ψ is a covering map; its kernel is a discrete central subgroup N ⊆ H2. The homomorphism ∂2 maps N homeomorphically onto the kernel of the covering map ϕ. Hence we are in the situation of Example 3.8, and C1 is equivalent to C2 = (G2, H2, ∂2, c2), where c2 lifts c1. (The composite C2 → C1 → C is also an equivalence, so we could have gone to C2 in only one step.) The Lie group G2 is simply connected. Let N2 be the connected component of the identity in H2. We claim that ∂2 maps N2 homeomorphically onto a closed subgroup of G2. We are going to prove this later, so let us assume this for a moment. Then we are again in the situation of Example 3.8. Letting G3 := G2/∂2(N2), H3 := H2/N2, and taking the induced maps ∂3 and c3, we get a crossed module equivalence C2 → C3 = (G3, H3, ∂3, c3). The Lie group H3 is discrete, and G3 is still a simply connected Lie group because it is a quotient of a simply connected group by a closed, connected, normal subgroup. The conjugation action c3 : G3 → Aut(H3) is a continuous action of a connected group on a discrete space. It must be trivial. Hence g∂3(h)g−1 = ∂3(h) for all h ∈ H3. Since we started with a thin crossed module, ∂3(H3) is dense in G3, so the above equality extends to all of G3, proving that G3 is Abelian. Any Abelian simply connected Lie group is of the form Rn for some n ∈ N. Thus C3 is a crossed module equivalent to C with G3 = Rn and discrete H3. Now we show that ∂2 is a homeomorphism from N2 onto a closed subgroup of G2 if G2 is simply connected. Let h and g be the Lie algebras of H2 and G2, re- spectively. The dense embedding ∂2 induces an injective map h → g, whose image is a Lie ideal. Hence there is a Lie algebra g/h and a Lie algebra homomorphism g → g/h. By Lie's theorems, there is a simply connected Lie group K with Lie algebra g/h and a Lie group homomorphism G2 → K that induces g → g/h on the Lie algebras. The kernel of this homomorphism is a connected, closed, normal sub- group, and its Lie algebra is h. This closed normal subgroup is ∂2(N2) because the exponential map is a local homeomorphism from the Lie algebra to the Lie group near the identity element. Moreover, the map ∂2 is indeed a homeomorphism from N2 onto this closed normal subgroup. To get another equivalent crossed module C4 with G4 ∼= Tn, we choose elements h1, . . . , hn ∈ H3 whose images form a basis in Rn, using the density of ∂(H3) in CROSSED PRODUCTS FOR CROSSED MODULES 19 G3 = Rn. These elements generate a subgroup N3 isomorphic to Zn in H3, which is mapped by ∂3 onto a discrete subgroup in G3. Using Example 3.8 once again, we find that C3 is equivalent to C4 with G4 = G3/N3 ∼= Tn, H4 = H3/N3 and ∂4 and c4 induced by ∂3 and c3. The quotient H4 is still discrete, and mapped by ∂4 onto a dense subgroup in the torus G4. Thus C4 has the desired form. Finally, we show that crossed modules with Gi ∼= Rni and discrete Hi are only equivalent when they are isomorphic through some invertible linear map G1 → G2 that maps ∂(H1) onto ∂(H2). We observe first that the quotient Lie algebra g/h for a crossed module C is invariant under equivalence. Any homomorphism of crossed modules C1 → C2 induces a pair of Lie algebra homomorphisms h1 → h2 and g1 → g2 that intertwine the differentials of ∂1 and ∂2 and hence induce a homomorphism g1/h1 → g2/h2. It is not hard to see that this map is invertible if the homomorphism is an equivalence. Roughly speaking, g/h is the tangent space at the unit element in π1(C), and it is invariant because π1(C) as a topological group is invariant. If X ∈ g/h and one representative X ∈ g has the property that exp( X) ∈ ∂(H), then this holds for all representatives. Let us denote this subset of g/h by T (C). A crossed module homomorphism C1 → C2 maps T (C1) → T (C2), and if it is an equivalence, then it maps T (C1) onto T (C2) because it induces a bijection between Gi/∂i(Hi) for i = 1, 2. Hence the pair consisting of the Lie algebra g/h and the subset T (C) is invariant under equivalence of crossed modules (in the sense that equivalent crossed modules have canonically isomorphic invariants). If a crossed module has discrete H and G = Rn, then we identify g/h = Rn in the obvious way and find that the subset of X with exp( X) ∈ ∂(H) is precisely ∂(H). This determines our crossed module because ∂ is injective, c is trivial, and H is discrete. Hence crossed modules of this form are only equivalent when they are isomorphic. (cid:3) Given a thin crossed module C = (G, H, ∂, c), the proof of Theorem 3.11 shows that g/h is the Abelian Lie algebra Rn for some n ∈ N, that T (C) ⊆ g/h is a dense subgroup, and that C is equivalent to the thin Abelian crossed module C′ with G′ = g/h, H ′ = T (C) with the discrete topology, ∂′ the inclusion map, and trivial c′. Hence we get an explicit Abelian replacement for C and do not have to follow the steps in the above proof to construct it. Example 3.12. Consider the thin crossed module from a dense embedding R ⊂ T2 given by a line of irrational slope. This is not yet in standard form because R is not discrete. Of course, the image of the connected component of the identity is not closed in this example. But when we pass to the universal covering R2 of T2, then we get the equivalent crossed module R + Z2 ⊂ R2. Now the image of the connected component is a line of irrational slope, which is closed in R2, and we may divide it out to get an equivalent crossed module Z2 → R. This is now in standard form, and indeed of the expected form G′ = g/h, H ′ = T (C). Example 3.13. Let G = T, and let H = G with the discrete topology. This is an example of a crossed module which is already in the second standard form of Theorem 3.11. Theorem 3.11 works also if our Lie groups have uncountably many components. 20 ALCIDES BUSS AND RALF MEYER The standard form with G = Tn in Theorem 3.11 is sometimes less useful because it is less unique, but it has the advantage that G is compact. Now we turn to general locally compact groups. Here a complete classification seems hopeless, and even Abelianness fails in general (see Example 3.16 below). The following theorem is what we can prove: Theorem 3.14. Any thin crossed module of locally compact groups is equivalent to a thin crossed module C′ = (G′, H ′, ∂′, c′) with compact G′ and discrete H ′. The following are equivalent: (1) C is equivalent to a thin crossed module C′ with compact Abelian G′, dis- crete H ′, and trivial c′; (2) C is equivalent to a crossed module C′ with trivial c′; (3) the commutator map G × G → G, (x, y) 7→ xyx−1y−1, factors as ∂ ◦ γ for a continuous map γ : G × G → H. Proof. We construct a finite chain of crossed modules Ci = (Gi, Hi, ∂i, ci) equiv- alent to C with increasingly better properties. Each step uses Example 3.7 or Example 3.8, which describe how the maps ∂i and ci and one of the groups in the next step are constructed. We must only describe the subgroup to which we restrict in Example 3.7 or the normal subgroup we divide out in Example 3.8. By the structure theory of locally compact groups, the group G contains an open, almost connected subgroup G1. Restricting to this subgroup as in Example 3.7, we get an equivalent crossed module C1 with almost connected G1. Next we divide out suitable compact, c1-invariant (hence normal) subgroups in H1. On such subgroups, the map ∂1 is automatically a homeomorphism onto a compact subgroup of G3, so that we may divide it out using Example 3.8. Let H 0 1 be the connected component of the identity in H1. This closed subgroup is in- trinsically defined and hence c1-invariant. Being connected, it contains a compact normal subgroup N so that H 0 1 /N is a Lie group. If N1 and N2 are such sub- groups, then so is N1 · N2 because of normality. Hence there is a maximal compact normal subgroup N1 in H 0 1 is intrinsically defined and hence c1-invariant. We may now pass to an equivalent quotient crossed module C2 with H2 = H1/N1. The group G2 is still almost connected, and H 0 1 /N1 is now a Lie group. 1 . This subgroup of H 0 2 = H 0 Since G2 is almost connected, it contains a decreasing net of compact normal subgroups Ki such that G2/Ki are Lie groups and T Ki = {1}. The map ∂1 induces Lie algebra homomorphisms from the Lie algebra h2 of H 0 2 to the Lie algebras of G2/Ki. Since ∂1 is injective and h2 is finite-dimensional, this map is injective for some i. Thus there is a compact normal subgroup K ⊆ G2 such that G2/K is a Lie group and the map h2 → g2/k is injective. Passing to a subgroup and a covering group G3 → G2 as in the proof of Theo- rem 3.11, we now find an equivalent crossed module C3 for which G3/K is simply connected. The Lie algebras of H2 and H3 are the same, so the map h3 → g3/k remains injective. Since G3/K is simply connected, there is a unique connected, closed, normal subgroup L3 of G3/K whose Lie algebra is the image of h3 (see the proof of Theorem 3.11); a long exact sequence of homotopy groups using that π2 vanishes for all Lie groups (such as G3/KL3) shows that L3 is simply connected. CROSSED PRODUCTS FOR CROSSED MODULES 21 The map H 0 3 → L3 is the identity on Lie algebras and hence a covering map. Since L3 is already simply connected, it is a homeomorphism. Then the map ∂3 : H 0 3 → G3 is a homeomorphism onto its image as well. Hence Example 3.8 gives us an equivalent crossed module C4 with H4 = H3/H 0 3 . Now H4 is totally disconnected, and G4 is still almost connected. Since H4 is totally disconnected, any action of a connected group on it is trivial. In particular, the conjugation action of G4 factors through G4/G0 4. Since this group is compact, an intersection U ′ := Tg∈G4 cg(U ) for an open subset U ⊆ H4 is again open. If U is a compact open subgroup, then this intersection is a compact, open, and c4-invariant subgroup in H4. Example 3.8 gives us an equivalent crossed module C5 with H5 = H4/U ′. Now H5 is discrete, and G5 is still almost connected. 5 is connected and H5 discrete, cg(h) = h for all h ∈ H5. Hence g∂5(h)g−1 = ∂5(h) for all h ∈ H5. Since ∂5(H) is dense in G5, this shows that g is central. Hence the connected component G0 5 belong to the connected component of the identity. Since G0 5 is a central subgroup in G5. Let g ∈ G0 The image of K3 ⊆ G3 in G5 is a compact normal subgroup K5 such that G5/K5 5 surjects onto it. Since G0 is connected, so that G0 is a commutative Lie group. As in the proof of Theorem 3.11, we now find a sub- group N5 ⊆ H5 isomorphic to Zl whose image in G5/K5 is a lattice. Dividing out this subgroup as in Example 3.8, we find an equivalent crossed module C6 where G6 is compact and H6 is still discrete. This proves our first statement. 5 is central, the quotient group G5/K5 Assume now that the commutator map in G factors through a continuous map G × G → H; since ∂ is injective, this just means that it is a map to H, and continuous as such. The crossed modules Ci constructed above inherit this property in each step. Thus the commutator map γ : G6 × G6 → G6 is a continuous map to H6. Since H6 is discrete, there is an open subset U ⊆ G6 with γ(x, y) = 1 for all x, y ∈ U , that is, xy = yx for all x, y ∈ U . This remains so for x, y in the subgroup G′ generated by U , which is open in G6 because U is open. Example 3.7 shows that G′ is part of a crossed module equivalent to C. This has Abelian compact G′, discrete H ′, and hence trivial c′. As a consequence, 3.14.3 implies 3.14.1. It is trivial that 3.14.1 implies 3.14.2. It remains to show that 3.14.3 follows, conversely, if C is equivalent to a crossed module C′ with trivial c′. Since C′ is again thin, this implies that G′ and hence coker ∂′ = π1(C′) is Abelian. Then π1(C) = G/∂(H) is Abelian as well because π1 is invariant under equivalence of crossed modules. This says that the commutator map of G factors through H; it does not yet give the continuity of the factorisation. We merely sketch how to prove this continuity. The commutator map of G′ is constant and hence clearly a continuous map to H ′. To finish the proof, we must show that the existence of a continuous commutator map G × G → H for thin crossed modules is invariant under equivalence of crossed modules. We turn the functor M : (G ⋊ H) × (G ⋊ H) → (G ⋊ H) on the arrow group- oids in (3.4) into a generalised morphism (bispace); two functors give isomorphic bispaces if and only if they are related by conjugation with a bisection. For a thin crossed module, a bisection that conjugates M onto M ◦ flip is exactly the same as a continuous commutator map G × G → H. Thus the property of having a continuous commutator map G × G → H is equivalent to the property that M 22 ALCIDES BUSS AND RALF MEYER and M ◦ flip are equivalent as generalised morphisms. This property is manifestly invariant under equivalence of crossed modules. (cid:3) Lemma 3.15. Let C = (G, H, ∂, c) be a thin crossed module with a compact subset K ⊆ H such that ∂(K) generates an open subgroup in G. Then C is equivalent to a crossed module C′ with trivial c′. Proof. Call a subgroup of a locally compact group compactly generated if it is generated by a compact subset. Our assumption is that there is a compactly generated subgroup A of H for which ∂(A) is open in G. If G1 ⊆ G is an open subgroup, then A1 := ∂−1(G1)∩A has the same property for the crossed module C1 constructed as in Example 3.7. If π : G2 → G is a quotient mapping for which ker π is compactly generated, then C2 constructed as in Example 3.8 also inherits this property, taking the preimage of A in A2, which is again finitely generated. The coverings needed in the proof of Theorem 3.14 have compactly generated kernels because we divide out either compact groups, connected Lie groups, or discrete subgroups in connected Lie groups, which are all compactly generated. Hence Theorem 3.14 provides an equivalent crossed module C′ with compact G′ and discrete H ′ and a compactly generated subgroup A ⊆ H ′ for which ∂(A) is dense in G′. Since H ′ is discrete, A is generated by a finite subset S. Since c′ is continuous, the set of g ∈ G′ with gh = hg for all h ∈ S is open in G′. Since S generates G′ topologically, this open subgroup is central in G′. Thus G′ has an open, finite-index centre Z. Now replace C′ by an equivalent crossed module C′′ as in Example 3.7 with G′′ = Z. This has commutative G′′ and hence trivial c′′ because ∂ is injective. (cid:3) Example 3.16. We construct a thin crossed module of locally compact groups where G/∂(H) is not Abelian. Since G/∂(H) = coker ∂ is invariant under equiva- lence, this crossed module cannot be equivalent to a commutative one. Let G1 be some finite group that is not Abelian. Let H := M G1, n∈N G := Y G1, n∈N where L G1 is the subgroup of all (gn) ∈ Q G1 with gn = 1 for all but finitely many entries. Hence H is a countable group; we give it the discrete topology. The group G is pro-finite and in particular compact. It contains H as a dense normal subgroup, and the conjugation map g 7→ ghg−1 for any fixed h ∈ H factors through a finite product and therefore is continuous. The constant embedding G1 → G remains an embedding into G/H. Hence G/H is not Abelian. Some of our simplifications also work for crossed modules that are not thin. We give one such statement: Proposition 3.17. Let C = (G, H, ∂, c) be a crossed module of Lie groups with connected G and injective ∂. Then C is equivalent to a crossed module C′ = (G′, H ′, ∂′, c′) where G′ is simply connected, H ′ is discrete, and c′ is trivial. Two crossed modules of this form are equivalent if and only if they are isomorphic. CROSSED PRODUCTS FOR CROSSED MODULES 23 Proof. As in the proof of Theorem 3.5, we may pass to an equivalent crossed module C1 where G1 is the universal covering of G; in this new crossed module, the image of the connected component of H1 is closed. Since ∂ is assumed injective, we may divide out this connected component and arrive at an equivalent crossed module with discrete H2 and simply connected G2. As in the proof of Theorem 3.5, it follows that the conjugation action on H2 is trivial. Hence C2 has the asserted properties. Now consider the invariant (g/h, T (C)) under equivalence used already in the proof of Theorem 3.5. If C1 and C2 are two crossed modules with discrete H, simply connected G and injective ∂, then an isomorphism g1/h1 → g2/h2 is an isomorphism g1 → g2 because H1 and H2 are discrete. This lifts to an isomor- phism G1 → G2 because G1 and G2 are simply connected. Furthermore, we claim that the exponential maps gi → Gi map T (Ci) onto ∂(Hi). This is be- cause ∂(Hi) is contained in the centre of Gi, the centre of Gi is isomorphic to Rn for some n because Gi is simply connected, and the exponential map for Rn is surjective. Therefore, an isomorphism between the invariants (g1/h1, T (C1)) and (g2/h2, T (C2)) already implies that C1 and C2 are isomorphic crossed modules. (cid:3) 4. Duality for Abelian crossed modules Definition 4.1. We call a crossed module (G, H, ∂, c) 2-Abelian if the action c is trivial, and Abelian if c is trivial and G is Abelian. For a 2-Abelian crossed module, H is Abelian because hkh−1 = c∂(h)(k) for all h, k ∈ H, and ∂(H) is a central subgroup of G because ∂(cg(h)) = g∂(h)g−1. For a thin crossed module, H is Abelian if and only if c is trivial, if and only if G is Abelian. Hence, for thin crossed modules, Abelianness and 2-Abelianness are equivalent conditions. Of course, in general a 2-Abelian crossed module need not be Abelian (for instance, any group G viewed as a crossed module (G, 0, 0, 0) is 2-Abelian). Proposition 3.17 implies that any crossed module of Lie groups with connected G and injective ∂ is equivalent to a 2-Abelian one. Let C = (G, H, ∂, c) be a 2-Abelian crossed module of locally compact groups. Let (A, υ) be a Fell bundle over C. Condition (2) in Definition 2.10 now says that a · υh = υh · a for all a ∈ A, h ∈ H. This extends to a ∈ Cc(A) and then to a ∈ C∗(A); here we embed υh into the multiplier algebra of C∗(A) using the universal representation ρu. Thus ρu now maps H to the centre of M(C∗(A)). This map integrates to a nondegenerate ∗-homomorphism (4.2) ∫ ρu : C∗(H) → ZM(C∗(A)). Identifying C∗(H) with C0( H) for the Pontryagin dual H of H, we get a structure of C0( H)-algebra on C∗(A). (We normalise the Fourier transform so that the unitary δh ∈ UM(C∗(H)) becomes the function h 7→ h(h) on H.) Proposition 4.3. The crossed product C∗(A, υ) is the fibre at 1 ∈ H for the C0( H)-C∗-algebra structure on C∗(A) just described. 24 ALCIDES BUSS AND RALF MEYER Proof. By construction, C∗(A, υ) is the quotient by the relation ρu(υh) = 1 for all h ∈ H. This is equivalent to ∫ ρu(f ) = f (1 H ) for all f ∈ C0( H) ∼= C∗(H). Hence C∗(A, υ) is the quotient by C0( H \ {1}) · C∗(A), that is, the fibre at 1. (cid:3) The above proposition is merely an observation, the main point is to see that C∗(A) has a relevant C0( H)-algebra structure. This proposition allows us to com- pute crossed products by 2-Abelian crossed modules in two more elementary steps: first take the crossed product C∗(A) by the locally compact group G; then take the fibre at 1 ∈ H for the canonical C0( H)-algebra structure on C∗(A). Example 4.4. Let θ ∈ R \ Q and let C∗(Tθ) be the associated noncommutative torus. It carries a strict action by the Abelian crossed module ∂ : Z → T with ∂(n) := exp(2πiθn); namely, T acts by part of the gauge action: αz(U ) := zU , αz(V ) := V , where U and V are the standard generators of C∗(Tθ), and Z acts by n 7→ V −n (see [1]). The crossed product for this action is already computed in [1]: it is the C∗-algebra of compact operators on the Hilbert space L2(T). This also follows from Proposition 4.3. First, the crossed product C∗(Tθ) ⋊ T turns out to be C(T) ⊗ K(L2T) because the T-action on C∗(Tθ) is a dual action if we interpret C∗(Tθ) as C(T) ⋊ Z with C(T) = C∗(V ∗) and U generating Z. The map C∗(Z) → C∗(Tθ) ⋊ T becomes an isomorphism onto C(T) ⊗ 1. Hence the fibre at 1 ∈ Z = T gives K(L2T) as expected. Next we generalise Takesaki -- Takai duality from Abelian groups to Abelian crossed modules. We first reformulate the classical statement in our setting of correspondence 2-categories. Let G be an Abelian locally compact group and let G be its Pontryagin dual. A crossed product for a G-action carries a canonical dual action of G and a crossed product for a G-action carries a canonical dual action of G. These constructions extend to map G-equivariant correspondences to G-equivariant correspondences, and vice versa, such that equivariant isomorphism of correspondences is preserved. That is, they provide functors Corr(G) ↔ Corr( G). Takesaki -- Takai duality implies that these functors are equivalences of 2-categories inverse to each other. Going back and forth gives equivariant isomorphisms A ⋊ G ⋊ G ∼= A ⊗ K(L2G) and A ⋊ G ⋊ G ∼= A ⊗ K(L2 G), respectively; we interpret these as equivariant Morita -- Rieffel equivalences A ⋊ G ⋊ G ≃ A and A ⋊ G ⋊ G ≃ A via A ⊗ L2G and A ⊗ L2 G, respectively. These equivariant Morita -- Rieffel equivalences are natural with respect to equivariant correspondences, hence the functors we get by composing the crossed product functors are naturally equivalent to the identity functors. Now let C = (G, H, ∂, 0) be an Abelian crossed module of locally compact groups; this is nothing but a continuous homomorphism between two locally com- Its dual crossed module C consists of the dual groups G pact Abelian groups. and H and the transpose ∂ : H → G of ∂. The bidual of C is naturally isomorphic to C because this holds for locally compact groups. Our duality theorem does not compare the action 2-categories of the crossed modules C and C, however; instead, the action 2-category of the arrow groupoid H ⋊ G of C appears. And the functors going back and forth are the crossed product functors for the groups G and G. CROSSED PRODUCTS FOR CROSSED MODULES 25 We now describe the equivalence of 2-categories Corr(C) ≃ Corr( H ⋊ G). Let (A, υ) be a saturated Fell bundle over C, encoding a C-action by correspondences. The cross-sectional C∗-algebra C∗(A) carries a dual action of G, where a character acts by pointwise multiplication on sections. Since Abelian crossed modules are 2-Abelian, (4.2) provides a nondegenerate ∗-homomorphism ∫ ρu : C0( H) ∼= C∗(H) → ZM(C∗A). Since υh is a multiplier of degree ∂(h), the dual action α of G acts on δh ∈ C∗(H) by αg(δh) = g(∂h) · δh. This defines an action of G on C∗(H), which corresponds to the action ρgf (h) = f (h · ∂(g)) on C0( H) coming from the right translation action of G on H through the homomorphism ∂ : G → H. Thus C∗(A) carries a strict, continuous action of the arrow groupoid H ⋊ G of C. The actions of H ⋊ G form a 2-category Corr( H ⋊ G) with actions by corre- spondences as objects, equivariant correspondences as arrows and isomorphisms of equivariant correspondences as 2-arrows. We claim that the construction on the level of objects above is part of a functor Corr(C) → Corr( H ⋊ G). Let E be a correspondence of Fell bundles (A, υA) → (B, υB). The space Cc(E) of compactly supported, continuous sections of E is a pre-Hilbert module over the ∗-algebra Cc(B) with a nondegenerate left action of Cc(A). Using the embedding Cc(B) ⊆ C∗(B), we may complete Cc(E) to a Hilbert C∗(B)-module C∗(E). The left action of Cc(A) on Cc(E) induces a left action of C∗(A) on C∗(E), turning it into a correspondence C∗(A) → C∗(B). Isomorphic correspondences of Fell bundles clearly give isomorphic correspondences C∗(A) → C∗(B). Thus taking cross-sectional C∗-algebras gives a functor Corr(C) → Corr (where Corr = Corr({1}) denotes the 2-category of C∗-algebras with correspondences as their morphisms); actually, since we did not use the H-action, this is just the crossed product functor Corr(G) → Corr, composed with the forgetful functor Corr(C) → Corr(G). Pointwise multiplication by characters defines a G-action on Cc(E). This ex- tends to the completion C∗(E), and turns it into a G-equivariant correspondence C∗(A) → C∗(B). This G-action is natural for isomorphisms of Fell bundle corre- spondences. Furthermore, C∗(E) is a C∗(H)-linear correspondence because υA h ·ξ = h for all h = cg(h) ∈ H by 2.13.(7). Thus taking the cross-sectional C∗-algebra ξ · υB (that is, the crossed product by G) gives a functor Corr(C) → Corr( H ⋊ G). Theorem 4.5. Let C be an Abelian crossed module. The functor Corr(C) → Corr( H ⋊ G) just described is an equivalence of 2-categories. Proof. To prove the existence of a quasi-inverse functor Corr( H ⋊ G) → Corr(C), it suffices to construct it for strict actions by automorphisms because arbitrary actions by correspondences are equivalent to strict actions by automorphisms (see [2, Theorem 5.3]), where "equivalent" means "isomorphic in the 2-category Corr( H ⋊ G)." Thus let B be a C∗-algebra with a continuous action of H ⋊ G in the usual sense. This consists of a strict action β of the group G and a G-equivariant nondegenerate ∗-homomorphism from C0( H) to ZM(B). The crossed product B ⋊ G carries a dual action β of G. For h ∈ H, define vh ∈ UM(C0( H)) by vh(h) = h(h)−1 for 26 ALCIDES BUSS AND RALF MEYER all h ∈ H. The right translation action of G acts on vh by g(vh) = g(∂(h))−1 · vh because g(vh)(h) = vh(h · ∂g) = ( ∂g)(h)−1h(h)−1 = g(∂h)−1vh(h) for all h ∈ H. Now map vh to UM(B⋊ G) using the homomorphism UM(C0( H)) → UM(B ⋊ G) induced by the C0( H)-C∗-algebra structure on B. These unitaries vh ∈ UM(B ⋊ G) commute with B because C0( H) is mapped to the centre of B, and they satisfy vhδgv∗ h = δg g(∂h) · vhv∗ h = g(∂h) · δg for g ∈ G. Hence β∂(h) = Ad(vh) for all h ∈ H. Since βg(vh) = vh for all g ∈ G, h ∈ H, the map h 7→ vh and the dual action β of G combine to a strict action of the crossed module C on B ⋊ G. The above constructions extend to equivariant correspondences in a natural way and thus provide a functor Corr( H ⋊ G) → Corr(C). We claim that this is quasi-inverse to the functor Corr(C) → Corr( H ⋊ G) constructed above. We must compose these functors in either order and check that the resulting functors are equivalent to the identity functors. Since all actions by correspondences are equiv- alent to strict actions by automorphisms, it is enough to verify the equivalence on those objects in Corr(C) and Corr( H ⋊ G) that are strict actions by automorphisms. Let α : G → Aut(A) and u : H → UM(A) be a strict action of C on A. Our functor to Corr( H ⋊ G) maps it to A ⋊ G equipped with the dual action of G and with the canonical map C0( H) ∼= C∗(H) → ZM(A ⋊ G) induced by the represen- tation h 7→ u∗ hδ∂(h) for h ∈ H; this is how the translation from strict actions to Fell bundles works on the level of the unitaries uh and υh (see Example 2.12). The map to actions of C takes this to (A ⋊ G) ⋊ G equipped with the dual action of G and the homomorphism H → UM(A ⋊ G ⋊ G), h 7→ (u∗ hδ∂(h))−1 ∈ UM(A ⋊ G) ⊆ UM(A ⋊ G ⋊ G) because the isomorphism C0( H) ∼= C∗(H) maps vh 7→ δh−1 . Takesaki -- Takai duality provides a canonical G-equivariant isomorphism A ⋊ G ⋊ G ∼= A ⊗ K(L2G) ∼= K(A ⊗ L2G). It extends the standard representation of A ⋊ G on A ⊗ L2G that maps a ∈ A to the operator ϕ(a) of pointwise multiplication by G ∋ g 7→ αg(a) and g ∈ G to the right translation operator ρgf (g′) = f (g′g) for all g′ ∈ G. Hence (u∗ hδ∂(h))−1 = δ∂(h)−1uh acts by the unitary operator (ρ∂(h)−1ϕ(uh)f )(g) = uh · f (g∂(h)−1) = uh · f (∂(h)−1g) because G is Abelian and αg(uh) = uh for all g ∈ G, h ∈ H. This gives the operator uh ⊗ λ∂(h) for the left regular representation λ on G. The imprimitivity bimodule A ⊗ L2G between K(A ⊗ L2G) ∼= A ⋊ G ⋊ G and A with the G-action g 7→ αg ⊗ λg is C-equivariant and thus provides a C-equivariant equivalence A ≃ A ⋊ G ⋊ G. Now let B carry a strict H ⋊ G-action. That is, G acts on B via an action β, and we have a nondegenerate G-equivariant ∗-homomorphism φ : C0( H) → ZM(B). Our functor to Corr(C) takes this to B ⋊ G equipped with the dual action of G and the homomorphism H → UZM(B ⋊ G) defined as the composite of the rep- resentation v : H → ZM(B), h 7→ φ(vh), with vh(h) := h(h)−1 and the canonical CROSSED PRODUCTS FOR CROSSED MODULES 27 embedding B → M(B ⋊ G). The functor that goes back to an H ⋊ G-action now gives B ⋊ G ⋊ G with the dual action of G and with the nondegenerate ∗-homomorphism C∗(H) → ZM(B ⋊ G ⋊ G) that maps h ∈ H to φ(v∗ h)δ∂(h) -- here φ(vh) is viewed as an element of M(B ⋊ G ⋊ G) using the canonical ho- momorphism B → M(B ⋊ G ⋊ G). Now we identify C0( H) ∼= C∗(H) as before, h to δh. Thus the resulting map C0( H) → ZM(B ⋊ G⋊G) mapping the character v∗ h)δ∂(h). On the other hand, via the isomorphism C0( H) ∼= C∗(H), maps v∗ the homomorphism φ : C0( H) → M(B) corresponds to the integrated form of the h). The G-equivariance of φ means that representation H → M(B), h 7→ φ(v∗ h) (recall that the G-action on C0( H) is induced by the βg(φ(v∗ right translation G-action on H via (h, g) 7→ h · ∂(g)). h)) = g(∂(h)) · φ(v∗ h to φ(v∗ Takesaki -- Takai duality gives a G-equivariant isomorphism B ⋊ G ⋊ G ∼= B ⊗ K(L2 G) ∼= K(B ⊗ L2 G). It restricts to the standard representation of B ⋊ G on B ⊗ L2 G, where b ∈ B acts by b · f (g) = βg(b)f (g) and g ∈ G acts by right translation; the representation of G lets g ∈ G act by g · f (g) = g(g)−1f (g) for all f ∈ Cc( G, B) ⊆ B ⊗ L2 G. Since βg(φ(v∗ h), we get h)) = g(∂(h)) · φ(v∗ (φ(v∗ h)δ∂(h)) · f (g) = g(∂(h)) · φ(v∗ h) · g(∂(h))−1 · f (g) = φ(v∗ h) · f (g). This means that the G-equivariant B⋊ G⋊G-B-equivalence bimodule B⊗L2 G also intertwines the representations of H (and hence the representations of C0( H)) on B ⋊ G ⋊ G and B. In other words, B ⊗ L2 G is an H ⋊ G-equivariant Morita -- Rieffel equivalence between B and B ⋊ G ⋊ G. (cid:3) Despite Theorem 4.5, there is an important difference between Corr(C) and Corr( H ⋊ G). Both 2-categories come with a natural tensor product structure: take the diagonal action on the tensor product in Corr(C), or the diagonal action on the tensor product over the base space H in Corr( H ⋊ G). These tensor products are quite different. In terms of Corr( H ⋊ G), the natural tensor product in Corr(C) does the following. Take two C∗-algebras A1 and A2 with actions of H ⋊ G. Their tensor product is a C∗-algebra over H × H with a compatible action of the group G × G. Restrict the group action to the diagonal (this gives the usual diagonal action). But instead of restricting the C∗-algebra to the diagonal in H × H as usual, give it a structure of C0( H)-C∗-algebra using the comultiplication C0( H) → C0( H × H). Remark 4.6. There is a more symmetric form of our duality where both partners in the duality are of the same form: both are length-two chain complexes of locally compact Abelian groups H d−→ G d−→ K, with the dual of the form K → G → H. An action of such a complex consists of an action of the crossed module H → G and an action of the transformation groupoid K ⋊ G, where both actions contain the same action of the group G; here the groupoid K ⋊ G is the transformation groupoid for the translation action k · g := k · d(g) of G on K. Actually, these actions are the actions of the 2-groupoid 28 ALCIDES BUSS AND RALF MEYER with object space K, arrows g : k → k · d(g), and 2-arrows H : g ⇒ g · d(h); the chain complex condition d2 = 0 ensures that this is a strict 2-groupoid. Our proof above shows that the 2-groupoids H → G → K and K → G → H have equivalent action 2-categories on C∗-algebras. Namely, an action of the crossed module from H → G on A induces an action of the groupoid from G → H on A ⋊ G, and an action of the groupoid from G → K induces an action of the crossed module from K → G on A ⋊ G. Since both constructions involve the same dual action of G, we get an action of K → G → H on A ⋊ G. Now the functor backwards has exactly the same form, and going back and forth is still a stabilisation functor. Since the stabilisation is compatible with the crossed module and groupoid parts of our actions, it is compatible with their combination to a 2-groupoid action. In Section 3 we have studied equivalence of general crossed modules via homo- morphisms. We end this section by a criterion for equivalences between Abelian crossed modules. Proposition 4.7. Let Ci = (Gi, Hi, ∂i, 0), i = 1, 2, be Abelian crossed modules and let (ϕ, ψ) : C1 → C2 be a homomorphism. The following are equivalent: (1) the homomorphism (ϕ, ψ) is an equivalence; (2) the diagram H1 ι1−→ G1 × H2 π2−→ G2 is an extension of locally compact Abelian groups, where ι1 := (∂−1 1 , ψ) : H1 → G1 × H2, π2 : G1 × H2 → G2, h1 7→ (∂1(h1)−1, ψ(h1)), (g1, h2) 7→ ϕ(g1)∂2(h2); (3) the dual diagram cH1 bι1←− cG1 × cH2 bπ2←− cG2 is an extension of locally compact Abelian groups; (4) the dual homomorphism ( ψ, ϕ) : C2 → C1 is an equivalence. Proof. We shall use Lemma 3.3. Since inversion on a topological group is a home- omorphism, the map in condition (1) in Lemma 3.3 is a homeomorphism if and only if ι1 is a homeomorphism onto ker π2. The map π2 is the same one appearing in condition (2) in Lemma 3.3. This is a homomorphism since all groups involved are Abelian, and its kernel is exactly the image of ι1. Therefore, the assump- tions in Lemma 3.3 hold if and only if H1 → G1 × H2 → G2 is a topological group extension. Since taking duals preserves this property, this is equivalent to cH1 ← cG1 × cH2 ← cG2 being a topological group extension. As above, this is equivalent to ( ψ, ϕ) being an equivalence. (cid:3) 5. Crossed products for crossed module extensions Now we come to the factorisation of the crossed product functor for an extension of crossed modules. Let Ci = (Gi, Hi, ∂i, ci) for i = 1, 2, 3 be crossed modules of locally compact groups. CROSSED PRODUCTS FOR CROSSED MODULES 29 Definition 5.1. A diagram C1 → C2 → C3 of homomorphisms of crossed modules is called a strict extension of crossed modules if the resulting diagrams H1 ψ1−−→ H2 ψ2−−→ H3 and G1 ϕ1−→ G2 ϕ2−→ G3 are extensions of locally compact groups. That is, ψ1 is a homeomorphism onto the kernel of ψ2 and ψ2 is an open surjection, and similarly for ϕ1 and ϕ2. Theorem 5.2. Let C1 ֌ C2 ։ C3 be a strict extension of crossed modules and let A be a C∗-algebra with an action of C2 by correspondences. Then A ⋊ C1 carries a canonical action of C3 by correspondences such that (A ⋊ C1) ⋊ C3 is naturally isomorphic to A ⋊ C2. Proof. Since strict actions by automorphisms are notationally simpler, we first prove the result in case A carries a strict action by automorphisms. Then we reduce the general case to this special case. A strict action by automorphisms is given by group homomorphisms α : G2 → Aut(A) and u : H2 → UM(A) that satisfy α∂2(h) = Aduh for all h ∈ H2 and αg(uh) = ucg(h) for all g ∈ G2, h ∈ H2. To simplify notation, we also view G1 and H1 as subgroups of G2 and H2, respectively, so that we drop the maps ϕ1 and ψ1. The crossed product A ⋊ C1 is a quotient of the crossed product A ⋊ G1 for the group G1 by the ideal generated by the relation uh ∼ δ∂1(h) for all h ∈ H1; that is, we divide A ⋊ G1 by the closed linear span of the subset {x · (uh − δ∂1(h)) · y : h ∈ H1, x, y ∈ A ⋊ G1}. A canonical action γ′ of G2 on A ⋊ G1 is defined by (γ′ g2 f )(g1) := αg2 (f (g−1 2 g1g2)) for g2 ∈ G2, g1 ∈ G1, f ∈ Cc(G1, A) on the dense subalgebra Cc(G1, A); this extends to the C∗-completion. The action c2 of G2 on H2 leaves H1 ⊆ H2 = ker ψ2 invariant because ψ2 is G2-equivariant. Since γ′ g for g ∈ G2 maps uh − δ∂1(h) to ucg(h) − δ∂1(cg(h)) and cG2(H1) ⊆ H1, the action γ′ descends to an action γ of G2 on A ⋊ C1. Let Ug for g ∈ G1 be the image of δg ∈ UM(A ⋊ G1) in UM(A ⋊ C1); this defines a homomorphism G1 → UM(A ⋊ C1) with γg = Ad(Ug) all g ∈ G1. Also let Uh ∈ UM(A ⋊ C1) for h ∈ H2 be the image of uh ∈ UM(A) under the canonical map A → M(A ⋊ G1) → M(A ⋊ C1). We claim that γ∂2(h) = Ad(Uh) for all h ∈ H2. To see this, we notice first that c2,g(h)h−1 ∈ H1 = ker ψ2 for g ∈ G1 because ϕ2(g) = 1 implies ψ2(c2,g(h)h−1) = ψ2(c3,ϕ2(g)(h)h−1) = 1. Hence Uc2,g−1 (h)h−1 = δ∂1(c2,g−1 (h)h−1) holds in UM(A ⋊ C1). This implies UhδgU ∗ h = δgUc2,g−1 (h)Uh−1 = δgUc2,g−1 (h)h−1 = δgδ∂1(c2,g−1 (h)h−1) = δg∂2(c2,g−1 (h))∂2(h)−1) = δ∂2(h)g∂2(h)−1 . h = α∂2(h)(a) for all a ∈ A, we get UhxU ∗ h = γ∂2(h)(x) Since we have assumed uhau∗ for all x ∈ A ⋊ C1. If g ∈ G1, h ∈ H2, then UgUhU ∗ g is the image of αg(uh) = uc2,g(h), that is, UgUhU ∗ g = Uc2,g(h). Thus the map (g, h) 7→ UgUh is a homomorphism G1 ⋉ H2 → UM(A ⋊ C1) where the semidirect product uses the action G1 ⊆ G2 → Aut(H2) 30 ALCIDES BUSS AND RALF MEYER given by restricting c2. Since Uh = U∂1(h) for h1 ∈ H1 ⊆ H2, we get a homo- morphism on H := G1 ⋉ H2/∆(H1) with the embedding ∆ : H1 → G1 ⋉ H2, h 7→ (∂(h)−1, h). The map G1 ⋉ H2 → G2, (g, h) 7→ g · ∂2(h), is a group ho- momorphism which vanishes on ∆(H1) and hence descends to a group homomor- phism ∂ : H → G2. We define a homomorphism c′ : G2 → Aut(G1 ⋉ H2) by g2 (g1, h2) := (g2g1g−1 c′ 2 , c2,g2 (h2)). This leaves ∆(H1) invariant and hence de- scends to a homomorphism c : G2 → Aut(H). Putting all this together gives a crossed module C := (G2, H, ∂, c) which acts on A ⋊ C1 by γ : G2 → Aut(A ⋊ C1) and U : H → UM(A ⋊ C1). ∼= H2/H1 The homomorphism ∂ maps G1 ⊆ H homeomorphically onto the closed normal subgroup G1 ⊆ G2. We have G2/∂(G1) ∼= G3 and H/G1 ∼= H3. Exam- ple 3.8 shows that C is equivalent to the crossed module C3. By Theorem 3.5, the action (γ, U ) of C on A ⋊ C1 is equivalent to an action of C3 on A ⋊ C1 by correspon- dences such that (A ⋊ C1) ⋊ C ∼= (A ⋊ C1) ⋊ C3. We claim that (A ⋊ C1) ⋊ C ∼= A ⋊ C2. By the universal property a morphism (A ⋊ C1) ⋊ C → D is equivalent to a C-covariant representation of A ⋊ C1 in M(D), that is, a morphism ρC : A ⋊ C1 → M(D) and a continuous homomorphism V : G2 → UM(D) with VgρC (c)V ∗ g = ρC (γg(c)) for all c ∈ A ⋊ C1, g ∈ G2 and V∂(h) = ρC(Uh) for all h ∈ H. By the universal property of A ⋊ C1, the representation ρC is equivalent to a morphism ρA : A → UM(D) and a continuous homomorphism W : G1 → UM(D) with WgρA(a)W ∗ g = ρA(αg(a)) for all a ∈ A, g ∈ G1 and W∂1(h) = ρA(uh) for all h ∈ H1. The assumptions on ρC are equivalent to VgρA(a)V ∗ g = ρA(αg(a)) for a ∈ A, g ∈ G2; Vg2 Wg1 V ∗ for g2 ∈ G2, g1 ∈ G1; Vg = Wg for g ∈ G1; and V∂2(h) = ρA(uh) for h ∈ H2. Thus the unitaries Wg for g ∈ G1 are redundant, and the conditions on ρA and the unitaries Vg for g ∈ G2 are precisely those for a covariant representation of A and C2. Hence the morphisms (A ⋊ C1) ⋊ C → M(D) are in natural bijection with morphisms A ⋊ C2 → M(D). This shows that A ⋊ C2 ∼= (A ⋊ C1) ⋊ C. Since (A ⋊ C1) ⋊ C ∼= (A ⋊ C1) ⋊ C3, this gives the desired isomorphism. g2 = Wg2g1g−1 2 We must show that the C3-action on A ⋊ C1 is natural, so that A 7→ (A ⋊ C1) ⋊ C3 is a functor. Here we still talk about functors defined on the full sub-2-category of Corr(C2) consisting of strict actions by automorphisms. The naturality of the C3-action is equivalent to the naturality of the C-action by Theorem 3.5, which is what we are going to prove. A C2-transformation between two strict actions on A1 and A2 by automorphisms is equivalent to a G-equivariant correspondence E from A1 to A2 in the usual sense, subject to the extra requirement uA1 for all h ∈ H2, ξ ∈ E. For such a correspondence, we get an action of G2 on the induced correspondence E ⋊ C1 from A1 ⋊ C1 to A2 ⋊ C2 by the same formulas as above, and this yields a C-equivariant correspondence from A1 ⋊ C1 to A2 ⋊ C2. Furthermore, isomorphic C2-equivariant correspondences induce isomorphic C-equivariant correspondences. Hence the C-action on A ⋊ C1 is natural on the 2-category of strict actions of C2 by automorphisms. h · ξ = ξ · uA2 h The isomorphism A ⋊ C2 → (A ⋊ C1) ⋊ C is natural in the sense that for any C2-equivariant correspondence E from A1 to A2, the square formed by the isomorphisms above and the induced correspondences A1 ⋊ C2 → A2 ⋊ C2 and CROSSED PRODUCTS FOR CROSSED MODULES 31 (A1 ⋊ C1) ⋊ C3 → (A2 ⋊ C1) ⋊ C3 commutes up to a canonical isomorphism of corre- spondences. This establishes the naturality of our isomorphism on the 2-category of strict actions of C2. By the Packer -- Raeburn Stabilisation Trick, any action of C2 by correspondences is equivalent to a strict C2-action by automorphisms ([2, Theorem 5.3]), where equivalence means an isomorphism (that is, equivariant Morita equivalence) in the 2-category Corr(C2). This equivalence means that a functor defined only on the subcategory of strict C2-actions may be extended to a functor on all of Corr(C2); all such extensions are naturally isomorphic; and a natural transformation between functors on the subcategory extends to a natural transformation between the ex- tensions. Hence the result for strict actions proves the more general result for actions by correspondences by abstract nonsense. (cid:3) Example 5.3. Let G be a locally compact group and let N be a closed normal sub- group of G so that we get a group extension N ֌ G ։ G/N . Viewing N , G and G/N as crossed modules C1 = (N, 0, 0, 0), C2 = (G, 0, 0, 0) and C3 = (G/N, 0, 0, 0), respectively, our result says that given an action α of G on a C∗-algebra A by correspondences, there is an action β of G/N on A ⋊α N by correspondences, where α denotes the restriction of α to N , such that A ⋊α G ∼= (A ⋊α N ) ⋊β G/N. We may also interpret everything in terms of Fell bundles: the action of G on A corresponds to a Fell bundle A over G with unit fibre A1 = A in such way that A⋊α G is (isomorphic to) the cross-sectional C∗-algebra C∗(A). The restricted crossed product A ⋊α N corresponds to the cross-sectional C∗-algebra C∗(AN ) of the restriction AN of A to N . Our theorem says that there is a Fell bundle B over G/N with unit fibre B1 = A ⋊α N ∼= C∗(AN ) such that C∗(B) ∼= C∗(A). Although our C∗-algebraic version appears to be new, a version for L1-cross-sectional algebras in proved by Doran and Fell in [3, VIII.6]. Even if we start with a strict action of G on A by automorphisms, the induced It action of G/N on A ⋊ N will usually not be an action by automorphisms. may be interpreted as a Green twisted action of (G, N ) on A ⋊ N , and the above decomposition corresponds to Green's decomposition of crossed products: A ⋊α G ∼= (A ⋊ N ) ⋊ (G, N ) (see [4, 7]). We may weaken the notion of strict extension by replacing the crossed modules involved by equivalent ones. We mention only one relevant example of this. Example 5.4. Let C = (G, H, ∂, c) be a crossed module. Let G2 := G ⋉c H be the semidirect product group. It contains H2 := H as a normal subgroup via ∂2 : H2 → G2, h 7→ (1, h), with quotient G2/H2 ∼= G. Let c2 : G2 → Aut(H2) be the result- ing conjugation action, c2,(g,h)(k) := cg(hkh−1). Then C2 = (G2, H2, ∂2, c2) is a crossed module of locally compact groups that is equivalent to (G, 0, 0, 0). Hence actions of C2 are equivalent to actions of the group G, with the same crossed prod- ucts on both sides (Theorem 3.5). Let C1 = (H, 0, 0, 0) be the group H turned into a crossed module and let C3 = C. We map C2 → C by ψ2 = Id : H → H and ϕ2 : G ⋉c H → H, (g, h) 7→ g · ∂(h). This is a homomorphism of crossed modules, and ψ2 and ϕ2 are open surjections. Their kernels are isomorphic to H1 := 0 32 ALCIDES BUSS AND RALF MEYER and G1 := H via ϕ1 : H → G ⋉c H, h 7→ (∂(h)−1, h), respectively. Thus we get a strict extension of crossed modules C1 → C2 → C with C1 = (H, 0, 0, 0) and C2 ≃ (G, 0, 0, 0). Hence the group G is equivalent to an extension of the group H by the crossed module C. Now let A carry an action of G, which we turn into an action of C2 via G ⋉ H → G, (g, h) 7→ g∂(h). When we apply Theorem 5.2 to this situation, we get back [1, Theorem 1]: A ⋊ G ∼= (A ⋊ H) ⋊ C. Example 5.5. Let θ be some irrational number and define an embedding θ : Z → R by n 7→ θn. Let C = (R, Z, θ, 0) be the resulting Abelian crossed module. This is equivalent to the group T (viewed as the crossed module (T, 0, 0, 0)) via the homomorphism (ϕ, ψ) : C → T with ϕ : R → T, t 7→ exp(2πiθt), and the trivial homomorphism ψ : Z → 0 (Example 3.9). Theorem 3.5 gives an equivalence of 2-categories Corr(T) ∼−→ Corr(C); it sends a T-algebra A to itself with Z acting trivially and R via ϕ and the given T-action. Now let C′ = (T, Z, ∂, 0) be the crossed module considered in Example 4.4, where ∂(n) = exp(2πiθn). View the group Z as a crossed module. There is an extension Z ֌ C ։ C′ described by the diagram: 0 Z θ Z R Id Id ϕ Z T ∂ Therefore, Theorem 5.2 gives a functor Corr(C) → Corr(C′) that sends a C-algebra A to the (restricted) crossed product A ⋊ Z with an induced C′-action, such that A ⋊ C ∼= (A ⋊ Z) ⋊ C′. Composing this with the equivalence Corr(T) ∼= Corr(C) we obtain a functor Corr(T) → Corr(C′) that sends a T-algebra A to A ⋊ Z with a C′-action such that (A ⋊ Z) ⋊ C′ ∼= A ⋊ T. As a simple example, we take the T-algebra C(T) with translation T-action. In this case, Z acts by irrational rotation by multiples of θ so that C(T) ⋊ Z ∼= C∗(Tθ) is the noncommutative torus and the induced C′-action is the same one considered in Example 4.4. Hence we get once again that C∗(Tθ) ⋊ C′ ∼= C(T) ⋊ T ∼= K(L2T). Theorem 4.5 shows that Corr(C) ∼= Corr( R ⋉ Z) ∼= Corr(R ⋉ T), where R ⋉ T denotes the transformation groupoid for the action t · z = exp(2πiθt)z for t ∈ R and z ∈ T. Composing Corr(T) ∼−→ Corr(C) with this equivalence, we get a functor Corr(T) ∼−→ Corr(R ⋉ T). This takes a T-algebra, views it as a C-algebra, and sends it to the crossed product A ⋊ R viewed as an R ⋉ T-algebra using the dual R-action and the structure of C(T)-algebra given by the homomorphism C(T) ∼= C∗(Z) → C∗(R) → M(A ⋊ R), which maps C(T) into the centre of M(A). Theorem 4.5 also gives Corr(C′) ∼= Corr(T ⋉ Z) ∼= Corr(Z ⋉ T), where Z acts by rotation by multiples of θ. The quotient map C → C′ becomes the forgetful functor that restricts an R ⋉ T-action to a Z ⋉ T-action on θZ ⊆ R. CROSSED PRODUCTS FOR CROSSED MODULES 33 6. Factorisation of the crossed product functor Now we put our results together to factorise the crossed product functor Corr(C) → Corr for a crossed module C = (G, H, ∂, c) of locally compact groups into "ele- mentary" constructions. First we give more details on the strict extensions that decompose C into simpler building blocks. The image ∂(H) is a normal subgroup in G because g∂(h)g−1 = ∂cg (h). Hence ∂(H) is a closed normal subgroup in G and is a locally compact group. The closed subgroup ¯π1(C) := G/∂(H) π2(C) := ker ∂ ⊆ H is Abelian because hkh−1 = c∂(h)(k) for all h, k ∈ H. Since π2(C) is an Abelian locally compact group, there is a crossed module C1 with H1 = π2(C) and trivial G1 (and hence trivial ∂1 and c1). The G-action c on H leaves π2(C) invariant and hence descends to an action c2 of G on H2 := H/π2(C). Of course, ∂ descends to a map ∂2 : H2 → G2 = G. This defines a crossed module of locally compact groups C2. The canonical maps C1 → C → C2 are homomorphisms of crossed modules, and they clearly form a strict extension of crossed modules, based on the extensions of locally compact groups π2(C) ֌ H ։ H/π2(C) and 0 ֌ G = G. There is a crossed module C3 with G3 := ∂(H), H3 = H2 = H/π2(C), and ∂3 : H3 → G3 and c3 : G3 → Aut(H3) induced by ∂ and c. Let C4 be the crossed module with G4 = ¯π1(C) and trivial H4, ∂4 and c4. The obvious maps give homomorphisms of crossed modules C3 → C2 → C4; these form a strict extension of crossed modules because we have extensions H3 = H2 ։ 0 and G3 ֌ G2 ։ G4 of locally compact groups. The strict extensions above show that the crossed product functor for C-actions factorises into the three crossed product functors with C1, C3 and C4. Now we analyse actions and crossed products for C1, C4, and C3, respectively. Since a crossed module C1 of the form (0, H1, 0, 0) is Abelian, Theorem 4.5 shows that Corr(C1) is equivalent to the 2-category of C0(cH1)-C∗-algebras, such that the crossed product by C1 corresponds to the functor that maps a C0(cH1)-C∗-algebra to its fibre at 1 ∈ cH1. The case at hand is much easier than the general case of Theorem 4.5 because G is trivial. We simply observe that a C1-action on A is exactly the same as a nondegenerate ∗-homomorphism from C0( H) ∼= C∗(H) to the central multiplier algebra of A. For crossed modules of the form C4 = (G4, 0, 0, 0), there is nothing to analyse: actions of this crossed module are the same as actions of the locally compact group G4, and the crossed product functor is also the same as for group actions. We already showed in [2] that group actions by correspondences are equivalent to saturated Fell bundles. The crossed product is the cross-sectional C∗-algebra of a Fell bundle. By the Packer -- Raeburn Stabilisation Trick (see also [2, Theorem 5.3]), we may replace G4-actions by correspondences by ordinary continuous group actions on a stabilisation. This replaces the crossed product functor for actions 34 ALCIDES BUSS AND RALF MEYER of C4 by a classical crossed product construction for actions of the locally compact group G4. Now we study crossed products by the thin crossed module C3 = (∂H, H/ ker ∂, ∂3, c3), using the results in Section 3.1 to replace C3 by an equivalent Abelian crossed module C5. By Theorem 3.11, such an Abelian model for C3 exists if both G and H are Lie groups. More generally, Lemma 3.15 gives an Abelian model if H/ ker ∂ has a compactly generated subgroup A for which ∂(A) is open in ∂(H); in particular, this happens if H itself is compactly generated. Theorem 3.14 also gives a necessary and sufficient condition for an Abelian model to exist; but this criterion does not explain why this happens so often. Assume that C3 is equivalent to an Abelian crossed module C5 = (G5, H5, ∂5, c5); even better, we can achieve that G5 is compact Abelian, H5 is discrete, and c5 is trivial. Since C3 is thin, so is C5, that is, ∂5 is an injective map with dense range. By Theorem 3.5, the 2-categories Corr(C3) and Corr(C5) of actions of C3 and C5 by correspondences are equivalent, in such a way that the crossed product functors on both categories are identified. Moreover, the proof shows immediately that the underlying C∗-algebra is not changed: an action of C3 becomes a C5-action on the same C∗-algebra. For crossed modules of Lie groups, we have explained in Section 3 how to construct C5 explicitly out of C3. Theorem 4.5 shows that Corr(C5) is equivalent to the 2-category of actions of the groupoid cG5 ⋉ cH5; this equivalence maps a C5-action to the crossed product by G5 equipped with a canonical C0(cH5)-C∗-algebra structure and the dual action of cG5; the crossed product by C5 corresponds to taking the fibre at 1 for the C0(cH5)-C∗-algebra structure. Thus after an equivalence Corr(C3) ≃ Corr(C5) ≃ Corr(cG5 ⋉ cH5) that on the underlying C∗-algebras takes a crossed product with the Abelian com- pact group G5, the crossed product with C3 becomes a fibre restriction functor. The following theorem summarises our factorisation of the crossed product: Theorem 6.1. Let C be a crossed module of Lie groups or, more generally, a crossed module of locally compact groups for which the associated thin crossed module H/ ker ∂ → ∂(H) is equivalent to an Abelian crossed module. There are a locally compact Abelian group X, compact Abelian groups Y and K, and a lo- cally compact group L, such that for any action of C by correspondences on a C∗-algebra A, (1) A carries a natural C0(X)-C∗-algebra structure; (2) the unit fibre A1 of A for this natural C0(X)-C∗-algebra structure carries a natural action of K by correspondences; (3) the crossed product A2 := A1 ⋊ K carries a natural C0(Y )-C∗-algebra structure; (4) the unit fibre A3 of A2 for this natural C0(Y )-C∗-algebra structure carries a natural action of L by correspondences; (5) the crossed product A3 ⋊ L is naturally isomorphic to A ⋊ C. CROSSED PRODUCTS FOR CROSSED MODULES 35 6.1. Computing K-theory of crossed module crossed products. In the localisation formulation of [11], the Baum -- Connes assembly map for a locally compact group G compares the K-theory of the reduced crossed product with a more topological invariant that uses only crossed products for restrictions of the action to compact subgroups of G. Its assertion is therefore trivial if G is itself compact. Crossed products for compact groups are an "elementary" operation for K-theory purposes in the sense that there is no better way to compute the K-theory than the direct one. Crossed products for non-compact groups are not "elementary" in this sense because the Baum -- Connes conjecture (if true) allows us to reduce the K-theory computation to K-theory computations for compact subgroups and some algebraic topology to assemble the results of these computations. Taking the fibre in a C0(X)-C∗-algebra seems to be an operation that is also "elementary" in the above sense. At least, we know of no better way to compute the K-theory of a fibre than the direct one. Notice that C0(X)-C∗-algebras need not be locally trivial. In the notation of Theorem 6.1, the functors A 7→ A1 7→ A2 7→ A3 are therefore "elementary" for K-theory purposes. The remaining fourth step A3 7→ A3 ⋊ L is the (full) crossed product by the locally compact group L = G/∂(H). Many results are available about the K-theory of such crossed products. Hence our decomposition of crossed module crossed products also gives us a useful recipe for computing their K-theory. This recipe is, however, quite different from the localisation approach for groups in [11]. References [1] Alcides Buss, Ralf Meyer, and Chenchang Zhu, Non-Hausdorff symmetries of C∗-algebras, Math. Ann. 352 (2012), no. 1, 73 -- 97, doi: 10.1007/s00208-010-0630-3. MR 2885576 [2] , A higher category approach to twisted actions on C∗-algebras, Proc. Edinb. Math. Soc. (2) 56 (2013), no. 2, 387 -- 426, doi: 10.1017/S0013091512000259. MR 3056650 [3] Robert S. Doran and James M. G. Fell, Representations of ∗-algebras, locally compact groups, and Banach ∗-algebraic bundles. Vol. 2, Pure and Applied Mathematics, vol. 126, Academic Press Inc., Boston, MA, 1988. MR 936629 [4] Siegfried Echterhoff, The primitive ideal space of twisted covariant systems with continuously varying stabilizers, Math. Ann. 292 (1992), no. 1, 59 -- 84, doi: 10.1007/BF01444609.MR 1141785 [5] Siegfried Echterhoff, Steven P. Kaliszewski, John Quigg, and Iain Raeburn, A categorical approach to imprimitivity theorems for C ∗-dynamical systems, Mem. Amer. Math. Soc. 180 (2006), no. 850, viii+169, doi: 10.1090/memo/0850. MR 2203930 [6] Siegfried Echterhoff and John Quigg, Induced coactions of discrete groups on C ∗-algebras, Canad. J. Math. 51 (1999), no. 4, 745 -- 770, doi: 10.4153/CJM-1999-032-1. MR 1701340 [7] Philip Green, The local structure of twisted covariance algebras, Acta Math. 140 (1978), no. 3-4, 191 -- 250, doi: 10.1007/BF02392308. MR 0493349 d'espaces [8] Michel Hilsum and Georges de feuilles d'A. Connes), Ann. Sci. École Norm. Sup. (4) 20 (1987), no. 3, 325 -- 390, available at http://www.numdam.org/item?id=ASENS_1987_4_20_3_325_0. MR 925720 Skandalis, Morphismes K-orientés (d'après une conjecture et fonctorialité en théorie de Kasparov [9] Tom Leinster, Basic Bicategories (1998), eprint. arXiv: math/9810017. [10] Saunders MacLane and John Henry Constantine Whitehead, On the 3-type of 41 -- 48, available at a complex, Proc. Nat. Acad. Sci. U. S. A. 36 (1950), http://www.pnas.org/content/36/1/41.full.pdf+html. MR 0033519 [11] Ralf Meyer and Ryszard Nest, The Baum -- Connes conjecture via localisation of categories, Topology 45 (2006), no. 2, 209 -- 259, doi: 10.1016/j.top.2005.07.001. MR 2193334 36 ALCIDES BUSS AND RALF MEYER [12] Paul S. Muhly, Bundles over groupoids, Groupoids in analysis, geometry, and physics (Boulder, CO, 1999), Contemp. Math., vol. 282, Amer. Math. Soc., Providence, RI, 2001, pp. 67 -- 82, doi: 10.1090/conm/282/04679. MR 1855243 [13] Paul S. Muhly, Jean N. Renault, and Dana P. Williams, Equivalence and isomor- phism for groupoid C ∗-algebras, J. Operator Theory 17 (1987), no. 1, 3 -- 22, available at http://www.theta.ro/jot/archive/1987-017-001/1987-017-001-001.pdf. MR 873460 [14] Paul S. Muhly and Dana P. Williams, Equivalence and disintegration theorems for Fell bundles and their C ∗-algebras, Dissertationes Math. (Rozprawy Mat.) 456 (2008), 1 -- 57, doi: 10.4064/dm456-0-1. MR 2446021 [15] May Nilsen, C ∗-bundles and C0(X)-algebras, Indiana Univ. Math. J. 45 (1996), no. 2, 463 -- 477, doi: 10.1512/iumj.1996.45.1086. MR 1414338 [16] Behrang Noohi, Notes on 2-groupoids, ules, Homology, Homotopy Appl. http://projecteuclid.org/euclid.hha/1175791088. MR 2280287 (2007), 9 2-groups 1, no. and 75 -- 106, crossed available mod- at [17] Shigeru Yamagami, On primitive ideal spaces of C ∗-algebras over certain locally com- pact groupoids, Mappings of operator algebras (Philadelphia, PA, 1988), 1990, pp. 199 -- 204. MR 1103378 E-mail address: [email protected] Departamento de Matemática, Universidade Federal de Santa Catarina, 88.040-900 Florianópolis-SC, Brazil E-mail address: [email protected] Mathematisches Institut, Georg-August-Universität Göttingen, Bunsenstrasse 3 -- 5, 37073 Göttingen, Germany
1710.08475
2
1710
2017-12-07T08:18:21
Composition of PPT Maps
[ "math.OA", "math-ph", "math.FA", "math-ph" ]
M. Christandl conjectured that the composition of any trace preserving PPT map with itself is entanglement breaking. We prove that Christandl's conjecture holds asymptotically by showing that the distance between the iterates of any unital or trace preserving PPT map and the set of entanglement breaking maps tends to zero. Finally, for every graph we define a one-parameter family of maps on matrices and determine the least value of the parameter such that the map is variously, positive, completely positive, PPT and entanglement breaking in terms of properties of the graph. Our estimates are sharp enough to conclude that Christandl's conjecture holds for these families.
math.OA
math
COMPOSITIONS OF PPT MAPS MATTHEW KENNEDY, NICHOLAS A. MANOR, AND VERN I. PAULSEN Abstract. M. Christandl conjectured that the composition of any trace preserving PPT map with itself is entanglement breaking. We prove that Christandl's conjecture holds asymptotically by showing that the distance between the iterates of any unital or trace preserving PPT map and the set of entanglement breaking maps tends to zero. Finally, for every graph we define a one-parameter family of maps on matrices and determine the least value of the parameter such that the map is vari- ously, positive, completely positive, PPT and entanglement breaking in terms of properties of the graph. Our estimates are sharp enough to conclude that Christandl's conjecture holds for these families. 1. General Introduction The usual mathematical model for a quantum channel is a completely positive trace-preserving (CPTP ) map between two matrix spaces. A com- pletely positive map is called a positive partial transpose (PPT ) map if the composition of the map with the transpose map on the range space is still completely positive. PPT maps play an important role in the study of en- tanglement. A completely positive map can also be identified with a state on the tensor product of the two matrix algebras, and the states corresponding to PPT maps are called PPT states. PPT states could play a role in quantum key distribution(QKD), which is the study of the use of various quantum mechanical systems to construct shared states that would be used to insure secure communication. These considerations lead Christandl to consider questions about how PPT maps behaved under composition and lead to the following conjecture [10]: Conjecture 1.1 (PPT-Squared Conjecture). The composition of a pair of PPT maps is always entanglement breaking. We will define and discuss entanglement breaking maps in the following section. From the point of view of shared states, this conjecture is equivalent to the following statement [1]: "Assume that Alice and Charlie share a PPT state and that Bob and Charlie share a PPT state; then the state of Alice and Bob, conditioned on any measurement by Charlie, is always separable." First author supported by NSERC Grant Number 418585 Second author supported by NSERC Grant Number 396164132 1 2 M. KENNEDY, N. A. MANOR, AND V. I. PAULSEN The PPT-squared conjecture is known to be true for maps on the 2 × 2 matrices. We prove that Christandl's intuition about the behaviour of PPT maps under composition is at least asymptotically true by showing that the distance between the iterates of a PPT channel and the set of entanglement breaking maps tends to zero. From the point of view of a quantum communication network this implies that if each pair of channels shares the same PPT state then, eventually, the network will behave as if they are sharing a state corresponding to an entanglement breaking map. In Section 2, we give precise definitions of the concepts introduced. In Sec- tion 3, we show using basic techniques from the theory of topological semi- groups that the conjecture holds asymptotically. In Section 4, we consider a new type of spectral graph theory problem; we associate a one-parameter family of maps to each graph and determine in terms of the graph the smallest values of the parameter for which the map is, variously, positive, completely positive, PPT, and entanglement breaking. Our estimates are sharp enough that we are able to show that whenever two maps in this family are PPT, then their composition is entanglement breaking, i.e., the conjecture holds for this family. We recall that by Chois theorem [2, Theorem 2], a map φ : Mp → Mq is 2. Basics CP if and only if its Choi matrix Cφ := (cid:0)φ(Ei,j)(cid:1) is a positive semidefinite We present here a very basic but useful result about PPT maps. matrix in Mp(Mq) = Mpq. Lemma 2.1. Let φ : Mp → Mq be a CP map. Then φ is PPT if and only if φ ◦ T is CP, where T denotes the transpose map on Mp. In other words, to check whether a CP map is PPT we may compose or precompose with the transpose map. Proof. The Choi matrix of T ◦ φ is ... ... φ(E11)T φ(En1)T · · · φ(E1n)T . . . · · · φ(Enn)T CT◦φ =  Since the transpose is a positive map, Cφ◦T is positive if and only if its transpose CT◦φ is. Corollary 2.2. The set of PPT maps is closed under composition by CP maps on the right and on the left. · · · φ(En1) . . . · · · φ(Enn) =    = Cφ◦T T . φ(E11) ... φ(E1n) (cid:3) ... T   By results of [5], φ is entanglement breaking if and only if it can be written as φ(X) =Xk vkw∗kXwkv∗k =Xk sk(X)Pk, for some set of vectors wk ∈ Cp and vk ∈ Cq, where sk(X) = hwk, Xwki and Pk = vkv∗k. By normalizing the wk's and vk's we extract weights dk = kwk · kvkk, and we may assume the sk are states and the Pk's are rank one projections. Thus, φ is entanglement breaking if and only if Cφ can be written as 3 Cφ =Xk dk(cid:0)sk(Ei,j)Pk(cid:1) =X dkQk ⊗ Pk, with the Qk ∈ Mp density matrices and the Pk ∈ Mq rank one projections. Moreover since every density matrix can be written as a sum of rank one projections, we have that φ is entanglement breaking if and only if tlRl ⊗ Sl, Cφ =Xl where Rℓ ∈ Mp and Sℓ ∈ Mq are rank one projections, and tℓ are positive weights. Another more recent characterization, given in [6], is that φ is entan- glement breaking exactly when it factors through ℓ∞k via positive maps, for some k. More precisely: there are positive maps ψ : Mp → ℓ∞k and γ : ℓ∞k → Mq so that φ = γ ◦ ψ. Remark 2.3. This characterization may be modified slightly so that instead of factoring through a finite-dimensional abelian C∗-algebra, any abelian C∗- algebra may be used. Simply note that, if φ = γ ◦ ψ for ψ : Mp → C(X) and γ : C(X) → Mq, then φ ◦ Θ is still completely positive for any positive map Θ on Mp. This simply follows from the fact that positive maps into or from an abelian C∗-algebra are necessarily completely positive. 3. The Asymptotic Result It turns out that Christandl's intuition holds asymptotically in the follow- ing sense: the sequence of iterates of a PPT channel φ on Mn approaches the set of entanglement breaking maps. To prove this we use some very basic results from the theory of abelian semigroups. Since these objects live in a finite dimensional space, convergence is independent of any particular metric. We first examine the case of an idempotent unital PPT map. Lemma 3.1. Let φ : Mn → Mn be an idempotent unital PPT map. Then the range of φ is an abelian C*-algebra with respect to the product a ∗ b = φ(ab), for a, b in φ(Mn). Proof. A result of Choi and Effros [3] implies that the range of φ is a C*- algebra with respect to the product given above. We must show that this C*-algebra is abelian. Supposing it is non-abelian: it has a direct summand isomorphic to Mk for some k ≥ 2. So by composing with the associated projection we get, by Corollary 2.2, an induced PPT map ψ which is surjective onto Mk. In this case, ψ ⊗ idk : Mn ⊗ Mk → Mk ⊗ Mk is also surjective. So there is a positive matrix (Aij) ∈ Mn ⊗ Mk with ψ ⊗ idk((Aij )) = (Eij), however 4 M. KENNEDY, N. A. MANOR, AND V. I. PAULSEN (T (Eij)) is not positive, so the composition T ◦ ψ is not completely positive, contradicting ψ being PPT. Therefore, we conclude that the range of φ is abelian. (cid:3) Proposition 3.2. Let φ be an idempotent unital PPT map on Mn. Then φ is entanglement breaking. Proof. By the previous lemma, φ factors through a finite dimensional abelian C*-algebra; so by Remark 2.3 we have the result. (cid:3) Lemma 3.3. If φ is a contractive map on Mn endowed with any norm, then there is an idempotent map ψ in the limit points of (φk)k≥1. Proof. Let S denote the closure of {φk : k ∈ N}. This is compact since φ is contractive, and it is an abelian semigroup under composition. Let K = ∩a∈S aS be the intersection of all singly generated ideals. Then we claim K is a minimal ideal in S. First of all, this set is non-empty by the finite intersection property and by the fact that the product of finitely many ideals is contained in their in- tersection. It is also clearly an ideal, as an intersection of ideals. Minimality follows from the fact that every ideal contains a singly generated ideal, and of course every singly generated ideal contains K. We will now show that K has a multiplicative identity, thus giving us an idempotent in S. Take k ∈ K. Since K is minimal we have k2S = K, so there is s ∈ K such that (sk)k = sk2 = k. We claim that sk is the identity in K. Taking any k′ ∈ K, again by minimality there is s′ ∈ S such that s′k = k′. From this we get that (sk)k′ = (sk)s′k = s′(sk)k = s′k = k′. So we may take ψ = sk. (cid:3) Note that if φ is trace preserving then so is the idempotent ψ. The next lemma tells us exactly how φk approaches the set of EB maps. Lemma 3.4. If φ and ψ are as above then kφk − φk ◦ ψk → 0. Proof. Since ψ is a limit of powers of φ, φ and ψ commute. Hence ran(ψ) and ker(ψ) = ran(id −ψ) are invariant for φ. It follows from the spectral mapping theorem that σ(φran(ψ)) ⊆ T and σ(φran(id −ψ)) ⊆ D. Since φ and ψ commute, this implies σ(φnran(id −ψ)) ⊆ D. Hence lim kφn − φn ◦ ψk = 0. (cid:3) Theorem 3.5. Every unital or trace preserving PPT map φ is asymptot- ically entanglement breaking, in the sense that d(φk, EB) → 0, where EB denotes the set of entanglement breaking maps on Mn. Proof. In the unital case, we know that the idempotent map ψ from the lemma above is PPT and hence entanglement breaking by Proposition 3.2. 5 So for every n the map φk ◦ ψ is entanglement breaking, and by the previous lemma this implies d(φk, EB) → 0. To retrieve the trace preserving case, recall that trace preserving maps are precisely the adjoints of unital ones, and the set of entanglement breaking maps is also *-symmetric. So, since the adjoint operation is an isometry we get the result. (cid:3) Remark 3.6. The above theorem can also be deduced from the work of Lami and Giovanetti[7] on asymptotically entanglement-saving channels. A channel is called asymptotically entanglement-saving if no limit point of its iterates is entanglement breaking, which is easily seen to be equivalent to the negation of our condition that limk d(φk, EB) = 0. Combining [7, Theo- rem 32.2] and [7, Theorem 12] shows that no PPT map can be asymptotically entanglement-saving. 4. Schur Product Maps In this section, we examine a class of maps for which it is possible to prove Christandl's conjecture; these form a one parameter family of maps defined from graphs. Determining for which values of the parameter these maps belong to the various classes of positive maps, completely positive maps, PPT maps, and entanglement breaking maps, leads to interesting spectral questions in combinatorics. For some of these families of maps we are able to answer these questions exactly, for others we can only give estimates on the parameter. However, our estimates are good enough to show that, for all graphs, the PPT-squared conjecture is true for maps in this family. Recall that given matrices A = (ai,j), B = (bi,j) of the same size, their Schur product is the matrix A ◦ B := (ai,j bi,j). Given A ∈ Mp then we set SA : Mp → Mp to be the map SA(B) = A ◦ B. It is well known that SA is CP if and only if A ≥ 0. Proposition 4.1. Let P be an n × n matrix. Then SP is PPT if and only if P ≥ 0 and P is diagonal. Proof. It is readily checked that if P ≥ 0 and P is diagonal then Sp is PPT. Assume that SP is PPT. Then since SP is CP, P ≥ 0. Let T denote the transpose map, and assume that T ◦ SP is CP. If P = (pi,j), then by Choi's theorem, If i 6= j, then the 2 × 2 block submatrix Ei,j ⊗ pi,jEj,i ≥ 0. Xi,j pi,jEi,j pj,jEj,j(cid:19) ≥ 0. (cid:18)pi,iEi,i pi,jEj,i But this is possible, only if pi,j = 0. Hence P is diagonal. (cid:3) 6 M. KENNEDY, N. A. MANOR, AND V. I. PAULSEN Thus, there are no "interesting" Schur product maps that are PPT. Let T r : Mp → C be the usual trace, and let tr(B) = 1 p T r(B) denote the normalized trace. We set δ : Mp → Mp to be δ(X) = tr(X)Ip. Note that δ ◦ δ = δ. Note that this CP map is entanglement breaking. Now let A = A∗ be a p × p matrix of 0's and 1's with the diagonal equal to 0. The set of (i, j) such that ai,j 6= 0 can be thought of as the edge set of a graph G = (V, E) on p vertices, in which case A is the adjacency matrix of the graph. The Schur product map SA is idempotent. We are interested in the one parameter family of maps, and in determining the following parameters of the graph G: γt = γt,A = tδ + SA, • tpos = min{t : γt is a positive map }, • tcp = min{t : γt is CP }, • tppt = min{t : γt is PPT }, • teb = min{t : γt is EB }. Clearly, we have that tpos ≤ tcp ≤ tppt ≤ teb. In general, we expect tpos < tcp. In fact for the the case of the complete graph on 2 vertices, i.e., an edge we have that γt(cid:0)(cid:18)p11 p12 p21 p22(cid:19)(cid:1) = t(p11+p22) 2 p21 p12 t(p11+p22) 2 ! . Using the determinant test and the root mean inequality, one sees that this map is positive for t = 1, since for any positive matrix, p11 + p22 2 2 ≥ p11p22 ≥ p12p21. However, one readily sees that for t = 1, the Choi matrix of this map is not positive. Thus, γ1 is positive but not CP. Note that γt ◦ γt = γt2 , so that the PPT-squared conjecture will hold for this family of maps if and only if teb ≤ t2 ppt, which we shall prove below. Given an adjacency matrix A we let λmin to denote the least (real) eigen- value of A. Since T r(A) = 0 this number will always be strictly negative, as long as A 6= 0. Proposition 4.2. If A is a non-zero p × p adjacency matrix, then tcp = tppt = −pλmin. Proof. To compute tcp, we must determine restrictions imposed by requir- ing the Choi matrix of the map is positive. This matrix is t p I ⊗ I +CSA where CSA is the Choi matrix of the map SA. For tppt we also need that 7 t p I ⊗ I +CSA◦T is positive, where T is the transpose map on Mp and CSA◦T is the Choi matrix of SA ◦ T . As for requiring t p I ⊗ I +CSA to be positive, we find the minimal eigen- value of CSA =P(i,j)∈E(G) Eij ⊗ Eij, where G is the associated graph of A. Notice that CSA is identically zero on the space spanned by ek ⊗ el, with k 6= l, and on the span of ek ⊗ ek it behaves exactly as A acting on Cp. We have that −λmin I ⊗ I +CSA◦T is positive, and it is non-positive for any strictly smaller multiple of I ⊗ I. Thus, tcp = −pλmin. For the second case, observe that CSA◦T = P(i,j)∈E Eji ⊗ Eij, so that (CSA◦T )2 = P(i,j)∈E Eii ⊗ Ejj is a diagonal matrix of only 1's and 0's. In particular, the spectrum of (CSA◦T )2 must be a subset of {0, 1}, but then CSA◦T may only have eigenvalues of -1, 0 and 1. So for t = p we will certainly have that γt,A ◦ T is completely positive, and it is minimal exactly when -1 is an eigenvalue of CSA◦T . In fact, this will always be the case; choose (k, l) such that ak,l = 1 and notice that CSA◦T (ek ⊗ el − el ⊗ ek) = el ⊗ ek − ek ⊗ el. Thus, tppt = p · min{1, −λmin}. It is easily checked that for any non-zero adjacency matrix, λmin ≤ −1 so that tppt = −pλmin, also. To verify this last claim, note that if Ai,j = 1 and we set v = ei−ej√2 then v is a unit vector and with hAv, vi = −1, from which it follows that λmin ≤ −1. (cid:3) We would now like to fully understand teb, although this will pose a greater issue as it is rarely clear when a matrix is separable in a tensor product. We present here a natural upper bound for teb via a simple computation. Recall that we view A as being the adjacency matrix of a graph G = (V, E). Lemma 4.3. Let A be the adjacency matrix of a graph G = (V, E) on p vertices. Then γpd = pd · δ + SA is entanglement breaking, where d denotes the maximum edge degree in G. In particular, teb ≤ pd. Proof. We proceed by considering the Choi matrix Cφ of φ = pd · δ + SA and showing it is separable. It is easy to see that Eij ⊗ Eij Cφ = d Ip ⊗ Ip + X(i,j)∈E = D + X(i,j)∈E and i<j Eii ⊗ Ejj + Ejj ⊗ Eii + Eij ⊗ Eij + Eji ⊗ Eji, where D is a diagonal matrix consisting of 1's and 0's (hence it is separable). So it suffices to show that matrices of the same form as the summand on the right hand side (where i and j may vary from 1 to p) are separable. 8 M. KENNEDY, N. A. MANOR, AND V. I. PAULSEN We use only the four following positive matrices in Mp to prove this fact: Q1,i,j = Ei,i + Ej,j + Eij + Eji Q2,i,j = Ei,i + Ej,j + Eij − Eji Q3,i,j = Ei,i + Ej,j + iEij − iEji Q4,i,j = Ei,i + Ej,j − iEij + iEji, where i < j vary from 1 to p. A routine computation shows that 4(cid:0)(Ei,i + Ej,j) ⊗ (Ei,i + Ej,j) + Eij ⊗ Eij + Eji ⊗ Eji(cid:1) = Q1,i,j ⊗ Q1,i,j + Q2,i,j ⊗ Q2,i,j + Q3,i,j ⊗ Q4,i,j + Q4,i,j ⊗ Q3,i,j. Summing over all edges we get that R = 4 X(i,j)∈E (Ei,i + Ej,j) ⊗ (Ei,i + Ej,j) + 8 X(i,j)∈E Ei,j ⊗ Ej,i, is separable. Now in the sum P(i,j)∈E(Ei,i + Ej,j) ⊗ (Ei,i + Ej,j), for k 6= l each term Ek,k ⊗ El,l appears at most once, while Ei,i ⊗ Ei,i occurs exactly 2di ≤ 2d times. Since each term Ek,k ⊗ El,l is separable, we see that we can add a separable term Q to R so that R + Q = 8d(cid:0)Xi,j Ei,i ⊗ Ej,j(cid:1) + 8 X(i,j)∈E Ei,j ⊗ Ej,i = 8Cφ, and it follows that Cφ is separable so that φ = γpd,A is entanglement break- ing. (cid:3) Before stating the next result note that the Schur product A ◦ B of two adjacency matrices is again an adjacency matrix. Corollary 4.4. If A and B are adjacency matrices and the maps γt1,A and γt2,B are PPT, then their composition is entanglement breaking. Proof. The composition evaluates to γt1t2,A◦B, so if either A or B is zero then the composition is the map X 7→ t1t2 tr(X). This map is clearly entanglement breaking. If both are non-zero matrices then t1t2 ≥ p2, and this is necessarily greater than teb for any adjacency matrix of size p since the degree of any vertex cannot exceed p − 1. The corollary follows. (cid:3) Numerically, it is possible to compute tpos. Indeed, to check if γt is positive it is enough to check that it is positive for all rank one positive matrices arising from unit vectors. For such a matrix we have that γt((αiαj)) = t p I +SA((αiαj)). This leads to tpos = −p min{λmin(SA(αiαj)) : α12 + · · · + αp2 = 1}. 9 On the other hand, if we consider SA as a map from Mp to Mp endowed with its trace norm, i.e., the Schatten one-norm, then the norm kSAk1 is attained on such rank one matrices, and so tpos ≤ pkSAk1. However, the adjoint of SA is again the map SA, so that one has kSAk1 = kSAk, where the latter norm is the norm of the linear map SA : Mp → Mp and Mp is endowed with its usual operator norm, i.e., kXk2 = λmax(X∗X). There is a well-known formula for computing the norm of such Schur product maps. See for example [9, Theorem 8.7]. Thus, tpos ≤ pkSAk gives an upper bound on this quantity. Next we turn to some lower bounds on tpos in terms of more familiar graph parameters. Proposition 4.5. Let A be the adjacency matrix of a graph G = (V, E) on p vertices and let ϑ = ϑ(G) denote the Lov´asz theta number of the graph and let ϑ = ϑ(G), denote the Lov´asz theta number of the graph complement of G. Then tpos ≥ max{1, −λmin(A), −pλmin(A) E , −pλmin(A) teb , λmax(A) ϑ − 1 }. Proof. Let r ≥ tpos, so that γr is a positive map. The first two inequalities come from applying γr to the positive matrices −λmin(A) I +A and the p × p matrix of all 1's. The third inequality comes from applying γr to the Laplacian matrix of the graph, L = A +Pi diEi,i, which is positive since it is diagonally dominant. To see the fourth inequality, note that if s ≥ teb, then γr ◦ γs = γrs is the composition of a positive map and an entanglement breaking map and so is CP. Hence, tebtpos ≥ tcp = −pλmin(A), and the inequality follows. For the final inequality, we use the fact that [8, Theorem 3], ϑ = min{λmax(H) : H = H∗, SI +A(H) = I I +H}. Let K = K∗ be the matrix such that SI +A(K) = I +A and λmax(K) = ϑ. Then we have that K ≤ ϑ I and so, r I +A = γr(K) ≤ rϑ I. Hence, A ≤ r(ϑ − 1) I and the last inequality follows. (cid:3) References [1] S. Bauml, M. Christandl, K. Horodecki, and A. Winter, Limitations on Quantum Key Repeaters, arXiv preprint arXiv:1402.5927 (2014). [2] M.D. Choi, Completely positive linear maps on complex matrices, Linear algebra and its applications 10, no. 3 (1975): 285-290. [3] M.D. Choi and E.G. Effros, Injectivity and Operator Spaces, J. Functional Anal. 24(1977), 156-209. [4] D. Chru´sci´nski and A. Kossakowski, On the structure of entanglement witnesses and new class of positive indecomposable maps, Springer (2007) [5] M. Horodecki, P.W. Shor, M.B. Ruskai, General entanglement breaking channels, Rev. Math. Phys. 15 (2003) 629-641. [6] N. Johnston, D. Kribs, V. Paulsen, and R. Pereira, Minimal and maximal operator spaces and operator systems in entanglement theory. Journal of Functional Analysis 260.8 (2011): 2407-2423. 10 M. KENNEDY, N. A. MANOR, AND V. I. PAULSEN [7] L. Lami and V. Giovannetti, Entanglement-Saving Channels, arXiv preprint, arXiv:1505.00461v1 (2015). [8] L. Lov´asz, On the Shannon capacity of a Graph, IEEE Transactions on Information Theory, Vol 25, No. 1, January 1979, 1-6. [9] V.I. Paulsen, Completely Bounded Maps and Operator Algebras, Cambridge Studies in Advanced Mathematics 78, Cambridge University Press (2002). [10] Mary Beth Ruskai, Marius Junge, David Kribs, Patrick Hayden, Andreas Winter, Operator structures in quantum information theory, Final Report, Banff International Research Station (2012). Department of Pure Mathematics, University of Waterloo, Waterloo, ON, Canada N2L 3G1 E-mail address: [email protected] Department of Pure Mathematics, University of Waterloo, Waterloo, ON, Canada N2L 3G1 E-mail address: [email protected] Institute for Quantum Computing and Department of Pure Mathematics, University of Waterloo, Waterloo, ON, Canada N2L 3G1 E-mail address: [email protected]
1401.1143
3
1401
2015-02-02T19:43:08
The reflexive closure of the adjointable operators
[ "math.OA" ]
Given a Hilbert module E over a C*-algebra A, we show that the collection of all bounded A-module operators acting on E forms the reflexive closure for the algebra of the adjointable operators. We also make an observation regarding the representation theory of the left centralizer algebra of a C*-algebra and use it to give an intuitive proof of a related result of H. Lin
math.OA
math
THE REFLEXIVE CLOSURE OF THE ADJOINTABLE OPERATORS E. G. KATSOULIS Abstract. Given a Hilbert C ∗-module E over a C*-algebra A, we give an explicit description for the invariant subspace lattice lat L(E) of all adjointable operators on E. We then show that the collection EndA(E) of all bounded A-module operators acting on E forms the reflexive clo- sure for L(E), i.e., EndA(E) = alg lat L(E). Finally we make an obser- vation regarding the representation theory of the left centralizer algebra of a C ∗-algebra and use it to give an intuitive proof of a related result of H. Lin. 1. Introduction In this note, A denotes a C*-algebra and E a Hilbert C*-module over A, i.e., a right A-module equipped with an A-valued inner product h , i so that the norm kξk ≡ k hξ, ξi1/2 k makes E into a Banach space. The collection of all bounded A-module operators acting on E is denoted as EndA(E). A linear operator S acting on E is said to be adjointable iff given x, y ∈ E there exists y′ ∈ E so that hSx, yi = hx, y′i. Elementary examples of adjointable operators are the "rank one" operators θη,ξ, defined by θη,ξ(x) ≡ η hξ, xi, where η, ξ, x ∈ E. The collection of all adjointable operators acting on E will be denoted as L(E) while the norm closed subalgebra generated by the rank one operators will be denoted as K(E). It is a well known fact that L(E) ⊆ EndA(E). However, the reverse inclusion is known to fail in general; this is perhaps the first obstacle one encounters when extending the theory of operators on a Hilbert space to that of operators on a Hilbert C ∗-module. This problem has been addressed since the beginning of the theory [21, page 447] and has influenced its sub- sequent development. The first few chapters of the monograph of Manuilov and Troitsky [19] and the references therein provide the basics of the theory and give a good account of what is known regarding that issue. (See also [4, 17].) The purpose of this note is to demonstrate that the inequality be- tween L(E) and EndA(E) is intimately related to another area of continuing mathematical interest, the reflexivity of operator algebras. If A is a unital operator algebra acting on a Banach space X, then lat A will denote the collection of all closed subspaces M ⊆ X which are left invariant 2010 Mathematics Subject Classification. 46L08, 47L10. Key words and phrases: Hilbert C ∗-module, adjointable operator, reflexive operator algebra, reflexive closure, invariant subspace, left centralizer, left multiplier. 1 2 E. G. KATSOULIS by A, i.e., A(m) ∈ M , for all A ∈ A and m ∈ M . Dually, for a collection L of closed subspaces of X, we write alg L to denote the collection of all bounded operators on X that leave invariant each element of L. The reflexive cover of an algebra A of operators acting on X is the algebra alg lat A; we say that A is reflexive iff A = alg lat A. Similarly, the reflexive cover of a subspace lattice L is the lattice lat alg L and L is said to be reflexive if L = lat alg L. A formal study of reflexivity for operator algebras and subspace lattices began with the work of Halmos [10], after Ringrose's proof [23] that all nests on Hilbert space are reflexive. Since then, the concept of reflexivity for operator algebras and subspace lattices has been addressed by various authors on both Hilbert space [1, 2, 3, 6, 9, 13, 15, 20, 24, 25] and Banach space [5, 7, 8], including in particular investigations on a Hilbert C ∗-module. The main results of this short note provide a link between the two areas of inquiry discussed above. In Theorem 2.5 we show that the presence of bounded but not adjointable module operators on a C ∗-module E is equiv- alent to the failure of reflexivity for L(E). (Here we think of L(E) simply as an operator algebra acting on E.) Actually, we do more: we explicitly describe lat L(E) and we show that as a complete lattice, lat L(E) is iso- morphic to the lattice of closed left ideals of hE, Ei (Theorem 2.3). A key step in the proof of Theorem 2.5 is a classical result of Barry Johnson [11, Theorem 1]. Actually, our Theorem 2.5 can also be thought of as a gener- alization of Johnson's result, since its statement reduces to the statement of [11, Theorem 1], when applied to the case of the trivial (unital) Hilbert C ∗-module. Another interpretation for the inequality between L(E) and EndA(E) comes from the work of H. Lin. Lin shows in [18, Theorem 1.5] that EndA(E) is isometrically isomorphic as a Banach algebra to the left centralizer algebra of K(E). Furthermore, the isomorphism Lin constructs extends the familiar ∗-isomorphism between L(E) and the double centralizer algebra of K(E). This shows that the gap between L(E) and EndA(E) is solely due to the presence of left centralizers for K(E) which fail to be double centralizers. In Proposition 3.3 we observe that the representation theory of the left centralizer algebra of a C ∗-algebra is flexible enough to allow the use of representations on a Banach space. This leads to yet another short proof of Lin's Theorem, which we present in Theorem 3.4. Our proof makes no reference to Cohen's Factorization Theorem and its only prerequisite is the existence of a contractive approximate identity for a C ∗-algebra. (Compare also with [4, Proposition 8.1.16 (ii)].) A final remark. Johnson's Theorem [11, Theorem 1], which plays a central role in this paper, may no longer be true for Banach algebras which are not semisimple. Nevertheless there are specific classes of (non-semisimple) THE REFLEXIVE CLOSURE OF THE ADJOINTABLE OPERATORS 3 operator algebras for which this theorem is actually valid. This is being explored in a subsequent work [16]. 2. the main result We begin by identifying a useful class of subspaces of E. Definition 2.1. Let E a Hilbert C*-module over a C*-algebra A. If J ⊆ A, then we define E(J ) := span{ξa ξ ∈ E, a ∈ J }. The correspondence J 7→ E(J ) of Definition 2.1 is not bijective. Indeed, if l(J ) is the closed left ideal generated by J ⊆ A, then it is easy to see that E(l(J )) = E(J ). Therefore we restrict our attention to closed left ideals of A. It turns out that an extra step is still required to ensure bijectivity. First we need the following. Lemma 2.2. Let E be a Hilbert C*-module over a C*-algebra A and let J ⊆ A be a closed left ideal. Then E(J ) = {ξ ∈ E hη, ξi ∈ J for all η ∈ E}. Proof. The inclusion E(J ) ⊆ {ξ ∈ E hη, ξi ∈ J for all η ∈ E} is obvious. The reverse inclusion follows from the well known fact [19, Lemma 1.3.9] that for any ξ ∈ E. ξ = lim ǫ→0 ξ hξ, ξi [hξ, ξi + ǫ]−1 The following gives now a complete description for the lattice of invariant subspaces of the adjointable operators. Theorem 2.3. Let E a Hilbert C*-module over a C*-algebra A. Then lat L(E) = {E(J ) J ⊆ hE, Ei closed left ideal } and the association J 7→ E(J ) establishes a complete lattice isomorphism between the closed left ideals of hE, Ei and lat L(E). In addition, lat K(E) = lat L(E) = lat EndA(E). Proof. First observe that if J ⊆ A is a closed left ideal, then the subspace E(J ) is invariant under L(E), because L(E) consists of A-module operators. Conversely assume that M ∈ lat L(E) and let J(M ) ≡ span{hη, mi η ∈ E and m ∈ M }. Clearly, J(M ) ⊆ hE, Ei and the identity a hη, mi = hηa∗, mi , a ∈ A, η ∈ E, m ∈ M, 4 E. G. KATSOULIS implies that J(M ) is a left ideal. We claim that M = E(J(M )). Indeed, if m ∈ M , then by the definition of J(M ) we have hη, mi ∈ J(M ), for all η ∈ E, and so Lemma 2.2 implies that m ∈ E(J(M )). On the other hand, any ξa, with ξ ∈ E and a ∈ J(M ) is the limit of finite sums of elements of the form ξ hη, mi, where η ∈ E and m ∈ M . However ξ hη, mi = θξ,η(m) ∈ M and so M = E(J(M )). This shows that J 7→ E(J ) is surjective. In order to prove that J 7→ E(J ) is also injective we need to verify that J = J(E(J )), for any closed ideal J ⊆ hE, Ei. Since J ⊆ hE, Ei is a left ideal, J(E(J )) ⊆ J . On the other hand, if (ei)i is a right approximate identity for J , then any element of J ⊆ hE, Ei can be approximated by elements of the form X k hηk, ξki ek = X k hηk, ξkeki , ηk, ξk ∈ E. However, ξkek ∈ E(J ), by Definition 2.1, and so sums of the above form belong to J(E(J )). Hence J ⊆ J(E(J )) and so J 7→ E(J ) is also injective with inverse M 7→ J(M ). The proof that J 7→ E(J ) respects the lattice operations follows from two successive applications of Lemma 2.2. Indeed, if (Ji)i is a collection of closed ideals of hE, Ei, then ξ ∈ ∩iE(Ji) is equivalent by Lemma 2.2 to hη, ξi ∈ ∩iJi which, once again by Lemma 2.2, is equivalent to ξ ∈ E(∩iJi). Therefore ∩iE(Ji) = E(∩iJi). The proof of ∨iE(Ji) = E(∨iJi) is immediate. For the final assertion of the theorem, first note that lat K(E) ⊇ lat L(E) ⊇ lat EndA(E). On the other hand, if M ∈ lat K(E), then an argument identical to that of the second paragraph of the proof shows that M = E(J(M )). Hence M ∈ lat EndA(E) and the conclusion follows. The following result was proved by B. Johnson [11, Theorem 1] for arbi- trary semisimple Banach algebras by making essential use of their represen- tation theory. One can adopt Johnson's original proof to the C*-algebraic context by using the GNS construction and Kadison's Transitivity Theorem wherever representation theory is required in the original proof. Theorem 2.4. Let A be a C ∗-algebra and let Φ be a linear operator acting on A that leaves invariant all closed left ideals of A. Then Φ(ba) = Φ(b)a, ∀ a, b ∈ A. In particular, if 1 ∈ A is a unit then Φ is the left multiplication operator by Φ(1). Note that the proof of Theorem 2.3 shows that any bounded A-module map leaves invariant lat L(E). This establishes one direction in the following, which is the main result of the paper. THE REFLEXIVE CLOSURE OF THE ADJOINTABLE OPERATORS 5 Theorem 2.5. Let E be a Hilbert module over a C*-algebra A. Then alg lat L(E) = EndA(E). In particular, EndA(E) is a reflexive algebra of operators acting on E. Proof. Let S ∈ alg lat L(E) and ξ, η ∈ E. Consider the linear operator Φη,ξ : A ∋ a 7−→ hη, S(ξa)i ∈ A We claim that Φη,ξ leaves invariant any of the closed left ideals of A. Indeed, if J ⊆ A is such an ideal and j ∈ J , then ξj ∈ E(J ) and since S ∈ alg lat L, S(ξj) ∈ E(J ). By Theorem 2.3, we have Φη,ξ(j) = hη, S(ξj)i ∈ J and so Φη,ξ leaves J invariant, which proves the claim. Hence Theorem 2.4, implies now that Φη,ξ(ba) = Φη,ξ(b)a, ∀ a, b ∈ A. Let (ei) be an approximate unit for A. By the above Φη,ξ(eia) = Φη,ξ(ei)a, ∀i, and so hη, S(ξa)i = lim i Φη,ξ(eia) i hη, S(ξei)i a hη, S(ξeia)i = lim = lim i Φη,ξ(ei)a = lim i = hη, S(ξ)i a Hence which establishes that S is an A-module map. hη, S(ξa)i = hη, S(ξ)ai , ∀a ∈ A, The above Theorem can also be thought as a generalization of Theo- rem 2.4 (Johnson's Theorem) since its statement reduces to the statement of Theorem 2.4 when applied to the case of the trivial unital Hilbert C ∗- module. Corollary 2.6. If E is a selfdual Hilbert C ∗-module, then L(E) is reflexive as an algebra of operators acting on E. In particular, the above Corollary shows that if A is a unital C ∗-algebra, then L(A(n)), 1 ≤ n < ∞, is a reflexive operator algebra. This is not necessarily true for L(A(∞)). Indeed in [19, Example 2.1.2] the authors give an example of a unital commutative C ∗-algebra A for which L(A(∞)) 6= EndA(A(∞)). By Theorem 2.5, L(A(∞)) is not reflexive. 3. Left Centralizers and a theorem of H. Lin An alternative description for the inclusion L(E) ⊆ EndA(E) has been given by H. Lin in [18]. Definition 3.1. If A is a Banach algebra then a linear and bounded map Φ : A → A is called a left centralizer if Φ(ab) = Φ(a)b, for all a, b ∈ A. If in addition there exists a map Ψ : A → A so that Ψ(a)b = aΦ(b), for all a, b ∈ A, then Φ is called a double centralizer. 6 E. G. KATSOULIS The collection of all left (resp. double) centralizers equipped with the supremum norm will be denoted as LC(A) (resp. DC(A)). Note that in the case where A has an approximate unit, the linearity and boundedness of centralizers does not have to be assumed a priori but instead follows from the condition Φ(ab) = Φ(a)b, for all a, b ∈ A. (See [12] for a proof; the unital case is of course trivial.) In [18, Theorem 1.5] Lin shows that EndA(E) is isometrically isomorphic as a Banach algebra to LC (K(E)). Furthermore, the isomorphism Lin constructs extends the familiar ∗-isomorphism of Kasparov [14] between L(E) and DC(K(E)). Lin's proof is similar in nature to that of Kasparov [14] for the double centralizers of K(E). However it is more elaborate and also requires some additional results of Paschke [21]. In what follows we give an elementary proof of Lin's Theorem. Our argument depends on the observation that the representation theory for the left centralizers of a C ∗- algebra A is flexible enough to allow the use of representations on a Banach space. Definition 3.2. Let X be a Banach space and let A be a norm closed sub- algebra of B(X), the bounded operators on X. The left multiplier algebra of A is the collection LMX(A) ≡ {b ∈ B(X) ba ∈ A, for all a ∈ A}. If b ∈ LMX(A), then Lb ∈ B(A) denotes the left multiplication operator by b. The following has also a companion statement for double centralizers, which we plan to state and explore elsewhere. Proposition 3.3. Let A be a C ∗-algebra and assume that A is acting iso- metrically and non-degenerately on a Banach space X. Then the mapping (1) LMX(A) −→ LC(A) : b 7−→ Lb establishes an isometric Banach algebra isomorphism between LMX(A) and LC(A). Proof. The statement of this Proposition is a well-known fact, provided that X is a Hilbert space. In that case, in order to establish the surjectivity of (1) one starts with a contractive approximate unit (ei)i for A. If B ∈ LC (A), then the net (B(ei))i is bounded and therefore has at least one weak limit point b ∈ B(X). The conclusion then follows by showing that b ∈ LMX(A). (See [22, Proposition 3.12.3] for a detailed argument.) Bounded nets of operators on a Banach space need not have weak limits. However, the non-degeneracy of the action and the identity B(ei)ax = B(eia)x, a ∈ A, x ∈ X, guarantees that the net (B(ei)x)i is convergent when x ranges over a dense subset of X. Since (B(ei))i is bounded, we obtain that (B(ei)x)i is Cauchy THE REFLEXIVE CLOSURE OF THE ADJOINTABLE OPERATORS 7 (and thus convergent) for any x ∈ X. This establishes that (B(ei))i con- verges pointwise to some bounded operator b ∈ B(X), even when X is as- sumed to be a Banach space. With this observation at hand, the rest of the proof now goes as in the Hilbert space case. We are in position now to give the promised proof for Lin's Theorem. Theorem 3.4. Let E be a Hilbert C ∗-module over a C ∗-algebra A. Then there exists an isometric isomorphism of Banach algebras φ : EndA(E) −→ LC (K(E)) , whose restriction φL(E) establishes a ∗-isomorphism between L(E) and DC(K(E)). Proof. In light of Proposition 3.3, it suffices to verify that LME(K(E)) = EndA(E). Clearly EndA(E) ⊆ LME(K(E)). Conversely, let S ∈ LME(K(E)). a ∈ A and η, ξ, ζ ∈ E, then If S(η hξ, ζi a) = Sθη,ξ(ζa) = Sθη,ξ(ζ)a = S(η hξ, ζi)a. However vectors of the form η hξ, ζi, η, ξ, ζ ∈ E, are dense in E by [19, Lemma 1.3.9] and so S is an A-module map, as desired. Specializing now the mapping of (1) to our setting, we obtain an isometric isomorphism (2) φ : EndA(E) −→ LC(K(E)) : S 7−→ LS. Furthermore, the restriction φL(E) coincides with Kasparov's map and the conclusion follows. Acknowledgements. The present paper grew out of discussions between the author and Aristides Katavolos during the International Conference on Op- erator Algebras, which was held at Nanjing University, China, June 20-23, 2013. The author would like to thank Aristides for the stimulating conver- sations and is grateful to the organizers of the conference for the invitation to participate and their hospitality. References [1] M. Anoussis, A. Katavolos and M. Lambrou, On the reflexive algebra with two in- variant subspaces, J. Operator Theory 30 (1993), 267-299. [2] A, Arias and G. Popescu, Factorization and reflexivity on Fock spaces, Integral Equa- tions Operator Theory 23 (1995), 268-286. [3] W.B. Arveson, Operator algebras and invariant subspaces, Ann. Math. (2) 100 (1974), 433 -- 532. [4] D. Blecher and C. Le Merdy, Operator algebras and their modules: an operator space approach, London Mathematical Society Monographs, New Series 30, Oxford Uni- versity Press, 2004. 8 E. G. KATSOULIS [5] J. Bracic, V. Muller and M. Zajac, Reflexivity and hyperreflexivity of the space of locally intertwining operators, J. Operator Theory 63 (2010), 101-114. [6] K. Davidson, E. Katsoulis and D. Pitts, The structure of free semigroup algebras, J. Reine Angew. Math. 533 (2001), 99-125. [7] J. Erdos, Reflexivity for subspace maps and linear spaces of operators, Proc. London Math. Soc. 52 (1986), 582-600. [8] D. Hadwin, A general view of reflexivity, Trans. Amer. Math. Soc. 344 (1994), 325- 360. [9] D. Hadwin, E. Nordgren, H. Radjavi and P. Rosenthal, Orbit-reflexive operators, J. London Math. Soc. (2) 34 (1986), 111-119. [10] P. Halmos, Reflexive lattices of subspaces, J. London Math. Soc. (2) 4 (1971), 257 -- 263. [11] B.E. Johnson, Centralisers and operators reduced by maximal ideals J. London Math. Soc. 43 (1968), 231 -- 233. [12] B.E. Johnson, Continuity of centralisers on Banach algebras, J. London Math. Soc. 41 (1966), 639-640. [13] E. Kakariadis, Semicrossed products and reflexivity, J. Operator Theory 67 (2012), 379 -- 395. [14] G. Kasparov, Hilbert C ∗-modules: theorems of Stinespring and Voiculescu, J. Oper- ator Theory 4 (1980), 133-150. [15] A. Katavolos and S.C. Power, Translation and dilation invariant subspaces of L2(R), J. Reine Angew. Math. 552 (2002), 101-129. [16] E. Katsoulis, Local maps and the representation theory of operator algebras, manu- script. [17] E.C. Lance, Hilbert C*-Modules: A Toolkit for Operator Algebraists. Lecture note series: London Mathematical Society. Cambridge University Press, 1995. [18] H. Lin, Bounded module maps and pure completely positive maps, J. Operator Theory 26 (1991), 121-138. [19] V. Manuilov, E. Troitsky, Hilbert C ∗-modules, Translations of Mathematical Mono- graphs, 226. American Mathematical Society, Providence, RI, 2005. viii+202 pp. [20] R. Olin and J. Thomson, Algebras of subnormal operators, J. Funct. Anal. 37 (1980), 271-301. [21] W. Paschke, Inner product modules over B ∗-algebras, Trans. Amer. Math. Soc. 182 (1973), 443 -- 468. [22] G. Pedersen, Gert K. C ∗-algebras and their automorphism groups, London Mathe- matical Society Monographs 14, Academic Press, Inc., London-New York, 1979. [23] J. Ringrose, On some algebras of operators, Proc. London Math. Soc. 15 (1965), 61-83. [24] D. Sarason, Invariant subspaces and unstarred operator algebras, Pacific J. Math. 17 (1966), 511-517. [25] V. Shulman and I. Todorov, On subspace lattices. I. Closedness type properties and tensor products, Integral Equations Operator Theory 52 (2005), 561-579. Department of Mathematics, East Carolina University, Greenville, NC 27858, USA E-mail address: [email protected]
1701.05414
1
1701
2017-01-19T13:53:14
Free quantitative fourth moment theorems on Wigner space
[ "math.OA", "math.PR" ]
We prove a quantitative Fourth Moment Theorem for Wigner integrals of any order with symmetric kernels, generalizing an earlier result from Kemp et al. (2012). The proof relies on free stochastic analysis and uses a new biproduct formula for bi-integrals. A consequence of our main result is a Nualart-Ortiz-Latorre type characterization of convergence in law to the semicircular distribution for Wigner integrals. As an application, we provide Berry-Esseen type bounds in the context of the free Breuer-Major theorem for the free fractional Brownian motion.
math.OA
math
FREE QUANTITATIVE FOURTH MOMENT THEOREMS ON WIGNER SPACE SOLESNE BOURGUIN AND SIMON CAMPESE Abstract. We prove a quantitative Fourth Moment Theorem for Wigner in- tegrals of any order with symmetric kernels, generalizing an earlier result from Kemp et al. (2012). The proof relies on free stochastic analysis and uses a new biproduct formula for bi-integrals. A consequence of our main result is a Nualart-Ortiz-Latorre type characterization of convergence in law to the semicircular distribution for Wigner integrals. As an application, we provide Berry-Esseen type bounds in the context of the free Breuer-Major theorem for the free fractional Brownian motion. 1. Introduction Let (A , ϕ) be a tracial W ∗-probability space, S be a semicircular random vari- able and F = In(f ) be a self-adjoint Wigner integral (for a simple example, take off-diagonal homogeneous sums of a semicircular system). Recently, Kemp et al. showed in [KNPS12] that for a sequence of such Wigner integrals, convergence of the fourth moment controls convergence in distribution towards the semicircular law. Moreover, they provided a quantitative bound in terms of the free gradient operator, which is of the form (all unexplained notation appearing in this section will be introduced in the sequel) (1) dC2 (F, S) ≤ 1 2 ϕ ⊗ ϕ(cid:18)(cid:12)(cid:12)(cid:12)(cid:12) Z ∇s(cid:0)N −1 0 F(cid:1) ♯ (∇sF )∗ ds − 1 ⊗ 1(cid:12)(cid:12)(cid:12)(cid:12) (cid:19) . Here, dC2 is a distance that metrizes free convergence in distribution (see Definition 2.4), ∇ denotes the free gradient operator first introduced by Biane and Speicher in [BS98] and N −1 stands for the pseudo-inverse of the number operator (see Section 2). showed in [KNPS12] that the gradient expression appearing in (1) can further be bounded by the fourth moment. To be more precise, it holds that In the special case of Wigner integrals of order two, Kemp et al. 0 dC2 (I2(f ), S) ≤ (2) ≤ 1 2 ϕ ⊗ ϕ(cid:18)(cid:12)(cid:12)(cid:12)(cid:12) Z ∇s(cid:0)N −1 2r 3 2pϕ (I2(f )4) − 2. 1 0 I2(f )(cid:1) ♯ (∇sI2(f ))∗ ds − 1 ⊗ 1(cid:12)(cid:12)(cid:12)(cid:12) (cid:19) 2010 Mathematics Subject Classification. 46L54, 68H07, 60H30. Key words and phrases. Free probability, Wigner integrals, free Malliavin calculus, free sto- chastic analysis, free quantitative central limit theorems, free Fourth Moment Theorems. 1 2 SOLESNE BOURGUIN AND SIMON CAMPESE A question left open in the aforementioned article is whether a similar fourth mo- ment bound holds for Wigner integrals of higher orders, as is the case in the commu- tative setting (see Nualart and Peccati [NP05] and Nourdin and Peccati[NP09b]). In this paper, we provide a positive answer to this question by proving fourth mo- ment bounds for Wigner integrals of any order with symmetric kernels. Our main result can be paraphrased as follows (see Theorem 3.7 for a precise statement). Theorem. For a Wigner integral F of order n with normalized symmetric kernel it holds that ϕ ⊗ ϕ (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)  ZR+ ∇s(cid:0)N −1 2 0 F(cid:1) ♯ (∇sF )∗ ds − 1 ⊗ 1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)  ≤ Cn(cid:16)ϕ(cid:0)F 4(cid:1) − 2(cid:17). The constant Cn grows asymptotically linearly with n and is a local maximum of a certain polynomial (see Theorem 3.7 for full details). Combined with (1), our result quantifies the free Fourth Moment Theorem [KNPS12, Theorem 1.3] for the case of Wigner integrals with symmetric kernels. In particular, √C2 = 1 2 , so that, by Cauchy-Schwarz, the bound (2) is included as a special case. 2q 3 It is well-known that in order to ensure that Wigner integrals are self-adjoint (and thus free random variables), the symmetry of the kernel can be relaxed to mirror- symmetry (see Definition 2.7). As our main bound is stated for symmetric kernels, the natural question arises whether or not it can be generalized to cover the mirror- symmetric case as well. The answer to this question is negative, as is shown by the counterexample in Remark 3.9. In the proof of our main result we use a new biproduct formula (see Theorem 3.5) for Wigner bi-integrals (see Subsection 2.2) which generalizes the product for- mula proved by Biane and Speicher in [BS98] for usual Wigner integrals. In this biproduct formula, the nested contractions become what we call bicontractions. As product formulae play a central role in free (and also classical) stochastic analysis, this might be of independent interest. Other ingredients include the free Malliavin calculus introduced by Biane and Speicher in [BS98] as well as a fine combinato- rial analysis. A direct consequence of our bound is a Nualart-Ortiz-Latorre type equivalent condition for convergence towards the semicircular law, which reads as follows (see Theorem 3.10 for a precise statement). Theorem. A sequence Fk of Wigner integrals of order n with normalized symmetric kernels converges in law to the standard semicircular distribution if, and only if, ZR+ (∇sFk) ♯ (∇sFk)∗ ds → n · 1 ⊗ 1 in L2 (A ⊗ A , ϕ ⊗ ϕ) . This is a free analogue of the main result of [NOL08]. Our findings contribute to the growing literature on free limit theorems obtained by means of free Malliavin calculus and free stochastic analysis. Earlier results include the already mentioned free Fourth Moment Theorem for multiple Wigner integrals [KNPS12], its multidimensional extension [NPS13], the free Fourth Mo- ment Theorem for free Poisson multiple integrals proved in [BP14b] and [Bou16], FREE QUANTITATIVE FOURTH MOMENT THEOREMS 3 free non-central limit theorems for Wigner and free Poisson integrals obtained in [DN12], [NP13] and [Bou15], as well as limit theorems for the q-Brownian mo- tion [DNN13] and convergence of free processes [NT14]. However, all these results, with the exception of [KNPS12] for the case of second order Wigner integrals, are not quantitative. In the commutative setting, which inspired this line of re- search in the context of free probability theory, the picture is much more com- plete. Here, quantitative limit theorems exist in the framework of Wiener integrals ([NP05, PT05, NOL08, NP09b, NPR10, NP09a] and references therein), Poisson integrals ([PSTU10, PZ10, Pec11, BP14a, PT13] and references therein) and eigen- functions of diffusive Markov generators ([Led12, ACP14, CNPP16]). The rest of this paper is organized as follows: Section 2 introduces the basic concepts of free probability theory and free stochastic analysis. The biproduct formula, our fourth moment bound, as well as the Nualart-Ortiz-Latorre characterization are presented and proved in Section 3. We conclude by providing a Berry-Esseen bound for the free Breuer-Major theorem for the free fractional Brownian motion in Section 4. 2. Preliminaries 2.1. Elements of free probability. In the following, a short introduction to free probability theory is provided. For a thorough and complete treatment, see [NS06], [VDN92] and [HP00]. Let (A , ϕ) be a tracial W ∗-probability space, that is A is a von Neumann algebra with involution ∗ and ϕ : A → C is a unital linear functional assumed to be weakly continuous, positive (meaning that ϕ (X) ≥ 0 whenever X is a non-negative element of A ), faithful (meaning that ϕ (XX ∗) = 0 ⇒ X = 0 for every X ∈ A ) and tracial (meaning that ϕ (XY ) = ϕ (Y X) for all X, Y ∈ A ). The self-adjoint elements of A will be referred to as random variables. Given a random variable X ∈ A , the law of X is defined to be the unique Borel measure on R having the same moments as X (see [NS06, Proposition 3.13]). The non- commutative space L2(A , ϕ) denotes the completion of A with respect to the norm kXk2 =pϕ (XX ∗). Definition 2.1. A collection of random variables X1, . . . , Xn on (A , ϕ) is said to be free if ϕ ([P1 (Xi1 ) − ϕ (P1 (Xi1 ))]··· [Pm (Xim ) − ϕ (Pm (Xim ))]) = 0 whenever P1, . . . , Pm are polynomials and i1, . . . , im ∈ {1, . . . , n} are indices with no two adjacent ij equal. Let X ∈ A . The k-th moment of X is given by the quantity ϕ(X k), k ∈ N0. Now assume that X is a self-adjoint bounded element of A (in other words, X is a bounded random variable), and write ρ(X) = kXk ∈ [0,∞) to indicate the spectral radius of X. Definition 2.2. The law (or spectral measure) of X is defined as the unique Borel probability measure µX on the real line such that RR P (t) dµX (t) = ϕ(P (X)) for every polynomial P ∈ R [X]. A consequence of this definition is that µX has support in [−ρ(X), ρ(X)]. 4 SOLESNE BOURGUIN AND SIMON CAMPESE The existence and uniqueness of µX in such a general framework are proved e.g. in [Tao12, Theorem 2.5.8] (see also [NS06, Proposition 3.13]). Note that, since µX has compact support, the measure µX is completely determined by the sequence (cid:8)ϕ(X k) : k ≥ 1(cid:9). Let {Xn : n ≥ 1} be a sequence of non -- commutative random variables, each possibly belonging to a different non-commutative probability space (An, ϕn). Definition 2.3. The sequence {Xn : n ≥ 1} is said to converge in distribution if to a limiting non-commutative random variable X∞ (defined on (A∞, ϕ∞)), limn→+∞ ϕn(P (Xn)) = ϕ∞(P (X∞)) for every polynomial P ∈ R[X]. If Xn, X∞ are bounded (and therefore the spectral measures µXn , µX∞ are well- defined), this last relation is equivalent to saying that ZR P (t) µXn (dt) →ZR P (t) µX∞ (dt). An application of the method of moments yields immediately that, in this case, one has also that µXn weakly converges to µX∞ , that is µXn (f ) → µX∞ (f ), for every f : R → R bounded and continuous (note that no additional uniform boundedness assumption is needed). Let h(x) = RR eixξν(dξ) be the Fourier transform of a complex measure ν on R. Note that, as ν is finite, h is continuous and bounded. For such functions h, define the seminorm I2(h) by I2(h) =ZR ξ2ν(dξ). Let C2 denote the set of those functions h for which I2(h) < ∞. Using the seminorm I2 and the set of functions C2, one can define a distance between two self-adjoint random variables. Definition 2.4. For two self-adjoint random variables X, Y , the distance dC2 (X, Y ) between X and Y is defined as dC2 (X, Y ) = sup{ϕ(h(X)) − ϕ(h(Y )) : h ∈ C2, I2(h) ≤ 1} . As is proved in [KNPS12], the distance dC2 is weaker than the Wasserstein distance but still metrizes convergence in law. Definition 2.5. The centered semicircular distribution with variance t > 0, de- noted by S(0, t), is the probability distribution given by S(0, t)(dx) = (2πt)−1p4t − x2dx, x < 2√t. Definition 2.6. A free Brownian motion S consists of: (i) a filtration {At : t ≥ 0} of von Neumann sub-algebras of A (in particular, As ⊂ At for 0 ≤ s < t), (ii) a collection S = {St : t ≥ 0} of self-adjoint operators in A such that: (a) S0 = 0 and St ∈ At for all t ≥ 0, (b) for all t ≥ 0, St has a semicircular distribution with mean zero and variance t, and (c) for all 0 ≤ u < t, the increment St − Su is free with respect to Au, and has a semicircular distribution with mean zero and variance t − u. FREE QUANTITATIVE FOURTH MOMENT THEOREMS 5 all complex-valued functions on Rn Lebesgue measure on Rn +. For every integer n ≥ 1, the space L2(cid:0)Rn Definition 2.7. Let n be a natural number and let f be a function in L2(cid:0)Rn +(cid:1). +(cid:1) denotes the collection of + that are square-integrable with respect to the (1) The adjoint of f is the function f ∗ (t1, . . . , tn) = f (tn, . . . , t1). +; C(cid:1) = L2(cid:0)Rn (2) The function f is called mirror-symmetric if f = f ∗, i.e., if f (t1, . . . , tn) = f (tn, . . . , t1) for almost all (t1, . . . , tn) ∈ Rn sure. + with respect to the product Lebesgue mea- (3) The function f is called (fully) symmetric if it is real-valued and, for any permutation σ in the symmetric group Sn, it holds that f (t1, . . . , tn) = + with respect to the prod- f(cid:0)tσ(1), . . . , tσ(n)(cid:1) for almost all (t1, . . . , tn) ∈ Rn uct Lebesgue measure. +(cid:1). Let p ≤ n∧m be a natural number. The p-th nested contraction f Definition 2.8. Let n, m be natural numbers and let f ∈ L2(cid:0)Rn L2(cid:0)Rm f and g is the L2(cid:16)Rn+m−2p p variables in f ⊗ g: ⌢ g(t1, . . . , tn+m−2p) = ZRp f +(cid:1) and g ∈ (cid:17) function defined by nested integration of the middle f (t1, . . . , tn−p, s1, . . . , sp) p ⌢ g of + p + In the case where p = 0, the function f g(sp, . . . , s1, tn−p+1, . . . , tn+m−2p)ds1 ··· dsp. ⌢ g is just given by f ⊗ g. 0 For f ∈ L2(cid:0)Rn +(cid:1), we denote by In(f ) the multiple Wigner integral of f with respect to the free Brownian motion as introduced in [BS98]. The space L2(S, ϕ) = {In(f ) : f ∈ L2(Rn +), n ≥ 0} is a unital ∗-algebra, with product rule given, for any n, m ≥ 1, +(cid:1), g ∈ L2(cid:0)Rm f ∈ L2(cid:0)Rn +(cid:1), by n∧m (3) In(f )Im(g) = Xp=0 In+m−2p(cid:16)f p ⌢ g(cid:17) and involution In(f )∗ = In(f ∗). For a proof of this formula, see [BS98]. Fur- thermore, as is well-known, multiple integrals of different orders are orthogonal in L2(A , ϕ), whereas for two integrals of the same order, the Wigner isometry (4) holds. ϕ (In(f )In(g)∗) = hf, giL2(Rn +) . Remark 2.9. Observe that it follows from the definition of the involution on the algebra L2(S, ϕ) that operators of the type In(f ) are self-adjoint if and only if f is mirror-symmetric. 6 SOLESNE BOURGUIN AND SIMON CAMPESE 2.2. Bi-integrals and free gradient operator. This subsection introduces the notion of bi-integral and the action of the free gradient operator on Wigner integrals. For a full treatment of these objects, see [BS98]. +(cid:1). Then, the +(cid:1) ∼= +(cid:1) ⊗ L2(cid:0)Rm (cid:1). From the Wigner isometry (4) for multiple integrals, we obtain the so +(cid:17)⊗ L2(cid:16)Rm′ + (cid:17) +(cid:1) ⊗ L2(cid:0)Rm Let n, m be two positive integers and f = g ⊗ h ∈ L2(cid:0)Rn Wigner bi-integral In ⊗ Im(f ) is defined as In ⊗ Im(f ) = In(g) ⊗ Im(h). This definition is extended linearly to generic elements f ∈ L2(cid:0)Rn L2(cid:0)Rn+m +(cid:1) and g ∈ L2(cid:16)Rn′ called Wigner bisometry: for f ∈ L2(cid:0)Rn ϕ⊗ϕ (In ⊗ Im(f )In′ ⊗ Im′ (g)∗) =(hf, giL2(Rn +(cid:1)⊗ L2(cid:0)Rm it holds that (5) if n = n′ and m = m′, + ) +)⊗L2(Rm otherwise 0 + Remark 2.10. Observe that, for any natural numbers n, m and any function g⊗ h ∈ L2(cid:0)Rn +(cid:1), it holds that +(cid:1) ⊗ L2(cid:0)Rm In ⊗ Im (g ⊗ h)∗ = (In (g) ⊗ Im (h))∗ = In (g)∗ ⊗ Im (h)∗ = In (g∗) ⊗ Im (h∗) = In ⊗ Im(cid:0)(g ⊗ h)∗(cid:1) , so that the operator In ⊗ Im (g ⊗ h) is self-adjoint if and only if both the function g and h are mirror-symmetric. By continuous extension (using the Wigner bisometry (5)), it holds that for any fully symmetric function f ∈ L2(cid:0)Rn +(cid:1), the operator In ⊗ Im (f ) is self-adjoint. Let (A , ϕ) be a W ∗-probability space. An A ⊗ A -valued stochastic process t 7→ Ut is called a biprocess. For p ≥ 1, U is an element of Bp, the space of Lp-biprocesses, if its norm +(cid:1) ⊗ L2(cid:0)Rm kUk2 Bp =Z ∞ 0 kUtk2 Lp(A ⊗A ,ϕ⊗ϕ) dt is finite. The free gradient operator ∇ : L2 (S, ϕ) → B2 is a densely-defined and closable operator whose action on Wigner integrals is given by ∇tIn(f ) = n Xk=1 t (cid:17) , Ik−1 ⊗ In−k(cid:16)f (k) where f (k) t (x1, . . . , xn−1) = f (x1, . . . , xk−1, t, xk, . . . , xn−1) is viewed as an element + (cid:1) ⊗ L2(cid:0)Rn−k + (cid:1). of L2(cid:0)Rk−1 Remark 2.11. For general elements of L2 (S, ϕ) in its domain, the free gradient is customarily defined via a Fock space construction (see [BS98]). This level of generality will not be needed in the sequel. We will also make use of the pseudo-inverse of the number operator N −1 action on a multiple Wigner integral of order n ≥ 1 is given by N −1 1 n In(f ). 0 , whose 0 In(f ) = FREE QUANTITATIVE FOURTH MOMENT THEOREMS 7 Before concluding this section, we introduce ♯ to be the associative action of A ⊗ A op (where A op denotes the opposite algebra) on A ⊗ A , as (6) Furthermore, we also write ♯ to denote the action of A ⊗ L2 (R+) ⊗ A op on A ⊗ L2 (R+) ⊗ A , as (A ⊗ B)♯(C ⊗ D) = (AC) ⊗ (DB). (A ⊗ f ⊗ B)♯(C ⊗ g ⊗ D) = (AC) ⊗ f g ⊗ (DB). The multiplication ♯ naturally appears in the following bound from [KNPS12] on the dC2 distance introduced above. Theorem 2.12 ([KNPS12]). Let S be a standard semicircular random variable and F ∈ L2(S, ϕ) be self-adjoint, in the domain of the free gradient ∇ and such that ϕ(F ) = 0. Then, (7) dC2 (F, S) ≤ 1 2 ϕ ⊗ ϕ (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ZR+ ∇s(cid:0)N −1 0 F(cid:1) ♯ (∇sF )∗ ds − 1 ⊗ 1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ! . 3. Main results 3.1. Bicontractions and biproduct formula. As announced in the introduc- tion, we will need an extension of the product formula (3) from [BS98]. To this end, we introduce the notion of bicontraction. Definition 3.1. Let n1, m1, n2, m2 be positive integers. Let f ∈ L2(cid:0)Rn1 + (cid:1) ∼= L2(cid:0)Rn1+m1 L2(cid:0)Rm1 p ≤ n1 ∧ n2, r ≤ m1 ∧ m2 be natural numbers. The (p, r)-bicontraction f is the L2(cid:16)Rn1+n2−2p + (cid:1) ∼= L2(cid:0)Rn2+m2 + (cid:1) ⊗ L2(cid:0)Rm2 (cid:1) ∼= L2(cid:16)Rn1+n2+m1+m2−2p−2r (cid:1) and g ∈ L2(cid:0)Rn2 (cid:17) ⊗ L2(cid:0)Rm1+m2−2r + (cid:1) ⊗ (cid:1) and let (cid:17) function defined by p,r ⌢ g + + + + + f p,r ⌢ g(t1, . . . , tn1+n2+m1+m2−2p−2r) =ZRp+r + f (t1, . . . , tn1−p,sp, . . . , s1, y1, . . . , yr, tn1+n2+m2−2p−r+1, . . . , tn1+n2+m1+m2−2p−2r) × g (s1, . . . , sp, tn1−p+1, . . . , tn1+n2+m2−2p−r, yr, . . . , y1) ds1 ··· dspdy1 ··· dyr. Remark 3.2. Observe that for f = f1 ⊗ f2 and g = g1 ⊗ g2 with f1 ∈ L2(cid:0)Rn1 + (cid:1), f2 ∈ L2(cid:0)Rm1 + (cid:1), the above definition reads + (cid:1), g1 ∈ L2(cid:0)Rn2 ⌢ g = (f1 ⊗ f2) + (cid:1) and g2 ∈ L2(cid:0)Rm2 ⌢ (g1 ⊗ g2) =(cid:16)f1 where the contractions appearing on the right-hand side are the nested contractions introduced in Definition 2.8. ⌢ g1(cid:17) ⊗(cid:16)g2 ⌢ f2(cid:17) , (8) p,r p p,r f r p,r ⌢ g and Remark 3.3. In what follows, for f, g as in Definition 3.1, we write f s ⌢ g to denote the bicontraction and contraction of f and g, respectively. Here, f we have somewhat abused notation by using the same symbol for a function living in L2(cid:0)Rn1 + (cid:1)⊗ L2(cid:0)Rm1 + (cid:1) or its identification in L2(cid:0)Rn1+m1 + (cid:1). However, it will always 8 SOLESNE BOURGUIN AND SIMON CAMPESE be clear from the type of contraction used which version of the function is being considered. The following result collects some properties of bicontractions in the case where both functions are symmetric. + (cid:1) ∼= L2(cid:0)Rn1+m1 Lemma 3.4. For n1, m1, n2, m2 ∈ N, let f ∈ L2(cid:0)Rn1 (cid:1) and g ∈ L2(cid:0)Rn2 (cid:1) be fully symmetric functions. Further- more, let p ≤ n1 ∧ n2 and r ≤ m1 ∧ m2 be natural numbers such that p + r = p′ + r′. Then, the following is true. + (cid:1) ∼= L2(cid:0)Rn2+m2 + (cid:1)⊗ L2(cid:0)Rm1 + (cid:1)⊗L2(cid:0)Rm2 + + (i) f p,r ⌢ g ∼= f p+r ⌢ g. (ii) f p,r ⌢ g = f p′,r′ ⌢ g. n1+n2−2p + )⊗L2(R m1 +m2−2r + ) n1+n2+m1+m2−2p−2r + ) . L2(R (iii) (cid:13)(cid:13)(cid:13) (iv) f 2 f p,r L2(R ⌢ g(cid:13)(cid:13)(cid:13) n1,m1⌢ f = kfk2 L2(cid:0)Rm1 + (cid:1). 2 f =(cid:13)(cid:13)(cid:13) p+r ⌢ g(cid:13)(cid:13)(cid:13) L2(R n1 + )⊗L2(R m1 + ) 1 ⊗ 1, which is a constant in L2(cid:0)Rn1 + (cid:1) ⊗ Proof. Just exploit the full symmetry of f in the above definition of contractions. (cid:3) We are now ready to state the biproduct formula, which will be a crucial tool in order to prove our main result. Theorem 3.5. For n1, m1, n2, m2 ∈ N, let f ∈ L2(cid:0)Rn1 and g ∈ L2(cid:0)Rn2 (9) In1 ⊗ Im1 (f ) ♯In2 ⊗ Im2 (g) = + (cid:1)⊗L2(cid:0)Rm1 + (cid:1) ∼= L2(cid:0)Rn2+m2 (cid:1). Then it holds that In1+n2−2p ⊗ Im1+m2−2r(cid:16)f Xr=0 + (cid:1) ∼= L2(cid:0)Rn1+m1 ⌢ g(cid:17) . + (cid:1) ⊗ L2(cid:0)Rm2 Xp=0 m1∧m2 n1∧n2 p,r + + (cid:1) +(cid:1) and b ∈ L2(cid:0)Rm Proof. Using a density argument together with the bisometry property of Wigner bi-integrals, it is enough to prove the claim for functions f and g of the type a ⊗ b where a ∈ L2(cid:0)Rn +(cid:1) as the subset of functions +(cid:1) , b ∈ L2(cid:0)Rm +(cid:1)(cid:9) + (cid:1), d ∈ L2(cid:0)Rm2 + (cid:1). It holds that In1 ⊗ Im1 (a ⊗ b) ♯In2 ⊗ Im2 (c ⊗ d) = In1 (a) ⊗ Im1 (b) ♯In2 (c) ⊗ Im2 (d) = In1 (a) · In2 (c) ⊗ Im1 (d) · Im2 (b) . (cid:8)a ⊗ b : a ∈ L2(cid:0)Rn + (cid:1) and g = c ⊗ d with c ∈ L2(cid:0)Rn2 +(cid:1)). Let therefore f = a ⊗ b with a ∈ L2(cid:0)Rn1 + (cid:1), is dense in L2(cid:0)Rn b ∈ L2(cid:0)Rm1 +(cid:1) ⊗ L2(cid:0)Rm FREE QUANTITATIVE FOURTH MOMENT THEOREMS 9 Using the usual multiplication formula for Wigner integrals on both sides of the tensor product, we get In1 (a) · In2 (c) ⊗ Im1 (d) · Im2 (b) = n1∧n2 = n1∧n2 Xp=0 Xp=0 Xp=0 n1∧n2 = m1∧m2 m1∧m2 r p ⌢ c(cid:17)! ⊗ m1∧m2 Xr=0 In1+n2−2p ⊗ Im1+m2−2r(cid:16)(cid:16)a In1+n2−2p ⊗ Im1+m2−2r(cid:16)(a ⊗ b) In1+n2−2p(cid:16)a Xr=0 Xr=0 Im1+m2−2r(cid:16)d ⌢ c(cid:17) ⊗(cid:16)d ⌢ b(cid:17)! ⌢ b(cid:17)(cid:17) ⌢ (c ⊗ d)(cid:17) , p,r p r where the last equality follows from the identity (8). (cid:3) Remark 3.6. 1. By taking m1 = m2 = 0, f = u ⊗ 1 and g = v ⊗ 1, we recover the usual product formula (3) for Wigner integrals. 2. Note that a similar version of the above biproduct formula also holds for the usual tensor product (with a slightly different definition for the bicontractions). Furthermore, using the same methodology, one could also define contractions and product formulae for higher order tensors. 3.2. Quantitative Fourth Moment Theorems. We are now in the position of stating the main result of this paper, namely a bound on the quantity appearing in the right hand side of (7) in terms of the fourth moment, which then leads to a quantitative Fourth Moment Theorem for multiple Wigner integrals. (10) L2(Rn +) = 1. Then, it holds that ZR+ ∇s(cid:0)N −1 Theorem 3.7. For n ∈ N, let F = In (f ) be a Wigner integral of order n with f ∈ L2(cid:0)Rn +(cid:1) symmetric and such that kfk2 2 ϕ ⊗ ϕ (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 0 F(cid:1) ♯ (∇sF )∗ ds − 1 ⊗ 1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)  ≤ Cn(cid:16)ϕ(cid:0)F 4(cid:1) − 2(cid:17),  n2 max {Pn (⌊u0⌋) , Pn (⌈u0⌉)} with 1 u2(n − u + 1)(cid:0)2(n − u)2 + 4(n − u) + 3(cid:1) , 3 5(cid:18)4(n + 1) − 1 2n2 + 4n − 3 3√2r(n) (cid:19) where Cn = 1 r(n) 3√4 − Pn(u) = u0 = (11) and r(n) = 3q4n3 + 12n2 + 5√2p4n4 + 16n3 + 20n2 + 8n + 5 + 22n + 14. Proof. In the following we will use the shorthand f (k) by s to denote the function given f (k) s (x1, . . . , xn−1) = f (x1, . . . , xk−1, s, xk+1, . . . , xn). 10 SOLESNE BOURGUIN AND SIMON CAMPESE Observe that ZR+ (∇sF ) ♯ (∇sF )∗ds = = n Xk,q=1ZR+ Xk,q=1ZR+ n Ik−1 ⊗ In−k(cid:16)f (k) s (cid:17) ♯(cid:16)Iq−1 ⊗ In−q(cid:16)f (q) s (cid:17)(cid:17)∗ ds Ik−1 ⊗ In−k(cid:16)f (k) s (cid:17) ♯Iq−1 ⊗ In−q(cid:16)f (q) s (cid:17) ds, where the last equality follows from the full symmetry of the function f . Using the product formula for bi-integrals proven in Theorem 3.5 yields Ik+q−2p−2 ⊗ I2n−k−q−2r(cid:16)f (k) s p,r ⌢ f (q) s (cid:17) ds, (∇sF ) ♯ (∇sF )∗ ds Xk,q=1ZR+ = n (k∧q)−1 n−(k∨q) Xp=0 Xr=0 and by a Fubini argument one gets ZR+ ZR+ (∇sF ) ♯ (∇sF )∗ ds Xp=0 Xk,q=1 (k∧q)−1 = n n−(k∨q) Xr=0 Ik+q−2p−2 ⊗ I2n−k−q−2r ZR+ s ds! . for any 1 ≤ k, q ≤ n, which together ⌢ f . Hence, s = f (q) s s ds = f p,r ⌢ f (q) f (k) s p+r+1 The full symmetry of f implies that f (k) p,r ⌢ f (q) f (k) s n p+r+1 (12) (k∧q)−1 n−(k∨q) Xr=0 Xp=0 Xk,q=1 Ik+q−2p−2⊗I2n−k−q−2r(cid:16)f with Lemma 3.4 yields RR+ ZR+ ⌢ f(cid:17) . (∇sF ) ♯ (∇sF )∗ ds = Exactly those summands Ik+q−2p−2 ⊗ I2n−k−q−2r(cid:16)f ⌢ f(cid:17) for which k + q − 2p − 2 = 0 and 2n − k − q − 2r = 0 yield the constant term kfk2 +) · 1 ⊗ 1 (i.e. a constant in L2 (R+) ⊗ L2 (R+)). These conditions, along with the ranges of summation, imply that k = q and p + r + 1 = n. Therefore, fixing k, for which we have n possibilities, fixes the other three indices q,p and r to take the values k, k − 1 and n − k, respectively. Recalling that kfk2 +) = 1, (12) can thus be rewritten as ZR+ (∇sF ) ♯ (∇sF )∗ ds L2(Rn L2(Rn p+r+1 = n · 1 ⊗ 1 + n Xk,q=1 (k∧q)−1 n−(k∨q) Xr=0 1{n−1−p−r>0} Xp=0 × Ik+q−2p−2 ⊗ I2n−k−q−2r(cid:16)f p+r+1 ⌢ f(cid:17) , FREE QUANTITATIVE FOURTH MOMENT THEOREMS 11 which, by using that N −1 0 F = 1 n F , gives 1 = ZR+ ∇s(cid:0)N −1 0 F(cid:1) ♯ (∇sF )∗ ds − 1 ⊗ 1 nZR+ Xk,q=1 (∇sF ) ♯ (∇sF )∗ ds − 1 ⊗ 1 Xp=0 n−(k∨q) (k∧q)−1 1 n = n 1{n−1−p−r>0} Xr=0 Ik+q−2p−2 ⊗ I2n−k−q−2r(cid:16)f ⌢ f(cid:17) 1{n−1−p−r>0} · Ik+q ⊗ I2(n−1−p−r)−k−q(cid:16)f p+r+1 p+r+1 ⌢ f(cid:17) . (13) = 1 n n−1 Xp,r=0 n−1−p−r Xk,q=0 Grouping all occuring bi-integrals by the order of the contraction, one arrives at (14) ZR+ ∇s(cid:0)N −1 0 F(cid:1) ♯ (∇sF )∗ ds − 1 ⊗ 1 Xu=1 1 n n−1 = 2(n−u) Xv=0 cu,vIv ⊗ I2(n−u)−v(cid:16)f u ⌢ f(cid:17) , where the cu,v are positive constants depending solely on u and v. Taking the trace of the square of (14) and using the Wigner bisometry (5) yields ϕ ⊗ ϕ (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ZR+ ∇s(cid:0)N −1 As is well known, = n−1 2(n−u) 1 n2 2(cid:17) 0 F(cid:1) ♯ (∇sF )∗ ds − 1 ⊗ 1(cid:12)(cid:12) u,v(cid:13)(cid:13)(cid:13) Xv=0 Xv=0 Xu=1 1≤u≤n−1 1 n2 max  2(n−u) ≤ c2 f n−1 Xu=1(cid:13)(cid:13)(cid:13) 2 u ⌢ f(cid:13)(cid:13)(cid:13) L2(R2n−2u + ) L2(R2n−2u + ) 2 u f c2 n−1 ⌢ f(cid:13)(cid:13)(cid:13) u,v Xu=1(cid:13)(cid:13)(cid:13)  = ϕ(cid:0)F 4(cid:1) − 2, f 2 u ⌢ f(cid:13)(cid:13)(cid:13) L2(R2n−2u + . ) so that it only remains to evaluate the maximum. To this end, the constants cu,v will be computed explicitly. By carefully comparing (13) with (14), one sees that cu,v is given by the cardinality of the set of all quadruples (p, r, k, q) satisfying the following conditions: 0 ≤ p, r ≤ n − 1; k + q = v; p + r < n − 1. 0 ≤ k, q ≤ n − 1 − p − r; p + r = u − 1; By reindexing the second sum in (13) via the transformation (k′, q′) = (n− 1 − p − r − k, n− 1− p− r− q), we see that cu,v = cu,2(n−u)−v, thus only the constants cu,v for which v ≤ n − u need to be computed explicitly. Fix u and v. Then, there are 12 SOLESNE BOURGUIN AND SIMON CAMPESE u couples (p, r) satisfying p + r = u − 1, namely (0, u − 1), (1, u − 2), . . . , (u − 1, 0). Likewise, there are v + 1 couples (k, q) satisfying k + q = v. Therefore, cu,v =(u(v + 1) u(2(n − u) − v + 1) if v ≤ n − u if v > n − u. This yields 2(n−u) n−u 2(n−u) c2 u,v = Xv=0 = = n−u u2(v + 1)2 + u2(2(n − u) − v + 1)2 Xv=0 Xv=0 u2 (n − u) (n − u + 1)(2(n − u) + 1) + u2(n − u + 1)2 Xv=n−u+1 Xv=0 u2(v + 1)2 + u2(v + 1)2 n−u−1 1 3 = Pn(u). Straightforward analysis shows that the polynomial Pn has exactly one maximum in the interval (1, n − 1) attained at u0 as defined in (11). Therefore, to maximize Pn, one has to select the closest integer to u0. (cid:3) Combining Theorem 3.7 with the bound appearing in (7) and applying the Cauchy- Schwarz inequality immediately yields the following quantitative free Fourth Mo- ment Theorem. Corollary 3.8. Let n ≥ 2 be a natural number and F = In (f ), where f is a symmetric function in L2(cid:0)Rn +) = 1. Let S be standard semi- circular random variable. Then it holds that, √Cn 2 pϕ (F 4) − 2, where Cn is the constant appearing in Theorem 3.7. +(cid:1) such that kfk2 dC2(F, S) ≤ L2(Rn Remark 3.9. 1. It holds that C2 = 3 2 , so that Corollary 3.8 becomes dC2 (I2(f ), S) ≤ 1 2r 3 2pϕ (I2(f )4) − 2, which is precisely the conclusion of [KNPS12, Corollary 1.12]. The next few values of Cn are given by C3 = 2, C4 = 19 4 yielding and dC2 (I3(f ), S) ≤ dC2 (I4(f ), S) ≤ of L2(cid:0)Rn +(cid:1) for each n = 2, 3, 4. 1 √2pϕ (I3(f )4) − 2, √19 4 pϕ (I4(f )4) − 2, where f is of course always chosen appropriately to be symmetric and an element FREE QUANTITATIVE FOURTH MOMENT THEOREMS 13 2. In general, a straightforward analysis shows that Cn grows with n. In the com- mutative case, when bounding the distance between a multiple Wiener integral of any order and the standard Gaussian distribution by means of the fourth mo- ment, the constants appearing in the bounds do not depend on the order of the multiple integral (see for example [NP09b]). If such a dimension-free bound also holds in the free case is not known and this question is left for future research. 3. As stated in the introduction, convergence of the fourth moment to 2 also im- plies convergence of multiple integral with mirror-symmetric kernels towards the semicircular distribution. As one needs the function f to be symmetric in The- orem 3.7, it is natural to ask if the bound (10) also holds for mirror-symmetric kernels. As the following counterexample shows, this is not true. Divide [0, 1] into N intervals I1, I2, . . . , IN of equal length 1 N and define the function fN (x1, x2, x3) = √N 1Ik×Ik (x1, x3) N Xk=1 on [0, 1]3. Observe that fN is a mirror-symmetric function in L2(cid:16)[0, 1]3(cid:17). Then, I3 (fN ) = √N N Xk=1 I1 (1Ik ) I1 (1) I1 (1Ik ) . It is easy to check that ϕ(cid:0)I3(fN )2(cid:1) = 1 and +(cid:13)(cid:13)(cid:13) ϕ(cid:16)I3 (fN )4(cid:17) − 2 =(cid:13)(cid:13)(cid:13) ⌢ fN(cid:13)(cid:13)(cid:13) L2([0,1]4) fN 1 2 fN 2 2 ⌢ fN(cid:13)(cid:13)(cid:13) implying (by the free Fourth Moment Theorem of [KNPS12]) that the sequence {I3 (fN ) : N ≥ 1} converges in distribution to the standard semicircular law. Furthermore, = 2 N , L2([0,1]2) ∇tI3 (fN ) = √N N Xk=1 [1Ik (t) ⊗ I1 (1) I1 (1Ik ) + I1 (1Ik ) ⊗ I1 (1Ik ) + I1 (1Ik ) I1 (1) ⊗ 1Ik (t)]. As fN is a sum of products of non-negative indicator functions, the quantity (15) ϕ ⊗ ϕ  (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ZR+ ∇s(cid:0)N −1 0 I3 (fN )(cid:1) ♯ (∇sI3 (fN ))∗ ds − 1 ⊗ 1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 2  N is a sum of non-negative terms. Hence, if one of these terms can be proven not to converge to zero, the entire quantity must be bounded away from zero as well. One of the summands appearing is √NPN k=1 I1 (1Ik ) ⊗ I1 (1Ik ). It holds that I1 (1Ik ) ⊗ I1 (1Ik )!∗ √N Xk=1 Xk,q=1 I1 (1Ik ) I1(cid:0)1Iq(cid:1) ⊗ I1(cid:0)1Iq(cid:1) I1 (1Ik ) I1 (1Ik ) ⊗ I1 (1Ik )! ♯ √N Xk=1 = N N N 14 SOLESNE BOURGUIN AND SIMON CAMPESE and a straightforward calculation shows that ϕ ⊗ ϕ  (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) N N Xk,q=1 2 I1 (1Ik ) I1(cid:0)1Iq(cid:1) ⊗ I1(cid:0)1Iq(cid:1) I1 (1Ik )(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)  = 1 + 3 N , which does not go to zero as N goes to infinity. In total, it holds that ϕ(cid:16)I3 (fN )4(cid:17)− 2 → 0 as N goes to infinity, but the quantity (15) is strictly greater than 1 for all N , hence proving that the quantity (15) can not be controlled by the fourth moment. Therefore, Theorem 3.7 can not be extended to mirror-symmetric ker- nels. In the commutative case, the classical Nualart-Ortiz-Latorre equivalence criterion for normal convergence of multiple Wiener integrals Fk = I W (DsFk)2 ds → n in L2 (Ω) , as k → ∞, n (fk) reads ZR+ where D denotes the Malliavin gradient and Ω stands for the underlying probability space (see [NOL08]). For Wigner integrals, an analogue of this criterion was only known to hold in the second chaos (see [KNPS12, Theorem 4.8]). Theorem 3.7 extends this to any order of chaos. Therefore, all equivalent criteria for normal convergence of Wiener integrals have now free analogues for convergence of Wigner integrals towards a semicircular distribution. For the sake of completeness, we collect these analogues in the following Theorem. Theorem 3.10. Let n ≥ 2 be a natural number and let {fk : k ≥ 1} be a sequences of symmetric functions in L2(cid:0)Rn +) = 1. For any k ≥ 1, denote Fk = In (fk). Then, the following conditions are equivalent: (i) The sequence {Fk : k ≥ 1} converges in law to the standard semicircular dis- (ii) As k tends to infinity, ϕ(cid:0)F 4 (iii) For all 1 ≤ p ≤ n − 1, as k tends to infinity, (cid:13)(cid:13)(cid:13) +(cid:1) such that, for all k ≥ 1, kfkkL2(Rn ⌢ fk(cid:13)(cid:13)(cid:13)L2(R2n−2p (iv) As k tends to infinity, k(cid:1) → 2. tribution. ) → 0. + p fk ZR+ (∇sFk) ♯ (∇sFk)∗ ds → n · 1 ⊗ 1 in L2 (S ⊗ S, ϕ ⊗ ϕ) . Proof. The equivalences (i) ⇔ (ii) ⇔ (iii) and the implication (iv) ⇒ (i) follow from [KNPS12, Theorem 1.6, Theorem 1.10]. The missing implication (ii) ⇒ (iv) follows from the main result of this section, namely Theorem 3.7. (cid:3) 4. Quantifying the free Breuer-Major theorem Our main results can be used to provide Berry-Esseen bounds for a free version of the Breuer-Major theorem (see [KNPS12]) for the free fractional Brownian motion. This can be regarded as a free analog of [NP09b, Theorem 4.1]. The free fractional FREE QUANTITATIVE FOURTH MOMENT THEOREMS 15 Brownian motion SH with index H ∈ (−1, 1) is defined as a centered semicircular process with covariance function t SH ϕ(cid:0)SH 1 s (cid:1) = 2(cid:0)t2H + s2H − t − s2H(cid:1) . As is well-known (see for example [BS98] or [NS06]), the orthogonal polynomials associated to the semicircular distribution are the Chebyshev polynomials Un of the second kind defined on [−2, 2] by the recurrence relations U0(x) = 1, U1(x) = x, and for n ≥ 2, For n ∈ N, define the increment sequence (cid:8)Xk = SH ward calculations show that the autocovariance function ρH(k) is given by k : k ≥ 0(cid:9). Straightfor- Un+1(x) = xUn(x) − Un−1(x). k+1 − SH ρH(k) = ϕ(X0Xk) = Furthermore, define {Vm : m ≥ 1} as 1 2(cid:0)k + 12H + k − 12H − 2k2H(cid:1) . Vm = 1 √m m−1 Xk=0 Un (Xk) . With these definitions in place, we can now state the announced Berry-Esseen bounds. Theorem 4.1. With the above notation prevailing, suppose that there exists an integer n ≥ 1 such that σ2 = Pk∈Z ρH (k)n < ∞. Then, there exists a positive constant Cn,H such that where the function α (n, H) is given by σ dC2(cid:18) Vm α (n, H) =  ,S(0, 1)(cid:19) ≤ Cn,H mα(n,H), 2 m− 1 mH−1 mnH−n+ 1 2 if H ∈(cid:0)0, 1 2(cid:3) , 2n−2i , if H ∈h 1 if H ∈h 2n−3 2 , 2n−3 2n−2 , 2n−1 2n (cid:17) . Proof. It is well known (see e.g. [Nou12, Proposition 2.5]), that the (non-free) fractional Brownian motion can be represented as a Wiener integral with respect to a standard Brownian motion as BH t =Z t 0 SH t =Z t 0 KH (t, u) dWu, KH (t, u) dSu. where the kernel KH(·,·) is explicit (see e.g. [Nou12, Proposition 2.5]). Using the correspondence between Wiener and Wigner integrals, it also holds that Indeed, this can be verified by checking that the covariance function of the above integral coincides with the one of the free fractional Brownian motion. Then, de- noting fk,m,H = mH(cid:18)KH(cid:18) k + 1 m ,·(cid:19) 1[0, k+1 m ] − KH(cid:18) k m ,·(cid:19) 1[0, k m ](cid:19) , 16 SOLESNE BOURGUIN AND SIMON CAMPESE it holds that m−1 1 √m Un (I1 (fk,m,H )) Vm = Xk=0 Observe that kfk,m,HkL2(R+) = 1 so that k,m,H(cid:17) = In 1 In(cid:16)f ⊗n 1 √m Vm = m−1 Xk=0 √m Define f ⊗n k,m,H! . m−1 Xk=0 gn,m,H = 1 σ√m f ⊗n k,m,H . m−1 Xk=0 σ dC2(cid:18) Vm ,S(0, 1)(cid:19) ≤ Applying Corollary 3.8 to Vm, we get √Cn 2 rϕ(cid:16)In (gn,m,H)4(cid:17) − 2 √Cn 2 vuut ⌢ gn,m,H(cid:13)(cid:13)(cid:13) Xu=1(cid:13)(cid:13)(cid:13) gn,m,H n−1 = u 2 L2(R2n−2u + . ) From here, one can evaluate and estimate the contraction norms similarly as in the proof of [NP09b, Theorem 4.1]. (cid:3) Acknowledgments. The authors wish to thank Roland Speicher for providing the counterexample appearing in Remark 3.9 and Tobias Mai and Roland Speicher for several stimulating discussions. S. Campese was partially supported by ERC grant 277742 Pascal. References [Bou16] [Bou15] [BP14a] [ACP14] Ehsan Azmoodeh, Simon Campese, and Guillaume Poly. Fourth Moment Theorems for Markov diffusion generators. Journal of Functional Analysis, 266(4):2341 -- 2359, 2014. Solesne Bourguin. Poisson convergence on the free Poisson algebra. Bernoulli. Of- ficial Journal of the Bernoulli Society for Mathematical Statistics and Probability, 21(4):2139 -- 2156, 2015. Solesne Bourguin. Vector-valued semicircular limits on the free Poisson chaos. Elec- tronic Communications in Probability, 21(55):1 -- 11, 2016. Solesne Bourguin and Giovanni Peccati. Portmanteau inequalities on the Poisson space: mixed regimes and multidimensional clustering. Electronic Journal of Probability, 19:no. 66, 42, 2014. Solesne Bourguin and Giovanni Peccati. Semicircular limits on the free Poisson chaos: Counterexamples to a transfer principle. Journal of Functional Analysis, 267(4):963 -- 997, August 2014. Philippe Biane and Roland Speicher. Stochastic calculus with respect to free Brow- nian motion and analysis on Wigner space. Probability Theory and Related Fields, 112(3):373 -- 409, 1998. [BP14b] [BS98] [CNPP16] Simon Campese, Ivan Nourdin, Giovanni Peccati, and Guillaume Poly. Multivariate Gaussian approximations on Markov chaoses. Electronic Communications in Probabil- ity, 21, 2016. Aur´elien Deya and Ivan Nourdin. Convergence of Wigner integrals to the tetilla law. ALEA. Latin American Journal of Probability and Mathematical Statistics, 9:101 -- 127, 2012. [DN12] [DNN13] Aur´elien Deya, Salim Noreddine, and Ivan Nourdin. Fourth moment theorem and q- Brownian chaos. Communications in Mathematical Physics, 321(1):113 -- 134, 2013. FREE QUANTITATIVE FOURTH MOMENT THEOREMS 17 [HP00] Fumio Hiai and D´enes Petz. The semicircle law, free random variables and entropy, volume 77 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2000. [KNPS12] Todd Kemp, Ivan Nourdin, Giovanni Peccati, and Roland Speicher. Wigner chaos and the fourth moment. The Annals of Probability, 40(4):1577 -- 1635, 2012. [Led12] M. Ledoux. Chaos of a Markov operator and the fourth moment condition. The An- nals of Probability, 40(6):2439 -- 2459, November 2012. Zentralblatt MATH identifier: 06114704. [Nou12] [NP05] [NP09a] [NP13] [NP09b] [NOL08] D. Nualart and S. Ortiz-Latorre. Central limit theorems for multiple stochastic integrals and Malliavin calculus. Stochastic Processes and their Applications, 118(4):614 -- 628, 2008. Ivan Nourdin. Selected aspects of fractional Brownian motion, volume 4 of Bocconi & Springer Series. Springer, Milan, 2012. David Nualart and Giovanni Peccati. Central limit theorems for sequences of multiple stochastic integrals. The Annals of Probability, 33(1):177 -- 193, 2005. Ivan Nourdin and Giovanni Peccati. Noncentral convergence of multiple integrals. The Annals of Probability, 37(4):1412 -- 1426, July 2009. Ivan Nourdin and Giovanni Peccati. Stein's method on Wiener chaos. Probability The- ory and Related Fields, 145(1-2):75 -- 118, 2009. Ivan Nourdin and Giovanni Peccati. Poisson approximations on the free Wigner chaos. The Annals of Probability, 41(4):2709 -- 2723, 2013. Ivan Nourdin, Giovanni Peccati, and Anthony R´eveillac. Multivariate normal approx- imation using Stein's method and Malliavin calculus. Annales de l'Institut Henri Poincar´e Probabilit´es et Statistiques, 46(1):45 -- 58, 2010. Ivan Nourdin, Giovanni Peccati, and Roland Speicher. Multi-dimensional Semicircular Limits on the Free Wigner Chaos. In Robert C. Dalang, Marco Dozzi, and Francesco Russo, editors, Seminar on Stochastic Analysis, Random Fields and Applications VII, number 67 in Progress in Probability, pages 211 -- 221. Springer Basel, January 2013. Alexandru Nica and Roland Speicher. Lectures on the combinatorics of free proba- bility, volume 335 of London Mathematical Society Lecture Note Series. Cambridge University Press, Cambridge, 2006. Ivan Nourdin and Murad S. Taqqu. Central and non-central limit theorems in a free probability setting. Journal of Theoretical Probability, 27(1):220 -- 248, 2014. Giovanni Peccati. The Chen-Stein method for Poisson functionals. arXiv:1112.5051, December 2011. [NPR10] [NT14] [Pec11] [NPS13] [NS06] [PSTU10] G. Peccati, J. L. Sol´e, M. S. Taqqu, and F. Utzet. Stein's method and normal approx- [PT05] [PT13] [PZ10] [Tao12] imation of Poisson functionals. The Annals of Probability, 38(2):443 -- 478, 2010. Giovanni Peccati and Ciprian A. Tudor. Gaussian limits for vector-valued multiple stochastic integrals. In S´eminaire de Probabilit´es XXXVIII, volume 1857 of Lecture Notes in Math., pages 247 -- 262. Springer, Berlin, 2005. Giovanni Peccati and Christoph Thale. Gamma limits and u-statistics on the Poisson space. ALEA. Latin American Journal of Probability and Mathematical Statistics, 10(1):525 -- 560, 2013. Giovanni Peccati and Cengbo Zheng. Multi-dimensional Gaussian fluctuations on the Poisson space. Electronic Journal of Probability, 15:no. 48, 1487 -- 1527, 2010. Terence Tao. Topics in random matrix theory, volume 132 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2012. [VDN92] D. V. Voiculescu, K. J. Dykema, and A. Nica. Free random variables, volume 1 of CRM Monograph Series. American Mathematical Society, Providence, RI, 1992. A noncommutative probability approach to free products with applications to random matrices, operator algebras and harmonic analysis on free groups. 18 SOLESNE BOURGUIN AND SIMON CAMPESE Boston University, Department of Mathematics and Statistics, 111 Cummington Mall, Boston, MA 02215, USA E-mail address: [email protected] University of Luxembourg, Mathematics Research Unit, 6, rue Richard Coudenhove- Kalergi, 1359 Luxembourg, Luxembourg E-mail address: [email protected]
1804.02203
2
1804
2019-03-27T15:03:23
The Category of Von Neumann Algebras
[ "math.OA", "cs.LO", "math.CT" ]
In this dissertation we study the category of completely positive normal contractive maps between von Neumann algebras. It includes an extensive introduction to the basic theory of $C^*$-algebras and von Neumann algebras.
math.OA
math
The Category of Von Neumann Algebras . A O h t a m [ 2 v 3 0 2 2 0 . 4 0 8 1 : v i X r a P ter aan op volgens in op dinsdag om 10.30 door Abraham Anton Westerbaan geboren te 1 Identifiers hdl: 2066/201611 arXiv: 1804.02203 isbn: 978-94-6332-484-7 Persistent links https://arxiv.org/abs/1804.02203 https://doi.org/2066/201611 https://hdl.handle.net/2066/201611 Source code LATEX https://github.com/westerbaan/theses https://github.com/westerbaan/ndpt cover Printed by GVO drukkers & vormgevers B.V., Ede, https://proefschriften.nl. Where applicable, © 2019 A.A. Westerbaan, c b available under cc by, [1]. The Category of Von Neumann Algebras Proefschrift ter verkrijging van de graad van doctor aan de Radboud Universiteit Nijmegen op gezag van de rector magnificus prof. dr. J.H.J.M. van Krieken, volgens besluit van het college van decanen in het openbaar te verdedigen op dinsdag 14 mei 2019 om 10.30 uur precies door Abraham Anton Westerbaan geboren op 30 augustus 1988 te Nijmegen 3 Promotor: Prof. dr. B.P.F. Jacobs Manuscriptcommissie: Prof. dr. J.D.M. Maassen Prof. dr. P. Panangaden (McGill University, Canada) Prof. dr. P. Selinger (Dalhousie University, Canada) Dr. C.J.M. Heunen (University of Edinburgh, Verenigd Koninkrijk) Dr. A.R. Kissinger The Category of Von Neumann Algebras Doctoral Thesis to obtain the degree of doctor from Radboud University Nijmegen on the authority of the Rector Magnificus prof. dr. J.H.J.M. van Krieken, according to the decision of the Council of Deans to be defended in public on Tuesday, May 14, 2019 at 10.30 hours by Abraham Anton Westerbaan born on August 30, 1988 in Nijmegen (the Netherlands) 5 Supervisor: Prof. dr. B.P.F. Jacobs Doctoral Thesis Committee: Prof. dr. J.D.M. Maassen Prof. dr. P. Panangaden (McGill University, Canada) Prof. dr. P. Selinger (Dalhousie University, Canada) Dr. C.J.M. Heunen (University of Edinburgh, United Kingdom) Dr. A.R. Kissinger 7 1 Introduction 2 C∗-algebras 2.1.1 Operators 2.1 Definition and Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2 The Basics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3 Positive Elements . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3.1 Holomorphic Functions . . . . . . . . . . . . . . . . . . . Spectral Radius . . . . . . . . . . . . . . . . . . . . . . . 2.3.2 2.3.3 The Square Root . . . . . . . . . . . . . . . . . . . . . . 2.4 Representation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.5 Matrices over C∗-algebras . . . . . . . . . . . . . . . . . . . . . . 2.6 Towards von Neumann Algebras . . . . . . . . . . . . . . . . . . 2.6.1 Directed Suprema . . . . . . . . . . . . . . . . . . . . . . 2.6.2 Normal Functionals . . . . . . . . . . . . . . . . . . . . . . . . by Continuous Functions Representation by Bounded Operators 2.4.1 2.4.2 3 Von Neumann Algebras 3.2.1 3.2.2 3.2.3 3.2.4 3.1 The Basics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1.1 Definition and Counterexamples . . . . . . . . . . . . . . Elementary Theory . . . . . . . . . . . . . . . . . . . . . 3.1.2 3.1.3 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2 Projections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Ceiling and Floor . . . . . . . . . . . . . . . . . . . . . . Range and Support . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Carrier and Commutant Central Support and Central Carrier . . . . . . . . . . . 3.3 Completeness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3.1 Closure of a Convex Subset . . . . . . . . . . . . . . . . . 3.3.2 Kaplansky's Density Theorem . . . . . . . . . . . . . . . 3.3.3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3.4 3.4 Division . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.4.1 (Approximate) Pseudoinverses . . . . . . . . . . . . . . . 3.4.2 Division . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.4.3 Polar Decomposition . . . . . . . . . . . . . . . . . . . . 3.4.4 Hereditarily Atomic Von Neumann Algebras . . . . . . . 3.5 Normal Functionals . . . . . . . . . . . . . . . . . . . . . . . . . 3.5.1 Ultraweak Boundedness . . . . . . . . . . . . . . . . . . . 3.5.2 Ultraweak Permanence . . . . . . . . . . . . . . . . . . . Closedness of Subalgebras Completeness 1 2 3 4 7 12 12 16 23 27 27 30 31 35 36 38 41 42 42 44 49 55 56 59 63 67 71 72 74 75 76 78 79 81 82 85 85 86 88 4 Assorted Structure in W∗cpsu 92 93 4.1 Measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94 Corner and Filter . . . . . . . . . . . . . . . . . . . . . . 4.1.1 99 Isomorphism . . . . . . . . . . . . . . . . . . . . . . . . . 4.1.2 Purity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100 4.1.3 4.1.4 Contraposition . . . . . . . . . . . . . . . . . . . . . . . . 101 4.1.5 Rigidity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102 4.1.6 ⋄-Positivity . . . . . . . . . . . . . . . . . . . . . . . . . . 103 4.2 Tensor product . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107 4.2.1 Definition . . . . . . . . . . . . . . . . . . . . . . . . . . . 108 4.2.2 Existence . . . . . . . . . . . . . . . . . . . . . . . . . . . 109 4.2.3 Universal Property . . . . . . . . . . . . . . . . . . . . . 112 4.2.4 Functoriality . . . . . . . . . . . . . . . . . . . . . . . . . 115 4.2.5 Miscellaneous Properties . . . . . . . . . . . . . . . . . . 116 4.2.6 Monoidal Structure . . . . . . . . . . . . . . . . . . . . . 119 4.3 Quantum Lambda Calculus . . . . . . . . . . . . . . . . . . . . . 120 First Adjunction . . . . . . . . . . . . . . . . . . . . . . . 122 4.3.1 Second Adjunction . . . . . . . . . . . . . . . . . . . . . 124 4.3.2 4.3.3 Free Exponential . . . . . . . . . . . . . . . . . . . . . . 125 4.3.4 Hereditarily Atomic Von Neumann Algebras . . . . . . . 126 4.4 Duplicators and Monoids . . . . . . . . . . . . . . . . . . . . . . 126 4.4.1 Duplicators . . . . . . . . . . . . . . . . . . . . . . . . . . 127 4.4.2 Monoids . . . . . . . . . . . . . . . . . . . . . . . . . . . 132 9 Chapter 1 Introduction 1 What does this Ph.D. thesis offer? Proof, perhaps, to the doctoral thesis com- mittee of passable academic work; an advertisement, as it may be, of my school's perspective to colleagues; a display, even, of intellectual achievement to friends and family. But I believe such narrow and selfish goals alone barely serve to keep a writer's spirits energised -- and are definitely detrimental to that of the readers. That is why I have foolhardily challenged myself not just to drily list contributions, but to write this thesis as the introduction, that I would have liked to read when I started research for this thesis back in May 2014. The topic is von Neumann algebras, the category they form, and how they may be used to model aspects of quantum computation. Let us just say for now that a von Neumann algebra is a special type of complex vector space endowed with a multiplication operation among some other additional structure. An important example is the complex vector space M2 of 2 × 2 complex matrices, because it models (the predicates on) a qubit; but all N × N -complex matrices form a von Neumann algebra MN as well. Using von Neumann algebras (and their little cousins, C∗-algebras) to describe quantum data types seems to be quite a recent idea (see e.g. [21, 29, 63], and [5] for an overview) and has two distinct features. Firstly, classical data types are neatly incorporated: C2 ≡ C ⊕ C models a bit, and the direct sum M2 ⊕ M3 models the union type of a qubit and a qutrit. Secondly, von Neumann algebras allow for infinite data types as well such as B(ℓ2(Z)), which represents a "quantum integer."∗ It should be said that this last feature is both a boon and a bane: it brings with it all the inherent intricacies of dealing with infinite dimensions; and it is no wonder that most authors choose to restrict themselves to finite dimensions, especially since ∗Though other methods of modelling infinite dimensional quantum computing have been proposed as well e.g. using non-standard analysis [24], pre-sheaves [52], the geometry of inter- action [27], and quantitative semantics [58]. 1.. 11 this seems to be enough to describe quantum algorithms, see e.g. [57]. II In this thesis, however, we do face infinite dimensions, because the two main results demand it: 1. For the first result, that von Neumann algebras form a model of Selinger and Valiron's quantum lambda calculus, as Cho and I explained in [11] and for which I'll provide the foundation here, we need to interpret function types, some of which are essentially infinite dimensional. 2. The second result, an axiomatisation of the map a 7→ √pa√p : A → A representing measurement of an element p ∈ [0, 1]A of a von Neumann algebra A was tailored by B.E. Westerbaan (my twin brother) and myself to work for both finite and infinite dimensional A . These results are part of a line of research that tries to find patterns in the cat- egory of von Neumann algebras, that may also be cut from other categories modelling computation -- ideally in order to arrive at categorical axioms for (probabilistic) computation in general. When I joined the fray the notion of effectus [30] had already been established by Jacobs, and the two results above offer potential additional axioms. The work in this area has largely been a col- laborative effort, primarily between Jacobs, Cho, my twin brother, and myself, and many of their insights have ended up in this thesis. Of this I'd say no more than that my work appears conversely, and propor- tionally, in their writings too, except that the close cooperation with my brother begs further explanation. Our efforts on certain topics have been like interleav- ing of the pages of two phone books: separating them would be nigh impossible, So that's why we decided to write our theses as two volumes of the same work; especially the work on the axiomatisation of a 7→ √pa√p and Paschke dilations. preliminaries on von Neumann algebras, and the axiomatisation of a 7→ √pa√p appear in this thesis, while the work on dilations, and effectus theory appear in my brother's thesis, [84]. III The two results mentioned above only make up about a third of this thesis; the rest of it is devoted to the introduction to the theory of von Neumann algebras needed to understand these results. My aim is that anyone with, say, a bachelor's degree in mathematics (more specifically, basic knowledge of linear algebra, analysis [66], topology [87] and set theory [17]) should at least be able to follow the lines of reasoning with only minimal recourse to external sources. But I hope that they will gain some deeper understanding of the material as well. To this end, and because I wanted to gain some of this insight for myself too, I've not just mixed and matched results from the literature, but I tailored a thorough treatise of everything that's needed, including proofs (except 121 III). Whenever possible, I've taken shortcuts (e.g. avoiding for example the theory of Banach algebras and locally convex spaces entirely) to prevent the mental tax the added concepts (and pages) would have brought. For the same reasons I've refrained from putting everything in its proper abstract (and categorical [51]) context trusting that it'll shine through of its own accord. I've however not been able to restrain myself in making perhaps frivolous variations on the existing theory whenever not strictly necessary, taking for example Kadison's characterisation [46] of von Neumann algebras as my definition, and developing the elementary theory for it; in my defence I'll just say this adds to the original element that is expected of a thesis. Advertisements Due to space -- time constraints this thesis is based only on a selection [8, 10, 11, 80, 81] of the works I produced under supervision of Jacobs, and while [9, 42, 82] are incorporated in my brother's thesis, this means [40, 41] are unfortunately left out. If you like this thesis, then you might also want to take a look at these [21,28,48,50,64,65] recent works on von Neumann algebras, and C∗-algebras. If you're curious about effectus theory and related matters, please have a look at [6 -- 9, 30 -- 44]. But if you'd like more pictures instead, I'd suggest [14]. Writing style I've replaced page numbers by paragraph numbers such as V for this paragraph. The numbers after 134 refer to paragraphs in my twin brother's thesis [84]. Definitions are set like that (i.e. in blue), and can be found in the index. Proofs of certain facts that are easily obtained on the back of an envelope, and would clutter this manuscript, have been left out. Instead these facts have been phrased as exercises as a challenge to the reader. The symbol ":=" should be interpreted as "is defined to be", while "≡" should be read as "being of the form". Sometimes "≡" is used to define something on its right-hand side, as in "let A ≡(cid:0) a b b∗ c(cid:1) be a self-adjoint matrix." Other times "≡" indicates a simple rewrite step, as in "since a = 2, we have a + 2 = 2 + 2 ≡ 4," where it's not suggested a = 2 implies 2 + 2 = 4. Acknowledgements The work in this thesis specifically has benefited greatly from discussions with John van de Wetering, Robert Furber, Kenta Cho, and Bas Westerbaan, but I've also had the pleasure of discussing a variety of other top- ics with Aleks Kissinger, Andrew Polonsky, Bert Lindenhovius, Frank Roumen, Hans Maassen, Henk Barendregt, Joshua Moerman, Martti Karvonen, Rob- bert Krebbers, Robin Adams, Robin Kaarsgaard, Sam Staton, Sander Uijlen, Sebastiaan Joosten, and many others. I'm especially honoured to have been received in Edinburgh by Chris Heunen and in Oberwolfach by Jianchao Wu. I'm very grateful to Arnoud van Rooij, Bas Westerbaan and John van de We- tering for proofreading large parts of this manuscript, without whose efforts even more shameful errors would have remained. I'm very grateful too for the manuscript committee's members' various suggestions and comments, and hope the improvements I made to this text do them justice. I should of course not forget to mention the contribution of friends (both close and distant), family, ..1.. 13 IV V Va VI and colleagues -- too numerous to name -- of keeping me sane these past years. This is the second dissertation topic I've worked on; my first attempt under different supervision was unfortunately cut short after 11/2 years. When Bart Jacobs graciously offered me a second chance, I initially had my reservations, but accepted on account of the challenging topic. Little did I know that behind the ambition and suit one finds a man of singular moral fibre, embodying what was said about von Neumann himself: "[he] had to understand and accept much that most of us do not want to accept and do not even wish to understand."† VII Funding was received from the European Research Council under grant agree- ment № 320571. †An excerpt from Eugene P. Wigner's writings, see page 130 of [85]. Chapter 2 C∗-algebras 2 We redevelop the essentials of the theory of (unital) C∗-algebras in this chapter. Since we are ultimately interested in von Neumann algebras (a special type of C∗-algebras) we will evade delicate topics such as tensor products (of C∗- algebras), quotients, approximate identities, and C∗-algebras without a unit. The zenith of this chapter is Gelfand's representation theorem (see 27 XXVII), the fact that every commutative (unital) C∗-algebra is isomorphic to the C∗- algebra C(X) of continuous functions on some compact Hausdorff space X -- it yields a duality between the category CH of compact Hausdorff spaces (and continuous maps) and the category cC∗miu of commutative C∗-algebras (and unital ∗-homomorphisms, the appropriate structure preserving maps), see 29. As the road to Gelfand's representation theorem is a bit winding -- involving intricate relations between technical concepts -- we have put emphasis on the invertible and positive elements so that the important theorems about them may serve as landmarks along the way: 1. first we show that the norm on a C∗-algebra is determined by the invertible elements (via the spectral radius), see 16 II; 2. then we construct a square root of a positive element in 23 VII; 3. and finally we show that an element of a commutative C∗-algebra is not invertible iff it is mapped to 0 by some multiplicative state, see 27 XV. At every step along the way the positive and invertible elements (and the norm, multiplicative states, multiplication and other structure on a C∗-algebra) are bound more tightly together until Gelfand's representation theorem emerges. To make this chapter more accessible we have removed much material from the ordinary development of C∗-algebras such as the more general theory of ..1 -- 2.. 15 Banach algebras (and its pathology). This forces us to take a slightly different path than is usual in the literature (see e.g. 16 VIII). After Gelfand's representation theorem we deal with two smaller topics: that a C∗-algebra may be represented as a concrete C∗-algebra of bounded operators on a Hilbert space (see 30 VI), and that the N × N -matrices with entries drawn from a C∗-algebra A form a C∗-algebra MN (A ) (see 33 I). We end with an overture to von Neumann algebras -- the topic of the next chapter. 2.1 Definition and Examples 3 Definition A C∗-algebra is a complex vector space A endowed with 1. a binary operation, called multiplication (and denoted as such), which is associative, and linear in both coordinates; 2. an element 1, called unit, such that 1 · a = a = a · 1 for all a ∈ A ; 3. a unary operation (· )∗, called involution such that (a∗)∗ = a, (ab)∗ = b∗a∗, (λa)∗ = ¯λa∗, and (a + b)∗ = a∗ + b∗ for all a, b ∈ A and λ ∈ C; 4. a complete norm k · k such that kabk 6 kakkbk for all a, b ∈ A , and ka∗ak = kak2 holds; this equality is called the C∗-identity. II Warning In the literature it is usually not required that a C∗-algebra possess a The C∗-algebra A is called commutative if ab = ba for all a, b ∈ A . unit; but when it does it is called a unital C∗-algebra. III Example The vector space C of complex numbers forms a commutative C∗- algebra in which multiplication and 1 have their usual meaning. Involution is given by conjugation (z∗ = ¯z), and norm by modulus (kzk = z). IV Example A C∗-subalgebra of a C∗-algebra A is a subset B of A , which is a linear subspace of A , contains the unit, 1, is closed under multiplication and involution, and is closed with respect to the norm of A ; such a C∗-subalgebra of A is itself a C∗-algebra when endowed with the operations and norm of A . V Example One can form products (in the categorical sense, see 20a I) of C∗- algebras as follows. Let Ai be a C∗-algebra for every element i of some index Ai set I. The direct sum of the family (Ai)i is the C∗-algebra denoted byLi∈I on the set Ai : (cid:8) a ∈Qi∈I supi∈I ka(i)k < ∞ (cid:9) whose operations are defined coordinatewise, and whose norm is a supremum Ai is norm given by kak = supi ka(i)k. If each Ai is commutative, then Li∈I commutative. In particular, taking Ai ≡ C, we see that the vector space ℓ∞(X) of bounded complex-valued functions on a set X forms a commutative C∗-algebra with pointwise operations and supremum norm. VI VII VIII 4 Example The bounded continuous functions on a topological space X form a commutative C∗-subalgebra BC(X) of ℓ∞(X) (see above). In particular, since a continuous function on a compact Hausdorff space is automatically bounded, we see that the continuous functions on a compact Hausdorff space X form a commutative C∗-algebra C(X) with pointwise operations and sup-norm. We'll see that every commutative C∗-algebra is isomorphic to a C(X) in 27 XXVII. Example An example of a non-commutative C∗-algebra is the vector space Mn of n × n-matrices (n > 1) over C with the usual (matrix) multiplication, the identity matrix as unit, and conjugate transpose as involution (so (A∗)ij = Aji). The norm kAk of a matrix A in Mn is less obvious, being the operator norm (cf. 4 II) of the associated linear map v 7→ Av, Cn → Cn, that is, kAk is the least number r > 0 with kAvk2 6 rkvk2 for all v ∈ Cn (where kwk2 = (Pi wi2)1/2 denotes the 2-norm of w ∈ Cn). It is not entirely obvious that kA∗Ak = kAk2 holds and that Mn is com- plete. We will prove these facts in the more general setting of bounded op- erators between Hilbert spaces, see 5 I. Suffice it to say, Cn is a Hilbert space with hv, wi = Pi viwi as inner product, each matrix gives a (bounded) linear map v 7→ Av, Cn → Cn, and the conjugate transpose A∗ is adjoint to A in the sense that hv, Awi = hA∗v, wi for all v, w ∈ Cn. Remark Combining V and VII we see that Lk Mnk is a finite-dimensional C∗-algebra for any tuple n1, . . . , nK of natural numbers. dimensional C∗-algebra is of this form as we'll see in 84 II.∗ In fact, any finite- 2.1.1 Operators Example Let us now turn to perhaps the most important and difficult exam- ple: we'll show that the vector space B(H ) of bounded operators on a Hilbert space H forms a C∗-algebra when endowed with the operator norm. Multipli- cation is given by composition, involution by taking the adjoint (see VIII), and unit by the identity operator. A concrete C∗-algebra or a C∗-algebra of bounded operators refers to a C∗-subalgebra of B(H ). We will eventually see that every C∗-algebra is isomorphic to a C∗-algebra of bounded operators in 30 XIV. ∗Although clearly related to the Wedderburn -- Artin theorem, see e.g. [56], this description of finite-dimensional C ∗-algebras does not seem to be an immediate consequence of it. ..2 -- 4.. 17 II Definition Let X and Y be normed vector spaces. We say that r ∈ [0,∞) is a bound for a linear map (=operator) T : X → Y when kT xk 6 rkxk for In all x ∈ X , and we say that T is bounded when there is such a bound. that case T has a least bound, which is called the operator norm of T , and is denoted by kTk. The vector space of bounded operators from X to Y is denoted by B(X , Y ), and the vector space of bounded operators from X to itself is denoted by B(X ). III Exercise Let X , Y and Z be normed complex vector spaces. 1. Show that the operator norm on B(X , Y ) is, indeed, a norm. 2. Let T : X → Y and S : Y → Z be bounded operators. Show that ST is bounded by kSkkTk, so that kSTk 6 kSkkTk. 3. Show that the identity operator id : X → X is bounded by 1. IV Exercise Let T : X → Y be a bounded operator between normed vector spaces, and let r ∈ [0,∞). Show that rkTk = supx∈(X )r kT xk, V where (X )r = {x ∈ X : kxk 6 r}. (The set (X )1 is called the unit ball of X .) Lemma The operator norm on B(X , Y ) is complete when Y is a complete normed vector space. VI Proof Let (Tn)n be a Cauchy sequence in B(X , Y ). We must show that (Tn)n converges to some bounded operator T : X → Y . Let x ∈ X be given. Since k Tnx − Tmxk = k (Tn − Tm) xk 6 kTn − Tmk kxk and kTn − Tmk → 0 as n, m → ∞ (because (Tk)k is Cauchy), we see that k Tnx − Tmxk → 0 as n, m → ∞, and so (Tnx)n is a Cauchy sequence in Y . Since Y is complete, (Tnx)n converges, and we may define T x := limn Tnx, giving a map T : X → Y , which is easily seen to be linear (by continuity of addition and scalar multiplication). It remains to be shown that T is bounded, and that (Tn)n converges to T with respect to the operator norm. Let ε > 0 be given, and pick N such that kTn − Tmk 6 1 2 ε for all n, m > N . Then for every x ∈ X we can find M > N with kT x − Tmxk 6 1 2 εkxk for all m > M , and so, for n > N , m > M , k(T − Tn)xk 6 kT x − Tmxk + kTmx − Tnxk 6 εkxk giving that T − Tn is bounded and kT − Tnk 6 ε for all n > N . Whence T is bounded too, and (Tn)n converges to T . (cid:3) From III and V it is clear that the complex vector space of bounded opera- tors B(X ) on a complete normed vector space X with composition as multi- plication and the identity operator as unit satisfies all the requirements to be a C∗-algebra that do not involve the involution, (· )∗ (that is, B(X ) is a Banach algebra). To get an involution, we need the additional structure provided by a Hilbert space as follows. Definition An inner product on a complex vector space V is a map h· , ·i : V × V → C such that, for all x, y ∈ V , hx, ·i : V → V is linear; hx, xi > 0; and hx, yi = hy, xi. We say that the inner product is definite when hx, xi = 0 =⇒ x = 0 for x ∈ V . A pre-Hilbert space H is a complex vector space endowed with a definite inner product. We'll shortly see that every such H carries a 1/2; if H is complete with respect to this norm, we norm given by kxk := hx, xi say that H is a Hilbert space. Let H and K be pre-Hilbert spaces. We say that an operator T : H → K is adjoint to an operator S : K → H when VII VIII hT x, yi = hx, Syi for all x ∈ H and y ∈ K . In that case, we call T adjointable. We'll see (in X) that such adjointable T is adjoint to exactly one S, which we denote by T ∗. Example We endow CN (where N is a natural number) with the inner product IX For an example of an infinite-dimensional Hilbert space, we'll have to wait The space c00 of sequences x1, x2, . . . for which xn is non-zero for finitely many n's is an example of a pre-Hilbert which is not complete when endowed given by hx, yi =Pi xiyi, making it a Hilbert space. with hx, yi =P∞n=0 xnyn as inner product. until 6 II where we'll show that the sequences x1, x2, . . . with Pn xn2 < ∞ form a Hilbert space ℓ2 with hx, yi =P∞n=0 xnyn as its inner product, because at this point it is not even clear that this sum converges. Exercise Let x and x′ be elements of a pre-Hilbert space H with hy, xi = hy, x′i for all y ∈ H . Show that x = x′ (by taking y = x − x′). Conclude that every operator between pre-Hilbert spaces has at most one adjoint. Remark Note that we did not require that an adjointable operator T : H → K between pre-Hilbert spaces be bounded, and in fact, it might not be. Take for example the operator T : c00 → c00 given by (T x)n = nxn, which is adjoint to itself, and not bounded. On the other hand, if either H or K is complete, then both T and T ∗ are automatically bounded as we'll see in 35 VI. X XI Exercise Let S and T be adjointable operators on a pre-Hilbert space. XII 1. Show that T ∗ is adjoint to T (and so T ∗∗ = T ). 2. Show that (T + S)∗ = T ∗ + S∗ and (λS)∗ = λS∗ for every λ ∈ C. ..4.. 19 3. Show that ST is adjoint to T ∗S∗ (and so (ST )∗ = T ∗S∗). We will, of course, show that every bounded operator on a Hilbert space is adjointable, see 5 XI. But let us first show that k · k defined in VIII is a norm, which boils down to the following fact about 2 × 2-matrices. v ) > 0 for all u, v ∈ C), XIII Lemma For a positive matrix A ≡(cid:0) p c XIV Proof Let u, v ∈ C be given. We have we have p, q > 0, and c2 6 pq. c q(cid:1) (i.e. ( u v ) A ( u 0 6 ( u v ) A ( u v ) = u2 p + uv c + uv c + v2 q. By taking u = 1 and v = 0, we see that p > 0, and similarly q > 0. The trick to see that c2 6 pq is to take v = 1 and u = tc with t ∈ R: 0 6 p c2 t2 + 2 c2 t + q. If p = 0, then −2 c2 t 6 q for all t ∈ R, which implies that c2 = 0 = pq. Suppose that p > 0. Then taking t = −p−1 we see that 0 6 c2 p−1 − 2 c2 p−1 + q = − c2 p−1 + q. (cid:3) Rewriting gives us c2 6 pq. XV Exercise Let h· , ·i be an inner product on a vector space V . Show that the for- mula kxk =phx, xi defines a seminorm on V , that is, kxk > 0, kλxk = λkxk, and -- the triangle inequality -- kx + yk 6 kxk + kyk for all λ ∈ C and x, y ∈ V . Moreover, prove that k · k is a norm when h· , ·i is definite; and for x, y ∈ V : 1. The Cauchy -- Schwarz inequality: hx, yi2 6 hx, xi hy, yi; 2. Pythagoras' theorem: kxk2 + kyk2 = kx + yk2 when hx, yi = 0; 3. The parallelogram law: kxk2 + kyk2 = 1 4P3 4. The polarisation identity: hx, yi = 1 (Hint: prove the Cauchy -- Schwarz inequality before the triangle inequality by applying XIII to the matrix(cid:0) hx,xi hx,yi using the inequalities hx, yi + hy, xi 6 2 hx, yi 6 2kxkkyk.) hy,xi hy,yi(cid:1). Then prove kx + yk2 6 (kxk +kyk)2 2 (kx + yk2 + kx − yk2 ); n=0 inkinx + yk2. XVI Lemma For an adjointable operator T on a pre-Hilbert space H kT ∗Tk = kTk2 and kT ∗k = kTk. Proof If T = 0, then T ∗ = 0, and the statements are surely true. XVII Suppose T 6= 0 (and so T ∗ 6= 0). Since kT xk2 = hT x, T xi = hx, T ∗T xi 6 kxk kT ∗T xk 6 kxk2kT ∗Tk for every x ∈ H by Cauchy -- Schwarz, we have kTk2 6 kT ∗Tk. Since kT ∗Tk 6 kT ∗kkTk and kTk 6= 0, it follows that kTk 6 kT ∗k. Since by a similar reasoning kT ∗k 6 kTk, we get kTk = kT ∗k. But then kTk2 6 kT ∗Tk 6 kT ∗kkTk = kTk2, and so kTk2 = kT ∗Tk. (cid:3) Exercise Given a Hilbert space H show that the adjointable operators form a closed subspace of B(H ). Exercise Let x and y be vectors from a Hilbert space H . XVIII XIX 1. Show that xihy : z 7→ hy, zi x defines a bounded operator H → H , and, moreover, that k xihy k = kxkkyk. 2. Show that xihy is adjointable, and (xihy)∗ = yihx. At this point it is clear that the vector space of adjointable operators on a Hilbert space forms a C∗-algebra. So to prove that B(H ) is a C∗-algebra, it remains to be shown that every bounded operator is adjointable (which we'll do in XI). We first show that each bounded functional f : H → C has an adjoint, see IX, for which we need the (existence and) properties of "projections" on (closed) linear subspaces: 5 Definition Let x be an element of a pre-Hilbert space H . We say that an element y of a linear subspace C of H is a projection of x on C if II kx − yk = min{ kx − y′k : y′ ∈ C }. (In other words, y is one of the elements of C closest to x.) Then y = he, xi e is the unique projection of x on eC. Exercise We'll see in VII that on a closed linear subspace every vector has a projection. For arbitrary linear subspaces this isn't so: show that the only vectors in ℓ2 having a projection on the linear subspace c00 (from 4 IX) are the vectors in c00 themselves. Lemma Let H be a pre-Hilbert space, and let x, e ∈ H with kek = 1. Proof Let y′ ∈ eC with y′ 6= y be given. To prove that y is the unique projection of x on eC it suffices to show that kx − yk < kx − y′k. Since y′ 6= y ≡ he, xi e, there is λ ∈ C, λ 6= 0 with y′ = (λ + he, xi)e. Note that he, yi = he,he, xi ei = he, xihe, ei = he, xi, and so he, x − yi = 0. Then y′−y ≡ λe and x−y are orthogonal too, and thus, by Pythagoras' theorem (see 4 XV), we have ky′− xk2 = ky′− yk2 +ky− xk2 ≡ λ2 +kx− yk2 > kx− yk2, because λ 6= 0. Hence ky′ − xk > ky − xk. (cid:3) ..4 -- 5.. 21 III IV V VI Exercise Let y be a projection of an element x of a pre-Hilbert space H on a linear subspace C. Show that y is a projection of x on yC. Conclude that y is the unique projection of x on C, and that hy, x − yi = 0. Show that y + c is the projection of x + c on C for every c ∈ C. Conclude that hy′, x − yi ≡ hy′, (x + y′ − y) − y′i = 0 for every y′ ∈ C. VII Projection Theorem Let C be a closed linear subspace of a Hilbert space H . VIII Proof We only need to show that there is a projection y of x on C, because VI Each x ∈ H has a unique projection y on C, and hy′, yi = hy′, xi for y′ ∈ C. gives us that such y is unique and satisfies hy′, yi = hy′, xi for all y′ ∈ C. Write r := inf{ kx − y′k : y′ ∈ C }, and pick a sequence y1, y2, . . . ∈ C such that kx − ynk → r. We will show that y1, y2, . . . is Cauchy. Let ε > 0 be given, and pick N such that kyn− xk2 6 r2 + 1 4 ε for all n > N . Let n, m > N be given. Then since 1 2 (yn +ym)−xk > 2r, and so by the parallelogram law (see 4 XV), 2 (yn +ym) is in C, we have kyn +ym−2xk ≡ 2k 1 kyn − ymk2 ≡ k(yn − x) − (ym − x)k2 = 2kyn − xk2 + 2kym − xk2 − kyn + ym − 2xk2 6 4r2 + ε − 4r2 6 ε. Hence y1, y2, . . . is Cauchy, and converges to some y ∈ C, because H is complete and C is closed. It follows easily that kx − yk = r, and thus y is the projection of x on C. (cid:3) IX Riesz' Representation Theorem Let H be a Hilbert space. For every bounded X linear map f : H → C there is a unique vector x ∈ H with hx, ·i = f . If f = 0, then x = 0 does the job. Suppose that f 6= 0. There is Proof an x′ ∈ H with f (x′) = 1. Note that ker(f ) is closed, because f is bounded. So by VII, we know that x′ has a projection y on ker(f ), and hx′, zi = hy, zi for all z ∈ ker(f ). Then for x′′ := x′ − y, we have f (x′′) = 1 and hx′′, y′i = 0 for all y′ ∈ ker(f ). Given z ∈ H , we have f ( z − f (z)x′′ ) = 0, so z − f (z)x′′ ∈ ker(f ), and thus 0 = hx′′, z − f (z)x′′i ≡ hx′′, zi − f (z)kx′′k2. Hence writing x := x′′kx′′k−2 we have f (z) = hx, zi for all z ∈ H . (cid:3) Finally, uniqueness of x follows from 4 X. XI Exercise Prove that every bounded operator T on a Hilbert space H is ad- jointable, as follows. Let x ∈ H be given. Prove that hx, T (· )i : H → C is a bounded linear map. Let Sx be the unique vector with hSx, ·i = hx, T (· )i, which exists by IX. Show that x 7→ Sx gives a bounded linear map S, which is adjoint to T . XII Thus the bounded operators on a Hilbert space H form a C∗-algebra B(H ) as described in 4 I. We will return to Hilbert spaces in 30 XIV, where we show that every C∗-algebra is isomorphic to a C∗-subalgebra of a B(H ). Here is a non-trivial example of a Hilbert space that will be used later on. Proposition Given a family (Hi)i∈I of Hilbert spaces, the vector space Li Hi = { x ∈Qi Hi : Pi kxik2 < ∞ }. is a Hilbert space when endowed with the inner product hx, yi =Pi hxi, yii. Proof To begin with we must show thatPi hxi, yii converges for x, y ∈Li Given ε > 0 we must find a finite subset G of I such that (cid:12)(cid:12)Pi∈F hxi, yii(cid:12)(cid:12) 6 ε for all finite F ⊆ I\G. Since an obvious application of the Cauchy -- Schwarz inequality gives us that for every finite subset F of I Hi. 2 (cid:12)(cid:12)(cid:12)Xi∈F hxi, yii(cid:12)(cid:12)(cid:12) 6 Xi∈F kxik2 Xi∈F kyik2, any G ⊆ I with Pi∈I\G kxik2 6 √ε and Pi∈I\G kyik2 6 √ε will do. It is easily seen that hx, yi := Pi hxi, yii gives a definite inner product on Li Hi; the remaining difficulty lies in showing that the resulting norm is complete. To this end, let x1, x2, . . . be a Cauchy sequence in Li∈I Hi; we must show that it converges to some x∞ ∈ Li Hi. We do the obvious thing: since for every i ∈ I the sequence (x1)i, (x2)i, . . . is Cauchy in Hi we may define (x∞)i := limn(xn)i, and thereby get an element x∞ ofQi Hi. Since for each fi- nite subset F of I we havePi∈F k(x∞)ik2 = limnPi∈F k(xn)ik2 6 limn kxnk2, we have Pi∈I k(x∞)ik2 6 limn kxnk2 < ∞, and so x∞ ∈Li It remains to be shown that x1, x2, . . . converges to x∞ (not only coordi- natewise but also) with respect to the inner product on Li Hi. Given ε > 0 pick N such that kxn − xmk 6 1 ε for all n, m > N . We claim that for such n 2√2 we have kx∞ − xk 6 ε. Indeed, first note that since the sum Xi∈I k(x∞)i − (xn)ik2 + Xi∈I\F converges (to kx∞ − xnk2), we can find a finite subset F (depending on n) such that second term in the right-hand side above is smaller than 1 2 ε2. To see that the first term is also below 1 k(x∞)i − (xn)ik2 ≡ Xi∈F 2 ε2, begin by noting that for every m, k(x∞)i − (xn)ik2 Hi. 1/2 1/2 (cid:16)Xi∈F 6 (cid:16)Xi∈F k(x∞)i−(xn)ik2(cid:17) k(x∞)i−(xm)ik2(cid:17) k(xm)i−(xn)ik2(cid:17) Since F is finite, and (xm) converges to x∞ coordinatewise we can find an m large enough that the first term on the right-hand side above is below 1 ε. If 2√2 we choose m > N we see that the second term is below 1 ε as well, and we 2√2 conclude that kx∞ − xnk 6 ε. (cid:3) +(cid:16)Xi∈F ..5 -- 6 23 6 II III 1/2 . 2.2 The Basics 7 Now that we have seen the most important examples of C∗-algebras, we can begin developing the theory. We'll start easy with the self-adjoint elements: II Definition Given an element a of a C∗-algebra A , 1. we say that a is self-adjoint if a∗ = a, and 2. we write aR := 1 2 (a + a∗) and aI := 1 part of a, respectively. 2i (a − a∗) for the real and imaginary The set of self-adjoint elements of A is denoted by AR. III Exercise Let a be an element of a C∗-algebra A . 1. Show that aR and aI are self-adjoint, and a = aR + iaI. 2. Show that if a ≡ b + ic for self-adjoint elements b, c of A , then b = aR and c = aI. 3. Show that (a∗)R = aR and (a∗)I = −aI. 4. Show that a is self-adjoint iff aR = a iff aI = 0. 5. Show that a 7→ aR and a 7→ aI give R-linear maps A → A . 6. Show that aI = −(ia)R and aR = (ia)I. 7. Show that a∗a is self-adjoint, and a∗a = a2 R + a2 I + i(aRaI − aIaR). 8. Give an example of A and a with aRaI 6= aIaR. (This inequality is a source of many technical difficulties.) 9. Show that a∗a + aa∗ = 2(a2 R + a2 I ). 10. The product of self-adjoint elements b, c need not be self-adjoint; show that, in fact, bc is self-adjoint iff bc = cb. 11. Show that ka∗k = kak. (Hint: kak2 = ka∗ak 6 ka∗kkak.) 12. Show that kaRk 6 kak and kaIk 6 kak. 13. Show that ka2k = kak2 when a is self-adjoint. However, show that ka2k 6= kak2 might occur when a is not self-adjoint. (Hint: (cid:0) 0 1 0 0(cid:1).) Notation Recall that (in this text) every C∗-algebra A has a unit, 1. Thus, for every scalar λ ∈ C, we have an element λ · 1 of A , which we will simply denote by λ. This should hardly cause any confusion, for while an expression of an element of A such as i + 2 + 5a (where a ∈ A ) may be interpreted in several ways, the result is always the same. Exercise There is a subtle point regarding the norm kλk of a scalar λ ∈ C inside a C∗-algebra A : we do not always have kλk = λ on the nose. 1. Indeed, show that k1k = 0 6= 1 when A = {0} is the trivial C∗-algebra. 2. Show that kλk 6 λ (in C). 3. Show that kλk = λ when kλk and λ are interpreted as elements of A . Let us now generalise the notion of a positive function in C(X) to a positive element of a C∗-algebra. There are several descriptions of positive functions in C(X) in terms of the C∗-algebra structure (see II) on which we can base such a generalisation, and while we will eventually see that these all yield the same notion of positive element of a C∗-algebra (see 25 I) we base our definition of positive element (IV) on the description that is perhaps not most familiar, but does give us the richest structure at this stage. 8 II 9 Exercise Let X be a compact Hausdorff space. Show that for self-adjoint f ∈ C(X), the following are equivalent. II 1. f (X) ⊆ [0,∞); 2. f ≡ g2 for some g ∈ C(X)R; 3. f ≡ g∗g for some g ∈ C(X); 4. kf − tk 6 t for some t ∈ R; 5. kf − tk 6 t for all t > 1 2kfk. (Hint: kf − tk 6 t iff −t 6 f − t 6 t iff 0 6 f 6 2t.) Exercise To see how condition 1 can be expressed in terms of the C∗-algebra structure of C(X), prove that λ ∈ f (X) iff f − λ is not invertible. Definition A self-adjoint element a of a C∗-algebra A is called positive if ka − tk 6 t for some t ∈ R. We write a 6 b for a, b ∈ A when b − a is positive, and we denote the set of positive elements of A by A+. Given elements a and b of a C∗-algebra A we denote by [a, b]A , or sometimes simply [a, b], the set of elements c of A with a 6 c 6 b. III IV IVa 7 -- 9.. 25 V Remark One advantage of this definition over, say, taking the elements of the form a∗a to be positive, is that it is immediately clear that an element b of a C∗-subalgebra B of a C∗-algebra A is positive in B iff b is positive in A -- that is, 'positive permanence' comes for free (cf. 11 XXIII). Another advantage is that it's also pretty easy to see that the sum of such positive elements is again positive, see VII. Va Remark Note that when an element a of a C∗-algebra is positive on the grounds that ka − tk 6 t for some t ∈ R, then this number t must be positive, and we even have t > 1 2kak, since kak − ktk 6 ka − tk 6 t. There's nothing special about this t: we'll see in 17 V that ka − sk 6 s for all s > 1 2kak and positive a. VI Example We'll see in 25 V, that a bounded operator T on a Hilbert space H is positive iff hx, T xi > 0 for all x ∈ H . VII Lemma Let a, b be positive elements of a C∗-algebra. Then a + b is positive. VIII Proof Since a > 0, there is t ∈ R with ka − tk 6 t. Similarly, there is s ∈ R (cid:3) with kb − sk 6 s. Then ka + b − (t + s)k 6 ka − tk + kb − sk 6 t + s. IX Exercise Given an element a of a C∗-algebra A with 0 6 a 6 1 (which is called an effect) show that the orthosupplement a⊥ := 1 − a is an effect too. Exercise Let A be a C∗-algebra. X 1. Show that A+ is a cone: 0 ∈ A+, a + b ∈ A+ for all a, b ∈ A+, and λa ∈ A+ for all a ∈ A+ and λ ∈ [0,∞). Conclude that 6 is a preorder. 2. Show that 1 is positive, and −kak 6 a 6 kak for every self-adjoint ele- ment a of A . (Thus 1 is an order unit of AR.) 3. The behaviour of positive elements may be surprising: give an example of positive elements a and b from a C∗-algebra such that ab is not positive. 4. Given a self-adjoint element a of A define kako = inf{ λ ∈ [0,∞) : − λ 6 a 6 λ }. Show that k−ko is a seminorm on AR, and that kako 6 kak for all a ∈ AR. Prove that 0 6 a 6 b implies that kako 6 kbko for a, b ∈ AR. 5. There is not much more that can easily be proven about positive elements, at this point, but don't take my word for it: try to prove the following facts about a self-adjoint element a of A directly. (a) a2 is positive; (b) if a is the limit of positive an ∈ A , then a is positive; (c) if a > − 1 n for all n ∈ N, then a > 0; (d) kak = kako; (e) a = 0 when 0 6 a 6 0. We will prove these facts when we return to the positive elements in 17. Let us spend some words on the morphisms between C∗-algebras. Definition A linear map f : A → B between C∗-algebras is called 10 II 1. multiplicative if f (ab) = f (a)f (b) for all a, b ∈ A ; 2. involution preserving if f (a∗) = f (a)∗ for all a ∈ A ; 3. unital if f (1) = 1; 4. subunital if f (1) 6 1; 5. positive if f (a) is positive for every positive a ∈ A , and 6. completely positive ifPi,j b∗i f ( a∗i aj ) bj is positive for all a1, . . . , an ∈ A , and b1, . . . , bn ∈ B (see Remark 5.1 of [60]). 7. (For normal maps, we refer to 38 I and 44 XV.) We use the bold letters as abbreviations, so for instance, f is pu if it is positive and unital, and a miu-map is a multiplicative, involution preserving, unital lin- ear map between C∗-algebras (which is usually called a unital ∗-homomorphism). We'll denote the category of C∗-algebras and miu-maps by C∗miu, and the subcategory of commutative C∗-algebras by cC∗miu. We'll use similar notation for the other classes of maps, but will, naturally, only mention C∗cpu after having established that cp-maps are closed under composition. The advantages of completely positive maps become apparent only later on when we start dealing with matrices (see 34 II) and the tensor product (see 115 II). Lemma ("p⇒i") A positive map f : A → B between C∗-algebras is involution preserving. Proof Let a ∈ A be given. We must show that f (a∗) = f (a)∗. But first we'll show that if a is self adjoint, then so is f (a). Indeed, since kak and kak − a are positive (see 9 X), we see that f (kak) and f (kak − a) are positive, and so f (a) = f (kak) − f (kak − a) being positive is self adjoint. It follows that f (a)R = f (aR) and f (a)I = f (aI) (for a ∈ A ), because f (a) ≡ f (aR) + if (aI), and f (aR) and f (aI) are self adjoint (see 7 III). Hence f (a∗) ≡ f (aR − iaI) = f (a)R − if (a)I ≡ f (a)∗. (cid:3) III IV V ..9 -- 10.. 27 VI Remark Other important relations between these types of morphisms can only be established later on once we have a firmer grasp on the positive elements. We will then see that every mi-map is completely positive (in 34 IV), and that every completely positive map is positive (in 25 II). VIa Note that we didn't bother to include an abbreviation for bounded linear maps in our list, II. That's because we'll see in 20 II that any positive map between C∗-algebras is automatically bounded. VII [Moved to 20a I.] VIII IX [Moved to 20a II.] [Moved to 20a III.] 11 II III After having visited the positive elements, let us explore our second landmark, the invertible elements of a C∗-algebra, whose role is as important as it is technical. This paragraph culminates in what is essentially spectral permanence (XXIII): the fact that if an element a of a C∗-subalgebra B of a C∗-algebra A is invertible in A , then a is already invertible in B, see XVI. Lemma Let a be an element of a C∗-algebra A with kak < 1. Then a⊥ ≡ 1− a has an inverse, namely (a⊥)−1 = P∞n=0 an. Moreover, this series converges absolutely, that is, P∞n=0 kank < ∞. Proof Note that (1 − kak) (1 + kak + kak2 + ··· + kakN ) = 1 − kakN +1, and so kakn = N Xn=0 1 − kakN +1 1 − kak for every N . Thus, since kakN converges to 0 (because† kak < 1), we get P∞n=0 kakn = (1−kak)−1. Note that since kank 6 kakn for every n, this entails that P∞n=0 kank 6 (1 − kak)−1 < ∞. IV Note that aN norm converges to 0, because kakN converges to 0. Also (but slightly less obvious), Pn an norm converges, because Pn kakn converges. V Thus, taking the norm limit on both sides of (1−a)(1+a+a2+··· aN ) = 1−aN +1, gives us (1− a)(Pn an) = 1. Since we can derive (Pn an)(1− a) = 1 in a similar manner, we see that Pn an is the inverse of 1 − a. (cid:3) 1. Show that a − λ is invertible for every λ ∈ C with kak < λ. 2. Show that a − b is invertible when b ∈ A is invertible and kak < kbk. 3. Show that U := { b ∈ A : b is invertible } is an open subset of A . † In case you've never seen the argument: the limit b := limN kakN exists, because kak > VI Exercise Let a be an element of a C∗-algebra A . kak2 > · · · > 0, and is zero because kakb = limN kakN +1 = b and kak < 1. and in that case converges absolutely. Lemma For a self-adjoint element a of A the seriesPn an converges iff kak < 1; Proof We have already seen in II thatPn an converges absolutely when kak < 1. Now, if Pn an converges, then kank (being the norm of the difference between consecutive partial sums of Pn an) converges to 0. In particular, kak2n (being equal to ka2n k by the C∗-identity) converges to 0 too, which only happens when kak < 1. (cid:3) Remark For non-self-adjoint elements a of A , the convergence of Pn an is a more delicate matter. Take for example the matrix A := (cid:0) 0 2 0 0(cid:1) for which the series Pn An converges (to 1 + A), while kAk = 2 -- the problem being that kA2k1/2 differs from kAk. In fact, we'll see from 13 II (although we won't need it) that Pn an converges absolutely when 1 > lim supn kank1/n, and diverges when 1 < lim supn kank1/n. This begs the question what happens when 1 = lim supn kank1/n -- which I do not know. Lemma Let A be a C∗-algebra. The assignment a 7→ a−1 gives a continuous map (from the set { b ∈ A : b is invertible} to A .) Proof (Based on Proposition 3.1.6 of [47].) 2 be given; we First we establish continuity at 1: let a ∈ A with k1 − ak 6 1 claim that a is invertible, and k1 − a−1k 6 2k1 − ak. 2 < 1, a is invertible by II, and a−1 =P∞n=0(1− a)n. Indeed, since k1− ak 6 1 Then k1− a−1k = kP∞n=1(1− a)nk 6P∞n=1 k1− akn = k1− ak (1−k1− ak)−1. Thus, as k1 − ak 6 1 2 , we get (1 − k1− ak)−1 6 2, and so k1 − a−1k 6 2k1− ak. Let a be an invertible element of A , and let b ∈ A with ka − bk 6 1 2ka−1k. We claim that b is invertible, and ka−1 − b−1k 6 2ka − bk ka−1k2. Since ka − bk 6 2ka−1k we have k1 − a−1bk 6 ka−1k ka − bk 6 1 2 . By XI, a−1b is invertible, and k1 − (a−1b)−1k 6 2k1 − a−1bk 6 2ka − bk ka−1k. Hence ka−1 − b−1k = k(1 − (a−1b)−1)a−1k 6 k1 − (a−1b)−1k ka−1k 6 2ka − bk ka−1k2. (cid:3) Lemma For a self-adjoint element a from a C∗-algebra, a − i is invertible. Proof (Based on Proposition 4.1.1(ii) of [47].) 1 The trick is to write a−i ≡ (a+ni) − (n+1)i for sufficiently large n, because then by VI a− i is invertible provided that n + 1 > ka + nik. Indeed, for n such that kak < 2n + 1, we have ka + nik2 = k(a + ni)∗(a + ni)k = ka2 + n2k 6 kak2 + n2 < 2n + 1 + n2 = (n + 1)2, and so ka + nik < n + 1. (cid:3) Exercise Let a be a self-adjoint element of a C∗-algebra. 1. Show that a − λ is invertible for all λ ∈ C\R. 2. Show that a2 − λ is invertible for all λ ∈ C\[0,∞). (Hint: first prove that a2 + 1 ≡ (a + i)(a − i) is invertible.) Conclude that an − λ is invertible for all λ ∈ C\[0,∞) and even n ∈ N. ..10 -- 11.. 29 VII VIII IX X XI XII XIII XIV XV 3. Let n ∈ N be odd. Show that an − λ is invertible for all λ ∈ C\[0,∞) if and only if a − λ is invertible for all λ ∈ C\[0,∞). (Hint: show that an + 1 =Qn k=1 a + ζ2k+1 where ζ = e πi n .) XVI Proposition Let A be a C∗-subalgebra of a C∗-algebra B. Let a be a self- adjoint element of A , which has an inverse, a−1, in B. Then a−1 ∈ A . XVII Proof While we do not know yet that a is invertible in A , we do know that a+i/n has an inverse (a + i/n)−1 in A by XV for each n (using that a is self-adjoint.) Since a+ i/n converges to a in B as n increases, we see that (a+ i/n)−1 converges to a−1 in B by X. Thus, as all (a + i/n)−1 are in A , and A is closed in B, we see that a−1 is in A . (cid:3) XVIII Exercise Show that the assumption in XVI that a is self-adjoint may be dropped. (Hint: consider a∗a, see Proposition VIII.1.14 of [15].) XIX Definition The spectrum, sp(a), of an element a of a C∗-algebra is the set of complex numbers λ for which a − λ is not invertible. XX Exercise Verify the following examples. 1. The spectrum of a continuous function f : X → R on a compact Hausdorff space X being an element of the C∗-algebra C(X) is the image of f , that is, sp(f ) = {f (x) : x ∈ X}. 2. The spectrum of a square matrix A from the C∗-algebra Mn is the set of eigenvalues of A. XXI Exercise Let a be an element of a C∗-algebra A . 1. Prove that sp(a) ⊆ R when a is self-adjoint (see XV). The reverse implication does not hold: show that sp((cid:0) 0 2 2. Show that sp(a2) ⊆ [0,∞) when a is self-adjoint (see XV). 3. Show that λ 6 kak for all λ ∈ sp(a) using VI. 0 0(cid:1)) = {0}. In fact, we will see in 16 II, that kak = sup{λ : λ ∈ sp(a)}. 4. Show that sp(a) is closed (using VI). Conclude that sp(a) is compact. 5. Show that sp(a + z) = {λ + z : λ ∈ sp(a)} for all z ∈ C. 6. Prove that sp(a−1) = {λ−1 : λ ∈ sp(a)} if a is invertible (and 0 /∈ sp(a)). On first sight, the spectrum sp(a) of an element a of a C∗-algebra A depends not only on a, but also on the surrounding C∗-algebra A for it determines for which λ ∈ C the operator a − λ is invertible. Thus we should perhaps write spA (a) instead of sp(a). However, such careful bookkeeping turns out be unnecessary by the following result. Theorem (Spectral Permanence) Let B be a C∗-subalgebra of a C∗-algebra A . Then spA (a) = spB(a) for every element a of B. Proof Let a be an element of B, and let λ ∈ C. We must show that a − λ is invertible in A iff a − λ is invertible in B. Surely, if a − λ has an inverse (a − λ)−1 in B, then (a − λ)−1 is also an inverse of a − λ in A , since B ⊆ A . The other, non-trivial, direction follows directly from XVI (and XVIII.) (cid:3) XXII XXIII XXIV 2.3 Positive Elements 2.3.1 Holomorphic Functions The next order of business is to show that the spectrum sp(a) of an element a of a C∗-algebra contains enough points, so to speak. One incarnation of this idea is that sp(a) is non-empty (see 16 V), but we will need more, and prove that kak = λ for some λ ∈ sp(a) (provided that a is self-adjoint). Somewhat bafflingly, the canonical and apparently easiest way to derive this fact is by considering the power series expansion of a cleverly chosen A -valued function (see 16 II). To this end, we'll first quickly redevelop some complex analysis for A - valued functions (instead of C-valued functions), which will only be needed to prove this fact. 12 Setting Fix a C∗-algebra A for the remainder of this paragraph. For brevity, we'll say that a function is a partially defined map f : C → A whose domain of definition dom(f ) is an open subset of C. Such a function is called holomorphic at a point x ∈ C if f is defined on x, and II f (x) − f (y) x − y converges (with respect to the norm on A ) to some element f′(x) of A as y ∈ dom(f )\{x} converges to x. and the function z 7→ f′(z) with dom(f′) = dom(f ) is called its derivative. Exercise Verify the following examples of holomorphic functions. We say that f is holomorphic if f is holomorphic at x for all x ∈ dom(f ), III 1. If f and g are holomorphic functions with dom(f ) = dom(g), then f + g and f · g are holomorphic, and (f + g)′ = f′ + g′ and (f · g)′ = f′g + g′f . ..11 -- 12.. 31 2. The function f given by f (z) = z and dom(f ) = C is holomorphic, and f′(z) = 1 for all z ∈ C. 3. Let a ∈ A . The constant function f given by f (z) = a for all z ∈ C is holomorphic, and f′(z) = 0 for all z ∈ A . 4. Any polynomial, that is, function f of the form f (z) ≡ anzn+···+a1z +a0 with ai ∈ A is holomorphic with f′(z) = nanzn−1 + ··· + 2a2z + a1. 13 We now turn to perhaps the most important example of a holomorphic A - valued function -- or at the very least the very source from which (as we'll see) all holomorphic functions draw their interesting and pleasant properties: the II Theorem Let a0, a1, a2, . . . ∈ A be given, and write R := (lim supn kank1/n)−1. holomorphic A -valued function given by a power series Pn anzn. Then for every z ∈ C, 1. Pn anzn converges absolutely when z < R, and 2. if Pn anzn converges, then z 6 R. (The number R ∈ [0,∞] is called the radius of convergence of the seriesPn anzn.) Proof Suppose that z < R. To show that the series Pn anzn converges 1/n z )n < ∞. If z = absolutely, we must show that Pn kank zn ≡ Pn(kank 0, this is obvious, so we'll assume that z > 0. Then, since z < R, we have R−1 z < 1 (and R−1 < ∞). Note that there is ε > 0 with (R−1+ε)z < 1. The point of this ε is that lim supn kank1/n < R−1 + ε, so that we can find N with kank1/n 6 R−1 + ε for all n > N . Then kank1/n z 6 (R−1 + ε)z < 1 for all n > N , and soPn kank zn 6PN−1 n=0 kank zn+P∞n=N ( (R−1 +ε)z )n < ∞ by convergence of the geometric series (c.f. 11 II). Suppose now instead thatPn anzn converges. Then kank zn converges to 0. In particular, there is N with kank zn 6 1 for all n > N . Then kank1/n z 6 1, and kank1/n 6 z−1 for all n > N , so that R−1 ≡ lim supn kank1/n 6 z−1, giving z 6 R. IV Proposition The A -valued function f given by a seriesPn anzn with radius of convergence R := ( lim supn kank1/n )−1 is holomorphic when defined on the disk dom(f ) = {z ∈ C : z < R}, and f′(z) =P∞n=1 nanzn−1 for all z ∈ dom(f ). Proof If R = 0, the statement is rather dull, but clearly true, so we assume that R 6= 0, that is, lim supn kank1/n < ∞. Note that the radius of convergence ofP∞n=1 nanzn−1 ≡P∞n=0(n+1)an+1zn is also R, because (cid:3) III V (cid:13)(cid:13) (n + 1) an+1(cid:13)(cid:13) 1/n = (n + 1)1/n kan+1k 1 n+1 (cid:0)kan+1k 1/n, 1 n+1(cid:1) 1 h 1/n con- 1 n+1(cid:1) n+1 , and both (n + 1)1/n and (cid:0)kan+1k and R−1 = lim supn kan+1k verge to 1 as n → ∞ (using here that lim supn kank1/n < ∞). Hence P∞n=1 nanzn−1 converges absolutely for every z ∈ C with z < R. Let z ∈ C with z < R be given. We must show that f is holomorphic at z with f′(z) =Pn nanzn−1. For this it suffices to show that − nzn−1(cid:12)(cid:12)(cid:12)(cid:12) converges to 0 as h ∈ C (with h 6= 0 and z + h < R) tends to 0. Pick r > 0 with z < r < R. With the appropriate algebraic gymnastics (involving the identity an − bn = (a − b)Pn k=1 an−kbk−1 and the inequalities z + h 6 r and z 6 r) we get, for every n and h ∈ C with h 6= 0 and z + h < r, kank(cid:12)(cid:12)(cid:12)(cid:12) (z + h)n − zn ∞ Xn=0 (2.1) (cid:12)(cid:12)(cid:12)(cid:12) (z + h)n − zn h − nzn−1 (cid:12)(cid:12)(cid:12)(cid:12) n = (cid:12)(cid:12)(cid:12)(cid:12) Xk=1(cid:0) (z + h)n−k − zn−k(cid:1)zk−1 (cid:12)(cid:12)(cid:12)(cid:12) 6 2nrn−1. (2.2) (2.3) On the one hand, we see from (2.2) that any term -- and thus any partial sum -- of the series from (2.1) converges to 0 as h tends to 0. On the other hand, so that the tails of the series in (2.1) vanish uniformly in h. All in all, the sum of the infinite series from (2.1) converges to 0 as h tends to 0. (cid:3) we see from (2.3) that the series from (2.1) is dominated by 2Pn kanknrn−1 (which converges because the radius of convergence of Pn annzn−1 is R > r), Exercise LetPn anzn be a power series over A with radius of convergence R > 0 such that Pn anzn = 0 for all z from some disk around 0 with radius r < R. Show that 0 = a0 = a1 = a2 = ··· . vanishes on the disk around 0 with radius r.) (Hint: clearly a0 = 0. Show that the derivative of the power series also VI All holomorphic functions are power series in the sense that any A -valued holo- 14 morphic function f defined on 0 is given by some power series Pn anzn on the largest disk around 0 that fits in dom(f ). This fact, which follows from 15 V and 15 VII below, is all the more remarkable, because here the pointwise ("lo- cal") property of being holomorphic entails the uniform ("global") property of being equal to a power series (on some disk). The device that bridges this gap is integration of A -valued holomorphic functions along line segments. Exercise We're going to define as quickly as possible an integral R f for every continuous map f : [0, 1] → A . Any interval I in [0, 1] is of one of the following forms II [s, t] [s, t) (s, t] (s, t) ..12 -- 14.. 33 where 0 6 s 6 t 6 1; we'll denote the length of an interval I -- being t − s in the four cases above -- by I. An A -valued step function is a function f : [0, 1] → A of the form f ≡Pn an1In for some a1, . . . , aN ∈ A and intervals I1, . . . , IN (where 1In is 1 is the indicator function of In which is 1 on In and 0 elsewhere); and the set of A -valued step functions is denoted by SA , which is a subset of the space of all bounded functions f : [0, 1] → A which we'll denote by BA . 1. Show that there is a unique linear map R : SA → A with R a1I = I a for every interval I in [0, 1] and a ∈ A . (Hint: the difficulty here is to show that no contradiction arises in the sense that Pn an In = Pm a′m I′m when Pn an1In = Pm a′m1I ′ n for intervals I1, . . . , IN , I′1, . . . , I′M in [0, 1] and a1, . . . , aN , a′1, . . . , a′M ∈ A .) 2. We endow BA with the supremum norm, viz. kfk = supt∈[0,1] kf (t)k for all f ∈ BA . Show that every A -valued step function f may be written as f ≡Pn an1In Show that for such a representation kfk = supn kank, and Pn In 6 1. Deduce that kR fk 6Pn kank In 6 kfk. Conclude that R : SA → A is a bounded linear map and can therefore be uniquely extended to a bounded linear map R : SA → A on the clo- 3. Show that every continuous function f : [0, 1] → A is the supremum norm limit of a sequence g1, g2, . . . of A -valued step functions (i.e. f ∈ SA ). 4. Show that R af = aR f when f : [0, 1] → C is continuous and a ∈ A . III Definition The integral of a holomorphic A -valued function f along a line segment [w, w′] ⊆ dom(f ) (where w and w′ are thus complex numbers) is now defined as where I1, . . . , IN are disjoint and non-empty intervals in [0, 1]. sure SA of SA . Z w′ w f = (w′ − w)Z 1 0 f ( w + t(w′ − w) ) dt. We'll also need integration along a triangle T , which is for this purpose a triple of complex numbers w0, w1, w2 (of which the order does matter) called the vertices of T . The boundary of such a triangle T is ∂T := [w0, w1] ∪ [w1, w2] ∪ [w2, w0], and given any A -valued holomorphic function f with ∂T ⊆ dom(f ) we define ZT f = Z w1 w0 f + Z w2 w1 f + Z w0 w2 f. We'll need some more terminology relating to our triangle T . Its closure, writ- ten cl(T ), is the convex hull of w0, w1, w2, and its interior is simply in(T ) = cl(T )\∂T . The length of T is given by length(T ) := w1 − w0 + w2 − w1 + w0 − w2. The number of times the triangle T winds around a point z ∈ C\∂T in the counterclockwise direction is called the winding number, and is written wnT (z), is either 1 or −1 when z ∈ in(T ) (depending on the order of the vertices), is 0 when z /∈ cl(T ), and undefined on ∂T . It is defined formally for z ∈ C\∂T by 2π wnT (z) = ∡(w0, z, w1) + ∡(w1, z, w2) + ∡(w2, z, w0), where ∡(w0, z, w1) denotes the number of radians in (−π, π) needed to rotate the line through z and w0 counterclockwise around z to hit w1, that is, the angle of the corner on the right when travelling from w0 to w1 via z. later on, see VIII. Goursat's Theorem Let f be a holomorphic function, and let T be a triangle The winding number wnT pops up in the value of the integralRT (z−z0)−1dz whose closure is entirely contained in dom(f ). Then RT f = 0. Proof (Based on [54].) If two vertices of T coincide the result is obviously true, so we may assume that they're all distinct, that is, in(T ) 6= ∅. Note that if f has an antiderivative, that is, f ≡ g′ for some holomorphic function g, then one can show that RT f = 0 (after deriving the fundamental theorem of calculus). Although it is true that every holomorphic function with simply connected domain has a antiderivative, this result is not yet available (and in fact usually depends on this very theorem). Instead we will approx- imate f by an affine function (which does have an antiderivative) using the derivative of f . But since such an approximation only concerns a single point, we first need to zoom in. If we split T into four similar triangles T i, T ii, T iii, T iv (cid:9) (cid:8) (cid:9) (cid:9) From this it is clear how to get a sequence of similar triangles T0, T1, T2, . . . n=iRT n f . There is T ′ among T i, T ii, T iii, T iv with kRT fk 6 we haveRT f =Piv 4kRT ′ fk. Clearly, length(T ) = 2 length(T ′). Write T0 := T and T1 := T ′. with kRT fk 6 4nkRTn sequence that converges to some point z0 ∈ C which lies in Tn cl(Tn). We can approximate f by an affine function at z0 as follows. For z ∈ dom(f ), If we pick a point on the closure cl(Tn) of each triangle Tn we get a Cauchy fk, and length(T ) = 2n length(Tn). f (z) = f (z0) + f′(z0) (z − z0) − r(z) (z − z0), ..14.. 35 IV V VI VII where r : dom(f ) → C is given by r(z) = f′(z0) − (f (z) − f (z0))(z − z0)−1 for z 6= z0 and r(z0) = 0. We see that r(z) converges to 0 as z → z0. Let ε > 0 be given. There is δ > 0 such that z ∈ dom(f ) and kr(z)k 6 ε for all z ∈ C with kz − z0k < δ. There is n such that the triangle Tn is contained f (z0) + f′(z0)(z − z0) dz = 0 by in the ball around z0 of radius δ. Note that RTn the discussion in V, because the integrated function is affine. Thus RTn f = −RTn r(z) (z − z0) dz. Note that for z ∈ Tn, we have kz − z0k 6 length(Tn), and kr(z)k 6 ε (because kz − z0k < δ), and so kr(z)(z − z0)k 6 ε length(Tn). Thus: kRTn fk = kRTn Using the inequalities from VI, we get r(z) (z − z0) dzk 6 ε length(Tn)2. kRT fk 6 4n kRTn fk 6 ε 4n length(Tn)2 ≡ ε length(T )2. Since ε > 0 was arbitrary, we see that RT f = 0. VIII Exercise The assumption in Goursat's Theorem (IV) that the holomorphic function f is defined not only on the boundary ∂T of the triangle T but also on the interior in(T ) is essential, for if only a single hole in dom(f ) is allowed (cid:3) within in(T ) the integralRT f can become non-zero -- which we will demonstrate here by computing RT (z − z0)−1dz. 1. Show that for a non-zero complex number z we have z−1 = zR − izI z2 R + z2 I . 2. Given real numbers a 6= 0 and b, show that Z a+ib a z−1 dz = iZ t = iZ b = i arctan( b/a ) + log a + ib − log ia , a − it a2 + t2 dt a2 + t2 dt + Z b a2 + t2 dt a t 0 0 0 and similarly, show that for real numbers a and b 6= 0 Z ib a+ib z−1 dz = i arctan( a/b ) + log ib − log a + ib . 3. Show that for complex numbers w, w′ and z0 with z0 /∈ [w, w′] (z − z0)−1 dz = i ∡(w, z0, w′) + log w′ − z0 w − z0 Z w′ w , where ∡(w, z0, w′) denotes the number of radians in (−π, π) needed to rotate the line through z0 and w counterclockwise around z0 to hit w′. (Hint: using Goursat's Theorem, IV, one may reduce the problem to inte- gration along horizontal and vertical line segments.) 4. Given a triangle T and z0 ∈ C\∂T , show that 1 2πiZT (z − z0)−1 dz = wnT (z0). Thus integration of z 7→ (z − z0)−1 along a triangle T detects the number of times T winds around z0. There is nothing special about a triangle: a similar result -- not needed here -- holds for a broad class of curves (c.f. Thm 2.9 of [15]). Integration along a curve can also be used to probe the value of a holomorphic function at a point z0. On this occasion we restrict ourselves to regular N -gons. Theorem (Cauchy's Integral Formula) Let f be a holomorphic A -valued func- tion which is defined on the interior and boundary of some regular N -gon with centre c ∈ C, circumradius r and vertices wn := c + r cos(2π/n) + ir sin(2π/n). Then for any complex number z0 in the interior of the N -gon we have IX 15 f (z0) = 1 2πi N−1 Xn=0Z wn+1 wn f (z) z − z0 dz Proof Since PN−1 n=0 R wn+1 wn f (z0) z−z0 dz = 2πif (z0) by 14 VIII it suffices to show that II N−1 Xn=0Z wn+1 wn f (z) − f (z0) z − z0 dz = 0. (2.4) Let ε > 0 be given. Since f is holomorphic at z0 we can find δ > 0 with III (cid:13)(cid:13)(cid:13)(cid:13) f (z) − f (z0) z − z0 (cid:13)(cid:13)(cid:13)(cid:13) 6 kf′(z0)k + 37 for all z ∈ dom(f ) with z − z0 6 δ. To use III, we must restrict our attention to a smaller polygon. Let T be a triangle that is entirely inside the N -gon such that wnT (z0) = −1, length(T ) 6 IV ..14 -- 15.. 37 ε, and kz0 − zk 6 δ for all z ∈ ∂T . By partitioning the area between T and the N -gon in the obvious manner into triangles T1, . . . , TM (for whichRTm f = 0 for all m by 14 IV) we see that N−1 Xn=0Z wn+1 wn f (z) − f (z0) z − z0 dz = ZT f (z) − f (z0) z − z0 dz. (2.5) Hence by III we have N−1 Xn=0Z wn+1 wn (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) f (z) − f (z0) z − z0 dz(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 6 length(T ) · sup 6 kf′(z0)kε + 37ε. z∈∂T (cid:13)(cid:13)(cid:13)(cid:13) f (z) − f (z0) z − z0 (cid:13)(cid:13)(cid:13)(cid:13) Since ε > 0 was arbitrary, (2.4) follows from (2.5). (cid:3) V Proposition Let f be a holomorphic A -valued function defined on the boundary and interior of a regular K-gon with vertices w0, . . . , wK−1, wK = w0 as in I. Then for every element z of an open disk in the interior of the K-gon with centre w, f (z) = ∞ Xn=0 1 2πi K−1 Xk=0 Z wk+1 wk f (u) (u − w)n+1 du! (z − w)n. VI Proof By I and some easy algebra we have 2πif (z) = K−1 Xk=0 Z wk+1 wk f (u) u − z du = K−1 Xk=0 Z wk+1 wk f (u) u − w 1 1 − z−w u−w du Note that z − w < u − w for all u ∈ [wk, wk+1] and k, because the open disk with centre w from which z came lies entirely in the K-gon. Hence, by 11 II, 2πif (z) = = wk K−1 f (u) u − w Xk=0 Z wk+1 Xn=0 Xk=0 Z wk+1 K−1 ∞ wk ∞ (z − w)n (u − w)n du Xn=0 (u − w)n+1 du (z − w)n, f (u) VII Proposition Let f be an A -valued holomorphic function that can be written (u−w)n+1 converges uniformly in u as N → ∞. where the interchange of "P" and "R " is allowed because the partial sum n=0 f (u) (z−w)n PN as a power series f (z) = Pn an(z − w)n where a0, a1, . . . ∈ A for all z from some disk in dom(f ) around w with radius r > 0. (cid:3) Then the formula f (z) = Pn an(z − w)n holds also for any z from a larger disk with radius R > r around w that still fits in dom(f ). Proof Let z with z − w < R be given. By choosing K large enough we can fit the boundary of a regular K-gon centred around w with vertices w0, . . . , wK−1, wK ≡ w0 inside the difference between the two disks, and we can moreover, by V, choose the polygon in such a way that f (z′) = Pn bn(z′ − w)n for all z′ ∈ C k=0 R wk+1 with z′ − w 6 z − w where bn =PK−1 Thus to show that f (z) = Pn an(z − w)n it suffices to show that an = bn for all n. This in turn follows by 13 VI from the fact that Pn an(z′ − w)n = Pn bn(z′ − w)n for all z′ ∈ C with z′ − w < r. (u−w)n+1 du. f (u) (cid:3) VIII wk 2.3.2 Spectral Radius Our analysis of A -valued holomorphic functions allows us to expose the follow- ing connection between the norm and the invertible elements in a C∗-algebra. Proposition For a self-adjoint element a of a C∗-algebra A , we have kak = sup{ λ : λ ∈ sp(a)}. (The quantity on the right hand-side above is called the spectral radius of a.) Proof Write r = sup{ λ : λ ∈ sp(a)\{0} } where the supremum is computed in [0,∞] so that sup ∅ = 0. Since λ 6 kak for all λ ∈ sp(a) (11 VI) we see that r 6 kak, and so we only need to show that kak 6 r. Note that this is clearly true if kak = 0, so we may assume that kak 6= 0. The trick is to consider the power series expansion around 0 of the holo- morphic function f defined on G := { z ∈ C : 1 − az is invertible} by f (z) = z(1 − az)−1. More specifically, we are interested in the distance R of 0 to the complement of G, viz. R = inf{λ : λ ∈ C\G} (where the infimum is computed in [0,∞] so that inf ∅ = ∞) because since 0 ∈ G and z /∈ G ⇐⇒ z−1 ∈ sp(a), we have R = r−1 (using the convention 0−1 = ∞). Note that f has the power series expansion f (z) =Pn anzn+1 for all z ∈ C with kzk < kak−1, because for such z we have Pn(az)n = (1 − az)−1 by 11 II, and thus f (z) = z(1 − az)−1 = zPn(az)n =Pn anzn+1. By 15 VII we know that f (z) = Pn anzn+1 is valid not only for z ∈ C with z < kak−1, but for all z with z < R. However, R cannot be strictly larger than kak−1, because for every z ∈ C with z > kak−1 the series Pn(az)n and thus Pn anzn+1 diverges (see 11 VII) -- using here that a is self-adjoint. Hence R = kak−1, and so r = kak. Remark For an arbitrary (possibly non-self-adjoint) element a of a C∗-algebra A the formula in II might be incorrect, e.g. (cid:13)(cid:13)(cid:0) 0 1 0 0(cid:1) ) = {0} 0 0(cid:1)k = 1 while sp((cid:0) 0 1 ..15 -- 16.. (cid:3) 39 16 II III IV cf. 11 IX. For such a the formula sup{ λ : λ ∈ sp(a)} = lim supn kank1/n can be derived (see e.g. Theorem 3.3.3 of [47]) -- which we won't need here. Exercise Given a self-adjoint element a of a C∗-algebra show that sp(a) 6= ∅. VI Exercise Given a self-adjoint element a of a C∗-algebra and λ ∈ R show that V sp(a) = {λ} iff a = λ. VIa Exercise Use the previous exercise to prove the following theorem. VII Theorem (Gelfand -- Mazur for C∗-algebras) If every non-zero element of a C∗-algebra A is invertible, then A = C or A = {0}. VIII Remark A logical next step towards Gelfand's representation theorem is to show that if λ ∈ sp(a) for some element a of a commutative C∗-algebra A , then there is an miu-map f : A → C with f (a) = λ. Here we have moved ourselves into a tight spot by evading Banach algebras, because the mentioned result is usually obtained by finding a maximal ideal I of A (by Zorn's Lemma) that contains λ − a, and then forming the Banach algebra quotient A /I. One then applies Gelfand -- Mazur's Theorem for Banach algebras, to see that A /I = C, and thereby obtain an miu-map f : A → C with f (a − λ) = 0. The problem here is that while A /I will turn out to be a C∗-algebra (indeed, be C) the formation of the C∗-algebra quotient is non-trivial and depends on Gelfand's representation theorem (see e.g. §VIII.4 of [15]) which is the very theorem we are working towards. The way out of this predicament is to avoid ideals and quotients of C∗- and Banach algebras altogether, and instead work with order ideals (and what are essentially quotients of Riesz and order unit spaces). To this end, we develop the theory of the positive elements of a C∗-algebra farther than is usually done for Gelfand's representation theorem. 17 We return to the positive elements in a C∗-algebra (see 9 IV). We'll see that the connection we have established between the norm and invertible elements of a C∗-algebra via the spectral radius (16 II) affects the positive elements as well, see V. Exercise Show that λ − t 6 t iff λ ∈ [0, 2t], where λ, t ∈ R. II III Proposition For a self-adjoint element a from a C∗-algebra, and t ∈ [0,∞], ka − tk 6 t ⇐⇒ sp(a) ⊆ [0, 2t]. IV Proof To begin, note that sp(a − t) = sp(a) − t ⊆ R by 11 XXI, because a is self-adjoint. Thus ka− tk = sup{ λ − t : λ ∈ sp(a)} by 16 II. Hence ka− tk 6 t iff λ − t 6 t for all λ ∈ sp(a) iff sp(a) ⊆ [0, 2t] (by II). (cid:3) Exercise Show (using III and 11 XXI) that for any self-adjoint element a of a C∗-algebra A , the following are equivalent. V 2kak; 1. ka − tk 6 t for some t > 1 2. ka − tk 6 t for all t > 1 2kak; 3. sp(a) ⊆ [0,∞); 4. a is positive. We will complete this list in 25 I. Exercise Let A be a C∗-algebra. 1. Show that 0 6 a 6 0 entails that a = 0 for all a ∈ A . 2. Show that A+ is closed. 3. Let a be a self-adjoint element of A . Show that −λ 6 a 6 λ iff kak 6 λ, for λ ∈ [0,∞). Conclude that kak = inf{λ ∈ R : − λ 6 a 6 λ}. (In other words AR is a complete Archimedean order unit space, see Defi- nition 1.12 of [2] -- a type of structure first studied in [45].) Show that 0 6 a 6 b entails kak 6 kbk for a, b ∈ AR. 4. Recall that ab need not be positive if a, b > 0. However: Show that a2 is positive for every self-adjoint element a of A . Show that an is positive for even n ∈ N and a ∈ AR. Show that an is positive iff a is positive for odd n ∈ N and a ∈ AR. Show that an is positive for every positive a from A and n ∈ N. 5. Let a be an invertible element of A . Show that a > 0 iff a−1 > 0. 6. Show that a positive element a of A is invertible iff a > 1 n for some n > 0. (Hint: show that sp(a) ⊆ [ 1 n ,∞) when a > 1 n .) [Moved to 20a I.] Although we can't quite yet see that a∗a is positive -- for this we need the ex- istence of the square root, 23 VII, -- we can already prove that a∗a can't be negative, see III. Lemma For elements a and b from a C∗-algebra, we have sp(ab)\{0} = sp(ba)\{0}. Proof Let λ ∈ C with λ 6= 0 be given. We must show that λ− ab is invertible iff λ − ba is invertible. Suppose that λ − ab is invertible. Then using the equality a(λ − ba) = (λ − ab)a one sees that (1 + b(λ − ab)−1a)(λ − ba) = λ. Since similarly (λ − ba)(1 + b(λ − ab)−1a) = λ, we see that λ−1(1 + b(λ − ab)a) is the inverse of λ − ba. (cid:3) ..16 -- 19.. 41 VI 18 19 Ia II 20 II III III But on the other hand, a∗a + aa∗ = 2(a2 IV Proof Suppose that a∗a 6 0. Then sp(a∗a) ⊆ (−∞, 0], almost by definition, I ) > 0, and so a∗a + aa∗ = 0. (cid:3) Lemma We have a∗a 6 0 =⇒ a = 0 for every element a of a C∗-algebra. and so sp(aa∗) ⊆ (−∞, 0] by Ia, giving aa∗ 6 0. Thus a∗a + aa∗ 6 0. R + a2 Then 0 > a∗a = −aa∗ > 0 gives a∗a = 0, and a = 0. Observe that the norm and order on (the self-adjoint elements of a) C∗-algebra A completely determine one another (using the unit): on the one hand kak = inf{λ > 0 : − λ 6 a 6 λ} by 17 VI, and on the other hand a > 0 iff ka − sk 6 s for some s ∈ R by definition (9 IV). This has some useful consequences. Lemma A positive map f : A → B between C∗-algebras is bounded. More specifically, we have kf (a)k 6 kf (1)k kak for all self-adjoint a ∈ AR, and we have kf (a)k 6 2kf (1)k kak for arbitrary a ∈ A . Proof Given a ∈ AR we have −kak 6 a 6 kak, and −kak f (1) 6 f (a) 6 kak f (1) (because f is positive), and thus kf (a)k 6 f (1)kak 6 kf (1)k kak by 17 VI. For an arbitrary element a ≡ aR + iaI of A we have kf (a)k 6 kf (aR)k + kf (aI)k 6 2kf (1)k kak. (cid:3) IV Remark It is a non-trivial theorem (see 34a VII) that the factor "2" in the statement above can be dropped, i.e. kfk = kf (1)k (c.f. Corollary 1 of [67]). We'll be using this improved bound mostly for completely positive maps, for which it's much easier to obtain (see 34 XVI). For miu-maps we can already obtain the improved bound here: V Lemma Any miu-map : A → B between C∗-algebras A and B is positive, bounded, and, in fact, kk 6 1. Va Proof Let a be a positive element of A , so sp(a) ⊆ [0,∞) by 17 V, To show that is positive, we must prove that (a) > 0, that is, sp((a)) ⊆ [0,∞). This follows immediately from the observation that sp((a)) ⊆ sp(a): when a − λ is invertible, so is ( a − λ ) ≡ (a) − λ, for any λ ∈ C. Hence is positive. It follows by II that is bounded, and k(b)k 6 kbk for self-adjoint b ∈ A . It remains to be shown that k(a)k 6 kak for arbitrary a ∈ A . But since a∗a is self-adjoint for such a, we have k(a)k2 ≡ k(a∗a)k 6 ka∗ak = kak2 by the C∗- identity and using that is an miu-map. Whence k(a)k 6 kak for all a ∈ A , and so kk 6 1. (cid:3) Lemma For a pu-map f : A → B the following are equivalent. VI 1. f is bipositive, that is, f (a) > 0 iff a > 0 for all a ∈ A ; 2. f is an isometry on AR, that is, kf (a)k = kak for all ∈ AR; 3. f is an isometry on A+. VII Proof It is clear that 2 implies 3. 1 0 1 2 (1=⇒2) Let a ∈ AR be given. Note that −λ 6 a 6 λ iff −λ 6 f (a) 6 λ for all λ > 0, because f is bipositive and unital. In particular, since −kak 6 a 6 kak, we have −kak 6 f (a) 6 kak, and so kf (a)k 6 kak. On the other hand, −kf (a)k 6 f (a) 6 kf (a)k implies −kf (a)k 6 a 6 kf (a)k, and so kak 6 kf (a)k. Thus kak = kf (a)k, and f is an isometry on AR. (3=⇒1) Let a ∈ A be given. We must show that f (a) > 0 iff a > 0. Since f is involution preserving (10 IV) a is self-adjoint iff f (a) is self-adjoint, and so we might as well assume that a is self-adjoint to start with. Since f is an isometry on A+, kak − a is positive, and f is unital, we have k kak − ak = kf (kak − a)k = k kak − f (a)k. Now, observe that 0 6 a iff k kak − ak 6 kak, and that k kak − f (a)k 6 kak iff 0 6 f (a), by 17 VI, because 1 2kak 6 kak and 2kf (a)k 6 kak (by II). (cid:3) Warning Such a map f need not preserve the norm of arbitrary elements: the map A 7→ 1 2 A + 1 1 2 AT : M2 → M2 is bipositive and unital, but (cid:13)(cid:13)(cid:13)(cid:13) 0 (cid:19) + (cid:18) 0 (cid:18)0 0 (cid:18)0 1/2 0(cid:19)(cid:13)(cid:13)(cid:13)(cid:13) = 1 6= (Hint: use here that the projections πj are bounded by 20 V.) Show that the same description applies to cC∗pu and cC∗pu. (Hint: first show (Even if f is completely positive, 34 IV, it might still only preserve the norm of self-adjoint elements cf. 21 IX.) Ai of a family (Ai)j∈I of C∗-algebras de- fined in 3 V is also the categorical product of these C∗-algebras in C∗miu and cC∗miu Ai → Aj given by πj (a) = a(j). = (cid:13)(cid:13)(cid:13)(cid:13) Exercise Show that the productLi∈I with as projections the maps πj : Li∈I that an element a of Li∈I Exercise Show that given miu-maps f, g : A → B between C∗-algebras the collection E := {a ∈ A : f (a) = g(a)} is a C∗-subalgebra of A (using the fact that f and g are bounded by 20 V to show that E is closed.) Show that the inclusion e : E → A is a (positive) miu-map that is in fact the equaliser of f and g in C∗miu and C∗pu. Show that the same description applies to cC∗miu and cC∗pu. Remark The assumption here that f and g are miu-maps is essential: the pair of pu-maps f, g : C4 → C given by 2 (a + b), Ai is positive iff a(i) is positive for every i ∈ I.) We'll return to the product of C∗-algebras a final time in 34 VI. . 0(cid:19)(cid:13)(cid:13)(cid:13)(cid:13) and g(a, b, c, d) = 1 2 (c + d), f (a, b, c, d) = 1 0 1/2 VIII IX X 20a II III for example, has no equaliser in C∗pu, as we'll show in 84a I. We just saw in 20 VI that a map on a C∗-algebra A that preserves and reflects the order determines the norm of the self-adjoint -- but not all -- elements of A . 21 ..19 -- 21.. 43 This theme, to what extent a linear map (or a collection of linear maps) on a C∗-algebra determines its structure, while tangential at the moment, will grow ever more important until it is essential for the theory of von Neumann algebras. That's why we introduce the four levels of discernment that a collection of maps on a C∗-algebra might have already here. II Definition A collection Ω of linear maps on a C∗-algebra A will be called 1. order separating if an element a of A is positive iff 0 6 ω(a) for all ω ∈ Ω; 2. separating if an element a of A is zero iff ω(a) = 0 for all ω ∈ Ω; 3. faithful if an element a of A+ is zero iff ω(a) = 0 for all ω ∈ Ω; and 4. centre separating if a ∈ A+ is zero iff ω(b∗ab) = 0 for all ω ∈ Ω and b ∈ A . (The "centre" in "centre separating" will be explained in 69 IX.) (Note that (1) =⇒ (2) =⇒ (3) =⇒ (4).) Examples We'll see later on that the following collections are order separating. III 1. The set of all pu-maps ω : A → C (called states) on a C∗-algebra (see 22 VIII). 2. The set of all miu-maps ω : A → C on a commutative C∗-algebra (see 27 XVIII). 3. The set of functionals on B(H ), where H is a Hilbert space, of the form hx, (· )xi : B(H ) → C where x ∈ H (see 25 III). We'll call these functionals vector functionals. (They are clearly bounded and involution preserving linear maps, and once we know that each pos- itive element of a C∗-algebra is a square, in 23 VII, it'll be obvious that vector functionals are positive too.) The unital vector functionals (called vector states) are order separating too. IV None of the four levels of separation coincide. This follows from the following examples, that we'll just mention here, but can't verify yet. 1. A single non-zero vector x from a Hilbert space H gives a vector functional hx, (· )xi on B(H ) that is centre separating on its own, but is not faithful when H has dimension > 2. 2. Given an orthonormal basis E of a Hilbert space H the collection { he, (· )ei : e ∈ E } of vector functionals on B(H ) is faithful, but not separating when E has more than one element. 3. Given Hilbert spaces H and K the set of vector functionals { h x ⊗ y, (· ) x ⊗ y i : x ∈ H , y ∈ K } on B(H ⊗ K ) is separating, but not order separating when both H and K are at least two dimensional. Exercise One use for a separating collection Ω of involution preserving maps on a C∗-algebra A is checking whether an element a ∈ A is self-adjoint: show that a ∈ A is self-adjoint iff ω(a) is self-adjoint for all ω ∈ Ω. An order separating collection senses the norm of a self-adjoint element: Proposition For a collection Ω of pu-maps on a C∗-algebra A the following are equivalent. 1. Ω is order separating; 2. kak = supω∈Ω kω(a)k for all a ∈ AR; 3. kak = supω∈Ω kω(a)k for all a ∈ A+. Bω (see 20a I). Proof Denoting the codomain of ω ∈ Ω by Bω (so that ω : A → Bω), ap- ply 20 VI to the pu-map hωiω∈Ω : A →Lω∈Ω (cid:3) Warning The formula kak = supω∈Ω kω(a)k need not be correct for an arbitrary (not necessarily self-adjoint) element a. Indeed, consider the matrix A :=(cid:0) 0 1 0 0(cid:1), and the collection Ω = { hx, (· )xi : x ∈ C2, kxk = 1 }, which will turn out to be order separating. We have kAk = 1, while hx, ω(A)xi = x1x2 never exceeds 1/2 for x ≡ (x1, x2) ∈ H with 1 = kxk. Exercise Show that any operator norm dense subset Ω′ of an order separating collection Ω of positive functionals on a C∗-algebra A is order separating too. We'll use 21 VII to show that the pu-maps ω : A → C on a C∗-algebra A (called states of A for short) are order separating by showing that for every self-adjoint element a ∈ A there is a state ω of A with ω(a) = kak or ω(a) = −kak. To obtain such a state we first find its kernel, which leads us to the following definitions. Definition An order ideal of a C∗-algebra A is a linear subspace I of A with b ∈ I =⇒ b∗ ∈ I and b ∈ I ∩ A+ =⇒ [−b, b] ≡ { a ∈ A : − b 6 a 6 b } ⊆ I. The order ideal I is called proper if 1 /∈ I, and maximal if it is maximal among all proper order ideals. Warning "Order ideals" like "subspaces" appear in relation to other structures as well, with appropriately varying meanings. Our definition for C∗-algebras is based on to the order ideals for order unit spaces from Definition 2.2 of [45]. ..21 -- 22.. 45 V VI VII VIII IX X 22 II IIa III Exercise Let A be a C∗-algebra. 1. Show that the kernel of a state is a maximal order ideal. (Hint: the kernel of a state is already maximal as linear subspace.) 2. Let I be a proper order ideal of A . Show that there is a maximal order ideal J of A with I ⊆ J. (Hint: Zorn's Lemma may be useful.) 3. Let a ∈ AR. Show that there is a least order ideal (a) that contains a, and that given b ∈ AR we have b ∈ (a) iff there are λ, µ ∈ R with λa 6 b 6 µa. Show that (a) = Ca when 0 66 a 66 0. Show that 1 ∈ (a) if and only if a is invertible and either 0 6 a or a 6 0. 4. Let a be a self-adjoint element of A which is not invertible. Show that there is a maximal order ideal J of A with a ∈ J. 5. Let a be a self-adjoint element of A . Show that kak − a or kak + a is not invertible (perhaps by considering the spectrum of a.) IV V VI Lemma For every maximal order ideal I of a C∗-algebra A , there is a state ω : A → C with ker(ω) = I. Proof Form the quotient vector space A /I with quotient map q : A → A /I. Note that since 1 /∈ I we have q(1) 6= 0 and so we may regard C to be a linear subspace of A /I via λ 7→ q(λ). We will, in fact, show that C = A /I. But let us first put an order on A /I: we say that a ∈ A /I is positive if a ≡ q(a) for some a ∈ A+, and write a 6 b if b − a is positive for a, b ∈ A /I. Note that the definition of "order ideal" is such that if both a and −a are positive, then a = 0. We leave it to the reader to verify that A /I becomes a partially ordered vector space with the order defined above. There is, however, one detail we'd like to draw attention to, namely that a scalar λ is positive in A /I iff λ is positive in C. Indeed, if λ > 0 in C, then λ > 0 in A , and so λ > 0 in A /I. On the other hand, if λ > 0 in A /I, but λ 6 0 in C, then λ 6 0 in A /I, and so λ = 0. This detail has the pleasant consequence that once we have shown that A /I = C, we automatically get that q : A → C is positive. Let a ∈ AR be given. Define α := inf{ λ ∈ R : q(a) 6 λ}. Note that −kak 6 α 6 kak. We will prove that q(a) = α by considering the order ideal J := { b ∈ A : ∃λ, µ ∈ R [ λ(α − q(a)) 6 bR 6 µ(α − q(a)) ]∧ ∃λ, µ ∈ R [ λ(α − q(a)) 6 bI 6 µ(α − q(a)) ]}. We claim that 1 /∈ J. Indeed, suppose not -- towards a contradiction. Then there is µ ∈ R with 1 6 µ(α − q(a)). What can we say about µ? If µ < 0, then 0 > 1/µ > α − q(a), so α − 1/µ 6 q(a), but q(a) 6 α + ε for every ε > 0, If µ = 0, then we get and so α − 1/µ 6 q(a) 6 α − 1/2µ, which is absurd. 1 6 µ(α − q(a)) ≡ 0, which is absurd. If µ > 0, then 1/µ 6 α − q(a), or in other words, q(a) 6 α − 1/µ, giving α 6 α − 1/µ by definition of α, which is absurd. Hence 1 /∈ J. But then since I ⊆ J, we get I = J, by maximality of I. Thus, as α− a ∈ J, we have α − a ∈ I, and so q(a) = α, as desired. Let a ∈ A be given. Then a = aR + iaI. By VI, there are α, β ∈ R with q(aR) = α, and q(aI) = β. Thus q(a) = α + iβ. Hence A /I = C. Since the quotient map q : A → A /I ≡ C is pu, and ker(q) = I, we are done. (cid:3) Exercise Show using IV that given a self-adjoint element a of a C∗-algebra A there is a state ω with ω(a) = kak. Conclude that the set of states of a C∗-algebra is order separating (see 21 II). 2.3.3 The Square Root The key that unlocks the remaining basic facts about the (positive) elements of a C∗-algebra is the existence of the square root √a of a positive element a, and its properties. For technical reasons, we will assume kak 6 1, and construct 1 − √1 − a instead of √a. Lemma Let a be an element of a C∗-algebra A with 0 6 a 6 1. Then there is a unique element b ∈ A with, 0 6 b 6 1, ab = ba, and (1 − b)2 = 1 − a. To be more specific, b is the norm limit of the sequence b0 6 b1 6 ··· given by b0 = 0 and bn+1 = 1 n). Moreover, if c ∈ A commutes with a, then c commutes with b, and if in addition c2 6 1 − a and c∗ = c, we have c 6 1 − b. Proof When discussing bn it is convenient to write bn ≡ qn(a) where q0, q1, . . . are the polynomials over R given by q0 = 0 and qn+1 = 1 n). For example, we have bn > 0, because all coefficients of qn are all positive, and a, a2, a3, . . . are positive by 17 VI. With a similar argument we can see that b0 6 b1 6 b2 6 ··· . Indeed, the coefficients of qn+1 − qn are positive, by induction, because 2 (a + b2 2 (x + q2 VII VIII 23 II III n+1) − 1 2 (x + q2 n) qn+2 − qn+1 = 1 = 1 = 1 = (qn + 1 2 (x + q2 2 (q2 2 (qn+1 + qn)(qn+1 − qn) n+1 − q2 n) 2 (qn+1 − qn))(qn+1 − qn), has positive coefficients if qn+1 − qn has positive coefficients, and q1 − q0 ≡ 1 2 x clearly has positive coefficients. Hence bn+1 − bn = qn+1(a) − qn(a) is positive. (Note that we have carefully avoided using the fact here that the product of positive commuting elements is positive, which is not available to us until V.) ..22 -- 23.. 47 2 (1 + qn(1)2) 6 1 if qn(1) 6 1, and clearly 0 ≡ q0(1) 6 1. Let us now show that b0 6 b1 6 ··· converges. Let n > N from N be given. Since the coefficients of qn − qN are positive, and kak 6 1, the triangle inequality gives us kbn − bNk ≡ k(qn − qN )(a)k 6 qn(1) − qN (1), and so it suffices to show that the ascending sequence q0(1) 6 q1(1) 6 ··· of real numbers converges, i.e. is bounded. Indeed, we have qn(1) 6 1, by induction, because qn+1(1) ≡ 1 Let b be the limit of b0 6 b1 6 ··· . Then b being the limit of positive elements is positive (see 17 VI), and if c ∈ A commutes with a, then c commutes with all powers of a, and therefore with all bn, and thus with b. Further, from the recurrence relation qn+1 = 1 2 (a+ b2), and so −a = −2b + b2, giving us (1 − b)2 = 1 − 2b + b2 = 1 − a. Let us prove that b 6 1. To begin, note that kbnk 6 1 for all n, by induction, because 0 ≡ kb0k 6 1, and if kbnk 6 1, then kbn+1k 6 1 2 (kak +kbnk2) 6 1, since kak 6 1. Since bn > 0, we get −1 6 bn 6 1 for all n, and so b 6 1. Let us take a step back for the moment. From what we have proven so far we see that each positive c ∈ A is of the form c ≡ d2 for some positive d ∈ A which commutes with all e ∈ A that commute with c. From this we can see that c1c2 > 0 for c1, c2 ∈ A+ with c1c2 = c2c1. Indeed, writing ci ≡ d2 i with di as above, we have d1c2 = c2d1 (because c1c2 = c2c1), and thus d1d2 = d2d1. It follows that d1d2 is self-adjoint, and c1c2 = (d1d2)2. Hence c1c2 > 0. n) we get b = 1 2 (a+ q2 IV V n) 6 1 2 (a + b2 We will also need the following corollary. For c, d ∈ A+ with c 6 d and cd = dc, we have c2 6 d2. Indeed, d2 − c2 ≡ d(d − c) + c(d − c) is positive by the previous paragraph. Let c ∈ AR be such that ca = ac and c2 6 1 − a, that is, a 6 1 − c2. We must show that c 6 1 − b, that is, b 6 1 − c. Of course, since b is the limit of b1, b2, . . . , it suffices to show that bn 6 1 − c, and we'll do this by induction. Since 0 6 c2 6 1 − a, we have kck2 6 k1 − ak 6 1, and so −1 6 c 6 1. Thus b0 ≡ 0 6 1 − c. Now, suppose that bn 6 1 − c for some n. Then bn+1 = 1 2 ((1 − c2) + (1 − c)2) = 1 − c, where we have used that b2 n 6 (1 − c)2, because bn 6 1 − c by IV. VI We'll now show that b is unique in the sense that b = b′ for any b′ ∈ A with 0 6 b′ 6 1, b′a = ab′ and (1− b′)2 = 1− a. Note that b′ 6 1, because k1− b′k2 = k1 − ak 6 1, From a = 1 − (1 − b′)2, we immediately get b 6 1 − (1 − b′) = b′ by V. For the other direction, note that (1−b′)2 = (1−b)2 ≡ (1−b′ +(b′−b))2 = (1− b′)2 + 2(1− b′)(b′ − b) + (b′− b)2, which gives 0 = 2(1− b′)(b′ − b) + (b′− b)2. Now, since 1− b′ and b′− b are positive, and commute, we see that (1− b′)(b′− b) is positive by V, and so 0 = 2(1 − b′)(b′ − b) + (b′ − b)2 > (b′ − b)2 > 0, which entails (b′ − b)2 = 0, and so k(b′ − b)2k = kb′ − bk2 = 0, yielding b = b′. (cid:3) VII Exercise Let a be a positive element of a C∗-algebra A . Show that there is a unique positive element of A denoted by √a (and by a1/2) with √a2 = a and a√a = √aa. Show that if c ∈ A commutes with a, then c√a = √ac, and if in addition c∗ = c and c2 6 a, then c 6 √a. Using this, verify: 1. If a, b ∈ A are positive, and ab = ba, then ab > 0. 2. Let a ∈ A+. If b, c ∈ AR commute with a, then b 6 c implies ab 6 ac. 3. If a, b ∈ AR commute, and a 6 b, then a2 6 b2. 4. The requirement in the previous item that a and b commute is essential: there are positive elements a, b of a C∗-algebra A with a 6 b, but a2 66 b2. In other words, the square a 7→ a2 on the positive elements of a C∗-algebra need not be monotone, (but a 7→ √a is monotone, see 28 III). (Hint: take a = ( 1 0 0 0 ) and b = a + 1 2 ( 1 1 1 1 ) from M2.) Definition Given a self-adjoint element a of a C∗-algebra A , we write 24 := √a2 a a+ := 1 2 (a + a) a− := 1 2 (a − a). We call a+ the positive part of a, and a− the negative part. Exercise Let a be a self-adjoint element of a C∗-algebra A . 1. Show that − a 6 a 6 a, and k a k = kak. 2. Prove that a+ and a− are positive, a = a+ − a− and a+a− = a−a+ = 0. 3. One should not read too much into the notation · in the non-commutative case: give an example of self-adjoint elements a and b of a C∗-algebra with a + b 66 a + b. (Hint: one may take a = 1 2 ( 1 1 1 1 ) and b = − ( 1 0 0 0 ).) The existence of positive and negative parts in a C∗-algebra has many pleasant and subtle consequences of which we'll now show one. Lemma Given an element a of a C∗-algebra A , we have a∗a > 0. Proof Writing b := a((a∗a)−)1/2, we have b∗b = ((a∗a)−)1/2a∗a((a∗a)−)1/2 = (a∗a)− a∗a = −((a∗a)−)2 6 0, and so b = 0 by 19 III. Hence ((a∗a)−)2 = 0, and thus (a∗a)− = 0 (by, say, the C∗-identity,) giving us a∗a = (a∗a)+ > 0. (cid:3) Exercise Round up our results regarding positive elements to prove that the following are equivalent for a self-adjoint element a of a C∗-algebra A . 1. a is positive, that is, ka − tk 6 t for some t ∈ R; 2. ka − tk 6 t for all t > 1 2kak; ..23 -- 25.. 49 II III IV V 25 3. a ≡ b2 for some self-adjoint b ∈ A ; 4. a ≡ c∗c for some c ∈ A ; 5. sp(a) ⊆ [0,∞). II Exercise The fact that a∗a is positive for an element a of a C∗-algebra A has some nice consequences of its own needed later on. 1. Show that b 6 c =⇒ a∗ba 6 a∗ca for all b, c ∈ AR and a ∈ A . 2. Show that every mi-map and cp-map is positive. 3. Show that a 6 b−1 iff √ba√b 6 1 iff k√a√bk 6 1 iff b 6 a−1 for positive invertible elements a, b of A (and so a 6 b entails b−1 6 a−1). 4. Prove that (1 + a)−1a 6 (1 + b)−1b for 0 6 a 6 b from A . (Hint: add (1 + a)−1 + (1 + b)−1 to both sides of the inequality.) III Proposition The vector states of B(H ) are order separating (see 21 II) for every Hilbert space H . IV Proof By 21 VII it suffices to show that kTk = supx∈(H )1 hx, T xi for given T ∈ B(H )+. Since hx, T xi = (cid:10)T 1/2x, T 1/2x(cid:11) = kT 1/2xk2 for all x ∈ H , we have kTk = kT 1/2k2 = ( supx∈(H )1(cid:13)(cid:13)T 1/2x(cid:13)(cid:13) )2 = supx∈(H )1 hx, T xi. Corollary For a bounded operator T on a Hilbert space H , we have (cid:3) V 1. T is self-adjoint iff hx, T xi is real for all x ∈ (H )1; 2. 0 6 T iff 0 6 hx, T xi for all x ∈ (H )1; 3. kTk = supx∈(H )1 hx, T xi when T is self-adjoint. VI Proof This follows from 21 V and 21 VII because the vector states on B(H ) are (cid:3) order separating by III. 26 II The interaction between the multiplication and order on a C∗-algebra can be subtle, but when the C∗-algebra is commutative almost all peculiarities disap- pear. This is to be expected as any commutative C∗-algebra is isomorphic to a C∗-algebra of continuous functions on a compact Hausdorff space (as we'll see in 27 XXVII). Exercise Let A be a commutative C∗-algebra. Let a, b, c ∈ AR. 1. Show that a is the supremum of a and −a in AR. 2. Show that if a and b have a supremum, a ∨ b, in AR, then c + a ∨ b is the supremum of a + c and b + c. (Hint: prove that 1 3. Show that AR is a Riesz space, that is, a lattice ordered vector space. 2 (a + b + a − b) is the supremum of a and b in AR.) 4. Show that an miu-map f : A → B between commutative C∗-algebras preserves finite suprema and infima. Exercise Prove the Riesz decomposition lemma: For positive elements a, b, c of a commutative C∗-algebra A with c 6 a+b we have c ≡ a′ +b′ where 0 6 a′ 6 a and 0 6 b′ 6 b. III 2.4 Representation 2.4.1 . . . by Continuous Functions Now that we have have a firm grip on the positive elements of a C∗-algebra we turn to what is perhaps the most important fact about commutative C∗- algebras: that they are isomorphic to C∗-algebras of continuous functions on a compact Hausdorff space, via the Gelfand representation. 27 Setting A is a commutative C∗-algebra. The Gelfand representation of A is the miu-map γ : A → C(sp(A )) given Definition The spectrum of A , denoted by sp(A ), is the set of all miu-maps f : A → C. We endow sp(A ) with the topology of pointwise convergence. by γ(a)(f ) = f (a). Exercise Verify that the map sp(A ) → C, f 7→ f (a) is indeed continuous for every a ∈ A , and that γ is miu. Remark One might wonder if there is any connection between the spectrum sp(A ) of a commutative C∗-algebra, and the spectrum sp(a) of one of A 's elements (from 11 XIX); and indeed there is as we'll see in XVII (and 28 II). Our program for this paragraph is to show that the Gelfand representation γ is an miu-isomorphism. In fact, we will show that it gives the unit of an equivalence between the category of commutative C∗-algebras (with miu-maps) and the opposite of the category of compact Hausdorff spaces (with continuous maps). The first hurdle we take is the injectivity of γ -- that there are sufficiently many points in the spectrum of a commutative C∗-algebra, so to speak -- , and involves the following special type of order ideal. Definition A Riesz ideal of A is an order ideal I such that a ∈ I ∩ AR =⇒ a ∈ I. A maximal Riesz ideal is a proper Riesz ideal which is maximal among proper Riesz ideals. II III IV V VI VII ..25 -- 27.. 51 VIII Lemma Let I be a Riesz ideal of A . For all a ∈ A and x ∈ I we have ax ∈ I. IX Proof Since x = xR + ixI, it suffices to show that axR ∈ I and axI ∈ I. Note that xR, xI ∈ I, so we might as well assume that x is self-adjoint to begin with. Similarly, using that x+ ∈ I (because x+ = 1 2 (x + x) and x ∈ I) and x− ∈ I, we can reduce the problem to the case that x is positive. We may also assume that a is self-adjoint. Now, since x > 0 and −kak 6 a 6 kak, we have −kakx 6 ax 6 kakx by 23 VII, and so ax ∈ I, because kakx ∈ I. (cid:3) Exercise Verify the following facts about Riesz ideals. X 1. The least Riesz ideal that contains a self-adjoint element a of A is (a)m := { b ∈ A : ∃n ∈ N [ bR , bI 6 na ]}. Moreover, (a)m = A iff a is invertible, and we have (a) = (a)m when a > 0 (where (a) is the least order ideal that contains a, see 22 III). For non- positive a, however, we may have (a) 6= (a)m. 2. I +J is a Riesz ideal of A when I and J are Riesz ideals. (Hint: use 26 III.) But I + J might not be an order ideal when I and J are order ideals. 3. Each proper Riesz ideal is contained in a maximal Riesz ideal. XI Lemma A maximal Riesz ideal I of A is a maximal order ideal. XII Proof Let J be a proper order ideal with I ⊆ J. We must show that J = I. Let a ∈ J be given; we must show that a ∈ I. Since aR, aI ∈ J, it suffices to show that aR, aI ∈ I, and so we might as well assume that a is self-adjoint to begin with. Similarly, since a ∈ J, and it suffices to show that a ∈ I, because then − a 6 a 6 a entails a ∈ I, we might as well assume that a is positive. Note that the least ideal (a) that contains a is also a Riesz ideal by X. Hence I + (a) is a Riesz ideal by X Since a ∈ J, we have (a) ⊆ J, and so I + (a) ⊆ J is proper. It follows that a ∈ I + (a) = I by maximality of I. (cid:3) XIII Lemma Let I be a maximal Riesz ideal of A . Then there is an miu-map f : A → C with ker(f ) = I. XIV Proof Since I is a maximal order ideal by XI, there is a pu-map f : A → C with ker(f ) = I by 22 IV. It remains to be shown that f is multiplicative. Let a, b ∈ A be given; we must show that f (ab) = f (a)f (b). Surely, since f is unital, we have f (b − f (b)) = f (b) − f (b) = 0, an so b − f (b) ∈ ker(f ) ≡ I. Now, since I is a Riesz ideal, we have a(b − f (b)) ∈ I ≡ ker(f ) by VIII, and so 0 = f ( a(b − f (b)) ) = f (ab) − f (a)f (b). Hence f is multiplicative. (cid:3) XV Proposition Let a be a self-adjoint element of a C∗-algebra. Then a is not invertible iff there is f ∈ sp(A ) with f (a) = 0. Proof Note that if a is invertible, then f (a−1) is the inverse of f (a) -- and so f (a) 6= 0 -- for every f ∈ sp(A ). For the other, non-trivial, direction, assume that a is not invertible. Then by X the least Riesz ideal (a)m that contains a is proper, and can be extended to a maximal Riesz ideal I. By XIII there is an miu-map f : A → C with ker(f ) = I. Then f ∈ sp(A ) and f (a) = 0. (cid:3) Exercise Show that sp(a) = {f (a) : f ∈ sp(A )} for each self-adjoint a ∈ A . Exercise Prove that kγ(a)k = kak for each a ∈ A where γ is from XXVII. general case, use the C∗-identity.) (Hint: first assume that a is self-adjoint, and use XVII and 16 II. For the Conclude that the Gelfand representation γ : A → C(sp(A )) is injective, and that its range {γ(a) : a ∈ A } is a C∗-subalgebra of C(sp(A )). To show that γ is surjective, we use the following special case of the Stone -- Weierstrass theorem. Theorem Let X be a compact Hausdorff space, and let S be a C∗-subalgebra of C(X) which 'separates the points of X', that is, for all x, y ∈ X with x 6= y there is f ∈ S with f (x) 6= f (y). Then S = C(X). Proof Let g ∈ C(X)+ and ε > 0. To prove that S = C(X), it suffices to show that g ∈ S , and for this, it suffices to find f ∈ S with kf − gk 6 ε, because S is closed. It is convenient to assume that g(x) > 0 for all x ∈ X, which we may, without loss of generality, by replacing g by 1 + g. Let x, y ∈ X with x 6= y be given. We know there is f ∈ S with f (x) 6= f (y). Note that we can assume that f (x) = 0 (by replacing f by f − f (x)), and that f is self-adjoint (by replacing f by either fR or fI), and that f is positive (by replacing f by f+ or f−), and that f (y) = g(y) > 0 (by replacing f by g(y) f (y) f ), and that f 6 g(y) (by replacing f by f ∧ g(y)). Let y ∈ X be given. We will show that there is f ∈ S with 0 6 f 6 g + ε and f (y) = g(y). Indeed, since g is continuous there is an open neighbourhood V of y with g(y) 6 g(x)+ε for all x ∈ V . For each x ∈ X\V there is fx ∈ [0, f (y)]S with fx(x) = 0 and fx(y) = g(y) by XXII. Since the open subsets Ux := { z ∈ X : fx(z) 6 ε } with x ∈ X\V form an open cover of the closed (and thus compact) subset X\V , there are x1, . . . , xN ∈ X\U with Ux1∪···∪UxN ⊇ X\V . Define f := fx1 ∧ ··· ∧ fxN . Then f ∈ S , 0 6 f 6 g(y), f (y) = g(y), and f (x) 6 ε for every x ∈ X\V . We claim that f 6 g + ε. Indeed, if x ∈ X\V , then f (x) 6 ε 6 g(x) + ε. If x ∈ V , then f (x) 6 g(y) 6 g(x) + ε (by definition of V ). Hence f 6 g + ε. Thus for each y ∈ X there is fy ∈ S with 0 6 fy 6 g + ε and fy(y) = g(y). Since fy is continuous at y, and fy(y) = g(y), there is an open neighbourhood Uy of y with g(y) − ε 6 fy(x) for all x ∈ Uy. Since these open neighbourhoods cover X, and X is compact, there are y1, . . . , yN ∈ X with Uy1 ∪···∪ UyN = X. Define f := fy1 ∨ ··· ∨ fyN . Then f ∈ S , and g − ε 6 f 6 g + ε, giving ..27.. 53 XVI XVII XVIII XIX XX XXI XXII XXIII XXIV kf − gk 6 ε. (cid:3) XXV Lemma The spectrum sp(A ) of A is a compact Hausdorff space. XXVI Proof Since for each a ∈ A and f ∈ sp(A ) we have kf (a)k 6 kak by 20 V z 6 kak }, and we see that f (a) is an element of the compact set { z ∈ C : so sp(A ) is a subset of Qa∈A { z ∈ C : z 6 kak }, which is a compact Hausdorff space (by Tychonoff's theorem, under the product topology it inherits from the space of all functions A → C). So to prove that sp(A ) is a compact Hausdorff space, it suffices to show that sp(A ) is closed. In other words, we must show that if f : A → C is the pointwise limit of a net of miu-maps (fi)i, then f is an miu-map as well. But this is easily achieved using the continuity of addition, involution and multiplication on C, because, for instance, for a, b ∈ A , we have f (ab) = limi fi(ab) = limi fi(a)fi(b) = (limi fi(a)) (limi fi(b)) = f (a) f (b). (cid:3) XXVII Gelfand's Representation Theorem For a commutative C∗-algebra A , the Gelfand representation, γ : A → C(sp(A )) defined in III is an miu-isomorphism. XXVIII Proof We already know that γ is an injective miu-map (see IV and XVIII). So to prove that γ is an miu-isomorphism, it remains to be shown that γ is surjective. Since sp(A ) is a compact Hausdorff space (by XXV), and γ(A ) ≡ {γ(a) : a ∈ A } is a C∗-subalgebra of C(sp(A )) (by XVIII), it suffices to show that γ(A ) separates the points of sp(X) by XX. This is obvious, because for f, g ∈ sp(A ) with f 6= g there is a ∈ A with f (a) ≡ γ(a)(f ) 6= γ(a)(g) ≡ g(a). (cid:3) 28 While Gelfand's representation theorem is a result about commutative C∗- algebras, it tells us a lot about non-commutative C∗-algebras too, via their commutative C∗-subalgebras. II Exercise Let a be an element of a (not necessarily commutative) C∗-algebra A . We are going to use Gelfand's representation theorem to define f (a) for every continuous map f : sp(a) → C whenever a is contained in some commutative C∗-algebra. This idea is referred to as the continuous functional calculus. 1. Show that there is a least C∗-subalgebra C∗(a) of A that contains a. Given b ∈ C∗(a) show that bc = cb for all c ∈ A with ac = ca. 2. We call a ∈ A normal when C∗(a) is commutative. Show that a is normal iff aa∗ = a∗a iff aRaI = aIaR. 3. From now on assume a is normal so that C∗(a) is commutative. Show that j : 7→ (a), sp(C∗(a)) → sp(a) is a continuous map. Denoting the composition of the miu-maps C(sp(a)) f7→f◦j / C(sp(C∗(a))) ∼=, 27 XXVII / C∗(a) inclusion / / A . by Φ, we write f (a) := Φ(f ) for all f ∈ C(sp(a)). We have hereby defined, for example, aα when a > 0 and α ∈ (0,∞). From the fact that Φ is miu some properties of f (a) can be derived. Show, for example, that aαaβ = aα+β for all α, β ∈ (0,∞) when a > 0. 4. Given f ∈ C(sp(a)), show that f (a) is the unique element of C∗(a) with ϕ(f (a)) = f (ϕ(a)) for all ϕ ∈ sp(C∗(a)). 5. (Spectral mapping thm.) Show that sp(f (a)) = f (sp(a)) for f ∈ C(sp(a)). 6. Show that sp((a)) ⊆ sp(a) and (f (a)) = f ((a)) for every f ∈ C(sp(a)) and miu-map : A → B into a C∗-algebra B. 7. Given f ∈ C(sp(a)) and g ∈ C(f (sp(a))) show that g(f (a)) = (g ◦ f )(a). Show that (aα)β = aαβ for α, β ∈ (0,∞) and a ∈ A+. n )α 6 (b + 1 n )α for all n, because (a + 1 Theorem We have 0 6 a 6 b =⇒ aα 6 bα for all positive elements a and b of a C∗-algebra A , and α ∈ (0, 1]. Proof (Based on [61].) Note that the result is trivial if a and b commute. It suffices to show that (a + 1 n )α norm converges to aα as n → ∞. In other words, it suffices to prove aα 6 bα under the additional assumption that a and b are invertible. Note that a0 and b0 are defined for such invertible a and b, because the function (· )0 : [0, 1] → C is only discontinuous at 0. Writing E for the set of all α ∈ [0, 1] for which b 7→ bα is monotone on positive, invertible elements of A we must prove that E ⊇ (0, 1], and we will in fact show that E = [0, 1]. Since clearly 0, 1 ∈ E it suffices to show that E is convex. We'll do this by showing that E is closed, and α, β ∈ E =⇒ 1 (E is closed) Let b be a positive and invertible element of A . A moment's thought reveals it suffices to prove that α 7→ bα, [0, 1] → A is continuous. And indeed it is being the composition of the map α 7→ bα : [0, 1] → C(sp(b)), which is norm continuous, and the functional calculus f 7→ f (b) : C(sp(b)) → A , which being an miu-map is norm continuous as well. (α, β ∈ E =⇒ 1 2 β ∈ E) Let α, β ∈ E. Let a, b ∈ A be positive . Since the map and invertible with a 6 b. We must show that a 2 β ∈ E. 2 α + 1 2 α + 1 α+β 2 6 b α+β 2 III IV V VI ..27 -- 28.. 55 / / 2 b α+β α+β α+β 4 b− α+β 4 (· )b is positive (by 25 II), it suffices to show that b− α+β 4 a that is, kb− α+β For this, it seems, we must take a look under the hood of the theory of C∗-algebras: writing (c) := supλ∈sp(c) λ for c ∈ A , we know that (c) 6 kck for any c, and (c) = kck for self-adjoint c by 16 II. Moreover, recall from 19 Ia that sp(cd)\{0} = sp(dc)\{0}, and so (cd) = (dc) for all c, d ∈ A . Hence α+β 2 b− α+β 4 6 1, 4 k 6 1. 4 a k b− α+β 4 a α+β 2 b− α+β 4 k = ( b− α+β = ( b− α+β 4 a 4 a α+β 2 α+β 2 4 b− α+β b− α+β 4 α+β ) b− α−β b− α+β 4 ) α−β 4 b b− α−β 4 α−β b− α+β 4 2 4 ) 4 a = ( b = ( b−β/2 aβ/2 aα/2 b−α/2 ) 6 k b−β/2 aβ/2 k k aα/2 b−α/2 k = k b−β/2 aβ b−β/2 k1/2 k b−α/2 aα b−α/2 k1/2 6 k b−β/2 bβ b−β/2 k1/2 k b−α/2 bα b−α/2 k1/2 = 1, 29 II III and so we're done. (cid:3) As a cherry on the cake, we use Gelfand's representation theorem 27 XXVII to get an equivalence between the categories (cC∗miu)op and CH of continuous maps between compact Hausdorff spaces. To set the stage, we extend X 7→ C(X) to a functor CH → (cC∗miu)op by sending a continuous function f : X → Y to the miu-map C(f ) : C(Y ) → C(X) given by C(f )(g) = g ◦ f for g ∈ C(Y ), and we extend A 7→ sp(A ) to a functor sp : (cC∗miu)op → CH by sending an miu-map ϕ : A → B to the continuous map sp(ϕ) : sp(B) → sp(A ) given by sp(ϕ)(f ) = f ◦ ϕ. The Gelfand representations γA : A → C(sp(A )) form a natural isomor- phism from C◦sp to the identity functor on (cC∗miu)op. So to get an equivalence, it suffices to find a natural isomorphism from the identity on CH to sp◦C, which is provided by the following lemma. Lemma Let X be a compact Hausdorff space, and let τ : C(X) → C be an miu-map. Then there is x ∈ X with τ (f ) = f (x) for all f ∈ C(X). Proof Define Z = { x ∈ X : h(x) 6= 0 for some h ∈ C(X)+ with τ (h) = 0}. We'll prove X\Z contains exactly one point, x0, and τ (f ) = f (x0) for all f . IV To see that X\Z contains no more than one point, let x, y ∈ X with x 6= y be given; we will show that either x ∈ Z or y ∈ Z. By the usual topological trickery, we can find f, g ∈ C(X)+ with f g = 0, f (x) = 1 and g(y) = 1. Then 0 = τ (f g) = τ (f ) τ (g), so either τ (f ) = 0 (and x ∈ Z), or τ (g) = 0 (and y ∈ Z). That X\Z is non-empty follows from the following result (by taking f = 1). For f ∈ C(X)+ with f (x) > 0 =⇒ x ∈ Z for all x ∈ X we have τ (f ) = 0. Indeed, for each x ∈ X with f (x) > 0 (and so x ∈ Z) we can find h ∈ C(X)+ with τ (h) = 0 and h(x) 6= 0. Then f (x) < g(x) and τ (g) = 0 for g := ( f (x) h(x) +1)h. By compactness, we can find g1, . . . , gN ∈ C(X)+ with τ (gn) = 0, such that for every x ∈ X there is n with g(x) < fn(x). Writing g := g1 ∨ ··· ∨ gN , we have 0 6 f 6 g and τ (g) = 0 (because by 26 II τ preserves finite infima). It follows that τ (f ) = 0. We now know that X\Z contains exactly one point, say x0. To see that τ (f ) = f (x0) for f ∈ C(X), write g := (f − f (x0))∗(f − f (x0)) and note that g(x) > 0 =⇒ x 6= x0 =⇒ x ∈ Z. Thus by V, we get 0 = τ (g) = τ (f ) − f (x0)2, and so τ (f ) = f (x0). (cid:3) Exercise Let X be a compact Hausdorff space. Show that for every x ∈ X the map δx : C(X) → C, f 7→ f (x) is miu, and that the map X → sp(C(X)), x 7→ δx is a continuous bijection onto a compact Hausdorff space, and thus a home- omorphism. V VI VII Exercise As an application of the equivalence between (cC∗MIU)op and CH, we will show that every injective miu-map between C∗-algebras is an isometry. VIII Show that an arrow f : X → Y in CH is mono iff injective, and epi iff surjective (using complete regularity of Y ). Conclude that f is both epi and mono in CH only if f is an isomorphism (i.e. homeomorphism). Let : A → B be an injective miu-map between C∗-algebras. Let a be a self-adjoint element of A . Show that can be restricted to an miu-map σ : C∗(a) → C∗((a)), which is both epi and mono in cC∗MIU. Conclude that σ is an isomorphism, and thus k(a)k = kak. Use the C∗-identity to extend the equality k(a)k = kak to (not necessarily self-adjoint) a ∈ A . Exercise Let : A → B be an injective miu-map. Show that (A ) is closed (using VIII). Conclude that (A ) is a C∗-subalgebra of B isomorphic to A . 2.4.2 Representation by Bounded Operators Let us prove that every C∗-algebra A is isomorphic to a C∗-algebra of bounded operators on some Hilbert space. We proceed as follows. To each p-map ω : A → C (see 10 II) we assign a inner product [· , · ]ω on A , which can be "completed" to a Hilbert space Hω. Every element a ∈ A gives a bounded operator on Hω via the action b 7→ ab, which in turn gives a miu-map ω : A → B(Hω). In general ω is not injective, but if Ω is a set of p-maps which separates the points of A , then the composition A hωiω∈Ω /Lω∈Ω B(Hω) ..28 -- 30.. Hω ) / B(Lω∈Ω 57 IX 30 / / does give an injective miu-map , which restricts to an isomorphism (29 IX) from A to the C∗-algebra (A ) of bounded operators on Lω∈Ω The creation of ω from ω is known as the Gelfand -- Naimark -- Segal (GNS) construction and will make a reappearance in the theory of von Neumann alge- bras (in 72 V). Hω, see 6 II. II We take a somewhat utilitarian stance towards the GNS construction here, but there is much more that can be said about it: in the first chapter of my twin brother's thesis, [84], you'll see that the GNS construction has a certain universal property, and that it can be generalised to apply not only to maps of the form ω : A → C, but also to maps of the form ϕ : A → B. Lemma For every p-map ω : A → C on a C∗-algebra A , [a, b]ω = ω(a∗b) defines an inner product [· , · ]ω on A (see 4 VIII). Proof Note that [a, a]ω ≡ ω(a∗a) > 0 for each a ∈ A , because a∗a > 0 (by 24 IV); and [a, b]ω = [b, a]ω for a, b ∈ A , because ω is involution preserving (by 10 IV). Finally, it is clear that [a, · ]ω ≡ ω(a∗ · ) is linear for each a ∈ A . (cid:3) IV Exercise Let ω : A → C be a p-map on a C∗-algebra. Let us for a mo- ment study the semi-norm k · kω on A induced by the inner product [· , · ]ω (so kakω = ω(a∗a)1/2), because it plays an important role here, and all through- out the next chapter. III 1. Use Cauchy -- Schwarz (4 XV) to prove Kadison's inequality: for all a, b ∈ A , ω(a∗b)2 6 ω(a∗a) ω(b∗b). 2. Show that kabkω 6 kωk kak kbkω for all a, b ∈ A (using a∗a 6 kak2). 0 1 ) and b = 1 Show that we do not always have kabkω 6 kωkkakωkbk. 1 1 ) from A = M2, and ω( ( c d (Hint: take a = ( 0 0 e f ) ) = c.) Show that neither always kabkω 6 kakωkbkω, or ka∗akω = kak2 ω. (Hint: take a = b = 1 e f ) ) = c.) Give a counterexample to ka∗kω = kakω. 1 1 ) from A = M2, and ω(( ( c d 2 ( 1 1 2 ( 1 1 V Exercise Let us begin by showing how a complex vector space V with inner product [· , · ] can be "completed" to a Hilbert space H . We will take for H the set of Cauchy sequences on V modulo the following equivalence relation. Two Cauchy sequences (an)n and (bn)n in V are considered equivalent iff limn kan−bnk = 0. We "embed" V into H via the map η : V → H which sends a to the constant sequence a, a, a, . . . . Note, however, that η need not be injective: show that η(a) = η(b) iff ka − bk = 0 for all a, b ∈ V . Show that d( (an)n, (bn)n ) = limn kan− bnk defines a metric on H , that H is complete with respect to this metric, and that if (an)n is a Cauchy sequence in V , then (η(an))n converges to the element (an)n of H (so V is dense in H ). Show that every uniformly continuous map f : V → X to a complete metric space X can be uniquely extended to a uniformly continuous map g : H → X. (We say that g extends f when f = g ◦ η.) Show that addition, scalar multiplication, and inner product on V (being uniformly continuous) can be uniquely extended to uniformly continuous oper- ations on H , and turn H into a Hilbert space. (Also verify that the extended inner product agrees with the complete metric we've already put on H .) Show that every bounded linear map f : V → K to a Hilbert space K can be uniquely extended to a bounded linear map g : H → K . (Categorically speaking, Hilbert spaces form a reflexive subcategory of the category of bounded linear maps between complex vector spaces with an inner product.) Definition (Gelfand -- Naimark -- Segal construction) Let ω : A → C be a p-map on a C∗-algebra A . Let Hω denote the completion of A endowed with the inner product [· , · ]ω (see II) to a Hilbert space as discussed in V. Recall that we have an "embedding" ηω : A → Hω with hηω(a), ηω(b)i = [a, b]ω for all a, b ∈ A . Since given a ∈ A the map b 7→ ab, A → A is bounded with respect to k · kω (because kabkω 6 kωkkakkbkω by IV), it can be uniquely extended to a bounded linear map Hω → Hω (by the universal property of Hω, see V), which we'll denote by ω(a). So ω(a) is the unique bounded linear map Hω → Hω with ω(a)(ηω(b)) = ηω(ab) for all b ∈ A . Proposition The map ω : A → B(Hω) given by VI is an miu-map. Proof Let a1, a2 ∈ A be given. Since ω(a1 + a2) ηω(b) = ηω((a1 + a2)b) = ηω(a1b) + ηω(a2b) = (ω(a1) + ω(a2)) ηω(b) for all b ∈ A , and {ηω(b) : b ∈ A } is dense in Hω, we see that ω(a1 + a2) = ω(a1) + ω(a2). Since similarly ω(λa) = λω(a) for λ ∈ C and a ∈ A , we see that ω is linear. and so ω is unital, ω(1) = 1. Since ω(1) ηω(b) = ηω(b) for all b ∈ A , we have ω(1) x = x for all x ∈ Hω, To see that ω is multiplicative, note that (ω(a1) ω(a2)) ηω(b) = ηω(a1a2b) = VI VII VIII ω(a1a2) ηω(b) for all a1, a2, b ∈ A . Let a ∈ A be given. To show that ω is involution preserving it suffices to prove that ω(a∗) is the adjoint of ω(a). Since hω(a∗) ηω(b), ηω(c)i ≡ [a∗b, c]ω = ω(b∗ac) = [b, ac]ω ≡ hηω(b), ω(a) ηω(c)i for all b, c ∈ A , and {ηω(b) : b ∈ A } is dense in Hω, we get hω(a∗)x, yi = hx, ω(a)yi for all x, y ∈ Hω, and so ω(a∗) = ω(a)∗. (cid:3) Definition Given a collection Ω of p-maps ω : A → C on a C∗-algebra A , let Ω : A → B(HΩ) be the miu-map given by Ω(a)x = Pω∈Ω ω(a)x(ω), where HΩ =Lω∈Ω Proposition For a collection Ω of positive maps A → C on a C∗-algebra A , Hω (and ω is as in VI). IX X ..30.. 59 the following are equivalent. XI Proof XII XIII 1. Ω : A → B(HΩ) is injective; 2. Ω is centre separating on A (see 21 II); 3. Ω′ = { ω(b∗(· )b) : b ∈ A , ω ∈ Ω} is order separating on A . In that case, Ω(A ) is a C∗-subalgebra of B(HΩ), and Ω restricts to an miu- isomorphism from A to Ω(A ). It is clear that 3 entails 2. (2=⇒1) Let a ∈ A with Ω(a) = 0 be given. We must show that a = 0 (in order to show that Ω is injective), and for this it is enough to prove that a∗a = 0. Let b ∈ A and ω ∈ Ω be given. Since Ω is centre separating, it suffices to show that 0 = ω(b∗a∗ab) ≡ kabk2 ω. Since Ω(a) = 0, we have ω(a) = 0, thus 0 = ω(a) ηω(b) = ηω(ab), and so kabkω = 0. Hence Ω is injective. (1=⇒3) Let a ∈ A with ω(b∗ab) > 0 for all ω ∈ Ω and b ∈ A be given. We must show that a > 0. Since Ω is injective, we know by 29 IX that Ω(A ) is a C∗-subalgebra of B(HΩ), and Ω restricts to an miu-isomorphism from A to Ω(A ). So in order to prove that a > 0, it suffices to show that Ω(a) > 0, and for this we must prove that ω(a) > 0 for given ω ∈ Ω. Since the vector states on Hω are order separating by 25 III, it suffices to show that hx, ω(a)xi > 0 for given x ∈ Hω. Since {ηω(b) : b ∈ A } is dense in Hω, we only need to prove that 0 6 hηω(b), ω(a)ηω(b)i ≡ ω(b∗ab) for given b ∈ A , but this is true by assumption. (cid:3) XIV Theorem (Gelfand -- Naimark) Every C∗-algebra A is miu-isomorphic to a C∗- algebra of operators on a Hilbert space. XV Proof Since the states on A are separating (22 VIII), and therefore centre sep- arating, the miu-map Ω : A → B(HΩ) (defined in IX) restricts to an miu- isomorphism from A onto the C∗-subalgebra (A ) of B(HΩ) by X. (cid:3) 2.5 Matrices over C∗-algebras 31 We have seen (in 4) that the N × N -matrices (N being a natural number) over the complex numbers C form a C∗-algebra (denoted by MN ) by interpreting them as bounded operators on the Hilbert space CN , and proving that the bounded operators B(H ) on any Hilbert space H form a C∗-algebra. In this paragraph, we'll prove the analogous and more general result that the N × N -matrices over a C∗-algebra A form a C∗-algebra by interpreting them as adjointable module maps on the Hilbert A -module A N , see 32 I and 32 XIII. 32 Definition An (A -valued) inner product on a right A -module X (A being a C∗- algebra) is a map h· , ·i : X×X → A such that, for all x, y ∈ X, hx, ·i : X → A is a module map, hx, xi > 0, and hx, yi = hy, xi∗. We say that such an inner product is definite if hx, xi = 0 =⇒ x = 0 for all x ∈ X. A pre-Hilbert A -module X (where A is always assumed to be a C∗-algebra) is a right A -module endowed with a definite inner product. Such X is called a Hilbert A -module when it is complete with respect to the norm we'll define in IX. Let X and Y be pre-Hilbert A -modules. We say that a map T : X → Y is adjoint to a map S : Y → X when hT x, yi = hx, Syi for all x ∈ X and y ∈ Y . In that case, we call T adjointable. It is not difficult to see that T must be linear, and a module map, and adjoint to exactly one S, which we denote by T ∗. (Note that we did not require that T is bounded, and in fact, it need not be, see 35 IX. However, if T is bounded, then so is T ∗, see X, and if either X or Y is complete, then T is automatically bounded, see 35 VI.) The vector space of adjointable bounded module maps T : X → Y is denoted by Ba(X, Y ), and we write Ba(X) = Ba(X, X). Example We endow A N (where A is a C∗-algebra and N is a natural number) II with the inner product hx, yi =Pn x∗nyn, making it a Hilbert A -module. Exercise Let S and T be adjointable operators on a pre-Hilbert A -module. 1. Show that T ∗ is adjoint to T (and so T ∗∗ = T ). 2. Show that (T + S)∗ = T ∗ + S∗ and (λS)∗ = λS∗ for λ ∈ C. 3. Show that ST is adjoint to T ∗S∗ (and so (ST )∗ = T ∗S∗). and thus a Hilbert C[0, 1]-module. Exercise Although a bounded linear map between Hilbert spaces is always adjointable (see 5), a bounded module map between Hilbert A -modules might have no adjoint as is revealed by the following example (based on [60], p. 447). Prove that J := { f ∈ C[0, 1] : f (0) = 0 } is a closed right ideal of C[0, 1], Show that the inclusion T : J → C[0, 1] is a bounded module map, which has no adjoint by proving that there is no b ∈ J with hb, ai = T a ≡ a for all a ∈ J (for if T had an adjoint T ∗, then hT ∗1, ai = h1, T ai = a for all a ∈ J). Remark Note that part of the problem here is the lack of the obvious analogue to Riesz' representation theorem (5 IX) for Hilbert A -modules. One solution (taken in the literature) is to simply add Riesz' representation theorem as ax- iom giving us the self-dual Hilbert A -modules. For those who like to keep Riesz' representation theorem a theorem, I'd like to mention that it is also pos- sible to assume instead that the Hilbert A -module is complete with respect to a suitable uniformity, as in done in my twin brother's thesis, [84], see 149 V. III IV V ..30 -- 32.. 61 VI Proposition (Cauchy -- Schwarz) We have hx, yi hy, xi 6 khy, yik hx, xi for ev- ery inner product h· , ·i on a right A -module X, and x, y ∈ X. VII Remark The symmetry-breaking norm symbols "k" cannot simply be removed from this version of Cauchy -- Schwarz, because 0 6 hx, yi hy, xi 6 hy, yihx, xi would imply that hy, yihx, xi is positive, and self-adjoint, and thus that hy, yi and hx, xi commute, which is not always the case. VIII Proof Let ω : A → C be a state of A . Since the states on A are order separating (22 VIII), it suffices to show that ω(hx, yi hy, xi ) 6 khy, yik ω(hx, xi). Noting that (u, v) 7→ ω(hu, vi) is a complex-valued inner product on X, we compute ω(hx, yi hy, xi )2 = ω(hx, y hy, xii )2 6 ω(hx, xi) ω(h y hy, xi , y hy, xi i ) = ω(hx, xi) ω(hx, yi hy, yi hy, xi ) 6 ω(hx, xi) ω(hx, yihy, xi ) khy, yik by Cauchy -- Schwarz, 4 XV since hy, yi 6 khy, yik. It follows (also when ω(hx, yi hy, xi ) = 0), that ω(hx, yi hy, xi ) 6 khy, yik ω(hx, xi), and so we're done. (cid:3) IX Exercise Let X be a pre-Hilbert A -module. Verify that 1. kxk = khx, xik 2. kxbk 6 kxkkbk and khx, yik 6 kxk kyk for all x, y ∈ X and b ∈ A . 1/2 defines a norm k·k on X, and X Lemma For a linear map T : X → Y between pre-Hilbert A -modules, and B > 0, the following are equivalent. 1. kT xk 6 B kxk for all x ∈ X (that is, T is bounded by B); 2. khy, T xik 6 B kykkxk for all x ∈ X, y ∈ Y . Moreover, if T is adjointable, and bounded, then kT ∗k = kTk. XI Proof If kT xk 6 Bkxk for all x ∈ X, then T is bounded, kTk 6 B, and therefore khy, T xik 6 kyk kT xk 6 Bkykkxk for all x ∈ X and y ∈ Y using VI. On the other hand, if 2 holds, and x ∈ X is given, then we have kT xk2 = khT x, T xik 6 B kT xkkxk, entailing kT xk 6 Bkxk (also when kT xk = 0). If T is adjointable, and bounded, then khx, T ∗yik = khy, T xik 6 kTkkykkxk for all x ∈ X, y ∈ Y , so kT ∗k 6 kTk, giving us that T ∗ is bounded. Since by a similar reasoning kTk 6 kT ∗k, we get kTk = kT ∗k. (cid:3) Exercise Show that kT ∗Tk = kTk2 for every adjointable bounded map T on a pre-Hilbert A -module. (Hint: adapt the proof of 4 XVI.) Proposition The adjointable bounded module maps on a Hilbert A -module form a C∗-algebra Ba(X) with composition as multiplication, adjoint as invo- lution, and the operator norm as norm. Proof Considering 4 VII and XII, the only thing that remains to be shown is that Ba(X) is closed (with respect to the operator norm) in the set of all bounded linear maps B(X). So let T : X → X be a bounded linear map which is the limit of a sequence T1, T2, . . . of adjointable bounded module maps. To see that T has an adjoint, note that kT ∗n − T ∗mk = k(Tn − Tm)∗k = kTn − Tmk for all n, m, and so T ∗1 , T ∗2 , . . . is a Cauchy sequence, and converges to some bounded operator S on X. Since for x, y ∈ X and n, XII XIII XIV khSx, yi − hx, T yik 6 kh(S − T ∗n)x, yik + khx, (Tn − T )yik 6 kS − T ∗nkkxkkyk + kTn − Tkkxkkyk, we see that hSx, yi = hx, T yi, so S is the adjoint of T , and T is adjointable. (cid:3) Exercise Let X be a Hilbert A -module. Show that the vector states of Ba(X) are order separating (see 21 II). Conclude that for an adjointable operator T on X XV 1. T is self-adjoint iff hx, T xi is self-adjoint for all x ∈ (X)1; 2. 0 6 T iff 0 6 hx, T xi for all x ∈ (X)1; 3. kTk = supx∈(X)1 k hx, T xik when T is self-adjoint. (Hint: adapt the proofs of 25 III and 25 V.) Corollary The operator T ∗T is positive in Ba(X) for every adjointable opera- tor T : X → Y between Hilbert A -modules. Proof hx, T ∗T xi = hT x, T xi > 0 for all x ∈ X, and so T ∗T > 0 by 25 V. Exercise Let us consider matrices over a C∗-algebra A . (cid:3) XVI XVII 33 1. Show that every N × M -matrix A (over A ) gives a bounded module map A : A N → A M via A(a1, . . . , aN ) = A(a1, . . . , aN ), which is adjoint to A∗ (where A∗ = (A∗ji)ij is conjugate transpose). 2. Show that A 7→ A gives a linear bijection between the vector space of N × M -matrices over A and the vector space of adjointable bounded module maps Ba(A N , A M ). 3. Show that A ◦ B = AB when A is an N × M and B an M × K matrix. ..32 -- 33.. 63 4. Conclude that the vector space MN A of N × N -matrices over A is a C∗- algebra with matrix multiplication (as multiplication), conjugate trans- pose as involution, and the operator norm (as norm, so kAk = kAk). II Exercise Let us describe the positive N × N matrices over a C∗-algebra A . 1. Show that an N × N matrix A over A is positive iff 0 6Pi,j a∗i Aijaj for all a1, . . . , aN ∈ A . (Hint: use 25 III.) 2. Show that the matrix (hxi, xji )ij is positive for all vectors x1, . . . , xN from a pre-Hilbert A -module X. 3. Show that the matrix (a∗i aj)ij is positive for all a1, . . . , aN ∈ A . III Exercise Let f : A → B be a linear map between C∗-algebras. 1. Show that applying f entry-wise to an N × N matrix A over A (yield- ing the matrix (f (Aij ))ij over B) gives a linear map, which we'll denote by MN f : MN A → MN B. 2. The map MN f inherits some traits of f : show that if f is unital, then MN f unital; if f is multiplicative, then MN f is multiplicative; and if f is invo- lution preserving, then so is MN f . 3. However, show that Mnf need not be positive when f is positive, and that Mnf need not be bounded, when f is. 34 II Let us briefly return to the completely positive maps (defined in 10 II), to show that a map f between C∗-algebras is completely positive precisely when MN f is positive for all N , and to give some examples of completely positive maps. We also prove two lemmas stating special properties of completely positive maps (setting them apart from plain positive maps), that'll come in very handy later on. The first one is a variation on Cauchy -- Schwarz (XIV), and the second one concerns the points at which a cpu-map is multiplicative (XVIII). Completely positive maps are often touted as good models for quantum pro- cesses (over plain positive maps) with an argument involving the tensor product, and while we agree, we submit that the absence of analogues of XIV and XVIII for positive maps is already enough to make complete positivity indispensable. Lemma For a linear map f : A → B between C∗-algebras, and natural num- ber N , the following are equivalent. 1. MN f : MN A → MN B is positive; 2. Pij b∗i f (a∗i aj)bj > 0 for all a ≡ (a1, . . . , aN ) ∈ A N and b ∈ BN ; 3. the matrix ( f (a∗i aj) )ij is positive in MN B for all a ∈ A N . 1 )∗aT N )∗aT 1 +···+(aT Proof Recall that MN f is positive iff (MN f )(C) is positive for all C ∈ (MN A )+. The trick is to note that such C can be written as C ≡ A∗A for some A ∈ MN A , and thus as C ≡ (aT N , where an ≡ (An1, . . . , AnN ) is the n-th row of A. Hence MN f is positive iff (MN f )( (aT )∗aT ) ≡ ( f (a∗i aj) )i,j is positive for all tuples a ∈ A N . Since B ∈ MN B is positive iff hb, Bbi > 0 for all b ∈ BN , we conclude: MN f is positive iff 0 6(cid:10)b, (MN f )( (aT )∗aT )b(cid:11) =Pij b∗i f (a∗i aj)bj for all a ∈ A N and b ∈ BN . (cid:3) Exercise Conclude from II that a linear map f between C∗-algebras is com- pletely positive iff MN f is positive for all N iff for all N and a ∈ A N the matrix ( f (a∗i aj) )i,j is positive in MN B. III IV Deduce that the composition of cp-maps is completely positive. Show that a mi-map f is completely positive. (Hint: MN f is a mi-map too.) Exercise Show that given a C∗-algebra A , the following maps are completely positive: V 1. b 7→ a∗ba : A → A for every a ∈ A ; 2. T 7→ S∗T S : Ba(X) → Ba(Y ) for every adjointable operator S : Y → X between Hilbert A -modules; VI VII VIII 3. T 7→ hx, T xi , Ba(X) → A for every element x of a Hilbert A -module X. Exercise Show that the product of a family of C∗-algebras (Ai)i in the cate- gory C∗cpsu (see 10 II) is given by Li Ai with the same projections as in 20a I. Show that the equaliser of miu-maps f, g : A → B in C∗cpsu is the inclusion of the C∗-subalgebra { a ∈ A : f (a) = g(a)} of A into A . Lemma Let A be a commutative C∗-algebra, and let N be a natural num- ber. The set of matrices of the form Pk akBk, where a1, . . . , aK ∈ A+ and B1, . . . , BK ∈ MN (C)+, is norm dense in (MN A )+. Proof Since A is isomorphic to C(X) for some compact Hausdorff space X (by 27 XXVII)), we may as well assume that A ≡ C(X). Let A ∈ MN (C(X))+ and ε > 0 be given. We're looking for g1, . . . , gK ∈ C(X)+ and B1, . . . , BK ∈ (MN )+ with kA −Pk gkBkk 6 ε. Since A(x) := (Aij (x))ij gives a continuous map X → MN , the sets Ux = { y ∈ X : kA(x) − A(y)k < ε } form an open cover of X. By compactness of X this cover has a finite subcover; there are x1, . . . , xK ∈ X with Ux1 ∪ ··· ∪ UxK = X. Let y ∈ X be given. Since y ∈ Uxk for some k, there is, by complete regularity of X, a function fy ∈ (C(X))+ with fy(y) > 0 and supp(fy) ⊆ Uxk . Since the open subsets supp(fy) cover X there are (by compactness of X) finitely many y1, . . . yL with X = supp(fy1) ∪ ··· ∪ supp(fyL ), and so Pℓ fyℓ > 0. Let ..33 -- 34.. 65 , and us group together the fyℓs: pick for each ℓ an kℓ with supp(fyℓ) ⊆ Uxkℓ let gk :=P{fℓ : kℓ = k}. Then gk ∈ (C(X))+, supp(gk) ⊆ Uk, and Pk gk > 0. Upon replacing gk with (Pℓ gℓ)−1gk if necessary, we see that Pk gk = 1. Since supp(gk) ⊆ Uxk , we have −ε 6 A(x) − A(xk) 6 ε for all x ∈ supp(gk), and so −gk(x)ε 6 gk(x)A(x) − gk(x)A(xk) 6 gk(x)ε for all x ∈ X, that is, −gkε 6 gkA − gkA(xk) 6 gkε. Summing yields −ε 6 A −Pk gkA(xk) 6 ε, and so kA −Pk gkA(xk)k 6 ε. (cid:3) IX Proposition Let f : A → B be a positive map between C∗-algebras. If either A or B is commutative, then f is completely positive. Proof Suppose that B is commutative, and let a1, . . . , aN ∈ A , b1, . . . , bN ∈ B be given. We must show that Pi,j b∗i f (a∗i aj)bj is positive. This follows from the observation that ω(Pi,j b∗i f (a∗i aj)bj ) = ω(f (Pi,j(aiω(bi))∗ ajω(bj) )) > 0 for every ω ∈ sp(B). Suppose instead that A is commutative, and let A ∈ (MN A )+ be given for some natural number N . We must show that (MN f )(A) is positive in MN B. By VII, the problem reduces to the case that A ≡ aB where a ∈ A+ and B ∈ (MN )+. Since (MN f )(aB) ≡ f (a)B is clearly positive in MN B, we are done. (cid:3) XI X XII Lemma For a positive matrix A ≡(cid:0) p a a∗ q(cid:1) over a C∗-algebra A we have a∗a 6 kpkq and aa∗ 6 kqkp. In particular, if p = 0 or q = 0, then a = a∗ = 0. XIII Proof Since (x, y) 7→ hx, Ayi gives an A -valued inner product on A 2, aa∗ = h ( 1 6 kh ( 0 0 ) , A ( 0 1 ) , A ( 0 1 )i h ( 0 1 )ik h ( 1 1 ) , A ( 1 0 )i 0 ) , A ( 1 0 )i = kqk p by Cauchy -- Schwarz (see 32 VI). By a similar reasoning, we get a∗a 6 kpkq. between C∗-algebras and a, b ∈ A , provided that M2f is positive. in M2A is positive, the 2 × 2 matrix T := (cid:0) f (a∗a) f (a∗b) Thus we get f (a∗b)f (b∗a) 6 kf (b∗b)k f (a∗a) by XII. (cid:3) XIV Lemma We have f (a∗b)f (b∗a) 6 kf (b∗b)k f (a∗a) for every p-map f : A → B XV Proof Since writing x ≡ (a, b) ∈ A 2, the 2 × 2 matrix (xT )∗xT ≡ (cid:0) a∗a a∗b b∗a b∗b(cid:1) f (b∗a) f (b∗b)(cid:1) in M2B is positive. XVI Corollary kfk = kf (1)k for every cp-map f : A → B between C∗-algebras. XVII Proof Let a ∈ A be given. It suffices to show that kf (a)k 6 kf (1)k kak so that kfk 6 kf (1)k, because we already know that kf (1)k 6 kfk k1k = kfk. Since kf (a∗a)k 6 kf (1)k ka∗ak by 20 II, we have kf (a)k2 = kf (a)∗f (a)k = kf (a∗1)f (1∗a)k 6 kf (1∗1)k kf (a∗a)k 6 kf (1)k kf (1)kka∗ak = kf (1)k2kak2 by XIV, and so kf (a)k 6 kf (1)k kak. (cid:3) (cid:3) Lemma (Choi [12]) We have f (a)∗f (a) 6 f (a∗a) for every cpu-map f : A → B between C∗-algebras, and a ∈ A . Moreover, if f (a∗a) = f (a)∗f (a) for some a ∈ A , then f (ba) = f (b)f (a) for all b ∈ A . Proof By XIV we have f (a)∗f (a) = f (a∗1)f (1∗a) 6 kf (1∗1)kf (a∗a) = f (a∗a), where we used that f is unital, viz. f (1) = 1. Instead of f (ba) = f (b)f (a) we'll prove that f (a∗b) = f (a)∗f (b) (but this is nothing Let a, b ∈ A be given, and assume that f (a∗a) = f (a)∗f (a). more than a reformulation). Since M2f is cp, we have, writing A ≡(cid:0) a b 0 0(cid:1), (cid:18) f (a)∗f (a) f (b)∗f (a) f (a)∗f (b) f (b)∗f (b)(cid:19) = (M2f )(A)∗ (M2f )(A) 6 (M2f )(A∗A) = (cid:18) f (a∗a) f (b∗a) f (a∗b) f (b∗b)(cid:19) . Hence (using that f (a∗a) = f (a)∗f (a)) the following matrix is positive. (cid:18) 0 f (b∗a) − f (b)∗f (a) f (a∗b) − f (a)∗f (b) f (b∗b) − f (b)∗f (b)(cid:19) (cid:3) But then by XII we have f (a∗b) − f (a)∗f (b) = 0. We've just seen in 34 XVI that the norm of a completely positive map f : A → B between C∗-algebras is given by kfk = kf (1)k. We'll show here that the same result holds when f is just positive. This result will be useful at the end of this thesis in 128 VI, where we'll try to consider the broadest possible class of duplicators δ : A ⊗ A → A (see 127 I) being a priori just positive, not completely positive. The proof consists of two ingredients: the fact, II, that kf (a)k 6 kf (1)kkak for all normal a ∈ A (see 28 II), and the result, known as Russo -- Dye's theorem, VII, that the convex combinations of unitaries (being normal) are norm dense in the unit ball (A )1 of A . Lemma We have kf (a)k 6 kf (1)k kak for every p-map f : A → B between C∗-algebras, and normal a ∈ A . Proof Since a is normal, the C∗-subalgebra C∗(a) of A generated by a is commutative (see 28 II), and so the restriction of f to a map f : C∗(a) → B is completely positive by 34 IX. Thus kf (a)k 6 kf (1)k kak by 34 XVI. (cid:3) Definition An element u of a C∗-algebra is unitary when u∗u = 1 and uu∗ = 1. In that case we also say that u is a unitary. Exercise Let A be a C∗-algebra. 1. Show that any λ ∈ C with λ = 1 is unitary in A . In particular, the unit, 1, of A is unitary. ..34, 34a.. 67 XVIII XIX 34a II III IV V 2. Show that a unitary u ∈ A is invertible with inverse u−1 = u∗, and that u∗ is a unitary as well. 3. Show that the product uv of unitaries u, v ∈ A is unitary. 4. Show that every unitary u of A is normal, that is, uu∗ = u∗u (see 28 II). Show that a normal element a of A is unitary iff a2 R + a2 I = 1. of some unitary u, so a = uR ≡ 1 5. Show that every self-adjoint element a of A with kak 6 1 is the real part 2 (u + u∗). (Hint: take u := a + i√1 − a2.) 6. Show that every invertible element a of A can be written as a = u√a∗a, where u is a unitary. (Hint: take u =pa−1(a−1)∗.) This is a variation on the polar decomposition we'll see in 82 I. VI Exercise (Based on II.3.2.14 -- 17 of [3].) Let A be a C∗-algebra. 1. Show that every invertible element a of A with kak 6 2 can be written as the sum of two unitaries. (Hint: write a = u√a∗a with u as above.) 2. Let u ∈ A be a unitary, and a ∈ A with kak < 1. Show that u + a is the sum of two unitaries. (Hint: write u+a = u(1+u∗a), and note that 1+u∗a is invertible by 11 II.) 3. Let a ∈ A be given, and let N be a natural number with kak < N . Show that a is the sum of N + 2 unitaries. (Hint: write a = 1 + (N + 1)b where b := a−1 N +1 , and show that kbk < 1.) 4. Prove the following theorem. VII Theorem (Russo -- Dye) An element a of a C∗-algebra A with kak < 1 − 2 N (u1 + ··· + uN ) for some for some natural number N can be written as a = 1 unitaries u1, . . . , uN ∈ A . N VIII Corollary The operator norm of a positive linear map f : A → B between C∗-algebras is given by kfk = kf (1)k. IX Proof We must show that kf (a)k 6 kf (1)k for every a ∈ A with kak 6 1. Since by Russo -- Dye's theorem every a ∈ A with kak 6 1 may be approximated with respect to the norm by a sequence of elements of the form b := 1 N (u1 +···+ uN ), where u1, . . . , uN are unitaries, it suffices to show that kf (b)k 6 kf (1)k for such b. Since un is normal, and thus kf (un)k 6 kf (1)k kunk 6 kf (1)k by II, we get kf (b)k 6 1 N (kf (u1)k + ··· + kf (uN )k) 6 kf (1)k, and so kfk = kf (1)k. (cid:3) 2.6 Towards von Neumann Algebras Let us work towards the subject of the next chapter, von Neumann algebras, by pointing out two special properties of B(H ) on which the definition of a von Neumann algebra is based, namely that 35 1. any norm-bounded directed subset of self-adjoint operators on H has a supremum (in B(H )R), and 2. all vector functionals hx, (· )xi : B(H ) → C preserve these suprema. We'll end the chapter by showing in 39 IX that every functional on B(H ) that preserves the aforementioned suprema is a (possibly infinite) sum of vector func- tionals. 2.6.1 Directed Suprema Theorem (Uniform Boundedness) A set F of bounded linear maps from a complete normed vector space X to a normed vector space Y is bounded in the sense that supT∈F kTk < ∞ provided that supT∈F kT xk < ∞ for all x ∈ X . Proof Based on [74]. Let r > 0 and T ∈ F be given. Writing Br(x) = { y ∈ X : kx − yk 6 r } for the ball around x ∈ X with radius r, note that rkTk = supξ∈Br (0) kT ξk almost by definition of the operator norm. We will need the less obvious fact that rkTk 6 supξ∈Br(x) kT ξk for every x ∈ X . To see why this is true, note that for ξ ∈ Br(0) either kT ξk 6 kT (x + ξ)k or kT ξk 6 kT (x − ξ)k, because we would otherwise have 2kT ξk = kT (x + ξ) − T (x− ξ)k 6 kT (x+ ξ)k +kT (x− ξ)k < 2kT ξk. Hence rkTk = supξ∈Br (0) kT ξk 6 supξ∈Br (x) kT ξk. Suppose towards a contradiction that supT∈F kTk = ∞, and pick T1, T2, . . . with kTnk > n3n. Using IV, choose x1, x2, . . . in X with kxn − xn−1k 6 3−n and kTnxnk > 2 3 3−nkTnk, so that (xn)n is a Cauchy sequence, and therefore 2 3−n (becauseP∞k=0 3−k = 3 converges to some x ∈ X . Note that kx−xnk 6 1 2 ), and so kTnxk > kTnxnk − kTn(xn − x)k > 2 6 n, which contradicts the assumption that supT∈F kT xk < ∞. (cid:3) Theorem Let T : X → Y be an adjointable map between pre-Hilbert A - modules. If either X or Y is complete, then T and T ∗ are bounded. Proof We may assume without loss of generality that X is complete (by swap- ping T for T ∗ and X with Y if necessary). 2 3−nkTnk > 1 3 3−nkTnk − 1 Note that for every y ∈ Y , the linear map hy, T ·i ≡ hT ∗y, ·i : Y → A is bounded, because khT ∗y, xik 6 kT ∗ykkxk for all x ∈ X (see 32 VI). ..34a -- 35.. 69 II III IV V VI VII Since on the other hand, khy, T xik 6 kyk kT xk 6 kT xk for all x ∈ X and y ∈ Y with kyk 6 1, we have supkyk61 k hy, T xik 6 kT xk < ∞ for all x ∈ X, and thus B := supkyk61 k hy, T ·ik < ∞ by II. It follows that k hy, T xik 6 Bkykkxk for all y ∈ Y and x ∈ X, and thus T and T ∗ are bounded, by 32 X. (cid:3) VIII Remark As a special case of the preceding theorem we get the fact, known as the Hellinger -- Toeplitz theorem, that every symmetric operator on a Hilbert space is bounded. Example The condition that either X or Y be complete may not be dropped: the linear map T : c00 → c00 given by T α = (nαn)n for α ∈ c00 is self-adjoint, but not bounded, because T maps (1, 1 to (1, 1, . . . , 1, 0, 0, . . . ), which has 2-norm equal to √n. n , 0, 0, . . . ) having 2-norm below π√6 IX 2 , . . . , 1 36 II III Definition A Hilbert A -module X is self-dual when every bounded module map r : X → A is of the form r ≡ hy, (· )i for some y ∈ X. Example By Riesz' representation theorem (5 IX) every Hilbert space is self- dual. Exercise Show that given a C∗-algebra A the Hilbert A -module A N of N - tuples is self dual. IV Definition Let us say that a (bounded) form on Hilbert A -modules X and Y is a map [· , · ] : X × Y → A such that [x, · ] : Y → A and [· , y]∗ : X → A are (bounded) module maps for all x ∈ X and y ∈ Y . V Proposition For every bounded form [· , · ] : X×Y → A on self-dual Hilbert A - modules X and Y there is a unique adjointable bounded module map T : X → Y . with [x, y] ≡ hT x, yi for all x ∈ X and y ∈ Y . VI Proof Let x ∈ X be given. Since [x, · ] : Y → A is a a bounded module map, and Y is self-dual, there is a unique T x ∈ Y with [x, y] = hT x, yi for all y ∈ Y , giving a map T : X → Y . For a similar reason we get a map S : Y → X with hSy, xi = [x, y]∗ for all x ∈ X and y ∈ Y . Since S and T are clearly adjoint, they are bounded module maps by 35 VI. (cid:3) 37 II III Another consequence of 35 II is this: Proposition Given a net (yα)α in a Hilbert space H for which hyα, xi is Cauchy and bounded ‡ for every x ∈ H , there is a unique y ∈ H with hy, xi = limα hyα, xi for all x ∈ H . Proof To obtain y, we want to apply Riesz' representation theorem (5 IX) to the linear map f : H → C defined by f (x) = limα hyα, xi, but must first show that f ‡Recall that while every Cauchy sequence is bounded, a Cauchy net need only be eventually bounded. is bounded. For this it suffices to show that supα khyα, (· )ik < ∞, and this follows by 35 II from the assumption that supα hyα, xi < ∞ for every x ∈ H . By Riesz' representation theorem (5 IX), there is a unique y ∈ H with hy, xi = f (x) ≡ limα hyα, xi for all x ∈ H , and so we're done. (cid:3) Remark The condition in II that the net (hyα, xi )α be bounded for every x may not be omitted (even though (hyα, xi )α being Cauchy is eventually bounded). To see this, consider a linear map f : H → C on a Hilbert space H which is not bounded. We claim that there is a net (yα)α in H with f (x) = limα hyα, xi for all x ∈ H , and so there can be no y ∈ H with hy, xi = limα hyα, xi for all x ∈ H , because that would imply that f is bounded. To create this net, note that f is bounded on the span hFi of every finite subset F ≡ {x1, . . . , xn} of vectors from H , and so by Riesz' representation theorem 5 IX applied to f restricted to closed subspace hFi of H there is a unique yF ∈ hFi such that f (x) = hyF , xi for all x ∈ hFi. These yF 's form a net in H (when we order the finite subsets F of H by inclusion), which approximates f in the sense that f (x) = limF hyF , xi for every x ∈ H , (because f (x) = hyF , xi for every F with {x} ⊆ F ). Definition Let H be a Hilbert space. 1. The weak operator topology (WOT) on B(H ) is the least topology with respect to which T 7→ hx, T xi , B(H ) → C is continuous for every x ∈ H . So a net (Tα)α converges to T in B(H ) with respect to the weak operator topology iff hx, Tαxi → hx, T xi as α → ∞ for all x ∈ H . 2. The strong operator topology (SOT) on B(H ) is the least topology with 1/2 is continuous for every x ∈ H . respect to which T 7→ kT xk ≡ hx, T ∗T xi So a net (Tα)α converges to T in B(H ) with respect to the strong operator topology iff kTαx − T xk → 0 as α → ∞ for all x ∈ H . Remark Although we'll only make use of the weak operator topology we have nonetheless included the definition of the strong operator topology here for comparison with the ultrastrong topology that appears in the next chapter. Lemma Let (Tα)α be a net of bounded operators on a Hilbert space H such that (hx, Tαxi ) is Cauchy and bounded for every x ∈ H . Then (Tα)α WOT-converges to some bounded operator T in B(H ). IV V VI VII Proof Let x, y ∈ H be given. Since by a simple computation (c.f. 4 XV(4)) VIII hy, Tαxi = 1 4P3 k=0 ik(cid:10) iky + x, Tα(iky + x)(cid:11) , (hy, Tαxi )α is bounded for every y ∈ H , and so by II there is T x ∈ H with hy, T xi = limα hy, Tαxi for all y ∈ H , giving us a linear map T : H → H . It is clear that (Tα)α WOT-converges to T , provided that T is bounded. ..35 -- 37.. 71 So to complete the proof, we must show that T is bounded, and we'll do this by showing that T has an adjoint (see 35 VI). Note that hx, T ∗αxi = hx, Tαxi is Cauchy and bounded (with α running), so by a similar reasoning as before (but with T ∗α instead of Tα) we get a map S : H → H with hx, Syi = limα hx, T ∗αyi for all x, y ∈ H , which will be adjoint to T , which is therefore bounded. (cid:3) IX Proposition Let H be a Hilbert space, and D an upwards directed subset of B(H )R with supT∈D hx, T xi < ∞ for all x ∈ H . Then 1. (T )T∈D converges in the weak operator topology to some T ′ in (B(H ))R, 2. T ′ is the supremum of D in (B(H ))R, and 3. hx, T ′xi = supT∈D hx, T xi for all x ∈ H . X Proof Let x ∈ H . Since hx, (· )xi : B(H ) → C is positive we see that (hx, T xi)T∈D is an increasing net in R, bounded from above (by assumption), and therefore converges to supT∈D hx, T xi. In particular, (T )T∈D is WOT- Cauchy, and "WOT-bounded", and thus (by VII) WOT-converges to some self- adjoint T ′ from B(H ). Since (hx, T xi )T∈D converges both to hx, T ′xi, and to supT∈D hx, T xi, we In particular, conclude that hx, T ′xi = supT∈D hx, T xi for every x ∈ H . hx, T xi 6 hx, T ′xi for all x ∈ H and T ∈ D, and thus T 6 T ′ for all T ∈ D. Let S be a self-adjoint bounded operator on H with T 6 S for all T ∈ D. To prove that T ′ is the supremum of D, we must show that T ′ 6 S. Let x ∈ H be given. Since hx, T xi 6 hx, Sxi for each T ∈ D (because T 6 S), we have hx, T ′xi ≡ supT∈D hx, T xi 6 hx, Sxi, and therefore T ′ 6 S by 25 V. (cid:3) directed subset D in (B(H ))R (which exists by IX) is denoted by W D. XI Definition Let H be a Hilbert space. The supremum of a (norm) bounded 2.6.2 Normal Functionals 38 Definition Given a Hilbert space H a p-map ω : B(H ) → C is called normal when ω(W D) =WT∈D ω(T ) for every bounded directed subset D of B(H )R. Ia Notation We use the letter "n" to abbreviate "normal" in line with 10 II. So an npu-map ω : B(H ) → C is a normal positive unital linear functional on B(H ). Example All vector functionals hx, (· )xi are normal by 37 IX. Exercise To show that a positive linear functional is normal, it suffices to show that it preserves directed suprema of effects: show that given a Hilbert space H II III a positive map ω : B(H ) → C is normal provided that ω(W D) = WT∈D ω(T ) for every directed subset D of [0, 1]B(H ). It is easy to see that ω is linear and positive, so we'll only show that ω Lemma Every sequence x1, x2, . . . in a Hilbert space H with Pn kxnk2 < ∞ gives an np-map ω : B(H ) → C defined by ω(T ) =Pn hxn, T xni. Proof Given T ∈ B(H ) we have hxn, T xni 6 kxnk2kTk by Cauchy -- Schwarz (4 XV), so Pn hxn, T xni 6 kTkPn kxnk2, which means that Pn hxn, T xni converges, and so we may define ω as above. is normal. We must prove that ω(W D) = WT∈D ω(T ) for every bounded di- rected subset of (B(H ))R. By III we may assume without loss of generality that D ⊆ [0, 1]B(H ). This has the benefit that hxn, T xni is positive for all n and T ∈ D, so that their sum (over n) is given by a supremum over partial sums, viz. Pn hxn, T xni = WNPN n=1 hxn, T xni. Completing the proof is now simply a matter of interchanging suprema, WT∈D ω(T ) = WT∈DWNPN = WNWT∈DPN = WNPN n=1 hxn, (· )xni is normal. n=1 hxn, T xni n=1 hxn, T xni n=1 hxn, (W D) xni = ω(W D), (cid:3) where we used that PN Exercise The following observations regarding a net (xα)α in a Hilbert space H will be useful later on. IV V VI respect to the operator norm to some bounded functional on B(H ). 1. Show that Pα kxαk2 < ∞ if and only if Pα hxα, (· )xαi converges with 2. Given some x ∈ H , show that xα converges to x if and only if hxα, (· )xαi operator-norm converges to hx, (· )xi. (For the "if" part it may be convenient to first prove that hxα, xi → hx, xi by considering the bounded operator xihx on B(H ).) The final project of this chapter is to show that each normal positive func- tional ω on a B(H ) is of the form ω ≡ P∞n=0 hxn, (· )xni for some x1, x2, . . . with Pn kxnk2 < ∞. For this we'll need some more nuggets from the theory of Hilbert spaces. Definition A subset E of a Hilbert space is called orthonormal if he, e′i = 0 for all e, e′ ∈ E with e 6= e′, and he, ei = 1 for all e ∈ E . We say that E is maximal when E is maximal among all orthonormal subsets of H ordered by inclusion, and in that case we call E an orthonormal basis for H for reasons that will be become clear in IV below. Remark Clearly, by Zorn's lemma, each Hilbert space has an orthonormal basis. Proposition Given an orthonormal subset E of a Hilbert space H , and x ∈ H , 39 II III IV ..37 -- 39.. 73 V 1. (Bessel's inequality) Pe∈E he, xi2 6 kxk2; 2. Pe∈E he, xi e converges in H , 3. Pe∈E he, xi e = x if E is maximal, and 4. (Parseval's identity) Pe∈E he, xi2 = kxk2 if E is maximal. Proof 1 Since for any finite subset F of E we have 0 6 kx−Pe∈F he, xi ek2 = kxk2−2Pe∈F he, xihx, ei+Pe,e′∈F hx, e′ihe′, eihe, xi = kxk2−Pe∈F he, xi2, and so Pe∈F he, xi2 6 kxk2, we get Pe∈E he, xi2 6 kxk2. 2 From the observation that kPe∈F he, xi ek2 = Pe∈F he, xi2 for any fi- nite F ⊆ E , and the fact thatPe∈E he, xi2 converges (by the previous point), one deduces that (Pe∈F he, xi e)F is Cauchy, and so Pe∈E he, xi e converges. 3 Writing y := Pe∈E he, xi e we must show that x = y. If it were not so, that is, x 6= y, then e′ := kx − yk−1(x − y) satisfies he′, e′i = 1 and he′, ei = 0 for all e ∈ E , and so may be added to E to yield an orthonormal basis E ∪ {e′} extending E contradicting E s maximality. 4 Finally, kxk2 = hx, xi =Pe,e′∈E hx, e′ihe′, eihe, xi =Pe∈E he, xi2. (cid:3) 1. Show that Pe∈E eihe converges to 1 in the weak operator topology. 2. Show thatPe∈E eihe = 1 also in the sense that the directed set of partial sumsPe∈F eihe, where F is a finite subset of E , has 1 as its supremum. 3. Conclude that ω(1) =Pe∈E ω(eihe) for every np-map ω : B(H ) → C. VI Exercise Let E be an orthonormal basis of a Hilbert space H . VII Lemma Given a Hilbert space H with orthonormal basis E , we have ω(A) = Xe,e′∈E he, Ae′i ω(eihe′ ). for every normal p-map ω : B(H ) → C and A ∈ B(H ). VIII Proof Let F be a finite subset of E , and write P = Pe∈F eihe. Since P AP = Pe,e′∈F he, Ae′i eihe′ it suffices to show that ω(A − P AP ) vanishes as F increases. Note that P ∗P = P and (P ⊥)∗P ⊥ = P ⊥. Further, since kPk 6 1, and A − P AP = P ⊥A + P AP ⊥, we have, by Kadison's inequality, ω(A − P AP ) 6 (cid:12)(cid:12)ω(P ⊥A)(cid:12)(cid:12) + (cid:12)(cid:12)ω(P AP ⊥)(cid:12)(cid:12) 6 ω(P ⊥)1/2 ω(A∗A)1/2 + ω(P AA∗P )1/2 ω(P ⊥)1/2 6 2kAkω(1)1/2 ω(P ⊥)1/2. But since Pe∈E ω(eihe) = ω(1) by VI we see that ω(P ⊥) → 0 as F → ∞. (cid:3) IX X 40 Theorem Let H be a Hilbert space. Every normal p-map ω : B(H ) → C is of the form ω =Pn hxn, (· )xni where x1, x2, . . . ∈ H with Pn kxnk2 = kωk. Proof By 36 V there is a unique ∈ B(H ) with ω(yihx) = hx, yi for all x, y ∈ H , because (x, y) 7→ ω(yihx), H × H → C is a bounded form in the sense of 36 IV. Note that is positive by 25 V because hx, xi = ω(xihx) > 0 for all x ∈ H . Now, let E be an orthonormal basis for H . Since ω is normal, VI Thus, we are done if can show that ω′ = ω, (because √e is non-zero for at gives us ω(1) = Pe∈E ω(eihe) = Pe∈E he, ei = Pe∈E k√ek2, so that ω′ := Pe∈E (cid:10)√e, (· )√e(cid:11) defines a normal positive functional on B(H ) by 38 VI. most countably many e ∈ E ). To this end, note that ω(xihx) =(cid:10)√x,√x(cid:11) = Pe∈E (cid:10)√x, e(cid:11)(cid:10)e,√x(cid:11) = Pe∈E (cid:10)√e,xihx√e(cid:11) = ω′(xihx) for each x ∈ H , and so ω(xihy) = ω′(xihy) for all x, y ∈ H by polarisation, and thus ω = ω′ by VII. (cid:3) In this chapter we've studied the algebraic structure of the space B(H ) of bounded operators on a Hilbert space H abstractly via the notion of a C∗- algebra. We've seen not only that every C∗-algebra is miu-isomorphic to a C∗-subalgebra of such a B(H ) (in 30 XIV), but also that any commutative C∗- algebra is miu-isomorphic to the space C(X) of continuous functions on some compact Hausdorff space (in 27 XXVII). But there's more to B(H ) than just being a C∗-algebra: it has the two additional properties of having suprema of bounded directed subsets (see 37 IX), and having a faithful collection of normal functionals (viz. the vector functionals, 25 III). This leads us to the study of von Neumann algebras -- the topic of the next chapter. ..39 -- 40 75 Chapter 3 Von Neumann Algebras We have arrived at the main subject of this thesis, the special class of C∗- algebras called von Neumann algebras (see definition 42 below) that are char- acterised by the existence of certain directed suprema and an abundance of functionals that preserve these suprema. While all C∗-algebras and the cpsu- maps between them may perhaps serve as models for quantum data types and processes, respectively, we focus for the purposes of this thesis our attention on the subcategory W∗cpsu of von Neumann algebras and the cpsu-maps between them that preserve these suprema (called normal maps, see 44 XV), because 1. W∗cpsu is a model of the quantum lambda calculus (in a way that C∗cpsu is 41 not, see 125 X), and 2. we were able to axiomatise the sequential product (b 7→ √ab√a) in W∗cpsu (but not in C∗cpsu) see 106 I. Both these are reserved for the next chapter; in this chapter we'll (re)develop the theory we needed to prove them. The archetypal von Neumann algebra is the C∗-algebra B(H ) of bounded operators on a Hilbert space H . In fact, the original [55,78] and common [16,47] definition of a von Neumann algebra is a C∗-subalgebra A of a B(H ) that is closed in a "suitable topology" such as the strong or weak operator topology (see 37 V). Most authors make the distinction between such rings of operators (called von Neumann algebras) and the C∗-algebras miu-isomorphic to them (called W ∗-algebras), but we won't bother and call them all von Neumann algebras. Partly because it seems difficult to explain to someone picturing a quantum data type the meaning of the weak operator topology and the Hilbert space H , we'll use Kadison's characterisation [46] of von Neumann algebras as C∗-algebras with a certain dcpo-structure (c.f. 37 IX) and sufficiently many Scott-continuous functionals (c.f. 38 I) as our definition instead, see 42. 41.. 77 But we also use Kadison's definition just to see to what extent the repre- sentation of von Neumann algebras as rings of operators (see 48 VIII) can be avoided when erecting the basic theory. Instead we'll put the directed suprema and normal positive functionals on centre stage. All the while our treatment doesn't stray too far from the beaten path, and borrows many arguments from the standard texts [47, 68]; but most of them had to be tweaked in places, and some demanded a complete overhaul. The material on von Neumann algebras is less tightly knit as the theory of C∗-algebras, and so after the basics we deal with four topics more or less in linear order (instead of intertwined.) The great abundance of projections (elements p with p∗p = p) in von Neu- mann algebras -- a definite advantage over C∗-algebras -- is the first topic. We'll see for example that the existence of norm bounded directed suprema in a von Neumann algebra A allows us to show that there is a least projection ⌈a⌉ above any effect a from A given by ⌈a⌉ =Wn a1/2 (see 56 I); and also that any element of a von Neumann algebra can be written as a norm limit of linear combina- tions of projections (in 65 IV). Many a result about von Neumann algebras can be proven by an appeal to projections. n The second topic concerns two topologies that are instrumental for the more delicate results and constructions: the ultraweak topology induced by the normal positive functionals ω : A → C, and the ultrastrong topology induced by the associated seminorms k · kω (see 42). We'll show among other things that a von Neumann algebra is complete with respect to the ultrastrong topology and bounded complete with respect to the ultraweak topology (see 77 I). This completeness allows us to define, for example, for any pair a, b of elements from a von Neumann algebra A with a∗a 6 b∗b an element a/b with a = (a/b) b (see 81 I) -- this is the third topic. Taking b = √a∗a we ob- tain the famous polar decomposition a = (a/√a∗a)√a∗a (see 82 I, which is usually proven for a bounded operator on a Hilbert space first). The fourth, and final topic, is ultraweakly continuous functionals on a von Neumann algebra: we'll show in 90 II that any centre separating collection (21 II) of normal positive functionals Ω on a von Neumann algebra completely deter- mines the normal positive functionals, which will be important for the definition of the tensor product of von Neumann algebras in the next chapter, see 108 II. 3.1 The Basics 3.1.1 Definition and Counterexamples 42 Definition (Kadison [46]) A C∗-algebra A is a von Neumann algebra when 1. every bounded directed subset D of self-adjoint elements of A (so D ⊆ 2. if a is a positive element of A with ω(a) = 0 for every normal (see below) AR) has a supremum W D in AR, and positive linear map ω : A → C, then a = 0.∗ for every bounded directed subset of self-adjoint elements of D which has a A positive linear map ω : A → C is called normal if ω(W D) = Wd∈D ω(d) supremum W D in AR. Recall that we use the letter "n" as abbreviation for "normal", see 38 Ia. The ultraweak topology on A is the least topology that makes all normal pos- itive linear maps ω : A → C continuous; the ultraweakly open subsets of A are exactly the unions of finite intersections of sets of the form ω−1(U ), where ω : A → C is an np-map, and U is an open subset of C. One can verify that a net (bα)α in A converges ultraweakly to some b in A iff ω(bα) → b for all np-maps ω : A → C. The ultrastrong topology on A is the topology in- duced by the seminorms k · kω associated to the np-maps ω : A → C (given by kakω ≡ ω(a∗a)1/2, see 30 IV); a subset of A is ultrastrongly open iff it is the union of a finite intersections of sets of the form { a ∈ A : ka − bkω 6 ε }, where b ∈ A , ω : A → C is an np-map, and ε > 0. One can prove that a net (bα)α in A converges ultrastrongly to an element b of A iff kbα − bkω → 0 for all np-maps ω : A → C. Remark We work with the ultraweak and ultrastrong topology in tandem, because neither is ideal, and they tend to be complementary: for example, a 7→ a∗ is ultraweakly continuous but not ultrastrongly (see 43 II, point 4), while a 7→ a is ultrastrongly continuous (see 74 III) but not ultraweakly (43 II, point 6). This doesn't prevent the ultraweak topology from being weaker than the ultrastrong topology: a net that converges ultrastrongly converges ultra- weakly as well, see 43 I. Examples 1. C and {0} are clearly von Neumann algebras. 2. The C∗-algebra B(H ) of bounded operators on a Hilbert space H is a von Neumann algebra: B(H ) has bounded directed suprema of self- adjoint elements by 37 IX, and the vector states (and thus all normal func- tionals) are order separating (and thus faithful) by 25 III. II IIa III IV V 3. The direct sumLi is itself a von Neumann algebra. Ai (see 3 V) of a family (Ai)i of von Neumann algebras ∗In other words, the collection of normal positive functionals should be faithful (see 21 II). Interestingly, it's already enough for the normal positive functionals to be centre separating, but since we have encountered no example of a von Neumann algebra where it wasn't already clear that the normal positive functionals are faithful instead of just centre separating we did not use this weaker albeit more complex condition. ..41 -- 42.. 79 (While we're not quite ready to define morphisms between von Neumann algebras, we can already spoil that the direct sum gives the categorical product of von Neumann algebras once we do, see 47 IV.) 4. A C∗-subalgebra B of a von Neumann algebra A is called a von Neumann subalgebra (and is itself a von Neumann algebra) if for every bounded directed subset D of self-adjoint elements from B we have W D ∈ B (where the supremum is taken in AR). 4a. Let S be a subset of a von Neumann algebra A . Since the intersection of an arbitrary collection of von Neumann subalgebras of A is a von Neu- mann subalgebra of A as well, there is a least von Neumann subalgebra, W ∗(S), that contains S. 5. We'll see in 65 III that given a subset S of a von Neumann algebra A the set S(cid:3) = { a ∈ A : ∀s ∈ S [ as = sa ]} called the commutant of S is a von Neumann subalgebra of A when S is closed under involution. 6. We'll see in 49 IV that the N × N -matrices over a von Neumann algebra A form a von Neumann algebra. 7. We'll see in 51 IX that the bounded measurable functions on a finite com- plete measure space X (modulo the negligible ones) form a commutative von Neumann algebra L∞(X). (Recall that a measure space X is called finite when µ(X) < ∞.) 43 Exercise Let A be a von Neumann algebra. 1. Show that ω(a) 6 kakωkωk1/2 for every np-map ω : A → C and a ∈ A . 2. Show that when a net (aα)α in A converges ultrastrongly to a ∈ A it does so ultraweakly, too. 3. Show that an ultraweakly closed subset C of A is also ultrastrongly closed. Ia II Exercise Note that given a von Neumann algebra A the map a 7→ −a : A → A is an order reversing isomorphism. Deduce from this that any bounded filtered† subset F of self-adjoint elements of A has as infimum V F := −W{ −d : d ∈ F }. Exercise We give some counterexamples in B(ℓ2) to plausible propositions to sharpen your understanding of the ultrastrong and ultraweak topologies, and so that you may better appreciate the strange manoeuvres we'll need to pull off later on. †'Filtered' is the order dual of 'directed': F is filtered when for all a, b ∈ F there is c ∈ F with c 6 a and c 6 b. 1. First some notation: given n, m ∈ N, we denote by nihm the bounded operator on ℓ2 given by (nihm)(f )(n) = f (m) and (nihm)(f )(k) = 0 for k 6= n and f ∈ ℓ2. Verify the following computation rules, where k, ℓ, m, n ∈ N. (nihm)∗ = mihn , nihmℓihk = ( nihk 0 if m = ℓ otherwise n=0 nihn = 1. 2. Show that WNPN Conclude that (nihn )n converges ultrastrongly (and ultraweakly) to 0. Thus ultrastrong (and ultraweak) convergence does not imply norm con- vergence, which isn't unexpected. But we also see that if a sequence (bn)n converges ultrastrongly (or ultraweakly) to some b, then (kbnk)n doesn't even have to converge to kbk. (Note that (nihn)n resembles a 'moving bump'.) 3. Note that when a net (aα)α converges ultrastrongly to a, then ( a∗αaα )α is norm-bounded and converges ultraweakly to a∗a. The converse does not hold: show that (already in C) ein does not converge ultraweakly (nor ultrastrongly) as n → ∞, while 1 ≡ e−inein is norm- bounded and converges ultraweakly to 1 as n → ∞. 4. Show that (0ihn )n converges ultrastrongly (and ultraweakly) to 0. Deduce that (nih0 )n converges ultraweakly to 0, but doesn't converge ultrastrongly at all. Conclude that a 7→ a∗ is not ultrastrongly continuous on B(ℓ2). (This has the annoying side-effect that it is not immediately clear that the ultrastrong closure of a C∗-subalgebra of a von Neumann algebra is a von Neumann subalgebra; we'll deal with this by showing that the ultrastrong closure coincides with the ultraweak closure in 73 VIII.) 5. Show that the unit ball ( B(ℓ2) )1 of B(ℓ2) is not ultrastrongly compact by proving that (0ihn )n has no ultrastrongly convergent subnet. (But we'll see in 77 III that the unit ball of a von Neumann algebra is ultraweakly compact.) 6. Show that nih0 + 0ihn converges ultraweakly to 0 as n → ∞, while (nih0 + 0ihn)2 ≡ 0ih0 + nihn converges ultraweakly to 0ih0. Conclude that a 7→ a2 is not ultraweakly continuous on B(ℓ2). Conclude that a, b 7→ ab is not jointly ultraweakly continuous on B(ℓ2). ..42 -- 43.. 81 Prove that nih0 + 0ihn = 0ih0 + nihn. Conclude that a 7→ a is not ultraweakly continuous on (B(ℓ2))R. (We'll see in 74 I that a 7→ a is ultrastrongly continuous on self-adjoint elements.) 7. Let us consider the two extensions of · to arbitrary elements, namely a 7→ √a∗a =: as and a 7→ √aa∗ =: ar (for support and range, c.f. 59 VII). Prove that 0ih0 + 0ihn converges ultrastrongly to 0ih0 as n → ∞. Show that 0ih0 + 0ihns = 0ih0 + 0ihn + nih0 + nihn converges ultraweakly to 0ih0s ≡ 0ih0 as n → ∞, but not ultrastrongly. Show that 0ih0 + 0ihnr = √20ih0. Conclude that ·s and ·r are not ultrastrongly continuous on B(ℓ2). 8. Show that 1 + nih0 + 0ihn is positive, and converges ultraweakly to 1 as n → ∞, while the squares 1 +nihn+0ih0+ 2 nih0+ 2 0ihn converge ultraweakly to 1 + 0ih0 (as n → ∞). Hence a 7→ a2 and a 7→ √a are not ultraweakly continuous on B(ℓ2)+. 9. For the next counterexample, we need a growing moving bump, which still converges ultraweakly. Sequences won't work here: Show that nnihn does not converge ultraweakly as n → ∞. Show that nf (n)ihf (n) does not converge ultraweakly as n → ∞ for every strictly monotone (increasing) map f : N → N. So we'll resort to a net. Let D be the directed set which consists of pairs (n, f ), where n ∈ N\{0} and f : N → N is monotone, ordered by (n, f ) 6 (m, g) iff n 6 m and f 6 g. Show that the net ( nf (n)ihf (n) )n,f∈D converges ultrastrongly to 0. So a net which converges ultrastrongly need not be bounded! (The cure for this pathology is Kaplansky's density theorem, see 74 IV.) Show that 1 Show that the product f (n)ih0 = ( nf (n)ihf (n) ) ( 1 not converge ultrastrongly as D ∋ (n, f ) → ∞. Conclude that multiplication a, b 7→ ab is not jointly ultrastrongly contin- uous on B(ℓ2), even when b is restricted to a bounded set. (Nevertheless we'll see that multiplication is ultrastrongly continuous when a is restricted to a bounded set in 45 VI.) n f (n)ih0 converges ultrastrongly to 0 as D ∋ (n, f ) → ∞. n f (n)ih0 ) does 10. Show that an,f = 1 n (f (n)ih0 + 0ihf (n)) + nf (n)ihf (n) converges ul- trastrongly to 0 as D ∋ (n, f ) → ∞, while a2 n,f does not. Hence a 7→ a2 is not ultrastrongly continuous on B(ℓ2)R. 11. Let us show that B(ℓ2) is not ultraweakly complete. Show that there is an unbounded linear map f : ℓ2 → C (perhaps using the fact that every vector space has a basis by the axiom of choice), and that for each finite dimensional linear subspace S of ℓ2 there is a unique vector xS ∈ S with f (x) = hxS, yi for all y ∈ S (using 5 IX). Consider the net (eihxS )S where S ranges over the finite dimensional subspaces of ℓ2 ordered by inclusion, and e is some fixed vector in ℓ2 with kek = 1. Let ω : B(ℓ2) → C be an np-map, so ω ≡Pn hyn, (· )yni for y1, y2, . . . ∈ ℓ2 with Pn kynk2 < ∞, see 39 IX. Show that ω(eihxS − eihxT ) = h xS − xT , Pn yn hyn, eii = 0 when vector Pn yn hyn, ei. Conclude that (eihxS )S is ultraweakly Cauchy. Show that if (eihxS )S converges ultraweakly to some A in B(ℓ2), then we have he, Ayi = f (y) for all y ∈ ℓ2. Conclude that (eihxS )S does not converge ultraweakly, and that B(ℓ2) is not ultraweakly complete. S and T are finite dimensional linear subspaces of ℓ2 which contain the (Nevertheless, we'll see that every von Neumann algebra is ultrastrongly complete, and that every norm-bounded ultraweakly Cauchy net in a von Neumann converges, in 77 I.) 3.1.2 Elementary Theory The basic facts concerning von Neumann algebras we'll deal with first mostly involve the relationship between multiplication and the order structure. For example, while it is clear that translation and scaling on a von Neumann algebra are ultraweakly (and ultrastrongly) continuous, the fact that multiplication is ultraweakly (and ultrastrongly) continuous in each coordinate is less obvious (see 45 IV). Quite surprisingly, this problem reduces to the ultraweak continuity of b 7→ a∗ba by the following identity. Exercise Show that for elements a, b, c of a C∗-algebra, 44 II a∗ c b = 1 k=0 ik (ika + b)∗ c (ika + b). 4 P3 (Note that this identity is a variation on the polarisation identity for inner products, see 4 XV.) ..43 -- 44.. 83 III Lemma Let (xα)α∈D be a net of effects of a von Neumann algebra A , which converges ultraweakly to 0. Let (bα)α∈D be a net of elements with kbαk 6 1 for all α. Then (xαbα)α converges ultraweakly to 0. IV Proof Let ω : A → C be an np-map. We have, for each α, since xα > 0 by Kadison's inequality, 30 IV since xα 6 1 since b∗αbα 6 1. ω(xαbα)2 = ω(√xα √xα bα )2 6 ω(xα) ω( b∗αxαbα ) 6 ω(xα) ω(b∗αbα) 6 ω(xα) ω(1) Thus, since (ω(xα))α converges to 0, we see that (ω(xαbα))α converges to 0, and so (xαbα)α converges ultraweakly to 0. (cid:3) V VI Exercise Let D be a bounded directed set of self-adjoint elements of a von Neumann algebra A , and let a ∈ A . Show that the net (d)d∈D converges ultraweakly to W D. converges ultraweakly to a∗(W D). VII Use III to show that (da)d converges ultraweakly to (W D)a, and that (a∗d)d VIII Proposition Let a be an element of a von Neumann algebra A . Then for every bounded directed subset D of self-adjoint elements of A . Wd∈D a∗ d a = a∗ (W D) a IX Proof If a is invertible, then the (by 25 II) order preserving map b 7→ a∗ba has an order preserving inverse (namely b 7→ (a−1)∗ba−1), and therefore preserves all suprema. X The general case reduces to the case that a is invertible in the following way. There is (by 11 VI) λ > 0 such that λ + a is invertible. Then as d increases a∗ d a ≡ (λ + a)∗ d (λ + a) − λ2d − λa∗d − λda that the ultraweak topology is Hausdorff. At the moment, however, we must converges ultraweakly to a∗ (W D) a, because ( (λ + a)∗ d (λ + a) )d converges ultraweakly to (λ + a)∗ (W D) (λ + a) by IX and VI, and (a∗d + da)d converges ultraweakly to a∗(W D) + (W D)a by VII. Since (a∗da)d converges to Wd∈D a∗da too, we could conclude thatWd∈D a∗ d a = a∗ (W D) a if we would already know content ourselves with the conclusion that ω( a∗(W D)a − Wd∈D a∗da ) = 0 for every np-functional ω on A . But since a∗(W D)a −Wd∈D a∗da happens to be positive, we conclude that a∗(W D)a −Wd∈D a∗da = 0 nonetheless. XI Exercise Show that the set of np-functionals on a von Neumann algebra A is (cid:3) not only faithful but also order separating using 30 X. Deduce 1. that the ultraweak and ultrastrong topologies are Hausdorff, 2. that A+, AR and [0, 1]A are ultraweakly (and ultrastrongly) closed, 3. and that the unit ball (A )1 is ultrastrongly closed. (We'll see only later on, in 73 VIII, that (A )1 is ultraweakly closed as well.) Exercise Let D be a directed subset of self-adjoint elements of a von Neumann algebra A , and let a ∈ A . Show that if ad = da for all d ∈ D, then a(W D) = (W D)a. Use III to show that (W D − d)2 converges ultraweakly to 0 as D ∋ d → ∞. Conclude that (d)d∈D converges ultrastrongly to W D. Exercise Show that for a positive linear map f : A → B between von Neumann algebras, the following are equivalent. XII XIII XIV XV 1. f is ultraweakly continuous; 2. f is ultraweakly continuous on [0, 1]A ; 3. f (W D) =Wd∈D f (d) for each bounded directed D ⊆ AR; 4. ω ◦ f : A → C is normal for each np-map ω : B → C. In that case we say that f is normal. (Note that this definition of "normal" extends the one for positive functionals from 42 II.) Conclude that b 7→ a∗ba, A → A is ultraweakly continuous for every ele- ment a of a von Neumann algebra A . The converse does not hold: give an example of a map f which is normal, Exercise Show that if a positive linear map f : A → B between von Neumann algebras is ultrastrongly continuous (on [0, 1]A ), then f is normal. (Hint: use that a bounded directed set D ⊆ AR converges ultrastrongly to W D.) but not ultrastrongly continuous. (Hint: transpose.) Proposition An ncp-map f : A → B between von Neumann algebras is ultra- strongly continuous. Proof Note that f is ultrastrongly continuous at a ∈ A iff f ((· )+ a) ≡ f + f (a) is ultrastrongly continuous at 0. Thus to show that f is ultrastrongly continuous it suffices to show that f is ultrastrongly continuous at 0. So let (bα)α be a net in A which converges ultrastrongly to 0; we must show that (f (bα))α converges ultrastrongly to 0, viz. that ( f (bα)∗f (bα) )α converges ultraweakly to 0. Since f (bα)∗f (bα) 6 f (b∗αbα)kf (1)k by 34 XIV, it suffices to show that ( f (b∗αbα) )α converges ultraweakly to 0, but this follows from the facts that f is ultraweakly continuous and (b∗αbα)α converges ultraweakly to 0 (since (bα)α converges ultra- strongly to 0). (cid:3) 45 II III ..44 -- 45.. 85 IV Exercise Let A be a von Neumann algebra. Conclude (using II and 34 V) that the map a 7→ b∗ab, A → A is ultrastrongly continuous for every element b ∈ A . Use this, and 44 II, to show that b 7→ ab, ba : A → A are ultraweakly and ultrastrongly continuous for every element a of a von Neumann algebra A . V We saw in 43 II that the multiplication on a von Neumann algebra is not jointly ultraweakly continuous, even on a bounded set. Neither is a, b 7→ ab jointly ultrastrongly continuous, even when b is restricted to a bounded set; but it is jointly ultrastrongly continuous when a is restricted to a bounded set: VI Proposition Let (aα)α and (bα)α be nets in a von Neumann algebra A with the same index set that converge ultrastrongly to a, b ∈ A , respectively. Then the net (aαbα)α converges ultrastrongly to ab provided that (aα)α is bounded. VII Proof Let ω : A → C be an np-functional. Since kab − aαbαkω 6 k(a − aα)bkω + kaα(b − bα)kω 6 ka − aαkω(b∗( · )b) + kaαkkb − bαkω vanishes as α → ∞, we see that (aαbα)α converges ultrastrongly to ab. (cid:3) 46 We can now prove a bit more about the ultrastrong and ultraweak topologies. II III Exercise Show that a net (bα)α in a von Neumann algebra A converges ultra- strongly to an element b of A if and only if both b∗αbα −→ b∗b and bα −→ b ultraweakly as α → ∞. Exercise Show that for a positive linear map ω : A → C on a von Neumann algebra A the following are equivalent 1. ω is normal; 2. ω is ultraweakly continuous; 3. ω is ultrastrongly continuous. (Hint: combine 44 XV and 45 II.) 47 Enter the eponymous hero(s) of this thesis. II Definition We denote the category of normal cpsu-maps by W∗cpsu, and its subcategory of nmiu-maps by W∗miu. (We omit the "N" for the sake of brevity.) III Though arguably W∗miu is a good candidate for being called the category of von Neumann algebra, the title of this thesis refers to W∗cpsu. Indeed, it's the ncpsu-maps between von Neumann algebras that stand to model the arbitrary quantum processes, and it's the category of these quantum processes we want to mine for abstract structure. This is mostly a task for the next chapter, though. For now we'll just establish that W∗cpsu has all products, IV, certain equalisers, V, and that (W∗cpsu)op is an effectus, see VI. IV V VI Exercise Show that given a family (Ai)i of von Neumann algebras the direct sum Ai → Aj Ai into the product of the Ai Ai from 3 V is a von Neumann algebra and the projections πj : Li Li are normal. Moreover, show that this makes Li in the categories W∗miu and W∗cpsu (see 20a I and 34 VI). Exercise Show that given nmiu-maps f, g : A → B between von Neumann al- gebras the set E := { a ∈ A : f (a) = g(a)} is a von Neumann subalgebra of A , and the inclusion e : E → A is the equaliser of f and g in the categories W∗miu and W∗cpsu (see 20a II and 34 VI). Let us briefly indicate what makes (W∗cpsu)op an effectus; for a precise for- mulation and proof of this fact we refer to [4, 9] (or 180 V, 180 VII, and 180 X ahead). Note that the sum f + g of two ncpsu-maps f, g : A → B between von Neumann algebras is again an ncpsu-map iff f (1) + g(1) 6 1. The partial addition on ncpsu-maps thereby defined has, aside from some fairly obvious properties (summarised by the fact that the category W∗cpsu is PCM-enriched, see [4]), the following special trait: given ncpsu-maps f : A → D and g : B → D with f (1) + g(1) 6 1 we may form an ncpsu-map [f, g] : A × B → D by [f, g](a, b) = f (a) + g(b), and, moreover, every ncpsu-map A × B → D is of this form. This observation, which gives the product of W∗cpsu a coproduct-like quality without forcing it to be a biproduct (which it's not), makes (W∗cpsu)op a FinPAC (see 180 VII). For (W∗cpsu)op to be an effectus, we need a second ingredient: the complex numbers, C. Since the ncpsu-maps p : C → A are all of the form λ 7→ λa for some effect a ∈ [0, 1]A , the ncpsu-maps p : C → A (called predicates in this context) are not only endowed with a partial addition, but even form an effect algebra. This, combined with the observation that an ncpsu-map f : A → B is constant zero iff f (1) = 0, makes (W∗cpsu)op an effectus in partial form (see 180 VII). As you can see, there's nothing deep underlying (W∗cpsu)op being an effectus. In that respect effectus theory resembles topology: just as a topology provides a basis for notions such as compactness, connectedness, meagerness, and homo- topy, so does an effectus provide a framework to study aspects of computation such as side effects (223 II) and purity (173 VII). Let us quickly prove that every von Neumann algebra is isomorphic to a von Neumann algebra of operators on a Hilbert space (see VIII). 48 Exercise Let Ω be a collection of np-functionals on a von Neumann algebra B that is faithful (see 21 II). Show that a positive linear map f : A → B is normal iff ω ◦ f is normal for all ω ∈ Ω. Proposition Given an np-map ω : A → C on a von Neumann algebra A , the map ω : A → B(Hω) from 30 VI is normal. Proof Since by definition of Hω the vectors of the form ηω(a) where a ∈ A are II III IV ..45 -- 48.. 87 dense in Hω, the vector functionals hηω(a), (· )ηω(a)i form a faithful collection of np-functionals on B(Hω). Thus by II it suffices to show given a ∈ A that hηω(a), ω(· )ηω(a)i ≡ ω(a∗(· )a) is normal, which it is, by 44 VIII. (cid:3) Exercise Show that the map Ω from 30 IX is normal for every collection Ω of np-maps A → C on a von Neumann algebra A . Lemma Let f : A → B be an injective nmiu-map between von Neumann algebras. Then the image f (A ) is a von Neumann subalgebra of B, and f restricts to an nmiu-isomorphism from A to f (A ). V VI VII Proof We already know by 29 IX that f (A ) is a C∗-subalgebra of A , and that f restricts to an miu-isomorphism f′ : A → f (A ). The only thing left to show is that f (A ) is a von Neumann subalgebra of B, because an miu- isomorphism between von Neumann algebras (being an order isomorphism) will automatically be an nmiu-isomorphism. Let D be a bounded directed subset of f (A ). Note that S := (f′)−1(D) is a bounded directed subset of A , and so W D ≡ W f ( S ) = f (W S), because f is normal. Thus W f (D) ∈ f (A ), and so f (A ) is a von Neumann subalgebra of B. (cid:3) VIII Theorem (normal Gelfand -- Naimark) Every von Neumann algebra A is nmiu- isomorphic to von Neumann algebra of operators on a Hilbert space. IX Proof Recall that an element a ∈ A is zero iff ω(a) = 0 for all np-maps ω : A → C. It follows that the collection Ω of all np-maps A → C obeys the condition of 30 X, and so the miu-map Ω : A → B(HΩ) (from 30 IX) is injective. Since Ω is also normal by V, we see by VI that Ω restricts to an nmiu-isomorphism from A to the von Neumann subalgebra Ω(A ) of B(HΩ). (cid:3) 3.1.3 Examples Matrices over von Neumann algebras 49 We'll show that the C∗-algebra of N × N -matrices MN (A ) over a von Neumann algebra A is itself a von Neumann algebra, and to this end, we prove something a bit more more general. III II Theorem Given a von Neumann algebra A , the C∗-algebra Ba(X) (32 XIII) of bounded adjointable module maps on a self-dual (36 I) Hilbert A -module X is a von Neumann algebra, and hx, (· )xi : Ba(X) → A is normal for every x ∈ X. Proof We'll first show that a bounded directed subset D of Ba(X)R has a supremum (in Ba(X)R). To obtain a candidate for this supremum, we first de- fine a bounded form [· , · ] : X × X → A in the sense of 36 IV and apply 36 V. To this end note that given x ∈ X the subset { hx, T xi : T ∈ D } of AR is bounded and directed, and so (since A is a von Neumann algebra) has a supremum. 4P3 Since the net (hx, T xi )T∈D converges ultraweakly to this supremum by 44 VI, we see that hy, T zi = 1 k=0 ik(cid:10)y + ikz, T (y + ikz)(cid:11) converges ultraweakly to some element [y, z] of A as T → ∞ for all y, z ∈ X, giving us a form [· , · ] on X. Since khy, T zik 6 supT ′∈D kT ′kkykkzk for all T ∈ D by 32 X, and thus k[y, z]k 6 supT ′∈D kT ′kkykkzk for all y, z ∈ X, we see that the form [· , · ] is bounded. Since X is self dual, there is by 36 V S ∈ Ba(X) with [y, z] = hy, Szi for all y, z ∈ X; we'll show that S is the supremum of D. To begin, given T ∈ D we have hx, T xi 6 WT ′∈D hx, T ′xi = [x, x] = hx, Sxi for all x ∈ X, and so T 6 S by 32 XV, that is, S is an upper bound for D. Given another upper bound S′ ∈ Ba(X)R of D (so T 6 S′ for all T ∈ D) we have hx, T xi 6 hx, S′xi and so hx, Sxi = [x, x] = WT∈D hx, T xi 6 hx, S′xi for all x ∈ X implying that S 6 S′. Hence S is the supremum of D in Ba(X)R. Note that since hx, Sxi = WT∈D hx, T xi we immediately see that hx, (· )xi : Ba(X) → A preserves bounded directed suprema for every x ∈ X. It remains to be shown that there are sufficiently many np-functionals on Ba(X) in the sense that T ∈ (Ba(X))+ is zero when ω(T ) = 0 for every np- functional ω : Ba(X) → C. This is indeed the case for such an operator T , because ξ(hx, (· )xi) is an np-functional on Ba(X) for every x ∈ X and an np-functional ξ : A → C, implying that ξ(hx, T xi) = 0, and hx, T xi = 0, and so T = 0. (cid:3) Exercise Let A be a von Neumann algebra, and let N be a natural number. IV a von Neumann algebra. 1. Show that the C∗-algebra MN (A ) of N × N -matrices over A (see 33 I) is 2. Show that the map A 7→ Pij a∗i Aij aj : MN A → A is normal and com- pletely positive, and that the map A 7→ Pij a∗i Aijbj : MN A → A is ul- trastrongly and ultraweakly continuous for all a1, . . . , aN , b1, . . . , bN ∈ A . In particular, A 7→ Aij : MN A → A is ultraweakly and ultrastrongly continuous for all i, j. Show that a net (Aα)α in MN A converges ultraweakly (ultrastrongly) to B ∈ MN A iff (Aα)ij converges ultraweakly (ultrastrongly) to Bij as α → ∞ for all i, j. 3. Given an ncp-map f : A → B between von Neumann algebras, show that the cp-map MN f : MN A → MN B from 33 III is normal. Commutative von Neumann algebras Another important source of examples of von Neumann algebras is measure theory: we'll show that the bounded measurable functions on a finite complete 50 ..48 -- 50.. 89 measure space X form a commutative von Neumann algebra L∞(X) when func- tions that are equal almost everywhere are identified (see 51 IX). In fact, we'll see in 70 III that every commutative von Neumann algebra is nmiu-isomorphic to a direct sum of L∞(X)s. This is not only interesting in its own right, but will also be used in the next chapter to show that the only von Neumann algebras that can be endowed with a 'duplicator' are of the form ℓ∞(X) for some set X (see 127 III). We should probably mention that L∞(X) can be defined for any measure space X, and is a von Neumann algebra precisely when X is localisable, see [69]. This has the advantage that any commutative von Neumann algebra is nmiu- isomorphic to a single L∞(X) for some localisable measure space X, but since it has no other advantages relevant to this text we restrict ourselves to complete finite measure spaces. We'll assume the reader is reasonably familiar with the basics of measure theory, and we'll only show a selection of results that we deemed important. For the other details, we refer to volumes 1 and 2 of [19]. Nevertheless, we'll recall some basic definitions to fix terminology, which is sometimes simpler than in [19] (because we're dealing with finite complete measure spaces), and sometimes modified to the complex-valued case (c.f. 133C of [19]). A motivated reader will have no problem adapting the results from [19] to our setting. 51 Let X be a finite and complete measure space. We'll denote the σ-algebra of measurable subsets of X by ΣX , and the measure by µX : ΣX → [0,∞) (or µ when no confusion is expected). That X is finite means that µ(X) < ∞ (which doesn't mean that the set X is finite), and that X is complete means that every subset A of a negligible subset B of X is itself negligible. (Recall that N ⊆ X is negligible when N ∈ ΣX and µ(N ) = 0.) A function f : X → C is measurable when the inverse image f−1(U ) of any open subset U of C is measurable (which happens precisely when both x 7→ f (x)R, x 7→ f (x)I : X → R are measurable in the sense of 121C of [19]). An important example of a measurable function on X is the indicator function 1A of a measurable subset A of X (which is equal to 1 on A and 0 elsewhere.) II The bounded measurable functions f : X → C form a C∗-subalgebra of CX that we'll denote by L∞(X). The space L∞(X) is not only closed with respect to the (supremum) norm on CX , but also with respect to coordinatewise lim- its of bounded sequences (c.f. 121F of [19]). As a result, the coordinatewise (countable) supremum Wn fn of a bounded ascending sequence f1 6 f2 6 ··· in L∞(X)R is again in L∞(X), and is fact the supremum of (fn)n in L∞(X). However L∞(X) might still not be a von Neumann algebra because not ev- ery bounded directed subset of L∞(X)R might have a supremum as we'll show presently; this is why we'll move from L∞(X) to L∞(X) in a moment. For a counterexample to L∞(X) being always a von Neumann algebra we III take X to be the unit interval [0, 1] with the Lebesgue measure. Let A be a non-measurable subset of [0, 1] (see 134B of [19]). The indicator functions 1F where F is a finite subset of A form a bounded directed subset D of L∞([0, 1])R that -- so we claim -- has no supremum. Indeed, note that since f ∈ L∞([0, 1])R is an upper bound for D iff 1A 6 f , the least upper bound h for D would be the least bounded measurable function above 1A. Surely, h 6= 1A for such h (because otherwise A would be measurable), so h(x) > 1A(x) for some x ∈ [0, 1]. But then h− (h(x)− 1A(x))1{x} < h is an upper bound for D too contradicting the minimality of h. Whence L∞([0, 1]) is not a von Neumann algebra. To deal with L∞(X) we need to know a bit more about L∞(X), namely that the measure on X can be extended to a an integral R : L∞(X) → C (see 122M of [19])‡ with the following properties. 1. R (1A) = µ(A) for every measurable subset A of X. 2. R : L∞(X) → C is a positive linear map (see 122O of [19]). 3. R Wn fn =WnR fn for every bounded sequence f1 6 f2 6 ··· in L∞(X)R. that do exist in L∞(X): for example, the set D := { f ∈ [0, 1]L∞(X) : R f = 0 } is directed, bounded, and has supremum 1, but Wf∈DR f = 0 < 1 = R W D. What is surprising is that the lifting of R to L∞(X) will be normal. But let us first define L∞(X). We say that f, g ∈ L∞(X) are equal almost everywhere and write f ≈ g when f (x) = g(x) for almost all x ∈ X (that is, { x ∈ X : f (x) 6= g(x)} is negligible). It is easily seen that ≈ is an equivalence relation; we denote the equivalence class of a function f ∈ L∞(X) by f◦, and the set of equivalence classes by L∞(X) := { f◦ : f ∈ L∞(X)}, which becomes a commutative C∗-algebra when endowed with the same operations as L∞(X), but with a slightly modified norm given by, for f ≡ f◦ ∈ L∞(X), (This is a special case of Levi's theorem, see 123A of [19].) Unsurprisingly, the integral interacts poorly with the uncountable directed suprema IV V kfk = min{ kgk : g ∈ L∞(X) and g◦ = f} = min{ λ > 0 : f (x) 6 λ for almost all x ∈ X }. This is called the essential supremum norm. To see that L∞(X) is complete one can use the fact that L∞(X) is complete in a slightly more general sense than discussed before: when a bounded sequence f1, f2, . . . in L∞(X) converges coordinatewise for almost all x ∈ X to some bounded function f : X → C, this function f is itself measurable (and so f ∈ L∞(X), c.f. 121F of [19]). Another consequence of this is that a bounded ascending sequence f◦1 6 f◦2 6 ··· in L∞(X) (so f1, f2, . . . ∈ L∞(X), and f1(x) 6 f2(x) 6 ··· for almost ‡Note that every element of L∞(X) being bounded is integrable by 122P of [19]. ..50 -- 51.. 91 all x ∈ X) has a supremumWn f◦n in L∞(X). Indeed, we'll haveWn f◦n = g◦ for any bounded map g : X → C with g(x) =Wn fn(x) for almost all x ∈ X. VI Now, let us return to the integral. Since R f = R g for all f, g ∈ L∞(X) with f ≈ g we get a map R : L∞(X) → C given by R f◦ =R f . Clearly, R is positive we see thatR Wn fn =WnR fn for any bounded ascending sequence f1 6 f2 6 ··· in L∞(X)R. Note thatR : L∞(X) → C is also faithful, because ifR f◦ =R f = 0 for some f ∈ L∞(X)+, then f (x) = 0 for almost all x ∈ X, and so f◦ = 0. Now, the fact that L∞(X) is a von Neumann algebra follows from the following general and rather surprising observation. and linear, and by (a slightly less special case of) Levi's theorem (123A of [19]) in D with b 6 b1 and an 6 bn Since τ is faithful and normal, A is a von Neumann algebra. (cid:3) VIII Proof Our first task is to show that a bounded directed subset D of self-adjoint VII Proposition Let A be a C∗-algebra, and let τ : A → C be a faithful positive map. If every bounded ascending sequence a1 6 a2 6 ··· of self-adjoint elements from A has a supremumWn an (in AR) and τ (Wn an) =Wn τ (an), then A is a von Neumann algebra, and τ is normal. elements of A has a supremum W D in AR. Since Wd∈D τ (d) is a supremum in R we can find a1 6 a2 6 ··· in D with Wn τ (an) = Wd∈D τ (d). We'll show that Wn an is the supremum of D. Surely, any upper bound of D being also is above Wn an, so the only thing that we an upper bound for a1 6 a2 6 ··· need to show is that Wn an is an upper bound of D. So let b ∈ D be given. The trick is to pick a sequence b1 6 b2 6 ··· for all n (which exists on account of D's directedness). Then Wn an 6 Wn bn, and Wd∈D τ (d) = Wn τ (an) = τ (Wn an ) 6 τ (Wn bn ) = Wn τ (bn) 6 Wd∈D τ (d), so τ (Wn an ) = τ (Wn bn ), which implies that Wn an = Wn bn as τ is faithful. Since then b 6 b1 6Wn bn = Wn an we see that Wn an is an upper bound (and thus the supremum) of D. Moreover, since Wd∈D τ (d) 6 τ (W D) = τ (Wn an) = Wn τ (an) 6 Wd∈D τ (d), we see that Wd∈D τ (d) = τ (W D), and so τ is normal. commutative von Neumann algebra, and the assignment f 7→R f gives a faithful normal positive map R : L∞(X) → C. 52 We'll show that any commutative von Neumann algebra A that admits a faithful np-functional ω : A → C is nmiu-isomorphic to L∞(X) for some finite complete measure space X. It makes sense to regard this result as a von Neumann algebra analogue of Gelfand's theorem for commutative C∗-algebras, (see 27 XXVII -- that any commutative C∗-algebra is miu-isomorphic to C(Y ) for some compact Hausdorff space Y .) But one should not take the comparison too far too lightly: while Gelfand's theorem readily yields a clean equivalence between commutative C∗-algebras and compact Hausdorff spaces (see 29), the fact that L∞(X1) ∼= L∞(X2) for finite complete measure spaces X1 and X2 does not even imply IX Corollary Given a finite complete measure space X the C∗-algebra L∞(X) is a that X1 and X2 have the same cardinality.§ Obtaining an equivalence between commutative von Neumann algebras and measure spaces is nonetheless possible after a suitable non-trivial modification to the category of measure spaces (as is shown by Robert Furber in as of yet unpublished work.) We obtain our finite complete measure space X from the commutative von Neumann algebra A by taking for X the compact Hausdorff space sp(A ) of all miu-functionals on A , and declaring that a subset A of X ≡ sp(A ) is measurable when A is clopen up to a meagre subset (defined below, II). It takes some effort to show that this yields a σ-algebra in sp(A ), and that the faithful np-functional ω : A → C gives a finite complete measure on sp(A ), but once this is achieved it's easily seen that A ∼= C(sp(A )) ∼= L∞(sp(A )). Definition Let X be a topological space. 1. A subset A of X is called meagre when A ⊆Sn Bn for some closed subsets B1 ⊆ B2 ⊆ ··· of X with empty interior (so B◦n = ∅ for all n.) 2. Given A, B ⊆ X we write A ≈ B when A ∪ B \ A ∩ B is meagre. 3. We say that A ⊆ B is almost clopen when A ≈ C for some clopen C ⊆ X. Exercise Given a topological space X, verify the following facts. 2. A subset of a meagre set is meagre. 1. A countable union Sn An of meagre subsets A1, A2, . . . ⊆ X is meagre. 3. U ≈ U for every open subset U of X. (Hint: show that U\U is closed with empty interior.) 4. Sn An ≈Sn Bn for all A1, A2, . . . , B1, B2, . . . ⊆ X with An ≈ Bn. 5. A\B ≈ A′\B′ for all A, A′, B, B′ ⊆ X with A ≈ A′ and B ≈ B′. 6. If A, B ⊆ X are almost clopen, then A ∪ B and A\B are almost clopen. Meagerness can be thought of as a topological analogue of negligibility. In fact, with respect to the measure we'll put on sp(A ) in 54 XI, meagerness and negligibility actually coincide. In general, however, the notions are disparate, as is demonstrated rather dramatically by the following example. §Indeed, one may take X1 to be a measure space consisting of a single non-negligible point ∗ (so X1 = {∗} and µ(X1) 6= 0), while letting X2 be a measure space on an uncountable set formed by taking for the measurable subsets of X2 the countable subsets and their comple- ments, by making the countable subsets negligible, and by giving all cocountable subsets the same non-zero measure. Then all measurable functions on X1 and on X2 are constant al- most everywhere, (because in X1 and X2 there are no two non-negligible disjoint measurable subsets,) so that L∞(X1) ∼= C ∼= L∞(X2). ..51 -- 52.. 93 II III IIIa IIIb Example of a meagre subset A of [0, 1] with Lebesgue measure 1. Given an enumeration q1, q2, . . . of the rational numbers in [0, 1], the set Bm := Sn(cid:0) qn − 1 2m 2−n, qn + 1 2m 2−n(cid:1) ∩ [0, 1] is open and dense in [0, 1], with a Lebesgue measure of at most 1/m. So the 53 II intersection B :=Tm Bm is Lebesgue negligible. On the other hand, [0, 1]\Bm is closed and has empty interior, so that A := [0, 1]\B ≡Sm Bm is meagre with Lebesgue measure 1. The fact that the almost clopen subsets of the spectrum sp(A ) of a commutative von Neumann algebra A are closed under countable unions (and thus form a σ-algebra) relies on a special topological property of sp(A ) that is described in III below. Exercise Let A be a commutative von Neumann algebra. Using the fact that the Gelfand representation γA : A → C(sp(A )) from 27 III is an miu- isomorphism by 27 XXVII and thus an order isomorphism, show that C(sp(A )) is a commutative von Neumann algebra that is nmiu-isomorphic to A via γA . III Proposition The spectrum sp(A ) of a commutative von Neumann algebra A is extremally disconnected: the closure U of an open subset U of sp(A ) is open. IV Proof (Based on §6.1 of [77].) that 1U is continuous, so that U is both open and closed. Let U be an open subset of sp(A ), and let 1U be the indicator function of U . The set D = { f ∈ C(sp(A )) : f 6 1U } is directed and bounded and so has a supremum W D in C(sp(A )) since C(sp(A )) is a von Neumann algebra by II. Note that 0 6W D 6 1. We'll prove that W D = 1U , because this entails Let x ∈ U be given. By Urysohn's lemma (see 15.6 of [87], using here that sp(A ) being a compact Hausdorff space, 27 XXV, is normal by 17.10 of [87]) there is f ∈ [0, 1]C(sp(A )) with f (x) = 1 and f (y) = 0 for all y ∈ sp(X)\U . It follows that f ∈ D, and f 6 W D 6 1, so that 1 = f (x) 6 (W D)(x) 6 1, and (W D)(x) = 1. By continuity of W D, we get (W D)(x) = 1 for all x ∈ U . Let y ∈ sp(A )\U be given. Again by Urysohn's lemma there is f ∈ [0, 1]C(sp(A )) with f (y) = 0 and f (x) = 1 for all x ∈ U . Since g 6 1U 6 f for every g ∈ D, we get W D 6 f , and so 0 6 (W D)(y) 6 f (y) = 0, which implies that (W D)(y) = 0. Hence (W D)(y) = 0 for all y ∈ sp(A )\U . All in all we have W D = 1U , and so U is open. (cid:3) V Corollary The almost clopen subsets of an extremally disconnected topological space X form a σ-algebra. VI Proof of almost clopen subsets A1, A2, . . . In light of 52 III it remains only to be shown that the union Sn An is almost clopen. Let C1, C2, . . . ⊆ X be clopen with An ≈ Cn for each n. Then Sn An ≈ Sn Cn, and C := Sn Cn is open (but not necessarily closed). Since C ≈ C (by 52 III), and C is clopen (as X is extremally disconnected) we get Sn An ≈ C, so Sn An is almost clopen.(cid:3) The final ingredient we need to prove the main result, XI, of this section is the observation that an almost clopen subset of a compact Hausdorff space is equivalent to precisely one clopen, which follows from the following famous theorem. Baire category theorem A meagre subset of a compact Hausdorff space has empty interior. Proof Let A be a meagre subset of a compact Hausdorff space X. So there are closed B1 ⊆ B2 ⊆ . . . with A ⊆Sn Bn and B◦n = ∅ for all n. Then Un := X\Bn is an open dense subset of X for each n. Since A◦ ⊆ (Sn Bn)◦ = X\(Tn Un ) it suffices to show that Tn Un is dense in X. That is, given a non-empty open subset V of X we must show that V ∩Tn Un 6= ∅. Write V1 := V . Since U1 is open and dense, and V1 is open and not empty, we have U1 ∩ V1 6= ∅. Since X is regular (see e.g. [87]) we can find an open and non-empty subset V2 of X with V 2 ⊆ U1 ∩ V1. Continuing this process we obtain non-empty open subsets V ≡ V1 ⊇ V2 ⊇ ··· of X with V n+1 ⊆ Un ∩ Vn for all n, and so V 1 ⊇ V1 ⊇ V 2 ⊇ V2 ⊇ ··· . Since X is compact, Tn V n can not be empty, and neither will be V ∩Tn Un ⊇Tn V n. (cid:3) Lemma For open subsets U and V of a compact Hausdorff space X, U ≈ V ⇐⇒ U ≈ V ⇐⇒ U = V . Proof As U ≈ U by 52 III the only thing that is not obvious is that U ≈ V =⇒ U = V . So suppose that U ≈ V . Then U\V is empty, because it is an open subset of the meagre set U ∪ V \ U ∩ V (which has empty interior by II.) In other words, we have U ⊆ V , and thus U ⊆ V . Similarly, V ⊆ U , and so V = U . (cid:3) Corollary Given an almost clopen subset A of a compact Hausdorff space X there is precisely one clopen C with A ≈ C. Proof When C ≈ A ≈ C′ for clopen subsets C, C′ ⊆ X, we have C ≈ C′, and so C = C′ by IV. (cid:3) Interestingly, a compact Hausdorff space is extremally disconnected iff each of its open subsets is "measurable" in the sense of being almost clopen: If X is extremally disconnected, and U is open subset of X, then U is Proposition A compact Hausdorff space X is extremally disconnected iff every open subset of X is almost clopen. Proof clopen, and U ≈ U by 52 III giving us that U is almost clopen. Conversely, suppose that each open subset of X is almost clopen. To show that X is extremally disconnected we must show that U is open given an open 54 II III IV V VI VII VIII IX X ..52 -- 54.. 95 subset U of X. Pick a clopen C with U ≈ C. Then U ≈ U ≈ C (by 52 III), and so U = C by IV. (cid:3) XI Theorem Let A be a commutative von Neumann algebra A . Recall that the Gelfand representation γA : A → C(sp(A )) is an nmiu-isomorphism (by 53 II), C(sp(A )) is a von Neumann algebra, and that the almost clopen subsets (see 52 II) of sp(A ) form a σ-algebra. Given a faithful np-functional ω : A → C there is a (unique) measure µ on the almost clopen subsets of sp(A ) such that µ(A) = 0 iff A is mea- gre, and µ(C) = ω(γ−1 A (1C)) for every clopen subset C of sp(A ); and this turns sp(A ) into a finite complete measure space. With respect to this measure space a bounded function f : sp(A ) → C is measurable iff f is continuous almost everywhere. Moreover, f 7→ f◦ : C(sp(A )) → L∞(sp(A )) is an nmiu-isomorphism, and R f◦ = ω(γ−1 A (f )) for all f ∈ C(X). All in all, we get the following commuting diagram. A C(sp(A )) / L∞(sp(A )) γA ∼= $❍❍❍❍❍❍❍❍❍❍❍❍ ω f7→f ◦ ∼= x♣♣♣♣♣♣♣♣♣♣♣♣♣♣ R C XII Proof By VII we know that given an almost clopen subset A of sp(A ) there is a unique clopen CA with A ≈ CA, and so we may define µ(A) := ω(γ−1 A (1CA)). It is easily seen that µ is finitely additive. Further µ(A) = 0 for every meagre A ⊆ X, and so µ(A) = µ(B) when A ≈ B. Conversely, an almost clopen subset A of A with µ(A) = 0 is meagre, because for the unique clopen C with A ≈ C, we have ω(γ−1 A (1C )) = µ(A) = 0, so that 1C = 0 and thus C = ∅ -- using here that ω is faithful. we must show that f 6 0. Note that for such f we have f (x) 6 0 for To show that µ is a measure, it suffices to prove thatVn µ(An) = 0 given A1 ⊇ A2 ⊇ ··· with Tn An = ∅. To do this, pick clopen subsets C1, C2, . . . of sp(A ) with An ≈ Cn for all n. Then Vn µ(An) = Vn µ(Cn) = ω(γ−1 A (Vn 1Cn)) -- using here that ω is normal. So to prove that Vn µ(An) = 0 it suffices to show that Vn 1Cn = 0, that is, given a lower bound f of the 1Cn in C(sp(A ))R all x ∈ X\Tn Cn. Then f (x) 6 0 for all x ∈ X if we can show that X\Tn Cn is dense in X. But this indeed the case since Tn Cn ≈ Tn An = ∅ is meagre, and therefore has empty interior (by II). Whence µ is a measure. Note that µ is finite, because µ(sp(A )) = ω(1) < ∞, and complete, because a subset of a meagre set is meagre. Let h : sp(A ) → C be a bounded function. We'll show that h is continuous almost everywhere iff h is measurable. Surely, if h is continuous (everywhere), then h is measurable (since every open subset U of sp(A ) is almost clopen, / / $ / x IX). So if h is continuous almost everywhere, then h is measurable too. For the converse, it suffices to show that : h 7→ h◦ : C(sp(A )) → L∞(sp(A )) is surjective. To this end, note first that is injective, because a continuous function on sp(A ) that is zero almost everywhere, is non-zero on a meagre set, and by II zero on a dense subset, and so is zero everywhere. Since the image of the injective miu-map is norm closed in order to show that is surjective it suffices to show that image of is norm dense in L∞(X). This is indeed the case since the elements of L∞(sp(A )) of the form Pn λn1◦An where λ1, . . . , λN ∈ C and A1, . . . , AN are measurable (i.e. almost clopen) subsets of sp(A ) are easily seen to be norm dense in L∞(sp(A )) (c.f. 243I of [19]), and are in the range of , because given an almost clopen A ⊆ sp(A ) and a clopen C with A ≈ C we have 1◦A = 1◦C and 1C ∈ C(sp(A )). Hence is surjective. It remains to be show that R f◦ = ω(γ−1 A (f )) for all f ∈ C(sp(A )), that is, R = ω ◦ γ−1 A ◦ −1. By the previous discussion the linear span of the elements of L∞(sp(A )) of the form 1◦C , where C is (not just measurable but) clopen, is norm dense in L∞(sp(A )). Since R 1C = µ(C) = ω(γ−1 A (−1(1◦C )) for all clopen C, and both R and ω ◦ γ−1 A ◦ −1 are linear and bounded, we conclude that R = ω ◦ γ−1 (cid:3) To deduce from this that all commutative von Neumann algebras (and not just the ones admitting a faithful np-functional) are nmiu-isomorphic to direct sums of the form Li L∞(Xi) where the Xi are finite complete measure spaces we first need some basic facts concerning the projections of a commutative von Neumann algebra. A ◦ −1, and so we are done. 3.2 Projections One pertinent feature of von Neumann algebras is an abundance of projections: above each effect a there is a least projection ⌈a⌉ we call the ceiling of a (56 I); for every np-map ω : A → B between von Neumann algebras there is a least projection p with ω(p⊥) = 0 called the carrier of ω (see 63 I); the directed supremum of projections is again a projection; the partial order of projections is complete (see 56 XIII); and each element of a von Neumann algebra is the norm limit of linear combinations of projections (see 65 IV). We'll prove all this and more in this section. Definition An element p of a C∗-algebra is a projection when p∗p = p. Examples 1. The only projections in C are 0 and 1. 2. Given a measurable subset A of a finite complete measure space X the indicator function 1A is a projection in L∞(X), and every projection ..54 -- 55.. 97 XIII 55 II III in L∞(X) is of this form. 3. Given a closed linear subspace C of a Hilbert space H the inclusion E : C → H is a bounded linear map, and PC := EE∗ : H → H is a projection in B(H ), and every projection in B(H ) is of this form. IV Exercise Show that in a C∗-algebra: 1. 0 and 1 are projections. 2. A projection p is an effect, that is, p∗ = p and 0 6 p 6 1. 3. The orthocomplement p⊥ ≡ 1 − p of a projection p is a projection. 4. An effect a is a projection iff aa⊥ = 0. V Lemma Let a be an element of a C∗-algebra A with kak 6 1, and let p and q be projections on A . Then a∗pa 6 q⊥ iff paq = 0 iff aqa∗ 6 p⊥. VI Proof Suppose that a∗pa 6 q⊥. Then we have qa∗paq 6 qq⊥q = 0 (see 25 II) and so paq = 0, because kpaqk2 = k(paq)∗paqk = 0 by the C∗-identity. Applying (· )∗, we get qa∗p = 0, and so both qa∗ = qa∗p⊥ and aq = p⊥aq, giving us aqa∗ = p⊥aqa∗p⊥ 6 p⊥, where we used that aqa∗ 6 aa∗ 6 kaa∗k = kak2 6 1. By a similar reasoning, we get aqa∗ 6 p⊥ =⇒ paq = 0 =⇒ a∗pa 6 q⊥. (cid:3) VII Exercise Let a be an effect of a C∗-algebra A , and p be a projection from A . VIII Show that a 6 p iff p√a = √a iff √ap = √a iff p⊥√a = 0 iff √ap⊥ = 0 iff a2 6 p iff pa = a iff ap = a iff p⊥a = 0 iff ap⊥ = 0 iff √a 6 p. Show that p 6 a iff p√a = p iff √ap = p iff p√a⊥ = 0 iff √a⊥p = 0 iff p 6 a2 iff ap = p iff pa = p iff pa⊥ = 0 iff a⊥p = 0 iff p 6 √a. Lemma An effect a of a C∗-algebra A is a projection iff the only effect below a and a⊥ is 0. IX X XI Proof On the one hand, if a is a projection, and b is an effect with b 6 a and b 6 a⊥, then a⊥b = 0 and ab = 0 by VIII, and so b = ab + a⊥b = 0. On the other hand, if 0 is the only effect below both a and a⊥, then aa⊥ ≡ √aa⊥√a being an effect below a, and below a⊥, is zero, and so a is projection, by IV. (cid:3) XII Definition We say that projections p and q from a C∗-algebra A are orthogonal when pq = 0, and we say that a subset of projections from A is orthogonal (and its elements are pairwise orthogonal) when all p and q from E are either equal or orthogonal. XIII Exercise Let A be a C∗-algebra. 1. Show that projections p and q from A are orthogonal iff pq = 0 iff qp = 0 iff pqp = 0 iff p + q 6 1 iff p 6 q⊥ iff p + q is a projection. 2. Show that a finite set of projections p1, . . . , pn from A is orthogonal iff Pi pi 6 1 iff Pi pi is a projection. Show that, in that case, Pi pi is the least projection above p1, . . . , pn. Exercise Let p and q be projections from a C∗-algebra with p 6 q. Show that q − p is a projection (either directly, or using XIII). XIV 3.2.1 Ceiling and Floor Proposition Above every effect b of a von Neumann algebra A , there is a 56 smallest projection, ⌈b⌉, which we call the ceiling of b, given by ⌈b⌉ =W∞n=0 b1/2n . Moreover, if a ∈ A commutes with b, then a commutes with ⌈b⌉. Proof Since 0 6 b 6 b1/2 6 b1/4 6 ··· 6 1, we may define p :=Wn b1/2n . To begin, note that if a ∈ A commutes with b, then a commutes with p. Indeed, for such a we have a√b = √ba by 23 VII, and so ab1/2n = b1/2n a for each n by induction. Thus ap = pa by 44 XIII. Let us prove that p is a projection, i.e. p2 = p. Since p 6 1, we already have p2 ≡ √pp√p 6 p by 25 II, and so we only need to show that p 6 p2. We have: p2 = Wm √p b1/2m √p = Wm b1/2m+1 p b1/2m+1 = WmWn b1/2m+1 b1/2n b1/2m+1 by 44 VIII by III and 23 VII by 44 VIII Thus p2 > b1/2k for each k (taking n = m = k + 1,) and so p2 > p. It remains to be shown that p is the least projection above b. Let q be a projection in A with b 6 q; we must show that q 6 p. We have b1/2 6 q by 55 VIII, and so b1/2n 6 q for each n by induction. Hence p 6 q. (cid:3) II III IV V VII VIII IX Proposition Below every effect b of a von Neumann algebra A , there is greatest VI projection, ⌊b⌋, we call the floor of b, given by ⌊b⌋ =V∞n=0 b2n Moreover, if a ∈ A commutes with b, then b commutes with ⌊b⌋. Proof Note that 1 > b > b2 > b4 > ··· > 0, and define p :=Vn b2n (see 43 Ia.) If a ∈ A commutes with b, then a commutes with p. Indeed, such a commutes with b2 (because ab2 = bab = b2a,) and so a commutes with b2n for each n by induction. Thus a commutes with p ≡Vn b2n To see that p is a projection, i.e. p2 = p, we only need to show that p 6 p2, (by a variation on 44 XIII.) . ..55 -- 56.. 99 because we get p2 ≡ √p p √p 6 p from p 6 1 (using 25 II.) Now, since p2 = Vm √p b2m√p = Vm b2m−1 = VmVn b2m−1 p b2m−1 b2n b2n by a variation on 44 VIII by VIII and 23 VII b2m−1 by 44 VIII, X b2m−1 for all n, m, we get p 6 p2. and p 6 b2m−1 It remains to be shown that p is the greatest projection below b. Let q be a projection in A with q 6 b. We must show that q 6 p. Since q 6 b, we have q 6 b2 (by 55 IX), and so q 6 b2n . (cid:3) for each n by induction. Thus q 6 p ≡Vn b2n XI Exercise Show that given an effect a and a projection p in a von Neumann algebra A we have 1. pa = a iff ap = a iff ⌈a⌉ 6 p, and 2. pa = p iff ap = p iff p 6 ⌊a⌋. Conclude that ⌈a⌉ is the least projection p with a = ap (or, equivalently, a = pa), and that ⌊a⌋ is the greatest projection p with p = ap (or, equivalently, p = pa.) In particular, a = a⌈a⌉ = ⌈a⌉ a and ⌊a⌋ = a⌊a⌋ = ⌊a⌋ a. XII Example Given a finite complete measure space X we have ⌈f◦⌉ = 1◦ {x∈X : f (x)>0} and ⌊f◦⌋ = 1◦ {x∈X : f (x)=1} for every f ∈ L∞(X) with 0 6 f◦ 6 1. XIII Exercise Let a, b be effects of a von Neumann algebra A , and let λ ∈ [0, 1]. 1. Show that ⌈a⌉⊥ =(cid:4)a⊥(cid:5) and ⌊a⌋⊥ =(cid:6)a⊥(cid:7). 2. Show that ⌈λa⌉ = ⌈a⌉ when λ 6= 0. Use this to prove that (cid:6)λa + λ⊥b(cid:7) is the supremum of ⌈a⌉ and ⌈b⌉ in the poset of projections of A when λ 6= 0 and λ 6= 1. 3. Show that ⌊a⌋ =(cid:4)a2(cid:5) and ⌈a⌉ =(cid:6)a2(cid:7). XIV Lemma The supremum of a directed set D of projections from a von Neumann algebra A is a projection. XV Proof Writing p = W D, we must show that p2 = p. Note that dp = d for all d ∈ D (by 55 IX because d 6 p.) Now, on the one hand, (d)d∈D converges ultraweakly to p. On the other hand, (dp)d∈D converges ultraweakly to p2 by 44 VII. Hence p = p2 by uniqueness of ultraweak limits. Exercise Deduce from this result that every set A of projections from A has a XVI supremum S A and an infimum T A in the poset of projections from A . (Hint: use XIII, and the fact that p 7→ p⊥ is an order anti-isomorphism on the poset of projections on A .) Exercise Let A be a von Neumann algebra. XVII 1. Show that ⌈W D⌉ =Sd∈D ⌈d⌉ for every directed set D of effects from A . 2. Show that ⌊V D⌋ =Td∈D ⌊d⌋ for every filtered set D of effects from A . 3. Show that ⌈·⌉ does not preserve filtered infima, and ⌊·⌋ does not preserve directed suprema. (Hint: 1, 1 Conclude that ⌈·⌉ and ⌊·⌋ are neither ultraweakly, ultrastrongly nor norm continuous as maps from [0, 1]A to [0, 1]A . 2 , 1 3 , . . . .) this end, recall that (by 55 IX) a projection e is below an effect c iff ec = e iff Exercise Show that for a family (pi)i∈I of pairwise orthogonal projections (with I potentially uncountable) the seriesPi pi converges ultrastrongly toSi pi. (Hint: use the fact thatPi∈F pi =Si∈F pi for finite subsets F of I by 55 XIII.) Lemma Let a, b be effects of a von Neumann algebra A . Then ⌊√ab√a⌋ is the greatest projection below a and b, that is, ⌊√ab√a⌋ = ⌊a⌋ ∩ ⌊b⌋. Proof Surely, ⌊√ab√a⌋ 6 √ab√a 6 a. Let us prove that ⌊√ab√a⌋ 6 b. To e√c = e. In particular, since ⌊√ab√a⌋ 6 √ab√a and ⌊√ab√a⌋ 6 a, we get (cid:4)√ab√a(cid:5) = (cid:4)√ab√a(cid:5) √ab√a(cid:4)√ab√a(cid:5) = (cid:4)√ab√a(cid:5) b(cid:4)√ab√a(cid:5) , and so ⌊√ab√a⌋ b⊥ ⌊√ab√a⌋ = 0, which implies that ⌊√ab√a⌋ 6 b by 55 V. Now, let e be a projection below a and b, that is, e√a = e and eb = e. We must show that e 6 ⌊√ab√a⌋, or equivalently, e 6 √ab√a, or put yet differently, e√ab√a = e. But this is obvious: e = e√a = eb√a = e√ab√a. Having seen that ⌊√ab√a⌋ = ⌊a⌋∩⌊b⌋ in 57 one might wonder whether there is a similar expression for ⌈√ab√a⌉, but this doesn't seem to exist. However, for projections p and q we have ⌈pqp⌉ = p ∩ (p⊥ ∪ q) as we'll show below. Lemma Let p be a projection, and let a be an effect of a von Neumann algebra with a 6 p. We have p − ⌈a⌉ = ⌊p − a⌋. Proof We must show that p − ⌈a⌉ is the greatest projection below p − a. To begin, p − ⌈a⌉ 6 p − a, because a 6 ⌈a⌉. Further, since a 6 p, we have ⌈a⌉ 6 p, and so p − ⌈a⌉ is a projection (by 55 XIV). Let q be a projection below p − a. We must show that q 6 p − ⌈a⌉. The trick is to note that a 6 p − q. Since p − q is a projection (by 55 XIV because q 6 p − a 6 p), we have ⌈a⌉ 6 p − q, and so q 6 p − ⌈a⌉. (cid:3) (cid:3) XVIII 57 II III 58 II III ..56 -- 58.. 101 V IV Proposition We have ⌈pqp⌉ = p ∩ (p⊥ ∪ q) for all projections p and q from a von Neumann algebra. Proof Observe that ( p ∩ (p⊥ ∪ q) )⊥ = p⊥ ∪ (p ∩ q⊥). Since p⊥ and p ∩ q⊥ are disjoint, we have p⊥ ∪ (p ∩ q⊥) = p⊥ + p ∩ q⊥, and so p ∩ (p⊥ ∪ q) = p − p ∩ q⊥. By point V, it suffices to show that ⌈pqp⌉ = p − p ∩ q⊥, that is, p − ⌈pqp⌉ = p ∩ q⊥. Since p − ⌈pqp⌉ = ⌊p − pqp⌋ by II and (cid:4)pq⊥p(cid:5) = p ∩ q⊥ by 57 we are (cid:3) done. 3.2.2 Range and Support 59 Notation Let A be a von Neumann algebra. Because it will be very convenient II III we extend the definition of ⌈b⌉ to all positive b from A by ⌈b⌉ := (cid:6)kbk−1b(cid:7) when b 66 1. Note that -- contrary to what the notation suggests -- we do not have b 6 ⌈b⌉, for those b 66 1. Now, given an arbitrary element b of A , we'll call ⌈b) := ⌈b∗b⌉ the support (projection) of b, and (b⌉ := ⌈bb∗⌉ the range (projection) of b. Remark Some explanation is in order here. We did not just introduce the range and support notation for its own sake, but will use it extensively in §3.4 thanks to calculation rules such as ab = 0 ⇐⇒ ⌈a) (b⌉ = 0 (see 60 VIII). The notation was chosen such that (b⌉ b = b = b ⌈b) (see VI). Good examples are ⌈xihy ) = yihy and (xihy⌉ = xihx for unit vectors x and y from a Hilbert space H . Exercise Let a and b be positive elements of a von Neumann algebra A . 1. Given a projection p in A show that pa = a iff ap = a iff ⌈a⌉ 6 p. (In particular, ⌈a⌉ is the least projection p of A with ap = a.) 2. Show that ⌈a⌉ a = a⌈a⌉, and if fact, if b ∈ A commutes with a then b commutes with ⌈a⌉. 3. Show that a = 0 iff ⌈a⌉ = 0. 4. Show that ⌈a⌉ = ⌈λa⌉ for every λ > 0. IV Exercise Let a be a self-adjoint element of a von Neumann algebra. Show that ⌈a + b⌉ = ⌈a⌉ ∪ ⌈b⌉. 5. Show that (cid:6)a2(cid:7) = ⌈a⌉. 1. Show that ⌈a+⌉⌈a−⌉ = 0. (Hint: recall from 24 II that a+a− = 0.) 2. Show that ⌈a+⌉ a = a⌈a+⌉ = a+ and ⌈a−⌉ a = a⌈a−⌉ = −a−. Exercise Show that ⌈W D⌉ =Sd∈D ⌈d⌉ for every bounded directed set of positive elements of a von Neumann algebra A . Exercise Let a and b be elements of a von Neumann algebra. V VI 1. Show that ⌈a) ≡ ⌈a∗a⌉ is the least projection p of A with ap = a. 2. Show that (a⌉ ≡ ⌈aa∗⌉ is the least projection p of A with pa = a. 3. Show that ⌈a∗) = (a⌉ and (a∗⌉ = ⌈a). 4. Show that ⌈ab) 6 ⌈b) and (ab⌉ 6 (a⌉. Exercise Let T be a bounded operator on a Hilbert space H . VII 1. Show that (T⌉ is the projection onto the closure Ran(T ) of the range of T . 2. Show that ⌈T ) is the projection onto the support of T , i.e. the orthocom- plement Ker(T )⊥ of the kernel of T . 3. Show that ⌊T⌋ is the projection on { x ∈ H : T x = x} when T is an effect. 60 II Lemma Given a positive element a of a von Neumann algebra A and an np- functional ω : A → C we have ω(a) = 0 iff ω(⌈a⌉) = 0. Proof Note that if a = 0, the stated result is clearly correct, and the other case, when kak 6= 0, the problem reduces to the case that 0 6 a 6 1 by replacing a by a . So let us just assume that a ∈ [0, 1]A to begin with. For similar reasons, kak we may assume that ω(1) 6 1. Now, since 0 6 a 6 ⌈a⌉ we have 0 6 ω(a) 6 ω(⌈a⌉), so ω(⌈a⌉) = 0 =⇒ ω(a) = 0 is obvious. It remains to be shown that ω(⌈a⌉) = 0 given ω(a) = 0. Since ⌈a⌉ = Wn a1/2n (by 56 I) and ω is normal, we have ω(⌈a⌉) = Wn ω(a1/2n ), and so it suffices to show that ω(a1/2n ) = 0 for each n. As a result of Kadison's inequality (see 30 IV) we have ω(√a)2 6 ω(a) = 0, and so ω(√a) = 0. Since then ω(p√a) = 0 by the same token, and so on, we get ω(a1/2n ) = 0 for all n by induction. (cid:3) Proposition For positive elements a and b of a von Neumann algebra A , III ⌈a⌉ 6 ⌈b⌉ ⇐⇒ ∀ω [ ω(b) = 0 =⇒ ω(a) = 0 ], where ω ranges over all np-functionals A → C. Proof When ⌈a⌉ 6 ⌈b⌉ and ω is an np-functional on A with ω(b) = 0, then 0 6 ω(⌈a⌉) 6 ω(⌈b⌉) = 0 (by I), and so ω(⌈a⌉) = 0, so that ω(a) = 0 (again by I). For the other direction, assume that ω(b) = 0 =⇒ ω(a) = 0 for ev- ery np-functional ω on A ; we must show that ⌈a⌉ 6 ⌈b⌉, or in other words, IV ..58 -- 60.. 103 ⌈b⌉⊥ ⌈a⌉⌈b⌉⊥ = 0. Let ω : A → C be an arbitrary np-functional; it suffices to show that ω(⌈b⌉⊥ ⌈a⌉⌈b⌉⊥ ) = 0. Since ⌈b⌉⊥b⌈b⌉⊥ = 0 we have ω(⌈b⌉⊥b⌈b⌉⊥) = 0 and so ω(⌈b⌉⊥a⌈b⌉⊥) = 0 (by assumption, because ω(⌈b⌉⊥(· )⌈b⌉⊥) is an np- functional on A as well), which implies that ω(⌈b⌉⊥ ⌈a⌉⌈b⌉⊥) = 0 by I. (cid:3) V Proposition Let f : A → B be an np-map between von Neumann algebras. Then ⌈f (a)⌉ = ⌈f (⌈a⌉)⌉ for every a ∈ A+. VI Proof By III it suffices to show that ω(f (a)) = 0 iff ω(f (⌈a⌉)) = 0 for every (cid:3) np-functional ω : B → C, and this is indeed the case by I. VII Exercise Let a and b be elements of a von Neumann algebra A . 1. Deduce from V that ⌈a∗ba⌉ = ⌈a∗ ⌈b⌉ a⌉ when b > 0. 2. Conclude that ⌈ab) = ⌈⌈a) b) and (ab⌉ = (a (b⌉⌉ (see 59 I). VIII Exercise Let a and b be elements of a von Neumann algebra A . 1. Show that cb = 0 iff ⌈c) (b⌉ = 0 iff ⌈c) 6 (b⌉⊥ for c ∈ A . (Hint: if cb = 0, then ⌈b∗c∗cb⌉ ≡ ⌈b∗ ⌈c∗c⌉ b⌉ = 0 by VII.) 2. Show that c1b = c2b =⇒ c1 = c2 for all c1, c2 ∈ A with ⌈ci) 6 (b⌉. 3. Show that b∗c1b = b∗c2b =⇒ c1 = c2 for all c1, c2 ∈ (b⌉ A (b⌉ IX Exercise Let f : A → B be an np-map between von Neumann algebras. 1. Show that ⌈f (p ∪ q)⌉ = ⌈f (p)⌉ ∪ ⌈f (q)⌉ for all projections p and q in A . (Hint: recall from 56 XIII that p ∪ q =(cid:6) 1 2. Deduce from this and V that ⌈f (S A)⌉ = Sa∈A ⌈f (a)⌉ for every set of projections A from A . 2 p + 1 2 q(cid:7).) 3. Show that there is a greatest projection e in A with f (e) = 0. 61 II Given the rule ⌈f (⌈a⌉)⌉ = ⌈f (a)⌉ for an np-map f and self-adjoint a one might surmise that the equation ⌈f (⌈a))⌉ = ⌈f (a)) holds for arbitrary a; but one would be mistaken to do so. We can, however, recover an inequality by assuming that f is completely positive, see II. One of its corollaries is that ncpsu-isomorphisms are in fact nmiu-isomorphisms (see 99 IX). Proposition Given an ncp-map f : A → B between von Neumann algebras we have, for all a ∈ A , ⌈f (⌈a) )⌉ 6 ⌈f (a) ) and ⌈f ( (a⌉ )⌉ 6 ( f (a)⌉ . Proof Since f (a)∗f (a) 6 kf (1)k2 f (a∗a) by 34 XIV, we get ⌈f (a) ) ≡ ⌈f (a)∗f (a)⌉ 6 ⌈kf (1)k2f (a∗a)⌉ 6 ⌈f (a∗a)⌉ = ⌈f (⌈a∗a⌉)⌉ ≡ ⌈f (⌈a))⌉. One obtains ⌈f ( (a⌉ )⌉ 6 ( f (a)⌉ along similar lines. (cid:3) III 62 II Proposition Let f : A → B be a ncpsu-map between von Neumann algebras. Then ⌊f (a)⌋ = ⌊f (⌊a⌋)⌋ for every effect a from A . Proof Since ⌊a⌋ 6 a, we have ⌊f (⌊a⌋)⌋ 6 ⌊f (a)⌋. Thus we only need to show that ⌊f (a)⌋ 6 ⌊f (⌊a⌋)⌋, or equivalently, ⌊f (a)⌋ 6 f (⌊a⌋). We have ⌊f (a)⌋ 56 XIII=== (cid:4)f (a)2(cid:5) 6 (cid:4)f (a2)(cid:5) 6 ⌊f (a)⌋ , and so ⌊f (a)⌋ =(cid:4)f (a2)(cid:5). By induction, and similar reasoning, we get ⌊f (a)⌋ = ) for every n, and so ⌊f (a)⌋ 6 Vn f (a2n (cid:4)f (a2n f (⌊a⌋), where we used that f is normal, and ⌊a⌋ =Vn a2n ) = f (Vn a2n )(cid:5) 6 f (a2n 3.2.3 Carrier and Commutant 4 XV (see 56 VI). ) = (cid:3) Definition The carrier of an np-map f : A → B between von Neumann algebras (written ⌈f⌉) is the least projection p with f (p⊥) = 0 (which exists by 60 IX.) Exercise Let f, g : A → B be np-maps between von Neumann algebras. 63 II 1. Show that ⌈λf⌉ = ⌈f⌉ for all λ > 0. 2. Show that ⌈f + g⌉ = ⌈f⌉ ∪ ⌈g⌉. 3. Show that ⌈f⌉ = 1 iff f is faithful. 4. Assuming f is multiplicative show that ⌈f⌉ = 1 iff f is injective. (There is more to be said about the carrier of an nmiu-map, see 69 IV.) Exercise III 1. Given an element a of a von Neumann algebra A show that ⌈a∗(· )a⌉ = ⌈aa∗⌉ ≡ (a⌉ where a∗(· )a is interpreted as an np-map A → A . 2. Given a bounded operator T : H → K between Hilbert spaces show that ⌈T ∗(· )T⌉ is the projection onto Ran(T ) when T ∗(· )T is interpreted as a map B(K ) → B(H ). ..60 -- 63.. 105 3. Show that ⌈hx, (· )xi⌉ = xihx for any unit vector x from a Hilbert space H when hx, (· )xi is interpreted as a map B(H ) → C. (But be warned: when A is a von Neumann subalgebra of B(H ) the car- rier of the restriction hx, (· )xi : A → C might differ from xihx because the former is in A , while the latter may not be, see 88 IV.) IV V Lemma Let f : A → B be a p-map between C∗-algebras, and let p be an effect of A with f (p⊥) = 0. Then f (a) = f (pa) = f (ap) = f (pap) for all a ∈ A . Proof Assume B = C for now. Since p⊥ 6 1, we have (p⊥)2 =pp⊥p⊥pp⊥ 6 p⊥, and so 0 6 f ( (p⊥)2 ) 6 f (p⊥) = 0, giving us f ( (p⊥)2 ) = 0. Since f (p⊥a)2 6 f ( (p⊥)2 ) f (a∗a) = 0 by Kadison's inequality, 30 IV, we get f (p⊥a) = 0, and so f (pa) = f (a) for all a ∈ A . In particular, f (ap) = f (pa∗)∗ = f (a∗)∗ = f (a) for all a ∈ A , and so f (pap) = f (pa) = f (a) for all a ∈ A . Letting B be again arbitrary, and given a ∈ A , note that since the states on B are separating (by 22 VIII) it suffices to show that ω(f (a)) = ω(f (ap)) = ω(f (pa)) = ω(f (pap)) for all states ω : B → C. But this follows from the previous paragraph since ω ◦ f is a p-map into C. (cid:3) VI Corollary Given an np-map f : A → B between von Neumann algebras we have f (a) = f (⌈f⌉ a) = f (a⌈f⌉) = f (⌈f⌉ a⌈f⌉) for all a ∈ A . 64 We turn to the task of showing that every element of a von Neumann algebra is the norm limit of linear combinations of projections in 65 IV. We'll deal with the commutative case first (see II). II III 65 Proposition Every element a of a commutative von Neumann algebra A is the norm limit of linear combinations of projections. Proof By 53 II it suffices to show that the linear span of projections is norm dense in C(sp(A )). For this, in turn, it suffices by Stone -- Weierstrass' theo- rem (see 27 XX) to show that the projections in C(sp(A )) separate the points of sp(A ) in the sense that given x, y ∈ sp(A ) with x 6= y there is a projection f in C(sp(A )) with f (x) 6= f (y). Since sp(A ) is Hausdorff there are for such x and y disjoint open subsets U and V of sp(A ) with x ∈ U and y ∈ V . Then f := 1U is a projection in C(sp(A )) (continuous because U is clopen by 53 III) with f (x) = 0 6= 1 = f (y) since x ∈ U ⊆ sp(A )\V , and so y /∈ U . (cid:3) To reduce the general case to the commutative case we need the following tool (that will be useful later on too for different reasons). II Definition Given a subset S of a von Neumann algebra A the commutant of S is the set, denoted by S(cid:3), of all a ∈ A with as = sa for all s ∈ S. The commutant of A itself is denoted by Z(A ) := A (cid:3) and is called the centre of A . (Its elements, called central, are the subjects of the next section.) Exercise Let S and T be subsets of a von Neumann algebra A . III 1. Show that S ⊆ T (cid:3) iff T ⊆ S(cid:3). Show that S ⊆ T entails T (cid:3) ⊆ S(cid:3). Show that S ⊆ S(cid:3)(cid:3), and S(cid:3)(cid:3)(cid:3) = S(cid:3). 2. Show that S(cid:3) is closed under addition, (scalar) multiplication, contains the unit of A , and is ultraweakly closed. 3. Show that the commutant S(cid:3) need not be closed under involution. (Hint: compute {(cid:0) 0 1 0 0(cid:1)}(cid:3) in M2.) Suppose S is closed under involution. Show S(cid:3) is closed under involution as well, and conclude that in that case S(cid:3) is a von Neumann subalgebra of A . Show that Z(A ) is a von Neumann subalgebra of A . Show that S(cid:3)(cid:3) is a von Neumann subalgebra of A with S ⊆ S(cid:3)(cid:3). Show that if S is commutative (i.e. S ⊆ S(cid:3)), then so is S(cid:3)(cid:3). 4. In particular, if B is a von Neumann subalgebra of A , then B(cid:3)(cid:3) is a von Neumann subalgebra of A with B ⊆ B(cid:3)(cid:3). Show that ( A ∩ C )(cid:3) = A , and so ( A ∩ C )(cid:3)(cid:3) = Z(A ). So in general B(cid:3)(cid:3) needn't equal B. Nevertheless, we'll see in 88 V that B(cid:3)(cid:3) = B when A is of the form A = B(H ) for some Hilbert space H . 5. Given a von Neumann subalgebra B of A verify that Z(B) = B ∩ B(cid:3). Proposition Every self-adjoint element a of a von Neumann algebra A is the norm limit of linear combinations of projections from {a}(cid:3)(cid:3). Proof Since a is an element of the by III commutative von Neumann subal- gebra {a}(cid:3)(cid:3) of A , a is the norm limit of linear combinations of projections from {a}(cid:3)(cid:3) by 64 II. (cid:3) The carriers of np-functionals play such an important role in the theory that we decided to give them a name. Definition We call a projection p of a von Neumann algebra A ultracyclic if p = ⌈ω⌉ for some np-map ω : A → C. Remark Some explanation of this terminology is in order. A projection E in a von Neumann subalgebra R of B(H ) is usually defined to be cyclic when E is the projection onto R(cid:3)x for some x ∈ H (see Definition 5.5.8 [47]). With 88 IV and 88 VI we'll be able to see that this amounts to requiring that E be the IV V 66 II III ..63 -- 66.. 107 carrier of the vector functional hx, (· )xi : R → C. So, loosely speaking, a cyclic projection is the carrier of a vector functional with respect to some fixed Hilbert space, while an ultracyclic projection is the carrier of a vector functional with respect to some arbitrary Hilbert space. IV Exercise Let A be a von Neumann algebra. Verify the following facts. 1. If p and q are ultracyclic projections in A , then p ∪ q is ultracyclic. 2. If p 6 q are projections in A , and q is ultracyclic, then p is ultracyclic. 3. Every projection p in A is a directed supremum of ultracyclic projections. In fact, p = Wω ⌈ω⌉ where ω ranges over the np-functionals on A with ω(p⊥) = 0. (Hint: first consider p = 1.) 4. Every projection p in A is the sum of ultracyclic projections: there are np-functionals (ωi)i on A with p =Pi ⌈ωi⌉. 3.2.4 Central Support and Central Carrier 67 II Definition An element a of a von Neumann algebra A is called central when ab = ba for all b ∈ A (that is, when a ∈ Z(A ), see 65 III). Examples 1. In a commutative von Neumann algebra every element is central. 2. An element a of a direct sum Li iff ai is central for each i. Ai of von Neumann algebras is central 3. In B(H ), where H is a Hilbert space, only the scalars are central. Indeed, given a positive central element A of B(H ), we have(cid:10)x, Akyk2x(cid:11) = hx, (Axihy) yi = hx, (xihy A) yi = hx,k√Ayk2xi for all x, y ∈ H , and so Akyk2 = k√Ayk2 for all y ∈ H . Hence A is (zero or) a scalar. III Remark A von Neumann algebra in which only the scalars are central -- of which a B(H ) is but the simplest example -- is called a factor. The classifica- tion of these factors is an important part of the theory of von Neumann algebras that we did not need in this thesis. IV Exercise Note that if a von Neumann algebra A can be written as a direct sum A ∼= B1 ⊕ B2, then (1, 0) ∈ B1 ⊕ B2 gives a central projection in A . The converse also holds: 1. Given a central projection c in A , show that cA ≡ { ca : a ∈ A } is a von Neumann subalgebra of A for all but the fact that 1 need not be in cA . Show cA is a von Neumann algebra with c as unit, and that a 7→ (ca, c⊥a) gives an nmiu-isomorphism A → cA ⊕ c⊥A . 2. Given a family of central projections (ci)i in A with Pi ci = 1 show that a 7→ (cia)i gives an nmiu-isomorphism A →Li ciA . Proposition Given a projection e of a von Neumann algebra A 68 ⌈⌈e⌉⌉ := [a∈A ⌈a∗ea⌉ is the least central projection above e. Proof Let us first show that ⌈⌈e⌉⌉ is central. Given b ∈ A we have ⌈⌈⌈e⌉⌉ b) = ⌈b∗ ⌈⌈e⌉⌉ b⌉ = Sa∈A ⌈b∗ ⌈a∗ea⌉ b⌉ = Sa∈A ⌈(ab)∗eab⌉ 6 ⌈⌈e⌉⌉ by 60 IX, which im- plies that ⌈⌈e⌉⌉ b ⌈⌈e⌉⌉ = ⌈⌈e⌉⌉ b. Since similarly (or consequently) ⌈⌈e⌉⌉ b ⌈⌈e⌉⌉ = b ⌈⌈e⌉⌉ we get b ⌈⌈e⌉⌉ = ⌈⌈e⌉⌉ b ⌈⌈e⌉⌉ = ⌈⌈e⌉⌉ b, and so ⌈⌈e⌉⌉ is central. Clearly e 6 ⌈⌈e⌉⌉. It remains to be shown that ⌈⌈e⌉⌉ 6 c given a central projection c with e 6 c. For this it suffices to show that ⌈ea) ≡ ⌈a∗ea⌉ 6 c given a ∈ A . Now, since e 6 c we have ec = e and so eac = eca = ea which implies that ⌈ea) 6 c. Thus ⌈⌈e⌉⌉ 6 c. (cid:3) Definition Let a be an element of a von Neumann algebra A . Since given a central projection c of A we have ⌈⌈⌈a)⌉⌉ 6 c iff ⌈a) 6 c iff ac = a iff ca = a iff ⌈⌈(a⌉⌉⌉ 6 c, we see that ⌈⌈a⌉⌉ := ⌈⌈⌈a)⌉⌉ = ⌈⌈(a⌉⌉⌉ is the smallest central projection p with pa = a, which we'll call the central support of a. Exercise Let A be a von Neumann algebra. II III IV 1. Show that ⌈⌈a⌉⌉ = ⌈⌈a∗⌉⌉ = ⌈⌈a∗a⌉⌉ = ⌈⌈aa∗⌉⌉ for all a ∈ A . 2. Show that ⌈⌈W D⌉⌉ =Sd∈D ⌈⌈d⌉⌉ for any bounded directed subset of A . Show that ⌈⌈S E⌉⌉ =Se∈E ⌈⌈e⌉⌉ for any collection of projections from A . Show that ⌈⌈a + b⌉⌉ = ⌈⌈⌈a⌉ ∪ ⌈b⌉⌉⌉ = ⌈⌈a⌉⌉ ∪ ⌈⌈b⌉⌉ for all a, b ∈ A . 3. Given a ∈ A and a central projection c of A show that ⌈⌈a⌉⌉ c = ⌈⌈ac⌉⌉. Conclude that ⌈⌈a⌉⌉⌈⌈b⌉⌉ = ⌈⌈a⌈⌈b⌉⌉⌉⌉ = ⌈⌈⌈⌈a⌉⌉ b⌉⌉ = ⌈⌈a⌉⌉ ∩ ⌈⌈b⌉⌉ for all a, b ∈ A . Definition Let f : A → B be an np-map between von Neumann algebras. Show that given a central effect c of A we have f (c⊥) = 0 iff ⌈f⌉ 6 c iff ⌈⌈⌈f⌉⌉⌉ 6 c, and so ⌈⌈f⌉⌉ := ⌈⌈⌈f⌉⌉⌉ is the least central effect (and central projection) p with f (p⊥) = 0, which we'll call the central carrier of f . 69 ..66 -- 69.. 109 II III Proposition Every two-sided ideal D of a von Neumann algebra A that is closed under bounded directed suprema of self-adjoint elements -- for example when A is ultrastrongly closed -- is of the form cA for some unique central projection c of A . Moreover, c is the greatest projection in D. Proof We'll obtain c as the supremum over all effects in D, and to this end we'll 2 a + 1 show first that D ∩ [0, 1]A is directed. Since ⌈a⌉ ∪ ⌈b⌉ ≡(cid:6) 1 2 b(cid:7) (see 56 XIII) is an upper bound for a, b ∈ D ∩ [0, 1]A it suffices to show that ⌈a⌉ ∈ D for all a ∈ D ∩ [0, 1]A , which, in turn, follows from ⌈a⌉ =Wn a1/2n , see 56 I. Hence D ∩ [0, 1]A is directed, and so we may define c := W D ∩ [0, 1]A . Since D is a von Neumann subalgebra of A , we'll have c ∈ D ∩ [0, 1]A , and so c is the greatest element of D ∩ [0, 1]A . In particular, c will be above ⌈c⌉ implying ⌈c⌉ = c and making c a projection -- the greatest projection in D. Given a ∈ A we claim that a ∈ D iff ca = a. Surely, if a = ca, then a = ca ∈ D, because D is a two-sided ideal of A . Concerning the other direction, note that given a ∈ D the equality ac = a holds when a is an effect by 55 VIII (because a 6 c), and thus when a is self-adjoint too (by scaling), and hence for arbitrary a ∈ D by writing a ≡ aR + iaI where aR and aI are self-adjoint. Note that this claim entails that D ⊆ cA . Since D is an ideal we also have cA ⊆ D, and so D = cA . The claim also entails that c is central. Indeed, given a ∈ A we have ac ∈ D (because D is an ideal) and so c(ac) = ac by the claim. Since similarly (ca)c = ca, we get ac = ca. The only thing that remains to be shown is that c is unique. To this end let c and c′ be central projections with cA = D = c′A . As c′ ∈ D = cA , there is a ∈ A with c′ = ca. Then c′ = c′(c′)∗ = caa∗c∗ 6 cc∗kaa∗k = ckak2, and so c′ 6 c. Since similarly c 6 c′, we get c = c′. (cid:3) IV Corollary The carrier ⌈f⌉ of an nmiu-map f : A → B between von Neumann algebras is central, so ⌈f⌉ = ⌈⌈f⌉⌉. Moreover, ker(f ) = ⌈⌈f⌉⌉⊥A . IVa Exercise Show using IV and 67 IV that an nmiu-map f : A → B factors as f B , <①①①①①①①①① h : a7→f (a) A #❋❋❋❋❋❋❋❋❋ g : a7→⌈⌈f⌉⌉a ⌈⌈f⌉⌉ A where the g an nmiu-surjection, and h is an nmiu-injection. IVb Use this, and 48 VI, to show that f (A ) is a von Neumann subalgebra of B. V Lemma We have ⌈⌈ω⌉⌉ = ⌈ω⌉ for every np-functional ω : A → C on a von Neumann algebra A , where ω is as in 30 VI. VI Proof Let e be a projection in A . Note that 0 = kω(e)(ηω(a))k2 ≡ ω(a∗ea) iff ⌈a∗ea⌉ 6 ⌈ω⌉⊥ iff ⌈a⌈ω⌉ a∗⌉ 6 e⊥ for all a ∈ A . So since the ηω(a)'s / / # < lie dense in Hω, we have ω(e) = 0 iff ω(e)(ηω(a)) = 0 for all a ∈ A iff Sa∈A ⌈a⌈ω⌉ a∗⌉ 6 e⊥. Hence ⌈ω⌉ = Sa∈A ⌈a⌈ω⌉ a∗⌉ ≡ Sa∈A ⌈a∗ ⌈ω⌉ a⌉ = ⌈⌈⌈ω⌉⌉⌉ = ⌈⌈ω⌉⌉ by 68 I. (cid:3) Proposition Given a collection of np-functionals Ω on a von Neumann alge- bra A we have ⌈Ω⌉ =Sω∈Ω ⌈⌈ω⌉⌉ for Ω : A → B(HΩ) from 30 VI. Proof Let e be a projection of A . Since Ω(e)(x) =Pω∈Ω ω(xω) by 30 VI for all x ∈ HΩ ≡ Lω∈Ω Hω, we have Ω(e) = 0 iff ω(e) = 0 for all ω ∈ Ω iff e 6 ⌈ω⌉⊥ ≡ ⌈⌈ω⌉⌉⊥ iff e 6 Tω∈Ω ⌈⌈ω⌉⌉⊥ ≡ (Sω∈Ω ⌈⌈ω⌉⌉)⊥. Hence ⌈Ω⌉ = Sω∈Ω ⌈⌈ω⌉⌉. (cid:3) Corollary For a collection Ω of np-functionals on a von Neumann algebra, the following are equivalent. 1. Ω is centre separating (see 21 II). 2. A central projection z of A is zero when ω(z) = 0 for all ω ∈ Ω. 3. The map Ω : A → B(HΩ) from 30 VI is injective. Proof We've seen in 30 X that 1 ⇐⇒ 3, and 1⇒2 is trivial, which leaves us with 2⇒3. So assume that ∀ω ∈ Ω [ ω(z) = 0 ] =⇒ z = 0 for every central projection z of A . Then since ⌈Ω⌉⊥ is a central projection by VII with ⌈Ω⌉⊥ = (cid:0)Sω∈Ω ⌈⌈ω⌉⌉(cid:1)⊥ =Tω∈Ω ⌈⌈ω⌉⌉⊥ 6 ⌈⌈ω⌉⌉⊥ 6 ⌈ω⌉⊥ and thus ω(⌈Ω⌉⊥) 6 ω(⌈ω⌉⊥) = 0 for all ω ∈ Ω we get ⌈Ω⌉⊥ = 0, and so Ω is injective by 63 II. With our new-found knowledge on central elements we can complete the classi- fication of commutative von Neumann algebras we started in 52. (cid:3) VII VIII IX X 70 Exercise Show that every central projection c of a von Neumann algebra is of II the form c ≡Pi ⌈⌈ωi⌉⌉ for some family of np-functionals (ωi)i on A . (Hint: take (ωi)i to be a maximal set of np-functionals for which the ⌈⌈ωi⌉⌉'s are orthogonal.) Theorem Every commutative von Neumann algebra is nmiu-isomorphic to a direct sum of the formLi L∞(Xi) where Xi are finite complete measure spaces. Proof By II we have 1 ≡ Pi ⌈⌈ωi⌉⌉ for some np-functionals ωi : A → C, and so A ∼= Li ⌈⌈ωi⌉⌉A by 67 IV. Since A is commutative, and so ⌈⌈ωi⌉⌉ = ⌈ωi⌉, we see that restricting ωi gives a faithful functional on ⌈⌈ωi⌉⌉A , which is therefore by 54 XI nmiu-isomorphic to L∞(Xi) for some finite complete measure space Xi. From this the stated result follows. (cid:3) III IV 3.3 Completeness We set to work on the ultrastrong and bounded ultraweak completeness of von Neumann algebras (see 77 I) and their precursors: 71 ..69 -- 71.. 111 1. A linear (not necessarily positive) functional on a von Neumann algebra is ultraweakly continuous iff it is ultrastrongly continuous (see 72 XI). 2. A convex subset of a von Neumann algebra is ultraweakly closed iff it is ultrastrongly closed (see 73 VIII). 3. (Kaplansky's density theorem) The unit ball (A )1 of a C∗-subalgebra A ¯A is of a von Neumann algebra B is ultrastrongly dense in ( ¯A )1 where the ultrastrong (=ultraweak, 73 VIII) closure of A (see 74 IV). 4. Any von Neumann subalgebra A of B is ultraweakly and ultrastrongly closed in B (see 75 VIII). 5. The von Neumann algebra B(H ) of bounded operators on a Hilbert space H is ultrastrongly (76 I) and bounded ultraweakly complete (76 III). 3.3.1 Closure of a Convex Subset 72 We saw in 46 III that a positive linear functional f on a von Neumann algebra is ultrastrongly continuous iff it is ultraweakly continuous. In this section, we'll show that the same result holds for an arbitrary linear functional f . Note that if f is ultraweakly continuous, then f is automatically ultrastrongly continuous (because ultrastrong convergence implies ultraweak convergence). For the other direction, we'll show that if f is ultrastrongly continuous, then f can be written k=0 ikfk of np-maps f0, . . . , f3, and must therefore II Definition Let A be a von Neumann algebra. Given an np-map ω : A → C, be ultraweakly continuous. We'll need the following tool. as a linear combination f ≡P3 and b ∈ A , define b ∗ ω : A → C by (b ∗ ω)(a) = ω(b∗ab) for all a ∈ A . Exercise Let ω : A → C be an np-map on a von Neumann algebra A . III 1. Note that b ∗ ω : A → C is an np-map for all b ∈ A . Show that ω(a∗bc) 6 kωk kakω kbk kckω for all a, b, c ∈ A . Deduce that kb∗ ω− b′∗ ωk 6 kωk kb− b′kω (kbkω +kb′kω) for all b, b′ ∈ A . 2. Let b1, b2, . . . be a sequence in A which is Cauchy with respect to k · kω. Show that the sequence b1 ∗ ω, b2 ∗ ω, . . . is Cauchy (in the operator norm on bounded linear functionals A → C), and converges to a bounded linear map f : A → C. Show that f is an np-map. IV Exercise Let f : A → C be an ultrastrongly continuous linear functional on a von Neumann algebra A . Show that there are an np-map ω : A → C and δ > 0 with f (a) 6 1 for all a ∈ A with kakω 6 δ. (Keep this in mind when reading the following lemma.) Lemma Let ω : A → C be an np-map, and let f : A → C be a linear map. The following are equivalent. V 1. f (a) 6 B for all a ∈ A with kakω 6 δ, for some δ, B > 0; 2. f (a) 6 Bkakω for all a ∈ A , for some B > 0; 3. f (a) = [b, a]ω for all a ∈ A , for some b ∈ Hω (where Hω is the Hilbert space completion of A with respect to the inner-product [· , · ]ω). 4. f ≡ f0 + if1 − f2 − if3 where f0, . . . , f3 : A → C are np-maps for which there is B > 0 such that fk(a) 6 Bω(a) for all a ∈ A+ and k. Proof We make a circle. (4=⇒1) For a ∈ A and k, we have fk(a)2 6 fk(1) fk(a∗a) 6 fk(1)B ω(a∗a), giving fk(a) 6 (fk(1)B)1/2kakω, and so f (a) 6 Bkakω, where VI VII B = B1/2P3 k=0 fk(1)1/2. Hence f (a) 6 B for all a ∈ A with kakω 6 1. (1=⇒2) Let a ∈ A , and ε > 0 be given. Then for a := δ(ε + kakω)−1 a, we have kakω 6 δ, and so f (a) ≡ δ(ε + kakω)−1 f (a) 6 B, which entails f (a) 6 Bδ−1(ε + kakω). Since ε > 0 was arbitrary, we get f (a) 6 Bδ−1kakω. (2=⇒3) Since f (a) 6 Bkakω for all a ∈ A , the map f can be extended to a bounded linear map f : Hω → C. Then by Riesz' representation theorem, 5 IX, there is b ∈ Hω with f (x) = [b, x]ω for all x ∈ Hω. In particular, f (a) = [b, a]ω for all a ∈ A . (3=⇒4) We know that f (a) ≡ [b, a]ω for all a ∈ A , for some b ∈ Hω. Then, in A which converges to b by definition of Hω, there is a sequence b1, b2, . . . in Hω. Then the maps [bn, · ]ω : A → C approximate f = [b, · ]ω in the sense that f (a) − [bn, a]ω = [b − bn, a]ω 6 kb − bnkωkakω 6 kb − bnkωkωk1/2kak for all a ∈ A . In particular, [b1, · ]ω, [b2, · ]ω, . . . converges to f (in the operator norm). By "polarisation" (c.f. 44 II), we have [bn, a]ω = 1 k=0 ikfk,n(a), where fk,n := (ikbn + 1) ∗ ω is an np-map. Since (ikbn + 1)n is Cauchy with respect to k · kω, we see by III that (fk,n)n converges to an np-map fk : A → C (with respect to the operator norm). It follows that f = 1 4P3 It remains to be shown that there is B > 0 with fk(a) 6 Bω(a) for all k and a ∈ A+. Note that since fk,n(a) 6 kikbn + 1kω ω(a) 6 (kbnkω + 1) ω(a), for all n, k, and a ∈ A+, the number B := limn kbnkω + 1 will do. (cid:3) Corollary For a linear map f : A → C on a von Neumann algebra A the following are equivalent. k=0 ikfk. 4P3 VIII IX X XI 1. f is ultrastrongly continuous; ..71 -- 72.. 113 2. f is ultraweakly continuous; 3. f ≡ f0 + if1 − f2 − if3 for some np-maps f0, . . . , f3 : A → C; 4. "f is bounded on some k · kω-ball," that is, sup{ f (a) : a ∈ A : kakω 6 δ } < ∞ for some δ > 0 and np-map ω : A → C; 5. f (a) 6 kakω for all a ∈ A , for some np-map ω : A → C. 73 We'll show that the ultrastrong and ultraweak closure of a convex set agree. For this we need the following proto-Hahn -- Banach separation theorem, which concerns the following notion of openness. II Definition A subset A of a real vector space V is called radially open if for all a ∈ A and v ∈ V there is t ∈ (0,∞) with a + sv ∈ A for all s ∈ [0, t). Exercise Let V be a vector space. III 1. Show that the radially open subsets of V form a topology. 2. Show that with respect to this topology, scalar multiplication and trans- lations x 7→ x + a by a fixed vector a ∈ V are continuous. 3. Show that the subset of R2 depicted below in blue , including the point in the middle but not the dashed borders, is radially open, but not open in the usual topology on R2. 4. Show that addition on R2 is not jointly radially continuous. 5. Show that nevertheless {s ∈ R : sx + (1 − s)y ∈ A} is open for every radially open A ⊆ V , and x, y ∈ V . 6. Show that A + B is radially open when A, B ⊆ V are radially open. Show that {λa : a ∈ A, λ > 0} is radially open when A is radially open. Theorem For every radially open convex subset K of a real vector space V with 0 /∈ K there is a linear map f : V → R with f (x) > 0 for all x ∈ K. Proof (Based on Theorem 1.1.2 of [47].) IV V By Zorn's Lemma we may assume without loss of generality that K is max- imal among radially open convex subsets of V that do not contain 0. 2 x + 1 2 y), and K is convex. We also assume that K is non-empty, because if K = ∅, the result is trivial. We will show in a moment that H := {x ∈ V : − x, x /∈ K} is a linear subspace and V /H is one-dimensional. From this we see that there is a linear map f : V → R with ker(f ) = H. Since f (K) is a convex subset which does not contain 0 (because H ∩ K = ∅) we either have f (K) ⊆ (0,∞) or f (K) ⊆ (−∞, 0). Thus, by replacing f by −f if necessary, we see that there is a linear map f : V → R with f (x) > 0 for all x ∈ K. (H is a linear subspace) Note that x ∈ K, λ > 0 =⇒ λx ∈ K, because the subset {λx : x ∈ K, λ ∈ (0,∞)} ⊇ K is radially open, convex, doesn't contain 0, and is thus K itself. Furthermore, x, y ∈ K =⇒ x + y ∈ K, because x + y = 2( 1 Let K be the set of all x ∈ V with x + y ∈ K for all y ∈ K. Then it is not difficult to check that K is a cone: 0 ∈ K, and x ∈ K, λ > 0 =⇒ λx ∈ K, and x, y ∈ K =⇒ x + y ∈ K. We claim that x ∈ K iff −x /∈ K. Indeed, if x ∈ K, then −x /∈ K, because otherwise −x ∈ K and so 0 = x + (−x) ∈ K, which is absurd. For the other direction, suppose that −x /∈ K. Then x + y ∈ K for all y ∈ K, because {λx + y : y ∈ K, λ > 0} ⊇ K is radially open, convex, doesn't contain 0, and is thus K. It follows that H = K ∩ −K. Since K is a cone, −K is a cone, and thus H is a cone. But then −H = H is a cone too, and thus H is a linear subspace. (V /H is one-dimensional) Note that H 6= V , because K ∩ H = ∅ and K is (assumed to be) non-empty. So to show that V /H is one-dimensional, it suffices to show that any x, y ∈ V are linearly dependent in V /H. We may assume It suffices to find s ∈ [0, 1] with 0 = sx + s⊥y. that x ∈ K and y ∈ −K. The trick is to consider the sets S0 = {s ∈ [0, 1] : sx + s⊥y ∈ −K} and S1 = {s ∈ [0, 1] : sx + s⊥y ∈ K}, which are open (because K and −K are radially open), non-empty (because 0 ∈ S0 and 1 ∈ S1), and therefore cannot cover [0, 1] (because [0, 1] is connected). So there must be s ∈ (0, 1) such that sx + s⊥y is neither in K nor in −K, and thus sx + s⊥y ∈ H (by definition of H). Whence x and y are linearly dependent in V /H (since s 6= 0). (cid:3) Exercise We will use IV to prove that an ultrastrongly closed convex subset K of a von Neumann algebra A is ultraweakly closed as well. Let us first simplify the problem a bit. If K is empty, the result is trivial, so we may as well assume that K 6= ∅. Note that we must show that no net in K converges ultraweakly to any element a0 ∈ A outside K, but by considering K − a0 instead of K, we see that it suffices to show that no net in K ..72 -- 73.. 115 VI VII VIII converges ultraweakly to 0 under the assumption that 0 /∈ K. To this end we'll find an ultraweakly continuous linear map g : A → C and δ > 0 with g(k)R > δ for all k ∈ K -- if a net (kα)α in K were to converge ultraweakly to 0, then g(kα)R would converge to 0 as well, which is impossible. 1. Show that there is an np-map ω : A → C and ε > 0 with kkkω > ε for all k ∈ K. (Hint: use that K is ultrastrongly closed). 2. Show that B := {b ∈ A : kbkω < ε} is convex, radially open, B ∩ K = ∅. Show that B − K is convex, radially open, and 0 /∈ B − K. 3. Use IV to show that there is an R-linear map f : A → R with f (b) < f (k) for all b ∈ B and k ∈ K. Show that f can be extended to a C-linear map g : A → C by g(a) = f (a) − if (ia) for all a ∈ A . 4. Show that f (b) 6 f (k) and g(b) 6 2f (k) for all b ∈ B and k ∈ K. (Hint: b ∈ B =⇒ −b ∈ B.) Conclude that g is ultraweakly continuous (using 72 XI and K 6= ∅). 5. It remains to be shown that there is δ > 0 with f (k) ≡ g(k)R > δ for all k ∈ K. Show that in fact there is b0 ∈ B with f (b0) > 0, and that f (k) > f (b0) > 0 for all k ∈ K. 3.3.2 Kaplansky's Density Theorem 74 II Proposition Let A be a von Neumann algebra, and let f : R → R be a contin- uous map with f (t) = O(t), that is, there are n ∈ N and b ∈ [0,∞) such that f (t) 6 b t for all t ∈ R with t > n. Then the map a 7→ f (a), AR → AR, see 28 II, is ultrastrongly continuous. Proof (An adaptation of Lemma 44.2 from [16].) Let S denote the set of all continuous g : R → R such that a 7→ g(a), AR → AR is ultrastrongly continuous. We must show that f ∈ S. Let us first make some general observations. The identity map t 7→ t is in S, any constant function is in S, and S is closed under addition, and scalar multiplication. In particular, any affine transformation (t 7→ at + b) is in S. Moreover, we have g ◦ h ∈ S when g, h ∈ S, and also gh ∈ S provided that g is bounded. Finally, S is closed with respect to uniform convergence. t2 1+t2 one can see from the remarks above 1+t2 is in S -- here we use that t 7→ 1+t2 is bounded. In other words, we may assume without loss of generality, that it suffices to show that t 7→ f (t) f (t) that f vanishes at infinity, i.e. limt→∞ f (t) = 0. Suppose for the moment that there is e ∈ S, e 6= 0, which vanishes at infinity. Let a, b ∈ R. Then ea,b : R → R, t 7→ e(at + b) -- an affine transformation Now, as f (t) = f (t) 1 1 1+t2 + f (t) t followed by e -- is also in S, vanishes at infinity, and can be extended to a continuous real-valued function on the one-point compactification R∪{∞} of R (by defining ea,b(∞) = 0). It is easy to see that the C∗-subalgebra of C(R∪{∞}) generated by these extended ea,b's separates the points of R ∪ {∞}, and is thus C(R ∪ {∞}) itself by the Stone -- Weierstrass theorem (see 27 XX). Since f vanishes at infinity, f can be extended to an element of C(R ∪ {∞}), and can thus be obtained (by taking real parts if necessary) from the extended ea,b's and real constants via uniform limits, addition and (real scalar) multiplication. Since S contains the ea,b's and constants and is closed under these operations (acting on bounded functions), we see that f ∈ S. To complete the proof, we show that such e indeed exists. Let e, s : R → R be given by e(t) = ts(t) and s(t) = 1 1+t2 . Clearly e and s are continuous and vanish at infinity. To see that e is ultrastrongly continuous, let (bα)α be a net of self-adjoint elements of A which converges ultrastrongly to a ∈ AR, and let ω : A → C be an npu-map. Unfolding the definitions of e and s yields the following equality. e(bα) − e(a) = s(bα) (bα − a) s(a) − e(bα) (bα − a) e(a). Since ks(bα)k 6 1, we have ks(bα)(bα − a)s(a)kω 6 k(bα − a)s(a)kω ≡ kbα − aks(a)∗ω. Similarly, since ke(bα)k 6 1, we get ke(bα) − e(a)kω 6 kbα − aks(a)∗ω + kbα − ake(a)∗ω. Thus e(bα) converges ultrastrongly to e(a), and so e is ultrastrongly continuous. (cid:3) Corollary Given a von Neumann algebra A the map a 7→ a : AR → AR is ultrastrongly continuous. Kaplansky's Density Theorem Let b be an element of a von Neumann alge- bra B which is the ultrastrong limit of a net of elements from a C∗-subalgebra A of B. Then b is the ultrastrong limit of a net (aα)α in A with kaαk 6 kbk for all α. Moreover, III IV 1. if b is self-adjoint, then the aα can be chosen to be self-adjoint as well; 2. if b is positive, then the aα can be chosen to be positive as well, and 3. if b is an effect, then the aα can be chosen to be effects as well. Proof Let (aα)α be a net in A that converges ultrastrongly to b. V Assume for the moment that b is self-adjoint. Then (aα)R converges ultra- weakly (but perhaps not ultrastrongly) to bR = b as α → ∞, and so b is in the ultraweak closure of the convex set AR. Since the ultraweak and ultrastrong ..73 -- 74.. 117 closure of convex subsets of A coincide (by 73 VIII), we see that b is also the ultra- strong limit of some net (a′α)α in AR. Since the map −kbk∨(· )∧kbk : BR → BR is ultrastrongly continuous by I we see that a′′α := −kbk ∨ a′α ∧ kbk gives a net (a′′α)α in [−kbk,kbk]A that converges ultrastrongly to b. If we assume in addition that b is positive, then a′′′α := (a′′α)+ gives a net (a′′′α )α in [0,kbk]A that converges ultrastrongly to b+ = b, because the map (· )+ : BR → BR is ultrastrongly continuous by I. Note that if b is an effect, then so are the a′′′α . This takes care of all the special cases. The general case in which b is an a∗ von Neumann algebra M2(B) is self-adjoint, and the ultrastrong limit of the net arbitrary element of B requires a trick: since the element B := (cid:0) 0 b b∗ 0(cid:1) of the (cid:0) 0 aα α 0 (cid:1) from the C∗-subalgebra M2(A ) of M2(B), there is, as we've just seen, a net (Aα)α in M2(A ) that converges ultrastrongly to B with kAαk 6 kBk ≡ kbk for all α. Since the upper-right entries (Aα)12 will then converge ultrastrongly to B12 ≡ b as α → ∞, and k(Aα)12k 6 kAαk 6 kbk for all α, we are done. (cid:3) VI Corollary Given ε > 0 and an ultraweakly dense ∗-subalgebra S of a von Neumann algebra A each element a of A is the ultrastrong limit of a net (sα)α from S with ksαk 6 kak(1 + ε) for all α. VII Proof As the norm closure C of S in A is an ultraweakly (and thus by 73 VIII ultrastrongly) dense C∗-subalgebra of A , the element a of A is by IV the ultrastrong limit of net (cα)α∈D in C with kcαk 6 kak for all α. Each element cα is in its turn the norm (and thus ultrastrong) limit of a sequence sα1, sα2, . . . in S , and if we choose the sαn such that kcα − sαnk 6 2−n, then sαn converge ultrastrongly to b as D × N ∋ (α, n) → ∞. Finally, since limn ksαnk = kcαk 6 kck 6 (1 + ε)kck we have ksαnk 6 (1 + ε)kck for sufficiently large n, and thus for all n if we replace (sαn)n by the appropriate subsequence. (cid:3) 3.3.3 Closedness of Subalgebras 75 Recall that according to our definition (42 V) a von Neumann subalgebra B of a von Neumann algebra A is a C∗-subalgebra of A which is closed under suprema of bounded directed sets of self-adjoint elements. We will show that such B is ultrastrongly closed in A . II III Lemma Let B be a von Neumann subalgebra of a von Neumann algebra A . Let ω0, ω1 : A → C be npu-maps, which are separated by a net (bα)α of effects of B in the sense that limα ω0(bα) = 0 and limα ω1(b⊥α ) = 0. Then ω0 and ω1 are separated by a projection q of B in the sense that ω0(q) = 0 = ω1(q⊥). Proof (Based on Lemma 45.3 and Theorem 45.6 of [16].) Note that it suffices to find an effect a in B with ω0(a) = 0 = ω1(a⊥), because then ω0(⌈a⌉) = 0 = ω1(⌈a⌉⊥) by 60 I and ⌈a⌉ ∈ B. Note that we can find a subsequence (bn)n of (bα)α such that ω0(bn) 6 n−12−n and ω1(b⊥n ) 6 n−1 for all n. For n < m, define anm = (1 +Pm k=n kbk)−1 Pm k=n kbk. Since we have seen in 25 II that the map d 7→ (1 + d)−1d is order preserving (on B+), we have 0 6 anm 6 1 2 and we get the formation a12 6 a13 6 a14 6 ··· 6 a1 , 6 6 a23 6 a24 6 6 a34 6 ··· ··· . . . 6 6 a2 6 6 a3 6 ... 6 a k=n kbk, we get ω0(anm) 6 where an :=Wm>n anm and a :=Vn an. We'll prove that ω0(a) = 0 = ω1(a⊥). (ω0(a) = 0) Since ω0(bn) 6 n−12−n and anm 6 Pm Pm k=n kω0(bk) 6 21−n, and so ω0(a) =VnWm>n ω0(anm) 6Vn 21−n = 0. (ω1(a⊥) = 0) Let m > n be given. Since Pm k=n kbk > mbm and d 7→ (1 + d)−1d is monotone on B+ we get anm > (1+mbm)−1mbm, and so a⊥nm 6 (1+mbm)−1. Observe that for a real number t ∈ [0, 1], we have tt⊥ > 0, and so (1 + mt)(1 + mt⊥) = 1 + m + m2tt⊥ > 1 + m. This yields the inequality (1 + mt)−1 6 (1 + m)−1(1 + mt⊥) for real numbers t ∈ [0, 1]. The corresponding inequality for effects of a C∗-algebra (obtained via Gelfand's representation theorem, 27 XXVII) gives us ω1(a⊥nm) 6 ω1((1 + mbm)−1) 6 (1 + m)−1(1 + mω1(b⊥m)) 6 2 m . Hence ω1(a⊥n ) = 1+m , where we have used that ω1(b⊥m) 6 1 Lemma Let B be a von Neumann subalgebra of a von Neumann algebra A . Let p be a projection of A , which is the ultrastrong limit of a net in B. 1+m = 0 for all n, and so ω1(a⊥) =Wn ω1(a⊥n ) = 0. (cid:3) Vm>n ω1(a⊥nm) 6Vm>n For all npu-maps ω0, ω1 : A → C with ω0(p) = 0 = ω1(p⊥) there is a projection q of B with ω0(q) = 0 = ω1(q⊥). Proof Let (bα)α be a net in B which converges ultrastrongly to p. We may assume that all bα are effects by Kaplansky's density theorem (74 IV). Note that (ω0(bα))α converges to ω0(p) ≡ 0, and (ω1(b⊥α ))α converges to ω1(p⊥) ≡ 0. Now apply II. (cid:3) 2 IV V VI VII ..74 -- 75.. 119 VIII Theorem A von Neumann subalgebra B of a von Neumann algebra A is ultrastrongly and ultraweakly closed. IX Proof It suffices to show that B is ultrastrongly closed, because then, by 73 VIII, B will be ultraweakly closed as well. Let p be a projection of A which is the ultrastrong limit of a net from B. It suffices to show that p ∈ B, because the ultrastrong closure of B being a von Neumann subalgebra of A is generated by its projections, see 65 IV. Note that given an np-map ω : A → C, the carrier ⌈ω⌉ of ω need not be equal to the carrier of ω restricted to B, which we'll therefore denote by ⌈ω⌉B; but we do have ⌈ω⌉ 6 ⌈ω⌉B. Then by 66 IV Wω1 ⌈ω1⌉B > Wω1 ⌈ω1⌉ = p = Vω0 ⌈ω0⌉⊥ > Vω0 ⌈ω0⌉⊥B , where ω0 ranges over np-maps ω0 : A → C with ω0(p) = 0, and ω1 ranges over np-maps ω1 : A → C with ω1(p⊥) = 0. Since for such ω0 and ω1 there is by VI a projection q in B with ω0(q) = 0 = ω1(q⊥), we get ⌈ω1⌉B 6 q 6 ⌈ω0⌉⊥B, and so Wω1 ⌈ω1⌉B 6 Vω0 ⌈ω0⌉⊥B. It follows that the inequalities in (3.1) are in fact equalities, and so p =Wω1 ⌈ω1⌉B ∈ B. 3.3.4 Completeness (3.1) (cid:3) 76 II Proposition The von Neumann algebra B(H ) of bounded operators on a Hilbert space H is ultrastrongly complete. Proof Let (Tα)α be an ultrastrongly Cauchy net in B(H ) (which must be shown to converge ultrastrongly to some operator T in B(H )). Note that given x ∈ H , the net (Tαx)α in H is norm Cauchy, because k(Tα − Tβ)xk = kTα − Tβkhx,( · )xi vanishes for sufficiently large α, β, and so we may define T x := limα Tαx, giving a map T : H → H . It is clear that T will be linear, but the question is whether T is bounded, and whether in that case (Tα)α converges ultrastrongly to T . Suppose towards a contradiction that T is not bounded. Then we can find x1, x2, . . . ∈ H with kxnk2 6 2−n and kT xnk2 > 1 for all n. Since ω := Pn hxn, (· )xni : B(H ) → C is an np-map by 38 IV, it follows that kTαk2 ω ≡ P∞n=1 kTαxnk2 converges to some positive number R. Since any partial sum PN n=1 kT xnk2 > N , we must conclude n=1 kTαxnk2 6 kTαk2 that R > N , for all natural numbers N , which is absurd. Hence T is bounded. It remains to be shown that (Tα)α converges ultrastrongly to T . So let ω : B(H ) → C be an arbitrary np-map, being of the form ω ≡Pn hxn, (· )xni for some x1, x2, . . . ∈ H withPn kxnk2 < ∞ by 39 IX. We must show that kT − Tαkω ≡ (Pn k(T − Tα)xnk2)1/2 converges to 0 as α → 0. Let ε > 0 be given, and pick α0 such that kTα − Tβkω 6 1 2√2 ε for all α, β > α0 -- this is possible because (Tα)α is ultrastrongly Cauchy. We claim that ω converges to PN kT − Tαkω 6 ε for any α > α0. Since for such α the sum ∞ Xn=N k(T − Tα)xnk2 = k(T − Tα)xnk2 + Xn=1 Xn=1 N−1 ∞ k(T − Tα)xnk2 converges (to kT −Tαk2 above is below 1 ω), we can find N such that the second term in the bound 2 ε2. The first term will also be below 1 2 ε2, because N−1 Xn=1 (cid:0) k(T −Tα)xnk2(cid:1) 1/2 6 (cid:0) N−1 Xn=1 k(T −Tβ)xnk2(cid:1) 1/2 + (cid:0) N−1 Xn=1 1/2 k(Tβ−Tα)xnk2(cid:1) for any β, and in particular for β large enough that the first term on the right- hand side above is below 1 ε. If we choose β > α0 the second term will be 2√2 below 1 ε)2 ≡ ε2 all in all. 2√2 (This reasoning is very similar to that in 6 II.) (cid:3) ε too, and we get kT − Tαk2 Hence B(H ) is ultrastrongly complete. 2 ε2 + ( 1 2√2 ε + 1 2√2 ω 6 1 1 Proposition The von Neumann algebra B(H ) of bounded operators on a Hilbert space H is bounded ultraweakly complete. Proof Let (Tα)α be a norm-bounded ultraweakly Cauchy net in B(H ). We must show that (Tα)α converges ultraweakly to some bounded operator T on H . Note that given x, y ∈ H the net (hx, Tαyi )α is Cauchy (because hx, (· )yi ≡ 4P3 k=0 ik(cid:10)ikx + y, (· )(ikx + y)(cid:11) is ultraweakly continuous), and so we may de- fine [x, y] = limα hx, Tαyi. The resulting 'form' [· , · ] : H ×H → C (see 36 IV) is bounded, because k[x, y]k 6 (supα kTαk)kxkkyk for all x, y ∈ H and supα kTαk < ∞ since (Tα)α is norm bounded. By 36 V, there is a unique bounded operator T with hx, T yi = [x, y] for all x, y ∈ H . By definition of T it is clear that limα hx, (T − Tα)xi = 0 for any x ∈ H , but it is not yet clear that (Tα)α converges ultraweakly to T . For this we must show that limα ω(T − Tα) = 0 for any np-map ω : B(H ) → C. By 39 IX, we know that such ω is of the form ω = Pn hxn, (· )xni for some x1, x2, . . . ∈ H withPn kxnk2 < ∞. Now, given N and α we easily obtain the following bound. kxnk2 ω(T − Tα) 6 hxn(T − Tα), xni + (cid:0)kTk + sup α kTαk(cid:1) Xn=N Xn=1 N−1 ∞ Since the first term of this bound converges to 0 as α → ∞, we get, for all N , III IV lim sup α ω(T − Tα) 6 (cid:0)kTk + sup α kTαk(cid:1) ∞ Xn=N kxnk2. Since the tailP∞n=N kxnk2 converges to 0 as N → ∞, lim supα ω(T − Tα) = 0. Hence ω(T ) = limα ω(Tα), and so (Tα)α converges ultraweakly to T . (cid:3) ..75 -- 76 121 77 II Theorem A von Neumann algebra A is ultrastrongly complete and bounded ultraweakly complete. Proof Let Ω be the set of all np-functionals on A . Recall from 48 IX that Ω gives an nmiu-isomorphism onto the von Neumann algebra R := Ω(A ) of operators on the Hilbert space HΩ. Since B(HΩ) is ultrastrongly complete (76 I), and R is ultrastrongly closed in B(HΩ) (see 75 VIII), we see that R is complete with respect to the ultrastrong topology of B(HΩ), but since any np-functional ω : R → C is of the form ω ≡ hx, (· )xi for some x ∈ HΩ, and therefore the ultrastrong topology on B(HΩ) coincides on R with the ultra- strong topology of R, we see that R (and therefore A ) is complete with respect to its own ultrastrong topology. Since similarly B(HΩ) is bounded ultraweakly complete (76 III), the ultraweak topology on B(HΩ) coincides on R with the ultraweak topology on R, and R is ultraweakly closed in B(HΩ) (by 75 VIII), we see that R is bounded ultraweakly complete. (cid:3) III Theorem The ball (A )1 of a von Neumann algebra A is ultraweakly compact. IV Proof Writing Ω for the set of npu-maps ω : A → C, the map κ : A → CΩ given by κ(a) = (ω(a))ω for all a ∈ A is clearly a linear homeomorphism from A with the ultraweak topology onto κ(A ) ⊆ CΩ endowed with the product topology. Since κ restricts to an isomorphism of uniform spaces (A )1 → κ( (A )1 ), and (A )1 is ultraweakly complete (being a norm-bounded ultraweakly closed subset of the bounded ultraweakly complete space A , see I), we see that κ( (A )1 ) is complete, and thus closed in CΩ. Now note that κ( (A )1 ) is a closed subset of the (by Tychonoff's theorem) compact space ((C)1)Ω, because ω(a) 6 1 for all a ∈ (A )1 and ω ∈ Ω. But then κ( (A )1 ), being a closed subset of a compact Hausdorff space, is compact, and so (A )1 (being homeomorphic to it) is compact too. (cid:3) V Proposition Given an ultraweakly dense ∗-subalgebra S of a von Neumann algebra A , any ultraweakly continuous and bounded linear map f : S → B can be extended uniquely to an ultraweakly continuous map g : A → B. Moreover, g is bounded, and in fact, kgk = kfk. VI Proof As the uniqueness of g is rather obvious we concern ourselves only with its existence. Let a ∈ A be given in order to define g(a). Let also ε > 0 be given. Note that by 74 VI there is a net (sα)α in S that converges ultrastrongly (and so ultraweakly too) to a with ksαk 6 (1 + ε)kak for all α. Now, since the net (sα)α is bounded an ultraweakly Cauchy, and f is bounded and (uniformly) ultraweakly continuous, the net (f (sα))α is bounded and ultraweakly Cauchy too, and thus converges (by I) to some element uwlimα f (sα) of B. VII Of course we'd like to define g(a) := uwlimα f (sα), but must first check that uwlimα f (s′α) = uwlimα f (sα) when (s′α)α is a second net with the same proper- ties as (sα)α. Let us for simplicity's sake assume that (s′α)α and (sα)α have the same index set -- matters can always be arranged this way. Then as the differ- VIII ence sα− s′α converges ultraweakly to 0 in A as α → ∞, uwlimα f (sα− s′α) = 0, implying that uwlimα f (sα) = uwlimα f (s′α). In this way we obtain a map g : A → B -- which is clearly linear. The map g is also bounded, because since ksαk 6 (1 + ε)kak for all α, where (sα)α and t are as before, we have kf (sα)k 6 (1 + ε)kfkkak for all α, and so kg(a)k = k uwlimα f (sα)k 6 (1 + ε)kfkkak. More precisely, kgk 6 (1 + ε)kfk, and -- as ε > 0 was arbitrary -- in fact kgk 6 kfk, and so kgk = kfk. That, finally, g is ultraweakly continuous follows by a standard but abstract argument from the fact that f is uniformly ultraweakly continuous. We'll give a concrete version of this argument here. To begin, note that it suffices to show that ω ◦ g is ultraweakly continuous at 0 where ω : B → C is an np-functional. Let ε > 0 be given. Since f is ultraweakly continuous, and thus ω◦f is too, there is δ > 0 and an np-functional ν : A → C such that ν(s) 6 δ =⇒ ω(f (s)) 6 ε for all s ∈ S . We claim that ν(a) 6 δ/2 =⇒ ω(g(a)) 6 2ε for all a ∈ A , which implies, of course, that ω ◦ g is ultraweakly continuous on 0. So let a ∈ A with ν(a) 6 δ/2 be given. Pick (as before) a bounded net (sα)α in S such that f (sα) converges to a as α → ∞, and observe that, for all α, ω(g(a)) 6 ω(g(a) − f (sα)) + ω(f (sα)) . The first term on the right-hand side above will vanish as α → ∞ (since g(a) = uwlimα f (sα)), and will thus be smaller than ε for sufficiently large α. Since limα ν(sα) = ν(a) 6 δ/2 < δ we see that for sufficiently large α we'll have ν(sα) 6 δ and with it ω(f (s)) 6 ε. Combined, we get ω(g(a)) 6 2ε, and so g is ultraweakly continuous. (cid:3) 3.4 Division Using the ultrastrong completeness of von Neumann algebras (see 77 I) we'll address the question of division: given elements a and b of a von Neumann algebra A , when is there an element c ∈ A with a = cb? Surely, such c can not always exist, because its presence implies 78 a∗a 6 B b∗b, (3.2) where B = kck2; but this turns out to be the only restriction: we'll see in 81 V that if (3.2) holds for some B ∈ [0,∞), then a = cb for some unique c ∈ A with ⌈c) 6 (b⌉, which we'll denote by a/b. The main application of this division in our work is a universal property for the map b 7→ √ab√a : A → ⌈a⌉A ⌈a⌉ where a is a positive element of a von Neumann algebra A . Indeed, we'll show that for every np-map f : B → A with f (1) 6 a there is a (unique) np-map g : B → ⌈a⌉A ⌈a⌉ with f (b) = √ag(b)√a 77 -- 78.. 123 for all b ∈ B -- by taking g(b) = √a\(f (b)/√a), see 96 V. This does not give a complete description of the map b 7→ √ab√a, though, since it shares its universal property with all the maps b 7→ c∗bc, A → ⌈a⌉A ⌈a⌉ where c ∈ A with c∗c = a, but that is a challenge for the next chapter. Returning to division again, another application is the polar decomposition of an element a of a von Neumann algebra A , see 82 I, which is simply a = (a/√a∗a)√a∗a. Before we get down to business, let us indicate the difficulty in defining a/b for a and b that obey (3.2). Surely, if b is invertible, then we could simply put a/b := ab−1; and also if b is just pseudoinvertible in the sense that b∼1b = ⌈b) and bb∼1 = (b⌉ for some b∼1 the formula a/b := ab∼1 would work. But, of course, b need not be pseudoinvertible. The ideal of b∼1 can however be ap- proximated in an appropriate sense by a formal series Pn tn (which we call an approximate pseudoinverse) so that we can take a/b := Pn atn (using ultra- strong completeness to see that the series converges.) 3.4.1 (Approximate) Pseudoinverses 79 Definition Let a be an element of a von Neumann algebra A . We'll say that a is pseudoinvertible if it has a pseudoinverse, that is, an element t of A with ta = ⌈a) = (t⌉ and at = ⌈t) = (a⌉. When such t exists, it is unique (by 60 VIII), and we'll denote it by a∼1. If a∼1 = a∗, we say that a is a partial isometry (see IV). II Lemma For elements a, t of a von Neumann algebra the following are equivalent. 1. ta is a projection, and ⌈t) = (a⌉. 2. ata = a, and ⌈t) 6 (a⌉ and (t⌉ 6 ⌈a). 3. at is a projection, and ⌈a) = (t⌉. 4. tat = t, and ⌈a) 6 (t⌉ and (a⌉ 6 ⌈t). 5. t is a pseudoinverse of a. 6. a is a pseudoinverse of t. III Proof (5 ⇐⇒ 6) is clear. For the remainder we make two loops. (1=⇒2) We have ⌈t) 6 (a⌉ by assumption, and (t⌉ = (t⌈t)⌉ = (t (a⌉⌉ = (ta⌉ = ta = ⌈ta) 6 ⌈a). Further, ata = a by 60 VIII, because tata = ta (since ta is a projection) and (ata⌉ 6 (a⌉ 6 ⌈t). (3=⇒4) follows along the same lines. (2=⇒5) We have ta = ⌈a) by 60 VIII, because ata = a = a⌈a), and (ta⌉ 6 (t⌉ 6 ⌈a). Also, at = (a⌉, (because ata = a = (a⌉ a, and ⌈at) 6 ⌈t) 6 (a⌉). Further, ⌈t) = (a⌉, because (a⌉ = at = ⌈at) 6 ⌈t) 6 (a⌉; and, similarly, ⌈a) = (t⌉. (4=⇒5) is proven by the same principles, and (5=⇒1,3) is rather obvious. (cid:3) Exercise Show that an element u of a von Neumann algebra is a partial isometry iff u∗u is a projection iff uu∗u = u iff uu∗ is a projection iff u∗uu∗ = u∗ iff u∗ is the pseudoinverse of u. (Hint: use II, or give a direct proof.) Exercise Let a and b be a elements of a von Neumann algebra A . 1. Show that a is pseudoinvertible iff a∗ is pseudoinvertible, and, in that case, IV V (a∗)∼1 = (a∼1)∗. 2. Assuming that a and b are pseudoinvertible, and (b⌉ = ⌈a), show that ab is pseudoinvertible, and (ab)∼1 = b∼1a∼1. 3. Show that a is pseudoinvertible iff a∗a is pseudoinvertible, and, in that case, a∼1 = (a∗a)∼1a∗ and (a∗a)∼1 = a∼1(a∼1)∗. Exercise Let a be a positive element of a von Neumann algebra A . VI 1. Show that a is pseudoinvertible iff a is invertible in ⌈a⌉A ⌈a⌉ iff at = ⌈a⌉ for some t ∈ A+. Show, moreover, that at = ta for such t. 2. Show that a is pseudoinvertible iff there is λ > 0 with λ⌈a⌉ 6 a. 3. Assume that a is pseudoinvertible. Show that (cid:6)a∼1(cid:7) = ⌈a⌉. Show that if b ∈ A commutes with a, then b commutes with a∼1. (In other words, a∼1 ∈ {a}(cid:3)(cid:3).) 4. Show that c∼1 6 b∼1 when b 6 c are pseudoinvertible positive commuting elements of A . (The statement is still true without the requirement that b and c commute, but also much harder to prove.) 5. Show that (0, 0, 1, 1 2 , 1 3 , . . . ) is not pseudoinvertible in ℓ∞(N). Remark Note that the obvious candidate for the pseudoinverse of (0, 0, 1, 1 from ℓ∞(N) being (0, 0, 1, 2, 3, . . . ) is not bounded, and therefore not an element of ℓ∞(N). We can nevertheless approximate (0, 0, 1, 2, 3, . . . ) by the elements 2 , 1 3 , . . . ) 80 (0, 0, 1, 0, 0, . . . ), (0, 0, 1, 2, 0, . . . ), . . . of ℓ∞(N) forming what we will call "approximate pseudoinverse" for (0, 0, 1, 1 2 , 1 That this can also be done for an arbitrary element of a von Neumann algebra is what we'll see next. 3 , . . . ). ..78 -- 80.. 125 II Definition An approximate pseudoinverse of an element a of a von Neumann al- gebra A is a sequence t1, t2, . . . of elements of A such that t1a, t2a, . . . , at1, at2, . . . III are projections with Pn tna = ⌈a) =Pn (tn⌉ and Pn atn = (a⌉ =Pn ⌈tn). Exercise Let b be an element of a von Neumann algebra A , and let t1, t2, . . . be an approximate pseudoinverse of b∗b. Show that t1b∗, t2b∗, . . . is an approx- imate pseudoinverse of b. IV Theorem Every element a of a von Neumann algebra A has an approximate V pseudoinverse. Proof By III, it suffices to consider the case that a is positive. When a = 0 the sequence 0, 0, 0, . . . clearly yields an approximate pseudoinverse for a, so let us disregard this case, and assume that a is positive and non-zero. 2 6 a − 1 1 n+1 en 6 aen 6 1 n qn 6 aqn for all n > 0. n )+⌉ -- and picturing it as the places where a > 1 n )+, and thus ⌈a⌉ =Sn(cid:6)(a − 1 the places where projections e1, e2, . . . involving the facts that Note that a − 1 6 a − 1 and so does (a − 1)+ 6 (a − 1 n ), so that a =Wn(a− 1 toWn(a− 1 Writing qn = ⌈(a − 1 n )qn = (a− 1 we have (a− 1 of a von Neumann algebra, by 59 IV), and so 1 3 6 ··· converges in the norm to a ≡ a+, 2 )+ 6 . . . , which converges also ultraweakly n )+(cid:7) by 56 XVII. n -- n )+ > 0 (because b ⌈b+⌉ = b+ for a positive element b Writing en = qn+1 − qn for all n (taking q0 := 0) -- and thinking of it as n -- we get a sequence of (pairwise orthogonal) in {a}(cid:3)(cid:3) with Pn en = ⌈a⌉. By an easy computation n en. Indeed, on the one hand aen = enaen 6 kaken (as en ∈ {a}(cid:3)(cid:3)) and so ⌈aen⌉ 6 ⌈kaken⌉ = en (using here that n+1 en⌉ 6 ⌈aen⌉. In kak 6= 0), while on the other hand, particular, n+1 en 6 aen, so that aen is pseudoinvertible (by 79 VI). Writing tn := (aen)∼1, we have ⌈tn⌉ = en (since ⌈aen⌉ = en). Then tna = tn ⌈tn⌉ a = tnena = ⌈aen⌉ = en, and similarly, atn = en, so that Pn atn = Pn tna = Pn en = ⌈a⌉ = Pn ⌈tn⌉, making t1, t2, . . . an approxi- mate pseudoinverse of a. (cid:3) We claim that ⌈aen⌉ = ⌈en⌉ for any n. 1 n+1 en 6 aen gives en ≡ ⌈ 1 n and aqn 6 aqn+1, we get 1 n+1 ⌈aen⌉ = 1 1 n+1 6 a < 1 1 n+1 6 1 3.4.2 Division 81 II Definition Let b be an element of a von Neumann algebra A , and let a be an element of A b -- so a ≡ cb for some c ∈ A . We denote by a/b the (by 60 VIII) unique element c of A (b⌉ with a = cb, and, dually, given an element a of bA we denote by b\a the unique element c of ⌈b)A with a = bc. Exercise Let a and b be elements of a von Neumann algebra A . 1. Show that c/b is an element of (c⌉A (b⌉ for every element c of bA . 2. Show that (ab)/b = a (b⌉ and b\(ba) = ⌈b) a. 3. Let c be an element of aA b. Show that a\c ∈ A b, and c/b ∈ aA , and (a\c)/b = a\(c/b) =: a\c/b. Show that a\c/b is the unique element d of ⌈a) A (b⌉ with c = adb. 4. Let c be an element of A b and let d be an element of aA . Show that dc ∈ aA b, and a\(dc)/b = (a\d) (c/b). 5. Let c be an element of A b. Show that c∗ ∈ b∗A and b∗\c∗ = (c/b)∗. Lemma Given elements a and b of a von Neumann algebra A with a∗a 6 b∗b we have a ∈ A b. Moreover, given an approximate pseudoinverse t1, t2, . . . of b, the series Pn atn converges ultrastrongly to a/b, and uniformly so in a. Proof To show that PN show that (PN (PN n=M atn )∗ PN n=0 atn converges ultrastrongly as N → ∞ it suffices to n=0 atn )N is ultrastrongly Cauchy (because A is ultrastrongly complete, by 77 I). To this end, note that n=M tn) n=M tn) n=M atn = (PN 6 (PN = PN = PN n=M t∗n) a∗a (PN n=M t∗n) b∗b (PN n,m=M t∗nb∗btm m=M btm, where we've used that bt1, bt2, . . . are pairwise orthogonal projections -- but m=M btm" that gave the bound. n=0 atn is ultrastrongly Cauchy, and there- fore converges ultrastrongly -- and even uniformly so in a, because "a" does then the seriesP∞n=0 btm converges ultraweakly by 56 XVIII. This, coupled with the inequality above, gives us thatPN not appear in the expression "PN Define c := P∞n=0 atn. Since a∗a 6 b∗b, we have ⌈a) 6 ⌈b), and so a = a⌈b) = aPn tnb = Pn atnb = cb. So to get c = a/b we only need to prove that ⌈c) 6 (b⌉, that is, c (b⌉ = c. To this end, recall that Pn ⌈tn) = (b⌉, so that ⌈tn) 6 (b⌉, and tn (b⌉ = tn, which implies that atn (b⌉ = atn, and so c (b⌉ =Pn atn (b⌉ =Pn atn = c. (cid:3) 1. Let λ > 0 be given, and recall that (A )λ = {c ∈ A : kck 6 λ}. Show that a is in (A )λb iff a∗a 6 λ2b∗b, and then ka/bk 6 λ. (Compare this with "Douglas' Lemma" from [18].) Exercise Let a and b be elements of a von Neumann algebra A . 2. Show that a ∈ A (b⌉ need not entail that a ∈ A b. ..80 -- 81.. 127 III IV V VI Exercise Let b be an element of a von Neumann algebra A . 1. Let a be a positive element of A , and let λ > 0. Show that a ∈ b∗(A )λb iff a 6 λb∗b, and then kb∗\a/bk 6 λ. 2. Show that b∗\a/b is positive for every positive element a of b∗A b. (Hint: prove that (b∗\√a) (√a/b) = b∗\a/b.) VII Exercise Given elements b and c of a von Neumann algebra A , an approximate pseudoinverse t1, t2, . . . of b, and an approximate pseudoinverse of s1, s2, . . . of c, m=1 tm), converges ultrastrongly to c\a/b as N → ∞ show that (PN (and uniformly so) for a ∈ c(A )1b. n=1 sn) a (PN VIII Exercise Show that for positive elements a and b of a von Neumann algebra A , the following are equivalent. 1. a 6 λb for some λ > 0; 2. a = √bc√b for some positive c ∈ A . In that case, there is a unique c ∈ A+ with a = √bc√b and ⌈c⌉ 6 ⌈b⌉. Moreover, if t1, t2, . . . is an approximate pseudoinverse of √b, then Pm,n tmatn converges ultraweakly to such c. IX Lemma Given elements b and c of a von Neumann algebra A the maps a 7→ a/b : (A )1b → A and a 7→ c\a/b : c(A )1b → A are ultrastrongly continuous (where (A )1 is the unit ball). X Proof By III the series Pn atn converges ultraweakly to a/b, where t1, t2, . . . is an approximate pseudoinverse of b, and in fact uniformly so for a ∈ (A )1b (be- cause a∗a 6 b∗b for such a). Since a 7→PN n=1 atn, (A )1b → A is ultrastrongly continuous (by 45 IV) -- and the uniform limit of continuous functions is con- tinuous -- we see that a 7→ a/b, (A )1b → A is ultrastrongly continuous. It follows that (· )/b : c(A )1b → c(A )1 and c\(· ) : c(A )1 → A are ultrastrongly continuous; as must be their composition c\ · /b : c(A )1b → A . (cid:3) XI Remark The map a 7→ a/b might not give an ultrastrongly continuous map on 2 , 1 3 , . . . ) in ℓ∞(N) . . . , which is not ultrastrongly the larger domain A b, because, for example, upon applying (· )/(1, 1 to the ultrastrongly Cauchy sequence (1, 0, 0, . . . ), (1, 1, 0, . . . ), we get the sequence (1, 0, 0, . . . ), (1, 2, 0, . . . ), Cauchy. . . . 3.4.3 Polar Decomposition Proposition (Polar Decomposition) Any element a of a von Neumann alge- bra A can be uniquely written as a = [a]√a∗a, where [a] is an element of A ⌈a). Moreover, 1. [a] is partial isometry with [a]∗[a] = ⌈a∗a⌉ ≡ ⌈a) and [a][a]∗ = ⌈aa∗⌉ ≡ (a⌉, 2. and [a∗] = [a]∗, so that √aa∗[a] = a = [a]√a∗a. Proof Since a∗a 6 √a∗a√a∗a, the existence and uniqueness of an element [a] of A with a = [a]√a∗a and ⌈ [a] ) 6 ⌈a) ≡(cid:0)√a∗a(cid:7) is provided by 81 V, and we get ( [a]⌉ 6 (a⌉ to boot! Note that [a]∗[a] = ⌈a∗a⌉, by 60 VIII, because √a∗a [a]∗[a]√a∗a = a∗a = √a∗a ⌈a∗a⌉ √a∗a, and ⌈ [a]∗[a]⌉ 6 ⌈a) =(cid:6)√a∗a(cid:7). In particular, [a] is a partial isometry (by 79 IV). Let us prove that [a][a]∗ = (a⌉. Note that [a][a]∗ is a projection (because [a] is a partial isometry, by 79 IV). We already know that [a][a]∗ = ( [a]⌉ 6 (a⌉. Concerning the other direction, aa∗ = [a]√a∗a√a∗a[a]∗ = [a] a∗a [a]∗, so that (a⌉ = ⌈aa∗⌉ = ⌈ [a]a∗a[a]∗ ⌉ 6(cid:6)kak2[a][a]∗(cid:7) = ⌈[a][a]∗⌉ 6 [a][a]∗. To prove that a = √aa∗[a], we'll first show that √aa∗ = [a]√a∗a[a]∗. Indeed, since [a]∗[a] = (cid:6)√a∗a(cid:7), we have [a]√a∗a[a]∗[a]√a∗a[a]∗ = [a]√a∗a√a∗a[a]∗ = aa∗ -- now take the square root. It follows that √aa∗[a] = [a]√a∗a[a]∗[a] = [a]√a∗a = a. Finally, upon applying (· )∗, we see that a∗ = [a]∗√aa∗, and thus [a∗] = [a]∗, by uniqueness of [a∗], because ⌈ [a]∗ ) = ( [a]⌉ = (a⌉ = ⌈a∗). (cid:3) Recall from 68 I that the least central projection ⌈⌈e⌉⌉ above a projection e of a von Neumann algebra A is given by ⌈⌈e⌉⌉ = Sa∈A ⌈a∗ea⌉. Using the polar decomposition we can give a more economical description of ⌈⌈e⌉⌉, see V. Proposition Given projections e′ and e of a von Neumann algebra A , the following are equivalent. 82 II 83 II 1. e′ = ⌈a∗ea⌉ for some a ∈ A ; 2. e′ = ⌈a) and (a⌉ 6 e for some a ∈ A ; 3. e′ = u∗u and uu∗ 6 e for some partial isometry u. In that case we write e′ . e (and say e′ is Murray -- von Neumann below e). Proof That 3 implies 2 is clear. (2⇒1) Since (a⌉ 6 e, we have ea = a, and so ⌈a∗ea⌉ = ⌈a∗a⌉ = ⌈a) = e′. (1⇒3) By the polar decomposition (see 82 I) we get a partial isometry u := [ea] for which u∗u = [ea]∗[ea] = ⌈(ea)∗ea⌉ = e′ and uu∗ = ⌈eaa∗e⌉ 6 e. (cid:3) III ..81 -- 83.. 129 IV Exercise Show that . preorders the projections of a von Neumann algebra. V Lemma Given a projection e of a von Neumann algebra A there is a family VI Proof Let (ei)i be a maximal set of non-zero pairwise orthogonal projections Let ui be a partial isometry with u∗i ui = ei and uiu∗i 6 e. Since ei = u∗i ui = (ei)i of non-zero projections with ⌈⌈e⌉⌉ =Pi ei, and ei . e for all i. in A with ei . e for all i. Our goal is to show that Pi ei ≡Si ei = ⌈⌈e⌉⌉. u∗i uiu∗i ui 6 u∗i eui 6Sa∈A ⌈a∗ea⌉ = ⌈⌈e⌉⌉, we have Si ei 6 ⌈⌈e⌉⌉. Suppose that Si ei < ⌈⌈e⌉⌉ (towards a contradiction). Then since p := ⌈⌈e⌉⌉ −Si ei is a non-zero projection, and p = p ⌈⌈e⌉⌉ p = Sa∈A ⌈p ⌈a∗ea⌉ p⌉ = Sa∈A ⌈(eap)∗eap⌉, there must be a ∈ A with (eap)∗eap 6= 0. The polar decomposition (see 82 I) of eap gives us a partial isometry u := [eap] with uu∗ = ⌈eap(eap)∗⌉ = ⌈eapa∗e⌉ 6 e and u∗u = ⌈(eap)∗eap⌉ 6 p, so that u∗u is a non-zero projection, orthogonal to all ei with u∗u . e. In other words, e could have been added to (ei)i, contradicting its maximality. Hence Si ei = ⌈⌈e⌉⌉. (cid:3) Using 83 II we can classify all finite-dimensional C∗-algebras. 84 II Theorem Any finite-dimensional C∗-algebra A is a direct sum of full matrix III algebras, that is, A ∼=Lm MNm for some N1, . . . , NM ∈ N. Proof Let e1, . . . , eN be a basis for A . We'll first show that A is a von Neumann algebra, and for this we'll need the fact that the unit ball (A )1 is compact with respect to the norm on A . For this it suffices to show that k · k is equivalent to the norm k · k′ on A given by kak′ =Pn zn for all a ≡Pn znen where z1, . . . , zN ∈ C, (because the unit k · k′-ball is clearly compact being homeomorphic to the unit ball of CN .) Since for such a ≡Pn znen we have we see that a 7→ a : A → A is continuous from k · k′ to k · k. For the converse it suffices to show that fm : a ≡Pn zn 7→ zm, A → C is bounded with respect to k · k, because then kak 6 Pn znkenk 6 Pn zn supn kenk = kak′ supn kenk kak′ ≡ kPn fn(a)enk′ 6 Pn fn(a) 6 (Pn kfnk)kak. In fact, we'll show that any linear functional on A is bounded. Since the bounded linear functionals form a linear subspace A ∗ of N -dimensional vector space of all linear functionals on A it suffices to show that A ∗ has dimension N . So let f1, . . . , fM be a basis for A ∗; we must show that N 6 M . Since the states of A (see 22 VIII) and thus all linear functionals on A form a separating collection, the functionals f1, . . . , fN form a separating set too; since therefore a 7→ (f1(a), . . . , fM (a)) : A → CM is a linear injection from the N -dimensional space A to the M -dimensional space CM we get N 6 M . Whence all linear functionals on A are bounded, the norms k · k and k · k′ are equivalent, and (A )1 is norm compact. IV V VI (A is a von Neumann algebra) First we need to show that every bounded directed set D of self-adjoint elements of A has a supremum (in AR). We may assume without loss of generality that kdk 6 1 for all d ∈ D, and so D ⊆ (A )1. Since (A )1 is norm compact there is a cofinal subset D′ of D that norm converges to some a ∈ A , and thus D norm converges to a itself. It's easily seen that a is the supremum of D. Indeed, given d0 ∈ D we have d0 6 d for all d > d0, and so d0 6 limd>d0 d = a. Hence a is an upper bound for D; and if b is an upper bound for D, then d 6 b for all d ∈ D, and so a = limd d 6 b. Since in this finite-dimensional setting W D is apparently the norm limit of (d)d∈D, any positive functional f on A will mapW D to the limit of (f (d))d∈D, which is Wd∈D f (d), and so f (W D) =Wd∈D f (d). Whence every positive func- tional on A is normal; and since the positive functionals on A form a separating collection, A is a von Neumann algebra. (Reduction to a factor) Since pairwise orthogonal non-zero projections are easily seen to be linearly independent, and A is finite dimensional, every orthogonal set of projections in A is finite. In particular, any descending sequence of non- zero projections must eventually become constant. It follows that below every (central) projection p in A there is a minimal (central) projection, and even that p is the finite sum of minimal (central) projections. In particular, the projections of A . By 67 IV we know that zmA is a von Neumann algebra for unit 1 of A can be written as 1 =Pn zn where z1, . . . , zM are minimal central each m, and that A is nmiu-isomorphic to the direct sumLm zmA of these von Neumann algebras via a 7→ (zma)m. Since zm is a minimal central projection, the von Neumann algebra znA has no non-trivial central projections. (When A is a factor) Let e be a minimal projection of A (which exists by the previous discussion). Since e 6= 0, and A has no non-trivial central pro- jections, we have ⌈⌈e⌉⌉ = 1. By 83 V we have 1 ≡ ⌈⌈e⌉⌉ = Pk ek for some non-zero projections e1, . . . , eK in A with ek . e. So there are partial isome- tries u1, . . . , uK ∈ A with u∗kuk = ek and uku∗k 6 e for all k. In fact, since e is minimal, we have uku∗k = e. Thinking of uk as 0ihk define ukℓ = u∗kuℓ; we'll show that : A 7→ Pkℓ Akℓukℓ : MK → A is an miu-isomorphism. It's easy to see that is linear, involution preserving and unital. To see that is multiplicative, first note that uju∗k equals e when j = k and is zero otherwise. It follows that uijukℓ equals uiℓ when k = j and is zero otherwise. Whence (A)(B) = Pijkℓ AijuijBkℓukℓ = Piℓ(Pk AikBkℓ)uiℓ = (AB) for all matrices A, B ∈ MK, and so is multiplicative. It remains to be shown that is a bijection. To see that is injective, first note that is normal, because using the fact that is positive and thus bounded, we can show that preserves suprema of bounded directed sets in much the same way we showed that all np-functionals on A are bounded. We can thus speak of the central carrier ⌈⌈⌉⌉ of , and thus to show that is injective it suffices ..83 -- 84.. 131 84a to show that ⌈⌈⌉⌉ = 1. Since MK is a factor (see 67 II) the only alternative is ⌈⌈⌉⌉ = 0 i.e. = 0, which is clearly absurd unless A = {0} in which case we'd already be done. Hence is injective. To see that is surjective let a ∈ A with a 6= 0 be given. Since a ≡ Pk,ℓ ekaeℓ =Pk,ℓ uk1u1kauℓ1u1ℓ, and uk1 and u1ℓ are in the range of it suffices to show that u1kauℓ1 is in the range of for all k and ℓ. In other words, we may assume without loss of generality that eae = a, where e is the minimal projection in A we started with. Since e(aR)+e = (aR)+, and so on, we may assume that a is positive. By scaling, we may also assume that kak 6 1/3. Since ⌈kake − a⌉ 6 e, and e is minimal, we either have ⌈kake − a⌉ = e or ⌈kake − a⌉ = 0. The former case is impossible: indeed, if e = ⌈kake − a⌉ ≡Wn(kake − a)1/2n (see 56 I), then (kake − a)1/2n norm converges to ⌈kake − a⌉ = e (cf. IV), and so kkake − ak1/2n converges to kek = 1. Then kkake − ak = 1, while kkake − ak 6 kakkek + kak 6 2 Hence ⌈kake − a⌉ = 0, and so a = kake. In particular, a is in the range of . Whence is surjective, and thus an miu-isomorphism MN → A . (cid:3) Example Using the description of all finite-dimensional C∗-algebras from 84 II we can prove the claim made at the start of this thesis, in 20a III, that in C∗pu there's no equaliser for the maps f, g : C4 → C given by 3 , which is absurd. f (a, b, c, d) = 1 2 (a + b), and g(a, b, c, d) = 1 2 (c + d). Indeed, suppose towards a contradiction that f and g do have an equaliser e : E → C4 in C∗pu, and let S denote the set-theoretic equaliser: S := { (a, b, c, d) ∈ C4 : f (a, b, c, d) = g(a, b, c, d)} = { (a, b, c, d) ∈ C4 : a + b = c + d}. Note that the elements s ∈ S with 0 6 s 6 1 form a convex subset of C4 that is isomorphic to an octahedron -- this will be essential later. We claim that the range of e : E → C4 is simply the set-theoretic equaliser, e(E ) = S . Indeed, surely, e(E ) ⊆ S . For the other direction, let v ∈ S be given; we must find a ∈ E with e(a) = v. Since e is involution preserving, and so vR, vI ∈ S , we may assume without loss of generality that v is self-adjoint. Since v + kvk > 0, and e is unital, we may assume that v is positive too. By scaling v if necessary, we may assume also that 0 6 v 6 1. Now, to use the universal property of e : E → C4 consider the unique pu-map p : C2 → C4 given by p(1, 0) = v. Since v ∈ S we have f ◦ p = g ◦ p, and so there is a unique q : C2 → E with p = e ◦ q. Then v = p(1, 0) = e(q(1, 0)), and so e(E ) = S . The next thing to note is that e is injective, and for this it suffices to show that e is injective on [0, 1]E . So let a, b ∈ [0, 1]E with e(a) = e(b) be given; we must show that e(a) = 0. Let p, q : C2 → C4 be the unique pu-maps given by p(1, 0) = a and q(1, 0) = b, and note that e ◦ p = e ◦ q. Since equalisers are mono, we get p = q, and so a = b. Thus e : E → C4 gives a linear isomorphism from E onto the 3-dimensional linear subspace S of A , so E is 3-dimensional too. By the classification of finite-dimensional C∗-algebras, 84 II, E must be miu-isomorphic to C3. The map e : E → C4 is not only injective, but in fact bipositive (see 20 VI). Indeed, if e(a) 6 0 for some a ∈ [0, 1]E we can, as before, find q : C2 → E with e(q(1, 0)) = e(a), and so a = q(1, 0) > 0. It follows that e gives a linear order isomorphism between E and the subspace S of C4, and so [0, 1]E is as convex space isomorphic to S ∩ [0, 1]C4. This is problematic, because on the one hand the convex space [0, 1]E being a cube (because E is miu-isomorphic to C3) has eight extreme points, while on the other hand S ∩ [0, 1]C4 being an octahedron has six extreme points: a contradiction. 3.4.4 Hereditarily Atomic Von Neumann Algebras 84b We've seen in 84 II that every finite-dimensional von Neumann algebra is the product of finitely many full matrix algebras. For our purposes this class is too small, not admitting interpretations of infinite-dimensional datatypes, so we've focused on all von Neumann algebras instead. There is, however, a rather modest but very promising subclass of von Neumann algebras that does sate our desire for the infinite: following Kornell we'll call a von Neumann algebra that is the product of a possibly infinite set of full matrix algebras hereditarily atomic, see II. In his recent paper [49] Kornell develops the position that these hereditar- ily atomic von Neumann algebras are the "correct" quantum generalisation of sets, and -- which is especially relevant to our work -- observes that the category of hereditarily atomic von Neumann algebras and the nmiu-maps between them endowed with the regular tensor is monoidal coclosed (see [49], Theorem 9.1.) This will allow us to build a model of the quantum lambda calculus not only using all von Neumann algebras, but also just from the hereditarily atomic ones. Hereditarily atomic von Neumann algebras have garnered attention for a completely different reason too: Selinger observed in Example 2.7 of [70] that the effects of MN (and thus of every hereditarily atomic von Neumann algebra) form a continuous dcpo. Furber and Weaver have proven recently that the converse also holds: that every von Neumann algebra A for which [0, 1]A as dcpo is continuous, is hereditarily atomic, see Theorem III.15 of [20]. Definition A von Neumann algebra is called hereditarily atomic if it is nmiu- II isomorphic to a direct sum Li∈I MNi of possibly infinitely many MNi's. We denote by haW∗miu and haW∗cpsu the full subcategories of W∗miu and W∗cpsu, respectively, of hereditarily atomic von Neumann algebras. Proposition A von Neumann subalgebra B of a hereditarily atomic von Neu- III ..84 -- 84b.. 133 mann algebra A is itself hereditarily atomic. IV Proof Since A is hereditarily atomic, we may assume without loss of generality and cj B be finite-dimensional for every j. Note that to show that B is hereditarily atomic, it suffices to find a orthog- cj B is finite-dimensional. Indeed, then each cj B is hereditarily atomic, and so that A ≡Li∈I MNi for some family of natural numbers (Ni)i∈I . onal family of central projections (cj)j∈J in B with Pj cj = 1 such that each will be B ∼=Lj∈J cj B (see 67 IV). It's even enough to find a family of central projections (dk)k∈K in B, not necessarily orthogonal, but withSk∈K dk = 1 and each dkB finite-dimensional. Indeed, any maximal orthogonal family (cj )j∈J of non-zero central projections in B for which each cj is below some dk will have the properties thatPj∈J cj = 1 Define dj := ⌈πj ◦ e⌉ to be the carrier (see 63 I) of the inclusion e : B → A followed by the j-th projection πj : A ≡ Li∈I MNi → MNj . Since πj ◦ e is an nmiu-map, dj is a central projection by 69 IV. Since there are fewer projection in B than in A , we have ⌈πj⌉ 6 ⌈πj ◦ e⌉ ≡ dj. Now, since clearlyPi∈I ⌈πi⌉ = 1, this implies that Si∈I di = 1. Let i ∈ I be given. It remains to be shown that diB is finite-dimensional. To see this, simply note that the restriction of πi ◦ e to a map diB → MNi is an injection (by 69 IV) into the finite-dimensional space MNi, and so diB is finite-dimensional too. (cid:3) V Corollary Given nmiu-maps f, g : A → B between hereditarily atomic von Neumann algebras A and B, the von Neumann subalgebra E := { a ∈ A : f (a) = g(a)} of A is hereditarily atomic, and the inclusion e : E → A is an equaliser of f and g both in haW∗miu and haW∗cpsu. VI Remark It follows that haW∗miu is the least full subcategory of W∗miu closed under limits that contains all finite-dimensional von Neumann algebras. 3.5 Normal Functionals 85 For our study of the category of von Neumann algebras we need two more technical results concerning the normal functionals on a von Neumann algebra. The first one, that a net (bα)α in a von Neumann algebra A is (norm) bounded provided that (ω(bα))α is bounded for each np-functionals ω : A → C (see 87 VIII), ultimately follows from a type of polar decomposition for ultra- weakly linear functionals (see 86 IX). The second one, that the ultraweak topology of a von Neumann subalgebra coincides with the ultraweak topology of the surrounding space (see 89 XI), is proven using the double commutant theorem (88 VI) and requires a lot of hard work. 3.5.1 Ultraweak Boundedness To get a better handle on the normal positive functionals on a von Neumann al- gebra, we first analyse the not-necessarily-positive normal functionals in greater detail. Lemma A linear map f : A → C on a C∗-algebra A is positive iff kfk 6 f (1). Proof (Based on Theorem 4.3.2 of [47].) If f (1) = 0, then f = 0 in both cases (viz. f is positive, and kfk 6 f (1)), so we may assume that f (1) 6= 0. The problem is easily reduced farther to the case that f (1) = 1 by replacing f by f (1)−1f (noting that f (1) > 0 in both cases), so we'll assume that f (1) = 1. (f positive =⇒ kfk 6 1) This follows immediately from 34 XVI and 34 IX, but here's a concrete proof: Let a ∈ A be given. Pick λ ∈ C with λ = 1 and λf (a) > 0. Then f (a) = f (λa) = f (λa)R = f ((λa)R) 6 f (kak) = kak, because (λa)R 6 k(λa)Rk 6 kλak = kak, and f is positive. Hence kfk 6 1. (kfk 6 1 =⇒ f is positive) Let a ∈ [0, 1]A be given. To prove that f is positive, it suffices to show that f (a) > 0. Since (f (a)R)⊥ = (f (a)⊥)R 6 (cid:12)(cid:12)f (a)⊥(cid:12)(cid:12) = (cid:12)(cid:12)f (a⊥)(cid:12)(cid:12) 6 1, and therefore f (a)R > 0, we just need to show that f (a)I = 0. Indeed, since (n + 1)2(f (a)I)2 = f (bn)2 6 kbnk2 = kb∗nbnk 6 ka − f (a)Rk2 + n2(f (a)I)2, one sees that (2n + 1)(f (a)I)2 6 ka − f (a)Rk2 for all n, which is impossible un- less (f (a)I)2 = 0, that is, f (a)I = 0. (cid:3) The trick is to consider bn := (a − f (a)R) + nif (a)I. Lemma An extreme point u of the unit ball (A )1 of a C∗-algebra A is a partial isometry with (uu∗)⊥A (u∗u)⊥ = {0}. Remark The converse (viz. every such partial isometry is extreme in (A )1) also holds, but we won't need it. Proof (Based on Theorem 7.3.1 of [47].) To show u is a partial isometry it suffices to prove that u∗u is a projection. Suppose towards a contradiction that u∗u is not a projection. Then u∗u, rep- resented as continuous function (on sp(u∗u) cf. 28 II), takes neither the value 0 nor 1 on a neighbourhood of some point, and so by considering a positive con- tinuous function, which is sufficiently small but non-zero on this neighbourhood and zero elsewhere, we can find a non-zero element a of the (commutative) C∗-subalgebra generated by u∗u with 0 6 a 6 u∗u and ku∗u(1 ± a)2k 6 1, so that ku(1± a)k 6 1. Since u is extreme in (A )1, and u = 1 2 u(1− a), we get ua = 0, and so 0 6 a2 6 √au∗u√a = u∗ua = 0, which contradicts a 6= 0. 2 u(1 + a) + 1 ..84b -- 86.. 135 86 II III IV V VI VII VIII 2 (u + a) + 1 Let a ∈ (uu∗)⊥A (u∗u)⊥ be given; we must show that a = 0. Assume (without loss of generality) that kak 6 1. We'll show that ku± ak 6 1, because, since u is extreme in (A )1, u ≡ 1 2 (u − a) implies that u = u + a, and so a = 0. Note that a∗a 6 (u∗u)⊥ (because a(u∗u)⊥ = a) and u∗a = 0 (because (uu∗)⊥a = a). Thus (u ± a)∗(u ± a) = u∗u ± u∗a ± a∗u + a∗a = u∗u + a∗a 6 u∗u + (u∗u)⊥ = 1, so ku ± ak 6 1. (cid:3) IX Theorem (Polar decomposition of functionals) Every functional f : A → C on a von Neumann algebra A which is ultraweakly continuous on the unit ball (A )1 is of the form f ≡ f (uu∗(· )) = f ((· )u∗u) for some partial isometry u on A such that f (u(· )) and f ((· )u) : A → C are positive. Proof (Based on Theorem 7.3.2 of [47].) X 2 a + 1 2 f (a) + 1 2 b ∈ F for some a, b ∈ (A )1, then 1 XI We'll first show that f takes the value kfk at some extreme point u of (A )1. To begin, since (A )1 is ultraweakly compact (77 III), and f is ultraweakly continu- ous the subset { f (a) : a ∈ (A )1 } of R is compact, and therefore has a largest el- ement, which must be kfk. Thus the convex set F := { a ∈ (A )1 : f (a) = kfk } is non-empty. Since F is ultraweakly compact (being an ultraweakly closed sub- set of the ultraweakly compact (A )1), F has at least one extreme point by the Krein -- Milman Theorem (see e.g. Theorem V7.4 of [15]), say u. Note that F is a face of (A )1: if 1 2 f (b) = kfk, so f (a) = f (b) = kfk (since kfk is extreme in (C)kfk) and thus a, b ∈ F . It follows that u is not only extreme in F , but also in (A )1, so that u is an partial isometry with (uu∗)⊥A (u∗u)⊥ = {0} by VI. Note that f (u(· )) is positive by II, because kf (u(· ))k 6 kfkkuk 6 kfk = f (u) = f (u(1)). By a similar argument f ((· )u) is positive. Let a ∈ A be given. It remains to be shown that f (a) = f (uu∗a) = f (au∗u). First note that u(u∗u)⊥ = 0 (since u is an isometry) and so f (u(u∗u)⊥) = 0, that is, u∗u > ⌈f (u(· ))⌉. This entails that f (ubu∗u) = f (ub) for all b ∈ A by 63 VI, and in particular f (uu∗au∗u) = f (uu∗a). Now, since (uu∗)⊥A (u∗u)⊥ = {0}, we have uu∗au∗u+a = uu∗a+au∗u, and thus f (a) + f (uu∗a) = f (a) + f (uu∗au∗u) = f (uu∗a) + f (au∗u), which yields f (a) = f (au∗u). By a similar reasoning we get f (uu∗a) = f (a). (cid:3) XII Corollary A functional f : A → C on a von Neumann algebra A is ultraweakly XIII Proof By IX there is a partial isometry u such that f (uu∗(· )) = f and f (u(· )) is positive. Recall from 44 XV that such a positive functional f (u(· )) is normal when it is ultraweakly continuous on [0, 1]A ; which it is, because a 7→ ua is ultraweakly continuous (see 45 IV), maps [0, 1]A into (A )1, and f is ultraweakly continuous on (A )1. But then f ≡ f (uu∗(· )) being the composition of the ultraweakly continuous maps f (u(· )) and a 7→ u∗a is ultraweakly continuous on A too. (cid:3) XIV Lemma Let f : A → C be a normal functional on a von Neumann algebra A , continuous when it is ultraweakly continuous on the unit ball (A )1. and let u be a partial isometry in A such that f (u(· )) is positive, and f = f (uu∗(· )). Then kfk = f (u). (cid:3) Proof Since f (u(· )) is positive, we have kf (u(· ))k = f (u) by 34 XVI; hence kfk = kf (uu∗(· ))k 6 kf (u(· ))kku∗k ≡ f (u) 6 kfk, and thus kfk = f (u). Definition Given a von Neumann algebra A , the vector space of ultraweakly continuous linear maps f : A → C endowed with the operator norm is denoted by A∗, and called the predual of A . Remark The reason that the space A∗ is called the predual of A is the non- trivial fact due to Sakai [68] (which we don't need and therefore won't prove), that the obvious map A → (A∗)∗, where (A∗)∗ is the dual of A∗ -- the vector space of bounded linear maps A∗ → C endowed with the operator norm -- , is a surjective isometry, and so A "is" the dual of A∗, (albeit only as normed space, because (A∗)∗ doesn't come equipped with a multiplication.) We will need this: Proposition The predual A∗ of a von Neumann algebra A is complete (with respect to the operator norm). Proof Let f1, f2, . . . be a sequence in A∗ which is Cauchy with respect to the operator norm. We already know (from 4 V) that f1, f2, . . . converges to a bounded linear functional f : A → C; so we only need to prove that f is ultraweakly continuous to see that A∗ is complete, and for this, we only need to show (by 86 XII) that f is ultraweakly continuous on the unit ball (A )1 of A . So let (bα)α be a net in (A )1 which converges ultraweakly to 0; we must show that limα f (bα) = 0. Now, note that for every n and α we have the bound f (bα) 6 (f − fn)(bα) + fn(bα) 6 kf − fnk + fn(bα) . From this, and limn kf − fnk = 0, and limα fn(bα) = 0 for all n, one easily deduces that limα f (bα) = 0. Thus f is ultraweakly continuous, and so A∗ is complete. (cid:3) Note that for a self-adjoint element a of a von Neumann algebra A we have kak = supω ω(a) where ω ranges over the npsu-functionals, but that the same identity does not need to hold for arbitrary (not necessarily self-adjoint) a ∈ A . The following lemma shows that this restriction to self-adjoint elements can be lifted by letting ω range over all of A∗. Lemma We have kak = supf∈(A∗)1 f (a) for every element a of a von Neumann algebra A . Proof For the other direction, write a ≡ [a]√a∗a (see 82 I) and note that kak = k√a∗ak = supω∈Ω(cid:12)(cid:12)ω(√a∗a )(cid:12)(cid:12), where Ω is the set of npu-maps A → C (which is order separating). Let ω ∈ Ω be given. Since [a]∗a = √a∗a we have It's clear that supf∈(A∗)1 f (a) 6 kak. ..86 -- 87.. 137 XV 87 II III IV V VI VII ω(√a∗a ) = ω([a]∗a) = f (a), where f := ω([a]∗(· )) ∈ (A∗)1, and so kak = supω∈Ω ω(√a∗a ) 6 supf∈(A∗)1 f (a). (cid:3) VIII Theorem A net (bα)α in a von Neumann algebra A is norm bounded (that is, supα kbαk < ∞) provided it is ultraweakly bounded, i.e., supα ω(bα) < ∞ for every np(u)-map ω : A → C. IX Proof Note that f 7→ f (bα) gives a linear map (· )(bα) : A∗ → C with k(· )(bα)k = kbαk by VI for each α. So to prove that (bα)α is norm bounded, viz. supα kbαk ≡ supα k(· )(bα)k < ∞, it suffices to show (by the principle of uniform bounded- ness, 35 II, using that A∗ is complete, III), that supα f (bα) < ∞ for all f ∈ A∗. Since such f ∈ A∗ can be written as f ≡ P3 k=0 ikωk where ωk : A → C are np-maps (by 72 V), we see that supα f (bα) 6 P3 k=0 supα ωk(bα) < ∞, (cid:3) because (bα)α is ultraweakly bounded. Thus (bα)α is norm bounded. 3.5.2 Ultraweak Permanence 88 We turn to a subtle, and surprisingly difficult matter: it is not immediately clear that the ultraweak topology on a von Neumann subalgebra A of a von Neumann algebra B, coincides (on A ) with the ultraweak topology on B. While it is easily seen that the former is finer (that is, a net in A which converges ultraweakly in A , converges ultraweakly in B too, because any np-map ω : B → C is also an np-map restricted to A ), it is not obvious that an np-map ω : A → C can be extended to an np-map on B -- but it can, as we'll see 89 XI. We'll call this independence of the ultraweak topology from the surrounding space ultraweak permanence being not unlike the independence of the spectrum of an operator from the surrounding space known as spectral permanence (11 XXIII). It is tempting to think that the extension of an np-map ω : A → C on a von Neumann subalgebra A of a von Neumann algebra B to B is simply a matter of applying Hahn -- Banach to ω, but this approach presents two problems: it yields a normal but not necessarily positive extension of ω; and it not clear that ω is ultraweakly continuous on A (that is, whether Hahn -- Banach applies). Instead of applying general techniques we feel forced to delve deeper into the particular structure provided to us by von Neumann algebras (namely the commutant, 65 II) to show that any np-map ω : A → C on a von Neumann algebra A of bounded operators on a Hilbert space H can be extended to an np-map on B(H ), and in fact, is of the form ω ≡ Pn hxn, (· )xni for some x1, x2, . . . ∈ H , see 89 IX. Proposition Let S be a subset of a von Neumann algebra A that is closed under multiplication, involution, and contains 1. Let e be a projection in A . Then ⌈e⌉S(cid:3) =Sa∈S ⌈a∗ea⌉ is the least projection in S(cid:3) above e. (Compare this with the paragraph "Subspaces" of §2.6 of [47].) II III IV V Proof Let us first show that p := ⌈e⌉S(cid:3) is in S(cid:3). Let b ∈ S be given; we must show that pb = bp. We may may assume without loss of generality that kbk 6 1. Since b∗(· )b : A → A is normal and completely positive, and p =Sa∈S ⌈a∗ea⌉, we have b∗pb 6 ⌈b∗pb⌉ = Sa∈S ⌈b∗ ⌈a∗ea⌉ b⌉ = Sa∈S ⌈(ab)∗ e ab⌉ 6 p by 60 IX and 60 V. Applying p⊥(· )p⊥, we get p⊥b∗pbp⊥ 6 p⊥pp⊥ = 0, so that pbp⊥ = 0, and thus pbp = pb. Since similarly pb∗ = pb∗p, we get bp = pbp = pb (upon applying (· )∗) and so p ∈ S(cid:3). Note that e 6 ⌈1∗e1⌉ 6 p, because 1 ∈ S. It remains to be shown that p is the least projection in S(cid:3) above e, so let q be a projection in S(cid:3) above e. Since for a ∈ S, we have aq⊥a∗ = q⊥aa∗q⊥ 6 kak2q⊥ 6 kak2e⊥, and so a∗ea 6 kak2q we get ⌈a∗ea⌉ 6 q for all a ∈ S, and thus p =Sa∈S ⌈a∗ea⌉ 6 q. (cid:3) Exercise Show that given a vector x of Hilbert space H , and a collection S of bounded operators on H that is closed under addition, (scalar) multiplication, involution, and contains the identity operator, the following coincide. 1. ⌈xihx⌉S(cid:3), the least projection in S(cid:3) above ⌈xihx⌉; 2. (cid:6)hx, (· )xi S(cid:3)(cid:7), the carrier of the vector functional on S(cid:3) given by x; 3. Sa∈S ⌈axihax⌉; and 4. the projection on Sx. Conclude that S(cid:3)(cid:3)x = Sx. (Hint: S(cid:3)(cid:3)(cid:3) = S(cid:3).) Now consider (instead of x) an np-map ω : B(H ) → C, which we know must be of the form ω ≡Pn hxn, (· )xni (by 39 IX) and is therefore given by an element x′ ≡ (x1, x2, . . . ) of the N-fold product H ′ :=Ln 1. Show that ω(t) = hx′, ′(t)x′i, where ′ : B(H ) → B(H ′) is the nmiu- map given by ′(t)y = (tyn)n for all t ∈ B(H ) and y ∈ H ′. Prove that ′(t) = Pn P ∗n tPn, where Pn := πn : H ′ ≡ Ln 2. Let t ∈ S(cid:3)(cid:3) be given (with S as above). Show that ′(t) ∈ ′(S)(cid:3)(cid:3). H → H is H of H . the n-th projection. (Hint: first show PnaP ∗m ∈ S(cid:3) for all m, n, and a ∈ ′(S)(cid:3).) Conclude that ′(t)x′ ∈ ′(S)(cid:3)(cid:3)x′ ≡ ′(S)x′. Whence for every ε > 0 one can find a ∈ S with kt − akω 6 ε. 3. Deduce that S(cid:3)(cid:3) is contained in the ultrastrong closure of S. Double Commutant Theorem For a collection S of bounded operators on a Hilbert space H that is closed under addition, (scalar) multiplication, involu- tion, and contains the identity operator the following are the same. VI ..87 -- 88.. 139 1. S(cid:3)(cid:3), the "double commutant" of S in B(H ); 2. us-cl(S), the ultrastrong closure of S in B(H ); 3. uw-cl(S), the ultraweak closure of S in B(H ); 4. W ∗(S), the least von Neumann subalgebra of B(H ) that contains S. VII Proof (Based on Theorem 5.3.1 of [47].) Note that: us-cl(S) ⊆ uw-cl(S), because ultrastrong convergence implies ultraweak convergence; and uw-cl(S) ⊆ W ∗(S), because W ∗(S) is ultraweakly closed in B(H ) by 75 VIII; and W ∗(S) ⊆ S(cid:3)(cid:3), because S(cid:3)(cid:3) is a von Neumann subalgebra of B(H ) by 65 III; and, finally, S(cid:3)(cid:3) ⊆ us-cl(S) by V. (cid:3) VIII Exercise Show that central elements of a von Neumann algebra A of bounded operators on a Hilbert space H coincide with the central elements of the com- mutant A (cid:3), that is, Z(A ) = Z(A (cid:3)). (Hint: A (cid:3)(cid:3) = A by VI.) 89 II Neumann algebra B. IX Deduce that ⌈⌈fA ⌉⌉ = (cid:6)(cid:6)fA (cid:3)(cid:7)(cid:7) for every np-map f : B(H ) → B into a von Lemma Let ω : A → C be an np-map on a von Neumann algebra A , which is represented by nmiu-maps : A → B(H ) and π : A → B(K ) on Hilbert spaces H and K . If hx, (· )xi = ω = hy, π(· )yi for some x ∈ H and y ∈ K , then there is a bounded operator U : K → H for which U U∗ is the projection on (A )x, U∗U is the projection on π(A )y, and U π(a) = (a)U for all a ∈ A . Proof (Compare this with Proposition 4.5.3 of [47].) Since k(a)xk2 = hx, (a∗a)xi = ω(a∗a) = hy, π(a∗a)yi = kπ(a)yk2 for all a ∈ A , there is a unique bounded operator V : π(A )y → (A )x with V π(a)y = (a)x for all a ∈ A . A moment's thought reveals that V is a unitary (and so V ∗V = 1 and V V ∗ = 1.) Now, define U := EV F ∗ where E : (A )x → H and F : π(A )y → K are the inclusions (and so E∗E = 1 and F ∗F = 1). Then U U∗ = EV F ∗F V ∗E∗ = EV V ∗E∗ = EE∗ is the projec- tion onto (A )x, and U U∗ = F F ∗ is the projection onto π(A )y. Let a ∈ A be given. It remains to be shown that U π(a) = (a)U . To this end, observe that V F ∗π(a)F = E∗(a)EV (because these two bounded linear maps are easily seen to agree on the dense subset π(A )y of π(A )y); and (a)E = EE∗(a)E (because (a) maps (A )x into (A )x); and similarly (a∗)F = F F ∗(a∗)F , so that F ∗(a) = F ∗(a)F F ∗ (upon application of the (· )∗). By these observations, U π(a) = EV F ∗π(a) = EV F ∗π(a)F F ∗ = EE∗(a)EV F ∗ = (a)EV F ∗ = (a)U . (cid:3) III Exercise It is not too difficult to see that the (ultraweak) sum Pi ui of a col- lection (ui)i of partial isometries from some von Neumann algebra is again a partial isometry, provided that the initial projections u∗i ui are pairwise orthog- onal, and the final projections uiu∗i are pairwise orthogonal. In this exercise, you'll establish a similar result, but for partial isometries between two different Hilbert spaces, and avoiding the use of an analogue of the ultraweak topology for such operators. Let H and K be Hilbert spaces, and let Ui : H → K be a bounded operator for every element i from some set I. Assume that the operators U∗i Ui are pair- wise orthogonal projections in B(K ), and that UiU∗i are pairwise orthogonal projections in B(H ). 1. Let x ∈ H and y ∈ K be given. Show that hx, Uiyi 6 kU∗i xkkUiyk for each i (perhaps by first proving that Ui = UiU∗i Ui). Show that Pi kUiyk2 6 kyk2 and Pi kU∗i xk2 6 kxk2, and deduce from this that Pi hx, Uiyi 6 kxkkyk. Now use 36 V to show that there is a bounded operator U : K → H with hx, U yi =Pi hx, Uiyi for all x ∈ H and y ∈ K . 2. Show that U∗i Uj = 0 when i 6= j. Deduce from this that U∗U =Pi U∗i Ui. Prove that U U∗ =Pi UiU∗i . Lemma Let Ω be a collection of np-maps ω : A → C on a von Neumann algebra A whose central carriers, ⌈⌈ω⌉⌉, are pairwise orthogonal to one another, and let H and K be Hilbert spaces on which A is represented such that each ω ∈ Ω is given by vectors xω ∈ H and yω ∈ K , that is, hxω, (· )xωi = ω = hyω, π(· )yωi, where : A → B(H ) and π : A → B(K ) are nmiu-maps. Then there is a bounded operator U : K → H which intertwines π and in the sense that U π(a) = (a)U for all a ∈ A such that U∗U is a projection in π(A )(cid:3) with ⌈⌈U∗U⌉⌉π(A )(cid:3) = π(Pω ⌈⌈ω⌉⌉), and U U∗ is projection in (A )(cid:3) with ⌈⌈U U∗⌉⌉(A )(cid:3) = (Pω ⌈⌈ω⌉⌉). Proof Given ω ∈ Ω, let σω : (A ) → C and σ′ω : (A )(cid:3) → C denote the restric- tions of the vector functional hxω, (· )xωi : B(H ) → C, and let τω : π(A ) → C and τ′ω : π(A )(cid:3) → C be similar restrictions of hyω, (· )yωi. We already know (by I and 88 IV) that there is a bounded operator Uω : K → H with U∗ωUω = ⌈τ′ω⌉, UωU∗ω = ⌈σ′ω⌉, and Uωπ(a) = (a)Uω for all a ∈ A . We'll combine these Uωs into one operator U using III, but for this we must verify that the projections UωU∗ω = ⌈σ′ω⌉ are pairwise orthogonal, and that the projections U∗ωUω are pairwise orthogonal too. To this end note that ⌈⌈σω⌉⌉ = ⌈⌈σ′ω⌉⌉ by 88 IX. Thus, since the projections ⌈⌈ω⌉⌉ are orthogonal to one another, and ⌈σ′ω⌉ 6 ⌈⌈σ′ω⌉⌉ = ⌈⌈σω⌉⌉ = (⌈⌈ω⌉⌉), we see that the projec- tions UωU∗ω ≡ ⌈σ′ω⌉ are indeed pairwise orthogonal. Since for a similar reason the projections U∗ωUω ≡ ⌈τ′ω⌉ are pairwise orthogonal too, there is by III a ..88 -- 89.. 141 IV V VI bounded operator U : K → H with U∗U =Pω U∗ωUω, U U∗ =Pω UωU∗ω, and hx, U yi =Pω hx, Uωyi for all x ∈ H and y ∈ K . Let us check that U has the desired properties. To begin, since the projec- tions ⌈⌈UωU∗ω⌉⌉ = ⌈⌈σ′ω⌉⌉ = (⌈⌈ω⌉⌉) are pairwise orthogonal, we have ⌈⌈U U∗⌉⌉ = Pω ⌈⌈UωU∗ω⌉⌉ = (Pω ⌈⌈ω⌉⌉) by 68 IV and 56 XVIII. Similarly, ⌈⌈U∗U⌉⌉ = π(Pω ⌈⌈ω⌉⌉). Finally, given a ∈ A we have U π(a) = (a)U , because hx, U π(a)yi = Pω hx, Uωπ(a)yi = Pω hx, (a)Uωyi = Pω h(a)∗x, Uωyi = h(a)∗x, U yi = hx, (a)U yi for all x ∈ H and y ∈ K . (cid:3) VII Corollary Let A be a von Neumann of bounded operators on some Hilbert space H , and let : A → B(H ) denote the inclusion. Let Ω be the collection of all np-maps A → C, and let Ω : A → B(HΩ) be as in 30 IX. There is a bounded operator U : HΩ → H such that U∗U is a projection in Ω(A )(cid:3) with ⌈⌈U∗U⌉⌉Ω(A )(cid:3) = 1 and U Ω(a) = (a)U for all a ∈ A . VIII Proof Let {xi}i be a maximal set of vectors in H such that the central car- riers ⌈⌈ωi⌉⌉ of the corresponding vector functionals ωi := hxi, (· )xii on A are pairwise orthogonal; so that we'll have Pi ⌈⌈ωi⌉⌉ = 1. Now, the point of HΩ is that there are vectors yi ∈ HΩ with ωi = hyi, Ω(· )yii for each i. Now apply V to get a map U : HΩ → H with the desired properties. (cid:3) IX Theorem Every np-map ω : A → C on a von Neumann subalgebra A of B(H ), where H is some Hilbert space, is of the form ω ≡ Pn hxn, (· )xni for some x1, x2, . . . ∈ H (with Pn kxnk2 < ∞). Let : A → B(H ) denote the inclusion, and let U : HΩ → H be as in VII. Since ω ∈ Ω, there is y ∈ HΩ with ω = hy, Ω(· )yi. We're going to 'transfer' y from HΩ to H using the following device. Since 1 = ⌈⌈U∗U⌉⌉Ω(A )(cid:3), we can (by 83 V) find partial isometries (vi)i in Ω(A )(cid:3) with 1 =Pi v∗i vi and viv∗i 6 U∗U for all i. Then for every a ∈ A , Proof (Based on Theorem 7.1.8 of [47].) X ω(a) = h y, Ω(a)y i = Pi h y, v∗i vi Ω(a)y i = Pi h y, v∗i U∗U vi Ω(a)y i = Pi h U viy, U Ω(a)viy i = Pi h U viy, (a) U viy i since 1 =Pi v∗i vi since viv∗i 6 U∗U since vi ∈ (A )(cid:3) since U Ω(a) = (a)U . XI Corollary Let A be a von Neumann subalgebra of a von Neumann algebra B. In particular, ω(1) = Pi kU viyk2, so at most countably many U viy's are non- zero; and denoting those by x1, x2, . . . , we get ω =Pn hxn, (· )xni. 1. For every np-map ω : A → C there is an np-map ξ : B → C with ξA = ω. (cid:3) 2. Ultraweak permanence: the restriction of the ultraweak topology on B to A coincides with the ultraweak topology on A . 3. Ultrastrong permanence: the restriction of the ultrastrong topology on B to A coincides with the ultrastrong topology on A . Show using 48 VI that any np-functional ω : A → C can be extended along , Exercise Let : A → B be an injective nmiu-map. that is, there is an np-functional ω′ : B → C with ◦ ω′ = ω. We end the chapter with another corollary to 89 IX: that the np-functionals on a von Neumann algebra are generated (in a certain sense) by any centre separating collection of functionals. This fact plays an important role in the next chapter for our definition of the tensor product of von Neumann algebras (on which the product functionals are to be centre separating, 108 II). XII 90 Proposition Given a centre separating collection Ω of np-functionals on a von Neumann algebra A , and an ultrastrongly dense subset S of A II III IV 1. Ω′ := { ω(s∗(· )s) : ω ∈ Ω, s ∈ S } is order separating, and 2. Ω′′ := {Pn ωn : ω1, . . . , ωN ∈ Ω′ } is operator norm dense in (A∗)+. Proof We tackle 1 first. We already know from 30 X that the collection Ξ := { ω(a∗(· )a) : ω ∈ Ω, a ∈ A }, which contains Ω′, is order separating; so to prove that Ω′ is itself order separating it suffices by 21 X to show that Ω′ is norm dense in Ξ. This is indeed the case since given a ∈ A and ω ∈ Ω, and a net (sα)α in S that converges ultrastrongly to a, the functionals sα ∗ ω ≡ ω(s∗α(· )sα) converge in norm to a ∗ ω as α → ∞ by 72 III. (Concerning 2) Let f : A → C be an np-map; we must show that f is in the norm closure Ω′′ of Ω′′. Note that since Ω is centre separating, the map Ω : A → B(HΩ) from 30 X is injective, and in fact restricts to an nmiu- isomorphism from A onto Ω(A ) (cf. 48 VIII). So by 89 IX f is of the form f ≡ Pn hxn, Ω(· )xni for some x1, x2, . . . ∈ HΩ with Pn kxnk2 < ∞, so that the partial sums PN n=1 hxn, Ω(· )xni converge with respect to the opera- tor norm to f (by 38 VI). Thus to show that f is in Ω′′ it suffices to show that each hxn, Ω(· )xni is in Ω′′ (since Ω′′ is clearly closed under finite sums and norm limits). In effect we may assume without loss of generality that f ≡ hx, Ω(· )xi for some x ∈ HΩ. We reduce the problem some more. By definition of Hω and Ω, we have f = hx, Ω(· )xi = Pω∈Ω hxω, ω(· )xωi; Hω ≡ Lω∈Ω and so we may, by the same token, assume without loss of generality that f = hx, ω(· )xi for some ω ∈ Ω and x ∈ Hω. Since such x (by definition of Hω, 30 VI) is the norm limit of a sequence ηω(a1), ηω(a2), ··· , where a1, a2, . . . ∈ A , the np-maps an ∗ ω ≡ hηω(an), ω(· )ηω(an)i converge to hx, ω(· )xi = f in the ..89 -- 90.. 143 operator norm as n → ∞ by 38 VI; and so we may assume without loss of gen- erality that f = a∗ ω for some a ∈ A and ω ∈ Ω. Since S is ultrastrongly dense in A we can find a net (sα)α in S that converges ultrastrongly to a. As the np-functionals sα ∗ ω in Ω′ ⊆ Ω′′ will then operator-norm converge to f = a ∗ ω as α → ∞ by 72 III, we conclude that f ∈ Ω′′. (cid:3) 91 With this chapter ends perhaps the most hairy part of this thesis: we've de- veloped the theory of von Neumann algebras starting from Kadison's charac- terisation (see 42) to the point that we have a sufficiently firm hold on the normal functionals (see e.g. 86 IX, 89 IX), the ultraweak and ultrastrong topolo- gies (e.g. 74 IV, 89 XI, 90 II), the projections (56 I, 59 I, 65 IV), and the division structure (81 V, 82 I) on a von Neumann algebra. In the next chapter we reap the benefits of our labour when we study an assortment of structures in the category W∗cpsu of von Neumann algebras and ncpsu-maps. Chapter 4 Assorted Structure in W∗cpsu In the previous two chapters we have travelled through charted territory when developing the theory of C∗-algebras and von Neumann algebras adding some new landmarks and shortcuts of our own along the way. In this chapter we properly break new ground by revealing two entirely new features of the cate- gory W∗cpsu of von Neumann algebras and the normal completely positive sub- unital linear maps between them, namely, 92 1. that the binary operation ∗ on the effects of a von Neumann algebra A given by p ∗ q = √pq√p (representing measurement of p) can be axioma- tised, and 2. that the category W∗cpsu has all the bits and pieces needed to be a model of Selinger and Valiron's quantum lambda calculus. We'll deal with the first matter directly after this introduction in Section 4.1. The second matter is treated in Section 4.3, but only after we have given the tensor product of von Neumann algebras a complete overhaul in Section 4.2. Fi- nally, as an offshoot of our model of the quantum lambda calculus we'll study all von Neumann algebras that admit a 'duplicator' in Section 4.4 -- surprisingly, they're all of the form ℓ∞(X). 4.1 Measurement The maps on a von Neumann algebra A of the form a 7→ √pa√p : A → A , where p is an effect of A , represent measurement of p, and are called assert maps in [30]. The importance of these maps to any logical description of quantum 93 ..90 -- 93.. 145 computation is not easily overstated. On the effects of A these maps are also studied in the guise of the binary operation p ∗ q = √pq√p called the sequential product (see e.g. [25]). We'll axiomatise this operation in this section in terms of the properties of the underlying assert maps. Our first observation to this end is that any assert map factors as A π : a7→⌈p⌉a⌈p⌉ / ⌈p⌉A ⌈p⌉ c : a7→√pa√p / A , where both π and c obey a universal property: c is a filter of p, see 96 I, and π is a corner of ⌈p⌉, see 95 I. Such maps that are the composition of a filter and a corner will be called pure, see 100 I, Since not only assert maps turn out to be pure, but also maps of the form b∗(· )b : A → A for an arbitrary element b of A , we need another property of assert maps, namely that √p e1 √p 6 e⊥2 ⇐⇒ √p e2 √p 6 e⊥1 for all projections e1 and e2 of A -- which we'll describe by saying that √p(· )√p : A → A is ⋄-self-adjoint. Judging only by the name it may not surprise you that the map b(· )b : A → A where b ∈ A is self-adjoint (but not necessarily positive) turns out to be ⋄-self-adjoint too, so that as a final touch we introduce the notion of ⋄-positive maps f : A → A that are simply maps of the form f ≡ gg for some ⋄-self-adjoint g. The main technical result, then, of this section is that any ⋄-positive map f : A → A is of the form f = √p(· )√p where p = f (1); and, accordingly, our axioms (in 106 I) that uniquely determine the sequential product ∗ on the effects of a von Neumann algebra A are: for every effect p of A , 1. p ∗ 1 = p, 2. p ∗ q = f (q) for all q ∈ [0, 1]A for some pure map f : A → A , 3. p = q ∗ q for some q from [0, 1]A , 4. p ∗ (p ∗ q) = (p ∗ p) ∗ q for all q ∈ [0, 1]A , 5. p ∗ e1 6 e⊥2 ⇐⇒ p ∗ e2 6 e⊥1 for all projections e1, e2 of A . While I would certainly not like to undersell the results mentioned above, I suspect that the notion of purity exposed along the way might turn out to be of far greater significance for the following reason. Our notion of purity can be described in wildly different terms: a map f : A → B is pure when given / B the map is surjective (see 171 VII its Paschke dilation A and [82]). Because of my faith in our notion of purity I've allowed myself to / P c / / / / address some theoretical questions concerning it here that are not required for the main results of this thesis, but suppose a general interest in purity: I'll show that every pure map f : A → B is extreme among the ncp-maps g : A → B with f (1) = g(1), and, in fact, enjoys the possibly stronger property of being rigid (see 102 II and 102 IX). 4.1.1 Corner and Filter Definition Given a projection e of a von Neumann algebra A , the corner of e is the subset eA e of A (consisting of the elements of A of the form eae with a ∈ A ). In this context, the obvious map eA e → A is called the inclusion and the map a 7→ eae, A → eA e is called the projection. Exercise Let e be a projection from a von Neumann algebra A . 94 II 1. Show that a ∈ A is an element of eA e iff eae = a iff (a⌉ ∪ ⌈a) 6 e. 2. Show that the corner eA e is closed under addition, (scalar) multiplication, and involution. 3. Show that e is a unit for eA e, that is, ea = ae = a for all a ∈ eA e. 4. Show that eA e is norm and ultraweakly closed. (Hint: use the fact that e(· )e : A → A is normal and bounded.) 5. Show that eA e -- endowed with the addition, (scalar) multiplication, involution and norm from A , and with e as its unit -- is a C∗-algebra. 6. Show that the supremum of a bounded directed set D of self-adjoint ele- ments of eA e computed in A is itself in eA e, and, in fact, the supremum of D in eA e. 7. Show that the inclusion eA e → A is an ncpsu-map. 8. Deduce from this that the restriction of an np-map ω : A → C to a map eA e → C is an np-map. Conclude that eA e is a von Neumann algebra. 9. Show that the projection a 7→ eae, A → eA e is an ncpu-map. 10. Show that every np-map ω : eA e → C is the restriction of the np-map ω(e(· )e) : A → C. Deduce from this that the ultraweak topology of eA e coincides (on eA e) with the ultraweak topology on A . Show that the ultrastrong topologies on eA e and A coincide in a similar fashion. ..93 -- 94.. 147 III Exercise Let a be an element of a von Neumann algebra A , and let p and q be projections of A with a∗pa 6 q. 1. Show that a∗ba ∈ qA q for every b ∈ pA p. 2. Show that a∗(· )a gives an ncp-map pA p → qA q. 95 II III 96 Ia Definition Let p be an effect of a von Neumann algebra A . A corner of p is an ncp-map π : A → C to a von Neumann algebra C with π(p⊥) = 0, which is initial among such maps in the sense that every ncp-map f : A → B with f (p⊥) = 0 factors as f = g ◦ π for some unique ncp-map g : C → B. While most corners that we'll deal with are unital, there are also corners which are not unital (because there are non-unital ncp-isomorphisms). When we write "corner" we shall always mean a "unital corner" unless explicitly stated otherwise. Proposition Given an effect p of a von Neumann algebra A , and a partial isometry u of A with ⌊p⌋ = uu∗, the map π : A → u∗uA u∗u given by π(a) = u∗au is a corner of p. Proof By 94 III, π is an ncp-map. To see that π(p⊥) ≡ u∗p⊥u = 0, note that since u∗u = u∗ uu∗ u, we have 0 = u∗(uu∗)⊥u = u∗⌊p⌋⊥u = u∗ ⌈p⊥⌉ u, and so 0 = ⌈u∗ ⌈p⊥⌉ u⌉ =(cid:6)u∗p⊥u(cid:7) by 60 VII, giving u∗p⊥u = 0 by 59 III. Let B be a von Neumann algebra, and let f : A → B be an ncp-map with f (p⊥) = 0. To show that π is a corner, we must show that there is a unique ncp-map g : u∗uA u∗u → B with f = g◦π. Uniqueness follows from surjectivity of π. Concerning existence, define g := f ◦ ζ, where ζ : u∗uA u∗u → A is the ncp-map given by ζ(a) = uau∗ for a ∈ A (see 94 III), so that it is immediately clear that g is an ncp-map. It remains to be shown f = g ◦ π, that is, f (a) = f (uu∗ a uu∗) for all a ∈ A . This follows from 63 IV because f ((uu∗)⊥) = 0, since ⌈f ( (uu∗)⊥ )⌉ = ⌈f (⌊p⌋⊥)⌉ = ⌈f (⌈p⊥⌉)⌉ = ⌈f (p⊥)⌉ = ⌈0⌉ = 0. (cid:3) Definition A filter is an ncp-map c : C → A between von Neumann algebras such that every ncp-map f : B → A with f (1) 6 c(1) factors as f = c ◦ g for some unique ncp-map g : B → C . We'll say that c is a filter for c(1). Remark In the abstract setting of effectus theory, it makes sense to call these filters "quotients", as we do in [8]; but since in the concrete setting of von Neumann algebras "quotient" has a pre-existing and unrelated meaning, we chose to use the word "filter" instead (as in "polarising filter"), an idea borrowed from [86]. II To show that there is a filter for every positive element of a von Neumann algebra we need the following result concerning ultraweak limits of ncp-maps. III Lemma Given von Neumann algebras A and B the pointwise ultraweak limit f : A → B of a net of positive linear maps fα : A → B is positive, and, 1. f is completely positive provided that the fα are completely positive, and 2. f is normal provided that the fα are normal and the ultraweak convergence of the fα to f is uniform on [0, 1]A . Proof Since given a ∈ A the element f (a) is the ultraweak limit of the positive elements fα(a), and therefore positive (by 44 XI), we see that f is positive. Suppose that each fα is completely positive. To show that f is completely positive, we must prove, given a1, . . . , an ∈ A and b1, . . . , bn ∈ B, that the elementPi,j b∗i f (a∗i aj)bj of B is positive. And indeed it is, being the ultraweak limit of the positive elements Pi,j b∗i fα(a∗i aj)bj, because fα(a∗i aj) converges ultraweakly to f (a∗i aj), and b∗i (· )bj : B → B is ultraweakly continuous (45 IV) for any i and j. If the fα are normal, and converge uniformly on [0, 1]A ultraweakly to f , then f is ultraweakly continuous on [0, 1]A (because the uniform limit of con- tinuous functions is continuous), and thus normal (by 44 XV). (cid:3) Proposition Given an element d of a von Neumann algebra A , the map c : (d⌉A (d⌉ → A given by c(a) = d∗ad is a filter. Proof Note that c is an ncp-map by 94 III. Let B be a von Neumann algebra, and let f : B → A be an ncp-map with f (1) 6 c(1). To show that c is a filter, we must show that there is a unique ncp-map g : B → (d⌉A (d⌉ with f = c ◦ g. Uniqueness of g follows from the observation that c is injective by 60 VIII. To establish the existence of such g, note that f (b) is an element of d∗A d, when b is positive by 81 VI because 0 6 f (b) 6 kbkf (1) 6 kbkc(1) = kbkd∗d, and thus for arbitrary b ∈ B too (being a linear combination of positive elements). We can thus define g : B → (d⌉A (d⌉ by g(b) = d∗\f (b)/d for all b ∈ B. It is clear that g is linear and positive, and c ◦ g = f . To see that g is normal, note that d∗\ · /d : d∗(A )1d → A is ultrastrongly continuous by 81 IX, as is f by 45 II (also) as map from (B)1 to d∗(A )1d, so that g is ultrastrongly continuous on (B)1, and therefore normal by 44 XV. IV V VI Finally, g is completely positive by III, because it is by 81 VII the uniform ul- trastrong limit of the by 94 III completely positive maps (PN where t1, t2, . . . is an approximate pseudoinverse of d. n=1 tn)∗ f (· ) (PN n=1 tn), (cid:3) Before exploring their more technical aspects, we'll explain how corners and filters can be made to appear at opposite ends of a chain of adjunctions: 97 Eff Filter ⊣ 0 ⊣ ⊣ 1 ⊣ Corner (W∗cpsu)op ..94 -- 97.. 149   * * t t 8 8 f f The category Eff has as objects pairs (A , p), where A is a von Neumann algebra, and p ∈ [0, 1]A is an effect from A . A morphism (A , p) −→ (B, q) in Eff is an ncpsu-map f : B → A with p 6 f (q) + f (1)⊥ -- that is, ω(p) 6 ω(f (q)) + ω(f (1))⊥ for every normal state ω : A → C. The functor Eff −→ (W∗cpsu)op in the middle of the diagram above maps a morphism f : (A , p) → (B, q) to the underlying map f : B → A . The func- tors 0 and 1 on its sides map a von Neumann algebra A to (A , 0) and (A , 1), respectively, and send an ncpsu-map f : A → B to itself; this is possible since 0 6 f (0) + f (1)⊥ and 1 6 f (1) + f (1)⊥. That 1 is right adjoint to the functor Eff −→ (W∗cpsu)op follows from the obser- vation that an ncpsu-map f : B → A is always a morphism (A , p) → (B, 1), whatever p ∈ [0, 1]A may be, because p 6 f (1) + f (1)⊥. For a similar reason 0 is left adjoint to Eff −→ (W∗cpsu)op. On the other hand, a morphism (A , 1) → (B, q) where q ∈ [0, 1]B is not just any ncpsu-map f : B → A , but one with 1 6 f (q) + f (1)⊥, that is, f (q⊥) = 0. It's no surprise then that a corner π : B → C for q ∈ [0, 1]B considered as morphism (C , 1) → (B, q) is a universal arrow from 1 to (B, q). On the other side there's a twist: a morphism (A , p) → (B, 0) where p ∈ [0, 1]A is an ncpsu-map f : A → B with p 6 f (0) + f (1)⊥, that is, f (1) 6 p⊥. It follows that any filter c : C → A for p⊥, when considered as morphism (A , p) → (C , 0), is a universal arrow from (A , p) to 0. This chain of adjunctions not only exposes a hidden symmetry between filters and corners, but such chains appear in many other categories as well, see [8]. 98 Definition Let A be a von Neumann algebra. 1. Given a positive element p of A we denote by cp : ⌈p⌉A ⌈p⌉ → A the standard filter for p given by cp(a) = √pa√p for all a ∈ ⌈p⌉A ⌈p⌉. 2. Given an effect p of A we denote by πp : A → ⌊p⌋A ⌊p⌋ the standard corner of p given by πp(a) = ⌊p⌋a⌊p⌋. II Exercise Let c : C → A be a filter, where C and A are von Neumann algebras. 1. Show that, writing p := c(1), there is a unique ncp-map α : C → ⌈p⌉A ⌈p⌉ with c = cp ◦ α; and that this α is a unital ncp-isomorphism. 2. Show that c is injective (by proving first that cp is injective using 60 VIII). Conclude that c is faithful (so ⌈f⌉ = 1), and that c is mono in W∗CP. 3. Show that c is bipositive (by proving first that cp is bipositive using 81 VI). Exercise Show that the composition d ◦ c of filters c : C → D and d : D → A between von Neumann algebras is again a filter. Exercise Let p be an effect of a von Neumann algebra A , and let π : A → C be a corner of p. 1. Show that there is a unique ncp-map β : ⌊p⌋A ⌊p⌋ → C with π = β ◦ πp; and that this β is unital and an ncp-isomorphism. 2. Show that π is surjective, and that π is epi in W∗cp. Exercise Show that an ncpu-map π : A → B between von Neumann algebras is a corner for an effect p of A iff π is a corner for ⌊p⌋; in which case ⌈π⌉ = ⌊p⌋. Thus a corner π is a corner for ⌈π⌉. Exercise Show that the composition τ ◦ π of corners π : A → B and τ : B → C between von Neumann algebras is again a corner. (Hint: prove and use the inequality ⌈τ⌉ 6 ⌈π(⌈τ ◦ π⌉⊥)⌉⊥.) Theorem Given an ncp-map f : A → B between von Neumann algebras, a projection e of A with ⌈f⌉ 6 e, and a positive element p of B with f (1) 6 p, there is a unique ncp-map g : eA e → ⌈p⌉B⌈p⌉ such that III IV V VI VII A πe f B cp eA e g / ⌈p⌉B⌈p⌉ commutes, and it is given by g(a) = √p\f (a)/√p for all a ∈ eA e. Proof Uniqueness of g follows from the facts that πe is epi and cp is mono in W∗cp, see IV and II. VIII Concerning existence, since πe is a corner of e, 95 I, and ⌈f⌉ 6 e, or in other words, f (e⊥) = 0, there is a unique ncp-map h : eA e → B with h ◦ πe = f . Note that h(a) = f (a) for all a from eA e. As h(1) = h(πe(1)) = f (1) 6 p = cp(1), and cp is a filter, 96 I, there is a unique ncp-map g : eA e → pBp with cp ◦ g = h, which is (by the proof of 96 V) given by g(a) = √p\h(a)/√p ≡ √p\f (a)/√p for all a from eA e. Then cp ◦ g ◦ πe = h ◦ πe = f . (cid:3) Corollary Given an ncp-map f : A → B between von Neumann algebras, there IX ..97 -- 98.. 151 / /   / O O is a unique ncp-map [f ]: ⌈f⌉A ⌈f⌉ → ⌈f (1)⌉B⌈f (1)⌉ such that A π⌈f ⌉ f B cf (1) ⌈f⌉A ⌈f⌉ [f ] / ⌈f (1)⌉B⌈f (1)⌉ X Moreover, [f ] is unital and faithful. commutes; and it is given by [f ](a) =pf (1)\f (a)/pf (1) for all a from ⌈f⌉A ⌈f⌉. Example For any faithful unital ncp-map f : A → B we have [f ] = f . Such a map need not be an isomorphism; as one may take f : (λ, µ) 7→ 1 2 (λ+µ), C2 → C. XI Example In the concrete case that f ≡ a∗(· )a : sA s → tA t, where a is an element of a von Neumann algebra, and s and t are projections of A with (a⌉ 6 s and ⌈a) 6 t, the map [f ] is closely related to the polar decomposition a ≡ [a]√a∗a = √aa∗[a] of a, where [a] = a/√a∗a (see 82 I). Indeed, since ⌈f⌉ = (a⌉, f (1) = a∗a, and [f ] ≡ √a∗a\a∗(· )a/√a∗a ≡ [a](· )[a]∗, the picture becomes: sA s π(a⌉ (a⌉A (a⌉ f = a∗ ( · ) a [f ] = [a] ( · ) [a]∗ tA t ca∗a / ⌈a)A ⌈a) Note that in this example [f ] is an ncpu-isomorphism, because [a] is a partial isometry with initial projection ⌈a) and final projection (a⌉. Thus one can think of the diagram above as an isomorphism theorem of sorts, which applies only to certain ncp-maps that'll be called pure in a moment (see 100 III). 4.1.2 Isomorphism 99 II In case you were wondering, the ncpu-isomorphism we encountered in 98 XI is simply an nmiu-isomorphism (see IX), which follows from the following charac- terisation of multiplicativity. Proposition For an ncpu-map f : A → B between von Neumann algebras the following are equivalent. 1. f is multiplicative. 2. f (a)f (b) = 0 for all a, b ∈ A with ab = 0. / /   / O O / /   / O O 3. ⌈f (p)⌉⌈f (q)⌉ = 0 for all projections p and q of A with pq = 0. 4. f maps projections to projections. 5. ⌈f (a)⌉ = f (⌈a⌉) for all a ∈ A+. Proof (Based in part on the work of Gardner in [22]). (1=⇒4 and 5=⇒4) are rather obvious. (4=⇒5) ⌈f (a)⌉ 60 V===⌈f (⌈a⌉)⌉ = f (⌈a⌉) since f (⌈a⌉) is a projection. (4=⇒3) Let p and q be projections of A with pq = 0. Then p 6 q⊥, and so f (p) 6 f (q⊥) = f (q)⊥, which implies that ⌈f (p)⌉⌈f (q)⌉ = f (p)f (q) = 0 since f (p) and f (q) are projections. (3=⇒2) Let a, b ∈ A with ab = 0 be given. We must show that f (a)f (b) = 0, and for this it suffices to show that ⌈f (a)) (f (b)⌉ = 0, because f (a)f (b) = f (a)⌈f (a)) (f (b)⌉ f (b). Since ab = 0, we have ⌈a) (b⌉ = 0 by 60 VIII, and so ⌈f (⌈a))⌉⌈f ((a⌉)⌉ = 0. Now, since ⌈f (⌈a))⌉ 6 ⌈f (a)) and ⌈f ((a⌉⌉ 6 (f (a)⌉ by 61 II, we get ⌈f (a)) (f (b)⌉ = ⌈f (a))⌈f (⌈a))⌉⌈f ((a⌉)⌉ (f (a)⌉ = 0. (2=⇒1) We must show that f (a)f (b) = f (ab) for all a, b ∈ A . Since the linear span of projections is norm-dense in A , it suffices to show that f (a)f (e) = f (ae) for any a ∈ A and a projection e of A . Given such a and e, we on the one hand have ae⊥ e = 0, so that f (ae⊥)f (e) = 0, that is, f (a)f (e) = f (ae)f (e); and on the other hand we have ae e⊥ = 0, so that f (ae)f (e⊥) = 0, that is, f (ae) = f (ae)f (e); so that we reach f (ae) = f (a)f (e) as sum total, and the result that f is multiplicative. (cid:3) Theorem An ncpsu-isomorphism f : A → B between von Neumann algebras (so both f and f−1 are ncpsu-maps) is an nmiu-isomorphism. Proof Since f−1(1) 6 1 and so 1 = f (f−1(1)) 6 f (1) 6 1, we see that f (1) = 1, so both f and f−1 are unital. It remains to be shown that f and f−1 are multiplicative. Since by 55 X an effect a of A is a projection iff 0 is the infimum of a and a⊥, and f (as ncpu-isomorphism) preserves (· )⊥ and order, we see that f maps projections to projections, and is thus multiplicative, by II. It follows automatically that f−1 is multiplicative too. (cid:3) III IV V VI VII VIII IX X Exercise Show that any filter of a projection is multiplicative. (Hint: the filter is a standard filter up to an ncpu-isomorphism, 98 II, which is an nmiu-isomorphism by IX.) Exercise Show that for an ncp-map f : A → B between von Neumann algebras the following are equivalent. XI XII 1. f is multiplicative. 2. f sends projections to projections. 3. ⌈f (a)⌉ = f (⌈a⌉) for all a ∈ A+. ..98 -- 99.. 153 (Hint: factor f = ζ ◦ h where ζ is a filter for f (1) and h is an ncp-map.) 4.1.3 Purity 100 Definition Filters, corners, and their compositions we'll call pure. II Exercise Show that the following maps are pure. 1. An ncp-isomorphism between von Neumann algebras. 2. The identity map id : A → A on a von Neumann algebra A . 3. The map a∗ (· ) a : A → A for any element a of a von Neumann alge- bra A . III Proposition For an ncp-map f : A → B between von Neumann algebras the following are equivalent. 1. f is pure, i.e., f is the composition of (perhaps many) filters and corners. 2. f = c ◦ π for a filter c : C → B and a corner π : A → C . 3. [f ] from 98 IX is an ncpu-isomorphism. IV Proof 3=⇒2 and 2=⇒1 are rather obvious. V (1=⇒2) Calling f properly pure when f ≡ c ◦ π for some filter c and corner π, we must show that every pure map is properly pure. For this it suffices to show that the composition of properly pure maps is again properly pure; which, since filters are closed under composition (by 98 III), and corners are closed under composition (by 98 VI), boils down to proving that the composition π ◦ c of a corner π and a filter c is properly pure. Since π ≡ α ◦ π⌈π⌉ and c ≡ cc(1) ◦ β for ncpu-isomorphisms α and β (see 98 II and 98 IV) it suffices to show that f := πscp is properly pure for a positive element p and a projection s of a von Neumann algebra A . Since such f is of the form f = s√p(· )√ps : ⌈p⌉A ⌈p⌉ → sA s, we know by 98 XI that [f ] is an ncpu-isomorphism, and thus that f ≡ cf (1)◦[f ]◦π⌈f⌉ is properly pure. (2=⇒3) Recall that [f ] is by definition the unique ncp-map with f = cf (1)[f ]π⌈f⌉, see 98 IX. Note that since f = c◦π, we have ⌈f⌉ = ⌈π⌉ (because ⌈c⌉ = 1 by 98 II), and f (1) = c(1) (because π(1) = 1). Since there are ncpu-isomorphisms α and β with π = απ⌈π⌉ and c = cc(1)β, we see that f = cc(1)βαπ⌈π⌉, and so [f ] = βα by definition of [f ], so [f ] is an ncpu-isomorphism. (cid:3) VI VII Exercise Use III to show that 1. a faithful pure map is a filter, 2. a unital pure map is a corner, and 3. a unital and faithful pure map is an ncpu-isomorphism. 4.1.4 Contraposition Definition Given an ncp-map f : A → B between von Neumann algebras we define f⋄ : Proj(A ) → Proj(B) by f⋄(e) = ⌈f (e)⌉ for all e ∈ Proj(A ). Remark The significance of the symbol "⋄" in f⋄ is in accommodating the notation f (e) := f⋄(e⊥)⊥ used in the next thesis, in 206 II. Proposition Given an ncp-map f : A → B between von Neumann algebras and a projection e from B there is a least projection f⋄(e) from A with(cid:6)f ( f⋄(e)⊥ )(cid:7) 6 e⊥, namely f⋄(e) = ⌈ ef (· )e ⌉ (being the carrier of the ncp-map ef (· )e from 63 I); giving a map f⋄ : Proj(B) → Proj(A ). Proof Since by definition ⌈ ef (· )e ⌉ is the greatest projection s of A with ef (s⊥)e = 0 (see 63 I); and ef (s⊥)e = 0 iff (cid:6)f (s⊥)(cid:7) 6 ⌈e(· )e⌉⊥ ≡ e⊥; the projection ⌈ ef (· )e ⌉ satisfies the description of f⋄(e). Exercise Let f : A → B be an ncp-map between von Neumann algebras. (cid:3) 1. Show that f⋄(s) 6 t⊥ iff f⋄(t) 6 s⊥, for all s ∈ Proj(A ) and t ∈ Proj(B). 2. Show that f⋄(S E ) =Se∈E f⋄(e) for every set of projections E from A . Exercise Show that for ncp-maps f, g : A → B between von Neumann algebras f⋄ = g⋄ iff f⋄ = g⋄. In that case we say that f and g are equivalent. Show that for ncp-maps f : A → B and g : B → A we have f⋄ = g⋄ iff f⋄ = g⋄ iff ⌈f (s)⌉ 6 t⊥ ⇐⇒ ⌈g(t)⌉ 6 s⊥ for all projections s from A and t from B. In that case we say that f and g are contraposed. Examples 101 Ia II III IV V VI VII 1. Given an element a of a von Neumann algebra A , the maps a∗(· )a and a(· )a∗ on A are contraposed. If p and q are projections of A with a∗pa 6 q (as in 94 III), then the maps a∗(· )a : pA p → qA q and a(· )a∗ : qA q → pA p are contraposed. In particular, the standard corner πs : A → sA s and the standard filter cs : sA s → A for a projection s from A are contraposed. 2. An ncp-isomorphism f : A → B between von Neumann algebras is con- traposed to its inverse f−1 : B → A . ..99 -- 101.. 155 3. There may be many maps equivalent to a given ncp-map f : A → B between von Neumann algebras: show that (zf )⋄ = f⋄ for every positive central element z of B with ⌈z⌉ = 1. VIII Exercise Let A f / B g / C be ncp-maps between von Neumann alge- bras A , B and C . 1. Show that (g ◦ f )⋄ = g⋄ ◦ f⋄ (using 60 V), and (g ◦ f )⋄ = f⋄ ◦ g⋄. 2. Assuming that f is equivalent to an ncp-map f′ : A → B and g is equiv- alent to an ncp-map g′ : B → C , show that g ◦ f is equivalent to g′ ◦ f′. 3. Assuming that f is contraposed to an ncp-map f′ : B → A and g is contraposed to an ncp-map g′ : C → B, show that g ◦ f is contraposed to f′ ◦ g′. IX Proposition Given ncp-maps f, g : A → B between von Neumann algebras (f + g)⋄(s) = f⋄(s) ∪ g⋄(s) and (f + g)⋄(t) = f⋄(t) ∪ g⋄(t) X XI for all s ∈ Proj(A ) and t ∈ Proj(B). Proof Note that (f + g)⋄(s) = ⌈(f + g)(s)⌉ = ⌈f (s) + g(s)⌉ = ⌈f (s)⌉∪⌈g(s)⌉ = f⋄(s) ∪ g⋄(s) by 59 III. Since (f + g)⋄(t) 6 s⊥ iff f⋄(s) ∪ g⋄(s) ≡ (f + g)⋄(s) 6 t⊥ iff both f⋄(s) 6 t⊥ and g⋄(s) 6 t⊥ iff both f⋄(t) 6 s⊥ and g⋄(t) 6 s⊥ iff f⋄(t) ∪ g⋄(t) 6 s⊥, we see that (f + g)⋄(t) = f⋄(t) ∪ g⋄(t). (cid:3) Lemma Given contraposed maps f : A → B and g : B → A between von Neumann algebras, we have ⌈f⌉ = ⌈gf⌉. XII Proof ⌈gf⌉ = (gf )⋄(1) = f⋄(g⋄(1)) = g⋄(⌈g⌉) = g⋄(1) = f⋄(1) = ⌈f⌉. (cid:3) 4.1.5 Rigidity 102 We now turn to a remarkable property shared by pure and nmiu-maps. II Definition We say that an ncp-map f : A → B between von Neumann algebras is rigid when the only ncp-map g : A → B with g(1) = f (1) and ⌈f (p)⌉ = ⌈g(p)⌉ for all projections p from A is f itself. among the ncp-maps g : A → B with g(1) = f (1). III Proposition A rigid map f : A → B between von Neumann algebras is extreme IV Proof Given f ≡ λg1 + λ⊥g2 where λ ∈ (0, 1) and g1, g2 : A → B are ncp- maps with gi(1) = f (1), we must show that f = g1 = g2. Note that for every projection s of A we have f⋄(s) = (λg1 + λ⊥g2)⋄(s) = g⋄1(s) ∪ g⋄2(s) by 101 IX and 101 VII; and in particular g⋄1(s) 6 f⋄(s). Then for h := λg1 + λ⊥f we have h(1) = f (1) and h⋄(s) = g⋄1(s) ∪ f⋄(s) = f⋄(s), so that λg1 + λ⊥f ≡ h = f = λg1 + λ⊥g2 by rigidity of f ; and thus f = g2. Similarly, f = g1. (cid:3) / / Proposition A nmiu-map : A → B between von Neumann algebras is rigid. Proof Let g : A → B be an ncpu-map with ⌈(p)⌉ = ⌈g(p)⌉ for every pro- jection p of A . To show that is rigid, we must show that g = , and for this, it suffices to prove that g(p) = (p) for every projection p of A . To this end, we'll show that g is multiplicative, because then g maps pro- jections to projections, so that g(p) = ⌈g(p)⌉ = ⌈(p)⌉ = (p) for every pro- jection p of A . We'll show that g is multiplicative using 99 II by proving that ⌈g(p)⌉⌈g(q)⌉ = 0 for projections p and q of A with pq = 0. Indeed, ⌈g(p)⌉⌈g(q)⌉ = ⌈(p)⌉⌈(q)⌉ = (p)(q) = (pq) = (0) = 0. (cid:3) Lemma Given an element b of a von Neumann algebra A the ncp-map a 7→ b∗ab, (b⌉A (b⌉ → A is rigid. Proof Let g : (b⌉A (b⌉ → A be an ncp-map with g(1) = b∗b and ⌈b∗pb⌉ = ⌈g(p)⌉ for every projection p of (b⌉A (b⌉. To prove that c := b∗(· )b : (b⌉A (b⌉ → A is rigid, we must show that g = c. Since c is a filter (by 96 V) and g(1) = b∗b there is a unique ncp-map h : (b⌉A (b⌉ → (b⌉A (b⌉ with g = c◦ h. Our task then is to show that h = id, and for this it suffices to show that, for all a ∈ (b⌉A (b⌉, V VI VII VIII en h( en a en ) en = en a en (4.1) for some sequence of projections e1, e2, . . . of (b⌉A (b⌉ that converges ultra- strongly to (b⌉, because by 45 VI the left-hand side of the equation above con- verges ultrastrongly to g(a), while the right-hand side converges ultrastrongly to a. We'll take eN :=PN n=1 ⌈tn), where t1, t2, . . . is an approximate pseudoin- verse for b, because (b⌉ =Pn ⌈tn). Since the identity on enA en is rigid by V, it suffices (for (4.1)) to show that enh(en)en = en and ⌈enh(p)en⌉ = p for every projection p from enA en. Writ- ing sN := PN n=1 tn, we have bsn = en, and so ⌈enh(p)en⌉ = ⌈s∗nb∗h(p)bsn⌉ = ⌈s∗ng(p)sn⌉ = ⌈s∗n ⌈g(p)⌉ sn⌉ = ⌈s∗n ⌈b∗pb⌉ sn⌉ = ⌈s∗nb∗pbsn⌉ = ⌈enpen⌉ for every projection p from (b⌉A (b⌉. In particular, ⌈enh(p)en⌉ = p when p is from enA en; and we see (cid:6)enh(e⊥n )en(cid:7) = (cid:6)ene⊥n en(cid:7) = 0 when we take p = e⊥n , so that enh(e⊥n )en = 0, which yields enh(en)en = en. (cid:3) Theorem Every pure map between von Neumann algebras is rigid. Proof Let f : A → B be a pure map between von Neumann algebras, and let g : A → B be an ncp-map with f (1) = g(1) and f⋄ = g⋄. To show that f is rigid, we must prove that f = g. We know by 98 IX that f can be written as f ≡ cf (1)◦[f ]◦π⌈f⌉, and that cf (1) is rigid, by VII, which we'll use shortly. To this end, note that since f⋄ = g⋄, we have f⋄ = g⋄, and so ⌈f⌉ = f⋄(1) = g⋄(1) = ⌈g⌉. As π⌈f⌉ is a corner of ⌈f⌉ = ⌈g⌉, there is a unique ncp-map h : ⌈f⌉A ⌈f⌉ → B , and π⋄ with h ◦ π⌈f⌉ = g. Since then h⋄ ◦ π⋄ ⌈f⌉ ⌈f⌉ is clearly surjective, we get h⋄ = c⋄f (1) ◦ [f ]⋄, and thus (h◦ [f ]−1)⋄ = c⋄f (1), using here that [f ] is invertible, because f is pure. Now, using that cf (1) is rigid, and = g⋄ = f⋄ = c⋄f (1) ◦ [f ]⋄ ◦ π⋄ ⌈f⌉ IX X ..101 -- 102.. 157 h([f ]−1(1)) = h(1) = h(π⌈f⌉(1)) = g(1) = f (1) = cf (1)(1), we get h ◦ [f ]−1 = cf (1), which yields g = h ◦ π⌈f⌉ = h ◦ [f ]−1 ◦ [f ] ◦ π⌈f⌉ = cf (1) ◦ [f ] ◦ π⌈f⌉ = f , and thus f is rigid. (cid:3) 4.1.6 ⋄-Positivity 103 Definition We'll call an ncp-map f : A → A between von Neumann algebras 1. ⋄-self-adjoint if f is pure and contraposed to itself (f⋄ = f⋄), and 2. ⋄-positive if f ≡ gg for some ⋄-self-adjoint map g : A → A . We added "⋄-" to "positive" not only to distinguish it from the pre-existing notion of positivity for maps between C∗-algebras, but also to contrast it with the notion of "†-positivity" that appears in the following thesis (see 214 I). Examples Let A be a von Neumann algebra. II 1. Given a ∈ AR the map a(· )a : A → A is ⋄-self-adjoint. 2. Given a ∈ A+ the map a(· )a : A → A is ⋄-positive. III Exercise Let f : A → A be an ncp-map, where A is a von Neumann algebra. 1. Show that ⌈f⌉ = ⌈f (1)⌉ when f is ⋄-self-adjoint. 2. Assuming f is ⋄-self-adjoint, show that f f is ⋄-self-adjoint, and show that ⌈f f⌉ = ⌈f⌉ (cf. 101 XI). 3. Show that f is ⋄-self-adjoint when f is ⋄-positive. 104 We now turn to the question roughly speaking to what extent a filter c is deter- mined by its action c⋄ : e 7→ ⌈c(e)⌉ on projections; we will see (essentially in VII) that two filters c1 and c2 are equivalent, c⋄1 = c⋄2, if and only if c1(1) and c2(1) are equal up to some central elements, that is, centrally similar. II Definition We say that positive elements p and q of a von Neumann algebra A are centrally similar if cp = dq for some positive central elements c and d of A with ⌈p⌉ 6 ⌈c⌉ and ⌈q⌉ 6 ⌈d⌉. Exercise Let p and q be positive elements of a von Neumann algebra A . III 1. Show that when p and q are centrally similar, every element a of A that commutes with p commutes with q too; and in particular, pq = qp. 2. Show that when p and q are centrally similar, ⌈p⌉ = ⌈q⌉. 2a. Assuming that p 6 Bq for some B ∈ [0,∞), show that p and q are centrally similar iff p/q is central. Show that p is centrally similar to 1 iff p is central. Show that p is centrally similar to p2 iff p is central. 3. Show that when p and q commute, and both p∧q p and p∧q q are central, p and q are centrally similar. 4. Show that when p and q are pseudoinvertible, then: p and q are centrally similar iff pq∼1 is central iff qp∼1 is central iff both (p∧q)p∼1 and (p∧q)q∼1 are central. 5. Assuming that p and q commute and e1 6 e2 6 ··· are projections com- muting with p and q with Sn en = ⌈p⌉ such that the enp and enq are pseudoinvertible, and centrally similar, show that p and q are centrally similar on the grounds that both p∧q p and p∧q q are central. (Hint: en p = (enp)∧(enq) p∧q enp are central, and converge ultraweakly to p∧q p .) Lemma Suppose that ⌈q ϑ(e) q⌉ 6 e and (cid:6)q ϑ(e⊥) q(cid:7) 6 e⊥, where e is a projec- tion of a von Neumann algebra A , q is a positive element of A , and ϑ : A → A is an miu-map. Then eq = qe and ϑ(e) = e. Proof We have ϑ(e)qe = ϑ(e)q, because e > ⌈q ϑ(e) q⌉ ≡ ⌈ϑ(e)q) (see 59 VI). Similarly, ϑ(e⊥)qe⊥ = ϑ(e⊥)q, because e⊥ > (cid:6)q ϑ(e⊥) q(cid:7) ≡ (cid:6)ϑ(e⊥)q(cid:1), and so ϑ(e⊥)qe = 0, which implies ϑ(e)qe = qe. Thus qe = ϑ(e)qe = ϑ(e)q, and so q2e = qϑ(e)q is self-adjoint, which gives us that q2e = (q2e)∗ = eq2. Since q2 commutes with e, q = pq2 commutes with e too (see 23 VII). Finally, ϑ(e)q = qe = eq and ⌈q⌉ = 1 imply that ϑ(e) = e by 60 VIII. (cid:3) Corollary A positive element q of a von Neumann algebra A with ⌈q⌉ = 1 is central provided there is an miu-map ϑ : A → A with ⌈q ϑ(e) q⌉ 6 e for every projection e from A ; and in that case ϑ = id. Proposition Positive elements p and q of a von Neumann algebra A with ⌈p⌉ = ⌈q⌉ = 1 are centrally similar when there is an miu-isomorphism ϑ : A → A with ⌈pep⌉ = ⌈q ϑ(e) q⌉ for all projections e of A ; and in that case ϑ = id. Proof Let e be a projection from A with ep = pe. Since 1 = ⌈p⌉ = ⌈p2⌉ we have e = ⌈e ⌈p2⌉ e⌉ = ⌈ep2e⌉ = ⌈pep⌉ = ⌈q ϑ(e) q⌉. Since e⊥ commutes with p too, we get e⊥ = ⌈q ϑ(e⊥) q⌉ by the same token; and thus eq = qe and ϑ(e) = e by IV. Since p is the norm limit of linear combinations of such projections e, we get pq = qp and ϑ(p) = p. IV V VI VII VIII Since p and q commute, we can find a sequence of projections e1 6 e2 6 ··· that commute with p and q with Sn en = ⌈p⌉ and such that pen and qen are pseudoinvertible -- one may, for example, take eN :=PN n=1 ⌈tn⌉ where t1, t2, . . . ..102 -- 104.. 159 is an approximate pseudoinverse of p ∧ q (see 80 IV). Note that to prove that p and q are centrally similar, it suffices to show that pen and qen are centrally similar, by III. Further, to prove that ϑ(a) = a for some a ∈ A , it suffices to show that ϑ(enaen) = enaen, because enaen converges ultrastrongly to a by 45 VI. Note that ϑ(en) = en, because enp = pen, and so ϑ maps enA en into enA en. Thus, by considering enA en instead of A , and the restriction of ϑ to enA en instead of ϑ, and pen and qen instead of p and q, we reduce the problem to the case that p and q are invertible; and so we may assume without loss of generality that p and q are invertible to start with. Given a projection e from A we have X to show that ϑ = id, and for this it suffices, by VII, to find some positive q in A Proof Note that f , being faithful and pure, is a filter (by 100 VII), and thus of (cid:6)p−1q ϑ(e) qp−1(cid:7) =(cid:6)p−1 ⌈q ϑ(e) q⌉ p−1(cid:7) =(cid:6)p−1 ⌈pep⌉ p−1(cid:7) = e; so by VI, we get that ϑ = id and p−1q is central; and so p and q are centrally similar (by III). (cid:3) IX Proposition A faithful ⋄-positive map f : A → A on a von Neumann algebra A is of the form f = √p(· )√p where p := f (1). the form f ≡ √p ϑ(· )√p for some isomorphism ϑ : A → A . Our task then is with ⌈q⌉ = 1 and f⋄(e) ≡(cid:6)√p ϑ(e)√p(cid:7) = ⌈qeq⌉ for all projections e in A . Since f is ⋄-positive, we have f ≡ ξξ for some ⋄-self-adjoint map ξ : A → A . Since 1 = ⌈f⌉ = f⋄(1) = ξ⋄(ξ⋄(1)) 6 ξ⋄(1) = ⌈ξ⌉ we have ⌈ξ⌉ = is a filter (by 100 VII). Furthermore, 1, and so, ξ, being pure and faithful, as ξ := pξ(1)(· )pξ(1) : A → A is a filter of ξ(1) too, there is an isomor- ξ⋄ = (α⋄)−1 ξ⋄ phism α : A → A with ξ = ξα. Now, ξ⋄α⋄ = ξ⋄ = ξ⋄ = α⋄ implies ξ⋄ = α⋄ ξ⋄α⋄, and f⋄ = (ξξ)⋄ = ξ⋄α⋄ ξ⋄α⋄ = ξ⋄ ξ⋄ = ( ξ ξ)⋄. In other words, (cid:6)√p ϑ(e)√p(cid:7) = f⋄(e) = ( ξ ξ)⋄(e) = ⌈ξ(1) e ξ(1)⌉ for all projections e of A , which implies that ϑ = id by VII, and hence that f = √p (· )√p. (cid:3) 105 To strip from 104 IX the assumption that f be faithful we employ this device: II Definition Given an ncp-map f : A → B between von Neumann algebras we denote by hfi : ⌈f⌉A ⌈f⌉ → ⌈f (1)⌉B⌈f (1)⌉ the unique ncp-map such that A π⌈f ⌉ ⌈f⌉A ⌈f⌉ f hfi B c⌈f (1)⌉ / ⌈f (1)⌉B⌈f (1)⌉ commutes. (Compare this with the definition of [f ] in 98 IX.) Exercise Let f : A → B be an ncp-map. III 1. Show that hfi = π⌈f (1)⌉ ◦ f ◦ c⌈f⌉ (using, perhaps, that π⌈f⌉ ◦ c⌈f⌉ = id). 2. Show that hfi = π⌈f (1)⌉ ◦ cf (1) ◦ [f ]. (Thus hfi(a) =pf (1) [f ](a) pf (1) for all a from ⌈f⌉A ⌈f⌉.) / /   / O O 3. Show that hfi is faithful, and hfi(1) = f (1). 4. Assuming that f is pure, show that hfi is pure, and hence a filter (by 100 VII). Exercise Let f : A → A be an ncp-map, where A is a von Neumann algebra. IV 1. Suppose that f is ⋄-self-adjoint. Recall that ⌈f⌉ = ⌈f (1)⌉, and so hfi : ⌈f⌉A ⌈f⌉ → ⌈f⌉A ⌈f⌉. Prove that hfi is ⋄-self-adjoint. 2. Suppose again that f is ⋄-self-adjoint, and recall from 103 III that f 2 is ⋄-self-adjoint, and (cid:6)f 2(cid:7) = ⌈f⌉. Show that (cid:10)f 2(cid:11) = hfi2. 3. Assuming that f is ⋄-positive, show that hfi is ⋄-positive. Theorem Given a positive element p of a von Neumann algebra A there is a unique ⋄-positive map f : A → A with f (1) = p, namely f = √p(· )√p. Proof We've already seen in 103 II that f = √p(· )√p : A → A is a ⋄- positive map with f (1) = p. Concerning uniqueness, (given arbitrary f ) the map hfi : ⌈p⌉A ⌈p⌉ → ⌈p⌉A ⌈p⌉ from II is ⋄-positive by IV, and faithful by III, and so of the form hfi = √p(· )√p by 104 IX (since hfi(1) = f (1) = p); implying that f = c⌈p⌉ ◦ hfi ◦ π⌈p⌉ = √p ⌈p⌉ (· )⌈p⌉√p = √p(· )√p. (cid:3) Corollary ("Square Root Axiom") Given a positive element p of a von Neu- mann algebra A there is a unique ⋄-positive map g : A → A with g(g(1)) = p, namely g = 4√p (· ) 4√p. Proof Any ⋄-positive map g : A → A with g(g(1)) = p will be of the form g =pg(1) (· )pg(1) by V; so that p = g(g(1)) = g(1)2 implies that g(1) = √p by 23 VII, thereby giving g = 4√p (· ) 4√p. (cid:3) Theorem On the effects of every von Neumann algebra A there is a unique binary operation ∗ such that for all p from [0, 1]A , A. p ∗ 1 = p, B. p ∗ q = f (q) for all q from [0, 1]A for some pure map f : A → A , C. p ∗ (p ∗ q) = (p ∗ p) ∗ q for all q from [0, 1]A , D. p = q ∗ q for some q from [0, 1]A , E. p ∗ e1 6 e⊥2 ⇐⇒ p ∗ e2 6 e⊥1 for all projections e1, e2 from A ; namely, the sequential product, given by p∗ q = √pq√p for all p, q from [0, 1]A . ..104 -- 106.. 161 V VI VII VIII 106 II Proof Let p from [0, 1]A be given. Pick p′ from [0, 1]A with p = p′ ∗ p′ using D, and find a pure map f : A → A with f (q) = p′ ∗ q for all q from [0, 1]A using B. Then f is ⋄-self-adjoint by E, and so f f is ⋄-positive. Since f (f (1)) = p′ ∗ (p′ ∗ 1) = p′ ∗ p′ = p by A, f f = √p(· )√p by 105 V, so p ∗ q = (p′ ∗ p′) ∗ q = p′ ∗ (p′ ∗ q) = f (f (q)) = √pq√p for all q ∈ [0, 1]A by C. (cid:3) III Exercise None of the axioms from I may be omitted (except perhaps D, see IV): except B. 1. Show that p ∗ q := ⌈p⌉ q ⌈p⌉ satisfies all axioms of I except A. 2. Show that p ∗ q := ⌊p⌋ q ⌊p⌋ + pp − ⌊p⌋ qpp − ⌊p⌋ satisfies all axioms 3. Show that if for every effect p of A we pick a unitary up from ⌈p⌉A ⌈p⌉ then ∗ given by p ∗ q = √pu∗p q up√p satisfies A and B. Show that this ∗ obeys C when u2 when u∗p = up. Conclude that when up is defined by up := g(p), where g : [0, 1] → {−1, 1} is any Borel function with g(2/3) = 1 and g(4/9) = −1 the operation ∗ (defined by up as above) satisfies all conditions of I except C. p = up2 , and D when pup = upp, and E 4. Show that there is a Borel function g : [0, 1] → S1 with g(1/2) 6= 1 and g(λ2) = g(λ)2 for all λ ∈ [0, 1], and that ∗ given by p ∗ q = √pg(p)∗ q g(p)√p sat- isfies all conditions of I except E. IV Problem Is there a binary operation ∗ on the effects [0, 1]A of a von Neumann algebra A that satisfies all axioms of I except D? V Remark The axioms for the sequential product (on a single von Neumann algebra) presented here (in I) evolved from the following axioms for all sequential products on von Neumann algebras (∗A )A we previously published in [81]. Ax.1 For every effect p of a von Neumann algebra A there is a filter c : C → A of p and a corner π : A → C of ⌊p⌋ with p∗A q = c(π(q)) for all q ∈ [0, 1]A . Ax.2 p ∗A (p ∗A q) = (p ∗A p) ∗A q for all effects p and q from a von Neumann algebra A . Ax.3 f (p ∗A q) = f (p) ∗B f (q) for every nmisu-map f : A → B between von Neumann algebras and all effects p and q from A . Ax.4 p ∗A e1 6 e⊥2 ⇐⇒ p ∗A e2 6 e⊥1 for every effect p from a von Neumann algebra A and projections e1 and e2 from A . Note that Ax.2 and Ax.4 are mutatis mutandis the same as axioms C and E, respectively, and Ax.1 is essentially the same as the combination of axioms A and B. In other words, we managed to get rid of Ax.3 -- and with it the need to axiomatise all sequential products simultaneously -- at the slight cost of adding axiom D, though that one might be superfluous as well (see IV). We refer to §VI of [81] for comments on the relation of our axioms with those of Gudder and Lat´emoli`ere [26] and for some more pointers to the literature. 4.2 Tensor product 107 The tensor product of von Neumann algebras A and B represented on Hilbert spaces H and K , respectively, is usually defined as the von Neumann subalge- bra of B(H ⊗ K ) generated by the operators on H ⊗ K of the form A ⊗ B where A ∈ A and B ∈ B. In line with the representation-avoiding treatment of von Neumann algebras from the previous chapter we'll take an entirely dif- ferent approach by defining the tensor product of von Neumann algebras A and B abstractly as an miu-bilinear map ⊗ : A × B → A ⊗ B whose range generates A ⊗ B and admits sufficiently many product functionals (see 108 II); we'll only resort to the concrete representation of the tensor product mentioned above to show that such an abstract tensor product actually exists (see 111 VII). Moreover, we'll show that the tensor product has a universal property 112 XI yielding bifunctors on W∗cpsu and W∗miu (see 115 IV) turning them into a monoidal categories (see 119 V). In the next chapter, we'll see that (W∗miu)op is even monoidal closed (see 125 VIII). This fact is one ingredient of our model for the quantum lambda calculus from [11] built of von Neumann algebras, but more of that later. 4.2.1 Definition Definition A bilinear map β : A × B → C between von Neumann algebras is 108 1. unital when β(1, 1) = 1, 2. multiplicative if β(ab, cd) = β(a, c)β(b, d) for all a, b ∈ A , c, d ∈ B, 3. involution preserving if β(a, b)∗ = β(a∗, b∗) for all a ∈ A , b ∈ B. 4. (This list is extended in 112 II.) We abbreviate these properties as in 10 II, and say, for instance, that β is miu- bilinear when it is unital, multiplicative and involution preserving. Definition A miu-bilinear map γ : A × B → T between von Neumann algebras is a tensor product of A and B when it obeys the following three conditions. II ..106 -- 108.. 163 1. The range of γ generates T (which means in this case that the linear span of the range of γ is ultraweakly dense in T .) This implies that for all f ∈ A∗ and g ∈ B∗ there is at most one h ∈ T∗ with, for all a ∈ A and b ∈ B, h(γ(a, b)) = f (a) g(b), which we'll call the product functional for f and g, and denote by γ(f, g) (when it exists). 2. For all np-functionals σ : A → C and τ : B → C the product functional γ(σ, τ ) : T → C exists and is positive. 3. The product functionals γ(σ, τ ) of np-functionals σ and τ form a faithful collection of np-functionals on T . (We'll see a slightly different characterisation of the tensor in which not all product functionals of np-functionals are required to exist upfront in 116 VII.) III Remark This compact definition of the tensor product leaves four questions unanswered: whether such a tensor product of two von Neumann algebras al- ways exists, whether it has some universal property, whether it is unique in some way, and whether it coincides with the usual definition. We'll shortly address all four questions. 4.2.2 Existence 109 We'll start with the existence of a tensor product of von Neumann algebras for which we'll first need the tensor product of Hilbert spaces. II Definition We'll call a bilinear map γ : H × K → T between Hilbert spaces a tensor product when it obeys the following two conditions. 1. The linear span of the range of γ is dense in T . 2. hγ(x, y), γ(x′, y′)i = hx, x′ihy, y′i for all x, x′ ∈ H and y, y′ ∈ K . III Exercise We're going to prove that every pair of Hilbert spaces H and K admits a tensor product. 1. Given sets X and Y show that γ : ℓ2(X) × ℓ2(Y ) → ℓ2(X × Y ) given by γ(f, g) = ( f (x) g(y) )x∈X,y∈Y is a tensor product of ℓ2(X) and ℓ2(Y ). 2. Show that a subset E of a Hilbert space H is an orthonormal basis (see 39 IV) iff the map T : ℓ2(E ) → H given by T (x) = Pe∈E xee is an isometric isomorphism. 3. Show that any pair H and K of Hilbert spaces has a tensor product (using the fact that every Hilbert space has an orthonormal basis). Proposition Let γ : H × K → T be a tensor product of Hilbert spaces. IV 1. We have kγ(x, y)k = kxkkyk for all x ∈ H and y ∈ K . 2. Given orthonormal bases E and F of H and K , respectively, the set G := { γ(e, f ) : e ∈ E , f ∈ F } is an orthonormal basis for T . Proof 1 We have kγ(x, y)k2 = hγ(x, y), γ(x, y)i = hx, xi hy, yi = kxk2kyk2. 2 Since hγ(e, e′), γ(f, f′)i = he, e′ihf, f′i where e, e′ ∈ E and f, f′ ∈ F , the set G is clearly orthonormal. To see that G is maximal (and thus a basis) it suffices to show that the span of G is dense in T , and for this it suffices to show that each γ(x, y) where x ∈ H and y ∈ K is in the closure of the span of G . Now, since y = Pf∈F hf, yi f , by 39 IV and hx, (· )i is bounded by 1 we have γ(x, y) =Pf∈F hy, fi γ(x, f ). Since similarly γ(x, f ) =Pe∈E he, xi γ(e, f ) for all f ∈ F , we see that γ(x, y) is indeed in the closure of the span of G . (cid:3) Definition We'll say that a bilinear map β : H × K → L between Hilbert spaces is ℓ2-bounded by B ∈ [0,∞) when kXi β(xi, yi)k2 6 B2Xi,j for all x1, . . . , xn ∈ H and y1, . . . , yn ∈ K . Remark We added the prefix "ℓ2-" to clearly distinguish it from the boundedness of (sesquilinear) forms from 36 IV, which one might call "ℓ∞-boundedness." hxi, xjihyi, yji This distinction is needed since for example given a Hilbert space H the bilinear map (f, x) 7→ f (x) : H ∗ × H → C is always ℓ∞-bounded in the sense that f (x) 6 kfkkxk for all f ∈ H ∗ and x ∈ H , but it is not ℓ2-bounded when H is infinite dimensional Theorem A tensor product γ : H × K → T of Hilbert spaces is ℓ2-bounded, and initial as such in the sense that for any by B ∈ [0,∞) ℓ2-bounded bilinear map β : H × K → L into a Hilbert space L there is a unique bounded linear map βγ : T → L with βγ(γ(x, y)) = β(x, y) for all x ∈ H and y ∈ K . Moreover, kβγk 6 B for such β. ..108 -- 110.. 165 V 110 II III IV Proof Note that γ is ℓ2-bounded, since for all x1, . . . , xn ∈ H , y1, . . . , yn ∈ K , we have kPi γ(xi, yi)k2 =Pi,j hγ(xi, yi), γ(xj, yj)i =Pi,j hxi, xjihyi, yji. Let E and F be orthonormal bases for H and K , respectively. Then since { γ(e, f ) : e ∈ E , f ∈ F } is an orthonormal basis for T by 109 IV, and βγ is fixed on it by βγ(γ(e, f )) = β(e, f ), uniqueness of βγ is clear. Concerning existence of βγ, note that since t =Pe∈E ,f∈F hγ(e, f ), ti γ(e, f ) for all t ∈ T by 39 IV, we'd like to define βγ by βγ(t) = Xe∈E , f∈F hγ(e, f ), ti γ(e, f ); (4.2) but before we can do this we must first check that this series converges. To this end, note that since β is ℓ2-bounded by B we have, given t ∈ T , (cid:13)(cid:13)Xe∈E, f∈F hγ(e, f ), ti β(e, f )(cid:13)(cid:13) 2 2 = (cid:13)(cid:13)Xe∈E, f∈F 6 B2 Xe′,e∈E, f ′,f∈F = B2 Xe∈E, f∈F β(e, hγ(e, f ), ti f )(cid:13)(cid:13) he′, ei ht, γ(e′, f′)i hf′, fi hγ(e, f ), ti hγ(e, f ), ti2 for all finite subsets E ⊆ E and F ⊆ F . Since ktk2 = Pe∈E , f∈F hγ(e, f ), ti2 by Parseval's identity (39 IV), we see that the series from (4.2) converges defin- ing βγ(t), and, moreover, that kβγ(t)k2 6 B2ktk2. The resulting map βγ : T → L is clearly linear, and bounded by B. Further, βγ(γ(e, f )) = β(e, f ) for all e ∈ E and f ∈ F implies that βγ(γ(x, y)) = β(x, y) for all x ∈ H and y ∈ K , and so we're done. (cid:3) Exercise Show that the tensor product of Hilbert spaces H and K is unique in the sense that given tensor products γ : H × K → T and γ′ : H × K → T ′ there is a unique isometric linear isomorphism ϕ : T → T ′ with γ′(x, y) = ϕ(γ(x, y)) for all x ∈ H and y ∈ K . V VI Notation Now that we've established that the tensor product of Hilbert spaces H and K exists and is unique (up to unique isomorphism) we just pick one and denote it by ⊗ : H × K → H ⊗ K . Essentially to turn ⊗ into a functor on the category of Hilbert spaces in V, we'll need the following result (known as part of Schur's product theorem), which will be useful several times later on. 111 II III Lemma For any natural number N the entrywise product (anmbnm) of positive N × N -matrices (anm) and (bnm) over C is positive. Proof Let z1, . . . , zN ∈ C be given. To show that (anmbnm) is positive, it suffices by 33 II to prove that Pn,m znanmbnmzm > 0 for all n, m. Since (anm) is a positive element of the C∗-algebra MN it's of the form (anm) = C∗C for some N × N -matrix C ≡ (cnm) over C, so anm = Pk cknckm for all n, m. Similarly, there a N × N -matrix (dnm) over C with bnm =Pℓ dℓndℓm for all n, m. Then zn cknckm dℓndℓmzm Xn,m znanmbnmzm = Xn,m,k,ℓ = Xk,ℓ(cid:16)Xn = Xk,ℓ (cid:12)(cid:12)(cid:12)Xn znckndℓn(cid:17)(cid:16)Xm znckndℓn(cid:12)(cid:12)(cid:12) > 0, 2 zmckmdℓm(cid:17) and so (anmbnm) is positive. (cid:3) Exercise Given square matrices (anm) 6 (anm) and (bnm) 6 (bnm) over C of the same dimensions, show that ( anmbnm ) 6 ( anmbnm ). Proposition Given bounded linear maps A : H → H ′ and B : K → K ′ between Hilbert spaces there is a unique bounded linear map A ⊗ B : H ⊗ K → H ′ ⊗ K ′ with (A ⊗ B)(x ⊗ y) = (Ax) ⊗ (By) for all x ∈ H and y ∈ K . Proof In view of 110 III the only thing we need to prove is that the bilinear map ⊗◦(A×B) : H ×K → H ⊗K is ℓ2-bounded (for then A⊗B = (⊗◦(A×B) )⊗.) So let x1, . . . , xn ∈ H and y1, . . . , yn ∈ K be given, and note that (Axi) ⊗ (Byi)k2 kXi (⊗ ◦ (A × B))(xi, yi)k2 = kXi = Xi,j 6 kAk2kBk2Xi,j hAxi, Axji hByi, Byji hxi, xjihyi, yji , IV V VI so ⊗ ◦ (A × B) is bounded by kAkkBk. The last step in the display above is justified by IV, and the inequalities (hAxi, Axji ) 6 (kAk2 hxi, xji ) and (hByi, Byji ) 6 (kBk2 hyi, yji ). (cid:3) Theorem Let A and B be von Neumann algebras of bounded operators on Hilbert spaces H and K , respectively. Sending operators A ∈ A and B ∈ B to A ⊗ B : H ⊗ K → H ⊗ K from V gives an miu-bilinear map VII ⊗ : A × B −→ B(H ⊗ K ). Letting T be the von Neumann subalgebra of B(H ⊗ K ) generated by the range of ⊗, the restriction γ : A × B → T of ⊗ is a tensor product of A and B. ..110 -- 111.. 167 VIII Proof We'll check that the three conditions of 108 II hold; we leave it to the IX X XI reader to verify that ⊗ is miu-bilinear. (Condition 1) The range of γ being the same as the range of ⊗ generates T simply by the way T was defined. (Condition 2) Let σ : A → C and τ : B → C be np-maps. We must find an np- functional ω on T with ω(A⊗ B) = σ(A)τ (B) for all A ∈ A , B ∈ B. Note that by 89 IX σ and τ are of the form σ ≡Pn hxn, (· )xni and τ ≡Pn hyn, (· )yni for some x1, x2, . . . ∈ H and y1, y2, . . . ∈ K withPn kxnk2 < ∞ andPm kymk2 < ∞. So as Pn,m kxn ⊗ ymk2 ≡Pn kxnk2 Pm kymk2 < ∞, we can define an np- functional ω on T by ω(T ) :=Pn,m hxn ⊗ ym, T xn ⊗ ymi; which does the job: ω(A ⊗ B) =Pn,m hxn, Axnihym, Bymi = σ(A)τ (B) for all A ∈ A and B ∈ B. Pm,n hxn ⊗ yn, (· ) xn ⊗ ymi for some x1, x2, . . . ∈ H and y1, y2, . . . ∈ K (and, conversely, it's easily seen that a functional of that form is a product functional). It suffices, then, to show that the subset of product functionals of the form hx ⊗ y, (· )x ⊗ yi where x ∈ H and y ∈ K is faithful. To this end, let T ∈ T+ with hx ⊗ y, T x ⊗ yi = 0 for all x ∈ H and y ∈ K be given in order to show that T = 0. Note that since k√T x ⊗ yk2 = hx ⊗ y, T x ⊗ yi = 0, and so √T x ⊗ y = 0 for all x ∈ H , y ∈ K , we have √T = 0 (since the linear span of the x ⊗ y is dense in H ⊗ K ), and thus T = 0. (Condition 3) It remains to be shown that the product functionals on T form a faithful collection. These functionals are -- as we've just seen -- all of the form (cid:3) XII Exercise Given von Neumann algebras A and B (which are not a priori rep- resented on Hilbert spaces) construct a tensor product γ : A × B → T of A and B using 48 VIII and VII. 4.2.3 Universal Property 112 Before we bring our categorical faculties to bear upon the tensor product for von Neumann algebras we quickly review the (algebraic) tensor product of plain vector spaces V and W first -- it is a vector space V ⊙ W equipped with a bilinear mapping ⊙ : V × W → V ⊙ W which is universal in the sense that for every bilinear mapping β : V × W → Z into some vector space Z there is a unique linear map β⊙ : V ⊙ W → Z with β⊙(v ⊙ w) = β(v, w) for all v ∈ V and w ∈ W . This property uniquely determines the algebraic tensor product in the sense that for any bilinear map ⊙ : V × W → V ⊙ W into a vector space V ⊙ W which shares this property there is a unique linear isomorphism ϕ : V ⊙ W → V ⊙ W with ϕ(v ⊙ w) = v ⊙ w for all v ∈ V and w ∈ W . In fact, one may take this property as a neat abstract definition of the algebraic tensor product. However, to see that the darn thing actually exists, one still needs a concrete description such as this one: take given a basis B of V and a basis C of W the bilinear map ⊙ on V × W to the vector space (B× C)· C with basis B × C determined by b ⊙ c = (b, c) for b ∈ B and c ∈ C. This shows us not only that the algebraic tensor product exists, but also that ⊙ is injective (among other things). This is all, of course, well known, and we already saw in 110 III that the tensor product for Hilbert spaces has a similar universal property; the interesting thing here is that with some work one can see that a tensor product γ : A × B → T of von Neumann algebras A and B has a similar universal property too! We'll see that any bilinear map β : A × B → C into a von Neumann algebra C which is sufficiently regular extends uniquely along γ to a ultraweakly continuous map βγ : T → C, where regular will mean that the extension β⊙ : A ⊙ B → C from the algebraic tensor product is ultraweakly continuous and bounded with respect to the norm and ultraweak topology induced on A ⊙ B by T via γ. To prevent a circular description here, we'll first describe the norm and ultraweak topology that the tensor product induces on A ⊙ B directly, which turns out to be independent (as it should) from the choice of γ. This description is essentially based on the fact that the product functionals on T are centre separating; and that this determines both norm and ultraweak topology is just a general observation concerning centre separating sets, as we saw in 90 II. Definitions Let A and B be von Neumann algebras. II 1. A basic functional is a map ω : A ⊙ B → C with ω ≡ (σ ⊙ τ )(t∗(· )t) for some np-maps σ : A → C, τ : B → C, and t ∈ A ⊙ B. A simple functional is a finite sum of basic functionals. 2. Each basic functional ω : A ⊙ B → C gives us an operation [· , · ]ω, that will turn out to be an inner product in V by [s, t]ω := ω(s∗t) (cf. 30 II), and an associated semi-norm denoted by ktkω := [t, t]1/2 ω = ω(t∗t)1/2. The tensor product norm on A ⊙ B is the norm (see VIII) given by ktk = supω ktkω, where ω ranges over all basic functionals on A ⊙ B with ω(1) 6 1. 3. Note that having endowed A ⊙ B with the tensor product norm we can speak of bounded functionals on A ⊙ B, and the operator norm on them; and note that the basic and simple functionals are bounded. The ultraweak tensor product topology is the least topology on A ⊙ B that makes all operator norm limits of simple functionals continuous. 4. A bilinear map β : A × B → C to a von Neumann algebra C is called (a) (continues the list from 108 I) (b) bounded when the unique extension β⊙ : A ⊙ B → C is bounded, ..111 -- 112.. 169 a1, . . . , aN ∈ A , b1, . . . , bN ∈ B, and c1, . . . , cN ∈ C . product topology on A ⊙ B and the ultraweak topology on C , (c) normal when β⊙ is continuous with respect to the ultraweak tensor (d) completely positive when Pi,j c∗i β(a∗i aj, b∗i bj) cj > 0 for all tuples IIa Warning While we'll be able to see shortly that any bilinear map β : A ×B → C between von Neumann algebras that is normal is jointly ultraweakly continuous as well, (as a consequence of X,) we do not know -- but doubt -- that the converse holds. So to clearly differentiate between these two possibly different properties, we decided to call the former "normality" instead of the more likely "ultraweak continuity", stretching the use of the word "normal" beyond its usual domain of positive (bilinear) maps. Lemma Given C∗-algebras A and B we have (σ⊙ τ )(t∗t) > 0 for all t ∈ A ⊙ B and p-maps σ : A → C and τ : B → C. IV Proof Note that writing t ≡Pn an⊙bn, where a1, . . . , aN ∈ A , b1, . . . , bN ∈ B, we have (σ ⊙ τ )(t∗t) = Pn,m σ(a∗nam) τ (b∗nbm). Since (a∗nam) is a positive matrix over A , and σ : A → C is completely positive (by 34 IX), the ma- trix (σ(a∗nam)) is positive. Since (τ (b∗nbm)) is positive by the same token, the entrywise product ( σ(a∗nam) τ (b∗nam) ) is positive too (by 111 II). Whence (cid:3) III VI (σ ⊙ τ )(t∗t) = Pn,m σ(a∗nam) τ (b∗nbm) > 0. Exercise Use III to show that [· , · ]ω from II is an inner product. Lemma Product functionals on A ⊙ B formed from separating collections Ω and Ξ of linear functionals on C∗-algebras A and B, respectively, are separating in the sense that given t ∈ A ⊙ B the condition that (σ⊙ τ )(t) = 0 for all σ ∈ Ω and τ ∈ Ξ entails that t = 0. VII Proof Write t ≡ Pn an ⊙ bn for some a1, . . . , aN ∈ A and b1, . . . , bN ∈ B. Note that (by replacing them if necessary) we may assume that the a1, . . . , aN are linearly independent. Let τ ∈ Ξ be given. Since 0 = (σ ⊙ τ )(t) = Pn σ(an)τ (bn) = σ(Pn anτ (bn) ) for all σ from the separating collection Ω, we have 0 = Pn anτ (bn), and so -- a1, . . . , aN being linearly independent -- we get 0 = τ (b1) = ··· = τ (bN ). Since this holds for any τ in the separating collection Ξ we get 0 = b1 = ··· = bN , and thus t =Pn an ⊙ bn = 0. (cid:3) VIII Exercise Show that the tensor product norm from II is, indeed, a norm. IX Exercise Note that given np-functionals σ : A → C and τ : B → C on von Neumann algebras, the functional σ ⊙ τ : A ⊙ B → C is ultraweakly continuous and bounded, almost by definition. Show that f ⊙ g is bounded and ultraweakly continuous too for all f ∈ A∗ and g ∈ B∗ (perhaps using 72 XI). Exercise We're going to show that the ultraweak tensor product topology and V X tensor product norm from II actually describe the norm and ultraweak topology on A ⊙ B induced by a tensor product A × B → T (via γ⊙) by establishing the two closely related facts that γ⊙ : A ⊙ B → T is an isometry and an ultraweak embedding, and that certain functionals ω : A ⊙ B → C can be extended uniquely to T along γ⊙. 1. Show using 90 II that the collection Ω of np-functionals on T of the form γ(σ, τ )(γ⊙(s)∗(· )γ⊙(s)), where σ : A → C, τ : B → C are np-functionals and s ∈ A ⊙ B, is order separating, and that every np-functional on T is the operator norm limit of finite sums of functionals from Ω. Show that ω ◦ γ⊙ is a basic functional (see II) for every ω ∈ Ω, and that every basic functional is of this form for some unique ω ∈ Ω. 2. Show that the subset Ω1 of Ω of unital maps is order separating, and so determines the norm on T via kak2 = ka∗ak = supω∈Ω1 ω(a∗a) for all a ∈ T (see 21 VII). Prove that kγ⊙(s)k = supω∈Ω1 ω(s∗s)1/2 = supω∈Ω1 kskω◦τ⊙ = ksk for all s ∈ A ⊙ B, and conclude that γ⊙ is an isometry. 3. Show that kf ◦ γ⊙k 6 kfk for every f ∈ T∗, and deduce from this that when ω : T → C is an np-functional its restriction ω ◦ γ⊙ is the operator norm limit of simple functionals on A ⊙ B implying that ω ◦ γ⊙ -- and thus γ⊙ itself -- is ultraweakly continuous. 4. In order to show that γ⊙ is an ultraweak embedding, we'll need the equality kf ◦ γ⊙k = kfk for all f ∈ T∗. In order to show this in turn, recall (from 86 IX) that there is a partial isometry u in T with f (u) = kfk (see 86 XIV). Show that given ε > 0 there is a net (sα)α in A ⊙ B with ksαk 6 1 + ε for all α such that γ⊙(sα) converges ultrastrongly to t as α → ∞ (cf. 74 VI). Deduce that kfk = f (u) = f (u) = limα f (γ⊙(sα)) 6 kf ◦ γ⊙k(1 + ε), and conclude that kfk = kf ◦ γ⊙k. 5. Show that any functional ω′ : A ⊙ B → C that is the operator norm limit of simple functionals on A ⊙ B can be extended uniquely along γ⊙ to an np-functional on T (using the fact that the operator norm limit of np-functionals is an np-functional again, see 87 III). Deduce from this that γ⊙ is a ultraweak topological embedding. (Note that by 77 V any bounded ultraweakly continuous functional on A ⊙ B can be extended uniquely to a normal functional on T .) ..112.. 171 XI Theorem A tensor product γ : A × B → T of von Neumann algebras A for every normal bounded bilinear map and B has this universal property: β : A × B → C to a von Neumann algebra C there is a unique ultraweakly continuous map βγ : T → C with βγ ◦ γ = β. Moreover, kβγk = kβ⊙k. XII Proof Since β⊙ : A ⊙ B → C is ultraweakly continuous and bounded, and A ⊙ B can by X be considered an ultraweakly dense ∗-subalgebra of T via γ⊙, the theorem follows from 77 V except for some trivial details. (cid:3) 113 We'll need some observations concerning completely positive bilinear maps. II Exercise Show that a mi-bilinear map β : A × B → C between von Neumann algebras is completely positive. III Notation Given a bilinear map β : A ×B → C between von Neumann algebras, we define MN β : MN A × MN B → MN C by (MN β)(A, B) = (β(Aij , Bij))ij for each N . IV Exercise Show that for a bilinear map β : A × B → C between von Neumann algebras the following are equivalent. 1. β is completely positive. 2. MN β is completely positive for each N . 114 3. (MN β)(A, B) > 0 for all A ∈ MN (A )+, B ∈ MN (B)+ and N . Deduce as a corollary that h◦β◦(f×g) is completely positive when f : A ′ → A , g : B′ → B and h : C → C ′ are cp-maps between von Neumann algebras. Exercise Let γ : A × B → T be a tensor product of von Neumann algebras, β : A × B → C a normal bounded bilinear map, and βγ : T → C its extension along γ⊙ from 112 XI. Show that 1. βγ is multiplicative iff β is multiplicative (see 112 II); 2. βγ is involution preserving iff β is involution preserving; 3. βγ is unital iff β is unital; 4. βγ is positive iff Pi,j β(a∗i aj, b∗i bj) > 0 for all tuples a1, . . . , aN from A and b1, . . . , bN from B; 5. βγ is completely positive iff β is completely positive. II Exercise Show that the tensor product of von Neumann algebras A and B is unique in the sense that when γ : A × B → T and γ′ : A × B → T ′ are tensor products of A and B, then there is a unique nmiu-isomorphism ϕ : T → T ′ with ϕ(γ(a, b)) = γ′(a, b) for all a ∈ A and b ∈ B. 4.2.4 Functoriality Notation Now that we've established that that the tensor product of von Neu- mann algebras A and B exists and is unique (up to unique nmiu-isomorphism) we just pick one and denote it by ⊗ : A × B → A ⊗ B. Proposition Given ncp-maps f : A → C and g : B → D between von Neumann algebras there is a unique ncp-map f ⊗ g : A ⊗ B → C ⊗ D with 115 II III (f ⊗ g)(a ⊗ b) = f (a) ⊗ f (b) for all a ∈ A and b ∈ B. Moreover, 1. f ⊗ g is multiplicative when f and g are multiplicative; 2. f ⊗ g is involution preserving when f and g are involution preserving; and 3. f ⊗ g is (sub)unital when f and g are (sub)unital. Proof As uniqueness of f ⊗ g is rather obvious, we leave it at that. To establish existence of f ⊗ g, it suffices to show that the bilinear map β : A × B → C ⊗ D given by β(a, b) = f (a) ⊗ g(b), which is completely positive by 113 IV, is bounded and normal; because then we may take f ⊗ g := β⊗ as in 112 XI and all the properties claimed for f ⊗ g will then follow with the very least of effort from 114 I. To see that β is bounded, we'll prove that kβ⊙(s)k 6 kfkkgkksk given an element s of A ⊗ B, and for this it suffices (by the definition of the tensor product norm, 112 II) to show that ω(β⊙(s)∗β⊙(s)) 6 kfk2kgk2ksk2 given a basic functional ω on A ⊙ B with ω(1) 6 1. We'll prove in a moment that kω ◦ β⊙k 6 kfkkgk and β⊙(s)∗β⊙(s) 6 kfkkgkβ⊙(s∗s), because with these two claims we get ω(β⊙(s)∗β⊙(s)) 6 kfkkgkω(β⊙(s∗s)) 6 kfkkgkkω ◦ β⊙kksk2 6 kfk2kgk2ksk2 -- which is the result desired. Concerning the first promise, that kω ◦ β⊙k 6 kfkkgk, note that writing ω ≡ (σ ⊙ τ )(t∗(· )t), where σ and τ are np-maps on C and D, respectively, and t ≡Pij ci ⊙ di is from C ⊙ D, we have and so ω ◦ β⊙ is ultraweakly continuous and bounded by 112 IX, because the σ(c∗i f (· )cj) and τ (d∗i g(· )dj) are bounded ultraweakly continuous functionals. Although the bound for ω ◦ β⊙ thus obtained is in all probability nowhere near kfkkgk, it does allow us by 112 XI to extend ω ◦ β⊙ to an ultraweakly continuous functional ω′ := (ω ◦ β)⊗ on C ⊗ D with the same norm, kω′k = kω◦β⊙k. Since this extension ω′ is completely positive (because β and thus ω◦β are completely positive, see 113 IV) its norm is by 34 XVI given by kω′k = ω′(1) ≡ ω ◦ β⊙ =Pij σ(c∗i f (· )cj) ⊙ τ (d∗i g(· )dj), ..112 -- 115.. 173 ω(f (1)⊗ g(1)) 6 kfkkgk, where we used that ω(1) 6 1. Thus kω◦ β⊙k = kω′k 6 kfkkgk, as was claimed. Incidentally, since each ω◦β⊙ is ultraweakly continuous, so is β⊙, and thus β is normal. The only thing that remains is to make good on our last promise, that β⊙(s)∗β⊙(s) 6 kfkkgkβ⊙(s∗s). To this end, write s ≡ Pi ai ⊙ bi, and consider the matrices A and B given by A :=   a1 a2 0 0 ... ... 0 0 ··· an ··· 0 ... . . . 0 ···   B :=   b1 0 ... 0 b2 0 ... 0 ··· ··· . . . ··· , bn 0 ... 0   and the cp-map h : Mn(C ⊗D) → C⊗D given by h(C) = h(1, . . . , 1), C(1, . . . , 1)i = Pij Cij . We make these arrangements so that we may apply the inequality (Mnf )(A)∗(Mnf )(A) 6 k(Mnf )(1)k(Mnf )(A∗A) easily derived from 34 XIV. Indeed, noting also k(Mnf )(1)k = kf (1)k = kfk, we have β⊙(s)∗β⊙(s) = Pij f (ai)∗f (aj) ⊗ g(bi)∗g(bj) = h( (Mnf )(A)∗(Mnf )(A) (Mn⊗) (Mng)(B)∗(Mng)(B) ) 6 kfkkgkh( (Mnf )(A∗A) (Mn⊗) (Mng)(B∗B) ) = kfkkgk Pij f (a∗i aj) ⊗ g(b∗i bj) = kfkkgk β⊙(s∗s), which concludes this proof. (cid:3) IV Exercise Show that the assignments (A , B) 7→ A ⊗ B, and (f, g) 7→ f ⊗ g give a bifunctor ⊗ : C × C → C where C can be W∗miu, W∗cp, W∗cpu or W∗cpsu. V Proposition Given injective nmiu-maps f : A → C and g : B → D, the nmiu- map f ⊗ g : A ⊗ B → C ⊗ D is injective. VI Proof The trick is to consider the von Neumann subalgebra T generated by the elements of C ⊗ D of the form f (a)⊗ g(b) where a ∈ A and b ∈ B, and to show that the miu-bilinear map γ : A × B → T given by γ(a, b) = f (a) ⊗ g(b) is a tensor product of A and B. Indeed, if this is achieved, then there is, by 114 II, a unique nmiu-map ϕ : A ⊗ B → T with ϕ(a ⊗ b) = γ(a, b) = f (a) ⊗ g(b), so that the following diagram commutes. f×g A × B ⊗ A ⊗ B γ &▼▼▼▼▼▼▼▼▼▼▼ ϕ / T ⊆ C × D ⊗ / C ⊗ D The map on the bottom side of this rectangle above is none other than f ⊗ g, and is thus, being the composition of the isomorphism ϕ with the inclusion T ⊆ C ⊗ D, injective. / / &     / / It remains to be shown that γ is a tensor product, that is, obeys the con- ditions from 108 II. Condition 1 holds simply by definition of T . To see that γ obeys condition 2, let np-functionals σ : A → C and τ : B → C be given; we must find an np-functional γ(σ, τ ) on T with γ(σ, τ )(a ⊗ b) = γ(a, b). By ultraweak permanence σ and τ can be extended along f and g, re- spectively, see 89 XII, giving us np-functionals σ : C → C and τ : D → C with σ = σ ◦ f and τ = τ ◦ g. Now simply take γ(σ, τ ) to be the restriction of σ ⊗ τ to T , which does the job. Finally, concerning condition 3, let z be a central projection of T with γ(σ, τ )(z) = 0 for all σ and τ of aforementioned type. We must show that z = 0, and for this it suffices to show that (σ ⊗ τ )(z) = 0 for all np-functionals σ and τ on C and D, respectively. Since for such σ and τ we have γ(σ, τ )(γ(a, b)) = σ(f (a)) τ (g(b)) = (σ ⊗ τ )(γ(a, b)) for all a ∈ A and b ∈ B, we have γ(σ, τ )(t) = (σ ⊗ τ )(t) for all t ∈ T , and, in particular, 0 = γ(σ, τ )(z) = (σ ⊗ τ )(z). Hence z = 0. (cid:3) 4.2.5 Miscellaneous Properties Lemma Given von Neumann algebras A and B, we have kf ⊗ gk = kfkkgk for all f ∈ A∗ and g ∈ B∗. Proof The trick is to use the polar decomposition for normal functionals, 86 IX. On its account we can find partial isometries u ∈ A and v ∈ B such that f (u(· )) and g(v(· )) are positive, and f ≡ f (uu∗(· )), g ≡ g(vv∗(· )). Then u ⊗ v is a partial isometry such that (f ⊗ g)((u ⊗ v)(· )) is positive, and f ⊗ g = (f ⊗g)( (u⊗v) (u⊗v)∗ (· ) ) so that kf ⊗gk = (f ⊗g)(u⊗v) = f (u)g(v) = kfkkgk by 86 XIV. (cid:3) 116 II Exercise There are some easily obtained facts concerning the tensor prod- uct A ⊗ B of von Neumann algebras that nevertheless deserve explicit mention. 1. Show that a ⊗ b > 0 for all a ∈ A+ and b ∈ B+; and conclude that a1 ⊗ b1 6 a2 ⊗ b2 for all a1 6 a2 from A and b1 6 b2 from B. III 2. Show that ka ⊗ bk = kakkbk for all a ∈ A and b ∈ B. Conclude that ⊗ : A × B → A ⊗ B is norm continuous. (Warning: as ⊗ is not linear this is not entirely trivial.) 3. Show that ⊗ : A∗ × B∗ → (A ⊗ B)∗ is norm continuous (using I). 4. Show that ⊗ : A × B → A ⊗ B is ultraweakly continuous. (Hint: since we already know that ⊗⊙ : A ⊙ B → A ⊗ B is ultraweakly continuous, by 112 X, an equivalent question is whether ⊙ : A × B → A ⊙ B is ultraweakly continuous, which may be boiled down to the fact ..115 -- 116.. 175 that (a, b) 7→ Pij σ(a∗i aaj) τ (b∗i bbj) : A × B → C is ultraweakly contin- uous, where σ and τ are np-functionals on A and B, respectively, and a1, . . . , an ∈ A , and b1, . . . , bn ∈ B.) 5. Show that a ⊗ (· ) : B → A ⊗ B is a ncp-map for every a ∈ A , and that 1 ⊗ (· ) : B → A ⊗ B is an nmiu-map. IIIa The following observation will come in very handy later on when we prove that A ⊗ (B ⊕ C ) ∼= A ⊗ B ⊕ A ⊗ C , and A ⊗ (B ⊗ C ) ∼= (A ⊗ B) ⊗ C , see 119 IV, and 117 III. IV Proposition Let A and B be von Neumann algebras. V 1. If S and T are subsets of A and B, respectively, whose linear span is ultra- weakly dense, then the linear span of { s⊗ t : s ∈ S, t ∈ T } is ultraweakly dense in A ⊗ B. 2. If Ω and Θ are centre separating collections of np-functionals on A and B, respectively, then { ω ⊗ ϑ : ω ∈ Ω, ϑ ∈ Θ } is centre separating for A ⊗ B. Proof Concerning 1: Let S′ and T ′ denote the linear spans of S and T , respec- tively. Since linear combinations of elements of A ⊗ B of the form a ⊗ b lie ultraweakly dense in A ⊗B where a ∈ A and b ∈ B, it suffices to show that such element a⊗ b is the ultraweak limit of elements of the form s′ ⊗ t′ where s′ ∈ S′ and t′ ∈ T ′ (because such s′⊗ t′ are, of course, a linear combinations of elements of the form s ⊗ t where s ∈ S and t ∈ T .) This is indeed the case as there are nets (s′α)α and (t′β)β in S′ and T ′ that converge ultraweakly to a and b, respectively, and so, because ⊗ is ultraweakly continuous by III, we see that s′α⊗t′β converges ultraweakly to a⊗b as α, β → ∞. Concerning 2, let t be a positive element of A ⊗ B with (ω⊗ ϑ)(s∗ts) = 0 for all ω ∈ Ω, ϑ ∈ Θ, and s ∈ A ⊗ B; we must show that t = 0. For this it suffices to show that (σ ⊗ τ )(t) = 0 for all np-functionals σ : A → C and τ : B → C (since the product functionals σ ⊗ τ form a faithful collection.) Now, since Ω is centre separating such σ may by 90 II be obtained as operator norm limit of finite sums of functionals of the form ω(a∗(· )a) where ω ∈ Ω and a ∈ A . Since an np-functional τ : B → C can be obtained in a similar fashion from Θ, and ⊗ : A∗ ⊗ B∗ → (A ⊗ B)∗ is operator norm continuous (by III), we see that a product functional σ ⊗ τ can be obtained as the operator norm limit of finite sums of functionals of the form ω(a∗(· )a) ⊗ ϑ(b∗(· )b) ≡ (ω ⊗ ϑ)( (a ⊗ b)∗ (· ) (a ⊗ b) ); and since those functionals map t to 0, by assumption, we conclude that (σ ⊗ τ )(t) = 0 too. (cid:3) VI To obtain certain examples the following characterisation of the tensor product of von Neumann algebras proves useful. Theorem Given centre separating collections Σ and Γ of np-functionals on von Neumann algebras A and B, respectively, an miu-bilinear map γ : A × B → T is a tensor product iff all of the following conditions hold. VII 1. The range of γ generates T . 2. For all σ ∈ Σ and τ ∈ Γ the product functional γ(σ, τ ) : T → C exists (see 108 II) and is positive. 3. The set { γ(σ, τ ) : σ ∈ Σ, τ ∈ Γ} is centre separating for T . Proof A tensor product γ obeys these conditions by definition and by IV, so we only need to show that a γ that obeys these conditions is a tensor product, and for this it suffices to show that γ can be extended to an nmiu-isomorphism γ⊗ : A ⊗ B → T . To extend γ to just an miu-map γ⊗ (to begin with) it suffices by 112 XI and 114 I to show that γ⊙ : A ⊙ B → T is bounded with respect to the tensor product norm on A ⊙ B and continuous with respect to the tensor product topology on A ⊙ B and the ultraweak topology on T . To see that γ⊙ is bounded, let t ∈ A ⊙B be given; we'll show that kγ⊙(t)k2 ≡ kγ⊙(t∗t)k 6 ktk2 where ktk is the tensor product norm of t. Since by 90 II the np-functionals on T of the form VIII γ(σ, τ )( γ⊙(s)∗ (· ) γ⊙(s) ) (4.3) where σ ∈ Σ, τ ∈ Γ and s ∈ A ⊙ B, are order separating, also with the restric- tion that 1 = γ(σ, τ )(γ⊙(s∗s)) ≡ (σ⊙ τ )(s∗s), and therefore determine the norm of t∗t as in 21 VII, it suffices to show that γ(σ, τ )(γ⊙(s)∗γ⊙(t∗t)γ⊙(s)) 6 ktk2 given such σ, τ , and s (with (σ⊙τ )(s∗s) = 1). But since γ(σ, τ )(γ⊙(s)∗γ⊙(t∗t)γ⊙(s)) = (σ⊙τ )(s∗( · )s) 6 ktk2 by the definition of the tensor product (σ ⊙ τ )(s∗t∗ts) = ktk2 norm (see 112 II), this is indeed the case. To see that γ⊙ : A ⊙B → T is ultraweakly continuous it suffices to show that ω ◦ γ⊙ is the operator norm limit of finite sums of basic functionals on A ⊙ B (see 112 II) given any np-functional ω : T → C. Since by 90 II such ω is the norm limit of finite sums of functionals on T of the form displayed in (4.3), and γ⊙ is bounded, we may assume without loss of generality that ω itself is as shown in (4.3). Since ω ◦ γ⊙ ≡ (σ ⊙ τ )(s∗(· )s) is then a basic functional γ⊙ is ultraweakly continuous. Having established boundedness and continuity of γ⊙ we obtain our nmiu- map γ⊗ : A ⊗ B → T with γ⊗(a ⊗ b) = γ(a, b) for all a ∈ A and b ∈ B. To show that γ is a tensor product, it suffices to show that γ⊗ is an nmiu- isomorphism, and for this, it suffices to show that γ⊗ is a bijection. In fact, we only need to show that γ⊗ is injective, because since the elements of T of the form γ(a, b) ≡ γ⊗(a ⊗ b) generate T (by assumption), and are in the range of γ⊗ (which is a von Neumann subalgebra of T by 48 VI), γ⊗ will be surjective. ..116.. 177 To show that γ⊗ is injective, it suffices to show that ⌈γ⊗⌉ ≡ ⌈⌈γ⊗⌉⌉ = 1 (see 69 IV). Since the product functionals on A ⊗B of the form σ⊗τ where σ ∈ Σ and τ ∈ Γ are centre separating (by IV), and ⌈⌈γ⊗⌉⌉ is central, it suffices to show that (σ ⊗ τ )(⌈⌈γ⊗⌉⌉⊥ ) = 0 given σ ∈ Σ and τ ∈ Γ. But this is easy -- (σ ⊗ τ )(⌈⌈γ⊗⌉⌉⊥ ) = γ(σ, τ )(γ⊗(⌈⌈γ⊗⌉⌉⊥ )) = 0. Whence γ is a tensor product. (cid:3) 117 Using the characterization from 116 VII it is pretty easy to see that the tensor product distributes over (infinite) direct sums (see III) after some unsurprising observations regarding direct sums (in II). II Exercise Let (Ai)i∈I be a collection of von Neumann algebras. 1. Show that given a generating subset Ai for each von Neumann algebra Ai Ai denotes Ai, where κi : Ai →Li∈I the set Si∈I κi(Ai) generates Li∈I the np-map given by (κi(a))i = a and (κi(a))j = 0 when j 6= i. 2. Show that given a centre separating collection Ωi of np-functionals on Ai for each i ∈ I the collection { ω ◦ πi : ω ∈ Ωi, i ∈ I } is centre separating for Li∈I Ai. III Proposition Given von Neumann algebras A and (Bi)i∈I the bilinear map γ : A ×Li A ⊗ Bi, (a, b) 7→ (ai ⊗ b)i Bi ∼=Li Bi −→Li is a tensor product. (Whence A ⊗Li IV Proof We use 116 VII to show that γ is a tensor product. Note that γ is clearly miu-bilinear, and that the elements of the form γ(a, κi(b)) = κ(a ⊗ b) from the range of γ where a ∈ A , i ∈ I, and b ∈ Bi generate Li A ⊗ Bi by II. Further, since given i ∈ I and np-functionals σ : A → C and τ : Bi → C the product functional γ(σ, τ ◦ πi) exists being simply (σ ⊗ τ ) ◦ πi : Li A ⊗ Bi → C, and such product functionals form a centre separating collection by II, we see that γ is indeed a tensor product. (cid:3) A ⊗ Bi.) 118 The tensor interacts with projections as expected. II Lemma Let A and B be von Neumann algebras. 1. We have ⌈a ⊗ b⌉ = ⌈a⌉ ⊗ ⌈b⌉ for all a ∈ A+ and b ∈ B+. 2. We have ⌈⌈a ⊗ b⌉⌉ = ⌈⌈a⌉⌉ ⊗ ⌈⌈b⌉⌉ for all a ∈ A and b ∈ B. III Proof Let a ∈ A+ and b ∈ B+ be given. Since the map (· )⊗ b : A → A ⊗ B is np, ⌈a ⊗ b⌉ 60 V===⌈⌈a⌉ ⊗ b⌉. Since similarly ⌈⌈a⌉ ⊗ b⌉ = ⌈⌈a⌉ ⊗ ⌈b⌉⌉ ≡ ⌈a⌉ ⊗ ⌈b⌉ using here that ⌈a⌉ ⊗ ⌈b⌉ is already a projection, we get ⌈a⌉ ⊗ ⌈b⌉ = ⌈a ⊗ b⌉. Let a ∈ A and b ∈ B be given in order to prove that ⌈⌈a ⊗ b⌉⌉ = ⌈⌈a⌉⌉ ⊗ ⌈⌈b⌉⌉. Since ⌈⌈a⌉⌉⊗ 1 commutes with all elements of A ⊗ B of the form a′⊗ b′, and thus IV with all elements of A ⊗B, we see that ⌈⌈a⌉⌉⊗1 is central. Since similarly 1⊗⌈⌈b⌉⌉ is central, we see that ⌈⌈a⌉⌉⊗ ⌈⌈b⌉⌉ = (⌈⌈a⌉⌉⊗ 1)⊗ (1 ⊗ ⌈⌈b⌉⌉) is central too. Since in addition ⌈⌈a⌉⌉⊗⌈⌈b⌉⌉ is a projection, and (⌈⌈a⌉⌉⊗⌈⌈b⌉⌉) (a⊗ b) = (⌈⌈a⌉⌉ a)⊗ (⌈⌈b⌉⌉ b) = a ⊗ b we see that ⌈⌈a ⊗ b⌉⌉ 6 ⌈⌈a⌉⌉ ⊗ ⌈⌈b⌉⌉ (by definition, see 68 III). So all that remains is to show that ⌈⌈a⌉⌉ ⊗ ⌈⌈b⌉⌉ 6 ⌈⌈a ⊗ b⌉⌉. Recall that ⌈⌈a⌉⌉ = Sa∈A ⌈a∗a∗aa⌉ by 68 I. Using this, a similar expression for ⌈⌈b⌉⌉, and 60 IX, we see that ⌈⌈a⌉⌉ ⊗ ⌈⌈b⌉⌉ = Sa∈A Sb∈B ⌈(a∗a∗aa) ⊗ (b∗b∗bb)⌉, and so it suffices to show that ⌈(a∗a∗aa) ⊗ (b∗b∗bb)⌉ 6 ⌈⌈a ⊗ b⌉⌉ given a ∈ A and b ∈ B. This is indeed the case since ⌈(a∗a∗aa) ⊗ (b∗b∗bb)⌉ = ⌈(a ⊗ b)∗ (a ⊗ b)∗(a ⊗ b) (a ⊗ b)⌉ 6 ⌈⌈a ⊗ b⌉⌉ (by 68 I, again.) (cid:3) Exercise Let f : A → B and g : C → D be np-maps between von Neumann algebras. We're going to prove that ⌈f ⊗ g⌉ = ⌈f⌉ ⊗ ⌈g⌉. 1. Show that (f⊗g)(⌈f⌉⊗⌈g⌉) = 1⊗1, and conclude that ⌈f ⊗ g⌉ 6 ⌈f⌉⊗⌈g⌉. 2. Assume for the moment that A and C are von Neumann algebras of bounded operators on Hilbert spaces H and K , respectively, and that f and g are vector functionals, that is, B = D = C, and f = hx, (· )xi for some x ∈ H , and g = hy, (· )yi for some y ∈ K . Show that ⌈f⌉ =Sa∈A (cid:3) ⌈ a∗ xihx a⌉ using 88 IV and 88 VI. 3. With the same assumptions as in the previous point, suppose, further- more, without loss of generality that A ⊗ B is given as the von Neumann subalgebra of B(H ⊗ K ) generated by the operators A⊗ B where A ∈ A and B ∈ B (cf. 111 VII). Show that f ⊗ g = hx ⊗ y, (· )x ⊗ yi. Given a ∈ A (cid:3) and b ∈ B(cid:3) show that a ⊗ b ∈ (A ⊗ B)(cid:3), and thus ⌈a∗ xihx a⌉ ⊗ ⌈b∗ yihy b⌉ 6 ⌈f ⊗ g⌉ . Deduce from this that ⌈f⌉ ⊗ ⌈g⌉ 6 ⌈f ⊗ g⌉, so ⌈f⌉ ⊗ ⌈g⌉ = ⌈f ⊗ g⌉. 4. Let f and g be arbitrary again, and assume now that f and g are func- when σ ranges over the np-functionals σ on B. tionals, that is, B = D = C. Show that ⌈f ⊗ g⌉ = ⌈f⌉ ⊗ ⌈g⌉. 5. Let f and g be arbitrary again, and recall from 66 IV that 1 = Sσ ⌈σ⌉ Show that 1⊗1 =Sσ,τ ⌈σ ⊗ τ⌉ where σ and τ range over the np-functionals on B and D, respectively. Show using 101 IV and 101 VIII that ⌈f ⊗ g⌉ ≡ (f ⊗ g)⋄(1⊗ 1) = ⌈f⌉⊗⌈g⌉. 6. Show that (f ⊗ g)⋄(s⊗ t) = f⋄(s)⊗ g⋄(t) for projections s ∈ B and t ∈ D. ..116 -- 118 179 4.2.6 Monoidal Structure 119 Up to this point we have only written about the tensor product A ⊗ B of two von Neumann algebras (to save ink), but all of it, as you will no doubt have observed already, can be easily adapted to deal with a tensor product ⊗ : A1 × . . . × An → A1 ⊗ ··· ⊗ An of a tuple A1, . . . , An of von Neumann algebras, which will then, of course, be a multilinear map instead of a bilinear map, etc.. II III What is less obvious is that there should be any relation between (A ⊗ B)⊗ C , and A ⊗ (B ⊗ C ) and A ⊗ B ⊗ C ; but there is. Proposition Given von Neumann algebras A , B and C , the trilinear map γ : (a, b, c) 7→ (a ⊗ b) ⊗ c, A × B × C → (A ⊗ B) ⊗ C is a tensor product. Proof We need to verify the three conditions from 108 II (adapted to trilinear maps). The first condition, that the elements of the form (a ⊗ b) ⊗ c generate (A ⊗B)⊗C follows by 116 IV since the elements of the form a⊗b generate A ⊗B (and C generates C ). The second condition is met by defining γ(σ, τ, υ) := (σ ⊗ τ )⊗ υ for all np-functionals σ : A → C, τ : B → C and υ : C → C. Finally, these product functionals γ(σ, τ, υ) are centre separating by 116 IV because the functionals on A ⊗ B of the form σ ⊗ τ are centre separating (and so is the set of all np-functionals on C ), which was the third condition. (cid:3) IV Corollary There is a unique nmiu-isomorphism αA ,B,C : A ⊗ (B ⊗ C ) −→ (A ⊗ B) ⊗ C , called an associator, with αA ,B,C ( a ⊗ (b ⊗ c) ) = (a ⊗ b) ⊗ c for all a ∈ A , b ∈ B, c ∈ C , for any von Neumann algebras A , B, C . If the point above means that ⊗ is associative, then the following two points mean that ⊗ has C as its unit, and ⊗ is commutative, respectively. IVa IVb Exercise Show that given a von Neumann algebra A the bilinear maps (z, a) 7→ za : C× A → A and (a, z) 7→ za : A × C → A are tensor products, and deduce from this that there are unique nmiu-isomorphisms λA : C ⊗ A −→ A , and, A : A ⊗ C −→ A , called a left and right unitor, respectively, with λA (z ⊗ a) = za = A (a⊗ z) for all a ∈ A and z ∈ C. IVc Exercise Show that given von Neumann algebras A and B the bilinear map (a, b) 7→ b ⊗ a : A ⊗ B −→ B ⊗ A is a tensor product, and deduce from this that there is a unique nmiu-isomorphism called a braiding, with γA ,B(a ⊗ b) = b ⊗ a for all a ∈ A and b ∈ B. γA ,B : A ⊗ B −→ B ⊗ A , Theorem Endowed with the tensor product, W∗miu, W∗cp, W∗cpu, and W∗cpsu are symmetric monoidal categories [51] with C as unit. Proof The first order of business is showing that the associators αA ,B,C (from IV) form a natural transformation in W∗cp (and thus in W∗miu, W∗cpu, and W∗cpsu too, as the αA ,B,C 's are nmiu), that is, that the following diagram commutes V Va A ⊗ (B ⊗ C ) αA ,B,C f⊗(g⊗h) A ′ ⊗ (B′ ⊗ C ′) αA ′,B′ ,C ′ (A ⊗ B) ⊗ C (f⊗g)⊗h / (A ′ ⊗ B′) ⊗ C ′ (4.4) for all ncp-maps f : A → A ′, g : B → B′, and h : C → C ′. Note that by both routes through this diagram a ⊗ (b ⊗ c) gets mapped to (f (a) ⊗ g(b)) ⊗ h(c) for all a ∈ A , b ∈ B, and c ∈ C . Since the linear span of such a ⊗ (b ⊗ c)'s is ultraweakly dense in A ⊗ (B⊗ C ) (by 116 IV,) this entails that (4.4) commutes. By a similar but simpler argument one sees that the braidings (γA ,B) and unitors (λA and A ) give natural transformations. It remains to be shown that the appropriate coherence relations hold. Given von Neumann algebras A , B, C , and D, the pentagon Vb (A ⊗ B) ⊗ (C ⊗ D) αA ,B,C⊗D 5❦❦❦❦❦❦❦❦❦❦❦❦❦❦❦❦❦❦❦❦❦ A ⊗ (B ⊗ (C ⊗ D)) ✼✼✼✼✼✼✼✼ idA ⊗αB,C ,D ✼✼✼✼✼✼✼✼ A ⊗ ((B ⊗ C ) ⊗ D)αA ,B⊗C ,D ((A ⊗ B) ⊗ C ) ⊗ D αA ⊗B,C ,D )❙❙❙❙❙❙❙❙❙❙❙❙❙❙❙❙❙❙❙❙❙ C✞✞✞✞✞✞✞✞ αA ,B,C⊗idD ✞✞✞✞✞✞✞✞ / (A ⊗ (B ⊗ C )) ⊗ D commutes, since by both routes the elements in A ⊗ (B ⊗ (C ⊗ D)) of the form a⊗(b⊗(c⊗d)) (whose linear span is ultraweakly dense) get sent to ((a⊗b)⊗c)⊗d. By similar arguments the diagrams A ⊗ (C ⊗ C ) '◆◆◆◆◆◆◆◆◆◆◆ idA ⊗λC αA ,C,C A ⊗ C (A ⊗ C) ⊗ C 7♣♣♣♣♣♣♣♣♣♣♣ A ⊗idC C ⊗ C C λC C 119.. 181 / /     / ) 5  / C / / '     7 commute, as does the diagram A ⊗ (B ⊗ C ) idA ⊗γB,C αA ,B,C (A ⊗ B) ⊗ C γA ⊗B,C / C ⊗ (A ⊗ B) , αC ,A ,B A ⊗ (C ⊗ B) αA ,C ,B / (A ⊗ C ) ⊗ B γA ,C⊗idB / (C ⊗ A ) ⊗ B and do the diagrams A ⊗ B γA ,B %❑❑❑❑❑❑❑❑❑❑ idA ⊗B B ⊗ A γB,A A ⊗ B B ⊗ C C ⊗ B . γB,C %❏❏❏❏❏❏❏❏❏❏ B λB B Thus W∗miu, W∗cp, W∗cpu, and W∗cpsu are symmetric monoidal categories. (cid:3) 4.3 Quantum Lambda Calculus 120 In this section we provide the parts needed to build a model of the quantum lambda calculus using von Neumann algebras. We will not venture to describe the quantum lambda calculus in all its details here, nor will we describe how to build the model from these parts (as we did in [11]); we'll just touch upon the two key ingredients: the interpretation of "!" and "⊸" -- with them the expert can easily produce the model. Let us, nevertheless, try to give some impression to those who are not familiar with the quantum lambda calculus. The quantum lambda calculus is a type theory proposed by Selinger and Valiron in [71, 72] to describe programs for quantum computers especially designed to include not only function types (⊸) and classical data types (such as bit), but also quantum data types (such as qubit), so that there can be a term such as new : bit ⊸ qubit that represents the program that initialises a qubit in the given state. There are of course also terms such as 0 : bit and 1 : bit, so that new 0 : qubit represents a qubit in state 0i. The addition of quantum data to a type theory is a very delicate matter for if one were to allow for example in this system a variable to be used twice (a thing usually beyond dispute) it would not take much more to construct a program that duplicates the contents of a qubit, which is nonphysical. Still, classical data such as a bit can be duplicated freely, so to accommodate this the type !bit is used. More precisely, the type !A represents that part of the type of A that is duplicable, so that !bit is the proper type for a bit, and !qubit is empty. For example, the term that represents the measurement of a / /   /   / / / / %   / / %   qubit is meas : qubit ⊸ !bit, where the ! indicates that the bit resulting from the measurement may be duplicated freely. The model we alluded to assigns to each type A a von Neumann algebra JAK, e.g. JqubitK = M2 and JbitK = C2. A (closed) term t : A is interpreted as an npsu-functional Jt : AK : JAK → C, so for example J0 : bitK : (x, y) 7→ x : C2 → C. When t : A has free variables x1 : B1, . . . , xN : BN the interpretation becomes an ncpsu-map JtK : JAK → JB1K ⊗ ··· ⊗ JBN K, so for example, 0 y(cid:1) : C2 → M2. J x : qubit ⊢ measx K : (x, y) 7→(cid:0) x 0 In short, there are no surprises here. As said, the difficulty lies in the definition of J!AK and JA ⊸ BK, for which we will provide the following three ingredients. • The observation (by Kornell, [48]) that the category (W∗miu)op is monoidal closed, that is, that for every von Neumann algebra B, the functor B ⊗ (· ) : W∗miu → W∗miu has a left adjoint (· )∗B. • The following two adjunctions. ℓ∞ Set ⊥ nsp:=W∗ miu(−,C) (W∗miu)op ⊆ ⊥ F (W∗cpsu)op The interpretation of J!AK and JA ⊸ BK will then be J!AK = ℓ∞(nsp(JAK)) and JA ⊸ BK = F (JBK)∗JAK. By the universal properties of F and (· )∗B, an ncpsu-map f : A → C ⊗ B cor- responds to unique nmiu-map Λ(f ) : F (A )∗B −→ C . This is used to interpret the "λ": JλxB.tK = Λ(JtK) : F (JAK)JBK −→ JB1K ⊗ ··· ⊗ JBN K for any term t : A with free variables x1 : B1, . . . , xN : BN , x : B, so JtK : JAK −→ JB1K ⊗ ··· ⊗ JBN K ⊗ JBK. Note that J!AK will always be a 'discrete' commutative von Neumann algebra no matter how complicated JAK may be, so that although this does the job perhaps a more interesting interpretation of ! may be chosen as well. This is not the case: in the next section we'll show that any von Neumann algebra that carries a ⊗-monoid structure (such as J!AK) is commutative and discrete, and that ℓ∞(nsp(A )) is moreover the free ⊗-monoid on A . In [73] the quantum lambda calculus is extended with recursion via the "let rec" operator; we don't know whether it's possible to interpret let rec in our model. II In this section, we'll need the following result from the literature on von Neu- mann algebras. 121 ..119 -- 121.. 183 , , , , j j l l II Proposition Given Hilbert spaces H and K , and von Neumann subalgebras A1 and A2 of B(H ) and von Neumann subalgebras B1 and B2 of B(K ), we have (A1 ⊗ B1) ∩ (A2 ⊗ B2) = (A1 ∩ A2) ⊗ (B1 ∩ B2). Here A1 ⊗ B1 denotes not just any tensor product of A1 and B1, but instead the "concrete" tensor product of A1 and B1: the least von Neumann subalgebra of B(H ⊗ K ) that contains all operators of the form A ⊗ B where A ∈ A1 and B ∈ B1. Proof See Corollary IV.5.10 of [75]. (cid:3) III 4.3.1 First Adjunction 122 Definition We write nsp := W∗miu(· , C) for the functor (W∗miu)op → Set which maps a von Neumann algebra A to its set of nmiu-functionals, nsp(A ), and sends an nmiu-map f : A → B to the map nsp(f ) : nsp(B) → nsp(A ) given by nsp(f )(ϕ) = ϕ ◦ f for ϕ ∈ nsp(B). Proposition Given a set X the map II η : X → nsp(ℓ∞(X)) given by η(x)(h) = h(x) is universal in the sense that for every map f : X → nsp(A ), where A is a von Neumann algebra, there is a unique nmiu-map g : A → ℓ∞(X) such that η X nsp(ℓ∞(X)) $❏❏❏❏❏❏❏❏❏❏ f nsp(g) nsp(A ) ℓ∞(X) g A III commutes. Moreover, and as a result, the assignment X 7→ ℓ∞(X) extends to a functor ℓ∞ : Set → (W∗miu)op that is left adjoint to nsp, and is given by ℓ∞(f )(h) = h ◦ f for any map f : X → Y and h ∈ ℓ∞(Y ). Proof Note that if we identify ℓ∞(X) with the X-fold product of C, we see that η(x) : ℓ∞(X) ≡ Lx∈X C → C is simply the x-th projection, and thus an nmiu-map (see 47 IV). Hence we do indeed get a map η : X → nsp(ℓ∞(X)). To see that η has the desired universal property, let f : X → nsp(A ) be given, and define g : A → ℓ∞(X) by g(a)(x) = f (x)(a). One can now either prove directly that g is nmiu, or reduce this in a slightly roundabout way from the known fact that ℓ∞(X) is the X-fold product of C with the η(x) as projec- tions; indeed g is simply the unique nmiu-map with η(x)◦g = f (x) for all x ∈ X, that is, g = hf (x)ix∈X . In any case, we see that nsp(g)(η(x)) ≡ η(x) ◦ g = f (x) for all x ∈ X, and so nsp(g)◦ η = f . Concerning uniqueness of such g, note that / / $   O O Ai of von Neumann algebras given an nmiu-map g′ : A → ℓ∞(X) with nsp(g′) ◦ η = f we have η(x) ◦ g′ = nsp(g′)(η(x)) = f (x) for all x ∈ X, and so g′ = hf (x)ix∈X = g. Hence η is a universal arrow from X to nsp. That as a result the assignment X 7→ ℓ∞(X) extends to a functor Set → (W∗miu)op by sending f : X → Y to the unique nmiu-map ℓ∞(f ) : ℓ∞(Y ) → ℓ∞(X) with nsp(ℓ∞(f )) ◦ ηX = ηY ◦ f is a known and easily checked fact (where ηX := η and ηY : Y → nsp(ℓ∞(Y )) is what you'd expect). Finally, applying x ∈ X and h ∈ ℓ∞(Y ) we get ℓ∞(f )(h)(x) = ηX (x)(ℓ∞(f )(h))) = nsp(ℓ∞(x))(ηX (x))(h) = ηY (f (x))(h) = h(f (x)). (cid:3) Lemma A nmiu-functional ϕ on a direct sum Li is of the form ϕ ≡ ϕ′ ◦ πi for some i and nmiu-functional ϕ′ on Ai. Proof Let ej denote the element of Li Ai given by ej(j) = 1 and ej(i) = 0 for all i 6= j. Note that given i and j with i 6= j we have eiej = 0 and so 0 = ϕ(eiej) = ϕ(ei)ϕ(ej); from this we see that there is at most one i with ϕ(ei) 6= 0. Since for this i we have e⊥i = Pj6=i ej and so ϕ(e⊥i ) = Pj6=i ϕ(ej) = 0, we see that ϕ(a) = ϕ(eia) for all a ∈ Li Aj be the nmisu- map given by κi(a)(i) = a and κi(a)(j) = 0 for j 6= i we have ϕ = ϕ ◦ κi ◦ πi. Hence taking ϕ′ := ϕ ◦ κi does the job. (cid:3) Exercise Deduce from IV that the functor nsp : (W∗miu)op → Set preserves coproducts, and that the map η : X → nsp(ℓ∞(X)) from II is a bijection. Show that ℓ∞ : Set → (W∗miu)op is full and faithful. Whence Set is (isomor- phic to) a coreflective subcategory of (W∗miu)op via ℓ∞ : Set → (W∗miu)op. Exercise We're going to prove that ℓ∞(X × Y ) ∼= ℓ∞(X) ⊗ ℓ∞(Y ). Ai. Letting κi : Ai → Lj 1. Given an element x of a set X let x denote the element of ℓ∞(X) that IV V VI 123 equals 1 on x and is zero elsewhere. Show that { x : x ∈ X } generates ℓ∞(X). 2. Show that the projections πx : ℓ∞(X) ≡ Ly∈X separating collection of nmiu-functionals on ℓ∞(X). C → C form an order 3. Using this, and 116 VII, prove that given sets X and Y the map ⊗ : ℓ∞(X) × ℓ∞(Y ) → ℓ∞(X × Y ) given by (f ⊗ g)(x, y) = f (x)g(y) is a tensor product. Conclude that ℓ∞(X × Y ) ∼= ℓ∞(X) ⊗ ℓ∞(Y ). (In fact, it follows that ℓ∞ is strong monoidal.) Exercise Let A and B be von Neumann algebras. We're going to show that nsp(A ⊗ B) ∼= nsp(A ) × nsp(B). II ..121 -- 123.. 185 1. Given an nmiu-functional ϕ : A ⊗ B → C show that σ := ϕ((· ) ⊗ 1) and τ := ϕ(1 ⊗ (· )) are nmiu-functionals on A and B, respectively; and show that ϕ = σ ⊗ τ (by proving that ϕ(a ⊗ b) = σ(a)τ (b).) 2. Show that σ, τ 7→ σ⊗ τ gives a bijection nsp(A )× nsp(B) → nsp(A ⊗ B). (This makes nsp strong monoidal.) 4.3.2 Second Adjunction 124 Lemma If a von Neumann algebra A is generated by S ⊆ A , then #A 6 22#C+#S , II where #S denotes the cardinality of S, and so on. Proof Note that the ∗-subalgebra S′ of A generated by S is ultraweakly dense in A . Since every element of S′ can be formed from the infinite set S ∪ C using the finitary operations of addition, multiplication, and involution, #S′ 6 #C+#S. Since every element of A is the ultraweak limit of a filter (see [87, §12]) on S′ of which there no more than 22#S′ (cid:3) III Theorem The inclusion W∗miu → W∗cpsu has a left adjoint F : W∗cpsu → W∗miu. IV Proof Note that since the category W∗miu has all products (47 IV), and equalis- ers (47 V), W∗miu has all limits (by Theorem V2.1 and Exercise V4.2 of [51]). Moreover, the inclusion U : W∗miu → W∗cpsu preserves these limits (see 47 IV and 47 V). So by Freyd's adjoint functor theorem (Theorem V6.1 of [51]) it suffices to check the solution set condition, that is, that , we conclude #A 6 22#C+#S . for every von Neumann algebra A there be a set I, and for each i ∈ I an ncpsu-map fi : A → Ai into a von Neumann algebra Ai such that every ncpsu-map f : A → B into some von Neumann algebra B is of the form f ≡ h ◦ fi for some i ∈ I and nmiu-map h : Ai → B. , define To this end, given a von Neumann algebra A , let κ := 22#C+#A I = { (C , γ) : C is a von Neumann algebra on a subset of κ, and γ : A → C is an ncpsu-map }, and set fi := γ for every i ≡ (C , γ) ∈ I. Let f : A → B be an ncpsu-map into a von Neumann algebra B. The von Neumann algebra B′ generated by f (A ) has cardinality below κ by I, and so by relabelling the elements of B′ we may find a von Neumann algebra C on a subset of κ isomorphic to B′ via some nmiu-isomorphism Φ : B′ → C . Then the map γ : A → C given by γ(a) = Φ(f (a)) for all a ∈ A is ncpsu, so that i := (C , γ) ∈ I, and, moreover, the assignment c 7→ Φ−1(c) gives an nmiu-map h : C → B with h ◦ fi ≡ h ◦ γ = f . Hence U : W∗miu → W∗cpsu obeys the solution set condition, and therefore has a left adjoint. (cid:3) Remark A bit more can be said about the adjunction between the inclu- sion U : W∗miu → W∗cpsu and F : since W∗miu has the same objects as W∗cpsu, the category (W∗cpsu)op is, for very general reasons, equivalent to the Kleisli cat- egory of the (by the adjunction induced) monad F U on (W∗miu)op in a certain natural way (see e.g. Theorem 9 of [80]). V 4.3.3 Free Exponential We'll prove Kornell's result (from [48]) that the functor B⊗ (· ) : W∗miu → W∗miu has a left adjoint (· )∗B for every von Neumann algebra B. Kornell original proof is rather complex, and so is ours, unfortunately, but we've managed to peel off one layer of complexity from the original proof by way of Freyd's Adjoint Functor Theorem, reducing the problem to the facts that B ⊗ (· ) : W∗miu → W∗miu preserves products, equalisers, and satisfies the solution set condition. 125 If A = {0}, then the result is obvious, so let us assume that A 6= {0}. Let Ω be the set of np-functionals on A . Recall that by the GNS-construction Lemma A von Neumann algebra A can be faithfully represented on a Hilbert space which contains no more than 2#A vectors. Proof Then A is infinite, and so ℵ0 · #A = #A . (see 48 VIII) A can be faithfully represented on the Hilbert space HΩ ≡Lω∈Ω Since every element of Hω is the limit of a sequence of elements from A , we have #Hω 6 ℵ#A 0 6 (2ℵ0)#A = 2#A , because ℵ0 · #A = #A . Since every normal state is a map ω : A → C, we have #Ω 6 #C#A = (2ℵ0 )#A = 2#A , because ℵ0 · #A = #A . Hence #H =Pω∈Ω #Hω 6 2#A · 2#A = 2#A . (cid:3) Lemma (Kornell) Every nmiu-map h : D → A ⊗ C , where A , C and D are von / A ⊗ C , where A is a Neumann algebras, factors as D von Neumann algebra, and ι and h are nmiu-maps, such that for all nmiu-maps f, g : A → B into some von Neumann algebra B with (f ⊗ id)◦ h = (g ⊗ id)◦ h we have f ◦ ι = g ◦ ι. Moreover, A can be generated by less than #D · 2#C elements. h / A ⊗ C ι⊗id / Proof Assume (without loss of generality) that C is a von Neumann algebra of operators on a Hilbert space H with no more than 2#C vectors, see II. For every vector ξ ∈ H let rξ : A ⊗ C → A be the unique np-map given by rξ(a⊗ c) = hξ, cξi a for all a ∈ A and c ∈ C (see 112 XI and 114 I), and let A be the least von Neumann subalgebra of A that contains S :=Sξ∈H rξ(h(D)), ..123 -- 125.. 187 Hω. II III IV V / and let ι : A → A be the inclusion (so ι is nmiu). Note that S (which gener- ates A ) has no more than #D · #H 6 #D · 2#C elements. Let f, g : A → B be nmiu-maps into a von Neumann algebra B such that (f ⊗ id) ◦ h = (g ⊗ id) ◦ h. We must show that f ◦ ι = g ◦ ι. By definition of A (and the fact that f and g are nmiu), it suffices to show that f ◦ rξ ◦ h = g ◦ rξ ◦ h for all ξ ∈ H . Note that given such ξ, we have f ◦ rξ = r′ξ ◦ (f ⊗ id), where r′ξ : B ⊗ C → B is the np-map given by r′ξ(b ⊗ c) = hξ, cξi b. Since similarly, g◦rξ = r′ξ◦(g⊗id), we get f ◦rξ◦h = r′ξ◦(f ⊗id)◦h = r′ξ◦(g⊗id)◦h = g ◦ rξ ◦ h. It remains only to be shown that h(D) ⊆ A ⊗ C , because we may then simply let h be the restriction of h to It is enough to prove that A ⊗ C = ( A ⊗ B(H )) ∩ (A ⊗ C ) (see 121 II) h(D) ⊆ A ⊗ B(H ), because and we already know that h(D) ⊆ A ⊗ C . Let (ek)k be orthonormal basis of H . Since 1 =Pk ekihek in B(H ), we have, for all d ∈ D, h(d) = (cid:0)Pk 1 ⊗ ekihek(cid:1) h(d) (cid:0)Pℓ 1 ⊗ eℓiheℓ(cid:1) = PkPℓ ( 1 ⊗ ekihek ) h(d) ( 1 ⊗ eℓiheℓ ). ( 1 ⊗ ξihξ ) h(d) ( 1 ⊗ ζihζ ) ∈ A ⊗ B(H ). We are done if we can prove that, for all ξ, ζ ∈ H , A ⊗ C . (4.5) By an easy computation, we see that, for all e ∈ A ⊗ C of the form e ≡ a ⊗ c, 3 1 4 ik rik ξ+ζ(e) ⊗ ξihζ . ( 1 ⊗ ξihξ ) e ( 1 ⊗ ζihζ ) = Xk=0 It follows that the equation above holds for all e ∈ A ⊗ C . Choosing e = h(d) we see that (4.5) holds, because rikξ+ζ(h(d)) ∈ A . (cid:3) VI Proposition Let e : E → A be an equaliser of nmiu-maps f, g : A → B between von Neumann algebras. Then e ⊗ id : E ⊗ C → A ⊗ C is an equaliser of f ⊗ id and g ⊗ id for every von Neumann algebra C . VII Proof Let h : D → A ⊗C be an nmiu-map with (f⊗id)◦h = (g⊗id)◦h. We must show that there is a unique nmiu-map k : D → E ⊗ C such that h = (e⊗ id)◦ k. Note that since the equaliser map e is injective, e⊗id : E ⊗C → A ⊗C is injective (by 115 V) and thus uniqueness of k is clear. Concerning existence, by IV, h / A ⊗ C where h and ι are nmiu-maps, and factors as D moreover, we have f◦ι = g◦ι. Since e is an equaliser of f and g, there is a unique nmiu-map ι : A → E with e ◦ ι = ι. Now, define k := (ι ⊗ id) ◦ h : D → E ⊗ C . Then (e ⊗ id) ◦ k = ((e ◦ ι) ⊗ id) ◦ h = (ι ⊗ id) ◦ h = h. (cid:3) VIIa So given a von Neumann algebra A the functor (· ) ⊗ A : W∗miu → W∗miu pre- serves all equalisers and products, thus all limits, and in particular, all pullbacks. This has the following pleasant consequence used later on. / A ⊗ C ι⊗id / h / Exercise Given a nmiu-map : B → C between von Neumann algebras B and C , and a von Neumann subalgebra S of C , show that ( ⊗ A )−1(S ⊗ A ) = −1(S ) ⊗ A VIIb for every von Neumann algebra A , where for the sake of simplicity we take −1(S ) ⊗ A to be the von Neumann subalgebra of B ⊗ A generated by { b ⊗ a : b ∈ −1(S ), a ∈ A }. (Hint: express −1(S ) as pullback in W∗miu of ◦ π1, e◦ π2 : B ⊕ S → C , where e : S → C is the inclusion.) Theorem (Kornell) The functor (· ) ⊗ A : W∗miu → W∗miu has a left adjoint (· )∗A for every von Neumann algebra A . Proof The category W∗miu is (small-)complete, and (−) ⊗ A : W∗miu → W∗miu preserves (small-)products and equalisers. Thus, by Freyd's (General) Adjoint Functor Theorem [51, Thm. V.6.2], it suffices to check the following Solution Set Condition (where we've used that W∗miu is locally small). VIII IX • For each B ∈ W∗miu, there is a small subset S of objects in W∗miu such that every arrow h : B → C ⊗A can be written as a composite h = (t⊗idA )◦f for some D ∈ S, f : B → D ⊗ A , and t : D → C . Let B be an arbitrary von Neumann algebra. We claim that the following set S satisfies the required condition: S = { D : D is a von Neumann algebra on a subset of κ}, where κ = 22#C·#B·2#A . Note that κ being an ordinal number is just the set of all ordinal numbers α < κ. To prove the claim, suppose that h : B → C ⊗ A is given. By IV, h factors as B / C ⊗ A ι⊗id / C ⊗ A , where C is a von Neumann algebra generated by no more than #B · 2#A elements. It follows that C has no more than κ elements (by 124 I). Thus we may assume without loss of generality that C is a subset of κ, that is, C ∈ S.(cid:3) Remark It should be noted that analogues of the first and second adjunctions can be found in the setting of C∗-algebras, which raises the question as to whether a variation on the free exponential exist for C∗-algebras, that is, is there a tensor ⊗ on C∗miu such that (−) ⊗ A : C∗miu → C∗miu has a left adjoint? Such a tensor does not exist if we require that on commutative C∗-algebras it is given by the product of the spectra (as is the case for the projective and injective tensors of C∗-algebras) in the sense that there is a natural isomor- / C(X × Y ) between the obvious functors of phism ΦX,Y : C(X) ⊗ C(Y ) ∼= / ..125.. 189 X / / type CH × CH → (C∗miu)op. Indeed, if (−) ⊗ A : C∗miu → C∗miu had a left adjoint and so would preserve all limits for all C∗-algebras A , then the func- tor (−)×X : CH → CH would preserve all colimits for every compact Hausdorff space X, which it does not, because if it did the square βN × βN of the Stone -- Cech compactification βN of the natural numbers (being the N-fold coproduct of the one-point space) would be homeomorphic to the Stone -- Cech compactifi- cation β(N × N) of N × N, which it is not (by Theorem 1 of [23]). the same way that W∗cpsu does. Whence C∗cpsu does not form a model of the quantum lambda calculus in 4.3.4 Hereditarily Atomic Von Neumann Algebras 125a We'll argue that it's possible to modify our model of the quantum lambda calculus from [11] to include only hereditarily atomic (84b II) von Neumann algebras (as suggested by Kornell on page 5 of [49].) To this end we must bring up that the types of the quantum lambda calculus are generated as follows: there's a type qubit, and a type ⊤; and from types A and B, we can form∗ A ⊕ B, A ⊗ B, !A, and A ⊸ B. Note that the interpretations, JqubitK = M2 and J⊤K = C, of the ground types are hereditarily atomic, and that the interpretation of the sum, JA ⊕ BK = JAK⊕JBK, and the tensor, JA⊗BK = JAK⊗JBK, are hereditarily atomic when JAK and JBK are hereditarily atomic. The interpretation J!AK = ℓ∞(nsp(JAK)) is hereditarily atomic regardless of whether JAK is hereditarily atomic, or not. So whether all von Neumann algebras in our model are hereditarily atomic hinges only on the interpretation of ⊸. As it turns out, the interpretation JA ⊸ BK = F (JBK)∗JAK we chose is not always hereditarily atomic when JAK and JBK are hereditarily atomic: we claim (without proof) that J⊤⊕3 ⊸ ⊤K ≡ C∗C3 has B(ℓ2) as factor, and that J⊤ ⊸ bitK ≡ F (C2) has L∞[0, 1] as summand. II The solution is obvious: show that the functor (· ) ⊗ A : haW∗miu → haW∗miu has a left adjoint (· )∗haA for every hereditarily atomic von Neumann algebra A , and show that the inclusion haW∗miu → haW∗cpsu has a left adjoint Fha. One may then define J· Kha exactly the same as J· K except for JA ⊸ BKha := Fha( JBKha )∗haJAKha . The benefit of using the hereditarily atomic model is that Fha and A ∗haB admit a significantly more concrete description see 125c III and 125e VII A potential drawback might be that the purely quantum mechanical is restricted to finite dimensions, so to speak. ∗The type bit discussed in 120 I is missing from this list, since it can defined by bit := ⊤⊕⊤. We establish the existence of Fha indirectly at first. Proposition The inclusion haW∗miu → haW∗cpsu has a left adjoint Fha(A ) : haW∗cpsu −→ haW∗miu. Proof Given our definition of hereditary atomicity, 84b II, it's pretty clear that the subcategory haW∗miu of W∗miu is closed under products, and that these prod- ucts are preserved by the inclusion functor haW∗miu → haW∗cpsu. Using 84b V one sees the same holds for equalisers. Whence the proof is completed by an application of Freyd's adjoint functor theorem, exactly as in 124 IV, but with as solution set for a hereditarily atomic von Neumann algebra A , the ncpsu-maps γ : A → C for which C is a hereditarily atomic von Neumann algebra on a subset of the cardinal κ ≡ 22#C+#A (cid:3) To give a concrete description of the functor Fha we need some notation first. Let A be a hereditarily atomic von Neumann algebra. We'll describe Fha(A ) in terms of ncpsu-maps f : A → MNf with W ∗(f (A )) = MNf . Let us say that are miu-equivalent when two such maps f1 : A → MNf1 with ϕ◦f1 = f2, (which implies there is an nmiu-isomorphism ϕ : MNf1 → MNf2 that Nf1 = Nf2.) Choose a set RA of representatives for this miu-equivalence. and f2 : A → MNf2 . 125b II III 125c II Theorem Given a hereditarily atomic von Neumann algebra A , the unique nmiu-map Φ that causes the diagram III A ηA (❘❘❘❘❘❘❘❘❘❘❘❘❘❘❘❘ hrir∈RA Fha(A ) Φ Lr∈RA MNr to commute is an nmiu-isomorphism. Here η denotes the unit of the adjunction between Fha and the inclusion haW∗miu → haW∗cpsu. Proof Since Fha(A ) is hereditarily atomic, it's nmiu-isomorphic to a direct sum of the formLi∈I MNi. We may as well assume that Fha(A ) ≡Li∈I MNi. We claim, writing ηA ≡ hsiii∈I : A → Li∈I MNi, that the si form a set of Indeed, if given r ∈ RA we denote by ir the unique element of I for which sir is miu-equivalent to r, and let ϕr : MNir → MNr be a corresponding nmiu-isomorphism with r = ϕr ◦ sir , then one easily sees using its defining property that Φ is the composition of The theorem follows easily from this claim. representatives for miu-equivalence as well. IV Fha(A ) ≡ Li∈I MNi hπir ir∈RA / /Lr∈RA MNri ..125 -- 125c.. Lr∈RA ϕr / Lr∈RA MNr. 191 / / (   / V VI VII Since r 7→ ir gives a bijection RA → I, the first map above is a nmiu- isomorphism. Since the second map is clearly a nmiu-isomorphism too, Φ is a nmiu-isomorphism. Let us begin by proving that W ∗(si(A )) = MNi for every i ∈ I. To this end, we'll first show that W ∗(ηA (A )) = Fha(A ). Let us denote by f : A → W ∗(ηA (A )) the restriction of ηA . By the universal property of ηA , there's a unique nmiu-map : Fha(A ) → W ∗(ηA (A )) with f = ◦ ηA . Letting e : W ∗(ηA (A )) → Fha(A ) be the inclusion, we have e ◦ ◦ ηA = e ◦ f = ηA . Since the identity idFha(A ) : Fha(A ) → Fha(A ) is the unique nmiu-map τ : Fha(A ) → Fha(A ) with τ ◦ ηA = ηA , we get idFha(A ) = e◦ . Since idFha(A ) is surjective, so is e, and thus W ∗(ηA (A )) = Fha(A ). Note that W ∗(ηA (A )) ⊆ Li∈I W ∗(si(A )) because Li∈I W ∗(si(A )) is a von Neumann subalgebra of Fha(A ) with ηA (A ) ⊆ Li∈I W ∗(si(A )). Since W ∗(ηA (A )) = Fha(A ), we get Li∈I W ∗(si(A )) = Fha(A ) ≡ Li∈I MNi, and so W ∗(si(A )) = MNi for all i ∈ I. It remains to be shown that given an ncpsu-map f : A → MNf with W ∗(f (A )) = MNf there's a unique i ∈ I such that si is miu-equivalent to f . (Uniqueness) Let i, j ∈ I such that si and sj are miu-equivalent be given, and let ϕ : MNi → MNj be the associated nmiu-isomorphism with ϕ ◦ si = sj. We must show that i = j. Recall that by the universal property of ηA , there's a unique nmiu-map : Fha(A ) → MNj with sj = ◦ ηA . Surely, πj : Fha(A ) ≡Li′∈I MNi′ −→ MNj fits this description; but so does ϕ◦πi, since ϕ◦πi◦ηA = ϕ◦si = sj. Hence πj = ϕ ◦ πi. This entails that the carriers (see 69 IV) of πj and πi are equal, so i = j. (Existence) Let f : A → MNf be an ncpsu-map with MNf = W ∗(f (A )). We must show that there is an i ∈ I such that si is miu-equivalent with f . By the universal property of ηA there's a unique nmiu-map : Fha(A ) → MNf with f = ◦ ηA . We claim that must be of the form ′ ◦ πi for some nmiu- isomorphism ′ : MNi → MNf . First note that is surjective: indeed, (Fha(A )) is a von Neumann subal- gebra of MNf by 69 IVb, that contains f (A ). Thus (Fha(A )) ⊇ W ∗(f (A )) ≡ MNf , which implies that is surjective. Since is surjective, it maps central projections of Fha(A ) to central projections of MNf . For each i ∈ I let ci denote the central projection in Fha(A ) given by ci(i) = 1 and ci(j) = 0 for all j 6= i. Then the ci form an orthogonal family of central projections with Pi∈I ci = 1. So the (ci) form an orthogonal family of central projections of MNf as well, with Pi∈I (ci) = 1. Since the only non-zero central projection in MNf (being a factor, 67 II) is 1, it follows that there is exactly one i ∈ I with (ci) = 1, and that (cj) = 0 for all j 6= i. From this one easily deduces (c.f. 69 IVa) that must be of the form = ′ ◦ πi for some injective nmiu-map ′ : MNi → MNf . Since is surjective, ′ is surjective too, and thus ′ is a nmiu-isomorphism. functor theorem, remains to be seen. Whether this indicates an error in our proof, or the power of the adjoint Now, since = ′ ◦ πi, and f = ◦ ηA , we get f = ′ ◦ πi ◦ ηA ≡ ′ ◦ si. Since ′ is a nmiu-isomorphism, we see that f is miu-equivalent to si. (cid:3) Remark Given the concrete description for Fha from III it seems tempting to prove directly that hrir∈RA : A →Lr∈RA MNr is a universal arrow from A to the inclusion haW∗miu → haW∗cpsu, without presupposing the existence of Fha. However, our attempts to do so have been thwarted by our inability to prove that W ∗(hrir∈RA ) =Lr∈RA MNr using elementary means. To allow interpretation of ⊸ in haW∗miu, we'll show that the functor (· ) ⊗ B : haW∗miu → haW∗miu has a left adjoint (· )∗haB for every hereditarily atomic von Neumann algebra B. This result has already been established by Kornell (in Theorem 9.1 of [49]); we improve upon it by giving a different, slightly more concrete, description. As was the case for Fha, we establish the existence of (· )∗haB indirectly at first. Proposition Given a hereditarily atomic von Neumann algebra B, the functor (· )⊗ B : haW∗miu −→ haW∗miu has a left adjoint (· )∗haB : haW∗miu → haW∗miu. Proof We already know from 125 VIII that (· ) ⊗ B preserves limits as functor W∗miu → W∗miu. Since the subcategory haW∗miu of W∗miu is closed under prod- ucts and equalisers, the restriction of (· ) ⊗ B to a functor haW∗miu → haW∗miu preserves limits as well. The proof is now completed by an application of Freyd's adjoint functor theorem, exactly as in 125 IX, but with a suitably modified so- lution set. (cid:3) To describe A ∗haB concretely we need some notation. Definition We say that a nmiu-map s : A → C ⊗ B, where A , B and C are von Neumann algebras, is (· ) ⊗ B-surjective when the only von Neumann sub- algebra S of C with s(A ) ⊆ S ⊗ B is S = C , where for the sake of simplicity we regard S ⊗ B to be a von Neumann subalgebra of C ⊗ B (c.f. 115 V). By inspecting the proof of 125 IV one sees that for any nmiu-map s : A → C ⊗B there is a von Neumann subalgebra C of C such that s(A ) ⊆ C ⊗ B, and the restriction of s to a a map s : A → C ⊗ B is (· ) ⊗ B-surjective. Lemma Given a (· )⊗ B-surjective nmiu-map s : A → C ⊗ B and a nmiu-map : C → D between von Neumann algebras, the composition A s / C ⊗ B is (· ) ⊗ B-surjective iff is surjective. Proof Suppose that is surjective, and let S be a von Neumann subalgebra of D with ( ⊗ B)(s(A )) ⊆ S ⊗ B. To prove that ( ⊗ B) ◦ s is (· ) ⊗ B- / D ⊗ B ⊗B ..125c -- 125e.. 193 VIII 125d II III 125e II IIa III IV / / surjective, we must show that S = D. Since (⊗ B)(s(A )) ⊆ S ⊗ B, we have s(A ) ⊆ ( ⊗ B)−1(S ⊗ B) ≡ −1(S ) ⊗ B, by 125 VIIb, and so −1(S ) = C , because s is (· ) ⊗ B-surjective. Whence S = (−1(S )) ≡ (C ) = D, using here that is surjective. For the other direction suppose that ( ⊗ B) ◦ s is (· ) ⊗ B-surjective. Since the range of ⊗ B is (C ) ⊗ B, we have ( ⊗ B)(s(A )) ⊆ (C ) ⊗ B, and so (C ) = D, because ( ⊗ B) ◦ s is (· ) ⊗ B-surjective. (cid:3) V VI Definition Let A and B be hereditarily atomic von Neumann algebras. We'll describe A ∗haB in terms of the (· )⊗B-surjective nmiu-maps f : A → MNf ⊗B. Let us say that two such maps f1 : A → MNf1 ⊗ B and f2 : A → MNf2 ⊗ B are with (· ) ⊗ B-equivalent when there is a nmiu-isomorphism ϕ : MNf1 → MNf2 (ϕ ⊗ B) ◦ f1 = f2 (which implies that Nf1 = Nf2.) Pick a set of representatives SA ,B for this (· ) ⊗ B-equivalence. unique nmiu-map Φ : A ∗haB −→Ls∈SA ,B MNs that makes the diagram VII Theorem Let A and B be hereditarily atomic von Neumann algebras. the ηA ,B A '◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆◆ hsis∈SA ,B A ∗haB ⊗ B Φ⊗B ∼= hπs⊗Bis∈SA ,B (cid:0)Ls∈SA ,B MNs(cid:1) ⊗ B Ls∈SA ,B MNs ⊗ B VIII Proof We follow roughly the same lines as the proof of 125c III. Since A ∗haB commute is a nmiu-isomorphism. Here η( · ),B denotes the unit of the adjunction between (· )∗haB and (· ) ⊗ B. is hereditarily atomic we write A ∗haB ≡ Li∈I MNi without loss of generality. Note that writing ei = (πi ⊗ B) ◦ ηA ,B the diagram ηA ,B A +❳❳❳❳❳❳❳❳❳❳❳❳❳❳❳❳❳❳❳❳❳❳❳❳❳❳❳❳❳ heiii∈I A ∗haB ⊗ B ≡ Li∈I MNi ⊗ B ∼= hπi⊗Bii∈I Li∈I Mi ⊗ B commutes. We claim that the ei are (· ) ⊗ B-surjective, and, moreover, form a set of representatives for (· ) ⊗ B-equivalence on the set of nmiu-maps f : A → MNf ⊗ B. From this claim the theorem follows with a reasoning similar to that in 125c IV, which we won't repeat here. / / '     / / +   To prove that ei ≡ (πi ⊗ B) ◦ ηA ,B is (· ) ⊗ B-surjective, it suffices, by III, to show that ηA ,B is (· )⊗ B-surjective (since πi is surjective.) So let S be a von Neumann subalgebra of A ∗haB such that ηA ,B(A ) ⊆ S ⊗ B. We must show that S = A ∗haB. Letting f : A → S ⊗ B denote the restriction of ηA ,B, there is, by the universal property of ηA ,B, a unique nmiu-map : A ∗haB → S such that f = ( ⊗ B) ◦ ηA ,B. Note that if we compose with the inclusion e : S → A ∗haB, then we get a nmiu-map σ := e◦ : A ∗haB → A ∗haB with the property that (σ ⊗ B) ◦ ηA ,B = ηA ,B. Since the identity on A ∗haB is the only map with this property, we get e ◦ = id. This implies that e is surjective, and thus S = A ∗haB. Whence ηA ,B is (· ) ⊗ B-surjective. It remains to be show that for every nmiu-map f : A → MNf ⊗ B there is a unique i ∈ I such that ei is (· ) ⊗ B-equivalent to f . (Uniqueness) Suppose that ei and ej are (· ) ⊗ B-equivalent for some i, j ∈ I; we must show that i = j. Let ϕ : MNi → MNj be an nmiu-isomorphism with (ϕ ⊗ B) ◦ ei = ej. Note that πj : A ∗haB ≡Lj′∈I MNj′ −→ MNj is the unique nmiu-map : A ∗haB → MNj with ej = ⊗B◦ηA ,B. Since (ϕ◦πi)⊗B◦ηA ,B = (ϕ ⊗ B) ◦ (πi ⊗ B) ◦ ηA ,B = ϕ ⊗ B ◦ ei = ej, we get ϕ ◦ πj = πi. Hence i = j. (Existence) Let f : A → MNf ⊗ B be a (· )⊗ B-surjective nmiu-map. We must show that there is a unique i ∈ I such that f is (· )⊗ B-equivalent to ei. By the universal property of ηA ,B, there's a unique nmiu-map : A ∗haB −→ MNf with ( ⊗ B) ◦ ηA ,B = f . Note that is surjective by III. Now, following the same reasoning as in 125c VII, we see that : A ∗haB ≡ Li∈I MNi −→ MNf must be of the form ≡ ′ ◦ πi for some i ∈ I and nmiu-isomorphism ′ : MNi → MNf . So f = (⊗ B)◦ ηA ,B = (′ ⊗ B) ◦ (πi ⊗ B) ◦ ηA ,B = (′ ⊗ B)◦ ei, and hence f is (· ) ⊗ B-equivalent to ei. (cid:3) IX X XI 4.4 Duplicators and Monoids When asked for an interpretation of the type !A as a von Neumann algebra 126 JAK⊗n J!AK = Mn (4.6) definitely seems like a suitable answer given the cue that !A should represent as many instances of A as needed, which makes the interpretation we actually use in our model of the quantum lambda calculus (namely J!AK = ℓ∞(nsp(JAK))) rather suspect. To address such concerns we'll show that any von Neumann algebra that carries a ⊗-monoid structure (in W∗miu as J!AK should) must be nmiu-isomorphic to ℓ∞(X) for some set X (see 127 III) ruling out the interpre- tation (4.6) for all but the most trivial cases. We'll show in fact that ℓ∞(nsp(A )) is the free ⊗-monoid over A in W∗miu (see 132 IV) exonerating it in our minds from all doubts. ..125e -- 126 195 4.4.1 Duplicators 127 Definition A von Neumann algebra A is duplicable if there is a duplicator on A , that is, an npsu-map δ : A ⊗ A → A with a unit u ∈ [0, 1]A satisfying δ(a ⊗ u) = a = δ(u ⊗ a) for all a ∈ A . (Note that we require of δ neither associativity nor commutativity.) II Remark The unit u can be identified with a positive subunital map u : C → A via u(λ) = λu. The definition is motivated by the fact that the interpretation of !A must carry a commutative monoid structure in W∗miu. The condition is weaker, requiring the maps to be only positive subunital, and dropping associa- tivity and commutativity. Nevertheless this is sufficient to prove the following. III Theorem A von Neumann algebra A is duplicable if and only if A is nmiu- In that case, the duplicator (δ, u) is isomorphic to ℓ∞(X) for some set X. unique, given by δ(a ⊗ b) = a · b and u = 1. IV Thus, to interpret duplicable types, we can really only use von Neumann alge- bras of the form ℓ∞(X). It also follows that a von Neumann algebra is duplica- ble precisely when it is a (commutative) monoid in W∗miu, or in the symmetric monoidal category W∗cpsu of von Neumann algebras and normal completely pos- itive subunital (CPsU) maps. V To prove III we proceed as follows. First we prove in 128 VIII every duplica- ble von Neumann algebra A is commutative (and that the duplicator is given by multiplication). This reduces the problem to a measure theoretic one, be- cause A ∼= Li L∞(Xi) for some finite complete measure spaces Xi (by 70 III). Since each of the L∞(Xi)s will be duplicable (see 128 XIII) we may assume without loss of generality that A ∼= L∞(X) for some finite complete measure space X. Since X splits into a discrete and a continuous part (see 129 VI), and the result is obviously true for discrete spaces, we only need to show that L∞(C) = {0} for any continuous complete finite measure space C for which L∞(C) is duplicable. In fact, we'll show that µ(C) = 0 for such C (see 129 VIII). VI Lemma Let δ be a duplicator with unit u on a von Neumann algebra A . Then u = 1 and δ(1 ⊗ 1) = 1. VII Proof Since 1 = δ(u ⊗ 1) 6 δ(1 ⊗ 1) 6 1, we have δ(1 ⊗ 1) = 1, and so δ(u⊥ ⊗ 1) = 0. But, because u⊥ = δ(u⊥ ⊗ u) 6 δ(u⊥ ⊗ 1) = 0, we have u⊥ = 0, and thus u = 1. Hence 1 = δ(1 ⊗ u) = δ(1 ⊗ 1). (cid:3) 128 To prove that a duplicable von Neumann algebra is commutative we'll need the following two classical theorems from the theory of C∗-algebras. Theorem (Tomiyama) Any linear surjection f : A → B of a von Neumann algebra A onto a von Neumann subalgebra B ⊆ A with f (f (a)) = f (a) and kf (a)k 6 kak for all a ∈ A obeys bf (a) = f (ba) for all a ∈ A and b ∈ B. Remark The usual (see e.g. 10.5.86 of [47]) and original [76] formulations of Tomiyama's theorem involve C∗-algebras instead of von Neumann algebras, and include the conclusion that f is positive. Since these improvements weren't necessary for our purposes, we've left them out, shortening the proof. Proof (Based on II.6.10.2 of [3].) Let a ∈ A and b ∈ B be given. Since b is the norm limit of linear com- binations of projections (cf. 65 IV), it suffices to show that ef (a) = f (ea) for every projection e from B and a ∈ A . For this, in turn, it suffices to show that e⊥f (ea) = 0 for every projection e from B, (and thus also ef (e⊥a) = 0,) because then f (ea) = ef (ea) = ef (a). Let λ ∈ R be given. The trick is to obtain the following inequality. II IIa III 2 (1 + 2λ)(cid:13)(cid:13)e⊥f (ea)(cid:13)(cid:13) 6 keak2 (4.7) Indeed, this inequality can only hold for all λ when ke⊥f (ea)k = 0. Work- ing towards (4.7), let us first note that f (e⊥f (ea)) = e⊥f (ea): indeed, since e⊥f (ea) ∈ B and f : A → B is surjective, there must be a′ ∈ A with f (a′) = e⊥f (ea), and thus e⊥f (ea) = f (a′) = f (f (a′)) = f (e⊥f (ea)). Then: 2 2 2 2 (1 + λ)2(cid:13)(cid:13)e⊥f (ea)(cid:13)(cid:13) = (cid:13)(cid:13)e⊥f ( ea + λe⊥f (ea) )(cid:13)(cid:13) 6 (cid:13)(cid:13) ea + λe⊥f (ea)(cid:13)(cid:13) = keak2 + λ2(cid:13)(cid:13)e⊥f (ea)(cid:13)(cid:13) Subtracting λ2(cid:13)(cid:13)e⊥f (ea)(cid:13)(cid:13) [Moved to 34a VII] [Removed] 2 since f (e⊥f (ea)) = e⊥f (ea) since ke⊥k 6 1 and kfk 6 1 using kck2 = kc∗ck and ee⊥ = 0 from both sides yields inequality (4.7). (cid:3) Lemma Let A be a von Neumann algebra, and let f : A ⊕ A → A be a pu-map with f (a, a) = a for all a ∈ A . Then p := f (1, 0) is central, and f (a, b) = ap + bp⊥ for all a, b ∈ A . Proof (Based on Lemma 8.3 of [30].) Note that (c, d) 7→ ( f (c, d), f (c, d) ) gives a pu-map f′ from A ⊕ A onto its von Neumann subalgebra { (a, a) : a ∈ A } with f′(f′(c, d)) = f′(c, d) for 127 -- 128.. 197 IV V VI VII all c, d ∈ A . Since kf′k 6 1 as a result of Russo -- Dye's theorem (see 34a VIII), Tomiyama's theorem (II) implies that for all a, c, d ∈ A , (a, a) f′(c, d) = f′( ac, ad ), and so a f (c, d) = f ( ac , ad ). Upon taking the adjoint we see that f (c, d)b = f (cb, db) for all b, c, d ∈ A (using here that f being positive is involution preserving, see 10 IV.) As a result of these observations, we get ap ≡ af (1, 0) = f (a, 0) = f (1, 0)a ≡ pa for all a ∈ A , and so p is central. Similarly, f (0, b) = bp⊥ for all b ∈ A . Then f (a, b) = f (a, 0) + f (0, b) = ap + bp⊥ for all a, b ∈ A . (cid:3) VIII Lemma Let δ : A ⊗ A → A be a duplicator on a von Neumann algebra A . Then A is commutative and δ(a ⊗ b) = a · b for all a, b ∈ A . IX Proof To prove A is commutative we must show that all a ∈ A are central, but, of course, it suffices to show that all p ∈ [0, 1]A are central (by the usual reasoning). Similarly, we only need to prove that δ(a ⊗ p) = a · p for all a ∈ A and p ∈ [0, 1]A . Given such p ∈ [0, 1]A define f : A ⊕ A → A by f (a, b) = δ(a ⊗ p + b ⊗ p⊥) for all a, b ∈ A . Then f is positive, unital, f (1, 0) = p, and f (a, a) = a for all a ∈ A . Thus by VI, p is central, and f (a, b) = ap + bp⊥ for all a, b ∈ A . Then a · p = f (a, 0) = δ(a ⊗ p). (cid:3) X Remark The special case of VIII in which δ is completely positive can be found in the literature, see for example Theorem 6 of [50] (where duplication is called copying,) and §3.3 of [13] (where it's called broadcasting.) Xa Remark Note that we've not yet used that a duplicator is normal. That a du- plicator is normal will come in to play momentarily when we show that L∞(X) is duplicable iff X is discrete. XI Corollary Let A be a von Neumann algebra. Then A is duplicable iff there is an np-map δ : A ⊗ A → A with δ(a ⊗ b) = a · b for all a, b ∈ A , (and in that case A is commutative.) XII Remark Thus for a non-commutative von Neumann algebra A multiplication (a, b) 7→ ab : A × A → A is not a normal bilinear map in the sense of 112 II. XIII Corollary When the direct sum A ⊕ B of von Neumann algebras A and B is duplicable, A and B are duplicable XIV Proof Let δ : (A ⊕ B) ⊗ (A ⊕ B) −→ A ⊕ B be a duplicator on A ⊕ B. By VIII A ⊕ B is commutative and δ((a1, b1) ⊗ (a2, b2)) = (a1a2, b1b2) for all a1, a2 ∈ A and b1, b2 ∈ B. Let κ1 : A → A ⊕ B be the nmiu- map given by κ1(a) = (a, 0) for all a ∈ A . Let δA be the composition of / A . Then δA is A ⊗ A κ1⊗κ1 / normal, positive, and δA (a1 ⊗ a2) = π1(δ((a1, 0) ⊗ (a2, 0))) = π1(a1a2, 0) = a1a2 for all a1, a2 ∈ A . Thus, by XI, A is duplicable. (cid:3) / (A ⊕ B) ⊗ (A ⊕ B) / A ⊕ B π1 δ / / We will now work towards the proof that if C is a continuous complete finite measure space, then L∞(C) cannot be duplicable unless µ(C) = 0, see X. Let us first fix some more terminology from measure theory (see 51 and [19]). Definition Let X be a finite complete measure space. 129 II 1. A measurable subset A of X is atomic if 0 < µ(A) and µ(A′) = µ(A) for all A′ ∈ ΣX with A′ ⊆ A and µ(A′) > 0. 2. X is discrete if X is covered by atomic measurable subsets. (This coincides with being "purely atomic" from 211K of [19].) 3. X is continuous (or "atomless") if X contains no atomic subsets. The following lemma, which will be very useful, is a variation on Zorn's Lemma (that does not require the axiom of choice). Lemma Let S be a collection of measurable subsets of a finite complete measure space X such that for every ascending countable sequence A1 ⊆ A2 ⊆ ··· in S there is A ∈ S with A1 ⊆ A2 ⊆ ··· ⊆ A. Then each element A ∈ S is contained in some B ∈ S that is maximal in S in the sense that µ(B′) = µ(B) for all B′ ∈ S with B ⊆ B′. Proof The trick is to consider for every C ∈ S the quantity βC = sup{ µ(D) : C ⊆ D and D ∈ S }. Note that µ(C) 6 βC 6 µ(X) for all C ∈ S, and βC2 6 βC1 for all C1, C2 ∈ S with C1 ⊆ C2. To prove this lemma, it suffices to find B ∈ S with A ⊆ B and µ(B) = βB. Define B1 := A. Pick B2 ∈ S such that B1 ⊆ B2 and βB1 − µ(B2) 6 1/2. Pick B3 ∈ S such that B2 ⊆ B3 and βB2 − µ(B3) 6 1/3. Proceeding in this fashion, we get a sequence B ≡ B1 ⊆ B2 ⊆ ··· in S with βBn − µ(Bn+1) 6 1/n for all n. By assumption there is a B ∈ S with B1 ⊆ B2 ⊆ ··· ⊆ B. Note that µ(B1) 6 µ(B2) 6 ··· 6 µ(B) 6 βB 6 ··· 6 βB2 6 βB1 . Since for every n ∈ N we have both µ(Bn+1) 6 µ(B) 6 βB 6 βBn and βBn − µ(Bn+1) 6 1/n, we get βB − µ(B) 6 1/n, and so βB = µ(B). (cid:3) Lemma Each finite complete measure space X contains a discrete measurable subset D such that X\D is continuous. Proof Since clearly the countable union of discrete measurable subsets of X is again discrete, there is by IV a discrete measurable subset D of X which is maximal in the sense that µ(D′) = µ(D) for every discrete measurable subset D′ of X with D ⊆ D′. To show that X\D is continuous, we must prove that X\D contains no atomic measurable subsets. If A ⊆ X\D is an atomic measurable ..128 -- 129.. 199 III IV V VI VII subset of X, then D ∪ A is a discrete measurable subset of X which contains D, and µ(D ∪ A) = µ(D) ∪ µ(A) > µ(D). This contradicts the maximality of D. Thus X\D is continuous. (cid:3) VIII Lemma Given a continuous finite complete measure space X, and r ∈ [0, µ(X)], there is a measurable subset A of X with µ(A) = r. IX Proof Let us quickly get rid of the case that µ(X) = 0. Indeed, then r = 0, and so A = ∅ will do. For the remainder, assume that µ(X) > 0. For starters, we show that for every ε > 0 and B ∈ ΣX with µ(B) > 0 there is A ∈ ΣX with A ⊆ B and 0 < µ(A) < ε. Define A1 := B. Since µ(B) > 0, and A1 is not atomic (because X is continuous) there is A ∈ ΣX with A ⊆ A1 and µ(A) 6= µ(A1). Since µ(A) + µ(A1\A) = µ(A1), either 0 < µ(A) 6 1 2 µ(A1) or 0 < µ(X\A) 6 1 2 µ(A1). In any case, there is A2 ⊆ A1 with A2 ∈ ΣX and 0 < µ(A2) 6 1 2 µ(A1). Similarly, since A2 is not atomic (because X is continuous), there is A3 ⊆ A2 with A3 ∈ ΣX and 0 < µ(A3) 6 1 2 µ(A2). Proceeding in a similar fashion, we obtain a sequence B ≡ A1 ⊇ A2 ⊇ ··· of measurable subsets of X with 0 < µ(An) 6 2−nµ(X). Then, for every ε > 0 there is n ∈ N such that 0 < µ(An) 6 ε and An ⊆ B. Now, let us prove that there is A ∈ ΣX with µ(A) = r. By IV there is a measurable subset A of X with µ(A) 6 r and which is maximal in the sense that µ(A′) = µ(A) for all A′ ∈ ΣX with µ(A) 6 r and A ⊆ A′. In fact, we claim that µ(A) = r. Indeed, suppose that ε := r − µ(A) > 0 towards a contradiction. By the previous discussion, there is C ∈ ΣX with C ⊆ X\A such that µ(C) 6 ε. Then A ∪ C is measurable, and µ(A ∪ C) = µ(A) + µ(C) 6 µ(A) + ε 6 r, which contradicts the maximality of A. (cid:3) X Lemma Let X be a continuous finite complete measure space for which L∞(X) is duplicable. Then µ(X) = 0. XI Proof Suppose that µ(X) > 0 towards a contradiction. Let δ be a duplicator Let ω : L∞(X) → C be given by ω(f◦) = 1 on L∞(X). By 128 VIII δ(f ⊗ g) = f · g for all f, g ∈ L∞(X). µ(X)R f dµ for all f ∈ L∞(X). Then ω is normal, positive, unital and faithful (cf. 51 IX). We'll use the product functional ω ⊗ ω : L∞(X) ⊗ L∞(X) → C, (which is also faithful, by 118 IV) to tease out a contradiction, but first we need a second ingredient. Since X is continuous, we may partition X into two measurable subsets of equal measure with the aid of VIII, that is, there are measurable subsets X1 and X2 of X with X = X1 ∪ X2, X1 ∩ X2 = ∅, and µ(X1) = µ(X2) = 1 2 µ(X). Similarly, X1 can be split into two measurable subsets, X11 and X12, of equal measure, and so on. In this way, we obtain for every word w over the alphabet {1, 2} -- in symbols, w ∈ {1, 2}∗ -- a measurable subset Xw of X such that Xw = Xw1 ∪ Xw2, Xw1 ∩ Xw2 = ∅, and µ(Xw1) = µ(Xw2) = 1 2 µ(Xw). It follows that µ(Xw) = 1 2#w µ(X), where #w is the length of the word w. Now, pw := 1◦Xw is a projection in L∞(X), and ω(pw) = 2−#w for every w ∈ {1, 2}∗. Moreover, pw = pw1 + pw2, and so pw ⊗ pw = pw1 ⊗ pw1 + pw1 ⊗ pw2 + pw2 ⊗ pw1 + pw2 ⊗ pw2 > pw1 ⊗ pw1 + pw2 ⊗ pw2. Thus, if we define qN := Pw∈{1,2}N pw ⊗ pw for every natural number N , where {1, 2}N is the set of words over {1, 2} of length N , then we get a descend- ing sequence q1 > q2 > q3 > ··· of projections in L∞(X)⊗ L∞(X). Let q be the infimum of q1 > q2 > ··· in the set of self-adjoint elements of L∞(X)⊗ L∞(X). Do we have q = 0 ? On the one hand, we claim that δ(q) = 1, and so q 6= 0. Indeed, δ(pw⊗pw) = pw · pw = pw for all w ∈ {1, 2}N . Thus δ(qN ) = Pw∈{1,2}N δ(pw ⊗ pw) = Pw∈{1,2}N pw = 1 for all N ∈ N. Hence δ(q) = Vn δ(qN ) = 1, because δ is normal. On the other hand, we claim that (ω⊗ω)(q) = 0, and so q = 0 since ω⊗ω Indeed, (ω ⊗ ω)(qN ) = Pw∈{1,2}N ω(pw) · ω(pw) = is faithful and q > 0. Pw∈{1,2}N 2−N·2−N = 2−N for all N ∈ N, and so (ω⊗ω)(q) =VN (ω⊗ω)(qN ) = VN 2−N = 0. Thus, q = 0 and q 6= 0, which is impossible. This takes care of the continuous case. To deal with the discrete case we first need some simple observations. Lemma Let A be an atomic measure space. Then L∞(A) ∼= C. Proof Let f ∈ L∞(A) be given. It suffices to show that there is z ∈ C such that f (x) = z for almost all x ∈ A. Moreover, we only need to consider the case that f takes its values in R (because we may split f in its real and imaginary parts, and in turn split these in positive and negative parts). (cid:3) 130 II III Let S be some measurable subset of A. Note that either µ(S) = 0 or µ(A\S) = 0. Indeed, if not µ(S) = 0, then µ(S) > 0, and so µ(S) = µ(A) (by atomicity of A), which entails that µ(A\S) = 0. In particular, for every real number t ∈ R one of the sets { x ∈ A : f (x) < t} { x ∈ A : t 6 f (x)} must be negligible. Whence either t 6 f◦ or f◦ 6 t. It follows that the two closed sets L := {t ∈ R : t 6 f◦} and U := {t ∈ R : f◦ 6 t} cover R. Since clearly −kfk ∈ L and kfk ∈ U , the sets L and U can't be disjoint, because they would partition R into two clopen non-empty sets. For an element t ∈ L ∩ U in the intersection we have t 6 f◦ 6 t, and so t = f◦. Hence L∞(X) ∼= C. (cid:3) Exercise Let X be a measure space with µ(X) < ∞. Show that L∞(X) ∼= L∞(A) for every countable partition A of X consisting of measurable LA∈A subsets. Corollary For every discrete measure space X with µ(X) < ∞ there is a set Y with L∞(X) ∼= ℓ∞(Y ). IV V ..129 -- 130 201 131 We are now ready to prove the main result of this section. II Proof of 127 III We have already seen that ℓ∞(X) can be equipped with a commutative monoid structure in W∗miu for any set X, and is thus duplicable. Conversely, let δ : A ⊗ A → A be a duplicator with unit u on a von Neumann algebra A . By 127 VI, we know that u = 1, and by 128 VIII, we know that A is commutative and δ(a ⊗ b) = a · b for all a, b ∈ A . Thus, the only thing that remains to be shown is that A is miu-isomorphic to ℓ∞(Y ) for some set Y . By 70 III A ∼= Li L∞(Xi) for some finite complete measure spaces Xi. So to prove that A ∼= ℓ∞(Y ) for some set Y it suffices to find a set Yi with L∞(Xi) ∼= ℓ∞(Yi) for each i, because then A ∼= Li∈I ℓ∞(Yi) ∼= ℓ∞(cid:0)Si∈I Yi(cid:1). Let i ∈ I be given. Since A ∼= L∞(Xi) ⊕ Lj6=i L∞(Xj) is duplica- ble, L∞(Xi) is duplicable by 128 XIII. By 129 VI there is a measurable sub- set D of Xi such that D is discrete, and C := X\D is continuous. We have L∞(Xi) ∼= L∞(D)⊕L∞(C) by 130 IV, and so L∞(D) and L∞(C) are dupli- cable (again by 128 XIII). By 129 X, L∞(C) can only be duplicable if µ(C) = 0, and so L∞(C) ∼= {0}. On the other hand, since D is discrete, we have L∞(D) ∼= ℓ∞(Y ) for some set Y (by 130 V). All in all, we have L∞(Xi) ∼= ℓ∞(Y ). (cid:3) 4.4.2 Monoids 132 We further justify our choice, J!AK = ℓ∞(nsp(JAK)), by proving that ℓ∞(nsp(A )) is the free (commutative) monoid on A in W∗miu. As a corollary, we also obtain that ℓ∞(W∗cpsu(A , C)) is the free (commutative) monoid on A in W∗cpsu. II Let us first recall some terminology. Given a symmetric monoidal category (SMC) C, a monoid in C is an object A from C endowed with a multiplication map m : A ⊗ A → A and a unit map u : I → A satisfying the associativity and the unit law, i.e. making the following diagrams commute. (A ⊗ A) ⊗ A α m⊗id / A ⊗ A m A ⊗ (A ⊗ A) / A ⊗ A id⊗m m / / A I ⊗ A / A ⊗ A m u⊗id &▼▼▼▼▼▼▼▼▼ λ A A ⊗ I id⊗u xqqqqqqqqq ρ Here α, λ, ρ respectively denote the associativity isomorphism, and the left and the right unit isomorphism. A monoid A is commutative if m ◦ γ = m, where γ : A⊗ A → A⊗ A is the symmetry isomorphism. A monoid morphism between monoids A1 and A2 is an arrow f : A1 → A2 that satisfies mA2◦(f⊗f ) = f◦mA1 and uA2 = f ◦ uA1. We denote the category of monoids and monoid morphisms in C by Mon(C). The full subcategory of commutative monoids is denoted by cMon(C). Recall that W∗miu and W∗cpsu are symmetric monoidal categories   /   / & /   o o x with C as tensor unit (see 119 V), and so we may speak about monoids in W∗miu and W∗cpsu. Exercise Let A be a von Neumann algebra. 1. Show that any monoid structure on A in W∗cpsu is a duplicator on A . 2. Deduce from this and 127 III that there is a monoid structure on A in W∗miu or in W∗cpsu iff A is duplicable iff A ∼= ℓ∞(X) for some set X; and that, in that case the multiplication m : A ⊗A → A of the monoid is commutative and uniquely being fixed by m(a ⊗ b) = a · b. 3. Show that the monoid morphisms in W∗miu and in W∗cpsu are precisely the nmiu-maps. 4. Conclude that cMon(W∗miu) = Mon(W∗miu) = cMon(W∗cpsu) = Mon(W∗cpsu). 5. Show that Mon(W∗miu) ∼= dW∗miu ≃ Setop, where dW∗miu denotes the full subcategory of W∗miu consisting of duplicable von Neumann algebras. (Hint: ℓ∞ : Set → (W∗miu)op is full and faithful by 122 VI.) Theorem Let A be a von Neumann algebra, and let η : A → ℓ∞(nsp(A )) be the nmiu-map given by η(a)(ϕ) = ϕ(a). Then ℓ∞(nsp(A )) is the free (commu- tative) monoid on A in W∗miu via η. Proof Let B be a monoid on W∗miu, and let f : A → B be an nmiu-map. We must show that there is a unique monoid morphism g : ℓ∞(nsp(A )) → B such that g ◦ η = f . Since the monoid structure on B is a duplicator on B we may assume, by 127 III, that B = ℓ∞(Y ) for some set Y . Since nsp : (W∗miu)op → Set is left adjoint to ℓ∞ : Set → (W∗miu)op with unit η (see 122 II), there is a unique map h : Y → nsp(A ) with ℓ∞(h)◦ η = f . Since ℓ∞ is full and faithful by 122 VI, the only thing that remains to be shown is that ℓ∞(h) is a monoid morphism. Indeed it is, since the monoid multiplication on ℓ∞(nsp(A )) and ℓ∞(Y ) is given by ordinary multiplication, which is preserved by ℓ∞(h) being an miu-map. (cid:3) III IV V Corollary Let A be a von Neumann algebra. Then ℓ∞(W∗cpsu(A , C)) is the free (commutative) monoid on A in W∗cpsu. Proof By IV ℓ∞ ◦ nsp is a left adjoint to the forgetful functor Mon(W∗miu) → W∗miu. Note that by III, the forgetful functor Mon(W∗cpsu) → W∗cpsu factors through W∗miu as: VI VII Mon(W∗cpsu) ℓ∞◦nsp Mon(W∗miu) ⊥ / / W∗miu F ⊥ / / W∗cpsu where F is from 124 III. Thus the free monoid on A in W∗cpsu is given by: (ℓ∞ ◦ nsp ◦ F )(A ) = ℓ∞(W∗miu(F A , C)) ∼= ℓ∞(W∗cpsu(A , C)) 131 -- 132.. 203 w w y y as was claimed. (cid:3) 133 Conclusion Here ends this thesis, but not the entire story. There's much more to be said about self-dual Hilbert A -modules, about dilations and their relation to purity, and about the abstract theory of corners, filters, and ⋄-positivity. You'll see all this, and more, in the sequel, "Dagger and dilations in the category of von Neumann algebras" [84], brought to you by my twin brother. 134 (Paragraphs numbered 134 and up can be found in [84].) Index :=, is defined to be, 1 Va ≡, being of the form, 1 Va ≈ space, 52 II A ≈ B, for subsets of a topological f ≈ g, for measurable functions, 51 V 6, order on a C∗-algebra, 9 IV ., Murray -- von Neumann preorder, 83 II (· )′ f′, derivative of a holomorphic func- tion, 12 II (· )I, imaginary part aI, of an element of a C∗-algebra, 7 II (· )R, real part AR, of a C∗-algebra, 7 II aR, of an element of a C∗-algebra, 7 II (· )⊥ (· )∗ a⊥, orthosupplement of an effect, 9 IX p⊥, orthocomplement of a projec- tion, 55 IV adjoint of an operator, 4 VIII involution on a C∗-algebra, 3 I (· )+, positive part A+, of a C∗-algebra, 9 IV a+, of a self-adjoint element of a C∗-algebra, 24 I (· )−, negative part a−, of a self-adjoint element of a C∗-algebra, 24 I (a), order ideal generated by a, 22 III ∡(w0, z, w1), angle between complex num- bers, 14 III A ∗B, free exponential, 125 VIII A ∗haB, 125d II a ∗ ω, 72 II (X )1, unit ball, 4 IV (X )r, r-ball, 4 IV !, 120 I T A, infimum of projections, 56 XVI S A, supremum of projections, 56 XVI [· ] [a], partial isometry from the po- lar decomposition of a, 82 I [f ], for an ncp-map, 98 IX ⌈⌈·⌉⌉ ⌈·⌉ ⌈⌈a⌉⌉, central support, 68 I, 83 V ⌈⌈f⌉⌉, central carrier, 69 I ⌈a⌉, ceiling, 56 I, 59 I ⌈f⌉, carrier of an np-map, 63 I ⌈a), support, 59 I (a⌉, range, 59 I ⌈e⌉S, 88 II f⋄, 101 I hfi, 105 II f⋄, 101 II eA e, corner, 94 I ⌊·⌋ γ(f, g), product functional, 108 II ⌊a⌋, of an effect, 56 VI in a von Neumann algebra, 43 Ia V F , infimum of F W D, supremum of D in B(H ), 37 XI in a von Neumann algebra, 42 I [· , · ]ω, given np-functional ω, 30 II xihy, with x, y ∈ H , 4 XIX nihm, with n, m ∈ N, 43 II ⊸, 120 I J· K, 120 I a∼1, pseudoinverse of a, 79 I a/b 205 in a von Neumann algebra, 81 I, C-valued, 4 VIII 81 III [a, b], interval a\c/b in a von Neumann algebra, 81 II, 81 VII √a, square root in a C∗-algebra, 23 VII S(cid:3), commutant of S, 65 II f◦, equivalence class of f , 51 V 115 I algebras, 115 I ⊗ H ⊗ K , of Hilbert spaces, 110 VI A ⊗ B, of von Neumann algebras, a⊗ b, of elements of von Neumann A⊗B, of operators between Hilbert bifunctor on W∗miu, . . . , 115 IV f ⊗ g, of normal functionals, 116 I f ⊗ g, of np-maps, 115 II x⊗y, of elements of Hilbert spaces, spaces, 111 V 110 VI ⊙, algebraic tensor product, 112 I β⊙, 112 I βγ, 112 XI β⊗, 115 III A∗, predual of A , 87 I L, direct sum Ai, of C∗-algebras, 3 V Ai, of von Neumann algebras, 42 V Hi, of Hilbert spaces, 6 II Li Li Li k · k, norm of an operator, 4 II on a C∗-algebra, 3 I on a pre-Hilbert A -module, 32 IX on a pre-Hilbert space, 4 XV supremum ∼, 3 V 1A, indicator function on [0, 1], 14 II h· , ·i, inner product A -valued, 32 I definite, 32 I [a, b]A , in a C∗-algebra, 9 IVa A A∗(· )A : Ba(X) → Ba(Y ) a∗(· )a : A → A is completely positive, 34 V is completely positive, 34 V is normal, 44 XV adjoint of a adjointable map between pre- Hilbert A -modules, 32 I of an operator, 4 VIII adjointable map between pre-Hilbert A -modules, 32 I operator, 4 VIII almost clopen, 52 II αA ,B,C , associator, 119 IV approximate pseudoinverse, 80 II associator, 119 IV atomic subset of a measure space, 129 II B BA , 14 II Baire's Category Theorem, 54 II Ba(X), 32 I as a von Neumann algebra, 49 II as C∗-algebra, 32 XIII Ba(X, Y ), 32 I BC(X) as a C∗-algebra, 3 VI Bessel's inequality, 39 IV B(H ) as a C∗-algebra, 4 I, 5 XII as a von Neumann algebra, 42 V Bicommutant Theorem, 88 VI bilinear map bounded, 112 II completely positive, 112 II normal, 112 II bipositive map between C∗-algebras, 20 VI bound for a linear map, 4 II braiding, 119 IVc B(X ), 4 II B(X , Y ), 4 II C C, the complex numbers as a C∗-algebra, 3 III as a Hilbert space, 4 IX as a von Neumann algebra, 42 V C(X), 3 VI as a C∗-algebra, 3 VI c00 as a pre-Hilbert space, 4 IX carrier, 63 I category of von Neumann algebras, 47 III Cauchy's Integral Formula, 15 I Cauchy -- Schwarz inequality for A -valued inner products, 32 VI for C-valued inner products, 4 XV cC∗: cC∗miu, cC∗pu,. . . , 10 III ceiling, 56 I, 59 I central element of a von Neumann alge- bra, 67 I central carrier, 69 I central support, 68 III centre of a von Neumann algebra, 65 II centre separating collection of maps on a C∗-algebra, 21 II of np-functionals, 90 II CH, 29 I Choi's Theorem, 34 XVIII commutant, 65 II completely positive bilinear map, 112 II, 113 I map between C∗-algebras, 10 II, 34 II contraposed, 101 VI corner (map), 95 I standard, 98 I unital, 95 I corner (von Neumann algebra), 94 I cp, standard filter for p, 98 I C∗: C∗miu, C∗cpu, . . . , 10 III C∗(a), C∗-subalgebra generated by a, 28 II C∗-algebra, 3 I commutative, 3 I, 27 XXVII concrete, 4 I finite dimensional, 3 VIII, 84 II of bounded operators, 4 I C∗-identity, 3 I C∗-subalgebra, 3 IV cyclic projection, 66 III D definitions, 1 V derivative of a holomorphic function, 12 II direct sum of C∗-algebras, 3 V of Hilbert spaces, 6 II of von Neumann algebras, 42 V dom(f ), domain of an A -valued par- tial function, 12 II Double Commutant Theorem, 88 VI Douglas' Lemma, 81 V duplicable von Neumann algebra, 127 I is commutative, 128 VIII duplicator, 127 I is multiplication, 128 VIII E Eff , 97 I effect in a C∗-algebra, 9 IX effectus, 47 VI equaliser in C∗cpsu, 34 VI in W∗miu and W∗cpsu, 47 V in C∗miu and C∗pu, 20a II (· ) ⊗ B-equivalent, 125e VI equivalent ncp-maps, 101 V essential supremum norm, 51 V 207 ηω, 30 VI extremal disconnectedness, 53 III F F : W∗cpsu → W∗miu, 124 III factor, 67 III faithful collection of maps on a C∗-algebra, 21 II Fha : haW∗cpsu −→ haW∗miu, 125b II filter, 96 I for p, 96 I standard, 98 I FinPAC, 47 VI floor of an effect, 56 VI form, between Hilbert A -modules, 36 IV bounded, 36 IV free exponential, 125 VIII free monoid in W∗cpsu, 132 VI in W∗miu, 132 IV function holomorphic (at x), 12 II A -valued & partial, 12 II functional basic, on A ⊙ B, 112 II positive, 86 II normal, 89 IX simple, on A ⊙ B, 112 II ultraweakly continuous, 86 XII vector, 21 III f (a), continuous functional calculus, 28 II G γA ,B, braiding, 119 IVc γ, Gelfand representation, 27 III Gelfand -- Mazur's Theorem, 16 VII Gelfand -- Naimark's Theorem, 30 XIV Goursat's Theorem, 14 IV H Hahn -- Banach's Theorem, 73 IV haW∗miu, haW∗cpsu, . . . , 84b II Hellinger -- Toeplitz's Theorem, 35 VIII hereditarily atomic von Neumann algebra, 84b II Hilbert A -module, 32 I self dual, 36 I Hilbert space, 4 VIII holomorphic function, 12 II HΩ, 30 IX Hω, 30 VI ∗-homomorphism, 10 III I imaginary part of an element of a C∗-algebra, 7 II inclusion of a corner, 94 I inner product C-valued, 4 VIII completion, 30 V definite, 4 VIII A -valued, 32 I invertible element of a C∗-algebra, 11 VI involution on a C∗-algebra, 3 I involution preserving bilinear map, 108 I map between C∗-algebras, 10 II K Kadison's inequality, 30 IV Kaplansky's Density Theorem, 74 IV for von Neumann algebras, 48 VIII Gelfand -- Naimark -- Segal (GNS), 30 VI Gelfand's Representation Theorem, 27 XXVII for von Neumann algebras, 53 II geometric series, 11 II, 11 VII L ℓ2 as a Hilbert space, 4 IX ℓ2-bounded bilinear map, 110 I λA , left unitor, 119 IVb L∞(X), 51 II ℓ∞(X), 3 V as a C∗-algebra, 3 V ℓ∞-bounded bilinear map, 110 II L∞(X), 51 V ℓ∞ : Set → (W∗miu)op, 122 II M Mn, the n × n-matrices MnA , the n × n-matrices over A as a C∗-algebra, 3 VII as a C∗-algebra, 33 I as a von Neumann algebra, 49 IV Mnf , 33 III is normal, 49 IV meagre, 52 II measurable function, 51 I measure complete, 51 I finite, 51 I measure space continuous, 129 II discrete, 129 II miu-bilinear, 108 I miu-equivalent, 125c II miu-map, 10 III injective is isometry, 29 VIII Mnβ, for bilinear β, 113 III monoid in an SMC, 132 II multiplicative bilinear map, 108 I map between C∗-algebras, 10 II ncp-map, 99 XII ncpsu-map, 99 II element of a C∗-algebra, 28 II functional, 42 II positive functional, 46 III, 89 IX on B(H ), 38 I, 39 IX positive map between von Neumann algebras, 44 XV, 48 II nsp : (W∗miu)op → Set, 122 I O operator, 4 II adjointable, 4 VIII bounded, 4 II operator norm, 4 II order ideal of a C∗-algebra, 22 II maximal, 22 II proper, 22 II order separating collection of maps on a C∗-algebras, 21 II of pu-maps on a C∗-algebra, 21 VII orthogonal projections, 55 XII orthonormal basis, for a Hilbert space, 39 II orthonormal, subset of a Hilbert space, 39 II maximal, 39 II orthosupplement operation in a C∗-algebra, 9 IX P parallelogram law, 4 XV Parseval's identity, 39 IV partial isometry in a von Neumann algebra, 79 I πj, projection in C∗miu, 20a I Murray -- von Neumann preorder, 83 II µX , measure, 51 I πp, standard corner of p, 98 I polar decomposition N negligible, subset of a measure space, 51 I normal bilinear map, 112 II of a functional, 86 IX of an element of a von Neumann algebra, 82 I polarisation identity for an inner product, 4 XV in a von Neumann algebra, 44 II 209 positive is normal, 48 V completely ∼ bilinear map, 112 II completely ∼ map between C∗-algebras, element of a C∗-algebra, 9 IV, 17 V, 10 II ω, 30 VI is normal, 48 III Riesz decomposition lemma, 26 III Riesz ideal, 27 VII 25 I map between C∗-algebras, 10 II ⋄-positive, 103 I power series, 13 II predual, 87 I pre-Hilbert A -module, 32 I pre-Hilbert space, 4 VIII Principle of Uniform Boundedness, 35 II product in C∗cpsu, 34 VI in W∗miu and W∗cpsu, 47 IV in C∗miu and cC∗miu, 20a I in C∗pu, 20a I product functional, 108 II projection in a C∗-algebra, 55 II of x on C, 5 II onto a corner, 94 I Projection Theorem, 5 VII pseudoinverse, 79 I approximate, 80 II pu-map, 10 III pure map, 100 I is rigid, 102 IX Pythagoras' theorem, 4 XV Q quantum lambda calculus, 120 I quotient -- comprehension chain, 97 I R RA , 125c II radially open set, 73 II radius of convergence, 13 II real part maximal, 27 VII Riesz' Representation Theorem, 5 IX rigid ncp-map, 102 II Russo -- Dye's Theorem, 34a VII S R , integral 14 II R f , of continuous f : [0, 1] → A , of continuous f : C → A RT f , over a triangle, 14 III R w′ w f , over an interval, 14 III SA , 14 II SA ,B, 125e VI Schur's Product Theorem, 111 I, 113 II self adjoint, 7 II ⋄-self-adjoint, 103 I separating collection of maps on a C∗-algebra, 21 II sequential product, 106 I ΣX , measurable subsets, 51 I SOT, strong operator topology, 37 V sp, spectrum sp(A ), of a C∗-algebra, 27 III is extremally disconnected for a von Neumann algebra, 53 III sp(a), of an element of a C∗-algebra, 11 XIX spacial tensor product, 111 VII Spectral Mapping Theorem, 28 II Spectral Permanence, 11 XXIII spectral radius, 16 II square root axiom, 105 VII state of a C∗-algebra, 21 III order separating, 22 VIII of an element of a C∗-algebra, 7 II A , right unitor, 119 IVb Ω, 30 IX Stone -- Weierstrass' Theorem, 27 XIX subunital map between C∗-algebras, 10 II supremum norm, 3 V (· ) ⊗ B-surjective, 125e II symmetric monoidal category (SMC), for a Hilbert A -module, 32 XV for a Hilbert space, 21 III is normal, 38 II is completely positive, 34 V is normal, 49 II vector state = unital vector functional, 21 III von Neumann algebra, 42 I category of, 47 III commutative, 70 III finite dimensional, 84 II hereditarily atomic, 84b II is bounded ultraweakly complete, 77 I is ultrastrongly complete, 77 I with a faithful np-functional, 51 VII von Neumann subalgebra, 42 V is ultraweakly closed, 75 VIII W W ∗(S), von Neumann subalgebra gen- erated by S, 42 V W∗miu, W∗cpsu, . . . , 47 II wnT , winding number, 14 III WOT, weak operator topology, 37 V Z Z(A ), centre of A , 65 II 132 II T tensor product algebraic = of vector spaces, 112 I of Hilbert spaces, 109 II exists, 109 III is ℓ2-bounded, 110 III universal property, 110 III of von Neumann algebras, 108 II exists, 111 XII functorial, 115 IV uniqueness, 114 II universal property, 112 XI, 114 I Tomiyama's Theorem, 128 II triangle, for our purposes, 14 III U ultracyclic projection, 66 II ultraweak and ultrastrong, 42 III completeness, 77 I convex ∼ly closed subset, 73 VIII permanence, 89 XI topologies are Hausdorff, 44 XI ultraweak tensor product topology, 112 II ultraweakly bounded net, 87 VIII Uniform Boundedness Theorem, 35 II unit of a C∗-algebra, 3 I unit ball, 4 IV of a C∗-algebra extreme points, 86 VI unital bilinear map, 108 I C∗-algebra, 3 II map between C∗-algebras, 10 II unitary in a C∗-algebra, 34a IV unitor, 119 IVb V vector functional 211 Bibliography [1] Creative commons 4.0 public https://creativecommons.org/licenses/by/4.0/legalcode. international attribution license. [2] Erik M. Alfsen and Frederic W. Shultz. State spaces of operator algebras. 2001. doi:10.1007/978-1-4612-0147-2. [3] Bruce E. Blackadar. Operator algebras. 2006. doi:10.1007/3-540-28517-2. [4] Kenta Cho. Total and partial computation in categorical quantum foun- In QPL 2015, volume 195 of EPTCS, pages 116 -- 135, 2015. dations. doi:10.4204/EPTCS.195.9 arXiv:1511.01569v1. [5] Kenta Cho. Semantics for a quantum programming language by op- New Generation Computing, 34(1-2):25 -- 68, 2016. erator algebras. doi:10.1007/s00354-016-0204-3. [6] Kenta Cho and Bart P.F. Jacobs. The EfProb library for probabilistic calculations. In CALCO 2017, volume 72 of LIPIcs, pages 25:1 -- 25:8, 2017. doi:10.4230/LIPIcs.CALCO.2017.25. [7] Kenta Cho and Bart P.F. Jacobs. Disintegration and Bayesian inver- sion via string diagrams. MSCS, 2019. doi:10.1017/S0960129518000488 arXiv:1709.00322v3. [8] Kenta Cho, Bart P.F. Jacobs, Abraham A. Westerbaan, and Bas E. In QPL 2015, vol- Westerbaan. ume 195 of EPTCS, pages 136 -- 147, 2015. doi:10.4204/EPTCS.195.10 arXiv:1511.01570v1. Quotient -- comprehension chains. [9] Kenta Cho, Bart P.F. Jacobs, Bas E. Westerbaan, and Abraham A. West- erbaan. An introduction to effectus theory. arXiv:1512.05813v1, 2015. [10] Kenta Cho and Abraham A. Westerbaan. Duplicable von Neumann alge- bras. arXiv:1903.02963v1, 2016. [11] Kenta Cho and Abraham A. Westerbaan. Von Neumann algebras form a model for the quantum lambda calculus. arXiv:1603.02133v1, 2016. [12] Man-Duen Choi. A Schwarz inequality for positive linear maps on Illinois Journal of Mathematics, 18(4):565 -- 574, 1974. C*-algebras. doi:10.1215/ijm/1256051007. [13] Rob Clifton, Jeffrey Bub, and Hans Halvorson. Characterizing quan- tum theory in terms of information-theoretic constraints. Foundations of Physics, 33(11):1561 -- 1591, 2003. doi:10.1023/A:1026056716397. [14] Bob Coecke and Aleks R. Kissinger. Picturing quantum processes. 2017. doi:10.1017/9781316219317. [15] John B. Conway. A course in functional analysis. 1985. doi:10.1007/978-1-4757-3828-5. [16] John B. Conway. A course in operator theory. 2000. doi:10.1090/gsm/021. [17] Keith J. Devlin. The joy of sets. 1993. doi:10.1007/978-1-4612-0903-4. [18] Ronald G. Douglas. On majorization, factorization, and range inclusion of operators on Hilbert space. Proceedings of the American Mathematical Society, 17(2):413 -- 415, 1966. doi:10.1090/S0002-9939-1966-0203464-1. [19] David H. Fremlin. Measure theory. Torres Fremlin, 2000. https://www1.essex.ac.uk/maths/people/fremlin/mt.htm. [20] Robert W.J. Furber. Continuous dcpos in quantum computing. http://people.cs.aau.dk/~furber/papers/contawconf.pdf, 2019. [21] Robert W.J. Furber and Bart P.F. Jacobs. From Kleisli cat- probabilistic Gelfand duality. volume 8089 of LNCS, pages 141 -- 157, 2013. to commutative C∗-algebras: egories In CALCO 2013, doi:10.1007/978-3-642-40206-7 12. [22] L. Terrell Gardner. Linear maps of C∗-algebras preserving the absolute value. Proceedings of the American Mathematical Society, 76(2):271 -- 278, 1979. doi:10.1090/S0002-9939-1979-0537087-0. [23] Irving L. Glicksberg. Stone -- Cech compactifications of products. Trans- the American Mathematical Society, 90(3):369 -- 382, 1959. actions of doi:10.1090/S0002-9947-1959-0105667-4. [24] Stefano Gogioso and Fabrizio Genovese. Infinite-dimensional categorical quantum mechanics. In QPL 2016, volume 236 of EPTCS, pages 51 -- 69, 2017. doi:10.4204/EPTCS.236.4 arXiv:1605.04305v2. [25] Stanley P. Gudder and Richard J. Greechie. Sequential products on effect algebras. Reports on Mathematical Physics, 49(1):87 -- 111, 2002. doi:10.1016/S0034-4877(02)80007-6. [26] Stanley P. Gudder and Fr´ed´eric Latr´emoli`ere. Characterization of the se- quential product on quantum effects. Journal of Mathematical Physics, 49(5):052106, 2008. doi:10.1063/1.2904475. 213 [27] Ichiro Hasuo and Naohiko Hoshino. Semantics of higher-order quantum computation via geometry of interaction. Annals of Pure and Applied Logic, 168(2):404 -- 469, 2017. doi:10.1016/j.apal.2016.10.010. [28] Chris J.M. Heunen and Bert J. Lindenhovius. Domains of com- In LICS 2015, pages 450 -- 461, 2015. mutative C∗-subalgebras. doi:10.1109/LICS.2015.49. [29] Bart P.F. Jacobs. On block structures in quantum computation. ENTCS, 298:233 -- 255, 2013. doi:10.1016/j.entcs.2013.09.016. [30] Bart P.F. Jacobs. probabilistic sical, doi:10.2168/LMCS-11(3:24)2015 arXiv:1205.3940v5. and quantum logic. New directions in categorical logic, LMCS, 11(3), for clas- 2015. [31] Bart P.F. Jacobs. Affine monads and side-effect-freeness. In CMCS 2016, volume 9608 of LNCS, pages 53 -- 72, 2016. doi:10.1007/978-3-319-40370-0 5. [32] Bart P.F. Jacobs. Effectuses from monads. ENTCS, 325:169 -- 183, 2016. doi:10.1016/j.entcs.2016.09.037. [33] Bart P.F. Jacobs. Hyper normalisation and conditioning for discrete proba- bility distributions. LMCS, 13(3), 2017. doi:10.23638/LMCS-13(3:17)2017 arXiv:1607.02790v3. [34] Bart P.F. Jacobs. Quantum effect logic in cognition. Journal of Mathemat- ical Psychology, 81:1 -- 10, 2017. doi:10.1016/j.jmp.2017.08.004. [35] Bart P.F. Jacobs. A recipe for state-and-effect triangles. LMCS, 13(2), 2017. doi:10.23638/LMCS-13(2:6)2017 arXiv:1703.09034v3. [36] Bart P.F. Jacobs. A channel-based perspective on conjugate priors. arXiv:1707.00269v2, 2018. [37] Bart P.F. Jacobs. From probability monads to commutative effectuses. Journal of Logical and Algebraic Methods in Programming, 94:200 -- 237, 2018. doi:10.1016/j.jlamp.2016.11.006. [38] Bart P.F. Jacobs and Jorik Mandemaker. Relating operator spaces via ad- junctions. Logic and Algebraic Structures in Quantum Computing, 45:123 -- 150, 2016. doi:10.1017/CBO9781139519687.008 arXiv:1201.1272v2. [39] Bart P.F. Jacobs, Jorik Mandemaker, and Robert W.J. Furber. The ex- pectation monad in quantum foundations. Information and Computation, 250:87 -- 114, 2016. doi:10.1016/j.ic.2016.02.009. [40] Bart P.F. Jacobs An effect- theoretic account of lebesgue integration. ENTCS, 319:239 -- 253, 2015. doi:10.1016/j.entcs.2015.12.015. and Abraham A. Westerbaan. [41] Bart P.F. Jacobs and Abraham A. Westerbaan. Distances between states and between predicates. arXiv:1711.09740v2, 2018. [42] Bart P.F. Jacobs, Bas E. Westerbaan, and Abraham A. Westerbaan. States In FoSSaCS 2015, volume 9034 of LNCS, pages 87 -- 101, of convex sets. 2015. doi:10.1007/978-3-662-46678-0 6. [43] Bart P.F. Jacobs and Fabio Zanasi. trans- former semantics for Bayesian learning. ENTCS, 325:185 -- 200, 2016. doi:10.1016/j.entcs.2016.09.038. A predicate/state [44] Bart P.F. Jacobs and Fabio Zanasi. A formal semantics of influence In MFCS 2017, volume 83 of LIPIcs, 2017. in Bayesian reasoning. doi:10.4230/LIPIcs.MFCS.2017.21. [45] Richard V. Kadison. A representation theory for commutative topological algebra. Number 7 in Memoirs of the American Mathematical Society. 1951. doi:10.1090/memo/0007. [46] Richard V. Kadison. Operator algebras with a faithful weakly- closed representation. Annals of Mathematics, 64(1):175 -- 181, 1956. doi:10.2307/1969954. [47] Richard V. Kadison and John R. Ringrose. Fundamentals of the The- doi:10.1090/gsm/015 & I & II. 1983. ory of Operator Algebras, doi:10.1090/gsm/016. [48] Andre Kornell. Quantum collections. Mathematics, 28(12):1750085, 2017. arXiv:1202.2994v2. International Journal of doi:10.1142/S0129167X17500859 [49] Andre Kornell. Quantum sets. arXiv:1804.00581v5, 2018. [50] Hans D.M. Maassen. Quantum probability and quantum informa- tion theory. In Quantum Information, Computation and Cryptogra- phy, volume 808 of Lecture Notes in Physics, pages 65 -- 108, 2010. doi:10.1007/978-3-642-11914-9 3. [51] Saunders Mac Lane. Categories for the working mathematician. 1971. doi:10.1007/978-1-4757-4721-8. [52] Octavio Malherbe. Categorical models of computation: partially traced cate- gories and presheaf models of quantum computation. PhD thesis, University of Ottawa, 2010. arXiv:1301.5087v1. 215 [53] Jurgen Mayer, Khaled Khairy, and Jonathon Howard. Drawing an elephant with four complex parameters. American Journal of Physics, 78(6):648 -- 649, 2010. doi:10.1119/1.3254017. [54] Eliakim H. Moore. A simple proof of the fundamental Cauchy -- Goursat theorem. Transactions of the American Mathematical Society, 1(4):499 -- 506, 1900. doi:10.1090/S0002-9947-1900-1500551-3. [55] Francis J. Murray and John von Neumann. On rings of operators. Annals of Mathematics, 37(1):116 -- 229, 1936. doi:10.2307/1968693. [56] William K. Nicholson. A short proof of the Wedderburn -- Artin theorem. New Zealand Journal of Mathematics, 22(1):83 -- 86, 1993. [57] Michael A. Nielsen and Isaac L. Chuang. Quantum computation and quan- tum information, 2010. doi:10.1017/CBO9780511976667. [58] Michele Pagani, Peter Selinger, and Benoıt Valiron. Applying quan- titative semantics to higher-order quantum computing. In POPL 2014, volume 49 of ACM SIGPLAN Notices, pages 647 -- 658, 2014. doi:10.1145/2535838.2535879 arXiv:1311.2290v1. [59] Robert A. Palais. π is wrong! The mathematical intelligencer, 23(3):7 -- 8, 2001. doi:10.1007/BF03026846. [60] William L. Paschke. Inner product modules over B*-algebras. Trans- the American Mathematical Society, 182:443 -- 468, 1973. actions of doi:10.1090/S0002-9947-1973-0355613-0. [61] Gert K. Pedersen. Pro- the American Mathematical Society, 36(1):309 -- 310, 1972. Some operator monotone functions. ceedings of doi:10.1090/S0002-9939-1972-0306957-4. [62] Plutarch. Live of Theseus. Parallel lives. [63] Mathys P.A. Rennela. Operator algebras in quantum computation. Mas- ter's thesis, Radboud University, 2015. arXiv:1510.06649v1. [64] Mathys P.A. Rennela and Sam Staton. Complete positivity and natu- ral representation of quantum computations. ENTCS, 319:369 -- 385, 2015. doi:10.1016/j.entcs.2015.12.022. [65] Mathys P.A. Rennela, Sam Staton, and Robert W.J. Furber. Infinite- dimensionality in quantum foundations: W*-algebras as presheaves over matrix algebras. In QPL 2016, volume 236 of EPTCS, 2017. doi:10.4204/EPTCS.236.11 arXiv:1701.00662v1. [66] Walter Rudin. Principles of Mathematical Analysis. McGraw -- Hill, 1976. [67] Bernard Russo and H. A. Dye. A note on unitary operators Duke Mathematical Journal, 33(2):413 -- 416, 1966. in C∗-algebras. doi:10.1215/S0012-7094-66-03346-1. [68] Shoichiro Sakai. C∗-algebras and W*-algebras. 1998. doi:10.1007/978-3-642-61993-9. [69] Irving E. Segal. Equivalences of measure spaces. American Journal of Mathematics, 73(2):275 -- 313, 1951. doi:10.2307/2372178. [70] Peter Selinger. Towards a semantics for higher-order quantum computation. In QPL 2004, volume 33 of TUCS General Publication, pages 127 -- 143. Turku Centre for Computer Science, 2004. [71] Peter Selinger and Benoıt Valiron. A lambda calculus for quantum compu- tation with classical control. In TLCA 2005, volume 3461 of LNCS, pages 354 -- 368, 2005. doi:10.1007/11417170 26. [72] Peter Selinger and Benoıt Valiron. A lambda calculus for quantum compu- tation with classical control. Mathematical Structures in Computer Science, 16(3):527 -- 552, 2006. doi:10.1017/S0960129506005238. [73] Peter Selinger and Benoıt Valiron. Quantum lambda calculus. https://www.mscs.dal.ca/∼selinger/papers/qlambdabook.pdf, >2008. [74] Alan D. Sokal. A really simple elementary proof of the uniform bounded- ness theorem. The American Mathematical Monthly, 118(5):450 -- 452, 2011. doi:10.4169/amer.math.monthly.118.05.450. [75] Masamichi Takesaki. Theory of operator algebras I. 1979. doi:10.1007/978-1-4612-6188-9. [76] Jun Tomiyama. On the projection of norm one in W*-algebras. Proceedings of the Japan Academy, 33(10):608 -- 612, 1957. doi:10.3792/pja/1195524885. [77] Arnoud C.M. van Rooij. Riesz spaces. https://github.com/awesterb/riesz-spaces, 2011. [78] John von Neumann. Zur Algebra der Funktionaloperationen und Theorie der normalen Operatoren. Mathematische Annalen, 102:370 -- 427, 1930. doi:10.1007/BF01782352. [79] Abraham A. Westerbaan. a generalisation of measure and integral. Master's thesis, Radboud University, 2012. arXiv:1903.06044v1. Lattice valuations: 217 [80] Abraham A. Westerbaan. Quantum programs as kleisli maps. In QPL 2016, volume 236 of EPTCS, pages 215 -- 228, 2016. doi:10.4204/EPTCS.236.14 arXiv:1501.01020v4. [81] Abraham A. Westerbaan and Bas E. Westerbaan. A universal property for sequential measurement. Journal of Mathematical Physics, 57(9):092203, 2016. doi:10.1063/1.4961526 arXiv:1603.00410v1. [82] Abraham A. Westerbaan and Bas E. Westerbaan. Paschke dilations. In QPL 2016, volume 236, pages 229 -- 244, 2017. doi:10.4204/EPTCS.236.15 arXiv:1603.04353v2. [83] Abraham A. Westerbaan, Bas E. Westerbaan, Rutger Kuyper, Carst Statman's hierar- doi:10.23638/LMCS-13(4:19)2017 Tankink, Remy Viehoff, and Henk P. Barendregt. chy theorem. arXiv:1711.05497v2. LMCS, 13(4), 2017. [84] Bas E. Westerbaan. Dagger and Dilations in the Category of Von Neu- mann Algebras. PhD thesis, Radboud University, 2019. arXiv:1803.01911 doi:2066/201785. [85] Eugene P. Wigner. John von neumann. In Historical and Biographical Reflections and Syntheses, volume VII of The Collected Works of Eugene Paul Wigner, pages 127 -- 130, 2001. doi:10.1007/978-3-662-07791-7 14. [86] Alexander Wilce. A royal road to quantum theory (or thereabouts). En- tropy, 20(4):227, 2018. doi:10.3390/e20040227. [87] Stephen Willard. General topology. Addison -- Wesley, 1970. Lekensamenvatting Wat is wiskunde? Volgens mij is ze niets anders dan een studie van de patronen in de wereld om ons heen in de meest strenge en formele zin. De wiskundige maakt zich bewust, bijna komisch, blind voor zaken die een ander z´o in het oog zouden springen. "Ja." antwoorden op de vraag "Wil je koffie of thee?" is voor een wiskundige volstrekt acceptabel. Maar door deze blindheid kan de wiskundige wel met een ongeevenaarde nauwkeurigheid, zekerheid en tijd- loosheid uitdrukking geven aan afzonderlijke aspecten van de werkelijkheid. Slechts een zeer beperkt aantal fenomenen is vatbaar voor wiskundige ana- lyse. Soms begrijpen we een fenomeen er nog niet goed genoeg voor. Soms is de benodigde wiskunde er nog niet voor ontwikkeld. Maar in de meeste gevallen is het zo dat het mes van de formele wiskunde simpelweg te scherp snijdt. Een wiskundige zou je niet kunnen vertellen of een schip hetzelfde blijft als je alle planken vervangt,† wat twee is w´el. Het is niet zozeer dat hij of zij zich niet bezig zou willen houden met dagelijkse begrippen, maar eerder dat deze be- grippen daartegen niet bestand zijn. Daarom gaat de wiskunde over abstracte en geıdealiseerde objecten zoals getallen, vierkanten, cirkels, distributies, ten- soren en varieteiten, in plaats van wereldse tegenhangers zoals respectievelijk logistiek, akkers, hemellichamen, toekomstverwachtingen, elektrische velden en sterrenstelsels. Het is wonderbaarlijk hoe behulpzaam wiskunde kan zijn ondanks en juist dankzij haar beperkingen. Niemand -- ook geen wiskundige -- kan zich het stilleven op de omslag van dit boekje vanuit een andere hoek nauwkeurig voorstellen. Maar als ik vertel dat het een plaatje is van een reflecterende bol in R12 (tussen twee afgeknotte hypervlakken, met een deels reflecterend schaakbordpatroon) en wanneer ik de relevante coordinaten en afmetingen geef, dan kan ieder die be- kend is met 'inproducten' (vgl. 4 VIII) een computerprogramma schrijven dat dit tafereel vanuit een andere hoek toont. Dankzij de wiskunde is bovenmenselijk inzicht in de twaalfde dimensie niet nodig! Dat sommige wiskundige begrippen zeer bruikbaar zijn, wil niet zeggen dat elk wiskundig begrip dit is. Integendeel zelfs: van alle denkbare figuren op pa- †Het schip van Theseus, zie §23 van [62]. 219 pier grijpt men ongetwijfeld het vaakst terug naar de cirkel en de lijn. Bepaalde constanten (2π, e, ϕ, √2, . . . ) verschijnen ook vaker in formules dan andere. Natuurlijk kan een nuttig wiskundig begrip ook slecht gebruikt worden. Men kan bijvoorbeeld de baan van een planeet met een groot aantal cirkels (Ptole- maeus' 'epicykels') omschrijven in plaats van met ´e´en ellips. Bij een dergelijk waardeoordeel als 'goed' of 'slecht' gebruik van een begrip is er zeker sprake van persoonlijke voorkeur, conventie en willekeur. Waarom gebruiken we niet 2π = 6,2831 . . . als constante in plaats van π = 3,1415 . . . ? Volgens mij zou dat veel formules eleganter maken; zie [59]. Dat een wiskundig begrip (zoals epicykels) in ongebruik geraakt is, wil bovendien niet zeggen dat het geen stijl- vol herintreden kan maken: de befaamde Fourier-reeks (die in feite bestaat uit oneindig veel epicykels) geeft de ellips weer het nakijken!‡ Dit proefschrift gaat over zo'n bruikbaar, abstract en geıdealiseerd wiskundig begrip: de von Neumann-algebra, bedacht door en vernoemd naar het Hongaarse genie en de alleskunner John von Neumann (geboren als Neumann J´anos Lajos). Aan hem hebben we niet alleen de moderne computer-architectuur te danken, maar bijvoorbeeld ook de springstoflenzen die in kernwapens gebruikt worden, de afschrikwekkende strategie van mutual assured destruction en de numerieke weersvoorspelling. Het is niet eenvoudig om uit te leggen wat von Neumann-algebra's pre- cies zijn (42), maar ik kan wel een indruk geven waarvoor ik ze gebruik. Uit- gangspunt van dit proefschrift is dat de elementaire systemen die je tegenkomt bij het ontwerpen van een algoritme voor een kwantumcomputer omschreven kunnen worden door von Neumann-algebra's. Zo wordt de klassieke bit (die 0 of 1 kan zijn) voorgesteld door de von Neumann-algebra "C2" en wordt de kwantumbit (die een complexe combinatie van 0 of 1 is totdat je haar meet) voorgesteld door de von Neumann-algebra "M2". Dit geeft de hoop dat in- gewikkeldere samengestelde systemen ook een bijpassende von Neumann-algebra hebben. Dat hangt natuurlijk af van de wijze van samenstelling: het systeem dat bestaat uit twee kwantumbits en ´e´en bit wordt voorgesteld met de von Neumann-algebra M2⊗ M2⊗ C2, terwijl het systeem dat een kwantumbit of een klassieke bit bevat omschreven wordt met de von Neumann-algebra M2 ⊕ C2. De bewerking "⊗" op von Neumann-algebra's is de interpretatie voor de "en"- samenstelling van systemen, terwijl "⊕" invulling geeft aan "of"-samenstelling van systemen. Een veel complexere samenstelling van twee systemen A en B bestaat uit alle 'processen' van A naar B. Meting is bijvoorbeeld zo'n proces van een kwantumbit naar een klassieke bit. Zulke processen worden in de wereld van von Neumann-algebra's voorgesteld door zogenaamde "ncpsu-afbeeldingen" (10 II) tussen von Neumann-algebra's. E´en van de hoofdresultaten van dit proefschrift is een interpretatie voor deze 'processen'-samenstelling, de bewerking ⊸. Door bepaalde formele kaders was ‡Of toch niet, [53]? het al van tevoren duidelijk dat er hoogstens ´e´en interpretatie mogelijk zou zijn; de vraag was alleen: welke? Sterker nog: is er uberhaupt eentje? Dit is vergelijkbaar met de vraag wat de kleinste§ grammaticaal correcte tekst is waarin alle Nederlandse woorden voorkomen. Om dit probleem op te lossen kun je niet zomaar alle woorden op een rij zetten -- de onderlinge samenhang moet immers kloppen. De crux was voor mij om niet te willen proberen om de bewerking ⊸ direct te omschrijven, maar om het bestaan ervan indirect aan te tonen, zoals je ook kunt laten zien dat de bovengenoemde Nederlandse tekst uit het voorbeeld bestaat, zonder te weten hoe hij precies is samengesteld. (Voor een beperkte klasse van von Neumann-algebra's, de 'hereditair atomische', bleek een directe omschrijving trouwens wel mogelijk, 125e VII.) Het tweede hoofdresultaat van dit proefschrift is de vondst van een abstracte omschrijving van de processen die bij een meting horen (105 VII). Het kenmer- kende aspect van deze omschrijving is dat het alleen gebruik maakt van zoge- naamd 'categorisch' jargon. Dat dit mogelijk is toont aan dat we in de categorie van von Neumann-algebra's niet op een te abstract niveau werken: we kunnen het nog steeds over meting hebben. Dankzij de categorische omschrijving wordt het bovendien mogelijk 'meting' in andere contexten te interpreteren. Bij de zoektocht naar deze categorische omschrijving was lange tijd het probleem om een manier te vinden om onderscheid te maken tussen de processen¶ √p(· )√p en √pu∗(· )u√p. De linker hoort bij een meting, de rechter niet. De oplossing was om een begrip uit de theorie van Hilbert-ruimten -- geadjungeerdeerdheid -- in een afgezwakte vorm over te nemen, namelijk ⋄-geadjungeerdheid (of contraposedness, 101 VI). Het linker proces blijkt een kwadraat te zijn van een aan zichzelf ⋄-geadjungeerd proces; de rechter niet. Beide hoofdresultaten staan in het laatste hoofdstuk. De rest van dit proefschrift bestaat -- enigszins ongebruikelijk -- uit een grondige introductie tot de benodigde, reeds bestaande, theorie van C∗-algebra's en van von Neumann-algebra's. Niet alleen bestond een geschikte introductie nog niet, maar het leek me ook een goede kans om me verder te verdiepen in de theorie van von Neumann-algebra's. In de eerste hoofdstukken ontwikkel ik de gehele benodigde theorie, inclusief bewijzen. Het is mijn bedoeling dat eenieder die een bachelorgraad in de wiskunde heeft behaald deze tekst zou moeten kunnen be- grijpen. Op een groot aantal plekken wijk ik af van het begane pad: soms om de tekst kort te houden, (zo ontwijk ik de theorie van Banach algebra's volkomen), maar meestal om te experimenteren met variaties. Zo gebruik ik Kadison's om- schrijving van von Neumann-algebra's om de theorie op te bouwen (wat niet eerder gedaan is.) Zo houdt men goede wiskundige begrippen levend. §Kleinste in de woordenboekordening. ¶Hier is u 6= 1 een unitaire en p een positief element van een von Neumann-algebra A met ⌈p⌉ = 1. 221 About the Author Bram Westerbaan, born August 30, 1988, enrolled as a physics and astronomy student at the Radboud University in 2006. He obtained a bachelor's degree in mathematics in 2012 (cum laude) with a thesis on the simply typed λ-calculus supervised by prof. dr. H.P. Barendregt (resulting in a publication [83].) The same year, he obtained a master's degree in mathematics (summa cum laude) with a thesis [79] on measure and integral under supervision of prof. dr. A.C.M. van Rooij. His doctoral studies started in 2013 on the topic of (co)algebra guided by dr. A. Silva, and shifted to the material presented in this thesis in 2014 under the auspices of prof. dr. B.P.F. Jacobs. From 2018 onward Bram works as a postdoctoral researcher at the digital security department on an NWA project applying polymorphic pseudonymisation to network traffic flow data. 223
1104.2626
1
1104
2011-04-13T21:30:48
Completely Bounded Characterization of Operator Algebras with Involution
[ "math.OA" ]
In this paper we study the completely bounded anti-isomorphisms on operator algebras, that work similarly to the involutions with the exception for the property of being completely isometric. We elaborate the Blecher's characterization theorem for operator algebras to make it applicable to the so-called operator $K$-algebras with completely bounded reflexive anti-isomorphism. We also establish a connection of this result with the notion of smooth $C^*$-modules, that play an important role in Mesland's approach to Baaj-Julg picture of $KK$-theory.
math.OA
math
Completely Bounded Characterization of Operator Algebras with Involution Nikolay P. Ivankov Max-Planck-Institut für Mathematik, Bonn November 13, 2018 Abstract In this paper we study the completely bounded anti-isomorphisms on operator alge- bras, that work similarly to the involutions with the exception for the property of being completely isometric. We elaborate the Blecher's characterization theorem for opera- tor algebras to make it applicable to the so-called operator K-algebras with completely bounded reflexive anti-isomorphism. We also establish a connection of this result with the notion of smooth C∗-modules, that play an important role in Mesland's approach to Baaj-Julg picture of KK-theory. 1 Introduction This article is supposed to support the theory developed in [6]. In [6], in turn, the author tries to enlarge a theoretical basis for the generalization of Kasparov product to the Baaj- Julg picture of KK-theory, which is being developed by Bram Mesland in [7]. The main idea of [7] is that under certain conditions called as transversality the Kas- parov product in KK-theory may be replaced by a simple formula involving the so-called unbounded KK-cycles inroduced by Baaj and Julg. However, in [7], if A, B, C are C∗- algebras and (E1, T) and (E2, S) are (A, B) and (B, C) unbounded KK-cycles respectively, then the transversality condition is given in terms of the two concrete unbounded opera- tors S and T. In [6] and the author's upcoming thesis we have proposed the way to justify the transversality of the operator T with respect to some class of unbounded (B, C)-KK- cycles. For that we have introduced the notion of abstract systems of smooth subalgebas of the C∗-algebra B (which differs from the approach of [7] where smooth systems are constructed by means of the unbounded operator S), and the Ck-algebras in these smooth systems are operator algebras that are either supposed to be given or are constructed in some way. We consider some simplest examples in the end of the article, a more explicit information may be found in [6] and [7]. In the case when the algebras are constructed we often encounter the point where one has an operator space which is an algebra, but not operator algebra. Moreover, the 1 construction of smooth algebras in [7] often uses the fact that the involution on the C∗- algebra A induces a completely isometric anti-isomorphism on its Ck subalgebras, and, again, this property may be not automatically fulfilled by the construction. Therefore we found a need for a result that would characterize the objects which can be completely boundedly isomorphic to operator algebras with a completely isometric involution. In this article we develop such a characterization. It is based on now classical Blecher's characterisation theorem for operator algebras [3], and incorporates an additional invo- lution structure. The result we present here is purely operator algebraic, and may prove itself to be useful in other fields concerning the operator algebra theory. 2 Preliminaries We recall the basic definitions from the theory of operator spaces. Definition 2.1. A (concrete) operator space is a linear subspace of B(H) for some Hilbert space H. A (concrete) operator algebra is a subalgebra of B(H). The map f : X → Y between two operator spaces is called completely bounded (cb-map) if there exists a positive constant C such that the natural extensons of f fn : Mn(X) → Mn(Y) 7→ ( f (xij)) (xij) has the norm less then C for all n ∈ N. The number k f kcb := sup n k fnk is called the cb-norm of f . The cb-map f is called a cb-isomorphism if it is an isomorphism and its inverse f −1 is also completely bounded. If f is a cb-isomorphism, such that k f kcb = k f −1kcb = 1, then f is called complete isometry. In what follows we require the operator spaces to be complete with respect to the operator norms on them. Remark 2.2. One may, of course, have an isomorphism f : X → Y of operator spaces, which is completely bounded, but its inverse f −1 is not. Therefore sometimes the term completely bicontinuous map is used instead of cb-isomorphism. Here, however, we are going to use the term cb-isomorphism only in the sense of the Definition 2.1, therefore avoiding the ambiguity. Definition 2.3. An operator space A which is also an algebra, such that the multiplica- tion map m : A × A → A is a completely bounded bilinear map with kmkcb ≤ K will be called an operator K-algebra (cf. [4]). We will use the term operator pseudoalgebra when the number K is not specified. The cb-homomorphism, cb-iomorphism and completely isometric isomorphism of between two operator pseudoalgebras are then algebra homo- morphism (isomorphism) which is completely bounded (isometric) as a map between operator spaces. 2 Remark 2.4. The notion of operator pseudoalgebra employed in this paper differs from the one given in [8]. In the notation of [4] these algebras would rather be called operator 1-algebras. There are two important results that give a characterization of operator spaces and operator K-algebras respectively. The first one is due to Effros and Ruan. Theorem 2.5 ([5]). Let X be a linear space with a set of matrix norms nk · k on Mn(X), satisfying the properties • n+mkx ⊕ yk = max{ nkxk, mkyk} • nkαxβk ≤ kαk nkxkkβk for all x ∈ Mn(X), y ∈ Mm(X) and α, β ∈ Mn(C). Then X is completely isometrically isomor- phic to a concrete operator space. Thus, we have a characterization of operator spaces up to a complete isometry. Another theorem is due to Blecher, and it would be the main point of our attention throughout the paper. Theorem 2.6 ([3],[4]). Let A be an operator K-algebra. Then there exists a (concrete) operator algebra A′, which is cb-isomorphic to A. Moreover, it may be chosen in such a way that if f : A → A′ is a cb-isomorphism, then max{k f kcb, k f −1kcb} ≤ max{K−1, 2K}. Remark 2.7. Obviously, all the concrete operator algebras are operator 1-algebras. The converse in general is not true. To have an operator 1-algebra being completely isomet- rically isomorphic to a concrete operator algebra, one has to add an assumption that A possesses a contractive approximate unit (cf. [8]) 3 Involution Recall that an involution on a Banach algebra A is an isometric anti-isomorphism ∗ : A → A, ∗ : a 7→ a∗ such that a∗∗ = a. Thus, if we want to specialize this notion for the case of operator algebras, we should first give a definition of a cb-anti-isomorphism. Definition 3.1. Let A be an operator pseudoalgebra. Then an anti-homomorphism f : A → B will be called cb-anti-homomorphism if there exists a positive number C such that nk( f (aji))ij ≤ C nk(aij)ijk is anti-isomorphic, and its inverse f −1 is also a cb-anti- for all (aij)ij ∈ Mn(A). homomorphism, then f will be called a cb-anti-isomorphism. One may analogously define completely isometric anti-isomorphisms. If f 3 Remark 3.2. Observe that, unlike the case of homomorphisms, we have to add a trans- position in matrix algebras to the definition of cb-anti-homomorphisms. This makes the notion of cb-anti-homomorphism much more subtle then the one of cb-homomorphism. It seems, although the author doesn't have a concrete example for now, that even for a general (concrete) operator algebra A there would not be any cb-anti-isomorphisms of A onto itself. However, as we have indicated in the introduction and will also see in the next section, the algebras having cb-anti-isomorphisms may often appear in applications. Definition 3.3. A cb-anti-isomorphism f : A → A such that f 2 = IdA would be called an (operator algebra) pseudo-involution on A. If, in addition, f is completely isometric then it will be called an (operator algebra) involution. An operator algebra possessing an involution will be called involutive. We are going to show that any pseudo-involution may in some sense be "updated" to become an involution. Proposition 3.4. Let A be an operator K-algebra with a pseudo-involution f . Then there is an operator pseudoalgebra B and a cb-isomorphism σ : A → B, such that σ f σ−1 is an involution on B. Proof. Let B = A as a algebras We define matrix norms on B as nk(aij)ijkB = max{ nk(aij)ijkA, nk( f (aji)ij)kA} The space B endowed with this norms is an operator pseudoalgebra. Indeed, we have that and n+mk(aij ⊕ bkl)kB = = max{max{ nk(aij)ijkA, nk( f (aji))ijkA}, max{ mk(bkl)lkkA, mk( f (bkl))lkkA}} = max{max{ nk(aij)ijkA, mk(bkl)lkkA}, max{ nk( f (aji))ijkA, mk( f (bkl))lkkA}} = max{ n+mk(aij)ij ⊕ (bkl)lkkA, n+mk( f (aji))ij ⊕ ( f (bkl))lkkA} = max{ nk(aij)kB, mk(bkl)kB} nkα(aij)βkB = max{ nkα(aij)ijβkA, nkβ⊺( f (aji))ijα⊺kA} ≤ max{kαk nk(aij)ijkAkβk, kβ⊺k nk( f (aji))ijkAkα⊺k} = kαkkβk max{ nk(aij)ijkA, nk( f (aji))ijkA} here we use the fact that α and β are scalar matrices. Thus, B is an operator space. To 4 prove that it is a pseudoalgebra, observe that nk(aij)(bkl)kB = max{ nk(aij)(bkl)kA, nk fn((aji)(bkl))kA} ≤ max{ nk(aij)(bkl)kA, k f kcb nk(aij)(bkl)kA} ≤ k f kcbK nk(aij)kA nk(bkl)kA ≤ k f kcbK · k f kcb max{ nk(aij)ijkA, nk( f (aji))ijkA}· · k f kcb max{ nk(bkl)klkA, nk( f (blk))klkA} = k f k3 cbK nk(aij)kB nk(bkl)kB Here we use the fact that since f 2 = 1 we have that k f kcb ≥ 1. Since f is a cb-anti-isomorphism and f 2 = 1, we have that k f k−1 cb nk · kA ≤ nk · kB ≤ k f kcb nk · kA so the algebras A and B are cb-isomorphic. Denote this isomorphism by σ. By the con- struction (σ f σ−1) = IdB. Now, for (aij) ∈ Mn(A) we have that nkσ f σ−1(σ(aij))kB = nkσ f (aij)kB = max{ nk f (aij)kA, nk f 2(aij)kA} = max{ nk f (aij)kA, nk(aij)kA} = nkσ(aij)kB Since σ is a cb-isomorphism, all the elements of Mn(B) have the form σ(aij). This last observation settles the proof. Remark 3.5. Observe that since f was an anti-isomorphism, we were not able to define σ as just σ : a 7→ a ⊕ f (a), since in this case σ(ab) = ab ⊕ f (ba). The result 3.4 gives us only an operator pseudoalgebra with involution. However, a closer look to the Theorem 2.6 lets us extend this result, making B into a (concrete) operator algebra with involution. In order to do this, we recall the construction from [3]. Let Γ be the set, n : Γ → N, γ 7→ nγ be a function. Let Λ be a set of formal symbols (variables) xγ ij, one variable for each γ ∈ Γ and each 1 ≤ i, j ≤ nγ. Denote by Φ a free associative algebra on Λ. In this case Φ consists of polynomials in the non-commuting variables with no constant term. One then defines a norm on Mn(Φ) by k(uij)kΛ := sup π (k(π(uij))k) (1) where π goes through all the representations of Φ on a separable Hilbert space satisfying the condition k(π(xγ ij))ijk ≤ 1 for all γ, where the latter matrix is indexed on rows by i and on columns by j for all 1 ≤ i, j ≤ nγ. 5 It is then shown in [3] that the map defined above indeed defines a norm on Mn(Φ) and that Φ becomes an operator algebra with respect to these operator norms. In the proof of the characterization theorem the set Γ is taken to be the collection of n × n matrices γ = (aij) with entries in A such that kγk = 1 2K , where K is a cb-norm of the multiplication in A. Then one takes Λ to be the collection of entries of these matrices xγ ij := aij, regarded as formal symbols indexed by γ and i, j, not identifying "equal" entries for different indexes. After that there is defined a map θ : Φ → A 7→ γij xγ ij which is then extended to general polynomials. It is proved in [3] that θ is a completely contractive homomorphism. One then puts B := Φ/ ker(θ), which is an operator algebra subject to the quotient operator norm, and is cb-isomorphic to A. Now let the pseudoalgebra A be involutive. Observe that since the involution on A is completely isometric, we have that nk(aij)∗k = nk(aij)k, and thus (aij)∗ ∈ Γ. Hence we have that a∗ ij ∈ Λ. This observation makes us able to define an involution the following way. On Φ we set (xγ1 i1j1 xγ2 i2j2 . . . xγk ikjk )∗ := (xγk ikjk )∗(xγk−1 ik−1jk−1 )∗ . . . (xγ1 i1j1 )∗ on the monomials, and then extend this to the whole Φ. Analogously, on Mn(Φ) we set ((Pij)ij)∗ = (P∗ ji)ij By the construction we have that θ((Pij)∗) = θ((P∗ ji)ij). Consequently, let π : Φ → B(H) be a representation of Φ satisfying the condition k(π(xγ ij)ij. Denote this set by Θ. We may define a representation π′ : Φ → B(H) by setting π′((Pij)∗) := (π(Pij))∗, where the latter involution is given by the one on the B(Hnγ). By the construction, we have that ij))k ≤ 1 for all (xγ nγkπ′(xγ ij)k = nγk(π((xγ ij)∗))∗k = nγkπ((xγ ij)∗)k ≤ 1 for all (xγ ij)ij since (xγ ij)∗ ∈ Γ, and so π′ ∈ Θ. Therefore we have that k(Pij)kΛ = sup π∈Θ = sup π′∈Θ = sup π′∈Θ (k(π(Pij))k) k((π′(Pij)∗))∗k kπ′(Pij)∗k = k(Pij)∗kΛ Hence we obtain that the map θ respects the involution, and thus the anti-isomorphism induced on B by the involution on Φ preserves the operator norms. Combining this observations with Proposition 3.4 we have the following 6 Theorem 3.6. Let A be an operator pseudoalgebra and let f be a pseudo-involution on A. Then there is a cb-isomorphism ρ : A → B, such that the map ρ f ρ−1 is an (operator algebra) involution on B. Proof. Put ρ = θσ. Remark 3.7. We may also estimate the cb-norm of ρ. Indeed, the map σ has the cb-norm k f kcb, and gives us a pseudoalgebra B′ with the cb-norm of multiplication bounded by k f k3 cbK. Thus, for K ≥ 1 the estimation from [3] shows us that the map θ has a cb-norm ≤ 2k f k3 cbK. Hence, kρkcb ≤ 2k f k4 cbK. 4 Application C1-Modules In this section we are going to show the relation of the construction of involutive operator algebras to the notion of smooth modules as they are defined in [7] and [6]. We will give here a simplified definition of smooth algebras and modules. For a more descriptive picture, see [7], [6]. Let A be a C∗-algebra E be a Hilbert C∗-module over A, and let A be an operator algebra, which is isomorphic to a pre-C∗-subalgebra of A abusively denoted by A. We define a C1 structure on E with respect to A by choosing a countable approximate unit un = ∑n j=1 xj ⊗ xj on KA(E) with a property that Then, a pre-C1-module over the C1 algebra A is defined as k(hxj, xki)jkk1,D ≤ C E = {e ∈ E hxj, ei ∈ A, k ∞ ∑ j=1 he, xjik1 < ∞} and it is a C1-module if it satisfies the Kasparov stabilization property. Now if the involution on A induces an operator algebra involution on A, then the space KA(E , A) is completely isometrically isomorphic to E . Thus, there is a well-defined inner product on E , which is a restriction of the inner product on E . The existence of this product then allows use to construct canonically the algebra CB∗ AE of completely bounded A-linear involutive of operators on E . Suppose now that there is another operator algebra A′ with the same properties as A, such that A ֒→ A′ as pre-C∗-algebras, and the inclusion map induces a completely bounded injective homomorphism of corresponding operator algebras. Then, by the con- struction, the smooth structure on E with respect to A will automatically be a smooth structure with respect to A′, and we obtain a completely bounded inclusion E → E ′, where E ′ is obtained form the approximate unit uk analogously to E . This observation also allows us to transfer additional structures which are involved in the construction of KK-product from E to E ′. 7 This construction may then become an intermediate step in the construction of KK- product in the Baaj-Julg picture of KK-theory. We briefly describe the simplest case. Let A, B be C∗-algebras and (E, D) be an unbounded (A, B)-KK-cycle on (see, [1], [2] for definition), and we suppose for a moment that D is selfadjoint. Denote A(1) D := {a ∈ A [D, a] extends to an element of CB∗ A(E)} Here we use the graded commutator. By the definition of an unbounded KK-cycle the algebra A(1) D is dense in A. We introduce a representation of A(1) D by setting D(a) = (cid:18) a π1 [D, a] 0 a(cid:19) with the operator norm k · k1,D defined by this representation. So, by definition A(1) D is a concrete operator algebra. The involution on A induces an operator algebra pseudoinvo- lution on A(1) D . Indeed, ka∗k1,D = (cid:13)(cid:13)(cid:13)(cid:13) = (cid:13)(cid:13)(cid:13)(cid:13) = (cid:13)(cid:13)(cid:13)(cid:13) = (cid:13)(cid:13)(cid:13)(cid:13) 0 0 (cid:18) a∗ a∗(cid:19)(cid:13)(cid:13)(cid:13)(cid:13) [D; a∗] a (cid:19)(cid:13)(cid:13)(cid:13)(cid:13) (cid:18)a ±[D; a] 0 (cid:19)(cid:18)a ±[D; a] (cid:18) 0 a(cid:19)(cid:13)(cid:13)(cid:13)(cid:13) (cid:18) a ±[D; a] IdE IdE 0 0 a (cid:19)(cid:18) 0 IdE IdE 0 (cid:19)(cid:13)(cid:13)(cid:13)(cid:13) where the sign is + when there the cycle (E, D) is even and − when it is odd. We will stick for now to the even case. Then the involution will obviously be isometric. To show that the involution is completely isometric, observe that for each m there is a "permutation" unitary Um, such that Um(π1 D(ajk))jkU−1 m = (cid:18) (ajk) [diagm(D); (ajk)] 0 (ajk)(cid:19) = π1 diagm(D)((ajk)) and so we can use the previous observation. In [6] we construct a kind of "universal" C1-subalgebra for a separable C∗-algebra A. More precisely for any set of C∗-algebras Ξ we construct an operator algebra A(1), such that for any C∗-algebra Bξ ∈ Ξ there is a set Ωξ of unbounded (A, B)-KK-cycles (Eξ,ω, Dξ,ω), for which the map Ωξ → KK0(A, Bξ) given by (Eξ,ω, Dξ,ω) 7→ [(Eξ,ω, Dξ,ω(1 + D∗ ξ,ωDξ,ω)− 1 2 )] 8 and there is a cb-inclusion A(1) ֒→ A(1) D which preserves the involution. The existence of such inclusion guarantees us that if an approximate unit uk defines a C1 structure on an Hilbert C∗ A-module E with respect to A(1), then so does it for all A(1) Dξ,ω . The idea of the construction is follows. We take a set of representatives (Eξ,ω, Fξ,ω) of the elements of KK0(A, Bξ), fix a total system {ai} on A and construct unbounded regular selfadjoint operators (Eξ,ω, Dξ,ω) such that [(Eξ,ω, Dξ,ω(1 + D∗ ξ,ωDξ,ω)− 1 2 )] = [(Eξ,ω, Fξ,ω)] We also define them in such a way that kaikDξ,ω ≤ Cj for all Dξ,ω for all the elements of the chosen total system on A. Then we define the algebra A(1) = {a ∈ A sup ξ,ω kak1,Dξ,ω < ∞} and with the collection of matrix norms on it defined as mk(akl)k1 := sup ξ,ω mk(akl)k1,Dξ,ω Since all the elements the total system {ai} lay in A(1), the algebra A(1) is dense in A. It is also shown in [6] that A(1) and A(1) are stable under holomorphic functional calculus Dξ,ω on A and therefore have the same K-theory as A. It is easy to check then that A(1) is then an operator 1-algebra with a completely isometric involution. We also have that there is a completely contractive inclusion A(1) ֒→ A(1) . However, in case when A is nonunital, A(1) may be not isomorphic to a concrete Dξ,ω operator algebra. Therefore, in order to make A(1) into an involutive operator algebra, we need to use Theorem 3.6. Observe that if we would like to incorporate the odd modules in the picture, the involution on the algebras A(1) Dξ,ω and we shall need to use the Theorem 3.6 to obtain involutive operator algebras. will not necessarily be completely isometric any more, Another example where the Theorem 3.6 may become useful arises when one con- siders almost selfadjoint unbounded operators instead of just selfadjoint ones. Let D be a selfadjoint regular operator on a Hilbert C∗-B-module E and suppose that b ∈ CB∗ B(E) and is even, but in general we do not demand the selfasjointness of b. We construct an operator algebra A(1) D . But now also in the case when we consider even unbounded KK-cycles the involution on A(1) D+b, although completely bounded, may be not isometric, since D+b analogously to A(1) ka∗k1,D+b = (cid:13)(cid:13)(cid:13)(cid:13) = (cid:13)(cid:13)(cid:13)(cid:13) = (cid:13)(cid:13)(cid:13)(cid:13) (cid:18) [D + b, a∗] (cid:18)a (cid:18) [D + b∗, a] 0 a a a∗ [D + b∗, a] 0 a∗(cid:19)(cid:13)(cid:13)(cid:13)(cid:13) (cid:19)(cid:13)(cid:13)(cid:13)(cid:13) a(cid:19)(cid:13)(cid:13)(cid:13)(cid:13) 0 9 and the latter norm should not in general be equal to kak1,D. The Proposition 3.4 then tells us that there is a (canonical) way to associate an involutive operator algebra to A(1) D+b and therefore simplify the consequent calculations. In the in the theory developed in [7] one may encounter other examples of operator algebras in which the involution should not necessarily be completely isometric, but only completely bounded. We have already mentioned one of the most important of them: these are the algebras of the form CB∗ A(E ) of completely bounded involutive A operators over C1-module E and their involutive subalgebras. The latter ones with operator norms induced by the norm on CB∗ A(E ) are used for the definition of subsequent C1-modules, which, in turn, are used for further construction of unbounded Kasparov product. Finally, it should be noted that in [6] and [7] there are considered higher orders of smoothness of algebras and modules, and the results we have presented in this paper may be also applicable to these cases. References [1] Saad Baaj and Pierre Julg, Théorie Bivariante de Kasparov et Opérateurs non Bornes dans les C∗-Modules Hilbertiens. C.R. Acad Sci. Paris, No. 296 (1983), Ser. I, pp. 875-878 [2] Bruice Blackadar. K-Theory for Operator Algebras. Springer-Verlag New York Inc., 1986 [3] David P. Blecher, A completely Bounded Characterisation of Operator ALgebras, Mathe- matische Annalen, No. 303 (1995), pp. 227-239. [4] David P. Blecher, Christian Le Merdy, Operator Algebras and Their Modules - An Oper- ator Space Approach, Oxford Univ. Press, 2004. [5] E. Effros, Zhong-Jin Ruan, On the Abstract Characterization of Operator Spaces, Proc. Amer. Math. Soc., No. 119 (1993), pp. 579-584 [6] N. P. Ivankov, Noncommutative Fréchet Spaces and Unbounded Bivariant K-Theory, arXiv:1103.4528v1 [math.KT] [7] Bram Mesland, Bivariant K-Theory of Groupoids and the Noncommutative Geometry of Limit Sets, arXiv:0904.4383v2 [math.KT] [8] Zhong-Jin Ruan, A characterization of nonunital operator algebras, Proc. of AMS, Vol. 121, No. 1 (May, 1994), pp. 193-198 10
1011.4154
1
1011
2010-11-18T08:57:22
Index maps in the K-theory of graph algebras
[ "math.OA", "math.KT" ]
Let $C^*(E)$ be the graph $C^*$-algebra associated to a graph E and let J be a gauge invariant ideal in $C^*(E)$. We compute the cyclic six-term exact sequence in $K$-theory of the associated extension in terms of the adjacency matrix associated to $E$. The ordered six-term exact sequence is a complete stable isomorphism invariant for several classes of graph $C^*$-algebras, for instance those containing a unique proper nontrivial ideal. Further, in many other cases, infinite collections of such sequences comprise complete invariants. Our results allow for explicit computation of the invariant, giving an exact sequence in terms of kernels and cokernels of matrices determined by the vertex matrix of $E$.
math.OA
math
INDEX MAPS IN THE K-THEORY OF GRAPH ALGEBRAS TOKE MEIER CARLSEN, SØREN EILERS, AND MARK TOMFORDE Abstract. Let C ∗(E) be the graph C ∗-algebra associated to a graph E and let J be a gauge-invariant ideal in C ∗(E). We compute the cyclic six-term exact sequence in K-theory associated to the extension 0 −→ J −→ C ∗(E) −→ C ∗(E)/J −→ 0 in terms of the adjacency matrix associated to E. The ordered six- term exact sequence is a complete stable isomorphism invariant for se- veral classes of graph C ∗-algebras, for instance those containing a unique proper nontrivial ideal. Further, in many other cases, finite collections of such sequences comprise complete invariants. Our results allow for explicit computation of the invariant, giving an exact sequence in terms of kernels and cokernels of matrices determined by the vertex matrix of E. The cyclic six-term exact sequence 1. Introduction (1.1) K0(J) ∂1 K1(cid:0)C ∗(E)/J(cid:1) ι∗ π∗ / K0(cid:0)C ∗(E)(cid:1) K1(cid:0)C ∗(E)(cid:1) π∗ / K0(cid:0)C ∗(E)/J(cid:1) ∂0 ι∗ K1(J) is a complete stable isomorphism invariant for a graph C ∗-algebra C ∗(E) of real rank zero containing a proper nontrivial ideal J when any of the following are satisfied • J is the unique proper nontrivial ideal of C ∗(E) ([7, Theorem 4.5]), • J is a smallest proper nontrivial ideal of C ∗(E), and C ∗(E)/J is AF ([6, Corollary 6.4]), • J is a largest proper nontrivial ideal of C ∗(E), and J is AF ([7, Theorem 4.7]). Date: October 25, 2018. 2000 Mathematics Subject Classification. 46L55. Key words and phrases. graph C ∗-algebras, classification, extensions, K-theory. This research was supported by the NordForsk Research Network "Operator Algebras and Dynamics" (grant #11580). The first named author was supported by the Research Council of Norway. The third author was supported by NSA Grant H98230-09-1-0036. 1 / /   O O o o o o 2 TOKE MEIER CARLSEN, SØREN EILERS, AND MARK TOMFORDE In other cases (cf. [6]) a complete invariant may be obtained by combining several six-term exact sequences associated to C ∗(E) and its ideals. It is therefore important to address how to compute sequences of the form in (1.1). In the existing literature it is shown that if E is a row-finite graph with no sinks, then K0(cid:0)C ∗(E)(cid:1) ∼= coker(At − I) and K1(cid:0)C ∗(E)(cid:1) ∼= ker(At − I), where At − I : ZE0 → ZE0 is the linear map given by the transpose of the vertex matrix A of E minus the identity matrix I. This description of the K0- group also includes a description of its order, and a similar computation ex- ists when sinks and infinite emitters are allowed. Since gauge-invariant ideals of graph C ∗-algebras and the corresponding quotients are naturally isomor- phic to graph C ∗-algebras, this allows one to compute the K0-groups and K1-groups in the above exact sequence. Moreover, since the C ∗-algebra of a graph satisfying Condition (K) has real rank zero [9, Theorem 3.5], it follows from [3] that the descending connecting map ∂0 : K0(C ∗(E)/J) → K1(J) is the zero map. All that remains is to describe a method for computing the other connecting group homomorphisms. The purpose of this paper is to provide explicit formulae for comput- ing the six-term exact sequence, the main challenge being to compute the connecting map ∂1 : K1(C ∗(E)/J) → K0(J). We shall also show that ∂0 : K0(C ∗(E)/J) → K1(J) is the zero map regardless of whether the graph E satisfies Condition (K) or not. All our calculations hold for an arbitrary graph algebra C ∗(E) and an arbitrary gauge-invariant ideal J in C ∗(E), even in the case of so-called breaking vertices. To compute ∂1, we need to choose generators for the K-groups involved. There is a canonical (and well-known) way to do this in K0; one can choose an isomorphism of K0(C ∗(E)) with coker(At − I) taking [pv] to ev + Im(At − I), where ev is the vector with a 1 in the vth position and zeroes elsewhere. However, for the K1-group the calculation is substantially harder. Descrip- tions of K1 can be found in [2] and [5], but we need a more explicit descrip- tion and therefore choose a different approach, choosing explicit generators for K1 based on a slightly intricate indexing of the entries in a matrix over C ∗(E). Although any quotient of a graph C ∗-algebra by a gauge-invariant ideal is isomorphic to a graph C ∗-algebra, it will be more convenient for us to use that such a quotient is isomorphic to a relative graph C ∗-algebra (cf. [11]), and we will therefore find generators of K0 and K1, not just for graph C ∗-algebras, but for relative graph C ∗-algebras. We prove that the generators we choose for K1 are indeed generators by computing the index map of the canonical Toeplitz extension of C ∗(E), using methods developed by Katsura in that framework. Our approach involves computing the index map using the canonical method (cf. [14]) of lifting the generating unitaries to partial isometries and computing defects. This method has similarities with the approach for Cuntz-Krieger algebras outlined by Cuntz himself in [4], and discussed with a few more details INDEX MAPS IN THE K-THEORY OF GRAPH ALGEBRAS 3 in [13]. After describing how to choose generators for K0 and K1 of any relative graph C ∗-algebra, we determine the index map ∂1 : K1(C ∗(E)/J) → K0(J) by, in a new extension, again lifting our generating unitaries to partial isometries, and computing defects. In Section 2 we briefly introduce graph C ∗-algebras, relative graph C ∗- algebras, and gauge-invariant ideals of graph C ∗-algebras. In Section 3 we find generators of K0 and K1 of any relative graph C ∗-algebra. Section 4 states the main result of the paper, allowing the computation of the index map ∂1 : K1(C ∗(E)/J) → K0(J) and the other maps in the six-term exact sequence (1.1), and this result is proved in Section 5. 2. Preliminaries A (directed) graph E = (E0, E1, r, s) consists of a countable set E0 of vertices, a countable set E1 of edges, and maps r, s : E1 → E0 identifying the range and source of each edge. A vertex v ∈ E0 is called a sink if s−1(v) = 0, and v is called an infinite emitter if s−1(v) = ∞. A graph E is said to be row-finite if it has no infinite emitters. If v is either a sink or an infinite emitter, then we call v a singular vertex. We write E0 sing for the set of singular vertices. Vertices that are not singular vertices are called regular vertices and we write E0 reg for the set of regular vertices. If E is a graph, a Cuntz-Krieger E-family is a set of mutually orthogonal projections {pv : v ∈ E0} and a set of partial isometries {se : e ∈ E1} with mutually orthogonal ranges which satisfy the Cuntz-Krieger relations: (CK1) s∗ ese = pr(e) for every e ∈ E1; (CK2) pv =Ps(e)=v ses∗ (CK3) ses∗ e ≤ ps(e) for every e ∈ E1. e for every v ∈ E0 reg; The graph algebra C ∗(E) is defined to be the C ∗-algebra generated by a universal Cuntz-Krieger E-family. It will in this paper also be relevant to work with relative graph C ∗- algebras introduced in [11]. To define a relative graph C ∗-algebra we must, in addition to a graph E, specify a subset R of E0 reg. A Cuntz-Krieger (E, R)-family is then a set of mutually orthogonal projections {pv : v ∈ E0} and a set of partial isometries {se : e ∈ E1} with mutually orthogonal ranges which satisfy the relations (CK1) and (CK3) above together with the following relative Cuntz-Krieger relation: e for every v ∈ R. The relative graph algebra C ∗(E, R) is defined to be the C ∗-algebra gen- erated by a universal Cuntz-Krieger (E, R)-family. reg, then a Cuntz-Krieger (E, R)-family is the same as a Cuntz-Krieger E-family and C ∗(E, R) = C ∗(E). If R = ∅, then C ∗(E, R) is the Toeplitz algebra T (E) defined in [8, Theorem 4.1]. We will call a Cuntz-Krieger (E, ∅)-family a Toeplitz-Cuntz-Krieger E-family. If R = E0 (RCK2) pv =Ps(e)=v ses∗ 4 TOKE MEIER CARLSEN, SØREN EILERS, AND MARK TOMFORDE the set of all paths of length n, and we let E∗ := S∞ A path in E is a sequence of edges α = α1α2 . . . αn with r(αi) = s(αi+1) for 1 ≤ i < n, and we say that α has length α = n. We let En denote n=0 En denote the set of finite paths in E. Note that vertices are considered paths of length zero. The maps r, s extend to E∗, and for v, w ∈ E0 we write v ≥ w if there exists a path α ∈ E∗ with s(α) = v and r(α) = w. Also for a path α := α1 . . . αn we define sα := sα1 . . . sαn, and for a vertex v ∈ E0 we let sv := pv. It is a consequence of the relations (CK1) and (CK3) that C ∗(E, R) = span{sαs∗ β : α, β ∈ E∗ and r(α) = r(β)}. We say that a path α := α1 . . . αn of length 1 or greater is a cycle if r(α) = s(α), and we call the vertex s(α) = r(α) the base point of the cycle. A cycle is said to be simple if s(αi) 6= s(α1) for all 1 < i ≤ n. The following is an important condition in the theory of graph C ∗-algebras. Condition (K): No vertex in E is the base point of exactly one simple cycle; that is, every vertex is either the base point of no cycles or at least two simple cycles. For any graph E a subset H ⊆ E0 is hereditary if whenever v, w ∈ E0 with v ∈ H and v ≥ w, then w ∈ H. A hereditary subset H is saturated reg with r(s−1(v)) ⊆ H, then v ∈ H. For any satu- if whenever v ∈ E0 rated hereditary subset H, the breaking vertices corresponding to H are the elements of the set BH :=(cid:8)v ∈ E0 : s−1(v) = ∞ and 0 < s−1(v) ∩ r−1(E0 \ H) < ∞(cid:9). An admissible pair (H, S) consists of a saturated hereditary subset H and a subset S ⊆ BH. For a fixed graph E we order the collection of admissible pairs for E by defining (H, S) ≤ (H ′, S′) if and only if H ⊆ H ′ and S ⊆ H ′ ∪ S′. For any admissible pair (H, S) we define J(H,S) to be the ideal in C ∗(E) generated by {pv : v ∈ H} ∪ {pH v0 : v0 ∈ S}, where pH v0 is the gap projection defined by pH v0 := pv0 − Xs(e)=v0 r(e) /∈H ses∗ e. Note that the definition of BH ensures that the sum on the right is finite. For any graph E there is a canonical gauge action γ : T → Aut C ∗(E) with the property that for any z ∈ T we have γz(pv) = pv for all v ∈ E0 and γz(se) = zse for all e ∈ E1. We say that an ideal J ⊳ C ∗(E) is gauge invariant if γz(J) ⊆ J for all z ∈ T. There is a bijective correspondence between the lattice of admissible pairs of E and the lattice of gauge-invariant ideals of C ∗(E) given by (H, S) 7→ J(H,S) [2, Theorem 3.6]. When E satisfies Condition (K), all ideals of C ∗(E) INDEX MAPS IN THE K-THEORY OF GRAPH ALGEBRAS 5 are gauge invariant [2, Corollary 3.8] and the map (H, S) 7→ J(H,S) is onto the lattice of ideals of C ∗(E). When BH = ∅, we write JH in place of J(H,∅) and observe that JH equals the ideal generated by {pv : v ∈ H}. Note that if E is row-finite, then BH is empty for every saturated hereditary subset H. 3. K-theory for relative graph algebras For a graph E, the adjacency matrix is the E0 × E0 matrix AE with AE(v, w) := #(cid:8)e ∈ E1 : s(e) = v and r(e) = w(cid:9). Note that the entries of AE are elements of {0, 1, 2, . . .} ∪ {∞}. Writing the adjacency matrix with respect to the decomposition E0 = E0 sing, where the regular vertices are listed first, we obtain a (possibly infinite) block matrix reg ⊔ E0 AE =(cid:20)A α H η(cid:21) in which all entries of A and α are finite, but the entries in H and η may be infinite. We will often just substitute "∗" for H and η, as they turn out to be irrelevant for the K-theory. Indeed, by [2] and [5] we know that the map contains the needed information, as (cid:20)At − I αt (cid:21) : ZE0 αt (cid:21) K0(cid:0)C ∗(E)(cid:1) ≃ coker(cid:20)At − I reg → ZE0 αt (cid:21) . K1(cid:0)C ∗(E)(cid:1) ≃ ker(cid:20)At − I for any v ∈ R by This result can be generalized to relative graph C ∗-algebras. In fact, we is the adjacency matrix of E written with respect to the decomposition E0 = R ⊔ (E0 \ R), where the vertices belonging to R are listed first, then prove in Proposition 3.8 that if E is a graph, R ⊆ Ereg, and AE =(cid:2) A α H η(cid:3) αt i → K0(cid:0)C ∗(E, R)(cid:1) given there exists a group isomorphism χ0 : cokerh At−I αt i(cid:17) = [pv]0, χ0(cid:16)ev + imh At−I αt i and and we construct a similar group isomorphism χ1 between kerh At−I K1(cid:0)C ∗(E, R)(cid:1). For this we first introduce some notation: αt i, first note that by definition x has only finitely Given x ∈ kerh At−I x :=(cid:8)(e, i) : e ∈ E1, 1 ≤ i ≤ −xs(e)} ∪ {(v, i) : v ∈ E0, 1 ≤ i ≤ xv(cid:9) x :=(cid:8)(e, i) : e ∈ E1, 1 ≤ i ≤ xs(e)} ∪ {(v, i) : v ∈ E0, 1 ≤ i ≤ −xv(cid:9) and note, using the convention that r(v) = v for any v ∈ E0, that many nonzero entries xv1, . . . xvk . We define L+ L− 6 TOKE MEIER CARLSEN, SØREN EILERS, AND MARK TOMFORDE Lemma 3.1. When x ∈ kerh At−I αt i, then for any vertex v ∈ E0 the sets and L+ L− v =(cid:8)(x, i) ∈ L+ v =(cid:8)(x, i) ∈ L− x : r(x) = v(cid:9) x : r(x) = v(cid:9) are finite and have the same number of elements. Proof. We need to consider three cases separately. Case I: v ∈ R and xv ≥ 0. The number of elements in L+ v is and the number of elements in L− #(cid:8)e ∈ E1 : s(e) = w, r(e) = v(cid:9) · (−xw) = xv − Xxw<0 xv + Xxw<0 Xxw>0 #(cid:8)e ∈ E1 : s(e) = w, r(e) = v(cid:9) · xw = Xxw>0 v is At v,w xw At v,w xw so the claim follows by inspecting the v coordinate of the equality Atx = x. Case II: v ∈ R and xv < 0. As above. Case III: v ∈ E0 \ R. The number of elements in L+ v is and the number of elements in L− #(cid:8)e ∈ E1 : s(e) = w, r(e) = v(cid:9) · (−xw) = − Xxw<0 Xxw<0 #(cid:8)e ∈ E1 : s(e) = w, r(e) = v(cid:9) · xw = Xxw>0 Xxw>0 v is αt v,w xw αt v,w xw so the claim follows by inspecting the v coordinate of the equality αtx = 0. (cid:3) Lemma 3.2. L+ elements. x and L− x are finite sets, and have the same number of Proof. This follows from Lemma 3.1, as indeed L+ the set {v : xv 6= 0} ∪(cid:8)v : xw 6= 0 for some w ∈ s(r−1(v))(cid:9) which is finite since no w is an infinite emitter. v 6= ∅ only when v lies in (cid:3) Denote the common number of elements in L+ x and L− x by h. Because of Lemma 3.1, we can define bijections [·] : L+ x → {1, . . . , h} h·i : L− x → {1, . . . , h} with the property that (3.1) [x, i] = hy, ji =⇒ r(x) = r(y) INDEX MAPS IN THE K-THEORY OF GRAPH ALGEBRAS 7 with the convention r(v) = v. When A is a C ∗-algebra then we let Mh(A) denote the C ∗-algebra of h × h-matrices over A. We are ready for our key definitions: Definition 3.3. Suppose that A is a C ∗-algebra which contains a Toeplitz- Cuntz-Krieger E-family {pv : v ∈ E0} ∪ {se : e ∈ E1}. With notation as above, we define the two elements V, P ∈ Mh(A) by s(e)=w V = X1≤i≤xw P = X1≤i≤xw se E[w,i],he,ii + X1≤i≤−xw pwE[w,i],[w,i] + X1≤i≤−xw s(e)=w s(e)=w,r(e)=v s∗ e E[e,i],hw,ii pv E[e,i],[e,i]. and (3.2) (3.3) (3.4) (3.5) Here E•,• denote the standard matrix units in Mh(M (A)) where M (A) is the multiplier algebra of A. Lemma 3.4. If {se, pv : e ∈ E1, v ∈ E0} is a Toeplitz-Cuntz-Krieger E- family, then P = X1≤i≤−xw V ∗ = X1≤i≤xw V V ∗ = X1≤i≤xw V ∗V = X1≤i≤−xw s(e)=w s(e)=w s(e)=w s∗ e ses∗ e s(e)=w,r(e)=v pwEhw,ii,hw,ii + X1≤i≤xw Ehe,ii,[w,i] + X1≤i≤−xw E[w,i],[w,i] + X1≤i≤−xw Ehw,ii,hw,ii + X1≤i≤xw ses∗ e s(e)=w s(e)=w,r(e)=v s(e)=w,r(e)=v pvEhe,ii,he,ii, seEhw,ii,[e,i], pvE[e,i],[e,i], pv Ehe,ii,he,ii. Proof. It follows from Lemma 3.2 and Equation (3.1) that pr(x)Ehx,ii,hx,ii, and it is easy to check that x x X(x,i)∈L+ pr(x)E[x,i],[x,i] = X(x,i)∈L− P = X(x,i)∈L+ pr(x)Ehx,ii,hx,ii = X1≤i≤−xw x pr(x)E[x,i],[x,i] and that X(x,i)∈L− x pwEhw,ii,hw,ii + X1≤i≤xw s(e)=w,r(e)=v pv Ehe,ii,he,ii 8 TOKE MEIER CARLSEN, SØREN EILERS, AND MARK TOMFORDE from which Equation (3.2) then follows. Equation (3.3) is straightforward to check. For Equation (3.4), using only (3.3) and the matrix unit relations we get that V V ∗ = X1≤i≤xw s(e)=w ses∗ e E[w,i],[w,i] + X1≤i≤−xw s∗ ese′ E[e,i],[e′,i] s(e)=w,r(e)=v s(e′)=w,r(e′)=v′ and (3.4) holds from (CK1) and the fact that the se's have mutually ortho- gonal ranges. The computation for V ∗V is similar. (cid:3) Lemma 3.5. If {se, pv : e ∈ E1, v ∈ E0} is a Toeplitz-Cuntz-Krieger E- family, then V is a partial isometry with P V = V P = V . Proof. Using Equation (3.5), the definition of V , and the fact that the se's are partial isometries, we see that V V ∗V = V , so that V is a partial isometry. Furthermore, (CK3) implies P V = V and V P = V by Equation (3.2). (cid:3) We now let {se, pv : e ∈ E1, v ∈ E0} be the universal Cuntz-Krieger x for the correspond- (E, R)-family generating C ∗(E, R) and write V the added subscript to emphasize the dependence of each of V and P on ing elements V and P in Mh(cid:0)C ∗(E, R)(cid:1) defined in Definition 3.3, using x ∈ kerh At−1 αt i. In addition, we define U x + (1 − P x and P x := V x). x = P x, and hence that U x is a unitary. Fact 3.6. We have that V V∗ x = V∗ x x V Proof. It follows from Equation (3.4) and (RCK2) that V x V∗ x = X1≤i≤xw = X1≤i≤xw = P x ses∗ e E[w,i],[w,i] + X1≤i≤−xw  Xs(e)=w pwE[w,i],[w,i] + X1≤i≤−xw pvE[e,i],[e,i] s(e)=w,r(e)=v s(e)=w,r(e)=v pvE[e,i],[e,i] showing the first claim. Likewise, Equation (3.5) and (RCK2) show that V∗ (cid:3) x is a unitary follows. x. The fact that U x = P V x Remark 3.7. Notice that although U x does depend on the choice of bijections h·i : L− x → {1, . . . , h}, be the adjacency matrix of E written with respect to the decomposition E0 = V ⊔ (E0 \ V ) where the vertices belonging to V are listed first. the element [U [·] : L+ x → {1, . . . , h} x]1 of K1(cid:0)C ∗(E, R)(cid:1) does not. AE =(cid:20)A α H η(cid:21) Proposition 3.8. Let E be a graph, let V be a subset of Ereg and let INDEX MAPS IN THE K-THEORY OF GRAPH ALGEBRAS 9 given for any v ∈ E0 by (1) There exists a group isomorphism χ0 : cokerh At−I αt (cid:21)(cid:19) = [pv]0. χ0(cid:18)ev + im(cid:20)At − I (3.6) αt i → K0(cid:0)C ∗(E, R)(cid:1) The preimage of the positive cone of K0(cid:0)C ∗(E, R)(cid:1) is generated by nev : v ∈ E0o ∪nev −Xe∈F sing, F ⊆ s−1(v), F finiteo. (2) The map χ1 : kerh At−I αt i → K1(cid:0)C ∗(E, R)(cid:1) given by er(e) : v ∈ E0 χ1(x) = [U x]1 is group isomorphism. Proof. As noted in [11], we can realize C ∗(E, R) as a relative Cuntz-Pimsner algebra over a Hilbert bimodule XE. It is not difficult to check that the corresponding Toeplitz algebra TXE is isomorphic to the Toeplitz algebra T (E). We let π : T (E) → C ∗(E, R) denote the canonical map, so that (3.7) 0 / ker π ι / T (E) π / C ∗(E, R) / 0 is exact. The associated K-theory is then K0(ker π) ι∗ ∂1 K1(cid:0)C ∗(E, R)(cid:1) π∗ / K0(cid:0)T (E)(cid:1) K1(cid:0)T (E)(cid:1) π∗ / K0(cid:0)C ∗(E, R)(cid:1) K1(ker π). ι∗ Now we appeal to Katsura's work. It follows from the results of [10, §8], that ker π and T (E) are KK-equivalent to the commutative AF -algebras c0(R) and c0(E0), respectively, and that there are group isomorphisms κ : K0(ker π) → ZR and λ : K0(cid:0)T (E)(cid:1) → ZE0 ι∗ / K0(ker π) 0 / K1(cid:0)C ∗(E, R)(cid:1) ∂1 κ K0(cid:0)T (E)(cid:1) ZR h At−I αt i / / ZE0 λ such that the diagram π∗ / K0(cid:0)C ∗(E, R)(cid:1) / 0 commutes with the top row exact. In [10] concrete ∗-homomorphisms are given inducing κ and λ, but we do not need them here. All we need is the fact that λ(pv) = ev and (3.8) κpw − Xs(e)=w ses∗ e0  = ew / / / / / /   O O o o o o / / / /     / / 10 TOKE MEIER CARLSEN, SØREN EILERS, AND MARK TOMFORDE x = V x + (1 − P x = x are the elements V and P in It follows that π∗ ◦ λ−1 is a surjective group for v ∈ E0 and w ∈ R. homomorphism from ZE0 [pv]0 and whose kernel is imh At−I χ0 : cokerh At−I to K0(cid:0)C ∗(E, R)(cid:1) which for any v ∈ E0 maps ev to αt i. The existence of a group isomorphism αt i → K0(cid:0)C ∗(E, R)(cid:1) which for any v ∈ E0 satisfies Equation (3.6) follows from this. The description of the positive cone in the row-finite case was given in [1, Theorem 7.1]. For the general situation, it is shown in [15, Theorem 2.2] that the process of desingularization can be used to extend the result from the row-finite case to the general case. x is a partial x is also a partial x as an element of K0(ker π). αt i → K1(cid:0)C ∗(E, R)(cid:1) is a group isomorphism, To see that χ1 : kerh At−I αt i and lift U fix x ∈ kerh At−I x) ∈ Mh(cid:0)C ∗(E, R)(cid:1) to eU x) ∈ Mh(cid:0)T (E)(cid:1) where eV x and eP x + (1 −eP eV Mh(cid:0)T (E)(cid:1) we get by using the universal Toeplitz-Cuntz-Krieger E-family which generates T (E) in Definition 3.3. By Lemma 3.5, eV It follows that eU x = eV xeP xeV x = eV isometry with eP isometry. We need to compute the defect of eU pw − Xs(e)=w x = X1≤i≤xw 1 −eU xeU∗ xeV∗ x −eV x =eP and a similar equation for 1 −eU∗ xeU xwpw − Xs(e)=w x(cid:3)0 = Xxw6=0 (cid:2)1 −eU x(cid:3)0 −(cid:2)1 −eU∗ xeU∗ xeU e E[w,i],[w,i] e0 αt i. This shows that χ1 is injective. Let us also prove that (cid:20)At − I αt (cid:21) ◦ κ ◦ ∂1(y) = λ ◦ ι∗ ◦ ∂1(y) = 0 χ1 is a group isomorphism. Fix y ∈ K1(C ∗(E, R)) and note that which together with Equation (3.8) and Equation (3.9) implies that for any x ∈h At−I x. Hence, in K0(ker π) we have that so that z := κ ◦ ∂1(y) lies in kerh At−I αt i. Since κ ◦ ∂1 is injective, it follows from Equation (3.10) that χ1(z) = y. We conclude that κ ◦ ∂1 is actually an inverse to χ1, and hence χ1 is a group isomorphism. (cid:3) (3.9) (3.10) ses∗ ses∗ We have by Lemma 3.4 that x. κ ◦ ∂1 ◦ χ1(x) = x 4. The index map Let E be a graph and let J be a gauge-invariant ideal in C ∗(E). It follows from [2] that J is of the form J(H,S) for an admissible pair (H, S). Writing INDEX MAPS IN THE K-THEORY OF GRAPH ALGEBRAS 11 the adjacency matrix of E with respect to the decomposition E0 reg ∩ H, E0 sing ∩ H, E0 reg \ H, E0 sing\(H ∪ S), S we arrive at the matrix ∗ 0 0 0 A α 0 ∗ 0 0 X ξ B β η ∗ ∗ ∗ ∗ ∗ Γ γ Z ∗ ∗  .  We are now ready to state our main result. Here and below, whenever T : G1 → G2 is a group homomorphism between abelian groups and H1 and H2 are subgroups of G1 and G2, respectively, such that T (H1) ⊆ H2, then we also use T to denote the group homomorphism from G1/H1 to G2/H2 induced by T , and we denote by Ia1···ak the canonical inclusion of the indicated components of a direct sum into a larger direct sum, and by Pa1···ak the corresponding projection. Theorem 4.1. Let E be a graph and let (H, S) be an admissible pair. The six term exact sequence in K-theory ι∗ ∂1 π∗ αt is isomorphic to K0(cid:0)J(H,S)(cid:1) K1(cid:0)C ∗(E)/J(H,S)(cid:1) / coker coker(cid:20) At−I 0 (cid:21) 0 I# " X t 0 ker ker(cid:20) Bt−I Γt ηt Z t−I(cid:21) where eI is given by the block matrix I1◦P2 eI ξt 0 βt γt βt ηt At−I X t ξt αt 0 Bt−I 0 0 / K0(cid:0)C ∗(E)(cid:1) K1(cid:0)C ∗(E)(cid:1)   = I125 − αt 0 Bt−I 0 0 At−I X t ξt  βt ηt I 0 0 0 I 0 0 0 −Γt 0 0 −γt 0 0 I − Z t  π∗ ι∗ ∂0 / K0(cid:0)C ∗(E)/J(H,S)(cid:1) K1(cid:0)J(H,S)(cid:1) /coker(cid:20) Bt−I Γt ηt Z t−I(cid:21) βt γt P345 0 ker(cid:20) At−I 0 (cid:21) αt I1 0 0 0 0 0 0 0 0 Γt 0 0 γt 0 0 Z t .  / /   O O o o o o / /   O O o o o o 12 TOKE MEIER CARLSEN, SØREN EILERS, AND MARK TOMFORDE Each cokernel is ordered as described in Theorem 3.8. We postpone the proof of the theorem to the ensuing section, but remark here that the iso- morphism between the two six term exact sequences is given by explicit defined maps which are described in the proof. For now, let us record a number of examples and specializations: Remark 4.2. If the saturated hereditary subset H has no breaking vertices (this is always the case if E is row-finite), or if S = ∅, then the six term exact sequence of Theorem 4.1 reduces to (4.1) αt i I12 cokerh At−I ξt# "X t kerh Bt−I βt i / coker" At−I X t βt # αt 0 Bt−I 0 ξt βt # ker" At−I X t αt 0 Bt−I 0 ξt P2 P34 /cokerh Bt−I βt i 0 αt i . kerh At−I I1 Remark 4.3. Let E be a row-finite graph with no sinks. Then any gauge- invariant ideal in C ∗(E) has the form JH for some saturated hereditary subset H and the six term exact sequence of Theorem 4.1 reduces in this case to coker(cid:2)At − I(cid:3) hX ti ker(cid:2)Bt − I(cid:3) I1 P2 0 / coker(cid:20)At − I ker(cid:20)At − I 0 X t Bt − I(cid:21) P2 Bt − I(cid:21) X t I1 / coker(cid:2)Bt − I(cid:3) ker(cid:2)At − I(cid:3) . 0 Corollary 4.4. Let E be a graph such that the associated graph C ∗-algebra C ∗(E) contains a unique proper nontrivial ideal. Then this ideal has the form JH for some saturated hereditary subset H with no breaking vertices. Consequently, the cyclic six term exact sequence determined by the short exact sequence 0 → JH → C ∗(E) → C ∗(E)/JH → 0 is isomorphic to the cyclic exact sequence described in (4.1). Proof. If E has a unique proper nontrivial ideal, then it follows from [7, Lemma 3.1] that the ideal has the form JH for a saturated hereditary subset H with no breaking vertices. (cid:3) Example 4.5. Consider the class of graphs Ex,y,z given by the adjacency matrix 0 0 0 0 x 1 1 0 y 1 1 1 z 0 1 1   / /   O O o o o o / /   O O o o o o INDEX MAPS IN THE K-THEORY OF GRAPH ALGEBRAS 13 where x, y, z ∈ N. These graphs all satisfy Condition (K) and have one non- trivial saturated hereditary subset (the subset consisting of the first vertex). Thus we are in the situation of Corollary 4.4, with E0 reg = {v2, v3, v4} and E0 reg = H = {v1}. Hence the adjacency matrix has the block form α ξ 0 B  / coker(cid:20) x y z 0 1 0(cid:21) ker(cid:20) x y z 0 1 0(cid:21) 0 1 0 1 0 1 0 1 0 1 0 1  P234 /cokerh 0 1 0 0 1 0i 1 0 1 0 ker 01×0 0 and the six-term exact sequence is coker 01×0 I1 [ x y z ] kerh 0 1 0 0 1 0i 1 0 1 P123 which simplifies to 0 x−z / Z / Z / Z/(x − z) ⊕ Z / Z / 0 when x 6= z and to 0 / Z Z 0 / Z / Z ⊕ Z / Z / 0 when z = x. The K0-group of the ideal is canonically ordered, and the order of the K0-group of the quotient is trivial, irrespective of x, y, z. We may hence apply [7] to prove that C ∗(Ex,y,z) ⊗ K ≃ C ∗(Ex′,y′,z ′) ⊗ K precisely when x − z = ±(x′ − z′). Example 4.6. Consider the class of graphs Fy,z given by the adjacency matrix where y, z ∈ N. These graphs all satisfy Condition (K) and have one non- trivial saturated hereditary subset {v1} for which {v3} is breaking. We furthermore have that E0 sing = {v1, v3}. If we consider the ideal J({v1},{v3}), then the adjacency matrix has the block form reg = {v2} and E0  0 y 0 0 3 1 ∞ z 3 0 ∗ 0 ξ B η ∗ Γ Z   / /   O O o o o o / / / / / / / / / / 14 TOKE MEIER CARLSEN, SØREN EILERS, AND MARK TOMFORDE which gives simplifying to coker 02×0 0 1i h y 0 ker [ 2 z 1 2 ] (cid:20) 1 0 0 −2(cid:21)/ 0 −z 2 / cokerh y 1i kerh y 1i 2 / coker [ 2 z 1 2 ] ker 02×0 0 4−zi/ / Z2h −1 −2y / Z2 0 / Zz−4 / 0 when z 6= 4 and to 0 when z = 4. 1 i / / Z h −2y 0 i/ / Z2h −1 −2y 0 / Z2 / Z / 0 In both cases, the K0-group of the ideal is ordered by (cid:8)(x1, x3) : x3 > 1 or [x3 = 0, x1 ≥ 0](cid:9), having only the trivial automorphism, so the computations combine with [7, Theorem 4.7] to show that C ∗(Fy,z) ⊗ K ≃ C ∗(Fy′,z ′) ⊗ K precisely when 4 − z = ±(4 − z′) and y − y′ ∈ (4 − z)Z. 5. Proof of main result The isomorphism of the two six-term exact sequences in Theorem 4.1 is given by the six group isomorphisms χ′ 1 defined as follows. If we let E(H,S) be the subgraph of E with vertices H ∪S and edges s−1(H)∪ 0, χ0, χ′′ 1, χ1, χ′′ 0, χ′ (cid:0)s−1(S)∩r−1(H)(cid:1), then the graph C ∗-algebra C ∗(cid:0)E(H,S)(cid:1) is isomorphic to a full corner of J(H,S) via an embedding φ : C ∗(cid:0)E(H,S)(cid:1) → J(H,S) with φ(pv) = φ∗ : K∗(cid:0)C ∗(E(H,S))(cid:1) → K∗(cid:0)J(H,S)(cid:1). Thus if we let χ pv for v ∈ H, φ(pv0 ) = pH (H,S) (cf. [2]). Notice that (E(H,S))0 sing ∩H)∪S. It follows (for example by [12, Proposition 1.2]) that φ induces an isomorphism denote the group isomorphisms given by Proposition 3.8 for C ∗(E(H,S)), then v0 for v0 ∈ S and φ(se) = se for e ∈ E1 sing = (E0 reg ∩H and that (E(H,S))0 reg = E0 E(H,S) ∗ and χ′ 0 := φ∗ ◦ χ E(H,S) 0 χ′ 1 := φ∗ ◦ χ E(H,S) 1 αt : coker(cid:20) At−I : ker(cid:20) At−I 0 (cid:21) → K0(cid:0)J(H,S)(cid:1) 0 (cid:21) → K1(cid:0)J(H,S)(cid:1) αt are group isomorphisms. Similarly, if we let E \ H be the subgraph of E with vertices E0 \ H and edges r−1(E0 \ H), then there is an isomorphism ψ : C ∗(E \ H, S) → C ∗(E)/J(H,S) which for any v ∈ E0 \ H maps pv to /   O O o o o o / / / / / / INDEX MAPS IN THE K-THEORY OF GRAPH ALGEBRAS 15 pv + J(H,S) and for any e ∈ r−1(E0 \ H) maps se to se + J(H,S) (cf. reg \ H and that (E \ H)0 Example 3.10]). Notice that (E \ H)0 sing \ H. Thus if we let χ(E\H,S) E0 Proposition 3.8 for C ∗(E \ H, S), then [11, sing = denote the group isomorphisms given by reg = E0 ∗ 0 := ψ∗ ◦ χ(E\H,S) χ′′ 0 and 1 := ψ∗ ◦ χ(E\H,S) χ′′ 1 βt : coker(cid:20) Bt−I Γt : ker(cid:20) Bt−I Γt ηt Z t−I(cid:21) → K0(cid:0)C ∗(E)/J(H,S)(cid:1) ηt Z t−I(cid:21) → K1(cid:0)C ∗(E)/J(H,S)(cid:1) βt γt γt are group isomorphisms. Finally we let χ∗ denote the group isomorphisms given directly by Proposition 3.8 for C ∗(E). The theorem then follows from the ensuing six claims. Claim 5.1. ι∗ ◦ χ′ Proof. If v ∈ H, then we have that 0 = χ0 ◦eI. (cid:18)ev + im(cid:20) At−I αt χ0 ◦ 0 0 I 0 0 I 0 0 −Γt 0 0 −γt 0 0 I−Z t and if v0 ∈ S, the left hand side equals   0 (cid:21)(cid:19) , At−I X t ξt αt 0 Bt−I 0 0 βt ηt = ι∗ ◦ χ′ 0 (cid:21)(cid:19) = χ0ev + im = [pv]0 =(cid:2)ι(φ(pv))(cid:3)0 0(cid:18)ev + im(cid:20) At−I   v0)(cid:3)0 =(cid:2)ι(φ(pv))(cid:3)0 αt 0 Bt−I 0 0 At−I X t ξt βt ηt αt r(e) /∈H χ0ev0 − Xs(e)=v0 er(e) + im = [pv0]0 − Xs(e)=v0 e]0 =(cid:2)ι(pH 0 (cid:21)(cid:19) . 0(cid:18)ev + im(cid:20) At−I = ι∗ ◦ χ′ [ses∗ r(e) /∈H αt Claim 5.2. π∗ ◦ χ0 = χ′′ 0 ◦ P345. Proof. As above, we check the claim of each class given by ev. If v ∈ H, then both sides vanish. If v /∈ H, both sides equal [pv]0. (cid:3) Claim 5.3. π∗ ◦ χ1 = χ′′ 1 ◦ I1 ◦ P2. (cid:3) 16 TOKE MEIER CARLSEN, SØREN EILERS, AND MARK TOMFORDE Proof. Fix x = [ y z ] ∈ ker At−I X t ξt αt 0 Bt−I 0 0 βt ηt . γt βt Then [ z 0 ] ∈ ker(cid:20) Bt−I Γt ηt Z t−I(cid:21) and we furthermore have that L+ x . Thus if we let h be the number of elements in L+ L− [ z 0 ] and we let h′ denote the number of elements in L+ [ z 0 ] can choose the bijections ⊆ L− ⊆ L+ x and [ z 0 ] x (and in L− x ), (and in L− [ z 0 ] ), then we [·] : L+ x → {1, . . . , h} h·i : L− x → {1, . . . , h} and [·]′ : L+ [ z 0 ] → {1, . . . , h′} h·i′ : L− [ z 0 ] → {1, . . . , h′} such that [·] is an extension of [·]′, and h·i is an extension of h·i′. We then have that π(V s(e)=w x) = π X1≤i≤xw = X1≤i≤zw = ψ X1≤i≤zw 0 ](cid:17) = ψ(cid:16)V[ z s(e)=w s(e)=w s(e)=w seE[w,i],he,ii + X1≤i≤−xw π(se)E[w,i],he,ii + X1≤i≤zw seE[w,i],he,ii + X1≤i≤−zw s(e)=w s(e)=w s∗ e s∗ e π(s∗ e)E[e,i],hw,ii E[e,i],hw,ii E[e,i],hw,ii 0 ](cid:17). x) = ψ(cid:16)P[ z since se ∈ J(H,S) = ker π when s(e) (and thus r(e)) lies in H, and zw = xw when w /∈ H. A similar computation for P x shows that π(P Thus π(U x) = ψ(cid:16)U[ z π∗ ◦ χ1(x) =(cid:2)π(U 0 ](cid:17) and x)(cid:3)1 =hψ(U[ z 0 ])i1 Claim 5.4. ι∗ ◦ χ′ 1 = χ1 ◦ I1. = χ′′(cid:0)[ z 0 ](cid:1) = χ′′ 1 ◦ I1 ◦ P2(x). (cid:3) INDEX MAPS IN THE K-THEORY OF GRAPH ALGEBRAS 17 Proof. Fix x ∈ ker(cid:20) At−I 0 (cid:21). This follows like in Claim 5.3 by choosing the αt bijections → {1, . . . , h} [·] : L+ x → {1, . . . , h} → {1, . . . , h} h·i : L− x → {1, . . . , h} and [·] : L+ [ x 0 ] h·i : L− [ x 0 ] to be pairwise equal. Claim 5.5. ∂0 = 0. (cid:3) Proof. It follows from Claim 5.4 that ι∗ : K1(cid:0)J(H,S)(cid:1) → K1(cid:0)C ∗(E)(cid:1) is injec- tive. Thus im(∂0) = 0 from which it follows that ∂0 = 0. (cid:3) Claim 5.6. ∂1 ◦ χ′′ Proof. Fix x = [ y and γt βt ξt 0 1 = χ′ 0 ◦(cid:20) X t 0 0 I(cid:21). ηt Z t−I(cid:21). We lift ψ(V z ] ∈ ker(cid:20) Bt−1 Γt se E[w,i],he,ii + X1≤i≤−xw x = X1≤i≤xw bV x = X1≤i≤xw pwE[w,i],[w,i] + X1≤i≤−xw bP s(e)=w,r(e) /∈H s(e)=w,r(e) /∈H s(e)=w,r(e)=v v /∈H x) and ψ(P x) to s∗ e E[e,i],hw,ii pv E[e,i],[e,i], index map on [U cf. [14, Proposition 9.2.2]. We have, using Lemma 3.4, that is a Toeplitz-Cuntz-Krieger (E \ H)-family, it follows from Lemma 3.5 that x. It follows that also x) is a partial isometry. Hence, to compute the value of the x in K0(J(H,S)), respectively, in Mh(cid:0)C ∗(E)(cid:1). Since(cid:8)se, pv : e ∈ r−1(E0 \ H), v ∈ E0 \ H(cid:9) xbP x = bV x is a partial isometry and that bP bV xbV x = bV x + (1−bP bU x :=bV x]1, we just need to compute the defect ofbU x −bV x = bP xbV∗ xbU∗ 1 −bU pw − Xs(e)=w,r(e)=v e E[w,i],[w,i] = X1≤i≤xw  Xs(e)=w,r(e)=v = X1≤i≤yw e − Xs(e)=w,r(e)=v e E[w,i],[w,i] ses∗ ses∗ ses∗ v6∈H v6∈H x 18 TOKE MEIER CARLSEN, SØREN EILERS, AND MARK TOMFORDE + X1≤i≤zv0 = X1≤i≤yw Xs(e)=w,r(e)=v v∈H pv0 − Xs(e)=v0,r(e) /∈H ses∗ e ses∗ e E[v0,i],[v0,i] E[w,i],[w,i] + X1≤i≤zv0 pH v0 E[v0,i],[v0,i]. Passing to the K0-group and using that se ∈ C ∗(EH ), we get [s∗ e]0 + X1≤i≤zv0(cid:2)pH v0(cid:3)0 ese]0 + X0<zv0 zv0(cid:2)pH v0(cid:3)0 [pv]0 + X0<zv0 zv0(cid:2)pH v0(cid:3)0. [pv]0 + Xzv0 6=0 zv0(cid:2)pH v0(cid:3)0. v∈H v∈H v∈H v∈H χ′ ξt 0 [ses∗ By comparison, x is similar, and we get (cid:2)1 −bU yw Xs(e)=w,r(e)=v x(cid:3)0 = X1≤i≤yw Xs(e)=w,r(e)=v xbU∗ yw Xs(e)=w,r(e)=v yw Xs(e)=w,r(e)=v = X0<yw = X0<yw The computation for 1 −bU∗ xbU x]0 = X06=yw x]0 − [1 −bU∗ xbU∗ [1 −bU xbU 0(cid:18)(cid:20) X t 0 0 I(cid:21) x(cid:19) = χ′ 0 (cid:21)(cid:19) 0(cid:18)(cid:20) X t z (cid:21) + im(cid:20) At−1 yw Xv∈E0 = Xyw6=0 Xv,w[pv]0 + Xv∈E0 + Xzv0 6=0 zv0(cid:2)pH v0(cid:3)0 yw Xs(e)=w,r(e)=v + Xzv0 6=0 zv0(cid:2)pH v0(cid:3)0 yw Xs(e)=w,r(e)=v [pv]0 + Xzv0 6=0 = Xxw6=0 = Xxw6=0 v∈E0 reg∩H reg∩H y ξt y αt v∈H [pv]0 + Xs(e)=w,r(e)=v v∈E0 sing∩H ξv,w[pv]0 [pv]0 zv0(cid:2)pH v0(cid:3)0, sing∩H completing the proof. (cid:3) INDEX MAPS IN THE K-THEORY OF GRAPH ALGEBRAS 19 References [1] P. Ara, M. A. Moreno, and E. Pardo. Nonstable K-theory for graph algebras. Algebr. Represent. Theory, 10(2):157 -- 178, 2007. [2] T. Bates, D. Pask, I. Raeburn, and W. Szymanski. C ∗-algebras of row-finite graphs. New York J. Math., 6:307 -- 324, 2000. [3] L.G. Brown and G.K. Pedersen. C ∗-algebras of real rank zero. J. Funct. Anal., 99:131 -- 149, 1991. [4] J. Cuntz. On the homotopy groups of the space of endomorphisms of a C ∗-algebra (with applications to topological Markov chains). In Operator algebras and group representations, Vol. I (Neptun, 1980), volume 17 of Monogr. Stud. Math., pages 124 -- 137. Pitman, Boston, MA, 1984. [5] D. Drinen and M. Tomforde. Computing K-theory and Ext for graph C ∗-algebras. Illinois J. Math., 46:81 -- 91, 2002. [6] S. Eilers, G. Restorff, and E. Ruiz. Classifying C ∗-algebras with both finite and infinite subquotients. Preprint, arXiv:0801.0324v3, 2010. [7] S. Eilers and M. Tomforde. On the classification of nonsimple graph algebras. Math. Ann., 346:393 -- 418, 2010. [8] N.J. Fowler and I. Raeburn. The Toeplitz algebra of a Hilbert bimodule. Indiana Univ. Math. J., 48(1):155 -- 181, 1999. [9] J.A Jeong. Real rank of C ∗-algebras associated with graphs. J. Aust. Math. Soc., 77(1):141 -- 147, 2004. [10] T. Katsura. On C ∗-algebras associated with C ∗-correspondences. J. Funct. Anal., 217(2):366 -- 401, 2004. [11] P.S. Muhly and M. Tomforde. Adding tails to C ∗-correspondences. Doc. Math., 9:79 -- 106, 2004. [12] W. Paschke. K-theory for actions of the circle group on C ∗-algebras. J. Operator Theory, 6(1):125 -- 133, 1981. [13] M. Rørdam. Classification of Cuntz-Krieger algebras. K-Theory, 9(1):31 -- 58, 1995. [14] M. Rørdam, F. Larsen, and N. Laustsen. An introduction to K-theory for C ∗-algebras, volume 49 of London Mathematical Society Student Texts. Cambridge University Press, Cambridge, 2000. [15] M. Tomforde. The ordered K0-group of a graph C ∗-algebra. C. R. Math. Acad. Sci. Soc. R. Can., 25(1):19 -- 25, 2003. Department of Mathematical Sciences, Norwegian University of Science and Technology, NO-7491 Trondheim, Norway E-mail address: [email protected] Department for Mathematical Sciences, University of Copenhagen, Uni- versitetsparken 5, DK-2100 Copenhagen Ø, Denmark E-mail address: [email protected] Department of Mathematics, University of Houston, Houston, TX 77204- 3008, USA E-mail address: [email protected]
1912.10700
1
1912
2019-12-23T09:37:37
Multipliers and Duality for Group Actions
[ "math.OA" ]
We define operator-valued Schur and Herz--Schur multipliers in terms of module actions, and show that the standard properties of these multipliers follow from well-known facts about these module actions and duality theory for group actions. These results are applied to study the Herz--Schur multipliers of an abelian group acting on its Pontryagin dual: it is shown that a natural subset of these Herz--Schur multipliers can be identified with the classical Herz--Schur multipliers of the direct product of the group with its dual group.
math.OA
math
MULTIPLIERS AND DUALITY FOR GROUP ACTIONS ANDREW MCKEE Abstract. We define operator-valued Schur and Herz -- Schur multipli- ers in terms of module actions, and show that the standard properties of these multipliers follow from well-known facts about these module actions and duality theory for group actions. These results are applied to study the Herz -- Schur multipliers of an abelian group acting on its Pontryagin dual: it is shown that a natural subset of these Herz -- Schur multipliers can be identified with the classical Herz -- Schur multipliers of the direct product of the group with its dual group. 1. Introduction Schur multipliers -- the scalar-valued functions on N × N for which the entrywise product maps B(ℓ2) into itself -- arose from Schur's work on the entrywise product of matrices in the early twentieth century. Their importance was recognised by Grothendieck [10] (see also Pisier [19, Chapter 5]), who used them to formulate his fundamental theorem. These classical Schur multipliers have been extended in several directions; see, for example, [19, Chapter 5]. Herz -- Schur multipliers, or completely bounded multipliers of the Fourier algebra of a group, originate in work of Herz [13] where they were viewed as a generalisation of the Fourier -- Stieltjes transform. They have proved useful in the study of approximation properties of operator algebras associated to groups; this was first made explicit by De Canni`ere and Haagerup [7], and has since been exploited by many other authors (see [6, Chapter 12] for further references). This utility has driven the development of several classes of Herz -- Schur multipliers, for example the radial multipliers which first appeared in [12]. Bozejko and Fendler [4] linked these two notions, using unpublished work of Gilbert (see also Jolissaint [14]) to give a 'transference' theorem, show- ing that every Herz -- Schur multiplier of G gives rise to a Schur multiplier acting on B(L2(G)). Moreover, one can characterise the Herz -- Schur multi- pliers as those Schur multipliers which are invariant, in the sense that they commute with conjugation by the right regular representation of G. We regard the transference and characterisation results as important goals of the generalised theory presented here. 1 2 A. MCKEE The importance of the theory of multipliers has led to several authors introducing operator-valued versions of Schur multipliers [3, 17] and Herz -- Schur multipliers [1, 2, 8, 17]. In particular our work with Todorov and Tur- owska [17] develops and studies C ∗-algebra-valued versions of Schur and Herz -- Schur multipliers, including both transference and characterisation theorems. The present work arose from an attempt to distill the essential features of some of the proofs given in that paper. Aspects of the theory of Schur and Herz -- Schur multipliers have also been generalised to quantum groups. For example, Junge -- Neufang -- Ruan [15] give a transference theorem in the setting of locally compact quantum groups, and Brannan [5] uses similar ideas when discussing approximation properties of quantum groups. This paper serves two purposes: firstly we show how to obtain the main results of [17] in the von Neumann algebra setting, and secondly we show how the definitions and important properties of (operator-valued) Schur and Herz -- Schur multipliers can be obtained from basic properties of group and module actions on operator algebras. More specifically, after preliminaries in Section 2, in Section 3 we define Schur multipliers as completely bounded maps commuting with a particular module action, and obtain a dilation-type characterisation of these multipliers in Theorem 3.3. Section 4 begins with the definition of a Herz -- Schur multiplier of a group action, so that the classical Herz -- Schur multipliers are the Herz -- Schur mul- tipliers of the trivial action of the group on C. We then prove the main results of the paper: Proposition 4.5 is a version of transference for our mul- tipliers, identifying the Herz -- Schur multipliers of a group action with certain Schur multipliers associated to the dual coaction, and a characterisation of the Schur multipliers which arise in this way in Theorem 4.6. In Section 5 we focus on abelian groups. When G is abelian the algebra B(L2(G)) is the crossed product formed by an action of either G or the dual group G, and in Theorem 5.1 we characterise the maps on B(L2(G)) which are Herz -- Schur multipliers of both actions simultaneously as the Herz -- Schur multipliers of G × G. Finally, we note that preliminary investigations have recovered some of the results of this paper for Kac algebras. This will be explored in a future work. 2. Preliminaries Throughout, G denotes a locally compact group, and M a von Neumann algebra acting on the Hilbert space HM . The normal spatial tensor product of von Neumann algebras will be denoted by ⊗. The unit of M will be written 1M , id denotes the identity representation of a von Neumann algebra, and IH the identity operator on the Hilbert space H. We follow Nakagami -- Takesaki [18] (except that we use the left group von Neumann algebra). An action of G on M is a homomorphism α : G → MULTIPLIERS AND DUALITY FOR GROUP ACTIONS 3 Aut(M ), continuous in the point-weak* topology. Equivalently, there is a normal ∗-isomorphism πα : M → M ⊗ L∞(G) satisfying (πα ⊗ id) ◦ πα = (id ⊗ παG) ◦ πα. Here αG denotes the action of G on L∞(G), so that παG : L∞(G) → L∞(G)⊗L∞(G); παG(f )(s, t) := f (st), which is the coproduct on L∞(G). Given an action α the corresponding isomorphism πα is defined by f ∈ L∞(G), s, t ∈ G, πα(a)ξ(s) := α−1 s (a)ξ(s), a ∈ M, s ∈ G, ξ ∈ L2(G, HM ). The crossed product associated to the action α, denoted M ⋊αG, is the von Neumann algebra on HM ⊗L2(G) generated by πα(M ) and C⊗vN(G). Note that vN(G) is the crossed product formed by the trivial action of G on C. The definitions for a coaction of G are identical to the above, except that the roles of vN(G) and L∞(G) are exchanged: a coaction δ of G on M is a normal ∗-isomorphism πδ : M → M ⊗ vN(G) satisfying Here δG denotes the coaction of G on itself, so that (πδ ⊗ id) ◦ πδ = (id ⊗ πδG ) ◦ πδ. πδG : vN(G) → vN(G) ⊗ vN(G); πδG (λr) := λr ⊗ λr, r ∈ G, which is the coproduct on vN(G). The crossed product associated to the coaction δ, denoted M ⋊δ G, is the von Neumann algebra on HM ⊗ L2(G) generated by πδ(M ) and C ⊗ L∞(G). Note that the crossed product formed by the trivial coaction of G on C is L∞(G). When G is abelian vN(G) can be identified with L∞( G), so in this case a coaction of G is an action of G. Given an action α of G on M there is a dual coaction α of G on M ⋊αG, given by π α(cid:0)πα(a)λr(cid:1) := πα(a)λr ⊗ λr, a ∈ M, r ∈ G. Similarly, given a coaction δ there is a dual action δ of G on M ⋊δ G. The Takai duality theorem for abelian groups can be generalised to this setting, and gives isomorphisms (cid:0)M ⋊α G(cid:1) ⋊ α G ∼= M ⊗ B(L2(G)) (cid:0)M ⋊δ G(cid:1) ⋊δ G ∼= M ⊗ B(L2(G)). and We will denote the first of these isomorphisms by Φ; generators of M ⊗ B(L2(G)) by (1) it is given on the where a ∈ M, r ∈ G, φ ∈ L∞(G) [18, page 8]. Under Φ the second dual Φ(cid:0)πα(a)(cid:1) = πα(a)⊗IL2(G), Φ(IH⊗λr) = IH⊗λr⊗λr, Φ(IH⊗φ) = IH⊗IL2(G)⊗φ, action α of G on (cid:0)M ⋊α G(cid:1) ⋊ α G is identified with the action α ⊗ Ad ρ on M ⊗ B(L2(G)), where ρ is the right regular representation of G. We use the basic theory of operator spaces and completely bounded maps, as found in [9] for example, without comment. The space of completely 4 A. MCKEE bounded, weak*-continuous maps on a von Neumann algebra M will be writ- ten CBσ(M ); if M is also a bimodule over A then the completely bounded, weak*-continuous A-bimodule maps on M will be denoted by CBA σ (M ). 3. Schur multipliers In this section we define generalised Schur multipliers. Throughout X = (X, µ) denotes a standard measure space for which the underlying topology is locally compact. Definition 3.1. A Schur X-multiplier of M is a completely bounded, weak*- continuous, L∞(X)-bimodule map on M ⊗ B(L2(X)). Given a Banach alge- bra A such that M is an A-(bi)module, equip M ⊗B(L2(X)) with the natural A-(bi)module structure. A Schur X-multiplier of M with respect to A is a Schur X-multiplier of M which is also an A-(bi)module map. Remarks 3.2. i. When M = C the Schur multipliers defined above are the classical Schur multipliers. In this case we need only require bound- edness of the L∞(X)-bimodule map, complete boundedness follows au- tomatically (see e.g. [20, Section 2]). ii. More generally, if N is matricially norming for M then any bounded map which is an N ⊗ L∞(X)-bimodule map is a Schur X-multiplier of M with respect to N , since such a map is automatically completely bounded [20]. iii. Choosing HM = L2(Y ), M = B(L2(Y )) and A = L∞(Y ), with Y a standard measure space with locally compact topology, the definition above becomes the completely bounded L∞(X × Y )-bimodule maps on B(L2(X × Y )), i.e. the classical Schur multipliers on B(L2(X × Y )). iv. It is clear that a classical Schur X-multiplier defines a Schur X-multiplier of M , and that the Schur multipliers of M of this form are module maps for any module structure on M . Recall from [17] that given k ∈ L2(X × X) ⊙ M one can associate a bounded operator Tk by (2) Tk : L2(X, HM ) → L2(X, HM ); Tkξ(y) := ZX k(y, x)ξ(x) dx, and that such operators are norm-dense in M ⊗min K(L2(X)). In [17, The- orem 2.6] we showed that the Schur multipliers defined there correspond to certain symbols ϕ : X × X → CB(M ) via Sϕ(Tk) := Tϕ·k where ϕ · k(x, y) := ϕ(y, x)(cid:0)k(x, y)(cid:1). In this paper we have defined Schur X-multipliers of M , which act on M ⊗ B(L2(X)); in the next result, which is based on [17, Theorem 2.6], we show that our definition of a Schur multiplier S determines how S acts on the operators Tk defined above, and use this to associate a symbol to S. Theorem 3.3. Let M be a von Neumann algebra on the separable Hilbert space HM . The following are equivalent: MULTIPLIERS AND DUALITY FOR GROUP ACTIONS 5 i. S is a Schur X-multiplier of M ; ii. there exists a bounded function ϕ : X × X → CBσ(M ), of the form ϕ(x, y)(a) = W (y)∗ρ(a)V (x), x, y ∈ X, a ∈ M, with ρ a normal representation of M and V, W ∈ L∞(X, B(HM , Hρ)), such that S = Sϕ. Moreover, if M is an A-(bi)module then S is an A-(bi)module map if and only if ϕ(x, y) is an A-(bi)module map for almost all x, y ∈ X. Proof. (i) =⇒ (ii) Write S = W ∗ 0 θ(·)V0, where θ is a normal representation of M ⊗B(L2(X)) on the Hilbert space Hθ and V0, W0 ∈ B(HM ⊗L2(X), Hθ). The map θ0(T ) := θ(1M ⊗ T ), T ∈ B(L2(X)), defines a normal representation of B(L2(X)) on Hθ. As is well known, this implies that we can write Hθ = Hρ ⊗ L2(X) for another Hilbert space Hρ and identify θ0(T ) with IHρ ⊗ T . Since θ(a ⊗ T ) = θ(a ⊗ IL2(X))θ(IHρ ⊗ T ) = θ(a ⊗ IL2(X))θ0(T ) and θ(a ⊗ T ) = θ0(T )θ(a ⊗ IL2(X)) for all a ∈ M and T ∈ B(L2(X)) we have that θ(a ⊗ I) commutes with C ⊗ B(L2(X)), so we obtain a representation ρ of M on Hρ such that θ(a ⊗ T ) = ρ(a) ⊗ T , acting on Hρ ⊗ L2(X). We now have that S = W ∗ 0 (ρ ⊗ id)(·)V0 (identifying the ranges of V0 and W0 with Hρ ⊗ L2(X)). Arguing as in [17, Theorem 2.6], using that S commutes with C ⊗ L∞(X), we can find projections P, Q so that V := P V0 and W := QW0 commute with C ⊗ L∞(X), so by Takesaki [21, Theorem 7.10] V, W ∈ L∞(X, B(HM , Hρ)). We can now conclude S = Sϕ with ϕ(x, y)(a) = W (y)∗ρ(a)V (x) as in [17, Theorem 2.6]. (ii) =⇒ (i) It is clear that if S = Sϕ with ϕ as in (ii) then S = W ∗(ρ ⊗ id)(·)V , so S is completely bounded. Since V, W ∈ L∞(X, B(HM , Hρ)) it follows that S is an L∞(X)-bimodule map, so S is a Schur X-multiplier of M . To show that ϕ(x, y) is an A-(bi)module map when S is take a ∈ M, b ∈ A, k ∈ L2(X × X). Then (b ⊗ id)(cid:0)S(a ⊗ Tk)(cid:1) = (b ⊗ id)(cid:0)S(Ta⊗k(cid:1) = (b ⊗ id)Tϕ·(a⊗k) = Tb·(ϕ·(a⊗k)), and Since S is a module map the last two displays are equal, which implies (cid:0)S(b · a ⊗ Tk)(cid:1) = Tϕ·(b·a⊗k). k(y, x)(cid:0)b · ϕ(x, y)(a)(cid:1) = k(y, x)ϕ(x, y)(b · a), and using the fact that they hold for all k ∈ L2(X × X), x, y ∈ X we conclude that ϕ(x, y)(b · a) = b · ϕ(x, y)(a) for all a ∈ M and b ∈ A. A similar calculation shows that ϕ respects the right module action when S 6 A. MCKEE does; thus ϕ(x, y) is a completely bounded A-(bi)module map on M . The converse follows similarly. (cid:3) Remarks 3.4. i. The above theorem reduces to a well-known character- isation of classical Schur multipliers when M = C [11, 16]. ii. If α is an action of G on M then the crossed product by the dual coaction α is identified with M ⊗ B(L2(G)), and Theorem 3.3 above identifies Schur multipliers on this space with functions G × G → CBσ(M ). In the next section we will define Herz -- Schur multipliers of α and identify them with a certain subspace of these Schur multipliers on G × G. 4. Herz -- Schur Multipliers We are now going to define Herz -- Schur multipliers for a group action on a von Neumann algebra. Throughout, G denotes a second-countable locally compact group. Definition 4.1. Let α be an action of G on M . We say that a map S : M ⋊αG → M ⋊αG is a Herz -- Schur multiplier of α if S is completely bounded, weak*-continuous, and (3) π α ◦ S = (S ⊗ id) ◦ π α. We will refer to a map S satisfying condition (3) by writing "S commutes with α". Remarks 4.2. i. Condition (3) is the same as the condition which de- fines a Fourier multiplier of a locally compact quantum group (see e.g. Brannnan [5, Proposition 4.5]). See also the definition of a (right) co- variant map by Junge -- Neufang -- Ruan [15, pg 391]. ii. In particular, it is straightforward to show that T : vN(G) → vN(G) defines a classical Herz -- Schur multiplier if and only if T∗ is a completely bounded map on A(G) such that (4) T∗(uv) = T∗(u)v, u, v ∈ A(G). If α is the trivial action of G on C then α = δG, which induces the product on A(G). If T satisfies (4) then, for x ∈ vN(G), u, v ∈ A(G), the calculation hπδG ◦ T (x), u ⊗ vi = hT (x), uvi = hx, T∗(u)vi = hπδG(x), T∗(u) ⊗ vi shows that T satisfies (3). A similar calculation shows (3) implies (4). iii. Observe that M ⋊α G carries an A(G)-module structure: for u ∈ A(G) define u ∗ x := (id ⊗ u)π α(x), x ∈ M ⋊α G, so that u ∗ (πα(a)λr) = u(r)πα(a)λr. It is easy to see that Definition 4.1 is equivalent to requiring that S commutes with this module action. MULTIPLIERS AND DUALITY FOR GROUP ACTIONS 7 iv. Given a Herz -- Schur multiplier of α, say S, equation (3) and (iii) above imply that S(πα(a)λr) ∈ πα(M )λr, so there is some aS,r ∈ M with S(πα(a)λr) = πα(aS,r)λr. Setting F (r)(a) := aS,r we obtain a function F on G such that F (r) is a linear map on M for each r ∈ G. Moreover, since S is completely bounded and weak*-continuous F (r) must be so too. This shows that for every Herz -- Schur multiplier of α S there is a symbol F : G → CBσ(M ) such that S(cid:0)πα(a)λr(cid:1) = πα(cid:0)F (r)(a)(cid:1)λr, a ∈ M, r ∈ G. v. Suppose that v : G → C is a classical Herz -- Schur multiplier of G. For any action α of G on M we can extend Sv to a completely bounded, weak*-continuous map on M ⋊α G by Sv(cid:0)πα(a)λr(cid:1) = v(r)πα(a)λr, a ∈ M, r ∈ G. It is easily checked that Sv commutes with α, so that Sv is a Herz -- Schur multiplier of α. vi. Let G be abelian, and consider the canonical action of G on L∞(G) = vN( G). In [17, Section 6] we showed that every element of B(G) ⊙ B( G) is a Herz -- Schur multiplier of this action; moreover, by symmetry, each such multiplier is also a Herz -- Schur multiplier of G on L∞( G) = vN(G), and the multipliers of this form are A(G) module maps on vN(G) (and A( G) module maps on L∞(G)). We will study these multipliers further below. If M has a (left) module structure over A we can introduce an A-module structure on M ⋊α G by (5) b · πα(a)λr := πα(b · a)λr, b ∈ A, a ∈ M, r ∈ G. It is easy to check that under the additional assumption (6) b · αr(a) = αr(b · a), r ∈ G, a ∈ M, b ∈ A σ (M ) then S is also an A-module map, since this module action is the one induced on M ⋊α G by the canonical module action of A on M ⊗ B(L2(G)). If S is a Herz -- Schur multiplier with symbol F : G → CBA S(cid:0)b · πα(a)λr(cid:1) = S(cid:0)πα(b · a)λr(cid:1) = πα(cid:0)b · F (r)(a)(cid:1)λr = b · (cid:0)S(πα(a)λr)(cid:1). Recall that for an action α of G on M the crossed product by the dual coaction α of G on M ⋊α G can be identified with M ⊗ B(L2(G)). Given a map R : M ⋊α G → M ⋊α G we define a map R on M ⊗ B(L2(G)) by R := Φ−1 ◦ (R ⊗ id) ◦ Φ, where Φ is the isomorphism (1). Observe that R is completely bounded (resp. completely positive) if R is. In the remainder of this section we explain how Herz -- Schur multipliers of α interact with the Schur multipliers of M ⊗ B(L2(G)). Lemma 4.3. Let α be an action of G on M . Fix a ∈ M, r ∈ G, φ ∈ L∞(G) and suppose (ui)i is a net of positive, compactly supported functions with kuik1 = 1 whose support shrinks to {r}. 8 A. MCKEE i. The kernels ki(s, t) := ui(st−1)αs−1(a) satisfy Tki ii. The kernels hi(s, t) := ui(st−1)(αr−1(a)⊗αG w∗ → πα(a)λr. r−1(φ)) satisfy Thi w∗ → a⊗φλr. Proof. Routine calculations show hTkiξ, ηi → hπα(a)λrξ, ηi and hThiξ, ηi → h(a ⊗ φλr)ξ, ηi for all ξ, η ∈ L2(G, H). The conclusion follows because the weak* topology coincides with the WOT on bounded sets. (cid:3) Lemma 4.4. Let S : M ⊗B(L2(G)) → M ⊗B(L2(G)) be a Schur multiplier, τ the trivial action of G on M , β := τ ⊗ αG and Ψ : (M ⊗ L∞(G)) ⋊β G → M ⊗ B(L2(G)) the canonical isomorphism. Then S := Ψ−1 ◦ S ◦ Ψ is a Herz -- Schur multiplier of β, i.e. π β ◦ S = ( S ⊗ id) ◦ π β. Proof. Let ϕ be the symbol of S, obtained in Theorem 3.3. It is straight- forward to check, using Lemma 4.3, that for r ∈ G we have S(a ⊗ φλr) = ϕr(a ⊗ φ)λr, where ϕr : G → CBσ(M ⊗ L∞(G)) is given by ϕr(s)(x) := ϕ(s, r−1s)(βr−1 (x)). Now we calculate, for a ∈ M and φ ∈ L∞(G), π β ◦ S(cid:0)πβ(a ⊗ φ)λr(cid:1) = π β ◦ Ψ−1(cid:0)ϕr(a ⊗ φ)λr(cid:1) = π β(cid:16)πβ(cid:0)ϕr(a ⊗ φ)(cid:1)λr(cid:17) = πβ(cid:0)ϕr(a ⊗ φ)(cid:1)λr ⊗ λr = S(cid:0)πβ(a ⊗ φ)λr(cid:1) ⊗ λr = ( S ⊗ id) ◦ π β(cid:0)πβ(a ⊗ φ)λr(cid:1), which proves the claim. (cid:3) First we have a version of the transference theorem (see also [4, 15, 17]). Proposition 4.5. Let α be an action of G on M and S a Herz -- Schur multiplier of α with symbol F : G → CBσ(M ). Then S is a Schur multiplier of α with symbol ϕ(s, t)(a) = αt−1(F (ts−1)(αt(a))). Moreover, if M has an A-module structure satisfying (6) and F (r) is an A-module map for all r ∈ G then ϕ(s, t) is an A-module map for all s, t ∈ G, so S is also an A-module map. Proof. Let S be a Herz -- Schur multiplier of α. For a ∈ M , r ∈ G and φ ∈ L∞(G) we have S(cid:16)(cid:0)πα(a)λr(cid:1)(IHM ⊗ φ)(cid:17) = Φ−1 ◦ (S ⊗ id)(cid:0)πα(a)λr ⊗ λrφ) = Φ−1(cid:16)πα(cid:0)F (r)(a)(cid:1)λr ⊗ λrφ(cid:17) = (cid:16)S(cid:0)πα(a)λr(cid:1)(cid:17)(IHM ⊗ φ); similarly S commutes with left multiplication by L∞(G). That S is a Schur multiplier follows by linearity and weak*-continuity. To calculate the symbol ϕ associated to the Schur multiplier S fix a ∈ M and r ∈ G. For k ∈ L2(G × G, M ) we define kr : G → M by kr(p) := k(p, r−1p). Let (ui)i∈I and (ki)i∈I be as in Lemma 4.3. Similarly one checks MULTIPLIERS AND DUALITY FOR GROUP ACTIONS 9 i )i∈I converge to πα(a) in the weak* topology of L∞(G, M ). Since that (kr S(Tki) → πα(F (r)(a))λr we have αt−1(cid:0)F (r)(a)(cid:1) = πα(cid:0)F (r)(a)(cid:1)(t) = lim i (ϕ · ki)r(t) = lim i (ϕ · ki)(t, r−1t) The claimed identity follows. = ϕ(r−1t, t)(cid:0)ki(t, r−1t)(cid:1) = ϕ(r−1t, t)(cid:0)αt−1(a)(cid:1). The statement about module maps is an easy calculation using (6). (cid:3) The following result characterises the Herz -- Schur multipliers of α among the Schur multipliers of α. We identify α with the action α ⊗ Ad ρ as in (1). Theorem 4.6. Let α be an action of G on M and R a Schur multiplier on M ⊗ B(L2(G)). The following are equivalent: i. π α ◦ R = (R ⊗ id) ◦ π α; ii. R = S for some Herz -- Schur multiplier S of α. Moreover, if M has an A-module structure satisfying (6) then R is an A- module map if and only if S is. Proof. (i) =⇒ (ii) Since R commutes with α we deduce that R defines a map on the fixed points of this action, which can be identified with M ⋊αG (see e.g. [18, Theorem II.1.1]). Observe that (Ψ ⊗ id) ◦ π β ◦ Ψ−1 restricts to the coaction π α of G on M ⋊α G. Now calculate, using Ψ R = RΨ and Lemma 4.4, π α ◦ R = (Ψ ⊗ id) ◦ π β ◦ Ψ−1 ◦ R = (Ψ ⊗ id) ◦ π β ◦ R ◦ Ψ−1 = (Ψ ⊗ id) ◦ ( R ⊗ id) ◦ π β ◦ Ψ−1 = (R ⊗ id) ◦ (Ψ ⊗ id) ◦ π β ◦ Ψ−1 = (R ⊗ id) ◦ π α. Hence the restriction of R to M ⋊α G is a Herz -- Schur multiplier. (ii) =⇒ (i) Suppose S is a Herz -- Schur multiplier of α with symbol F . Then, for any a ∈ M, r ∈ G, φ ∈ L∞(G), using the equivalent of (1) for dual actions [18, Theorem 2.7], (S ⊗ id) ◦ π α(cid:0)πα(a)λr(IHM ⊗ φ)(cid:1) = (S ⊗ id)(cid:16)(cid:0)πα(a)λr ⊗ IL2(G)(cid:1)(cid:0)IHM ⊗ παG(φ)(cid:1)(cid:17) = (cid:16)πα(cid:0)F (r)(a)(cid:1)λr ⊗ IL2(G)(cid:17)(cid:0)IHM ⊗ παG(φ)(cid:1) = π α(cid:16)(cid:16)πα(cid:0)F (r)(a)(cid:1)λr(cid:17)(IHM ⊗ φ)(cid:17) = π α ◦ S(cid:0)πα(a)λr(IHM ⊗ φ)(cid:1), so the claim follows by linearity and continuity. If S is a module map then S is also a module map by Proposition 4.5. On the other hand, if S is a module map then S(cid:0)b · πα(a)λr(cid:1) = S(cid:0)(b ⊗ IL2(G)) · πα(a)λr(cid:1) = (b ⊗ IL2(G)) · S(cid:0)πα(a)λr(cid:1) = b · S(cid:0)πα(a)λr(cid:1), 10 A. MCKEE so S is also a module map. (cid:3) Remark 4.7. When M = C and α is trivial the above results recover the known fact [4] that a Schur multiplier S on B(L2(G)) restricts to a Herz -- Schur multiplier on vN(G) if and only if S commutes with the action Ad ρ (the second dual of the trivial action). In this classical case Lemma 4.4 states that every Schur multiplier of B(L2(G)) can be identified with a Herz -- Schur multiplier of αG. In this section we have been careful to keep track of multipliers which respect an additional module structure. The reason is that the Herz -- Schur multipliers of a semidirect product H ⋊ G have an obvious identification with Herz -- Schur multipliers of vN(H) ⋊G, and become A(H)-module maps under this identification. In the next section we will make use of multipliers respecting this extra module structure. 5. Abelian Groups We now assume that G is abelian, with dual group G. By Takai duality B(L2(G)) is isomorphic to the crossed product formed by the coaction δG dual to the trivial action of G on C, or the action αG dual to the trivial action of G on C. For a map S on B(L2(G)) we write SαG for the corresponding δG G, map on vN( G) ⋊ so for example SδG = Φ ◦ S ◦ Φ−1. for the corresponding map on vN(G) ⋊ αG G and SδG In [17, Section 6] we raised the question of how the Herz -- Schur multipliers of G acting on vN( G) are related to B(G) ⊙ B( G); note that the convolution multipliers considered there are precisely those appearing in (i) below. Theorem 5.1. Let S be a completely bounded, weak*-continuous map on B(L2(G)). The following are equivalent: i. SαG is a Herz -- Schur multiplier of αG and is an A( G)-module map; ii. SδG is a Herz -- Schur multiplier of δG and is an A(G)-module map. Moreover, the set of all S satisfying the equivalent conditions can be identi- fied with the space B( G × G). Proof. Observe that under the identifications of each crossed product with δG G (see (5)) is carried B(L2(G)) the module action · of A(G) on vN(G) ⋊ to the action ∗ on vN( G) ⋊ αG G of Remark 4.2(iv), and the corresponding statement holds for the module actions of A( G). Assume (i) holds, take u ∈ A(G), r ∈ G and γ ∈ G, and write mγ for the multiplication operator on L2(G) associated to γ ∈ G. Then SδG(cid:0)u · πδG (λr)mγ(cid:1) = S(cid:0)u(r)λrmγ(cid:1) = hγ, riS(cid:0)u(r)mγλr(cid:1) = hγ, riSαG(cid:0)u ∗ παG(mγ)λr(cid:1) = hγ, riu ∗ SαG(cid:0)παG(mγ)λr(cid:1) = hγ, ri hγ, ri u · SδG(cid:0)πδG(λr)mγ(cid:1) = u · SδG(cid:0)πδG(λr)mγ(cid:1), MULTIPLIERS AND DUALITY FOR GROUP ACTIONS 11 (v ∗ πδG(λr)mγ) = v ∗ SδG δG G. Similarly we calculate so S defines an A(G)-module map on vN(G) ⋊ (πδG(λr)mγ) for each v ∈ A( G), so SδG that SδG is a Herz -- Schur multiplier of the action α by Remark 4.2(iii). We have now shown that (i) implies (ii); by Pontryagin duality the same proof shows (ii) implies (i). Now let S satisfy (i); if F denotes the symbol of SαG then, for any r ∈ G, γ ∈ G, F (r) is an A( G)-module map, and therefore a Herz -- Schur multiplier of G, so we identify F with a map G × G → C. Consider the Schur multiplier SαG; it will be convenient to regard SαG as acting on L∞(G)⊗B(L2(G)). The restriction of SαG to L∞(G)⊗vN(G) is a completely bounded, weak*-continuous map; to see that it preserves L∞(G) ⊗ vN(G) we calculate, using the modularity of SαG, SαG(mγ ⊗ λr) = (1 ⊗ mγ−1)SαG(mγ ⊗ mγλr) = (1 ⊗ mγ−1)Φ−1(cid:0)(SαG = (1 ⊗ mγ−1)Φ−1(cid:0)F (γ, r)mγ ⊗ mγλr ⊗ λr)(cid:1) ⊗ id)(mγ ⊗ mγλr ⊗ λr)(cid:1) = F (γ, r)(mγ ⊗ λr). From this calculation we also see that this restriction is an A( G×G)-module map on L∞(G) ⊗ vN(G), and therefore a Herz -- Schur multiplier. Conversely, if S ∈ B( G × G), with symbol u : G × G → C, consider the associated Schur multiplier S acting on B(L2(G) ⊗ L2(G)). The restriction of S to L∞(G) ⋊ αG G is a Herz -- Schur multiplier of αG, since S (παG(mγ)λr) = S(mγ ⊗ mγλr) = (1 ⊗ mγ)S(mγ ⊗ λr) = u(γ, r)παG(mγ)λr. That S commutes with the A( G)-module action also follows easily. (cid:3) References [1] Claire Anantharaman-Delaroche, Syst`emes dynamiques non commutatifs et moyennabilit´e, Math. Ann. 279 (1987), no. 2, 297 -- 315. MR919508 [2] Erik B´edos and Roberto Conti, Fourier series and twisted C∗-crossed products, J. Fourier Anal. Appl. 21 (2015), no. 1, 32 -- 75. MR3302101 [3] Oscar Blasco and Ismael Garc´ıa-Bayona, Schur product with operator valued entries, 2019. Preprint. arXiv: 1804.03432 [math.FA]. [4] Marek Bozejko and Gero Fendler, Herz-Schur multipliers and completely bounded multipliers of the Fourier algebra of a locally compact group, Boll. Un. Mat. Ital. A (6) 3 (1984), no. 2, 297 -- 302. MR753889 [5] Michael Brannan, Approximation properties for locally compact quantum groups, Topological quantum groups, 2017, pp. 185 -- 232. MR3675051 [6] Nathanial P. Brown and Narutaka Ozawa, C ∗-algebras and finite-dimensional approx- imations, Graduate Studies in Mathematics, vol. 88, American Mathematical Society, Providence, RI, 2008. MR2391387 [7] Jean De Canni`ere and Uffe Haagerup, Multipliers of the Fourier algebras of some simple Lie groups and their discrete subgroups, Amer. J. Math. 107 (1985), no. 2, 455 -- 500. MR784292 12 A. MCKEE [8] Zhe Dong and Zhong-Jin Ruan, A Hilbert module approach to the Haagerup property, Integral Equations Operator Theory 73 (2012), no. 3, 431 -- 454. MR2945214 [9] Edward G. Effros and Zhong-Jin Ruan, Operator spaces, London Mathematical Soci- ety Monographs. New Series, vol. 23, The Clarendon Press, Oxford University Press, New York, 2000. MR1793753 [10] Alexander Grothendieck, R´esum´e de la th´eorie m´etrique des produits tensoriels topologiques, Resenhas 2 (1996), no. 4, 401 -- 480. Reprint of Bol. Soc. Mat. Sao Paulo 8 (1953), 1 -- 79 [ MR0094682 (20 #1194)]. MR1466414 [11] Uffe Haagerup, Decomposition of completely bounded maps on operator algebras. Un- published manuscript. [12] , An example of a nonnuclear C ∗-algebra, which has the metric approximation property, Invent. Math. 50 (1978/79), no. 3, 279 -- 293. [13] Carl Herz, Une g´en´eralisation de la notion de transform´ee de Fourier-Stieltjes, Ann. Inst. Fourier (Grenoble) 24 (1974), no. 3, xiii, 145 -- 157. MR0425511 [14] Paul Jolissaint, A characterization of completely bounded multipliers of Fourier alge- bras, Colloq. Math. 63 (1992), no. 2, 311 -- 313. MR1180643 [15] Marius Junge, Matthias Neufang, and Zhong-Jin Ruan, A representation theorem for locally compact quantum groups, Internat. J. Math. 20 (2009), no. 3, 377 -- 400. MR2500076 [16] Aristides Katavolos and Vern I. Paulsen, On the ranges of bimodule projections, Canad. Math. Bull. 48 (2005), no. 1, 97 -- 111. MR2118767 [17] Andrew McKee, Ivan G. Todorov, and Lyudmila Turowska, Herz -- Schur multipliers of dynamical systems, Adv. Math. 331 (2018), 387 -- 438. MR3804681 [18] Yoshiomi Nakagami and Masamichi Takesaki, Duality for crossed products of von Neumann algebras, Lecture Notes in Mathematics, vol. 731, Springer, Berlin, 1979. MR546058 [19] Gilles Pisier, Similarity problems and completely bounded maps, Lecture Notes in Mathematics, vol. 1618, Springer-Verlag, Berlin, 1996. MR1441076 [20] Florin Pop, Allan M. Sinclair, and Roger R. Smith, Norming C ∗-algebras by C ∗- subalgebras, J. Funct. Anal. 175 (2000), no. 1, 168 -- 196. MR1774855 [21] Masamichi Takesaki, Theory of operator algebras. I, Encyclopaedia of Mathematical Sciences, vol. 124, Springer-Verlag, Berlin, 2002. Reprint of the first (1979) edition, Operator Algebras and Non-commutative Geometry, 5. MR1873025 Department of Mathematical Sciences, Chalmers University of Technol- ogy and the University of Gothenburg, Gothenburg SE-412 96, Sweden E-mail address: [email protected]
1609.01097
1
1609
2016-09-05T11:04:19
$C^*$-simplicity and Ozawa conjecture for groupoid $C^*$-algebras, part I: injective envelopes
[ "math.OA" ]
This paper studies injective envelopes of groupoid dynamical systems and the corresponding boundaries. Analogue to the group case, we associate a bundle of compact Hausdorff spaces to any (discrete) groupoid (the Hamana boundary of the groupoid). We study bundles of compact topological spaces equipped with an action of a groupoid. We show that any groupoid has a minimal boundary (the Furstenberg boundary of the groupoid). We prove that the Hamana and Furstenberg boundaries are the same, for (discrete) groupoids. We find the relation between the reduced crossed product of the $\G$-injective envelop of a groupoid dynamical system and the injective envelope of the reduced crossed product of the original system.
math.OA
math
C ∗-simplicity and Ozawa conjecture for groupoid C ∗-algebras, part I: injective envelopes Massoud Amini, Farid Behrouzi Department of Mathematics, Faculty of Mathematical Sciences, Tarbiat Modares School of Mathematics, Institute for Research in Fundamental Sciences (IPM) University, Tehran 14115-134, Iran Tehran 19395-5746, Iran [email protected], [email protected] Faculty of Mathematics, Alzahra University, Vanak, Tehran 19938-91176, Iran f [email protected] Abstract This paper studies injective envelopes of groupoid dynamical systems and the corresponding boundaries. Analogue to the group case, we associate a bundle of compact Hausdorff spaces to any (discrete) groupoid (the Hamana boundary of the groupoid). We study bundles of compact topological spaces equipped with an action of a groupoid. We show that any groupoid has a minimal boundary (the Furstenberg boundary of the groupoid). We prove that the Hamana and Furstenberg boundaries are the same, for (discrete) groupoids. We find the relation between the reduced crossed product of the G-injective envelop of a groupoid dynamical system and the injective envelope of the reduced crossed product of the original system. Keywords: groupoid dynamical system, injective envelop, boundary 2008 MSC: 46l35, 20F65 The first author was partly supported by grants from IPM (94430215) Contents 1 Introduction 2 Groupoid dynamical systems 2 4 Preprint submitted to Documenta Mathematica August 3, 2018 3 Injective envelopes and Hamana boundary 4 Furstenberg boundary 5 The reduced crossed product and its injective envelope 6 13 18 1. Introduction Let G be a discrete group and V be an operator system, i.e., a unital self- adjoint closed subspsace of a unital C*-algebra. We say that V is a G-module (G-operator system) if there is a homomorphism α from G into the group of all unital complete order isomorphisms of V . In this case, we say that the triple (G, V, α) is a (group) dynamical system. Hamana in [8] studied injectivity in the category whose objects are G-operator systems and whose morphisms are completely positive unital G-homomorphisms. In this setting, he proved that every G-operator system V has a unique G-injective envelope IG(V ), i.e., a minimal G-injective G-operator system containing V as sub- G-operator system. Hamana obtained the G-injective envelope of V by first embedding V into a G-injective operator system W . He then obtained a minimal V -projection of W and proved that the G-injective envelope is the rang of this projection. For C with the trivial action G, the G-injective envelope of C is C(X) for a compact Hausdorff space X [6]. Moreover, X is a G-space, called the Hamana boundary of G, denoted by ∂HG. It is proved in [9] that the action of G on ∂H G is a boundary action, i.e., it is minimal and strongly proximal (an action is minimal if it has dense orbits, and strongly proximal if for any probability measure µ on X, the weak*-closure of the orbit G · µ contains a point mass; see [4, 5] for more details). On the other hand, Furstenberg in [4] proved that for any discrete group G, there is a unique (up to G -isomorphism) maximal G-boundary ∂F G. Maximality here means that every G-boundary is a quotient of ∂F G. This is called the Furstenberg bounadry of G. Kalantar and Kennedy in [9] proved that for a discrete group G, the Furstenberg and Hamana boundaries are G-isomorphic. They also relate this to the notion of exactness of groups, introduced by Kirchberg and Wasserman [11]. Ozawa proved in [13, Theorem 3] that a discrete group is exact if and only it acts amenably on its Stone- Cech compactification, and authors in [9] show the same for the action on the Furstenberg boundary. More generally, Wasserman [16] showed that a C ∗-algebra is exact if it can be embedded into a nuclear C ∗-algebra (the 2 converse is also true by a result of Kirchberg [10]). Ozawa conjectured in [13] that for an exact C ∗-algebra A, there is a nuclear C ∗-algebra between A and its injective envelope. One of the main objectives of [9] was to prove this for reduced C ∗-algebras of discrete exact groups. Along the way, they also contributed to the C ∗-simplicity problem by showing that the reduced crossed product of C(∂F G) by the canonical action of G is simple if and only if the action is topologically free. This paper seeks an appropriate extension of these notions and results to (discrete) groupoids. As far as we know, none of the above results is explored for groupoids. The motivation of the paper is two folds. First we want to introduce appropriate notions of G-boundary for groupoids and show that the Hamana and Furstenberg boundaries are the same (Theorem 4.9). Also, we would like to make tools for checking the C ∗-simplicity and Ozawa conjecture for groupoid C ∗-algebras (and crossed products), something which is pursued in a forthcoming paper [1], in which the nuclearity of crossed products under exact groupoid actions is studied. The paper is organized as follows. In Section 2, we introduce basic notions of groupoids and groupoid dynamical systems. For an r-discrete groupoid G, we describe notions such as G-essentiality, G-rigidity, G-injectivity and prove the existence and uniqueness of the injective envelope of groupoid dy- namical systems. In Section 3, we proved that for any groupoid dynamical system A, there is a minimal injective dynamical system IG(A) "containing" A. We prove this by showing that every groupoid dynamical system A can be embedded into an injective dynamical system, and then find a minimal A-projection with IG(A) as its range. As an important example, the injec- tive envelope of the trivial groupoid dynamical system with one dimensional fibers, gives a bundle {Xu}u∈G(0) of compact Hausdorff spaces, called the Hamana boundary of G. In section 4, we consider an r-discrete groupoid G acting on a bundle of compact Hausdorff spaces over the unit space G(0) of G, and study minimality, strong proximality and boundary actions in this case. We show that there is a unique maximal G-boundary, called the Fursten- berg boundary. We show that the Hamana boundary is G-isomorphic to the Furstenberg boundary. We also study the (reduced) crossed products of the groupoid dynamical systems and show that the reduced crossed product of the G-injective envelop of such a system is included in the injective envelope of the reduced crossed product. This is essential in [1], where we want to prove the Ozawa conjecture for groupoid crossed products. 3 2. Groupoid dynamical systems We review basic facts on groupoids. For more details we refer the reader to [12, 14, 15]. Definition 2.1. A groupoid is a set G endowed with a product map: G2 −→ G; (g, h) 7→ gh, where G2 is a subset of G × G, called the set of composable pairs, and an inverse map: G −→ G; g 7→ g−1 such that 1. (g−1)−1 = g, 2. if (g, h) ∈ G2 and (h, k) ∈ G2, then (gh, k), (g, hk) ∈ G2 and (gh)k = g(hk), 3. (g−1, g) ∈ G2 and if (g, h) ∈ G2, then g−1(gh) = h, 4. (g, g−1) ∈ G2 and if (h, g) ∈ G2, then (hg)g−1 = h, for each f, g, h ∈ G. The unit space G0 is the subset of elements gg−1, where g ranges over G. The range and source maps r : G −→ G0 and d : G −→ G0 are defined by r(g) = gg−1 and d(g) = g−1g. A pair (g, h) belongs to G2 if and only if d(g) = r(h). For each u ∈ G0, the subsets Gu and Gu are defined by Gu = d−1({u}) and Gu = r−1({u}). An operator system is a closed, self-adjoint subspace of a unital C ∗- algebra containing its unit, or equivalently, a closed, self-adjoint subspace of B(H) containing the identity operator on the Hilbert space H. In the latter case, we say that A is an operator system on H (see [3] for more details). Definition 2.2. A groupoid dynamical system is a triple A = (G, {Au}u∈G(0), α), such that 1. G is a groupoid, 2. for each u ∈ G(0), Au is an operator system, 3. for each g ∈ G, αg : Ad(g) −→ Ar(g) is a complete order isomorphism, 4. for each (g, h) ∈ G(2), αgαh = αgh, for g ∈ G and a ∈ Ad(g), we write g · a for αg(a). 4 The following definition uses the notion of completely positive (c.p.) maps between C ∗-algebras. The notion is also meaningful for maps between opera- tor systems (see [3] for more details). We assume that all completely positive maps are unital. Definition 2.3. A G-morphism between systems A = (G, {Au}u∈G(0), α) and B = (G, {Bu}u∈G(0), β) is a family {ϕu}u∈G(0) of maps such that 1. for any u, ϕu : Au −→ Bu is a c.p. map, 2. for any g ∈ G and a ∈ Gd(g), βg(ϕd(g)(a)) = ϕr(g)(αg(a)), i.e., for each g, the following diagram is commutative: Ad(g) αg Ar(g) ϕd(g) Bd(g) . βg ϕr(g) / Br(g) The composition of G-morphisms {ϕu}u∈G(0) and {ψu}u∈G(0) (if it makes sense) is defined by {ψu}u∈G(0) ◦ {ϕu}u∈G(0) = {ψu ◦ ϕu}u∈G(0). Let A = (G, {Au}u∈G(0), α), B = (G, {Bu}u∈G(0), β), C = (G, {Cu}u∈G(0), γ), and W = (G, {Wu}u∈G(0), ω) be dynamical systems and Φ = {ϕu}u∈G(0), Ψ = {ψu}u∈G(0), and Θ = {θu}u∈G(0) be G-morphisms. A G-morphism Φ : A −→ B is a G-injection (resp., a G-isomorphism) if for any u, ϕu : Au −→ Bu is an injection (resp., a complete order isomorphism). Definition 2.4. A G-extension of a groupoid dynamical system A is a pair (W, Θ) such that Θ : A −→ W is a G-embedding. Moreover, 1. (W, Θ) is G-essential if for any G-morphism Φ : W −→ C, Φ is a G-injection whenever Φ ◦ Θ is a G-injection, 2. (W, Θ) is G-rigid if for any G-morphism Φ : W −→ W, Φ ◦ Θ = Θ implies ϕu = idWu, for all u ∈ G(0). 5 / /     / Definition 2.5. A groupoid dynamical system W is called G-injective if for systems A and B and G-injective morphism Φ : A −→ B and arbitrary G- morphism Ψ : A −→ W, there exists a G-morphism Θ : B −→ W such that Θ ◦ Φ = Ψ, that is, for any u ∈ G(0), the following diagram is commutative: Au ϕu / Bu ψu θu Wu. 3. Injective envelopes and Hamana boundary In this section we explore the existence and uniqueness of injective objects in the category of groupoid dynamical systems. An operator system is a closed, self-adjoint subspace of a unital C ∗-algebra containing its unit, or equivalently, a closed, self-adjoint subspace of B(H) containing the identity operator on the Hilbert space H. In the latter case, we say that A is an operator system on H (see [3] for more details). The injective envelopes of operator systems are studied by Hamana [7]. Let X, X0 be sets and s : X −→ X0 be a surjective function. Let A = {Au}u∈X0 be a family of operator spaces. A section of A is a function f : X −→Sx∈X0 such that f (x) ∈ As(x) and kf k∞ = sup x∈X kf (x)k < ∞. We denote the set of all sections of A by Γ∞(X, s∗A). The space Γ∞(X, s∗A) of sections form an is an operator system with pointwise operations and involution and the above norm. We denote the set of functions with finite support in Γ∞(X, s∗A) by Γc(X, s∗A). This is a *-subspace of Γ∞(X, s∗A). If each Au is a C ∗-algebra, then Γ∞(X, s∗A) is a C ∗-algebra and Γc(X, s∗A) is a *-subalgebra. Lemma 3.1. If A = {Au}u∈X0 is a family of injective operator systems and s : X −→ X0 is surjective, then Γ∞(X, s∗A) is an injective operator system. Proof. Let B and C be operator spaces and θ : B −→ C be an injective com- pletely positive map. To each completely positive map ϕ : B −→ Γ∞(X, s∗A) and x ∈ X we associate the map ϕx : B −→ As(x) by ϕx(b) = ϕ(b)(x). By injectivity of As(x), there exists a completely positive map ψx : C −→ As(x) 6 /   } } such that ψx ◦ θ = ϕx. Define ψ : C −→ Γ∞(X, s∗A) by ψ(c)(x) = ψx(c), for c ∈ C and x ∈ X. Then ψ ◦ θ(b)(x) = ψx(θ(b)) = ϕx(b) = ϕ(b)(x), for b ∈ B and x ∈ X. Suppose that G is a groupoid and A = {Au}u∈G(0) is a family of operator systems. Then (G, {Γ∞(Gu, s∗A)}u∈G(0), ℓ) becomes a groupoid dynamical system with the action ℓg : Γ∞(Gd(g), s∗A) −→ Γ∞(Gr(g), s∗A), For u ∈ G(0), define ℓg(f )(x) = f (g−1x). Imu : Au −→ Γ∞(Gu, s∗A) by Imu(a)(g) = g−1.a. Then {Imu}u∈G(0) is an injective G-morphism. The next lemma extends [8, Lemma 2.2]. Lemma 3.2. If A = {Au}u∈G(0) is a family of injective operator systems, then is a G-injective groupoid dynamical system. (G, {Γ∞(Gu, s∗A)}u∈G(0), ℓ) Proof. Let Θ : B −→ C be a G-injective morphism and let ϕ : B −→ (G, {Γ∞(Gu, s∗A)}u∈G(0), ℓ) be a G -morphism. For any u ∈ G(0), define bϕu : Bu −→ Γ∞(Gu, s∗A) by bϕu(b) = ϕu(a)(u). By the injectivity of A , there exists a completely positive map bψu : Cu −→ Au such that bψu ◦ θu = bϕu. Define a completely positive map ψu : Cu −→ Γ∞(Gu, s∗A)} by ψu(b)(g) = bψd(g)(g−1.b). For any u ∈ G(0) , g ∈ Gu and b ∈ Bu, we have ψu ◦ θu(b)(g) = bψd(g)(g−1.θu(b)) = bψd(g)(θd(g)(g−1.b)) = bϕd(g)(g−1.b) = ϕd(g)(g−1.b)(d(g)) = (g−1.ϕu)(d(g)) = ϕu(b)(gd(g)) = ϕu(b)(g). 7 Thus {ψu}u∈G(0) ◦ {θu}u∈G(0) = {ϕu}u∈G(0). We show that {ψu}u∈G(0) G-morphism: is a g.ψd(g)(c)(h) = ψd(g)(c)(g−1h) = bψd(h)(h−1.g.c) = ψr(g)(g.c)(h), for g ∈ G , h ∈ Gr(g) and c ∈ Cd(g). Thus g.bϕd(g)(c) = bϕr(g)(g.c). The next result follows from the above two lemmas and some routine algebraic manipulations. Proposition 3.3. Let A be groupoid dynamical system. Then A is injective if and only if for each u ∈ G(0), Au is an injective operator system and there exists a G-morphism {ϕu}u∈G(0) : (G, {Γ∞(Gu, s∗A)}u∈G(0), ℓ) −→ A such that ϕu ◦ Imu = IdAu, for any u ∈ G(0). We say that A is a G-dynamical subsystem of B, if for any u ∈ G(0), Au ⊆ Bu, and for any g ∈ G, βgAd(g) = αg. Definition 3.4. Let A be a G dynamical subsystem of B. 1. A G-morphism Φ : A −→ B is called an A-projection if for any u, φu ◦ φu = φu and φuAu = idAu, 2. A family {pu}u∈G(0) of seminorms is called an A-seminorm if there exists a G-morphism {ϕu}u∈G(0) such that, for any u, pu(.) = kφu(.)k and φuAu = idAu. Definition 3.5. Let P (resp. Pr) be the set of all A-seminorms (resp., all A-projections) on B. We define partial orders on P and Pr as follows: {pu}u∈G(0) ≤ {qu}u∈G(0), if for any u ∈ G(0) and b ∈ Bu, pu(b) ≤ qu(b), and {ϕu}u∈G(0) (cid:22) {ψu}u∈G(0), if for any u ∈ G(0), ψu ◦ φu = φu ◦ ψu = φu. The next lemma extends [7, Lemma 3.4]. Lemma 3.6. Let A be a G-dynamical subsystem of an injective groupoid dynamical system B. Then there exists a minimal A-seminorm on B. 8 Proof. Since {idBu}u∈G(0) induces an A-seminorm on B, by Zorn lemma, it is enough to show that every decreasing net of A-seminorms on B has a lower bound. Suppose that {{pi,u}u∈G(0)}i∈I is such a decreasing net. For any i ∈ I, there exists a G-morphism {ϕi,u}u∈G(0) such that, for any u ∈ G(0), pi,u(.) = kϕi,u(.)k and ϕi,uAu = idAu. Put B = ⊕u∈G(0)Bu and let H be a Hilbert space such that B ⊆ B(H). Define J : B −→ ℓ∞(G, B(H)) by J((bu)u∈G(0))(g) = (bu)u∈G(0), where bu =  bu u 6= d(g) g−1 · br(g) u = d(g). Then J is an imbedding, and we may regard B as an operator subsystem of ℓ∞(G, B(H)). The restriction of J to Bu is the imbedding Ju : Bu −→ ℓ∞(Gu, B(H)); Ju(b)(g) = g−1 · b. For any i ∈ I, define ϕi : B −→ B ⊆ ℓ∞(G, B(H)) by ϕi(bu)u∈G(0) = (ϕi,u(bu))u∈G(0). Then {ϕi}i∈I is a net in the unit ball of B(B, ℓ∞(G, B(H))), which is compact in the point-weak∗ topology, thus there exists a subnet {ϕj}j∈I ′, point-weak∗- converging to some ϕ0 in B(B, ℓ∞(G, B(H))), that is, for any v ∈ G(0) and b ∈ Bv, ϕj,v(b) −→ ϕ0((bu(b))u∈G(0))(v), where bu(b) =  b 0 u = v u 6= v. By the injectivity of B, there exists a G-morphism Ψ : (G, ℓ∞(Gu, B(H)), ℓ) −→ B such that for any u ∈ G(0), ψu ◦ Ju = idBu. Define ϕ0,v : Bv −→ ℓ∞(Gv, B(H)) by ϕ0,v(b) = ϕ0((bu(b))u∈G(0))Gv and ϕv = ψv ◦ ϕ0,v. Then ϕv is a c.p. map from Bv into Bv and ϕvBv = idBv. Let us observe that Φ is a G-morphism. Since Ψ is a G-morphism, it is enough to show that 9 {ϕ0,u}u∈G(0) is a family of G-morphisms. To see this, suppose that b ∈ Bd(g), for some g ∈ G. If h ∈ Gr(g), then g · ϕ0,d(g)(b)(h) = ϕ0((bu(b))u∈G(0))(g−1h) = lim j ϕj((bu(b))u∈G(0))(g−1h) = lim j ϕj,d(g)(b)(g−1h) = lim j h−1 · g · ϕj,d(g)(b) = lim j h−1 · ϕj,r(g)(g · b) = h−1 · ϕ0((bu(g.b))u∈G(0))(r(g)) = ϕ0((bu(g · b))u∈G(0))(hr(g)) = ϕ0((bu(g · b))u∈G(0))(h) = ϕ0,r(g)(g · b)(h). Put pu(·) = kϕu(·)k, tnen {pu}u∈G(0) is a lower bound for {{pi,u}u∈G(0)}i∈I. For b ∈ Bv, pv(b) = kϕv(b)k = kψv ◦ ϕ0,v(b)k ≤ kϕ0,v(b)k = kϕ0((bu(b))u∈G(0))k ≤ lim sup j kϕj((bu(b))u∈G(0))k = lim sup j kϕj,v(b)k = lim i pi,v(b). Now we are able to extend [7, Lemma 3.5]. Theorem 3.7. Let A be a G-dynamical subsystem of B and B is G-injective. Then there is a minimal A-projection on B. Proof. By Lemma 3.6, there exists a minimal A-seminorm {pu}u∈G(0). Thus, for any u ∈ G(0), the exists eϕu : Bu −→ Bu such that eϕuAu = idAu and pu(.) = keϕu(.)k. Let ϕ(n) u = 1 n (eϕu +fϕu 2 + · · · +fϕu n). u }u∈G(0)}j∈N and a G-morphism Φ such that ϕ(nj ) Then {{ϕ(n) u }u∈G(0)}n∈N is a net of G-morphisms from B into itself. A similar argument as in the proof of Lemma 3.6 shows that there exist a subnet {{ϕ(nj) (b) −→ ϕu(b), for all u ∈ G(0) and b ∈ Bu, in the weak∗-topology. Take a G-morphism Ψ which is an idempotent from B into (G, ℓ∞(Gu, B(H)), ℓ), where H is a Hilbert space with ⊕uBu ⊆ B(H). For u ∈ G(0), u kψu ◦ ϕu(b)k ≤ kϕu(b)k ≤ lim sup j 10 kϕ(nj) u (b)k ≤ keϕu(b)k = pu(b). By the minimality of {pu}u∈G(0), kψu ◦ ϕu(b)k = pu(b), thus Therefore, lim sup j kϕ(nj) u (b)k = keϕu(b)k. keϕu(x) − eϕ2 u(x)k = keϕu(x − eϕu)k = lim sup kϕ(nj )(x − eϕu(x)k keϕ(x) − eϕ nj+1 = lim sup 1 n u (x)k = 0. HenceeΦ = {eϕu}u∈G(0) is an A-projection. To see the minimality ofeΦ, suppose that Θ is any A-projection with Θ (cid:22) eΦ. Then, for u ∈ G(0), θu◦eϕu = eϕu◦θu = θu. Thus kθu(b)k ≤ keϕu(b)k = pu(b). The minimality of {pu}u∈G(0) implies that, for any u, kθu(b)k = keϕu(b)k, in particular, ker θu = kereϕu. For b ∈ Bu, eϕu(b) = eϕu((b − θu(b)) + θu(b)) = eϕ(θu(b)) = θu(b). The two next lemmas are proved similar to Lemma 3.11 and [7, Lemma 3.6]. Lemma 3.8. Let A be a groupoid G-dynamical subsystem of B and Φ : B −→ B be a G-morphism which induces a minimal A-seminorm. Then the extension IM(Φ) = ((G, {ϕu(Bu)}u∈G(0), β), {iu}u∈G(0)), is G-rigid, where iu : Au −→ ϕu(Bu) is inclusion map. Lemma 3.9. Let (B, Φ) be a G-injective G-extension of A. Then (B, Φ) is G-rigid if and only if it is G-essential. Lemma 3.10. Let A be an injective groupoid dynamical system and Φ be an idempotent G-morphism of A. Then (G, {ϕu(Au)}u∈G(0), α) is injective. Proof. Let B and C be groupoid dynamical systems, Ψ : B −→ C be a G- injective morphism and Θ : B −→ (G, {ϕu(Au)}u∈G(0), α) be a G-morphism. Suppose that {iu}u∈G(0) : (G, {ϕu(Au)}u∈G(0), α) −→ A is the inclusion mor- phism. Since A is injective, there exists a G-morphism { ψu}u∈G(0) from C into A such that for any u, ψu ◦ ψu = iu ◦ θu. 11 Hence, {ϕu ◦ ψu}u∈G(0) is a G-morphism from (G, {Cu}u∈G(0), γ) into A such that, for any u, (ϕu ◦ ψu) ◦ ψu = ϕu ◦ ( ψu ◦ ψu) = ϕu ◦ (iu ◦ θu) = θu. We are ready to prove the main result of this section. Theorem 3.11. Any groupoid dynamical system A has a G-injective enve- lope (IG(A), Υ), which is unique up to G-isomorphism. Proof. Let H be a Hilbert space with ⊕u∈G(0)Au ⊆ B(H), and put W = (G, {ℓ∞(Gu, B(H))}u∈G(0), ℓ), {u}u∈G(0)). For u ∈ G(0), define Ju : Au −→ ℓ∞(Gu, B(H)) by Ju(a)(g) = g−1 · a. We may regard A as a G-dynamical subsystem of W. By Theorem 3.11, there exists a minimal A-projection Θ on W, and IG(A) = (G, {θu(ℓ∞(Gu, B(H)))}u∈G(0), ℓ) is injective by Lemma 3.10. Suppose that iu : Au −→ θu(ℓ∞(Gu, B(H))) is the inclusion map and Υ = {iu}u∈G(0). Then (IG(A), Υ) is a G-injective envelope of A, by Lemma 3.8. Now if (B, Φ) is any other G-injective envelop of A, then there exist G-morphisms Ψ from IG(A) into B and { ψu}u∈G(0) from B into IG(A) such that, for u ∈ G(0), ψu ◦ iu = ϕu and ψu ◦ ϕu = iu, hence ψu◦ψu◦ϕu = ϕu and ϕ◦ψu◦iu = iu. By the rigidity, ψu◦ψu = idθ(ℓ∞((Gu,B(H)))) and ψu ◦ ψu = idBu. Our next step is to find an analog for the Hamana boundary. We first need the following result. Proposition 3.12. Let (G, {Au}u∈G(0), α) be a G-injective groupoid dynam- ical system. Then for any u ∈ G(0), there is a unique multiplication · : Au × Au −→ Au making Au a C ∗-algebra in its given ∗-operation and norm, and for any g ∈ G, αg : Ad(g) −→ Ar(g) is an isomorphism of C ∗-algebras. Moreover, if for any u ∈ G(0), Au is an operator system in a commutative C ∗-algebra, then under this multiplication, each Au becomes a commutative C ∗-algebra. 12 Proof. Similar to the proof of Lemma 3.6, there is a Hilbert space H such that W = (G, {ℓ∞(Gu, B(H))}u∈G(0), ℓ) has (G, {Au}u∈G(0), α) as a G-dynamical subsystem. By injectivity, there is a G-morphism φu : W −→ A such that, for any u ∈ G(0), φu is c.p. and φuAu = idAu. Given x, y ∈ Au = φu(ℓ∞(Gu, B(H))), put x ◦ y = φu(xy). By [3, Theorem 6.1.3.], this operation defines a multiplication on Au, making Au a C*-algebra. For g ∈ G and x, y ∈ Ad(g), g · (x ◦ y) = g · φd(g)(xy) = φr(g)(g · (xy)) = φr(g)((g · x)(g · y)) = (g · x) ◦ (g · y). Moreover, if for any u ∈ G(0), there is a compact Hausdorff space Xu such that Au ⊆ C(Xu), put X =Fu Xu, then we may regard C(Xu) as a C ∗-subalgebra of ℓ∞(X), and (G, {Au}u∈G(0), α) is a subsystem of (G, {ℓ∞(Gu, ℓ∞(X))}, ℓ). If we define the multiplication on Au as above, then Au is a commutative C ∗-algebra. Corollary 3.13. Let (G, {Bu}u∈G(0), β) be the injective envelope of a groupoid dynamical system (G, {Au}u∈G(0), α) such that, for any u ∈ G(0), Au is an operator system in a commutative C ∗-algebra. Then, for any u ∈ G(0), Bu is commutative C ∗-algebra. H is a compact Hausdorff space. We call {∂u The bundle with one dimensional fibres is of special interest. Let G be a groupoid and for u ∈ G(0), set Cu = C. Then (G, {Cu}u∈G(0), γ) is a groupoid dynamical system where γ is the trivial action, that is, γg : Cd(g) −→ Cr(g) is the identity. Then the injective envelope of C is of the form (G, C(∂u H), β), where ∂u H}u∈G(0) the Hamana boundary of G and denote it by ∂H (G). For any g ∈ G, γg induces a home- omorphism γ∗ H ), γg(f )(x) = f (γ∗ g−1 ·x). The groupoid dynamical system C is G-injective if and only if there are states φu : ℓ∞(Gu) −→ C such that for any g ∈ G and f ∈ ℓ∞(Gd(g)), φr(g)(g · f ) = φd(g)(f ). such that, for any f ∈ C(∂d(g) g : ∂r(g) H −→ ∂d(g) H 4. Furstenberg boundary The notion of groupoid action on sets (or topological spaces) generalizes the concept of group action by considering partially defined maps. 13 Definition 4.1. Let G be a groupoid. A G-space is a bundle of locally compact Hausdorff spaces X = {Xu}u∈G(0) and a bundle of maps {αg}g∈G such that 1. for any g ∈ G, αg is a homeomorphism from Xd(g) onto Xr(g), 2. for any u ∈ G(0), αu is the identity map idXu, 3. for any (g, h) ∈ G(2), αg ◦ αh = αgh. We denote αg(x) by g·x. A G-subspace of X is a bundle of locally compact Hausdorff spaces Y = {Yu}u∈G(0) such that, for each u, Yu ⊆ Xu and, for each g, the restriction of αg to Yd(g) is a homeomorphism onto Yr(g). For the rest of this section, all topological spaces are assumed to be com- pact and Hausdorff. Definition 4.2. A G-map between G-spaces X and Y is a family of maps {φg}g∈G such that 1. for any g ∈ G, φg : Xd(g) −→ Yd(g) is continuous, 2. for any g ∈ G and x ∈ Xd(g), g · φd(g) = φr(g)(g · x). We denote the space of complex finite Radon measures on X by M(X) and the subset of probability measures by P (X), equipped with the weak∗- topology. There is natural embedding of X into P (X) as point masses. If X = {Xu}u∈G(0) is a G-space, then P(X ) = {P (Xu)}u∈G(0) is a G-space. For g ∈ G and µ ∈ P (Xd(g)) define g · µ(E) = µ(g−1 · E), for Borel subsets E of Xr(g). Let X be a G-space. Then for any g ∈ G, the map x 7→ g · x is a homeomorphism from Xd(g) onto Xr(g). This induces an ∗-isomorphism αg : C(Xd(g)) −→ C(Xr(g)); αg(f )(x) = f (g−1 · x), and (G, {C(Xu)}, α) is a groupoid dynamical system. Conversely, given the groupoid dynamical system (G, {C(Xu)}, α), for g ∈ G, αg : C(Xd(g)) −→ C(Xr(g)) is an *-isomorphism, and by Banach-Stone theorem, there exists a homeomorphism αg : Xr(g) −→ Xd(g) such that, for f ∈ C(Xd(g)), αg(f )(x) = f (eαg(x)). For {Xu}u∈G(0) is a G-space. g : Xd(g) −→ Xr(g); α∗ α∗ g(x) = αg−1(x), Let X and Y be G-spaces. There is a one-to-one correspondence between G-morphisms Φ : (G, {C(Xu)}u∈G(0), α) −→ (G, {C(Yu)}u∈G(0), β) and G-maps 14 Φ∗ : {Yu}u∈G(0) −→ {P (Xu)}u∈G(0), given by φg(f )(y) = φ∗ restriction of the adjoint map φ∗ P (Xd(g)). g(y)(f ). Here, the g to Yd(g) is a continuous map from Yd(g) into Definition 4.3. A G-space X is called minimal if there is no nontrivial G- subspace, and strongly proximal if for every u, v ∈ G(0), with Gv u 6= ∅, and µ ∈ P (Xu), Gv u.µ ∩ Xv 6= ∅. A compact G-space X is called a G-boundary if it is minimal and strongly proximal, or equivalently, if X is the unique minimal closed G-subspace of P(X ). By Zorn lemma, every G-space has a minimal G-subspace. Also, every G-subspace of a strongly proximal G-space is again strongly proximal. u}u∈G(0)}i∈I be a family of G-spaces. The product space u }u∈G(0) is a G-space with the diagonal G-action. Lemma 4.4. If {Xi}i∈I is a family of compact strongly proximal G-spaces, w∗ P =\F F , P 15 Let {X i = {X i Qi∈I X i = { Qi∈I X i then Qi∈I Xi is also strongly proximal. Proof. For the case where I is finite, it suffices to prove check the claim when I has two elements. Let X = {Xu}u∈G(0) and Y = {Yu}u∈G(0) be two strongly proximal G-spaces. Let us show that X × Y = {Xu × Yu}u∈G(0) is strongly proximal. Define Λu : P (Xu × Yu) −→ P (Xu) by Λu(µ)(E) = µ(E × Yu). Take u, v ∈ G(0) with Gv u 6= ∅ and µ ∈ P (Xu × Yu). It is easy to see that u · Λu(µ). Since X is strongly proximal, there exists x ∈ Xv Λu(Gv u · µ). An straightforward measure theory argument such that δx ∈ Λu(Gv shows that there exists ν ∈ P (Xv) such that δx × ν ∈ Gv u · µ. Since Y is strongly proximal, there exists a net {gi} in Gv v and y ∈ Xv such that gi · ν −→ y. By compactness, we may assume that there is x′ ∈ Xv such that gi · x −→ x′. Therefore, δx′ ⊗ δy ∈ Gv u · µ) = Gv u · Λu(µ) ⊆ Gv u · Λu(µ). v · Gv For the general case, we need to use the idea of functions depending on finitely many variables. More precisely, let {Xi}i∈I be a family of com- pact Hausdorff spaces. For any finite subset F ⊆ I, let CF be the set of dexed by F , i.e., f ∈ CF if and only if f ((xi)i∈I) = f ((yi)i∈I), whenever all continuous functions in C(Qi∈I Xi) that depend only on variables in- (xi)i∈I, (yi)i∈I ∈ Qi∈I Xi with xi = yi, for all i ∈ F . By Stone-Weierstrass theorem, SF CF is dense in C(Qi∈I Xi), where the union is taken over all finite subsets of I. Therefore, if P ⊆ P (Qi∈I Xi), then F where P is the closure of P in the weak topology on P (Qi∈I Xi) induced by CF . Let F ⊆ I be a finite subset and let (xi)i∈I\F ∈ Qi∈I\F Xi. µ ∈ P (Qi∈I Xi), then there exists µF ∈Qi∈F Xi such that, for all f ∈ CF , If ZQi∈I Xi f dµ =ZQi∈I Xi f d(µF × δ(xi)i∈I\F ). Next, if I is an arbitrary set and u, v ∈ G(0) such that Gv u 6= ∅, and u), by the above observation, for any finite subset F ⊆ I, v 6= ∅, by the Cantor v 6= ∅ and hence Gv u · µ µ ∈ P (Qi∈I X i F TQi∈I X i Gv intersection theorem. u · µ w∗TQi∈I X i Now we could extend [4, Proposition 4.2]. Lemma 4.5. Let X be a G-boundary and Y be a minimal compact G-space. Then every continuous G-map {φu}u∈G(0) from Y into P(X ) has X as its range, i.e., for all u ∈ G(0), φu(Yu) = Xu. Equivalently, every G-morphism from the groupoid dynamical system {C(Xu)}u∈G(0) into the groupoid dynam- ical system {C(Yu)}u∈G(0) is an ∗-isometric G-morphism. Moreover, there is at last one such map. Proof. Let Φ = {φu}u∈G(0) : Y −→ P(X ) be a G-map. The G-subspace {φu(Yu)}u∈G(0) of P(X ) contains X . Since Y is minimal, the G-subspace {φ−1 u (Xu)}u∈G(0) coincides with Y. Therefore, for any u ∈ G(0), φu(Yu) = Xu and the G-morphism Φ from {C(Xu)}u∈G(0) into {C(Yu)}u∈G(0) is a G-isometry. If there are two such maps Φ and Ψ, then {(φu + ψu)/2} is also a G-map and hence has X as its range. Since δx is an extreme point of P (Xu), for any u, φu = ψu on Xu. Definition 4.6. The Furstenberg boundary ∂F G is a G-boundary which is universal in the sense that it has every G-boundary as a G-quotient. Such a maximal G-boundary exists: Take the family {Xi}i∈I of all G- boundaries (up to G-isomorphism). By an argument similar to the one in the group case [4], one can show that this forms a set, and we could consider Cartesian products. By Lemma 4.4, Qi∈I Xi is strongly proximal. Suppose that ∂F G is a minimal G-subspace of Qi∈I Xi, which exists by Zorn lemma. Since every G-subspace of a strongly proximal G-space is strongly proximal, ∂F G is a G-boundary and every G-boundary is a quotient of ∂F G. Also, by Lemma 4.5, such a maximal G-boundary is unique. 16 w w) if Gu 6= ∅ if and only if Gr(g) Let X = {Xu}u∈G(0) be a G-space and w ∈ G(0). For u ∈ G(0), put Su = ℓ∞(Gu w 6= ∅, and Su = C(Xu), otherwise. Note that, for g ∈ G, Gd(g) 6= ∅. Define αg : C(Xd(g)) −→ C(Xr(g)) by αg(f )(x) = f (g−1 · x), for f ∈ C(Xd(g)) and x ∈ Xr(g), and αg : ℓ∞(Gd(g) w ) −→ ℓ∞(Gr(g) w . Then (G, {Su}u∈G(0), α) is a groupoid dynamical system. For µ ∈ P (Xw), define P u w ) by αg(f )(h) = f (g−1h) for f ∈ ℓ∞(Gd(g) w ) and h ∈ Gr(g) µ : C(Xu) −→ ℓ∞(Gu w) by w µ (f )(g) =ZXw P u f (g · x) dµ(x), if Gu w 6= ∅, and by P u Take g ∈ G with Gd(g) w µ = idSu, otherwise. Then {P u w}u∈G(0) is a G-morphism. 6= ∅. For h ∈ Gr(g) w and f ∈ C(Xr(g)), (g · P d(g) µ )(h) = P d(g) µ (g−1h) =ZXw f ((g−1h) · x) dµ(x) =ZXw (g · f )(h · x) dµ(x) = P r(g) µ (g · f )(h). w = ∅, P d µ (g) = idSd(g) and P r(g) If Gu We call the G-morphism Pµ = {P u X and µ . µ = idSr(g). Also, g · P d(g) (g · f ). µ }u∈G(0) the Poisson G-map associated to (f ) = P r(g) µ µ The next two results extend [9, Lemma 3.6] and [9, Proposition 3.4, 3.6]. Lemma 4.7. Let G be a groupoid, let ∂HG = {∂u HG}u∈G(0) be the Hamna boundary of G and w ∈ G(0). Then for every µ ∈ P (Xw), the Poisson G-map Pµ associated to ∂HG is a G-isometry. Proof. Since {C(∂u {C}u∈G(0) and P u {C(∂u HG)}u∈G(0) is the injective envelope of the trivial system µ : C(Xu) −→ Su is a u.c.p. G-map, the G-essentiality of HG)}u∈G(0) implies that P u µ is an isometry for each u ∈ G(0). Proposition 4.8. The action of G on the Hamana boundary ∂HG is minimal and strongly proximal. Proof. Suppose Y = {Yu}u∈G(0) is a G-subspace of ∂HG and suppose that iu : C(∂u HG) −→ C(Yu) 17 is the restriction map. By the essentiality of {C(∂u and hence Yu = ∂u HG, for all u ∈ G(0). HG)}u∈G(0), iu is an isometry HG), we show that, for every w · µ. Otherwise, there Given u, w ∈ G(0) with Gu w 6= ∅ and µ ∈ P (∂w x ∈ ∂u exists f ∈ C+(∂u HG, δx is in the weak∗-closed convex hull of Gu HG) and r > 0 such that, for all g ∈ Gu w, µ (f )(g) =ZXw P u f (g · x) dµ(x) = hf, g · µi ≤ f (x) − r ≤ kf k − r, µ (f )k ≤ kf k − r. By Lemma 4.7, P u hence kP u contradiction. Hence ∂u w · µ. But the weak∗-closed convex hull of ∂u Gu Krein-Milman theorem, ∂u µ is an isometry, which is a HG is contained in the weak∗-closed convex hull of HG), thus by the HG is P (∂u HG ⊆ Gu w · µ. Now we are ready to prove the main result of this section, which extends [9, Theorem 3.11]. Theorem 4.9. For every groupoid G, ∂F G = ∂HG. Proof. Since ∂HG is a G-boundary, by the universal property of the Fursten- berg boundary, there exists a surjective G-map Q = {qu}u∈G(0) : ∂F G −→ ∂HG. The injectivity of the system {C(∂u H)}u∈G(0), gives a G-map Φ = {ϕu}u∈G(0) : ∂HG −→ ∂F G such that Q ◦ Φ = id∂H G. By Lemma 4.5, the only G-map from ∂F G into itself is the identity map. Hence Φ ◦ Q = id∂F G. 5. The reduced crossed product and its injective envelope Let H be a Hilbert space and X be a set. Suppose that ℓ2(X, H) is the set of all functions ξ : X −→ H with Px∈X kξ(x)k2 < ∞, then ℓ2(X, H) with the pointwise operations and the inner product hξ, ζi = Xx∈X hξ(x), ζ(x)i is a Hilbert space. For x ∈ X and h ∈ H define δx,h(t) =(cid:26) h t = x 0 t 6= x. 18 ϕ = supnfφF : F ∈ F (X)o < ∞. (5.2) Then δx,h ∈ ℓ2(X, H) and kδx,hk = khk. Any T ∈ B(ℓ2(X, H)) induces a bounded function ϕ = φT : X × X −→ H by ϕ(x, y)h = T (δy,h)(x). Clearly kϕk ≤ kT k and φ(x, t)(ξ(x)). (5.1) (T ξ)(x) =Xt∈X Conversely, if ϕ : X × X −→ H is a bounded function, formula (5.1) defines a bounded operator T = Tϕ on B(ℓ2(X, H)) with ϕTϕ = ϕ, that is, TϕT = T . Suppose that F (X) is the family of all finite subsets of X. For any F ∈ F (X) and ϕ ∈ ℓ∞(X × X, B(H)), the restriction ϕF of ϕ to F × F induces an operator on H F . Moreover, such a ϕ induces an operator on B(H) if and only if Net let us remind some basic facts about monotone completion of C*- algebras. Let A be a C*-algebra then its self-adjoint part Asa has a natural If each norm bounded, increasing net in Asa has a least partial ordering. upper bound then A is said to be monotone complete. In this case, A is unital. Let A and B be C*-algebras. A positive linear map φ : A −→ B is called normal if for every norm bonded increasing net {ai}i∈I in Asa with a = supi∈I ai we have supi∈I φ(ai) = φ(a). Each C*-algebra A has a unique regular monotone completion A and injective envelope I(A) with A ⊆ A ⊆ I(A), such that the inclusion maps A ֒→ A ֒→ I(A) are normal. i }i∈I , {b(k) i − a(k) ≤ b(k) i ց 0(O) and ai = P3 Let A be a monotone complete C*-algebra. For an increasing net {ai}i∈I in Asa with a = supi∈I, we write ai ր a(O) or −ai ց −a(O). A net {ai}i∈I in A order-converges to a , written O-limi ai = a, if there are bounded nets {a(k) i }i∈I in Asa and elements a(k) in Asa, k = 0, 1, 2, 3, such that 0 ≤ a(k) k=0 ika(k). Given a monotone complete C*-subalgebra A of B(H) and von Neumann subalgebra M of B(K), the monotone tensor product A⊗M of A and M is the monotone closure of A ⊙ M in the Fubini product F(A, M), that is, the smallest monotone closed C*-subalgebra containing A ⊙ M in F(A, M). The monotone complete tensor product A⊗M does not depend on the underling Hilbert space H and K. It is the monotone closure of A⊙M. More generally, if A is a monotone closed C*-subalgebra of a monotone complete C*-algebra B and M is a von Neumann subalgebra of a von Neumann algebra N, then A⊗M is the monotone closure of A ⊙ M in B⊗N. , a = P3 k=0 ika(k) i 19 Suppose that A is a monotone complete C*-subalgebra of B(H) and X be any set. We consider the monotone tensor product A⊗B(ℓ2(X)). For our purposes here, it is enough to observe that each element of the monotone tensor product A⊗B(ℓ2(X)) has a representation as a matrix over A, that is, each element of A⊗B(ℓ2(X)) is in the form ϕ : X × X −→ A ⊆ B(H) sat- isfying (5.2). The involution and multiplication in A⊗B(ℓ2(X)) are defined as follows: ϕ∗(x, y) = ϕ(x, y)∗, ϕ ◦ ψ(x, y) = O-Xt∈X ϕ(x, t)ψ(t, y). Definition 5.1. Let (G, {Au}u∈G(0), α) be a groupoid dynamical system. De- fine A(α) =(cid:8)f ∈ Γ∞(G(0), A) : f (r(g)) = αg(f (s(g)))(cid:9) . We call A(α) the fixed point algebra associated to the system (G, {Au}u∈G(0), α). Proposition 5.2. Let (G, {Au}u∈G(0), α) be a groupoid dynamical system such that for each u ∈ G(0), Au is a monotone complete C*-algebra. Then the fixed point algebra A(α) is a monotone complete C*-algebra. Proof. It is clear that A(α) is a C*-subalgebra of Γ∞(G(0), A). For the mono- tone completeness, suppose that {fj}j∈J is a norm bounded increasing net in A(α). Then, for each u ∈ G(0) , {fj(u)}j∈J is an increasing net in Au,, and hence it has the least upper bound, say f (u) ∈ Au. Then f ∈ Γ∞(G(0), A) and for any g ∈ G, αg(f (s(g))) = αg(sup fj(s(g))) j αg(fj(s(g))) = sup j = sup fj(r(g)) = f (r(g)). j Proposition 5.3. Let A = (G, {Au}u∈G(0), α) be an injective groupoid dy- namical system. Then A(α) is an injective C*-algebra. Proof. By assumption, there exists a G -morphism {ϕu}u∈G(0) from the dy- namical system (G, {Γ∞(Gu, s∗A)}u∈G(0), ℓ) into A such that ϕu ◦ Imu = IdAu, for any u ∈ G(0). Suppose that eA = {Γ∞(Gu, s∗A)}u∈G(0) and define τ : Γ∞(G(0), A) −→ Γ∞(G(0), eA); τ (f )(u)(x) = f (s(x)), 20 and Ψ : Γ∞(G(0), A) −→ Γ∞(G(0), A); Ψ(f )(u) = ϕu(τ (f )(u)). For g ∈ G, x ∈ Gr(g) and f ∈ Γ∞(G(0), A), we have ℓg(τ (f )(u))(x) = τ (f )(u)(g−1x)) = f (s(g−1x)) = f (s(x)) = τ (f )(r(g))(x), thus ℓg(τ (f )(u)) = τ (f )(r(g)). Therefore, αg(Ψ(f )(s(g))) = αg(ϕs(g)(τ (f )(s(g)))) = ϕr(g)(ℓg(τ (f )(s(g)))) = ϕr(g)(τ (f )(r(g))) = Ψ(f )(r(g)), that is, Ψ(f ) ∈ A(α). For f ∈ A(α), τ (f )(u)(x) = f (s(x)) = α−1 Imu(f )(x), for each x ∈ Gu, thus τ (f )(u) = Imu(f (u)). Hence x (f (u)) = Ψ(f )(u) = ϕu(τ (f )(u)) = ϕu(Imu(f (u)) = f (u). Therefore, Ψ is a conditional expectation form Γ∞(G(0), A) onto A(α). Since Γ∞(G(0), A) = ℓ∞Mu∈G(0) Au and each Au is an injective C*-algebra, A(α) is an injective C*-algebra. Let (G, {Au}u∈G(0), α) be a groupoid dynamical system such that for any u, Au is monotone complete. For u ∈ G(0), set A(u) = Au⊗B(ℓ2(Gu)). So each element of A(u) is represented by a function ϕ = ϕu : Gu × Gu −→ Au such that (5.2) is satisfied. For g ∈ G, define eαg : A(s(g)) −→ A(r(g)) by It is not hard to see that eαg(ϕ)(x, y) = αg(ϕ(x.g, y.g)). A⊗ = (G, {A(u)}u∈G(0),eα) 21 is a groupoid dynamical system. The fixed point subalgebra associated to this goupoid dynamical system is called the monotone crossed product of A by G and is denoted by M(G, A). By Proposition (5.2), M(G, A) is monotone complete. Next we show that if A is an injective groupoid dynamical system, then M(G, A) is an injective C*-algebra. Let u ∈ G(0) and be the embedding defined as above, that is, fImu : A(u) −→ Γ∞(Gu, s∗A⊗) fImu(F )(x) = eα−1 x (F ). Theorem 5.4. Let A = (G, {Au}u∈G(0), α) be an injective groupoid dynamical system. Then M(G, A) is an injective C*-algebra. Proof. Since M(G, A) is the fixed point algebra associated to the groupoid dynamical system A⊗, it is sufficient to prove that A⊗is injective. Since A is G-injective, there exists a G-map {ϕu}u∈G(0) from (G, {Γ∞(Gu, s∗A)}u∈G(0), ℓ) into A such that for any u ∈ G(0), φu ◦ Imu = IdAu. Define the map by Idu ⊗1 : Au −→ A(u), Idu ⊗1(a)(x, y) =(cid:26) a x = y 0 x 6= y. Since A is G-injective, there exists a G-map {θu}u∈G(0) from A⊗ into A such that, for each u ∈ G(0), θu ◦ Idu ⊗1 = IdAu. Also, define Θu : Γ∞(Gu, s∗A⊗) −→ Γ∞(Gu, s∗A), by Θu(F )(g) = θs(g)(F (g)). Then, for u ∈ G(0), θu ◦ fImu ◦ Idu ⊗1 = Imu. Set eϕu = Idu ⊗1 ◦ ϕu ◦ θu. Then eΦ = {eϕu}u∈G(0) is a G-map from the dynamical system (G, {Γ∞(Gu, s∗A⊗)}u∈G(0), ℓ) into A⊗ such that eϕu ◦ fImu = IdA(u). Thus A(α) is G-injective. Let s : X −→ X0 be a surjective map and {Hu}u∈X0 be a family of Hilbert spaces. Let ℓ2(X, s∗H) be the set of all functions ξ : X −→ ∪u∈X0Hu such that ξ(x) ∈ Hs(x) and kf k =Xx∈X f (x)2 < ∞. 22 Then ℓ2(X, s∗H) is a Hilbert space with the pointwise operations and the inner product hξ, ηi = Xx∈X hξ(x), η(x)i, which is canonically isomorphic to the ℓ2-direct sum Mu∈X0 ℓ2(s−1(u), Hu). Suppose that A = (G, {Au}u∈G(0), α) is a groupoid dynamical system and suppose that for any u ∈ G(0), Hu is a Hilbert space such that Au ⊂ B(Hu). For f ∈ Γc(G, r∗A) define Π(f ) : ℓ2(G, s∗H) −→ ℓ2(G, s∗H) by (Π(f )ξ)(x) = Xt∈Gr(x) α−1 x (f (t))ξ(t−1x). (5.3) This is a faithful representation for Γc(G, r∗A). The norm closure A ⋊r G of Π(Γc(G, r∗A)) in B(ℓ2(G, s∗H)) is called the reduced crossed product of A by G. Let A = (G, {Au}u∈G(0), α) be a groupoid dynamical system such that for any u ∈ G(0), Au is monotone complete. For u, v ∈ G(0) and a ∈ Av, define u(a) : Gu × Gu −→ Au πv by u(a)(x, y) =(cid:26) α−1 0, πv x (a), x = y ∈ Gv u otherwise, Also, for g ∈ G, define λu(g) : Gu × Gu −→ Au λu(g)(x, y) =(cid:26) 1, xy−1 = g 0, otherwise. Then πv = (πv Av and ξ = (ξu)u∈G(0) ∈ ℓ2(G, s∗H), we have πv(a)ξ = (πv u(a)ξu(x) = α−1 πv u)u∈G(0) and λ(g) = (λu(g))u∈G(0) are in M(G, A). For a ∈ u(a)ξu)u∈G(0) and x (a)ξu(x). Also, for g ∈ G, λgξ = (λu(g)ξu)u∈G(0) and λu(g)ξu(x) =(cid:26) ξu(g−1x) r(x) = r(g) otherwise. 0, 23 Hence, u(a) ◦ λu(g))ξu(x) =(cid:26) α−1 (πv x (a)ξu(g−1x), r(x) = r(g) = v 0, otherwise. Since the reduced crossed product A ⋊r G is generated by the set (cid:8)πv(a)λ(g) : v ∈ G(0)a ∈ Av, g ∈ G,(cid:9) , M(A, G) contains A ⋊r G . For each v ∈ G(0), πv(Av) is a C*-subalgebra of M(G, A) and for g ∈ G and a ∈ As(g) we have λ(g)πs(g)(a)λ∗(g) = πr(g)(αg(a)). g(πs(g)(a)) = πr(g)(αg(a)), then (G, {πv(Av)}v∈G(0), α′) is a groupoid Let α′ dynamical system, isomorphic to the original system A. Now we are ready to prove the main result of this section, which is essen- tial in further development of the theory [1]. Theorem 5.5. Suppose that A = (G, {Au}u∈G(0), α) be a groupoid dynamical system. Then IG(A) ⋊r G ⊆ I(A ⋊r G). Proof. Since A ⋊r G ⊆ IG(A) ⋊r G ⊆ M(G, IG(A)) and M(G, IG(A)) is an in- jective C*-algebra, we may consider I(A ⋊r G) ⊆ M(G, IG(A)). The inclusion map j : A ⋊r G −→ IG(A) ⋊r G extends to a completely positive map ψ : IG(A) ⋊r G −→ I(A ⋊r G) ⊆ M(G, IG(A)). Since ψ is completely positive and preserves λ(G), by [2, 3.1], for any f ∈ IG(A)⋊rG, ψ(λ(g)f λ(g)∗) = λ(g)ψ(f )λ(g)∗. For v ∈ G(0), set ψv = ψπv(IG (A)v ) and suppose that ρv : M(G, IG(A)) −→ πv(IG(A)v) ⊆ M(G, IG(A)) is the map defined by ρv((Fu)u∈G(0)) = πv(Fv(v, v)). Thus ρv is a conditional expectation and if ρv((Fu)∗ u∈G(0)(Fu)u∈G(0)) = 0 then for any u ∈ G(0) and x ∈ Gv u, y ∈ Gu, Fu(x, y) = 0. 24 For g ∈ G and F ∈ M(G, IG(A)), ρr(g)(λ(g)(Fu)u∈G(0)λ(g)∗) = πr(g)(λr(g)(g) ◦ Fr(g)λr(g)(g)∗) = πr(g)(Fr(g)(g−1, g−1)) = πr(g)(αg(Fs(g)(s(g), s(g))) = λ(g)πs(g)(Fs(g)(s(g), s(g)))λ(g)∗ = λ(g)ρs(g)λ(g)∗. Since (G, {πv(IG(A)v)}v∈G(0), α′) is a G-essential extension of the dynamical system (G, {πv(Av)}v∈G(0), α′) and ρv and ψv preserve πv(Av), eρv ◦ ψv(x) = x, for each x ∈ πv(IG(A)v). By a similar arguments as in the proof of [8, Lemma 3.3], we conclude that ρv((ψ(πv(a)) − πv(a))∗(ψ(πv(a)) − πv(a))) = 0, thus, for any u ∈ G(0) and x ∈ Gv u, y ∈ Gu, ψ(πv(a))u(x, y) = πv u(a)(x, y), and hence ψ(λvπv(a)) = λvψ(πv(a)) = πv(a). This means that πv(a) is in the range of ψ . Since IG(A) ⋊r G is generated by the operators λg and πv(a), we get I(A ⋊ G) ⊇ IG(A) ⋊r G. References References [1] M. Amini, F. Behrouzi, C ∗-simplicity and Ozawa conjecture for groupoid C ∗-algebras, part II: crossed products, preprint. [2] M.D. Choi, A Schwarz inequality for positive linear maps on C*-algebras, Illinois J. Math. 18 (1974), 565-574. [3] E.G. Effros and Z.J. Ruan, Operator spaces. (Oxford Univ. Press, Ox- ford, 2000). [4] H. Furstenberg, Boundary theory and stochastic processes on homo- geneous spaces. Harmonic Analysis on Homogeneous Spaces, (Proc. Sympos. Pure Math., Vol. XXVI, Williams Coll., Williamstown, Mass., 1972), pp. 193-229, Amer. Math. Soc., Providence, R.I., 1973. 25 [5] S. Glanser, Proximal flows, Lectures Notes in Mathematics, vol. 517. (Springer-Verlag, Berlin, Heidelberg, New York, 1976). [6] D. Hadwin, I. V. Paulsen, Injectivity and projectivity in analysis and topology. Science China Math. 54 (2011), 2347 -- 2359. [7] M. Hamana, Injective envelopes of operator systems. Publ. Res. Inst. Math. Sci. 15 (1979), 773 -- 785. [8] M. Hamana, Injective envelopes of C ∗-dynamical systems. Tohoku Math. J. 37 (1985), 463 -- 487. [9] M. Kalantar and M. Kennedy, Boundaries of Reduced C ∗-algebras of discrete groups. J. Reine Angew. Math., to appear. (arXiv:1405.4359). [10] E. Kirchberg, On subalgebras of the CAR algebra. J. Funct. Anal. 129 (1995), 35 -- 63. [11] E. Kirchberg and S. Wassermann, Exact groups and continuous bundles of C*-algebras. Math. Annalen. 315 (1999), 169 -- 203. [12] P. Muhly, Coordinates in operator algebra, unpublished. [13] N. Ozawa, Amenable actions and exactness for discrete groups. Comptes Rendus de lAcad´emie des Sciences. 330 (2000), 691 -- 695. [14] A. L. T. Paterson. Groupoids, inverse semigroups, and their operator algebras, Progress in Mathematics 170. (Birkhauser, Boston, MA, 1999). [15] J. Renault, A groupoid approach to C∗-algebras. Lecture Notes in Math. 793. (Springer-Verlag, Berlin, 1980). [16] S. Wasserman, Tensor products of free group C*-algebras. Bull. London Math. Soc. 22 (1990), 375 -- 380. 26
1701.04127
1
1701
2017-01-15T23:49:15
Around trace formulas in non-commutative integration
[ "math.OA", "math.QA" ]
Trace formulas are investigated in non-commutative integration theory. The main result is to evaluate the standard trace of a Takesaki dual and, for this, we introduce the notion of interpolator and accompanied boundary objects. The formula is then applied to explore a variation of Haagerup's trace formula.
math.OA
math
AROUND TRACE FORMULAS IN NON-COMMUTATIVE INTEGRATION SHIGERU YAMAGAMI Graduate School of Mathematics Nagoya University Nagoya, 464-8602, JAPAN Abstract. Trace formulas are investigated in non-commutative integration theory. The main result is to evaluate the standard trace of a Takesaki dual and, for this, we introduce the notion of interpolator and accompanied bound- ary objects. The formula is then applied to explore a variation of Haagerup's trace formula. Introduction The Haagerup's trace formula in non-commutative integration is a key to his whole theory of non-commutative Lp-spaces (see [6] and [9]). Our purpose here is to analyse it from the view point of modular algebras ([10], [11]), which was originally formulated in terms of Haagerup's Lp-theory itself. So, to circumvent tautological faults and also to fix notations, we first describe modular algebras as well as standard Hilbert spaces in terms of basic ingredients of Tomita-Takesaki theory. The semifiniteness of Takesaki's duals is then established by constructing relevant Hilbert algebras as a collaboration of modular algebras and complex analysis. Note that the known proofs of the existence of standard traces are not direct; for example, it is usually deduced from the innerness of modular automorphism groups combined with a reverse Radon-Nikodym theorem such as Pedersen-Takesaki's or Connes'. Since our construction of the Hilbert algebras is based on complex analysis, the associated trace can be also described in a calculational way. To make the setup transparent, we introduce the notion of interpolators together with associ- ated boundary operators and vectors. Viewing things this way, the main trace formula turns out to be just a straightforward consequence of definitions. The Haagerup's trace formula is then derived in a somewhat generalized form as a con- crete application of our formula. The Haagerup's correspondence between normal functionals and relatively invari- ant measurable operators on Takesaki's duals is also established on our streamlines. Recall that the standard approach to these problems is by the theory of operator- valued weights ([4], [5]) coupled with dual weights ([2], [3]), which is based on ex- tended positive parts, a notion of metaphysical flavor, and somewhat elaborate. Our method may not provide an easy route either but can be applied rather straightfor- wardly; it is just a simple combination of elementary Fourier calculus and complex analytic nature of modular stuffs. 1 2 SHIGERU YAMAGAMI The presentation below originates from the author's old work in 1990, which was addressed on the occasion of a satellite meeting of ICM90 held at Niigata University. The author would like to express hearty gratitude to Kichisuke Saito for his organization of the meeting and these records. Notation and Convention: The positive part of a W*-algebra M (resp. its predual M∗) is denoted by M+ (resp. M + For a positive element p in M+ or M + ∗ ). ∗ , its support projection in M is denoted by [p]. For a functional ϕ ∈ M + ∗ , the associated GNS-vector in the standard Hilbert space L2(M ) of M is denoted by ϕ1/2 (natural notation though not standard) and the modular operator by ∆ϕ so that ∆ϕ(aϕ1/2) = ϕ1/2a for a ∈ [ϕ]M [ϕ]. ∗ , σϕ,ψ For ϕ, ψ ∈ M + stands for the relative modular group of [ϕ]M [ψ], which is t simply denoted by σϕ t and expresses a modular automorphism group of the reduced algebra [ϕ]M [ϕ] when ϕ = ψ. For convergence in M , w*-topology (resp. s-topology or s*-topology) means weak operator topology (resp. strong operator topology or *strong operator topology) as a von Neumann algebra on the standard Hilbert space L2(M ). Direct integrals are indicated by H instead of ordinary R ⊕. This is to avoid The notion of weights is used in a very restrictive sense: weights are orthogonal duplication of sum meanings. sums of functionals in M + ∗ . For an interval I contained in [0, 1], TI expresses the tubular domain based on an imaginary trapezoid {(x, y) ∈ R2; x ≤ 0, y ≤ 0, −(x + y) ∈ I}: TI = {(z, w) ∈ C2; Im z ≤ 0, Im w ≤ 0, −( Im z + Im w) ∈ I}. A function f : D → M with D ⊂ C is said to be w*-analytic (s*-analytic) if it is w*-continuous (s*-continuous) and holomorphic when restricted to the interior D◦. Note that topologies are irrelevant for holomorphicity because weaker one implies power series expansions in norm. For real numbers α, β, α ∨ β = max{α, β}, α ∧ β = min{α, β.}. 1. Standard Hilbert Spaces Given a faithful ω ∈ M + ∗ , we denote the associated GNS-vector by ω1/2 and iden- tify the left and right GNS-spaces by the relation ∆1/2 ω (xω1/2) = ω1/2x, resulting in an M -bimodule L2(M, ω) = M ω1/2M with the positive cone L2(M, ω)+ and the compatible *-operation given by L2(M, ω)+ = {aω1/2a∗; a ∈ M } and (aω1/2b)∗ = b∗ω1/2a∗ in such a way that these constitute a so-called standard form of M . The dependence on ω as well as its faithfulness is then removed by the matrix ∗ , let M ⊗ ϕ1/2 ⊗ M be a dummy of the ampliation technique: For each ϕ ∈ M + algebraic tensor product M ⊗ M , which is an M -bimodule in an obvious manner with a compatible *-operation defined by the relation (a⊗ϕ1/2⊗b)∗ = b∗⊗ϕ1/2⊗a∗. On the algebraic direct sum M ⊗ ϕ1/2 ⊗ M Mϕ∈M + ∗ of these *-bimodules, introduce a sesquiliear form by TRACE FORMULAS 3  nMj=1 xj ⊗ ω1/2 j ⊗ yj(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) nMk=1 which is positive because of k ⊗ ω1/2 x′ k ⊗ y′ k=Xj,k ([ωk](x′ k)∗xj ω1/2 j ω1/2 k y′ ky∗ j [ωj]), ([ωk]x∗ kxj ω1/2 j Xj,k ω1/2 k yky∗ j [ωj]) = (Xω1/2ω1/2Y ) = (X 1/2ω1/2Y 1/2X 1/2ω1/2Y 1/2) ≥ 0. Here ω = diag(ω1, . . . , ωn) denotes a diagonal functional on the n-th matrix ampli- ation Mn(M ) of M and X = [ω] x∗ 1 ... x∗ n (cid:0)x1 . . . xn(cid:1) [ω] and Y = [ω] y1 ... yn (cid:0)y∗ 1 . . . y∗ n(cid:1) [ω] are positive elments in [ω]Mn(M )[ω]. Recall that [ω] = diag([ω1], . . . , [ωn]). The associated Hilbert space is denoted by L2(M ) and the image of a ⊗ ϕ1/2 ⊗ b in L2(M ) by aϕ1/2b. Here the notation is compatible with the one for L2(M, ϕ) because [ϕ]M [ϕ] ⊗ ϕ1/2 ⊗ [ϕ]M [ϕ] ∋ a ⊗ ϕ1/2 ⊗ b 7→ aϕ1/2b ∈ L2(M, ϕ) gives an isometric map by the very definition of inner products. Similar remarks are in order for left and right GNS spaces. The left and right actions of M are compatible with taking quotients and they are bounded on L2(M ): For a ∈ M , with 2 j ⊗ yj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)  a∗a(cid:0)x1 = (ω1/2ZJY Jω1/2) . . . xn(cid:1) [ω] ≤ kak2X. axj ⊗ ω1/2 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Mj 0 ≤ Z = [ω] x∗ 1 ... x∗ n Moreover, these actions give *-representations of M : (aξη) = (ξa∗η) and (ξaη) = (ξηa∗) for ξ, η ∈ L2(M ) and a ∈ M , which is immediate from the definition of inner product. 4 SHIGERU YAMAGAMI The *-operation on L2(M ) is also compatible with the inner product: j ⊗ ω1/2 y∗ = (Y ω1/2ω1/2X) = ((ω1/2X)∗(Y ω1/2)∗) = (Xω1/2ω1/2Y ) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Mj xj ⊗ ω1/2 j ⊗ yj 2 ∗(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Mj =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Mj 2 j ⊗ x∗ j(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) j ⊗ yj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 2 . xj ⊗ ω1/2 In this way, we have constructed a *-bimodule L2(M ) of M in such a way that L2(M, ϕ) ⊂ L2(M ) for each ϕ ∈ M + ∗ and the closed subspaces M ϕ1/2, ϕ1/2M in L2(M ) are naturally identified with the left and right GNS spaces of ϕ respectively. Moreover, for ϕ, ψ ∈ M + ∗ , we have [ϕ]M ψ1/2 = ϕ1/2M [ψ] in L2(M ), which is just a reflection of the fact that the same identification inside L2(Mn(M ), ω) is used in the definition of inner product. 2. Modular Algebras Recall the definition of (boundary) modular algebra which was introduced in [10] to 'resolve' various cocycle relations in modular theory. We shall here describe it without essential use of the notion of weights. Let M be a W*-algebra which is assumed to admit a faithful ω ∈ M + ∗ for the moment. The modular algebra M (iR) of M is then the *-algebra generated by elements in M and symbols ϕit for ϕ ∈ M + ∗ and t ∈ R under the conditions that (i) M (iR) contains M as a *-subalgebra, (ii) {ϕit}t∈R is a one-parameter group of partial isometries satisfying ϕi0 = [ϕ], (iii) ϕita = σϕ,ψ (a)ψit for ϕ, ψ ∈ M + ∗ , a ∈ [ϕ]M [ψ] and t ∈ R. t By utilizing a faithful ω ∈ M + algebraic crossed product of M by {σω sum of M (it) = M ωit = ωitM , where M (it) =Pϕ∈M + M (−it), M (is)M (it) = M (i(s + t)) and M (i0) = M . ∗ ∗ , it turns out that M (iR) is *-isomorphic to the t } and therefore M (iR) is an algebraic direct M ϕitM . Thus the modular algebra M (iR) is iR-graded in the sense that M (it)∗ = We say that an element a ∈ M is finitely supported if a = [ϕ]a[ϕ] for some ϕ ∈ M + ∗ . Let Mf be the set of finitely supported elements in M . Lemma 2.1. Mf is a w*-dense *-subalgebra of M and closed under sequential w*-limits in M . Moreover, Mf = Xϕ∈M + ∗ M [ϕ] = Xϕ∈M + ∗ [ϕ]M. Proof. Clearly Mf is closed under the *-operation and Mf is a subalgebra in view of [ϕ] ∨ [ψ] ≤ [ϕ + ψ]. The *-subalgebra Mf is then w*-dense in M in view of [ϕ] = 1. If a = a[ϕ], [aϕa∗] is the left support of a and [ϕ+ aϕa∗]a[ϕaϕa∗] = ∨ϕ∈M + a. Let a be a w*-limit of {an}n≥1 in Mf with [ϕn][an][ϕn] = an for n ≥ 1. Then, (cid:3) ∗ n=1 2−nϕn/ϕn(1) ∈ M + ∗ , [ϕ]a[ϕ]. for ϕ =P∞ TRACE FORMULAS 5 Now we relax the existence of faithful functionals in M + ∗ and set Mf (iR) = [ϕ∈M + ∗ [ϕ]M [ϕ](iR), ∗ satisfying where the natural inclusions [ϕ]M [ϕ](iR) ⊂ [ψ]M [ψ](iR) for ϕ, ψ ∈ M + [ϕ] ⊂ [ψ] are assumed in the union. j for families {ωj ∈ M + ∗ }j∈I of mutually orthogonal supports and allow products with elements in M to get {M (it)}t∈R so that Mf (it) ⊂ M (it) and M (0) = M . In what follows, a Finally we add formal expressions of the form ωit =Pj∈I ωit formal sum ω = Pj∈I ωj is referred to as a weight of M . A weight ω = P ωj is said to be faithful if 1 =P[ωj] in M . Note that any weight is extended to a a faithful ω and, for another choice of a faithful weight φ =Pk∈J φk and a ∈ M , faithful one and {ωit} is a one-parameter group of unitaries in M (iR) = ⊕M (it) for φitaω−it =Xj,k k aω−it φit j defines a continuous family of elements in M so that it consists of unitaries when a = 1 and σω t (a) = ωitaω−it gives an automorphic action of R on M . Remark 1. Here weights are introduced in a formal and restricted way. At this stage, we introduce two more classes of modular algebraic stuffs: M (iR + 1/2) =Xt∈R M (it + 1/2), M (iR + 1) =Xt∈R M (it + 1), with and M (it + 1/2) = Xϕ∈M + M (it + 1) = Xϕ∈M + M ϕit+1/2 = Xϕ∈M + M ϕit+1 = Xϕ∈M + ∗ ∗ ∗ ∗ ϕit+1/2M ϕit+1M so that M (1/2) = L2(M ) and M (1) = M∗. These are iR-graded *-bimodules of M (iR) in an obvious way and we have a natural module map M (iR + 1/2) ⊗M(iR) M (iR + 1/2) → M (iR+ 1) which respects the grading in the sense that M (is + 1/2)M (it + 1/2) = M (i(s + t) + 1) for s, t ∈ R. In particular, given a weight ω on M , we have M (it + s) = M (s)ωit = ωitM (s) for s = 1/2, 1. The evaluation of ϕ ∈ M∗ at the unit 1 ∈ M is called the expectation of ϕ and denoted by hϕi. Note that the expectation satisfies the trace property for various combinations of multiplications such as haϕi = hϕai and hϕitξψ−itηi = hψ−itηϕitξi for a ∈ M , ϕ, ψ ∈ M + ∗ and ξ, η ∈ L2(M ). The scaling ϕ 7→ e−sϕ on M + ∗ gives rise to a *-automorphic action θs of s ∈ R (called the scaling automorphisms) on these modular stuffs: θs(xϕit+r) = e−ist−srxϕit+r for x ∈ M , r ∈ {0, 1/2, 1} and t ∈ R. Remark 2. Since elements in L2(M ) and M∗ are always 'finitely supported', we can decribe M (it + s) (s = 1/2, 1) without referring to weights. 6 SHIGERU YAMAGAMI 3. Analytic Properties We here collect well-known analytic properties of modular stuffs (proofs can be found in [1] and [8] for example). Lemma 3.1 (Modular Extension). For ϕ, ψ ∈ M + ∗ and a ∈ M , R ∋ t 7→ ϕitaψ1−it ∈ M∗ is extended analytically to a norm-continuous function ϕizaψ1−iz on the strip −1 ≤ Im z ≤ 0 with a bound kϕit+raψ−it+1−rk ≤ kϕakrkaψk1−r (0 ≤ r ≤ 1) in such a way that (ϕiz aψ1−iz)z=t−i = ϕ1+itaψ−it, (ϕizaψ1−iz)z=t−i/2 = ϕit+1/2aψ−it+1/2. + and a ∈ [ϕ]M [ψ]. Then the (a)ψ1/2 = ϕitaψ−itψ1/2 of t ∈ R is analytically extended to an L2(M )- Corollary 3.2 (KMS condition). Let ϕ, ψ ∈ M ∗ function σϕ,ψ valued continuous function ϕizaψ−iz+1/2 of z ∈ R−i[0, 1/2] so that (ϕizaψ−iz+1/2)z=t−i/2 = ϕ1/2ϕitaψ−it = ϕ1/2σϕ,ψ (a). t t Lemma 3.3. Let ω ∈ M + are equivalent. ∗ be faithful and let a ∈ M . Then the following conditions (i) The inequality a∗ωa ≤ ω holds in M + ∗ . (ii) We can find a function a(z) ∈ M of z ∈ R − i[0, 1/2] such that a(t) = ωitaω−it for t ∈ R, a(z)ξ ∈ L2(M ) is norm-analytic in z for any ξ ∈ L2(M ) and ka(−i/2)k ≤ 1. (iii) We can find an element b ∈ M satisfying kbk ≤ 1 and ω1/2a = bω1/2. Moreover, if this is the case, with the notation in (ii), ξa(z) ∈ L2(M ) is norm- continuous in z for every ξ ∈ L2(M ). Corollary 3.4. For ϕ, ψ ∈ M + ∗ , the following conditions are equivalent. (i) The inequality ϕ ≤ ψ holds in M + ∗ . (ii) [ϕ] ≤ [ψ] and the function ϕitψ−it of t ∈ R is analytically extended to an M -valued function ϕizψ−iz of z ∈ R − i[0, 1/2] so that ϕizψ−izξ ∈ L2(M ) is norm-continuous in z for any ξ ∈ L2(M ) and kϕ1/2ψ−1/2k ≤ 1. (iii) We can find an element c ∈ M satisfying kck ≤ 1 and ϕ1/2 = cψ1/2. Moreover, if this is the case, ξϕizψ−iz ∈ L2(M ) is norm-continuous in z ∈ R − i[0, 1/2] for any ξ ∈ L2(M ). Remark 3. Under the above majorization conditions, the relevant analytic exten- sions are norm-bounded as M -valued functions of z ∈ R−i[0, 1/2] thanks to Banach- Steinhaus theorem. 4. Sectional Continuity We now describe continuity properties of families {M (it + r)}t∈R for r = 0, 1/2 and 1. Let us begin with a simple observation on the continuity of modular actions: Let ϕ =P ϕj and ψ =P ψk be weights on M in our sense. For ξ ∈ L2(M ), ϕitξψ−it =Xj,k j ξψ−it ϕit k TRACE FORMULAS 7 is norm-continuous in t ∈ R as an orthogonal sum of L2(M )-valued norm-continuous ∗ has an expression ξη with ξ, η ∈ L2(M ), one functions ϕit sees that . As any φ ∈ M + j ξψ−it k ϕitφψ−it = (ϕitξω−it)(ωitηψ−it) (ω being an auxiliary faithful weight) is an M∗-valued norm-continuous function of t ∈ R as a product of L2(M )-valued norm-continuous functions. The following facts on continuity of sections of {M (it + r)} are then more or less straightforward from this observation. Lemma 4.1. For a section x = {x(t)} of {M (it)}, the following conditions are equivalent. (i) There exists a faithful weight ω on M such that ω−itx(t) ∈ M is w*- continuous in t ∈ R. (ii) There exists a faithful weight ω on M such that x(t)ω−it ∈ M is w*- continuous in t ∈ R. (iii) For any faithful weight ω on M , ω−itx(t) ∈ M is w*-continuous in t ∈ R. (iv) For any faithful weight ω on M , x(t)ω−it ∈ M is w*-continuous in t ∈ R. (v) For any φ ∈ M + (vi) For any φ ∈ M + ∗ , φ−itx(t) ∈ M is w*-continuous in t ∈ R. ∗ , x(t)φ−it ∈ M is w*-continuous in t ∈ R. Moreover, if {x(t)} satisfies these equivalent conditions, kx(t)k is locally bounded in t ∈ R. We say that a section {x(t)} is w*-continuous if it satisfies any of these equiv- alent conditions. We here introduce the *-operation on sections by x∗(t) = x(−t)∗ ∈ M (it + s) for a section {x(t) ∈ M (it + r)}. As a consequence of the above lemma, for a section x(t) ∈ M (it), x∗(t) as well as ax(t)b with a, b ∈ M are w*-continuous if so is x(t). Lemma 4.2. Let p = 1 or 2 with the notation L1(M ) = M∗ for p = 1. Then the following conditions on a section {ξ(t)} of {M (it + 1/p)} are equivalent. (i) There exists a faithful weight ω on M such that ω−itξ(t) ∈ Lp(M ) is norm-continuous in t ∈ R. (ii) There exists a faithful weight ω on M such that ξ(t)ω−it ∈ Lp(M ) is norm-continuous in t ∈ R. (iii) For any faithful weight ω on M , ω−itξ(t) ∈ Lp(M ) is norm-continuous in t ∈ R. (iv) For any faithful weight ω on M , ξ(t)ω−it ∈ Lp(M ) is norm-continuous in t ∈ R. (v) For any φ ∈ M + (vi) For any φ ∈ M + ∗ , φ−itξ(t) ∈ Lp(M ) is norm-continuous in t ∈ R. ∗ , ξ(t)φ−it ∈ Lp(M ) is norm-continuous in t ∈ R. We say that a section {ξ(t)} is norm-continuous if it satisfies any of these equiv- alent conditions. Notice here that ξ∗(t) = ξ(−t)∗ is norm-continuous if so is ξ(t). Definition 4.3. A section {x(t)} of {M (it)}t∈R is said to be s*-continuous if x(t)ξ and ξx(t) are norm-continuous for any ξ ∈ L2(M ). Notice that x∗(t) is s*- continuous if and only if so is x(t) in view of x∗(t)ξ = (ξ∗x(−t))∗ and ξx∗(t) = (x(−t)ξ∗)∗. Clearly s*-continuous sections are w*-continuous. 8 SHIGERU YAMAGAMI To control the norm of a w*-continuous section x = {x(t) ∈ M (it)}, two norms are introduced by kxk∞ = sup{kx(t)k; t ∈ R}, kxk1 =ZR kx(t)k dt and x(t) is said to be bounded if kxk∞ < ∞ and integrable if kxk1 < ∞. Note here that kx(t)k is locally bounded and lower-semicontinuous. Lemma 4.4. The following conditions on a section {x(t)} of {M (it)} are equiva- lent. (i) For any ξ ∈ L2(M ), {x(t)ξ} is a norm-continuous section of {M (it+1/2)}. (ii) For any ϕ ∈ M + ∗ and any ξ ∈ L2(M ), x(t)ϕ−itξ ∈ L2(M ) is norm- continuous in t ∈ R. (iii) The norm function kx(t)k is locally bounded and, for a sufficiently large ∗ , x(t)φ−it+1/2 ∈ L2(M ) is norm-continuous in t ∈ R, i.e., given any φ ∈ M + ∗ such that ϕ ≤ φ and x(t)φ−it+1/2 ∈ L2(M ) ∗ , we can find φ ∈ M + ϕ ∈ M + is norm-continuous in t ∈ R. Corollary 4.5. A section x(t) ∈ M (it) is s*-continuous if and only if kx(t)k is locally bounded and L2(M )-valued functions x(t)φ−it+1/2, φ−it+1/2x(t) are norm- continuous for a sufficiently large φ ∈ M + ∗ . A section {x(t) ∈ M (it)}t∈R is said to be finitely supported if we can find ∗ so that x(t) = [φ]x(t)[φ] for every t ∈ R. We say that {x(t)} is locally φ ∈ M + bounded (bounded) if so is the function kx(t)k of t. 5. Convolution Algebra Consider a bounded, s*-continuous and integrable section {f (t) ∈ M (it)} and f (t) dt, which is compatible with the *-operation by Moreover, a formal rewriting identify it with a formal expression like ZR (cid:18)ZR f (t) dt(cid:19)∗ f (t)∗ dt =ZR =ZR g(t) dt =ZR(cid:18)ZR f (s) dsZR ZR f (s)g(t − s) ds =ZR (f g)(t) =ZR suggests to define a product of f and g by f ∗(t) dt. f (−t)∗ dt =ZR f (s)g(t − s) ds(cid:19) dt f (t − s)g(s) ds. It is then a routine work to check that the totality of such sections constitutes a normed *-algebra in such a way that kf gk∞ ≤ (kf k1kgk∞) ∧ (kf k∞kgk1), kf gk1 ≤ kf k1 kgk1. We notice that the scaling automorphism θs on {M (it)} induces a *-automorphic action on the *-algebra of sections by (θsf )(t) = e−istf (t). Here we shall apply formal arguments to illustrate how tracial functionals can be associated to this kind of *-algebras. TRACE FORMULAS 9 Formal manipulation is an easy business: Imagine that a section f (t) has an analytic extension to the region −1 ≤ Im z ≤ 0 in some sense so that f ∗(z) = f (−z)∗ and define a linear functional by τ(cid:18)ZR f (t) dt(cid:19) = hf (−i)i. Note that f (−i) in the right hand side belongs to M (1) = M∗. We then have τ (f ∗f ) =ZR =ZR hf ∗(s)f (−i − s)i ds =ZR hf (−s − i/2)∗f (−s − i/2)i ds ≥ 0, hf ∗(s − i/2)f (−s − i/2)i ds where Cauchy's integral theorem is formally used in the first line. The trace prop- erty is seen from τ (f g) =Z hf (s)g(−i − s)i ds =Z hf (t − i)g(−t)i dt =Z hg(−t)f (t − i)i dt =Z hg(t)f (−t − i)i dt = τ (gf ), where Cauchy's integral theorem is again used formally in the first line. Going back to the sane track, it turns out that it is not easy to make all of the above formal arguments rigorous at least in a reference-weight-free fashion. Instead we shall construct a Hilbert algebra as a halfway business in what follows, which is enough to extract the tracial functional. 6. Hilbert Algebras Definition 6.1. A section {f (t) ∈ M (it)} is said to be half-analytic if, for a sufficiently large φ ∈ M + of (t′, t′′) ∈ R2 is analytically extended to a bounded M -valued s*-continuous function fφ(z′, z′′) = φ−iz ′ ∗ , the function fφ(t′, t′′) = φ−it′ of (z′, z′′) ∈ T[0,1/2]. f (z′ + z′′)φ−iz ′′ f (t′ + t′′)φ−it′′ Note here that sufficient largeness in the condition has a meaning: For a φ ∗ , φitω−it is analytically extended to a s*-continuous function majorized by ω ∈ M + φizω−iz of z ∈ R− i[0, 1/2] (Corollary 3.4) and therefore ω−it′ has an analytic extension of the form (ω−iz ′ ), which is s*-continuous as a product of s*-continuous locally bounded operator-valued functions. f (t′ + t′′)ω−it′′ ω−iz ′′ f (z′ + z′′)φ−iz ′′ )(φ−iz ′ )(φiz ′′ φiz ′ Note also that the s*-continuity of φ−iz ′ continuity of L2-valued functions (φ−iz ′ z′′)φ−iz ′′ and φ1/2−it′ φ1/2−iz ′ f (z′ + z′′)φ−iz ′′ f (t′ + t′′)φ−it′′ respectively. ) (Lemma 4.4). These are analytic extensions of φ−it′ , whence simply denoted by φ−iz ′ f (z′+z′′)φ−iz ′′ f (z′ + z′′)φ−iz ′′ )φ1/2 and φ1/2(φ−iz ′ is equivalent to the norm- f (z′ + f (t′ + t′′)φ1/2−it′′ and f (z′ + z′′)φ1/2−iz ′′ Warning: No separate meaning of f (z) is assigned here. It is immediate to see that f (t) is half-analytic if and only if so is f ∗(t) = f (−t)∗ in such a way that (1) φ−iz ′ f ∗(z′ + z′′)φ−iz ′′ =(cid:0)φiz ′′ f (−z′′ − z′)φiz ′(cid:1)∗ . 10 SHIGERU YAMAGAMI To get the convolution product in a manageable way, we impose the following decaying condition. For a half-analytic section f (t) ∈ M (it), the obvious identity fφ(z′ + s′, z′′ + s′′) = φ−is′ shows that kfφ(z′, z′′)k depends only on r′ = − Im z′, r′′ = − Im z′′ and t = Re (z′ + z′′), which enables us to introduce fφ(z′, z′′)φ−is′′ f φ(t) = sup{kfφ(z′, z′′)k; r′ ≥ 0, r′′ ≥ 0, r′ + r′′ ≤ 1/2}. A half-analytic section f (t) is said to be of Gaussian decay if, for a sufficiently large φ ∈ M + ∗ , we can find δ > 0 so that f φ(t) = O(e−δt2 ). Now let N be the vector space of half-analytic sections of Gaussian decay, which is closed under taking the *-operation by (1). It is immediate to see that the scaling automorphisms leave N invariant so that φ−iz ′ = e−is(z ′+z ′′)φ−iz ′ (θsf )(z′ + z′′)φ−iz ′′ f (z′ + z′′)φ−iz ′′ . Let f, g ∈ N. Thanks to the Gaussian decay assumption, the convolution product f g has a meaning and (f g)(t) is an s*-continuous section. To see f g ∈ N, we therefore need to check that it admits a half-analytic extension of Gaussian decay. Choose an auxiliary weight ω which supports both f and g. Then φ−iz ′ ((f g)(z′ + z′′)φ−iz ′′ ) =ZR φ−iz ′ f (z′ + s)ω−isσω s (g(z′′ − s)φ−i(z ′′−s))ωisφ−is ds gives the s*-continuous analytic extension with its norm estimated by kφ−iz ′ (f g)(z′ + z′′)φ−iz ′′ k ≤ZR f φ( Re z′ + s) gφ( Re z′′ − s) ds = O(e−ǫδt2/(ǫ+δ)) for t = Re (z′ + z′′) if f φ(t) = O(e−ǫt2 ) and gφ(t) = O(e−δt2 ). So far N is shown to be a *-algebra with an automorphic action of R by scaling automorphisms. We next introduce an inner product which makes N into a Hilbert algebra. Lemma 6.2. The following identity holds for f ∈ N and sufficiently large φ, ϕ ∈ M + ∗ . [ϕ](cid:0)f (t − i/2)φ−it−1/2(cid:1)φit+1/2 = ϕit+1/2(cid:0)ϕ−it−1/2f (t − i/2)(cid:1)[φ] (the left hand side is therefore depends only on [ϕ] while the right hand side depends only on [φ] and the common element in M (it + 1/2) is reasonably denoted by [ϕ]f (t − i/2)[φ]). Proof. For a ∈ M , the identity h(f (t)φ−it)φ1/2σφ,ϕ t (a)ϕ1/2i = hϕ1/2(ϕ−itf (t))φ1/2ai is analytically continued from t to t − i/2 to get h(f (t − i/2)φ−it−1/2)φ1/2φ1/2σφ,ϕ t (a)i = hϕ1/2(ϕ−it−1/2f (t − i/2))φ1/2ai (use the KMS-condition at σφ,ϕ t (a)ϕ1/2) and, after a simple rewriting, h(f (t − i/2)φ−it−1/2)φit+1/2φ1/2aϕ−iti = hϕit+1/2(ϕ−it−1/2f (t − i/2))φ1/2aϕ−iti. (cid:3) TRACE FORMULAS 11 Since [ϕ]f (t−i/2)[φ] = [ϕ]([ϕ′]f (t−i/2)[φ′])[φ] whenever [ϕ] ≤ [ϕ′] and [φ] ≤ [φ′], k[ϕ]f (t − i/2)[φ] is increasing in [ϕ] and [φ]. We claim that f (t − i/2) = lim [ϕ]→1 [φ]→1 [ϕ]f (t − i/2)[φ] exists in M (it + 1/2). In fact, if not, we can find increasing sequences ϕn and φn in M + k[ϕn]f (t − i/2)[φn]k = ∞, which contradicts with ∗ so that lim n→∞ k[ϕn]f (t − i/2)[φn]k ≤ k[ϕ]f (t − i/2)[φ]k < ∞ for the choice ϕ =P ϕn/2nkϕnk, φ =P φn/2nkφnk. Moreover, the same reasoning reveals that we can find ϕ, φ ∈ M + ∗ so that f (t − i/2) = [ϕ]f (t − i/2) = f (t − i/2)[φ]. Consequently, {f (t − i/2) ∈ M (it + 1/2)} is a norm-continuous section of Gaussian decay from the expression f (t − i/2) = f (t − i/2)[φ] =(cid:0)f (t − i/2)φ−it−1/2(cid:1)φit+1/2 which is valid for a sufficiently large φ. Remark 4. By an analytic continuation, one sees that any half-analytic section {f (t)} of {M (it)} is finitely supported in the sense that there exists φ ∈ M + ∗ satisfying f (t) = [φ]f (t)[φ] for every t ∈ R. Example 6.3. Let ϕ ∈ M + t . Then, for α > 0 and β ∈ C, f (t) = e−αt2+βtaϕitb belongs to N and its boundary section is f (t − i/2) = e−α(t−i/2)2+β(t−i/2)aϕit+1/2b. ∗ and a, b ∈ [ϕ]M [ϕ] be entirely analytic for σϕ The inner product is now introduced by (f g) =ZR (f (t − i/2)g(t − i/2)) dt =ZR hf (t − i/2)∗g(t − i/2)i dt, which is clearly positive-definite and the completed Hilbert space H is naturally identified with the direct integral M (it + 1/2) dt H =IR ξ(t) dt =IR aωisIR (cid:18)I ξ(t) dt(cid:19)∗ =IR aωisξ(t − s) dt ξ(−t)∗ dt because N provides a dense set of measurable sections in the right hand side. The Hilbert space H is then made into a *-bimodule of M (iR) by and in such a way that actions of M (it) on H are s*-continuous. Since the family {M (it+1/2)} is trivialized by obvious isomorphisms L2(M )ωit ∼= L2(M ) ∼= ωitL2(M ) in terms of a faithful weight ω on M , we have identifications H ∼= L2(M ) ⊗ L2(R) in two ways, which transforms left and right multiplications of ωit into a translational unitary by t ∈ R. Recall that our weights are orthogo- nal direct sums of bounded functionals and the multiplication of ωis on H gives a continuous one-parameter group of unitaries. With these observations in mind, it is immediate to check the axioms of Hilbert algebra: the left and right multiplications are bounded with respect to the inner product, N2 is dense in H and (f ∗g∗) = (gf ) for f, g ∈ N. 12 SHIGERU YAMAGAMI Remark 5. Note that the scaling automorphism θs satisfies (θsf )(t−i/2) = e−ist−s/2f (t− i/2) and hence scales the inner product: (θsf θsg) = e−s(f g) for f, g ∈ N. In this way, we have constructed a Hilbert algebra N. The associated von Neu- mann algebra is denoted by N = M ⋊ R and referred to as the Takesaki dual of M in what follows. The scaling automorphisms θs of N induce a *-automorphic action (also denoted by θs) of R on N by θs(l(f )) = l(θsf ), which is referred to as the dual action. Here l(f ) denotes a bounded operator on H defined by l(f )g = f g for g ∈ N. Let ω be a faithful weight on M . From the convolution form realization of N on H, one sees that N contains M as well as ωit as operators by left multiplication and these in turn generates N . Likewise right multiplications of M and ωit generates the right action of N on H. Thus the Takesaki dual of M is isomorphic to the crossed product of M with respect to the modular automorphism group {σω t }, which justifies our notation M ⋊ R for N . We record here the following well-known fact for later use together with a proof to illustrate how the essence can be easily captured in the modular algebra formalism. Theorem 6.4 (Takesaki). The fixed-point algebra N θ of N under the dual action θ is identified with M . Proof. Through H ∼= L2(M )⊗L2(R) adapted to the trivialization M (it+1/2)ω−it = L2(M ) of M (it + 1/2), the right action of ωis is realized on L2(R) by translations whereas θs by multiplication of e−ist on L2(R). Since these generate B(L2(R)) (Stone-von Neumann), N θ is identified with (B(L2(M )) ⊗ 1) ∩ End(HM ). Let a ∈ M and f, g ∈ L2(R). For ξ, η ∈ L2(M ), (η ⊗ g(ξ ⊗ f )a) = (ηξσω gf ) with σω g(t)f (t)ωitaω−it ∈ M shows that T ∈ B(L2(M )) belongs to N θ if and only if it is in the commutant of the right action of {σω h (a); a ∈ M, h ∈ L1(R)} generates M , this implies N θ ⊂ M . (cid:3) h (a); h ∈ L1(R)} on L2(M ). Since {σω gf =ZR Now we introduce some notations and conventions in connection with our Hilbert algebra: N is regarded as a *-subalgebra of N and we write Nτ 1/2 = τ 1/2N to M (it + 1/2) dt, where τ 1/2 is just a indicate the corresponding subspace in H =IR for f ∈ N, whereas f τ 1/2 = τ 1/2f =IR f (t − i/2) dt. dummy symbol but its square τ will be soon identified with the standard trace on N . Thus h ∈ N is identified with an operator on H satisfying h(f τ 1/2) = (hf )τ 1/2 Let B ⊃ N be a dense *-ideal of N such that Bτ 1/2 = τ 1/2B is the set of bounded vectors in H; y ∈ N belongs to B if and only if there exists a vector η ∈ H satisfying ηf = y(f τ 1/2) = y(τ 1/2f ) for any f ∈ N and, if this is the case, we write η = yτ 1/2 = τ 1/2y. Recall that the standard trace τ on N+ is defined by τ (y∗y) = (yτ 1/2yτ 1/2) if y ∈ B and τ (y∗y) = ∞ otherwise. Note that, for f, g ∈ N, f ∗g ∈ N2 is in the trace class and its trace is calculated by τ (f ∗g) =ZR which justifies our notation f τ 1/2. (f (t − i/2)g(t − i/2)) dt = (f τ 1/2gτ 1/2), TRACE FORMULAS 13 From the scaling relation (θsf )(t − i/2) = e−ist−s/2f (t − i/2), the inner product is scaled by a factor e−s under the *-automorphism of N and hence the associated trace τ scales like τ (θs(y∗y)) = e−sτ (y∗y) for y ∈ N . To each ξ, η ∈ H, a sesquilinear element ξ∗η ∈ N∗ is associated by hξ∗η, xi = (ξηx) and a∗bτ = τ a∗b ∈ N∗ is defined to be (aτ 1/2)∗(bτ 1/2) for a, b ∈ B. As a square root of this correspondence, we have a unitary map H → L2(N ) in such a way that aτ 1/2 7→ (a∗aτ )1/2 for a ∈ B. Therefore, if we set B+ = B ∩ N+, the closure of B+τ 1/2 = τ 1/2B+ in H corresponds to the positive cone L2(N )+. Related to these, we recall the following well-known and easily proved fact (cf. [7] Corollary 19.1). Lemma 6.5. The Hilbert space H is canonically isomorphic to the vector space of Hilbert-Schmidt class operators with respect to τ in such a way that τ (y∗y) = (yτ 1/2yτ 1/2). Note that a closed operator y affiliated to N is in the Hilbert-Schmidt class if and only if τ (y∗y) < ∞. We shall now utilize the Hilbert algebra structure behind N to set up a method 7. Trace Formula modeled after N to calculate the standard trace τ on N . Given an open interval I ⊂ [0, 1/2], let eFI be the set of M -valued analytic [φ]eFI [φ]. We write fφ(z)φiz for φ ∈ M + functions of z ∈ R−iI and set FI = ∪φ∈M + ∗ and fφ ∈ [φ]FI [φ] to indicate dummies of elements in FI . All such dummies are then identified by the relation ϕiz = (ϕiz ψ−iz)ψiz whenever ϕ ≤ ψ and the obtained quotient set (which is a kind of inductive limit of dummy elements) is denoted by LII and an element in LII is called a left interpolator on I. ∗ Thus each left interpolator is of the form f (z) = fϕ(z)ϕiz and we say that f (z) ∗ mojorizing ϕ, f (z) is supported by φ and is supported by ϕ. Then, for φ ∈ M + fφ(z) = fϕ(z)(ϕizφ−iz ), which is also denoted by f (z)φ−iz. Clearly we have a similar notion of right interpolators with the obvious notations for them. These are related by the *-operation defined by f ∗(z) = f (−z)∗: If f ∈ LII , f ∗ ∈ RII so that φ−izf ∗(z) = (f (−z)φiz)∗. A pair (l(z), r(z)) of left and right interpolators on I is called an interpolator if one can find φ ∈ M + ∗ which supports l, r and interrelates them in the following sense: For each w ∈ R − iI, the function σφ t (φ−iwr(w)) of t ∈ R is analytically extended up to the horizontal line w + R so that the function σφ z (φ−iwr(w)) is w*-analytic on D = {(z, w) ∈ C2; w ∈ R − iI, Im w ≤ Im z ≤ 0} and satisfies σφ w(φ−iwr(w)) = l(w)φ−iw. Here, for z ∈ C \ R and a ∈ M , σz(a) means that σt(a) (t ∈ R) is analytically extended to a w*-continuous function of ζ ∈ R + i Im z[0, 1] and it is evaluated at ζ = z. Since analytical extensions are moved back to the starting horizontal lines, the condition is symmetrical in the left-and-right: σφ −t(l(w)φ−iw) is analytically ex- tended to σφ w−z(φ−iwr(w)), which is w*-continuous in (z, w) ∈ D. For (z, w) ∈ TI, the relation σφ z+w(φ−i(z+w)r(z + w)) = l(z + w)φ−i(z+w) is then rewritten into w(φ−i(z+w)r(z + w)) = σφ σφ −z(l(z + w)φ−i(z+w)), which is a w*-analytic function of (z, w) ∈ TI and denoted by φ−izf (z + w)φ−iw when (l(z), r(z)) is symbolically expressed by f (z). 14 SHIGERU YAMAGAMI Moreover, the interrelating condition is compatible with the majorization changes: Let φ ≤ ω and z ∈ R − iI. Then t (ω−izr(z)) = (ω−i(z−t)φi(z−t))σφ σω t (φ−izr(z))φitω−it is analytically continued from t to z to get (l(z)φ−iz)(φiz ω−iz) = l(z)ω−iz. We say that an interpolator f (z) = (l(z), r(z)) is supported by φ ∈ M + ∗ if both l(z) and r(z) are supported by φ and, in that case, we write φf (z) = φ−izf (z) = φ−izr(z) and fφ(z) = f (z)φ−iz = l(z)φ−iz. Let II be the set of interpolators on I. By restriction or extension, IJ ⊂ II if I ⊂ J ⊂ (0, 1/2). The *-operation on II is defined by (l(z), r(z))∗ = (r∗(z), l∗(z)) so that it is compatible with the inclusions IJ ⊂ II . Notice that N can be regarded as a *-subspace of I(0,1/2). Given an asymptotic function ρ : R \ [−R, R] → [0, ∞) with R > 0 a positive real, an interpolator f (z) on I is said to have a ρ-growth and denoted by f (z) = O(ρ( Re z)) if we can find C > 0 so that kφ−izf (z + w)φ−iwk ≤ Cρ( Re (z + w)) for any (z, w) ∈ TI satisfying z + w ∈ R \ [−R, R] − iI. Note that the growth condition is well-defined thanks to the half-power analyticity for majorization. f (z)φ−iz = O(eǫ( Re z)2 An interpolator f is said to be of sub-gaussian growth if, for any small ǫ > 0, I be the set of interpolators of sub-gaussian growth. I with I = (α, β) ⊂ [0, 1/2], we here introduce a sesqui-linear form on For f ∈ Ig ). Let Ig N as follows. Continuous functions F (s, t) =(cid:0)h(t − i/2)fφ(s − ir)φit+1/2 φg(t − s + ir − i/2)(cid:1) of (s, t) ∈ R2 parametrized by r ∈ I are of Gaussian decay with their absolutely convergent integrals independent of r ∈ I owing to Cauchy's integral theorem. Moreover F (s, t) does not depend on the choice of supporting φ either. Thus a sesqui-linear form h if on N is well-defined by hhgif =ZR2 =ZR2 dsdt(cid:0)h(t − i/2)fφ(s − ir)φit+1/2 dsdt(cid:0)h(t − i/2)(g∗)φ(−t + s + ir − i/2)φi(s−t)fφ(s − ir)φis+1/2(cid:1) φg(t − s + ir − i/2)(cid:1) as far as r ∈ I and φ ∈ M + supports f and g, which behaves well under the ∗ *-operation: hghif ∗ = hhgif . Notice that, when f ∈ N, hhgif is reduced to (hτ 1/2f gτ 1/2). We interprete the sequilinear form h if as defining an operator K in a kernel form by (hτ 1/2K(gτ 1/2)) = hhgif , which is referred to as the virtual operator of f (z) and denoted by f itself. Note that the *-operation on interpolators is compatible with the associated vir- tual operators; (hτ 1/2f (gτ 1/2)) = (gτ 1/2f ∗(hτ 1/2)) for g, h ∈ N, and virtual op- erators are affiliated to N in the sense that (hk∗τ 1/2f (gτ 1/2)) = (hτ 1/2f (gkτ 1/2)) for g, h, k ∈ N. Let D(f ) be the set of vectors gτ 1/2 ∈ Nτ 1/2 which makes the conjugate-linear functional hτ 1/2 7→ (hτ 1/2f (gτ 1/2)) bounded. For gτ 1/2 ∈ D(f ), if the vector ξ ∈ H satisfying (hτ 1/2ξ) is denoted by f (gτ 1/2), then we obtain a linear operator on H by D(f ) ∋ gτ 1/2 7→ f (gτ 1/2) ∈ H. A virtual operator is said to be densely defined if D(f ) is dense in H. When the sesqui-linear form h if itself is bounded, D(f ) = Nτ 1/2 and the associated linear TRACE FORMULAS 15 operator Nτ 1/2 → H is bounded and identified with an element y ∈ N in such a way that hhgif =(cid:0)hτ 1/2y(gτ 1/2)(cid:1) for g, h ∈ N. We next introduce the virtual vector as a conjugate-linear form on N2τ 1/2. ∗ supports g, h ∈ N, then vector-valued functions (hg∗)φ(s)φ1/2 Lemma 7.1. If φ ∈ M + and φ1/2 continuous functions (hg∗)φ(z)φ1/2 and φ1/2 these are of Gaussian decay and, for 0 ≤ r ≤ 1/2, satisfy φ(hg∗)(s) of s ∈ R are analytically continued to L2(M )-valued norm- φ(hg∗)(z) of z ∈ R − i[0, 1] so that (hg∗)φ(s − i(1 − r))φ1/2 =ZR φ(hg∗)(s − i(1 − r)) =ZR φ1/2 respectively. h(t − i/2)(g∗)φ(−t + s + ir − i/2)φ−it dt, φit φh(t + s + ir − i/2)g∗(−t − i/2) dt Proof. We already know that (hg∗)φ(s) has an s*-continuous analytic extension (hg∗)φ(z) ∈ M to z ∈ R − i[0, 1/2] so that (hg∗)φ(s − i/2)φ1/2 = f (s − i/2)φ−is, whereas (hg∗)φ(s − i/2)φ1/2 =Z h(t − i/2)g∗ φ(−t + s)φ−it dt is analytically continued to the norm-continuous function Z h(t − i/2)g∗ φ(−t + z)φ−it dt of z ∈ R − i[0, 1/2], which is of Gaussian decay as a convolution of functions of Gaussian decay. (cid:3) The sesqui-linear form hhgif is now expressed by hhgif =ZR ds(cid:0)(hg∗)φ(s − i(1 − r))φis+1/2fφ(s − ir)φis+1/2(cid:1), whenever 0 < r < 1 and φ supports g, h as well as f , which reveals that a conjugate- linear form f τ 1/2 on N2τ 1/2 is well-defined by the relation (hg∗τ 1/2f τ 1/2) = hhgif and called the vitual vector of f . Note that the virtual vector of f ∗ is given by (f τ 1/2)∗ which is defined by (ξ(f τ 1/2)∗) = (ξf τ 1/2) for ξ ∈ N2τ 1/2: hhgif ∗ = hghif = (gh∗τ 1/2f τ 1/2) = (hg∗τ 1/2(f τ 1/2)∗). These are also referred to as a boundary operator and a boundary vector for I = (0, ν) and I = (ν, 1/2) with additional notationsR f (t) dt andH f (t− i/2) dt respectively. We now focus on these. Boundary Operator: boundary operators, the following illustrates the meaning of boundary (limit). In extracting linear operators from the kernel form of Let D be the set of s*-continuous sections of {M (it + 1/2)} of Gaussian de- cay, which is a topological vector space of inductive limit of Banach spaces Dδ = {{ξ(t)} ∈ {M (it + 1/2)}; kξkδ < ∞} with kξkδ = sup{eδt2 kξ(t)k; t ∈ R}. The is norm-convergent in Dδ′ for any δ′ < δ and gives a bounded linear map Dδ → Dδ′, which depends continuously on r ∈ I in the norm-topology of B(Dδ, Dδ′). The induced continuous linear operator on D is then denoted by RR fϕ(s − ir)ϕis ds. We say that RR fϕ(s − ir)ϕis ds is bounded if it is bounded as a densely defined Note that, ifRR fϕ(s − ir)ϕis ds ∈ B(H) is locally norm-bounded for r ∈ I, it is s-continuous in r ∈ I by the density of D in H. Lemma 7.2. Let f ∈ Ig linear operator on H. I be supported by ϕ ∈ M + ∗ . Assume that D ∋ ξ 7→ZR fϕ(s − ir)ϕisξ ds ∈ H gives rise to a bounded linear operator yr =RR fϕ(s − ir)ϕis on H and f (s − i0) ds = lim fϕ(s − ir)ϕis ds y =ZR r→+0ZR 16 SHIGERU YAMAGAMI embedding Dδ → H is norm-continuous and therefore so is D → H. For f ∈ Ig I with I = (0, ν) and ξ ∈ Dδ ZR fϕ(s − ir)ϕitξ ds exists in the w*-topology of N . Then the boundary operator of f (z) is bounded and given by the above limit. Proof. Given g ∈ N and ϕ ∈ M + majorizes ϕ. Then, ∗ , choose φ ∈ M + ∗ so that it supports g and R2 ∋ (s, t) 7→ ϕisf (t − s − i/2)φ−it ∈ L2(M ) is analytically extended to an L2(M )-valued norm-continuous function (ϕiz φ−iz)(cid:0)φizg(t− z − i/2)(cid:1) of z ∈ R − i[0, 1/2] and t ∈ R, which is denoted by ϕizg(t − z − i/2). Since ϕizg(t − z − i/2) = (ϕizφ−iz)φit+1/2 φg(t − z − i/2), (ϕizg)(t − i/2) = ). Thus, ϕizg(t − z − i/2) belongs to Dδ as a function of t ∈ R if gφ(t) = O(e−δt2 ξr(t) = ϕrg(t + ir − i/2) is a Dδ-valued norm-analytic function of r. By our assumptions, s*-continuous family {yr}r∈I in N converges to y in w*- topology as r → +0, whence the operator norm kyrk is bounded in a neighborhood of r = 0 and we see that yrξr = lim r→0 yrξr = lim (r′,r′′)→(0,0) yr′ξr′′ = lim r′→0 yr′ξ0 = yξ0. Now the identity Z fφ(s − ir)φit+1/2 is used to get φg(t − s + ir − i/2) ds = (yrξr)(t) hhgif = (hτ 1/2yrξr) = (hτ 1/2yξ0) = (hτ 1/2y(gτ 1/2)). (cid:3) Corollary 7.3. Let f (z) be an interpolator on I = (0, ν) (0 < ν ≤ 1/2) and suppose that f is supported by a φ ∈ M + ∗ so that fφ(z) = f (z)φ−iz is a scalar operator of polynomial growth with its horizontal Fourier transform RR fφ(s − TRACE FORMULAS 17 r → 0, then the boundary operator of f (z) is a bounded operator ir)eisλ ds being in L∞(R) for a small r > 0 and w*-converging to cfφ ∈ L∞(R) as Here E(·) denotes the spectral measure of φit: φit =RR eitλE(dλ). cfφ(log φ) =ZRcfφ(λ)E(dλ) ∈ N. Proof. Due to the left trivialization [φ]L2(N ) ∼= L2(R) ⊗ [φ]L2(M ), the whole thing is reduced to L∞(R) on L2(R) and the classical harmonic analysis on the real line works. (cid:3) Example 7.4. If f (z)φ−iz extends to a bounded w*-continuous M -valued function of z ∈ R−i[0, ν) in such a way that there exists an integrable function ρ(t) satisfying kf (t − ir)φ−i(t−ir)k ≤ ρ(t) for t ∈ R and 0 ≤ r < ν, then the boundary operator is bounded and hence belongs to N . Example 7.5. For φ ∈ M + on I with I specified according to µ as follows: ∗ and µ ∈ C, consider an interpolator f (z) = 1 µ+iz φiz (i) I = (0, 1/2) ( Re µ ≥ 0). The boundary operator is given by 2π(1 ∨ φ)−µ. (ii) Either I = (0, − Re µ) (−1/2 < Re µ < 0) or I = (0, 1/2) ( Re ≤ −1/2). Then the boundary operator is given by −2π(1 ∧ φ)−µ for Re µ < 0. Here, with the help of a spectral decomposition φit =RR eitλE(dλ), e−µλE(dλ), e−µλE(dλ). (1 ∧ φ)−µ =Z 0 −∞ (1 ∨ φ)−µ =Z ∞ 0 Boundary Vector: We next look into boundary vectors. Let f (z) ∈ Ig I = (ν, 1/2) and g, h ∈ N. In the expression I with (hg∗τ 1/2f τ 1/2) =ZR ds(cid:0)(hg∗)ϕ(s − i(1 − r))ϕis+1/2fϕ(s − ir)ϕis+1/2(cid:1) ∗ ), notice that the norm-convergence (hg∗)ϕ(s − i(1 − r))ϕ1/2 = (hg∗)(s − i/2)ϕ−is in L2(M ) is uniformly in s ∈ R (g, h and f (z) being supported by ϕ ∈ M + lim r→1/2 and the domination k(hg∗)ϕ(s − i(1 − r))ϕ1/2k ≤ Ce−δs2 whereas kfϕ(s − ir)k = O(eǫs2 ) uniformly in r for any ǫ > 0. holds uniformly in r, Thus, if f (z) satisfies the condition that (i) ρϕ(s) = sup{kfϕ(s − ir)k; r ∈ (ν, 1/2)} is a locally integrable function of s ∈ R for some supporting ϕ and (ii) we can find a locally integrable measurable section η(s) ∈ M (is + 1/2) so that, for a sufficiently large φ and for almost all s, fφ(s − ir)φis+1/2 converges weakly to η(s) in M (is + 1/2) as r → 1/2, then we have the expression (hg∗τ 1/2f τ 1/2) =ZR(cid:0)(hg∗)(s − i/2)η(s)(cid:1) ds, which shows that the boundary vector of f (z) is represented by the measurable section η(s) ∈ M (is + 1/2). Note here that kη(s)k is of sub-gaussian growth. Example 7.6. If fφ(z) is extended to an M -valued w*-continuous function of z ∈ R − i(ν, 1/2], then ρφ(s) is locally bounded and η(s) = fφ(s − i/2)φis+1/2 = f (s − i/2) meets the requirements. 18 SHIGERU YAMAGAMI Example 7.7. Again consider f (z) = 1 if −1/2 < Re µ < 0 and I = (0, 1/2) otherwise. µ+iz ϕiz on I but this time I = (− Re µ, 1/2) Then, for Re µ 6= −1/2, the boundary vector of f belongs to H and is given by f (t − i/2) = (µ + it + 1/2)−1ϕit+1/2. When Re µ 6∈ [−1, −1/2], the expression (kτ 1/2f τ 1/2) =ZR 1 it + µ + 1/2 hk∗(−t − i/2)ϕit+1/2i dt for k ∈ N2 is analytically changed in the integration variable to get (kτ 1/2f τ 1/2) =ZR φ(k∗(−t)ϕit) it + µ + 1 dt. Thus the parametric limit of f τ 1/2 exists in simple convergence as µ approaches to a point in Re µ = −1/2 from the right ( Re µ > −1/2). Now let µ = im − 1/2 (m ∈ R) be on the critical line Re µ = −1/2 and set ǫ = 1/2 − r. By Lemma 7.1, we have hhgif =ZR 1 i(s + m) + 1/2 φ(cid:0)φis(gh∗)(−s)(cid:1) ds, which reveals that the boundary vector of f (z) coincides with ǫ→+0IR lim 1 i(t + m) + ǫ φit+1/2 dt. We now generalize the notion of interpolators on I = (0, 1/2) so that f (z) is allowed to be not defined on a compact subset K of R − i(0, 1/2). The various analyticity is then defined just avoiding K. Since the growth condition is about horizontal asymptotics, it remains having a meaning as well. We introduce the residue operator Rf =HK f (z) dz : Nτ 1/2 → H by Rf (gτ 1/2) =IR(cid:18)I f (z)g(t − z − i/2) dz(cid:19) dt. Here f (z)g(t − z − i/2) = fφ(z)φit+1/2 φg(t − z − i/2) is an M (it + 1/2)-valued analytic function of z ∈ (R − i[0, 1/2]) \ K and the coutour integral is performed by surrounding K. Theorem 7.8 (Trace Formula). Let f (z) be an interpolator on (0, 1/2) of sub- gaussian growth and assume that the boundary vector f τ 1/2 = HR f (t − i/2) dt Then the sum of the boundary operator f and the residue operator Rf is τ - exists in H. measurable and we have τ ((f + Rf )∗(f + Rf )) = (f τ 1/2f τ 1/2) =ZR(cid:0)f (t − i/2)f (t − i/2)(cid:1) dt. Proof. Let Vf be the virtual operator of f (z) (z ∈ R − i(1/2 − ǫ, 1/2)). By the residue formula, Vf = f + Rf and, for g, h ∈ N, (hτ 1/2Vf (gτ 1/2)) = (hτ 1/2(f τ 1/2)g) = (h1/2τ 1/2l(f τ 1/2)(gτ 1/2)) shows that the virtual operator Vf is closable with its closure given by l(f τ 1/2). Lemma 6.5 is then the applied to get the assertion. (cid:3) TRACE FORMULAS 19 Corollary 7.9. If f (z) is analytic on the whole R − i(0, 1/2) additionally, then the boundary operator f is τ -measurable and we have τ (f ∗f )) = (f τ 1/2f τ 1/2) =ZR(cid:0)f (t − i/2)f (t − i/2)(cid:1) dt. Example 7.10. Let G ∈ L2(R) and suppose that its Fourier transform bG(λ) = RR G(t)e−iλt dt is integrable and satisfiesR ∞ Then the inverse Fourier transform Gw of bG(λ)eiwλ belongs to L2(R) ∩ C0(R) and depends on w ∈ R− i[0, 1/2] norm-continuously for both k ·k∞ and k ·k2. Since, for F ∈ L2(R), 0 bG(λ)2eλ dλ < ∞. 2πZR bF (λ)bG(λ)eiwλ dλ 1 is analytic in w and Gs is reduced to the translation G(t + s) of G(t), F (t) is analytically extended to F (z) so that Gw(t) = G(t + w) for w ∈ R − i[0, 1/2] and t ∈ R. Now, for φ ∈ M + ∗ , g(z) = G(z)φit defines an interpolator on (0, 1/2) which vanishes at Re z = ±∞. Since φit on H is given by translation on L2(R) ⊗ [φ]L2(M )[φ] ∼= [φ]H[φ], the associated boundary operator is bounded and the boundary vector is given byHR G(t − i/2)φit+1/2 dt so that (g(t − i/2)f (s)g(t − s − i/2)) dsdt τ (g∗f g) =ZR2 (F Gw) = G(t − i/2)F (s)G(t − s − i/2) dsdt = φ(1)ZR2 2π ZR bF (λ)bG(λ)2eλ dλ. φ(1) = Here, for F ∈ L1(R), an L1-section {f (t)} of {M (it)} is defined by f (t) = F (t)φit and f =RR f (t) dt ∈ N . Thus, letting A be the W*-subalgebra of [φ]N [φ] generated by {φit; t ∈ R}, L2(A, τ ) is identified with L2(R, eλ dλ) by a unitary map Uφ : L2(A, τ ) ∋ gτ 1/2) 7→r φ(1) 2π bG(λ) ∈ L2(R, eλ dλ) so that φis on L2(A, τ ) is realized by a multiplication of the function e−isλ of λ ∈ R. 1 Example 7.11. For −1/2 < Re β < 0 and φ ∈ M + ∗ , the interpolator f (z) = β+iz φiz has −2π(1 ∧ φ)−β ∈ N as the boundary operator. The residue operator is calculated by the realization L∞(A) on L2(A) as Zz−iβ=ǫ 1 β + iz eiλz dz = 2πe−βλ, which is therefore 2πφ−β. Adding these, we see that (1 ∨ φ)−β is in the Hilbert- Schmidt class and hence, for x ∈ M and µ = −r + is ∈ −(0, 1) + iR, x(1 ∨ φ)−µ = x(1 ∨ φ)r/2−is(1 ∨ φ)r/2 is in the trace class with 2πτ (x(1 ∨ φ)−µ) = φ(x)ZR 1 1 −it + (1 − r)/2 i(t + s) + (1 − r)/2 dt = φ(x) is − r + 1 = φ(x) µ + 1 . 20 SHIGERU YAMAGAMI Although Haagerup deals only with the case µ = 0 and its scaled variation, the following generalization should also be attributed to him. Theorem 7.12 (Haagerup's Trace Formula). Let ω be a weight on M in our sense. The trace of a positive operator (1 ∨ω)−µ with µ ∈ R, which belongs to N for µ ≥ 0 and is affiliated to N for µ < 0, is given by τ ((1 ∨ ω)−µ) =( ω(1) 2π(µ+1) ∞ if µ > −1, otherwise. Moreover, when ω ∈ M + ∗ , for any x ∈ M and µ ∈ (−1, ∞)+iR, the τ -measurable operator x(1 ∨ ω)−µ is in the trace class and we have τ (x(1 ∨ ω)−µ) = ω(x) 2π(µ + 1) . ∗ . Then ωit is realized as a multiplication operator on Proof. Assume ω ∈ M + L2(R, eλ dλ) by a function e−itλ of λ ∈ R. Consequently (1 ∨ ω)−µ is represented by the function 1(−∞,0](λ)eλµ of λ, which is integrable relative to the measure eλ dλ if and only if Re µ > −1 with Z 0 −∞ eλµeλ dλ = 1 µ + 1 . M + Since our weights are orthogonal sums of elements in M + ∗ remains valid for weights. The remaining part is already covered in Example 7.11. ∗ , the formula for ω ∈ (cid:3) Remark 6. (i) By the integral expression ZR 1 µ + it ωit dt of 2π(1 ∨ ω)−µ, the formula coincides with the one obtained from the formal argument. (ii) The normalization of our trace is different from that in [6] and [9] by a factor 2π. Thus, for ω ∈ M + ∗ , the analytic generator h of ωit as a positive operator on H, which satisfies θs(h) = e−sh (called relative invariance of degree −1), is τ - measurable in the sense that limr→∞ τ ([r ∨ h]) = 0. Haagerup's ingeneous ob- servation is that the whole Lp(M )'s are captured as measurable operators on H satisfying relative invariance of degree −1/p. We now go into the reverse problem of characterizing τ -measurable positive op- erators satisfying relative invariance of degree −1, which is the heart of Haagerup's correspondence. Recall the original approach to this problem: First establish a one-to-one cor- respondence between normal weights on M and θ-invariant normal weights on N . Second the latter is then paraphrased into positive operators of relative invariance of degree −1 by taking Radon-Nikodym derivative with respect to τ . Finally, posi- tive operators associated to M + ∗ are characterised as τ -measurable operators among these. Formally the whole processes look natural and seem harmless but it is in fact supported by clever and effective controls over infinities based on extended positive parts. We shall here present an inelegant but down-to-earth proof by continuing ele- mentary Fourier calculus. TRACE FORMULAS 21 8. Haagerup Correspondence Let h ≥ 0 be a τ -measurable operator on H satisfying θs(h) = e−sh for s ∈ R. Our first task here is to identify hit with ϕit for some ϕ ∈ M + ∗ . Let e = [1 ∨ h] be the support projection of 1 ∨ h. By the relative invariance of h, θs(e) is the support projection of es ∨ h and we have a Stieltjes integral representation of h and set −∞ es dθs(e) =Z ∞ h = −Z ∞ (1 ∨ h)−µ = −Z ∞ 0 −∞ e−µs dθs(e), e−sdθ−s(e) which is τ -measurable for any µ ∈ C in view of τ (e) < ∞. Notice that θs(e) is continuous in s ∈ R and dθs has no spectral jumps. Let x ∈ M and start with the computation τ (hx(1 ∨ h)−µ) = τ (x(1 ∨ h)−µh) = τ (x(1 ∨ h)1−µ) = −Z ∞ = τ (xe)Z ∞ 0 0 e(1−µ)sdτ (xθs(e)) = −Z ∞ 0 e−µsds = τ (xe), 1 µ e(1−µ)sd(e−s)τ (xe) which is valid for Re µ > 0. For t ∈ R, σt(x) = hitxh−it (x ∈ M ) defines an automorphic action of R on M because hitxh−it is θ-invariant in view of θs(hit) = e−isthit. We claim that ϕ(x) = 2πτ (xe) satisfies the KMS-condition for the automorphic action σt. First notice that [h] = [ϕ]. In fact, from the definition of ϕ and the faithfulness of the standard trace, (1 − [ϕ])e = 0, which means that e ≤ [ϕ] and then [h] = lims→−∞ θs(e) ≤ θs([ϕ]) = [ϕ]. Conversely, from (1 − [h])e = 0, 1 − [h] ≤ 1 − [ϕ] gives the reverse inequality. Now consider ϕ(x∗σt(x)) = τ (x∗hitxh−ite) with x ∈ M . If the Stieltjes integral expression for h is used as in xh−ite = −Z ∞ 0 e−istdθs(xe), we have − 0 1 2π ϕ(x∗σt(x)) =Z ∞ =Z ∞ =Z ∞ −τ (x∗hitxe) =Z 0 −∞ 0 0 and then together with e−istdτ (x∗hitθs(xe)) e−istdτ(cid:16)θs(cid:0)x∗θ−s(hit)xe(cid:1)(cid:17) e−istd(e−seist)τ (x∗hitxe) = (it − 1)τ (∗hitxe) eistdτ(cid:16)x∗θs(e)xe(cid:17) +Z ∞ 0 eistdτ(cid:16)x∗θs(e)xe(cid:17), eistd(cid:16)e−sτ(cid:0)x∗exθ−s(e)(cid:1)(cid:17) eiste−sτ(cid:0)x∗exθ−s(e)(cid:1) ds, 0 Z ∞ eistdτ(cid:16)x∗θs(e)xe(cid:17) =Z ∞ eiste−sdτ(cid:16)x∗exθ−s(e)(cid:17) −Z ∞ =Z ∞ 0 0 0 22 SHIGERU YAMAGAMI reveals that −τ (x∗hitxe) is analytically extended to a bounded continuous function −τ (x∗hizxe) =Z ∞ −∞ eiszdτ(cid:16)x∗θs(e)xe(cid:17) =Z ∞ eisze−sdτ(cid:16)x∗exθ−s(e)(cid:17) −Z ∞ eiszd(cid:16)e−sτ (x∗exθ−s(e))(cid:17) eisze−sτ(cid:0)x∗exθ−s(e)(cid:1) ds. −∞ −∞ =Z ∞ −∞ of z = t − ir ∈ R − i[0, 1]. In these and the following calculations, note that τ (x∗exθ−s(e)) (τ (x∗θs(e)xe)) is Consequently, with the notation ϕ(x∗σz(x)) for the analytic continuation of positive, increasing (decreasing) and continuous in s ∈ R, whence both dτ(cid:16)x∗exθ−s(e)(cid:17) and −dτ(cid:16)x∗θs(e)xe(cid:17) give rise to positive finite measures on R. e(it+r−1)sdτ(cid:16)x∗exθ−s(e)(cid:17) ϕ(x∗σt−ir(x)) = (it + r − 1)Z ∞ ϕ(x∗σt(x)) and, with the help of integration-by-parts, we get the expression 1 2π −∞ = (it + r)Z ∞ For 0 < r < 1, we see lim s→∞ −∞ −∞ − (it + r − 1)Z ∞ e(it+r−1)sτ(cid:0)x∗exθ−s(e)(cid:1) ds e(it+r−1)sdτ(cid:16)x∗exθ−s(e)(cid:17) −he(it+r−1)sτ(cid:0)x∗exθ−s(e)(cid:1)i∞ e(it+r)sτ(cid:0)x∗θs(e)xe(cid:1) = 0 e(it+r−1)sτ(cid:0)x∗exθ−s(e)(cid:1) = lim ϕ(x∗σt−ir(x)) = (it + r)Z ∞ e(it+r−1)sdτ(cid:16)x∗exθ−s(e)(cid:17). e(it+r−1)sτ(cid:0)x∗exθ−s(e)(cid:1) = 0 and s→−∞ −∞ . −∞ lim s→−∞ 1 2π at the boundary values and therefore Since both sides are continuous in r ∈ [0, 1], the equality holds at the boundary as well. We now compare this expression with 1 2π ϕ(σt(x)x∗) = τ (ehitxh−itx∗) = −Z ∞ 0 eistdτ(cid:16)θs(e)xh−itx∗(cid:17) 0 = −Z ∞ = −Z ∞ = (it + 1)Z ∞ eistdτ(cid:16)θs(cid:0)exθ−s(h−it)x∗(cid:1)(cid:17) eistdτ(cid:16)exθ−s(e)x∗(cid:17) −∞ 0 eistd(e−s−ist)τ (exh−itx∗) = (it + 1)τ (exh−itx∗) to conclude that ϕ(x∗σt−i(x)) = ϕ(σt(x)x∗) for t ∈ R. So far we have checked that hitxh−it = ϕitxϕ−it for x ∈ [ϕ]M [ϕ]. Then u(t) = hitϕ−it is a unitary in the center of [ϕ]M [ϕ]. Since each ϕit commutes with the reduced center, {u(t)} is a one-parameter group of unitaries in the reduced algebra. Let u(t) = RR eist E(ds) be the spectral decomposition in [ϕ]M [ϕ]. Then an = R[−n,n] es/2 E(ds) is an increasing sequence of positive elements in the reduced center n = hit[an] = [an]hit for t ∈ R. Set hn = h[an] = and ϕn = anϕan ∈ M + ∗ satisfies ϕit TRACE FORMULAS 23 [an]h, which is also τ -measurable and satisfies θs(hn) = e−shn. From the equalities ϕn(x) 2πµ = τ (x(1 ∨ ϕn)1−µ) = τ (x(1 ∨ hn)1−µ) = τ (x[an](1 ∨ h)1−µ) = ϕ(x[an]) 2πµ for x ∈ M and µ ≥ 1, one sees that ϕn = ϕ[an] = [an]ϕ and then ϕit t ∈ R. Finally we have hit = limn→∞ hit n = limn→∞ ϕit[an] = ϕit. n = ϕit[an] for We next check the additivity of the correspondence hϕ ↔ ϕ. To see this, we first establish the following relation. Lemma 8.1. Let ω ∈ M + ∗ and µ > 0. Then (1 ∨ ω)−µ = 1 2πZR 1 µ + it ωit dt is in the τ -trace class and, for x ∈ [ϕ]M , we have 1 τ (hx∗(1 ∨ ω)−µx) = ϕ(x∗x). Recall here that (1 ∨ ω)−µ/2 = way that (1 ∨ ω)−µ/2τ 1/2 = ωit dt belongs to B+ in such a 2πµ 1 it + µ/2 1 1 2πZR 2πIR 1 it + (µ + 1)/2 ωit+1/2 dt. τ (hy) = − lim The identity is checked as follows: Letting y = x∗(1 ∨ ω)−µx, we have = lim = lim −n n→∞Z n n→∞(cid:18)Z n n→∞(cid:18)Z n n→∞(cid:18)Z n −n −n −n = lim esdτ(cid:0)θs(e)y(cid:1) esτ(cid:0)θs(e)y(cid:1) ds − enτ(cid:0)θn(e)y(cid:1) + e−nτ(cid:0)θ−n(e)y(cid:1)(cid:19) τ(cid:0)θs(e)x∗ωitx(cid:1) − τ(cid:0)eθ−n(y)(cid:1)(cid:19) ds esZR dsZR 2πZR τ (ex∗ωitx) − eint µ + it 2π(µ + it) 2π(µ + it) eist dt dt 1 1 τ (ex∗ωitx) dt(cid:19) . By the lemma below, the function τ (ex∗ωitx)/(µ + it) is integrable, whence Lemma 8.2. We have n→∞ZR lim eint µ + it τ (ex∗ωitx) dt = 0. τ (ex∗ωitx) = 1 2π(1 − it) ϕ(x∗ωitxϕ−it). Proof. From the expression τ (ex∗ωisx) = (xeτ 1/2ωisxeτ 1/2) with ωisxeτ 1/2 = dt 1 i(t − s) + 1/2 ωisxϕ−isϕit+1/2, 1 2πIR τ (ex∗ωisx) = = = 1 (2π)2ZR (2π)2ZR 1 1 1 2π 1 − is dt dt 1 1 −it + 1/2 i(t − s) + 1/2 1 1 −it + 1/2 i(t − s) + 1/2 ϕ(x∗ωisxϕ−is), (xϕit+1/2ωisxϕ−isϕit+1/2) ϕ(x∗ωisxϕ−is) 24 SHIGERU YAMAGAMI To deal with the first term in the last expression of τ (hy), we use the relation (cid:3) 2π(µ + it)−1 = g∗ ∗ g for g(t) = 1/(it + µ/2) to see that ZR eist µ + it ωit dt =ZR g(t′)ω−it′ZR dt′ e−ist′ dt eistg(t)ωit and hence 2πZR eist µ + it (xξ0ωitxξ0) dt dt′ eist′ g(t′)ωit′ dt eistg(t)ωitxξ0) = (ZR (ZR =Xj =Xj ZR g(t′)ωit′ xξ0ZR xξ0δj)(δjZR cFj(s)2, j ∗ Fj)(t) =Xj dt′ eist′ dt eist(F ∗ dt eistg(t)ωitxξ0) belong to L2(R). where {δj} is an orthonormal system in H supporting vectors {ωitxξ0}t∈R and The Plancherel formula is then applied to each Fj to get cFj(s)2 ds =Xj Z ∞ Fj(t) = g(t)(δj ωitxξ0) together with their Fourier transformscFj(s) =RR eistFj (t) dt (2π)2τ (hy) =Z ∞ −∞Xj = 2πZRXj (ξθs(y)ξ) =ZR 2πZR =ZR cFj(s)2 ds = 2πXj ZR Similarly and more easily, the side identity follows from dsZR dsZR µ + it e−ist µ + it (xξωitxξ) = (ξx∗xξ) Fj (t)2 dt = (xξ0xξ0) = τ (ex∗x). (xξθs(ωit)xξ) Fj(t)2 dt dt dt 2π µ 2π µ −∞ 1 2π µ for each ξ ∈ L2(N ). Theorem 8.3 (Haagerup correspondence). There is a linear isomorphism between M∗ and the linear space of τ -measurable operators h on L2(N ) satisfying θs(h) = e−sh and so that ϕ ∈ M + ∗ corresponds to the analytic generator hϕ of the one- parameter group {ϕit} of partial isometries in N . Moreover the correspondence preserves N *-bimodule structures as well as posi- tivity. Proof. The correspondence is already established for positive parts and Lemma 8.1 is used to get the additivity by 1 2πµ φ(x∗x) = h(hϕ + hψ)x∗(1 ∨ ω)−µxi = 1 2πµ (ϕ(x∗x) + ψ(x∗x)). Here ϕ, ψ ∈ M + ∗ and φ ∈ M + ∗ is specified by hφ = hϕ + hψ. Once the semilinearity is obtained, the other part is almost automatic. The linear extension is well-defined by hϕ = hϕ1 −hϕ2 +ihϕ3 −ihϕ4 for ϕ = ϕ1−ϕ2+iϕ3−iϕ4 ∈ TRACE FORMULAS 25 M∗ with ϕj ∈ M + Lemma 8.1 as ∗ . The identity ahϕa∗ = haϕa∗ for a ∈ M follows again from τ (aha∗x∗(1 ∨ ω)−µx) = τ (ha∗x∗(1 ∨ ω)−µxa) = and then ahϕb∗ = haϕb∗ by polarization. 2π µ ϕ(a∗x∗xa) = 2π µ (aϕa∗)(x∗x) (cid:3) References [1] O. Bratteli and D.W. Robinson, Operator algebras and quantum statistical mechanics, Vol.1, Springer, 1979. [2] U. Haagerup, On the dual weights for crossed products of von Neumann algebras I, Math. Scand., 43(1978), 99 -- 118. [3] U. Haagerup, On the dual weights for crossed products of von Neumann algebras II, Math. Scand., 43(1978), 119 -- 140. [4] U. Haagerup, Operator valued weights in von Neumann algebras I, J. Funct. Anal., 32(1979), 175 -- 206. [5] U. Haagerup, Operator valued weights in von Neumann algebras II, J. Funct. Anal., 33(1979), 339 -- 361. [6] U. Haagerup, Lp-spaces associated with an arbitrary von Neumann algebra, Colloques Inter- nationaux CNRS, No. 274, 175 -- 184, 1979. [7] I.E. Segal, A non-commutative extension of abstract integration, Annal Math., 57(1953), 401 -- 457. [8] M. Takesaki, Theory of operator algebras, Vol.2, Springer, 2003. [9] M. Terp, Lp-spaces associated with von Neumann algebras, Kbenhavns Univ., 1981. [10] S. Yamagami, Algebraic aspects in modular theory, Publ. RIMS, 28(1992), 1075 -- 1106. [11] S. Yamagami, Modular theory for bimodules, J. Funct. Anal., 125(1994), 327 -- 357.
1701.02052
2
1701
2017-06-27T21:48:41
Bicommutant categories from conformal nets
[ "math.OA", "math-ph", "math-ph", "math.RT" ]
We prove that the category of solitons of a finite index conformal net is a bicommutant category, and that its Drinfel'd center is the category of representations of the conformal net. In the special case of a chiral WZW conformal net with finite index, the second result specializes to the statement that the Drinfel'd center of the category of representations of the based loop group is equivalent to the category of representations of the free loop group. These results were announced in [arXiv:1503.06254].
math.OA
math
Bicommutant categories from conformal nets Andr´e Henriques Abstract We prove that the category of solitons of a finite index conformal net is a bicom- mutant category, and that its Drinfel'd center is the category of representations of the conformal net. In the special case of a chiral WZW conformal net with finite index, the second result specializes to the statement that the Drinfel'd center of the category of representations of the based loop group is equivalent to the category of representations of the free loop group. These results were announced in [Hen15]. Contents 1 Introduction and statement of results 1.1 Motivations from Chern-Simons theory . . . . . . . . . . . . . . . . . . . 1.2 Representations and solitons . . . . . . . . . . . . . . . . . . . . . . . . . 1.3 Bicommutant categories . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.4 Main results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2 Bicommutant categories 2.1 The commutant of a tensor category . . . . . . . . . . . . . . . . . . . . 2.2 Bi-involutive tensor categories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3 Bim(R) 3 Conformal nets . . . . . . . . . . . . . . . . . . . . . . . 3.1 Coordinate free conformal nets 3.2 Fusion of representations . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3 The braiding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4 Solitons . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.1 Solitons as bimodules 4.2 The braiding between T − A and T +A . . . . . . . . . . . . . . . . . . . . . 4.3 The absorbing object . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.4 The Drinfel'd center . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2 3 4 6 7 8 8 9 11 11 12 14 15 20 20 22 25 27 1 1 Introduction and statement of results In [Hen15], we made the announcement that, at least for G = SU (n), the Drinfel'd center of the category of locally normal1 representations of the based loop group is equivalent, as a braided tensor category, to the category of locally normal representa- tions of the free loop group: Z(cid:0)Repk(ΩG)(cid:1) ∼= Repk(LG). (1) One of the main goals of this paper is to establish the above relation (see Theorem 1.1 for a precise statement). It should be noted that the representation theory of based loop groups had not been considered before. The mere fact that the fusion product makes sense for these representations is, in itself, remarkable. The broader relevance of the above result comes from topological quantum field theory (TQFT), specifically from Chern-Simons theory. There are two main classes of topological quantum field theories in dimension three: theories of Turaev-Viro type, associated to fusion categories [TV92, BW96], and theories of Reshetikhin-Turaev type, associated to modular tensor categories [RT91, BK01] (Chern-Simons theories are of the latter kind). Since the groundbreaking work of Jacob Lurie on the classification of extended TQFTs [Lur09], it has been an important question to determine which theories fit into that formalism; a theory for which that is the case is said to "extend down to points". It is broadly accepted (even though this has not yet been proven) that theories of Turaev-Viro type extend down to points [DSPS13, Wra10]. On the other hand, for a typical Reshetikhin-Turaev theory, it was generally thought that this should not be possible (the results in [DMNO13, §5.5] can be interpreted as a no-go theorem -- see [Hen15, Rem. 5] for a discussion). The theory of bicommutant categories (which still needs to be developed) promises to achieve two things. First, it shows that, contrary to general expectations, Reshetikhin- Turaev theories do seem to extend down to points (at least the ones coming from conformal nets). Second, and more importantly, it puts Turaev-Viro theories and Reshetikhin-Turaev theories on an equal footing, by providing a unified language that applies to both of them. The expected relations are summarised in the following dia- (cid:40) (cid:41) , 2 Conformal nets gram: (cid:40) (cid:41) Unitary fusion category 1 (cid:40) (cid:40) (cid:41) (cid:41) Bicommutant categories 3 Extended 3-dim. TQFTs Turaev -- Viro construction2 Reshetikhin -- Turaev construction3 applied to Repf (A) The arrow labelled 1 was constructed in our earlier paper [HP17]. The arrow labelled 2 1Local normality is a technical condition which might be equivalent to the positivity of the energy [Hen17b, Conj. 22 & 34]. 2The Turaev-Viro construction requires the choice of a pivotal structure on the fusion category. A unitary fusion category admits a canonical pivotal structure [ENO05, Prop. 8.23]. 2 is the content of the present paper (see Corollary 1.8 below for a precise statement). The arrow labelled 3 is still conjectural and is only expected to exist when the bicommutant category satisfies certain finiteness conditions (ensuring that it is fully dualisable). 1.1 Motivations from Chern-Simons theory By the celebrated cobordism hypothesis [BD95, Lur09], a topological field theory is entirely determined by its value on a point. The present line of research was motivated by the quest for a mathematical object that one may reasonably declare to be the value of Chern -- Simons theory on a point. +(BG, Z) be the subset of elements k ∈ H 4(BG, Z) whose image under the Chern -- Weil homo- morphism Given a compact connected Lie group G, with classifying space BG, let H 4 H 4(BG, Z) → Sym2(g∗)G In our earlier paper [Hen16], given G and k ∈ H 4 (2) are positive definite metrics (cid:104)· ,·(cid:105)k on g. By [Hen16, Thm. 6], the map (2) is injective +(BG, Z) under that map is, up to a scalar, the set of invariant and the image of H 4 metrics on g such that (cid:107)X(cid:107)2 ∈ Z for all X in {X ∈ g : exp(X) = e}. +(BG, Z) as above, we constructed a vertex operator algebra VG,k and a chiral conformal net AG,k, called the chiral WZW vertex algebra and the chiral WZW conformal net, respectively.4 A bijective correspon- dence was established in [CKLW15] between a certain class of unitary vertex algebras and a certain class of chiral conformal nets. We conjecture that VG,k and AG,k map to each other under that correspondence, and that there is an equivalence of modular tensor categories Repf (VG,k) ∼= Repf (AG,k). Here, Repf denotes the category of representations which are finite direct sums of irreducible ones. Assuming the above conjectures, we define Repk f (LG), the modular tensor category of positive energy representations of the loop group LG at level k, to be the category Repf (VG,k), equivalently Repf (AG,k). Let CSG,k be the Chern -- Simons theory associated to the gauge group G and the level k [DW90, Wit89]. This is a 3-dimensional topological field theory with action functional given, up to a scalar, by:5 (cid:90) (cid:10)A ∧ dA(cid:11) S = (cid:10)A ∧ [A ∧ A](cid:11) k . k + 1 3 (3) In [Hen15], we argued that a necessary condition for a tensor category T to be the value of CSG,k on a point is for its Drinfel'd center Z(T ) to be braided equivalent to Repk f (LG), or possibly Repk(LG) (see Section 1.3 for a definition of the Drinfel'd center). We proposed the category Repk(ΩG) of locally normal representations of the Repf (A) is modular -- see Remark 1.2. 3For this construction to work, one needs to assume that the conformal net A has finite index, so that 4Earlier references on these models include [DLM96][Li01][FGK88][MS89, §2][FSS96, §6]. 5When G is not simply connected, one cannot use the formula (3) to define the action. See [CJM+05, DW90, FSS15] for ways to overcome this difficulty. 3 based loop group as a candidate for the value of Chern -- Simons theory on a point (see [FHLT10, Wra10] for previous work in that direction), and offered the relation (1) as evidence for our claim. For the remainder of this section, let us commit to the following definitions: Repk f (LG) := Repf (AG,k) Repk(LG) := Rep(AG,k). (4) Let us also define Repk(ΩG) to be the category of solitons of AG,k (see Definition 1.4, in the next section). We call it the category of locally normal representations of the based loop group6 It is widely believed that the chiral WZW conformal nets AG,k satisfy a certain finiteness condition called finite index, or complete rationality (see Section 1.4 for a definition). This property is known to hold for G = SU (n) [Was98, Xu00], and in a few other cases. Theorem 1.1. Let Repk(LG) be as in (4). If AG,k has finite index, then Z(Repk(ΩG)) ∼= Repk(LG). Proof. This is a special case of Theorem A, in Section 1.4. Remark 1.2. If a conformal net A has finite index, then Repf (A) is a modular tensor category, and Rep(A) = Hilb ⊗Vec Repf (A) [KLM01, Cor. 37][BDH17, Thm. 3.9].7 The latter implies that every object of Rep(A) is a (potentially infinite) direct sum of simple objects. Remark 1.3. When G is not simply connected, we presently do not know, in general, whether the vertex algebra VG,k is unitary. When G (cid:54)= SU (n), it is not known whether AG,k is completely rational or whether Repf (AG,k) is modular, except in some isolated cases. Even when G = SU (n), where it is known that Repf (AG,k) is modular, it is not known whether Repf (VG,k) ∼= Repf (AG,k) as (modular) tensor categories, except when n = 2 [Hen15, §3]. Establishing the above properties are important open problems. 1.2 Representations and solitons INT → VN from the category Conformal nets [BDH15, Def. 1.1] are functors A : of intervals (an interval is a manifold diffeomorphic to [0,1]) to the category of von Neumann algebras (see Definition 3.1 for the axioms that such a functor should satisfy). Let S1 := {z ∈ C : z = 1} be the standard circle. A representation of a conformal net consists of a Hilbert space H and a collection of compatible actions ρI : A(I) → B(H) of the algebras A(I), where I ranges over all subintervals of S1. We write Rep(A) for the category of representations of A whose underlying Hilbert space is separable. 6This category is equivalent to the version of Repk(ΩG) defined in [Hen15, §4] ([Hen17b, Thm. 31]). 7 The braiding on Rep(A) defined in [BDH17, Sec. 3B] has not been compared to the one in [KLM01]. We can therefore not exclude the possibility that, when µ(A) < ∞, the category Rep(A) has two distinct modular structures. The braided structure used in Theorem A is the one used in [KLM01]. 4 (Throughout this work, all Hilbert spaces are assumed to be separable. This will be important for the results in Section 4.3 to hold, see Remark 4.17.) The monoidal structure on Rep(A) is defined as follows. Let H and K be repre- sentations. Let I+ be the upper half of S1, and let I− be its lower half. Precomposing the left action of A(I+) on H by the map A( z (cid:55)→ ¯z : I− → I+) : A(I−)op → A(I+) (5) yields a right action of A(I−) on H. We let H (cid:2) K := H (cid:2)A(I−) K. Here, the symbol (cid:2) denotes Connes' relative tensor product (see Section 2.3 for a definition). The algebra A(I−) acts on K in the usual way, and it acts on H on the right as described above. The left actions of A(I−) on H and of A(I+) on K induce corresponding actions on H (cid:2) K. For every interval I ⊂ S1, the actions of 8 A(I ∩ I−) and A(I ∩ I+) on H (cid:2) K extend to an action ρI : A(I) → B(H (cid:2) K). Together, these equip H (cid:2) K with the structure of a representation. We refer the reader to Section 3.2 for more details. There is also a braiding on Rep(A), discussed in Section 3.3. A soliton of a conformal net is something akin to a representation [BE98, Kaw02, LR95, LX04] (the usage of the term 'soliton' in algebraic quantum field theory goes back to at least [Fro76]): Definition 1.4 ([LX04, §3.0.1]). A soliton of A is a Hilbert space (always assumed separable) equipped with compatible actions of the algebras A(I), where I ranges over all subintervals of the standard circle whose interior does not contain the base point 1 ∈ S1. We write TA for the category of solitons of A. Equivalently, a soliton is a Hilbert space equipped with compatible actions of all the algebras A(I) as I ranges over all subintervals I (cid:40) S1 cut is the manifold obtained from the standard circle by removing its base point and replacing it by two points: cut, where S1 S1 : S1 cut : Further down, we sometimes write T +A in place of TA, for reasons that will become clear later on. The monoidal structure on TA is defined in the same way as the one of Rep(A). Given two solitons H and K, we consider the right action A(I−)op → B(H) given as the composite of the map (5) with the left action A(I+) → B(H), and we let H (cid:2) K := H (cid:2)A(I−) K. (6) 8Here, we use the convention A(I1 (cid:116) I2) := A(I1) ¯⊗A(I2), where ¯⊗ denotes the spatial tensor product, to define the value of A on disjoint unions of intervals. 5 The left actions of A(I−) on H and of A(I+) on K induce corresponding actions on H (cid:2) K. Finally, for any interval I ⊂ S1, 1 (cid:54)∈ I, the actions of A(I ∩ I−) and A(I ∩ I+) extend to an action ρI : A(I) → B(H (cid:2) K). The details of his construction can be found in Section 4.1. Remark 1.5. We remind the reader that, by definition, when A = AG,k, the category of solitons agrees with the category Repk(ΩG) of locally normal representations of the based loop at level k. 1.3 Bicommutant categories Bicommutant categories are higher categorical analogs of von Neumann algebras. They are obtained by replacing the algebra B(H), in the definition of a von Neumann algebra, by the tensor category Bim(R) of all bimodules over a hyperfinite factor. Let R be a hyperfinite factor, and let Bim(R) be its category of bimodules, equipped with the monoidal structure given by Connes' relative tensor product (we insist that all Hilbert spaces be separable). The category Bim(R) admits an antilinear involution at the level of objects (the conjugate of a bimodule) and a second involution at the level of morphisms (the adjoint of a linear map). Together, these two involutions equip this category with the structure of a bi-involutive tensor category (Definition 2.3). A bicommutant category is a particular kind of bi-involutive tensor category. Given a bi-involutive functor ι : T → B between bi-involutive tensor categories, one may consider the commutant ZB(T ) of T inside B. The objects of ZB(T ) are pairs (X, e) with X ∈ B and e = (eY )Y ∈T a unitary half-braiding eY : X ⊗ ι(Y ) → ι(Y ) ⊗ X, natural in Y , and subject to the 'hexagon' axiom eY1⊗Y2 = (idι(Y1)⊗eY1)◦(eY1 ⊗idι(Y2)) (see Section 2.1 for more details). The category ZB(T ) is again bi-involutive, and is equipped with a bi-involutive functor (X, e) (cid:55)→ X to B: T → B ← ZB(T ). The Drinfel'd center is a special case of the above notion: Definition 1.6. The Drinfel'd center Z(T ) of a bi-involutive tensor category T is the commutant of T inside itself. The Drinfel'd center of a bi-involutive tensor category is braided and bi-involutive. When B = Bim(R), we write C(cid:48) := ZBim(R)(T ) for the commutant of T inside Bim(R). There is an obvious 'inclusion' functor T → T (cid:48)(cid:48) from any category to its bicommutant which sends an object Y ∈ T to the object (ι(Y ), e(cid:48)), with half-braiding e(cid:48) given by e(cid:48) for (X, e) ∈ C(cid:48). (X,e) := e−1 Y Definition 1.7. A bicommutant category is a bi-involutive tensor category T for which there exists a hyperfinite factor R and a bi-involutive functor T → Bim(R) such that the inclusion functor T → T (cid:48)(cid:48) is an equivalence of (bi-involutive tensor) categories. 6 The category of solitons of a conformal net is bi-involutive in the following way. cut, its conjugate H is the Given H ∈ TA, with actions ρI : A(I) → B(H) for I (cid:40) S1 complex conjugate Hilbert space equipped with the actions A(I) A(z(cid:55)→¯z) −−−−−→ A( ¯I)op ∗−−→ A( ¯I) ρ ¯I−−−→ B(H) = B(H). (7) Here, ¯I denotes the image of I ⊂ S1 under the complex conjugation map S1 → S1. The conjugation operation on TA squares to the identity, and satisfies H (cid:2) K ∼= K (cid:2) H. Given a conformal net A, set R := A(I−). Then there is an obvious fully faithful bi-involutive functor TA → Bim(R). (8) It sends a soliton H to the R-R-bimodule with left action given by the usual left action of A(I−) on H, and right action given by the left action of A(I+) precomposed by the map (5). One of our main results (Corollary 1.8) is that when A has finite index, the above functor exhibits TA as a bicommutant category. 1.4 Main results Recall that TA = T +A is the category whose objects are Hilbert spaces equipped with compatible actions of the algebras A(I), for I ⊂ S1, 1 (cid:54)∈ I. A denote the category whose objects are Hilbert spaces equipped with com- patible actions of A(I), for I ⊂ S1, −1 (cid:54)∈ I. Letting R := A(I−), the same formulas (6) and (7) endow T − A with the structure of a bi-involutive tensor category, and we have a bi-involutive functor Let T − T − A → Bim(R). Theorem A. Let A be a conformal net with finite index and let R := A(I−). Let TA = T +A be its category of solitons, with canonical inclusion T +A → Bim(R) as in (8). Then: tensor category. A )(cid:48) → Bim(R) is fully faithful and we have (T − • The canonical map (T +A )(cid:48) → Bim(R) is fully faithful and we have (T +A )(cid:48) = T − A . • The canonical map (T − A )(cid:48) = T +A . • The Drinfel'd center of T +A is equivalent to Rep(A) as a braided bi-involutive Corollary 1.8. If A is a conformal net with finite index, then TA is a bicommutant category. Remark 1.9. The main theorem in [Hen15, §5] is stated as an equivalence of balanced tensor categories (a balanced tensor category is a braided tensor categories with twists [JS91]). When X is a dualizable object, the twist θX : X → X is expressible in terms of the braiding and the dagger structure as θX := (evX ⊗ id)(id ⊗ βX,X )(ev∗ X ⊗ id), where evX : X⊗X → 1 and coevX : 1 → X⊗X are solutions to the normalized duality equations [HP17, §2.2]. This can then be extended to arbitrary objects by additivity (see Remark 1.2). Remark 1.10. If we do not assume that A has finite index, then we can still define the tensor functor T − A → (T +A )(cid:48) and the braided tensor functor Rep(A) → Z(TA), but we do not know whether they are equivalences. 7 Acknowledgments I am deeply indebted to Arthur Bartels and Christopher Douglas for the long-time collaboration that set the foundations on which the present work is resting, and I am grateful to Jacob Lurie for suggesting, back in 2008, that one could use our work with A. Bartels and C. Douglas to understand what Chern-Simons theory assigns to a point. This research was supported by the ERC grant No 674978 under the European Union's Horizon 2020 research innovation programme. 2 Bicommutant categories Bicommutant categories are higher categorical analogs of von Neumann algebras. They were introduced in [Hen15], and the first examples were constructed in [HP17]. Let R be a hyperfinite factor, and let R-Mod be the category of R-modules whose underlying Hilbert space is separable. We think of R-Mod as a higher categorical analog of an infinite dimensional Hilbert space. Our slogan is: von Neumann algebras act on Hilbert spaces; bicommutant categories act on categories like R-Mod. In this context, the higher categorical analog of B(H) is the tensor category End(R- Mod) of completely additive endofunctors of R-Mod (see [Lur11, Lecture 21] or [BDH16, §B.VIII] for a definition of completely additive functors). The latter is equivalent to the tensor category Bim(R) of all R-R-bimodules. such that the natural inclusion A → A(cid:48)(cid:48) into its bicommutant is an isomorphism: Recall that a von Neumann algebra is an algebra which admits a map to B(H) A → B(H) A = A(cid:48)(cid:48). Analogously, a bicommutant category is a tensor category T which admits a bi-involutive functor to Bim(R) such that the natural inclusion functor T → T (cid:48)(cid:48) of T into its bicom- mutant is an equivalence of categories: T → Bim(R) T ∼= T (cid:48)(cid:48). 2.1 The commutant of a tensor category Let T be a tensor category. The Drinfel'd center Z(T ) of T is the category whose objects are pairs (X, e), where X is an object of T and e = (eY : X ⊗ Y ∼=−→ Y ⊗ X)Y ∈T is a family of isomorphisms called a half-braiding. The half-braiding is required to be natural in Y , and to make the following diagram9 commute for every Y, Z ∈ T : eY ⊗idZ X ⊗ Y ⊗ Z Y ⊗ X ⊗ Z eY ⊗Z idY ⊗eZ Y ⊗ Z ⊗ X. (9) A morphism (X 1, e1) → (X 2, e2) in the Drinfel'd center is a morphism f : X 1 → X 2 in T such that (idY ⊗ f ) ◦ e1 Y ◦ (f ⊗ idY ) for every Y ∈ T . The tensor product Y = e2 9Here, we have suppressed associators for brevity. By adopting this simplified notation, we do not mean to imply that our tensor categories are strict. 8 of two objects of Y ⊗ idX 2) ◦ (idX 1 ⊗ e2 (e1 Y ). Finally, Z(T ) is equipped with a braiding Z(T ) is given by (X 1, e1) ⊗ (X 2, e2) := (X 1 ⊗ X 2, e12) with e12 Y := β : (X 1, e1) ⊗ (X 2, e2) ∼=−→ (X 2, e2) ⊗ (X 1, e1) given by e1 X 2. Basic references include [JS91, Maj91, Mug03]. The above definition can be relativized to the case when T is a subcategory of some bigger tensor category B (or, more generally, when T is equipped with a functor ι : T → B, not necessarily an inclusion). Definition 2.1. Let ι : T → B be a tensor functor between tensor categories. The ZB(T ) of T inside B is the category whose objects are pairs (X, e), where commutant X is an object of B and e = (eY : X ⊗ ιY → ιY ⊗ X)Y ∈T (10) is a collection of isomorphisms, called a half-braiding. The half-braiding is required to be natural in Y , and to satisfy the following analog of (9) for every Y, Z ∈ T : eY ⊗idιZ ιY ⊗ X ⊗ ιZ idιY ⊗eZ X ⊗ ιY ⊗ ιZ X ⊗ ι(Y ⊗ Z) ∼= eY ⊗Z ιY ⊗ ιZ ⊗ X ι(Y ⊗ Z) ⊗ X. ∼= (11) A morphism (X 1, e1) → (X 2, e2) in ZB(T ) is a morphism f : X 1 → X 2 in B satisfying (idιY ⊗ f ) ◦ e1 Y ◦ (f ⊗ idιY ) for every Y ∈ B. The tensor product in ZB(T ) is given by the same formula as for the Drinfel'd Y = e2 center: (X 1, e1) ⊗ (X 2, e2) := (X 1 ⊗ X 2, e12), e12 Y := (e1 Y ⊗ idX 2) ◦ (idX 1 ⊗ e2 Y ). Finally, there is a tensor functor ZB(T ) → B given by (X, e) (cid:55)→ X. In the presence of dagger structures, the definitions of Drinfel'd center and of commu- tant of a tensor category inside another tensor category can be modified by insisting that the half-braidings be unitary. We reserve the notations Z(T ) and of ZB(T ) for the unitary versions. 2.2 Bi-involutive tensor categories A dagger category is a linear category over C equipped with an antilinear map ∗ : Hom(X, Y ) → Hom(Y, X) which satisfies f∗∗ = f and (f ◦ g)∗ = g∗ ◦ f∗. An invertible morphism in a dagger category is called unitary if f∗ = f−1. Definition 2.2. A dagger tensor category is a dagger category equipped with a monoidal structure whose associator and unitor isomorphisms are unitary, and which satisfies (f ⊗ g)∗ = f∗ ⊗ g∗. 9 A dagger functor F between dagger tensor categories is a dagger tensor functor if it comes along with a unitary natural transformation µX,Y : F (X) ⊗ F (Y ) → F (X ⊗ Y ) and a unitary i : 1 → F (1) such that µX,Y ⊗Z◦(idF (X)⊗µY,Z) = µX⊗Y,Z◦(µX,Y ⊗idF (Z)) and µ1,X ◦ (i ⊗ idF (X)) = idF (X) = µX,1 ◦ (idF (X) ⊗ i). A dagger functor between dagger tensor categories is a dagger anti-tensor functor if it comes with a unitary natural transformation νX,Y : F (X) ⊗ F (Y ) → F (Y ⊗ X) and a unitary j : 1 → F (1) such that νX,Z⊗Y ◦ (idF (X) ⊗ νY,Z) = νY ⊗X,Z ◦ (νX,Y ⊗ idF (Z)) and ν1,X ◦ (j ⊗ idF (X)) = idF (X) = νX,1 ◦ (idF (X) ⊗ j). involution, denoted X (cid:55)→ X, which is a dagger anti-tensor functor: Bi-involutive tensor categories are dagger tensor categories equipped with a second Definition 2.3. A bi-involutive tensor category is a dagger tensor category T equipped with a covariant anti-linear dagger anti-tensor functor · : T → T called the conjugate. The structure data of this anti-tensor functor are denoted νX,Y : X ⊗ Y (cid:39)−→ Y ⊗ X and j : 1 → 1. This functor is involutive, meaning that for every X ∈ T , we are given unitary natural isomorphisms ϕX : X → X satisfying ϕX = ϕX . Finally, we require the compatibility conditions ϕ1 = j ◦ j and ϕX⊗Y = νY,X ◦ νX,Y ◦ (ϕX ⊗ ϕY ). Definition 2.4. A dagger tensor functor F between bi-involutive tensor categories is called a bi-involutive functor if it comes equipped with a unitary natural transformation γX : F (X) → F (X) satisfying γX = γX µX,Y ◦ νF (Y ),F (X) ◦ (γY ⊗ γX ) ◦ µ−1 Y ,X −1 ◦ ϕF (X) ◦ F (ϕX )−1, γ1 = i ◦ j ◦ i−1 ◦ F (j)−1, and γX⊗Y = ◦ F (νY,X )−1. bimodules over a von Neumann algebra R. The prototypical example of a bi-involutive category is the category Bim(R) of all Given a bi-involutive functor T → B between bi-involutive tensor categories, we let ZB(T ) be the full subcategory of the category described in Definition 2.1 where the half-braidings (10) are unitary. This category has the advantage of being, once again, a bi-involutive tensor category. The dagger structure is inherited from B, and the conjugate of an object (X, e) ∈ ZB(T ) is given by (X, e(cid:48)), with Y : X ⊗ Y e(cid:48) id⊗ϕY−−−−→ X ⊗ Y νX,Y−−−→ Y ⊗ X −1−−−→ X ⊗ Y eY ν−1 Y ,X−−−→ Y ⊗ X ϕ−1 Y ⊗id −−−−−→ Y ⊗ X. The Drinfel'd center Z(T ) of a bi-involutive tensor category T is the commutant of T inside itself. Remark 2.5. The categories ZB(T ) and ZB(T ) need not, in general, be equivalent. However, in the cases studied in this paper, they will turn out equivalent (see Re- mark 4.23). 10 2.3 Bim(R) Let R be a hyperfinite factor, and let Bim(R) be the category of all R-R-bimodules whose underlying Hilbert spaces is separable. It is a dagger category by means of the operation that sends a bimodule map to its adjoint. Let L2(R) be the non-commutative L2-space of R [Haa75, Yam92] (for any faithful state φ : R → C, there is a canonical identification between L2R and the GNS Hilbert space associated to φ [Tak03, Def IX.1.18]). By Tomita-Takesaki theory, this Hilbert space is equipped with two actions of R that are each other's commutants, and an antilinear involution J that satisfies J(xξy) = y∗J(ξ)x∗. The tensor structure (cid:2)R : Bim(R) × Bim(R) → Bim(R). (12) on Bim(R) is known as Connes fusion, or relative tensor product. The bimodule L2(R) is the unit object for that operation. Given two bimodules H and K, their fusion H(cid:2)RK is the completion of HomMod-R( L2R, H) ⊗R K with respect to the inner product (cid:104)ϕ ⊗ ξ, ψ ⊗ η(cid:105) := (cid:104)λ−1(ψ∗ ◦ ϕ)(ξ), η(cid:105), where λ : R → EndMod-R(L2R) denotes the left action of R on its L2-space. The left ac- tion of R on H and the right action of R on K equip H(cid:2)RK, once again, with the struc- ture of a bimodule. The Connes fusion can be equivalently described as a completion of H ⊗R HomR-Mod(L2R, K), or HomMod-R(L2R, H) ⊗R L2R ⊗R HomR-Mod(L2R, K). Basic references include [BDH14] [Con94, V.B.δ] [Sau83] [Tho11]. Remark 2.6. Using the last description of the fusion, and the fact that L2(R) ∼= L2(Rop), we see that there is a canonical isomorphism H (cid:2)R K ∼= K (cid:2)Rop H. Given H ∈ Bim(R), its complex conjugate H is a bimodule by means of the actions a ¯ξb := b∗ξa∗. We call it the conjugate bimodule. This operation comes with canonical isomorphisms ν : H (cid:2)R K → K (cid:2)R H and j : L2(R) → L2(R) reviewed in [HP17, §2.4]. All together, these operations endow Bim(R) with the structure of a bi-involutive tensor category. By definition, a bicommutant category is a bi-involutive tensor cat- egory T for which there exists a hyperfinite factor R and a bi-involutive functor T → Bim(R) such that the inclusion functor T → T (cid:48)(cid:48) = ZBim(R)(ZBim(R)(T )) is an equivalence of categories. 3 Conformal nets In this section, we recall the definition of conformal net from [BDH15], along with the notion of representation of a conformal net, the fusion product (cid:2) : Rep(A) × Rep(A) → Rep(A), and the braiding of representations βH,K : H (cid:2) K → K (cid:2) H. 11 3.1 Coordinate free conformal nets Let us define an interval to be an oriented manifold diffeomorphic to [0, 1]. We write Diff +(I) for the group of orientation preserving diffeomorphisms of an interval I. Let INT be the category whose objects are intervals and whose morphisms are embeddings, not necessarily orientation preserving, and let VN be the category whose objects are von Neumann algebras and whose morphisms are normal maps which are either ∗-algebra homomorphisms or ∗-algebra anti-homomorphisms. Definition 3.1 ([BDH15, Def. 1.1]). A conformal net is a covariant functor A : INT → VN from the category of oriented intervals and embeddings to the category of von Neumann algebras. It sends orientation-preserving embeddings to injective homomor- phisms and orientation-reversing embeddings to injective antihomomorphisms. More- over, for any intervals I and J, the natural map HomINT(I, J) → HomVN(A(I),A(J)) should be continuous for the C∞ topology on HomINT(I, J) and Haagerup's u-topology10 on HomVN(A(I),A(J)). In addition to that, a conformal net should satisfy the follow- ing five axioms: i. Locality: If I, J ⊂ K have disjoint interiors, then A(I) and A(J) are commuting subalgebras of A(K). ii. Strong additivity: If K = I∪J, then A(K) is generated as a von Neumann algebra by its two subalgebras: A(K) = A(I) ∨ A(J). iii. Split property: If I, J ⊂ K are disjoint, then the natural map from the algebraic tensor product A(I)⊗alg A(J) → A(K) extends to a map from the spatial tensor product A(I) ¯⊗A(J) → A(K). iv. Inner covariance: If ϕ ∈ Diff +(I) restricts to the identity in a neighbourhood of the boundary of I, then A(ϕ) : A(I) → A(I) is an inner automorphism. v. Vacuum sector: Let J (cid:40) I contain the boundary point p ∈ ∂I, and let ¯J denote J with reversed orientation; A(J) acts on L2(A(I)) via the left action of A(I), and A( ¯J) ∼= A(J)op acts on L2(A(I)) via the right action of A(I). In that case, we require that the action of A(J)⊗alg A( ¯J) on L2(A(I)) extends to an action of A(J ∪p ¯J)11: A(J) ⊗alg A( ¯J) B(L2A(I)) A(J ∪p ¯J) (13) A conformal net A is called irreducible if the algebras A(I) are factors. We will always assume that our conformal nets are irreducible. Remark 3.2. For any interval I, the identity map ¯I → I (which is orientation reversing) induces an isomorphism A( ¯I) ∼= A(I)op. This was used above, in the formulation of the vacuum sector axiom. 10Topology of pointwise convergence on the preduals. 11Here, J ∪p ¯J is equipped with any smooth structure extending the given smooth structures on J and ¯J, and for which the orientation-reversing involution that exchanges J and ¯J is smooth. 12 Let S1 := {z ∈ C : z = 1}. A representation of A consists of a Hilbert space H (always assumed separable) and a collection of homomorphisms ρI : A(I) → B(H) for every interval I ⊂ S1, subject to the compatibility condition ρIA(J) = ρJ whenever J ⊂ I. The vacuum representation, or vacuum sector, is a representation of A on the Hilbert space H0 = HA 0 := L2(A( )). ) acts via the isomorphism A(b) : A( The algebra A( ) acts via the usual left multiplication on its L2-space. The algebra A( )op, where b(z) = ¯z, followed by right multiplication of A( ) on its L2-space. The vacuum sector axiom then ensures that those two actions uniquely extend to actions of A(I) for every I ⊂ S1 [BDH15, §1.b]. Given an interval I ⊂ S1, let I(cid:48) := S1\I denote the complement of the interior of I. The representation H0 satisfies the important property of Haag duality [BDH15, Prop 1.18]: ) → A( ρI(cid:48)(A(I(cid:48))) = ρI (A(I))(cid:48). We note that Definition 3.1 is rigged in such a way so as to have Haag duality essen- tially built into it. Using the classical definition of conformal nets, Haag duality is an important theorem [GF93, §II.2][Lon08, Thm 6.2.3][BSM90]. Recall that P SU (1, 1) is the group of Mobius transformations, acting on S1 by ¯b ¯a ) : z (cid:55)→ az+b for a2 − b2 = 1. Then the vacuum sector admits a continuous ( a b representation ¯bz+¯a P SU (1, 1) → U (H0) ϕ (cid:55)→ uϕ (14) ϕ for every I ⊂ S1 and a ∈ A(I) which satisfies the covariance property A(ϕ)(a) = uϕau∗ 0 e−it/2) ∈ P SU (1, 1) denote rotation by t, and let [BDH15, Thm 2.13]. Let rt = ( eit/2 0 Rt = urt be its image under the above homomorphism. The unbounded self-adjoint operator L0 := −i d rotations in the sense that Rt = etiL0. We call a conformal net chiral if the energy operator L0 has positive spectrum and the P SU (1, 1)-invariant subspace H0 is one dimensional (equivalently, the subspace invariant under all the Rt's is one dimensional). (cid:12)(cid:12)t=0 Rt is called the energy operator ; it generates the subgroup of dt Remark 3.3. Our definition of conformal net (Definition 3.1) is different from the one usually encountered in the literature. If a conformal net in the sense of [GF93, Lon08] satisfies the additional assumptions of strong additivity and diffeomorphism covari- ance12, then it induces a conformal net in the sense of Definition 3.1 [BDH15, Prop 4.9]. Conversely, a conformal net (in the sense of Definition 3.1) which is chiral induces a conformal net in the classical sense by restricting it to the circle. This establishes a bijective correspondence between chiral conformal nets in the sense described above, and conformal nets in the sense of [GF93, Lon08] subject to the additional assumptions of strong additivity and diffeomorphism covariance. 12The split property was recently shown to be a consequence of diffeomorphism covariance [MTW16]. 13 3.2 Fusion of representations The standard monoidal structure (cid:2) : Rep(A) × Rep(A) → Rep(A) (15) on the category of representations of a conformal net is called fusion. There are two main approaches for defining that operation [GF93, Sec. IV.2] and [Was98, Sec. 30] (see [Con94, Prop V.B.δ.17] for the equivalence between the two). We will follow the latter. let I0 be a fixed 'standard interval' (in Section 1.2, this was taken to be the lower half of S1, but I0 = [0, 1] is also a pleasant choice), and let R := A(I0) be the value of our conformal net on that standard interval. A representation H ∈ Rep(A) has commuting left actions of the algebras A( ) ∼= R and A( ) ∼= Rop, constant-speed orientation preserving parametrization by I0 constant-speed orientation reversing parametrization by I0 (16) so we get commuting left and right actions of R, making H into an R-R-bimodule. The above construction gives a functor Rep(A) → Bim(R). We will see later, in Section 4.1, that this functor is fully faithful (Lemma 4.1), and that its image is closed under the operation (12) of Connes fusion (Lemma 4.4). This will allow us to define the fusion product on Rep(A) as the restriction of the corresponding operation on Bim(R). Remark 3.4. In our situation of interest, there is an alternative way of defining Connes fusion that avoids L2-spaces and Tomita-Takesaki theory, and avoids the parametriza- tions (16). It is defined directly as an operation on the category of A( )op- bimodules, equivalently, A( )-bimodules. All that we use is the fact that there is a distinguished bimodule H0 with the property that the actions of A := A( ) and of B := A( ) are each other's commutants (Haag duality). The fusion of H and K is the completion of )-A( )-A( HomB(H0, H) ⊗A K under the inner product (cid:104)ϕ ⊗ ξ, ψ ⊗ η(cid:105) := (cid:104)(ψ∗ϕ)ξ, η(cid:105), where ψ∗ϕ ∈ HomB(H0, H0) is identified with an element of A. Equivalently, the fusion is the completion of H ⊗B HomA(H0, K), or the completion of HomB(H0, H) ⊗A H0 ⊗B HomA(H0, K). There is also a 'coordinate free' version of the operation of fusion, that goes as follows. Recall that a representation of a conformal net is a Hilbert space equipped with compatible actions of the algebras A(I) for all the subintervals of the standard circle. More generally, for any circle S (a circle is an oriented 1-manifold diffeomorphic to S1) there is a notion of S-sector of A that generalises that of a representation [BDH15, Def 1.7]: Definition 3.5. Let A be a conformal net. An S-sector of A is a Hilbert space H and a collection of homomorphisms ρI : A(I) → B(H), I (cid:40) S 14 subject to the compatibility condition ρIA(J) = ρJ whenever J ⊂ I. We write SectS(A) for the category of S-sectors of A. The category SectS(A) contains a distinguished object H0(S,A), well defined up to non-canonical isomorphism, called the vacuum sector.13 By definition [BDH15, Def 1.17], for every interval I ⊂ S and every orientation reversing involution j : S → S that fixes ∂I, the vacuum sector H0(S,A) is isomorphic to L2(A(I)) via an isomorphism (17) well defined up to phase, that intertwines the two left actions of A(I) and satisfies v(A(j)(x)ξ) = (v(ξ))x for all x ∈ A(I) and ξ ∈ L2(A(I)). v : H0(S,A) → L2(A(I)), Let Θ be any theta-graph, and let S1, S2, S3 be its three circle subgraphs, oriented as follows: Θ : S1 : S2 : S3 : ) and local coordinates (cid:8)fi : [0, ε[ → Y(cid:9) The circles S1, S2, S3 are equipped with smooth structures which are compatible in the following sense: around each of the trivalent vertices of Θ, it is possible to pick a neighbourhood Y ⊂ Θ, (Y ∼= i=1,2,3 of the three legs of Y so that for each pair i, j ∈ {1, 2, 3} of distinct indices, the map (−fi) ∪ fj : ]−ε, ε[ → Y is smooth when viewed as a map to the relevant circle. Let I := S1 ∩ S2 be the central interval, equipped with the orientation inherited from I2. Then Connes fusion along A(I) defines a functor [BDH15, Def 1.31]: (cid:2)A(I) : SectS1(A) × SectS2(A) → SectS3(A). We denote the fusion of representations graphically by: H (cid:2) K = H (cid:2)A( ) K = K H (18) 3.3 The braiding In the literature on algebraic quantum field theory, the braiding on Rep(A) is usually defined as follows [FRS89, §2][Lon89, §7][GF93, Def 4.16]. First of all, the objects of Rep(A) are represented by localised endomorphisms of the ∗-algebra A := colim I⊂R A(I). Here, a ∗-algebra endomorphism ρ : A → A is said to be localised in a bounded region O ⊂ R if it acts as the identity on the subalgebras A(I) ⊂ A for every I disjoint from O. 13Since H0(S,A) is only well defined up to non-canonical isomorphism, it would be more correct to say a vacuum sector as opposed to the vacuum sector. 15 (We refer the reader to [GF93, §IV] for an explanation of the bijective correspondence between non-zero representations of A and localised endomorphisms of A.) Given localised endomorphisms ρ1 and ρ2, one picks unitaries U1, U2 ∈ A such that ρ1 := Ad(U1)ρ1 is localised in a region which is to the right of the region where ρ2 := Ad(U2)ρ2 is localised. The element ε := ρ2(U1)−1U−1 2 U1ρ1(U2) is then an intertwiner from ρ1ρ2 to ρ2ρ1 (it satisfies ερ1ρ2(x) = ρ2ρ1(x)ε), which is called the braiding of ρ1 and ρ2. It is independent of the choice of unitaries U1 and U2, provided Ad(U1)ρ1 is localised to the right of the localisation region of Ad(U2)ρ2. Here, the intertwining property is best explained by noting that ε is a composite of intertwiners ρ1(U2)−−−−→ ρ1 ρ2 U1−→ ρ1 ρ2 = ρ2 ρ1 ρ1ρ2 U−1 2−−−→ ρ2 ρ1 ρ2(U1)−1 −−−−−−→ ρ2ρ1. We now adapt the above definition14 of the braiding to the case of the fusion product (18). Given a representation H and an element x ∈ A(I) for some interval I ⊂ S1, we write ρH (x) for the action of x on H. Let In := {e2πiθ : θ ∈ [ n−1 4 ]} for n = 1, 2, 3, 4: 4 , n I2 I3 I1 I4 (19) Let us adopt the notations I123 := I1 ∪ I2 ∪ I3, I234 := I2 ∪ I3 ∪ I4, I34 = I3 ∪ I4, etc. By definition, the fusion of representations is given by H (cid:2) K = H (cid:2)A(I34) K. Let ϕ1 : I341 → I341 be a diffeomorphism that sends 1 to −i and is the identity near the boundary of that interval. Given two representations H, K ∈ Rep(A), we define H (cid:2)ϕ1 K := H (cid:2)A(I34) (ϕ1K), (20) where ϕ1K is the Hilbert space K with action of A(I34) twisted by A(ϕ1) : A(I34) → A(I3). We equip H (cid:2)ϕ1 K with the following actions of A(I412) and of A(I3). The algebra A(I412) acts on H (cid:2)ϕ1 K by means of its usual action on K. The algebra A(I3) acts on H (cid:2)ϕ1 K by first applying A(ϕ1)−1 : A(I3) → A(I34) and then using the action of A(I34) on H. We will see later that those actions extend, by strong additivity, to the structure of a representation on H (cid:2)ϕ1 K. Pick a unitary u1 ∈ A(I341) such that Ad(u1) = A(ϕ1) (Definition 3.1.iv). Lemma 3.6. The isomorphism U1 = U (H,K) 1 :=(cid:0)idH (cid:2) ρK(u1)(cid:1) ◦ ρH(cid:2)K(u1)−1 : H (cid:2) K → H (cid:2)ϕ1 K (21) intertwines the actions of A(I3) and of A(I412). 14Note that we also have ε = U−1 2 ρ2(U1)−1 ρ1(U2)U1. It is that second formula which most closely resembles our working definition (24) of the braiding. 16 Proof. We write ϕ, u, and U in place of ϕ1, u1, and U1. For x ∈ A(I3), we have: U ◦ ρH(cid:2)K(x) =(cid:0)idH (cid:2) ρK(u)(cid:1) ◦ ρH(cid:2)K(u−1x) =(cid:0)idH (cid:2) ρK(u)(cid:1) ◦ ρH(cid:2)K(cid:0)A(ϕ)−1(x)u−1(cid:1) =(cid:0)idH (cid:2) ρK(u)(cid:1) ◦ ρH(cid:2)K(cid:0)A(ϕ)−1(x)(cid:1) ◦ ρH(cid:2)K(u−1) =(cid:0)idH (cid:2) ρK(u)(cid:1) ◦(cid:0)ρH (A(ϕ)−1(x)) (cid:2) idK =(cid:0)ρH (A(ϕ)−1(x)) (cid:2) idK U ◦ ρH(cid:2)K(x) =(cid:0)idH (cid:2) ρK(u)(cid:1) ◦ ρH(cid:2)K(u−1x) (cid:1) ◦ ρH(cid:2)K(u−1) (cid:1) ◦(cid:0)idH (cid:2) ρK(u)(cid:1) ◦ ρH(cid:2)K(u−1) = ρH(cid:2)ϕK(x) ◦ U. =(cid:0)idH (cid:2) ρK(u)(cid:1) ◦ ρH(cid:2)K(cid:0)A(ϕ)−1(x)u−1(cid:1) =(cid:0)idH (cid:2) ρK(u)(cid:1) ◦ ρH(cid:2)K(cid:0)A(ϕ)−1(x)(cid:1) ◦ ρH(cid:2)K(u−1) =(cid:0)idH (cid:2) ρK(u)(cid:1) ◦(cid:0)idH (cid:2) ρK(A(ϕ)−1(x))(cid:1) ◦ ρH(cid:2)K(u−1) =(cid:0)idH (cid:2) ρK(uA(ϕ)−1(x))(cid:1) ◦ ρH(cid:2)K(u−1) =(cid:0)idH (cid:2) ρK(xu)(cid:1) ◦ ρH(cid:2)K(u−1) For x ∈ A(I412), we have: (22) = ρH(cid:2)ϕK(x) ◦ U. Corollary 3.7. The actions of A(I3) and A(I412) on H (cid:2)ϕ1 K endow it with the structure of an object of Rep(A). The map (21) is an isomorphism of representations. Let us now consider a diffeomorhpism ϕ2 : I234 → I234 that sends −1 to −i and is the identity near the boundary, and let u2 ∈ A(I341) be such that Ad(u2) = A(ϕ2). We can then define H (cid:2)ϕ2 K analogously to (20), and we have an isomorphism of representations :=(cid:0)idH (cid:2) ρK(u2)(cid:1) ◦ ρH(cid:2)K(u2)−1 : H (cid:2) K → H (cid:2)ϕ2 K. U2 = U (H,K) 2 (23) We are now ready to translate the classical definition of the braiding into the language of Connes fusion: Definition 3.8. Given two representations H and K, the braiding isomorphism βH,K : H (cid:2) K → K (cid:2) H is the composite βH,K : H (cid:2) K ∼= H (cid:2) K (cid:2) H0 U1 (cid:2) id−−−−→ H (cid:2)ϕ1 K (cid:2) H0 id (cid:2) U2−−−−→ H (cid:2)ϕ1 (K (cid:2)ϕ2 H0) ∼= K (cid:2)ϕ2 (H (cid:2)ϕ1 H0) id (cid:2) U∗ 1−−−−−→ K (cid:2)ϕ2 (H (cid:2) H0) (cid:2) id−−−−−→ K (cid:2) (H (cid:2) H0) ∼= K (cid:2) H, U∗ 2 (24) where H0 denotes the vacuum sector of the conformal net. The middle isomorphism is explained below. 17 Pictorially, we like to represent the isomorphisms (24) as the following sequence of moves: βH,K : K H ∼= H0 → K H H H0 → H0 K H K → H0 → H K H0 H K ∼= H K The isomorphism H (cid:2)ϕ1 (K (cid:2)ϕ2 H0) ∼= K (cid:2)ϕ2 (H (cid:2)ϕ1 H0) which occurs in (24) is symmetry isomorphism s : X (cid:2)R Y ∼= Y (cid:2)Rop X (Remark 2.6). The isomorphism in requires some explanation. Recall that for a right module X and a left module Y there the middle of (24) is the composite H (cid:2)R ϕ1(K (cid:2)R ϕ2H0) s−→ ϕ1(K (cid:2)R ϕ2H0) (cid:2)Rop H ϕ2(ϕ1H0 (cid:2)Rop H) id(cid:2)s−−−→ K (cid:2)R where the arrow labelled a is the associator of Connes fusion. a−→ K (cid:2)R ϕ2(H (cid:2)R ϕ1H0), (25) Remark 3.9. At this point, it is not clear whether the braiding (24) depends on the choice of diffeomorphisms ϕ1 and ϕ2, or whether it depends on our convention to add the vacuum sector on the top as opposed to the bottom. In Section 4.4, we will show that it is independent of all these choices, by using the fact that it extends to the case when H and K are solitons (H ∈ T − A , K ∈ T +A , see Corollary 4.22). Lemma 3.10. Let H, K, and L be representations. Then the following diagram is commutative: (H (cid:2)ϕ1 K) (cid:2) L a H (cid:2)ϕ1 (K (cid:2) L). (26) (H (cid:2) K) (cid:2) L a H (cid:2) (K (cid:2) L) U (H,K) 1 (cid:2) idL U (H,K(cid:2)L) 1 A corresponding property holds for U2. Proof. The maps involved only depend on L as an A(I341)-module. Let M := {L ∈ A(I341)-Mod : (26) holds}. That category contains the vacuum sector H0 as, in that : H (cid:2) case, the horizontal arrows in (26) can be both identified with the map U (H,K) K → H (cid:2)ϕ1 K. The category M is closed under taking direct sums and taking direct summands. H0 generates A(I341)-Mod under those operations. Therefore M = A(I341)-Mod. Lemma 3.11. Let H, K, and L be representations. Then the following diagram is commutative: 1 H (cid:2) K (cid:2) L 1 U (H(cid:2)K,L) (H (cid:2) K) (cid:2)ϕ1 L A corresponding property holds for U2. H (cid:2)ϕ1 (K (cid:2) L) idH (cid:2) U (K,L) H (cid:2)ϕ1 (K (cid:2)ϕ1 L). 1 U (H,K(cid:2)L) 1 a 18 (cid:1) ◦ U (H,K(cid:2)L) (cid:0)idH (cid:2) U (K,L) = idH (cid:2)(cid:16)(cid:0)idK (cid:2) ρL(u1)(cid:1) ◦ ρK(cid:2)L(u1)−1(cid:17) ◦(cid:16)(cid:0)idH (cid:2) ρK(cid:2)L(u1)(cid:1) ◦ ρH(cid:2)K(cid:2)L(u1)−1(cid:17) =(cid:0)idH(cid:2)K (cid:2) ρL(u1)(cid:1) ◦ ρH(cid:2)K(cid:2)L(u1)−1 1 1 (27) Proof. By definition, = U (H(cid:2)K,L) . 1 Proposition 3.12. The isomorphism (24) satisfies the 'hexagon' axioms βH,K(cid:2)L = (idK (cid:2) βH,L)(βH,K (cid:2) idL) and βH(cid:2)K,L = (βH,L (cid:2) idK)(idH (cid:2) βK,L) (we omit the associators for brevity). Proof. We only prove the first axiom. To keep notations short, we drop the symbol (cid:2), we write H1K for H (cid:2)ϕ1 K, we write H2K for H (cid:2)ϕ2 K, and we omit the vacuum sector H0. The definition (24) of the braiding then becomes: βH,K : HK → H1K → H1K2 → K2H1 → K2H → KH, where the isomorphism H1K2 → K2H1 is the map constructed in (25). Consider the following diagram, where all expressions are associated to the right unless otherwise indicated (for example, HKL stands for H (cid:2) (K (cid:2) (L (cid:2) H0)), H1K2L2 stands for H (cid:2)ϕ1 (K (cid:2)ϕ2 (L (cid:2)ϕ2 H0)), and H1(KL)2 stands for H (cid:2)ϕ1 ((K (cid:2) L) (cid:2)ϕ2 H0)): KHL K2HL KH1L K2H1L a H1K2L K2H1L2 a K2L2H1 a KH1L2 a KL2H1 KL2H HKL H1KL ((cid:63)) H1K2L2 a a K2L2H a K2LH ((cid:63)) KLH H1(KL)2 a (KL)2H1 (KL)2H The arrows labelled a involve associators. One reads βH,K (cid:2) idL along the top left (this is a consequences of Lemma 3.10 given our convention that all expressions are associated to the right), one reads idK (cid:2) βH,L along the top right, and one reads βH,K(cid:2)L along the bottom. In order to show that the desired equation βH,K(cid:2)L = (idK (cid:2) βH,L)(βH,K (cid:2) idL) holds, it is therefore enough to argue that each individual cell in the above diagram is commutative. The commutativity of the cells marked by a little star is the content of Lemma 3.11. The other cells are easily seen to be commutative. 19 4 Solitons Let A be a conformal net. Recall that a soliton is a Hilbert space (always assumed separable) equipped with compatible actions of the algebras A(I), for all subintervals of S1 whose interior does not contain the point 1. We write TA = T +A for the category of solitons. The category T − A is defined similarly, using the point −1 instead of the point 1. We also call the elements of T − A solitons, when this creates no confusion. Our first goal is to identify the categories T +A and T − A with subcategories of Bim(R). 4.1 Solitons as bimodules Let I1, . . . , I4 be the subintervals of S1 depicted in (19), and let us adopt the same notations as in the previous section: I12 = I1 ∪ I2, I23 = I2 ∪ I3, etc. Let I0 be the standard interval which we use to parametrize the lower and upper halves of the standard circle, as in (16). Let A be a conformal net, and let R := A(I0). The parametrizations I0 → I34 and ¯I0 → I12 induce an equivalence of categories between the category Bim(R) and the category whose objects are separable Hilbert spaces equipped with commuting left actions of the algebras A(I12) and A(I34). This allows us to identify T +A , T − A , and Rep(A) with subcategories of Bim(R): ι+ : T +A → Bim(R) ι− : T − A → Bim(R) ι : Rep(A) → Bim(R). (28) Lemma 4.1. The functors ι+, ι−, and ι are fully faithful. Proof. The functors ι+, ι−, ι are clearly faithful. We prove that they are full. Let H, K ∈ T +(A) (respectively H, K ∈ T −(A), respectively H, K ∈ Rep(A)), and let f : H → K be a morphism in Bim(R). By definition, f commutes with the actions of A(I12) and of A(I34). Let I ⊂ S1 be an interval which does not contain 1 in its interior (respectively an interval such that −1 (cid:54)∈ I, respectively any subinterval of S1). By assumption, f commutes with A(I ∩ I12) and A(I ∩ I34). By strong additivity (Definition 3.1.ii ), these two algebras generate a dense subalgebra of A(I). So f commutes with A(I). This being true for any I, f is a morphism in T +(A) (respectively a morphism in T −(A), respectively a morphism in Rep(A)). For a Hilbert space equipped with commuting left actions of A(I12) and of A(I34), consider the following properties: (a) The actions of A(I2) and A(I3) extend to an action of A(I23). (b) The actions of A(I4) and A(I1) extend to an action of A(I41). By using the parametrizations (16), we may treat (a) and (b) as conditions on R-R- bimodules. 20 Lemma 4.2. The essential images of the functors ι+, ι−, and ι are given by: Im(ι+) = {H ∈ Bim(R) condition (a) holds}, Im(ι−) = {H ∈ Bim(R) condition (b) holds}, Im(ι) = {H ∈ Bim(R) both (a) and (b) hold}. In order to establish this lemma, we will need the following technical result: Lemma 4.3. Let M be a connected 1-manifold (either a circle or an interval), and let {Ii ⊂ M}i∈I be a collection of intervals that satisfy Ii = M and Ii = M . (cid:91) i∈I (cid:91) i∈I Let H be a Hilbert space equipped with actions ρi : A(Ii) → B(H) for i ∈ I which are compatible in the sense that: 1. ρiA(Ii∩Ij ) = ρjA(Ii∩Ij ) : A(Ii ∩ Ij) → B(H). 2. For every j, k ∈ I and every intervals J ⊂ Ij, K ⊂ Ik with disjoint interiors, the algebras ρj(A(J)) and ρk(A(K)) commute. Then for every interval I (cid:40) M , the actions extend uniquely to an action of A(I) on H. ρiA(I∩Ii) : A(I ∩ Ii) → B(H) Let S1 cut be the manifold described in Section 1.2. Proof. The case when M is a circle was proved in [BDH15, Lem. 1.9]. The case when M is an interval was proved in [Hen17a, Lem. 4]. (And the two proofs are essentially the same.) Proof of Lemma 4.2. Clearly, every H ∈ T +(A) satisfies (a), every H ∈ T −(A) satis- fies (b), and every H ∈ Rep(A) satisfies both. If a bimodule H ∈ Bim(R) satisfies condition (a), then we may apply Lemma 4.3 with M = S1 cut. The Hilbert space H admits actions of the algebras A(I) for all I (cid:40) S1 cut, and is therefore a soliton. The argument for T −(A) is identical. If a bimodule H ∈ Bim(R) satisfies both (a) and (b), then we can apply Lemma 4.3 with M = S1. The Hilbert space H admits actions of the algebras A(I) for all I (cid:40) S1, and is therefore a representation of A. Lemma 4.4. The subcategories T +A , T − A , and Rep(A) of Bim(R) are closed under Connes fusion. Proof. In view of Lemma 4.2, it is enough to show that the properties (a) and (b) are preserved under fusion. By symmetry, it is enough to treat just one of them. Let H and K be bimodules that satisfy property (a). Then, by [BDH15, Cor. 1.29], the actions of A(I2) on K and A(I3) on H extend to an action of A(I23) on H (cid:2) K. That is, H (cid:2) K satisfies (a). (The intervals which were denoted by I, Il, Ir, and Il (cid:126)I Ir in [BDH15, Cor. 1.29] correspond to I0, I123, I234, and I23, respectively.) 21 4.2 The braiding between T − Given two solitons H ∈ T − A and K ∈ T +A , we can use the inclusions (28) to define their fusion H (cid:2) K ∈ Bim(R). The goal of this section is to extend the braiding on Rep(A) to a braiding A and T +A which is defined for all H ∈ T − Let ϕ1 : I341 → I341 and βH,K : H (cid:2) K → K (cid:2) H A and K ∈ T +A . H (cid:2)ϕ1 K := H (cid:2)R ϕ1K be as in Section 3.3. Here, as before, ϕ1K denotes the Hilbert space K with action of A(I34) twisted by A(ϕ1) : A(I34) → A(I3). We equip H (cid:2)ϕ1 K with the following actions of the algebras A(I12), A(I3), and A(I4). The algebras A(I12) and A(I4) act on H (cid:2)ϕ1 K by their usual action on K. The algebra A(I3) acts by first applying A(ϕ1)−1 : A(I3) → A(I34) and then using the usual action of A(I34) on H. We find it useful to represent the Hilbert spaces H (cid:2) K and H (cid:2)ϕ1 K by the following pictures: H (cid:2) K = ∗ ∗ K H H (cid:2)ϕ1 K = ∗ K ∗ H Here, the little star is a reminder that H and K are solitons, as opposed to represen- tations. We will see later, in Corollary 4.10, that the actions of A(I3) and A(I4) on H (cid:2)ϕ1 K extend, by strong additivity, to an action of A(I34). Let u1 ∈ A(I341) be such that Ad(u1) = ϕ1. The unitary =(cid:0)idH (cid:2) ρK(u1)(cid:1) ◦ ρH(cid:2)K(u1)−1 : H (cid:2) K → H (cid:2)ϕ1 K (29) U (H,K) 1 that was used in the definition (24) of the braiding no longer makes sense when H and K are solitons, because the actions of A(I34) and A(I1) on H (cid:2) K might not extend to an action of A(I341). We circumvent this difficulty by a trick that is based on the following lemma: Lemma 4.5 ([BDH16, Lem. B.24]). Let R be a factor, let A be any von Neumann algebra, and let F, G : R-Mod → A-Mod be completely additive functors. Let M ⊂ R-Mod be a full subcategory with only one object, which is not the zero object. Then a natural transformation τ : F → G is entirely determined by its restriction to M . Conversely, any natural transformation FM → GM extends to a natural transformation F → G. Proof. By complete additivity, τM determines τ on the subcategory of R-modules which are direct sums of the object of M . Every R-module is a direct summand of one of the above form, so τ is determined on all of R-Mod. (The proof is even simpler when R is a type III factor as, in that case, M is equivalent to the subcategory of all non-zero modules.) 22 Fix H ∈ T − A and consider the functors H (cid:2) − and H (cid:2)ϕ1 − from Rep(A) to the category C whose objects are Hilbert spaces equipped with commuting actions of the : H (cid:2) K → H (cid:2)ϕ1 K makes algebras A(I3) and A(I4). The definition (29) of U (H,K) sense in that context, so we get a natural transformation 1 U (H,−) 1 : Rep(A) C from K (cid:55)→ H (cid:2) K to K (cid:55)→ H (cid:2)ϕ1 K. We now observe that H (cid:2)− and H (cid:2)ϕ1 − also make sense as functors from A(I34)-Mod to C. Let M ⊂ A(I34)-Mod be the full subcategory consisting of only the vacuum Hilbert space. : H (cid:2) H0 → H (cid:2)ϕ1 H0 defined in (29) is a natural Lemma 4.6. The map U (H,H0) transformation M C. Proof. By Haag duality, EndM (H0) = EndA(I34)-Mod(H0) = A(I12). For every endo- morphism x ∈ A(I12) of H0, we need to show that the diagram 1 H (cid:2) H0 id (cid:2) x H (cid:2) H0 U (H,H0) 1 U (H,H0) 1 H (cid:2)ϕ1 H0 id (cid:2)ϕ1 x H (cid:2)ϕ1 H0 commutes. That computation was performed in (22). The category C is of the form A-Mod for some von Neumann algebra ([Gui66, §8]). By Lemma 4.5, we therefore get: Corollary 4.7. There exists a unique natural transformation U (H,−) 1 : A(I34)-Mod C (30) whose value on the vacuum sector H0 ∈ A(I34)-Mod is given by the map (29). We now have two definitions of U (H,−) 1 that we need to reconcile: A and K ∈ A(I341)-Mod. Then the map U (H,K) Lemma 4.8. Let H ∈ T − H (cid:2)ϕ1 K defined by (29) agrees with the one given by Corollary 4.7. Proof. Let M ⊂ A(I341)-Mod be the subcategory on which the two definitions of U (H,K) agree. By definition, M contains the vacuum sector. Since M is closed under direct sums and direct summands and the vacuum sector generates A(I341)-Mod under those operations, M = A(I341)-Mod. : H (cid:2) K → 1 1 We now restrict the natural transformation (30) along the functor T +A → A(I34)-Mod, to get a natural transformation U (H,−) 1 C from H (cid:2) − to H (cid:2)ϕ1 −. : T +A 23 Lemma 4.9. Let H ∈ T − A and K ∈ T +A be solitons. Then the map U (H,K) 1 : H (cid:2) K → H (cid:2)ϕ1 K (31) 1 1 defined above intertwines the actions of A(I12), A(I3), and A(I4). Proof. The map (31) intertwines the actions of A(I3) and A(I4) because it is a mor- phism in C. Recall from (30) that U (H,−) is natural with respect to all morphisms of : H (cid:2) K → H (cid:2)ϕ1 K intertwines the two actions A(I34)-modules. So the map U (H,K) of EndA(I34)-Mod(K). The actions of A(I12) on the source and on the target of (31) fac- tor through the aforementioned actions of EndA(I34)-Mod(K). The map (31) therefore intertwines the actions of A(I12). Corollary 4.10. The actions of A(I3) and of A(I4) on H (cid:2)ϕ1 K extend, by strong additivity, to an action of A(I34). Given two solitons H ∈ T − A and K ∈ T +A , we have upgraded H (cid:2)ϕ1 K to an object of : H (cid:2)K → H (cid:2)ϕ1 K. Similarly, Bim(R), and we have made sense of the unitary U (H,K) given a diffeomorphism ϕ2 : I234 → I234 as in Section 3.3, we can define K (cid:2)ϕ2 H and make sense of U (K,H) Definition 4.11. Let H ∈ T − βH,K : H (cid:2) K → K (cid:2) H is the composite A and K ∈ T +A be solitons. The braiding isomorphism : K (cid:2) H → K (cid:2)ϕ2 H. 2 1 βH,K : H (cid:2) K ∼= H (cid:2) K (cid:2) H0 −−−−−→ H (cid:2)ϕ1 K (cid:2) H0 U1 (cid:2) id −−−−−→ H (cid:2)ϕ1 (K (cid:2)ϕ2 H0) ∼= K (cid:2)ϕ2 (H (cid:2)ϕ1 H0) id (cid:2) U2 id (cid:2) U−1 1−−−−−→ K (cid:2)ϕ2 (H (cid:2) H0) U−1 −−−−−→ K (cid:2) (H (cid:2) H0) ∼= K (cid:2) H, (cid:2) id 2 (32) where H0 denotes the vacuum sector. We represent this isomorphism graphically as follows: βH,K : ∗ ∗ ∼= ∗ K H H0 K H ∗ → ∗ H0 K H ∗ → H0 ∗ H K → ∗ ∗ H0 H K → ∗ ∗ H0 H K ∼= ∗ H∗ K ∗ We have the following analogs of Lemma 3.10 and Lemma 3.11: Lemma 4.12. Let H ∈ T − commutative: A and K, L ∈ T +A be solitons. Then the following diagram is (H (cid:2) K) (cid:2) L U (H,K) 1 (cid:2) idL a H (cid:2) (K (cid:2) L) A similarly diagram holds for U2. U (H,K(cid:2)L) 1 24 (H (cid:2)ϕ1 K) (cid:2) L a H (cid:2)ϕ1 (K (cid:2) L). (33) Proof. By Corollary 4.7, the maps in the above diagram only depend on L as an A(I34)- module. We can therefore proceed as in Lemma 3.10. Let M := {L ∈ A(I34)-Mod : (33) holds}. If L = H0, then the horizontal arrows in (33) can be both identified with U (H,K) Lemma 4.13. Let H, K ∈ T − is commutative: . So M contains the vacuum sector. So M is all of A(I34)-Mod. 1 A , and L ∈ T +A be solitons. Then the following diagram H (cid:2) K (cid:2) L 1 U (H(cid:2)K,L) (H (cid:2) K) (cid:2)ϕ1 L U (H,K(cid:2)L) 1 a H (cid:2)ϕ1 (K (cid:2) L) idH (cid:2) U (K,L) H (cid:2)ϕ1 (K (cid:2)ϕ1 L), 1 (34) where the top horizontal arrow is the one constructed in (30). A similarly diagram holds for U2. Proof. Once again, the maps in (34) only depend on L as an A(I34)-module. Let M := {L ∈ A(I34)-Mod : (34) holds}. By Lemma 4.8, we may use the computation (27) to deduce that M contains the vacuum sector. As before, M is closed under taking direct sums and direct summands, so M is all of A(I34)-Mod. Finally, we have: Proposition 4.14. The braiding isomorphism (32) satisfies the two 'hexagon' axioms βH,K(cid:2)L = (idK (cid:2) βH,L)(βH,K (cid:2) idL) and βH(cid:2)K,L = (βH,L (cid:2) idK)(idH (cid:2) βK,L) (once again, we omit the associators for brevity). Proof. The proof of Proposition 3.12 applies word for word (use Lemma 4.12 in place of Lemma 3.10, and Lemma 4.13 in place of Lemma 3.11). We will show later, in Proposition 4.21, that there exists a unique braiding β : T − A × T +A Bim(R) that satisfies the hexagon axiom βH,K(cid:2)L = (idK(cid:2)βH,L)(βH,K(cid:2)idL). As a consequence, the braiding (32) is independent of the various choices that we made (e.g., the choice of diffeomorphisms ϕ1 and ϕ2). 4.3 The absorbing object In this section, we recall the results of our earlier paper [Hen17a], according to which the category of solitons admits an absorbing object. This is the only place where the condition that A has finite index is needed. We start by recalling the definition of an absorbing object: 25 Definition 4.15. An object Ω of a tensor category (T,⊗) is called absorbing if it is non-zero and satisfies (X (cid:54)= 0) ⇒ (X ⊗ Ω ∼= Ω ∼= Ω ⊗ X) ∀X ∈ T. If T admits a conjugation (in particular, if T is bi-involutive), then Ω ∈ T is absorbing if and only if it is non-zero and satisfies X ⊗ Ω ∼= Ω for every X (cid:54)= 0 (see the comments after [HP17, Def. 5.3]). Consider the following manifold (an equilateral triangle): ∆ := equipped with the smooth structure given by constant speed parametrization. We call the upper left side of this triangle ∆+, the lower left side ∆−, and the right side ∆free. Let Ω := H0(∆,A) be the vacuum sector of A associated to ∆, let S1 cut be as in Section 1.2, and let ϕ∆ : S1 cut → ∆− ∪ ∆+ be the constant speed parametrization that sends the lower half of S1 upper half of S1 A(∆− ∪ ∆+) on Ω to an action of A(S1 soliton. Note that, by Haag duality, cut to ∆− and the cut to ∆+. We use the diffeomorphism ϕ∆ to pull back the action of cut), and thus endow Ω with the structure of a EndTA(Ω) = A(∆free). The following important result was proven in [Hen17a, Thm. 9]: Proposition 4.16. If A is a conformal net with finite index, then the object Ω ∈ TA is absorbing. Remark 4.17. It is for the above proposition to hold that it was important to insist that all Hilbert spaces be separable. If we allow Hilbert spaces of arbitrarily large cardinalities, then the tensor category TA does not have an absorbing object. Remark 4.18. In the absence of the finite index condition, we do not know whether Ω is absorbing. Absorbing objects are important because they control half-braidings: Proposition 4.19 ([HP17, Prop. 5.9]). Let T be a category equipped with a tensor functor to Bim(R). Let Ω ∈ T be an absorbing object, and let (X, e) be in T (cid:48). Then e is completely determined by its value on Ω. Proof. Let Y be a non-zero object of T . Since e is a half-braiding, we have a commu- tative diagram Y (cid:2) X (cid:2) Ω eY (cid:2) idΩ X (cid:2) Y (cid:2) Ω eY (cid:2)Ω 26 idY (cid:2) eΩ Y (cid:2) Ω (cid:2) X. Pick an isomorphism φ : Y (cid:2) Ω → Ω. The following square is commutative eY (cid:2)Ω X (cid:2) (Y (cid:2) Ω) idX(cid:2) φ (Y (cid:2) Ω) (cid:2) X φ (cid:2)idX X (cid:2) Ω eΩ Ω (cid:2) X and so we get an equation eY (cid:2) idΩ = (idY (cid:2) e−1 Ω ) ◦ (φ−1 (cid:2) idX ) ◦ eΩ ◦ (idX (cid:2) φ). In particular, we see that eY (cid:2) idΩ is completely determined by eΩ. Since − (cid:2) Ω is a faithful functor, eY is completely determined by eY (cid:2) idΩ. Putting those two facts together, we see that eY is completely determined by eΩ. 4.4 The Drinfel'd center This section is devoted to the proof of our main theorem: Theorem 4.20. If A is a conformal net with finite index. Then the canonical map (T +A )(cid:48) → Bim(R) is fully faithful, and (T +A )(cid:48) = T − A . The statements in Theorem A are easy consequences of Theorem 4.20. The second A , and the bullet point in Theorem A is obtained by exchanging the roles of T +A and T − third bullet point is the following computation: Z(T +A ) = ZT +A (T +A ) = ZBim(R)(T +A ) ∩ T +A = T − A ∩ T +A = Rep(A), where we have used Lemma 4.2 for the last equality. We note that Lemma 4.1 (ac- cording to which T − A , T +A and Rep(A) are full subcategories of Bim(R)) has been used implicitly here, in the usage of the symbol ∩. Proposition 4.21. An object H ∈ Bim(R) admits at most one half-braiding with TA: e = (eK : H (cid:2) K → K (cid:2) H)K∈TA . Proof. Let e and e(cid:48) be half-braidings of H with TA. We wish to show that e = e(cid:48). By Proposition 4.16 and Proposition 4.19, it is enough to show that eΩ = e(cid:48) Ω. Consider the following R-R-bimodule map: Ω ◦ e−1 e(cid:48) (35) By the naturality of e and of e(cid:48), this map is equivariant for the actions of EndTA(Ω) = A(∆free). We may therefore treat (35) as a map of A(∆− ∪ ∆free)-R-bimodules. By Haag duality, Ω is an invertible A(∆− ∪ ∆free)-R-bimodule [BDH14, Prop. 3.10] (here, R is identified with A(∆+)op). So there exists an invertible R-R-bimodule map u : H → H that satisfies Ω : Ω (cid:2)R H → Ω (cid:2)R H. Ω ◦ e−1 e(cid:48) Ω = idΩ (cid:2) u. 27 Pick an isomorphism ω : Ω (cid:2) Ω → Ω in TA and consider the following diagram: e(cid:48) Ω (cid:2)id H (cid:2) Ω (cid:2) Ω (cid:2)id eΩ id H (cid:2) Ω (cid:2) Ω id(cid:2)ω id(cid:2)ω H (cid:2) Ω id H (cid:2) Ω Ω (cid:2) H (cid:2) Ω id(cid:2)u(cid:2)id id Ω (cid:2) H (cid:2) Ω eΩ(cid:2)Ω eΩ e(cid:48) Ω id(cid:2)e(cid:48) Ω id(cid:2)eΩ id(cid:2)id(cid:2)u Ω (cid:2) Ω (cid:2) H Ω (cid:2) Ω (cid:2) H ω(cid:2)id ω(cid:2)id Ω (cid:2) H id(cid:2)u Ω (cid:2) H The middle pentagon commutes because e is a half-braiding. The outer pentagon commutes by the corresponding property of e(cid:48). All the quadrilaterals are visibly com- mutative. It follows that idΩ (cid:2) u (cid:2) idΩ = idΩ(cid:2)H(cid:2)Ω. The functors Ω (cid:2) − and − (cid:2) Ω being faithful, we conclude that u = idH . Corollary 4.22. The braiding β : T − A × T +A Bim(R) defined in (32) is independent of the choices of diffeomorphisms ϕ1 and ϕ2. The same holds true for its restriction (24) to Rep(A). Proof of Theorem 4.20. The half-braiding constructed in Section 4.2 provides a functor that fits in a diagram T − A −→ (T +A )(cid:48) (T +A )(cid:48) T − A Bim(R) (36) We need to show that the functor (36) is an equivalence of categories. It is clearly faithful as T − A → Bim(R) and (T +A )(cid:48) → Bim(R) are both faithful functors. Recall from Lemma 4.1 that the functor T − A → Bim(R) is fully faithful. In order to check that the functor (36) is an equivalence of categories, it is therefore enough to show that: • The functor (T +A )(cid:48) → Bim(R) is full (and thus fully faithful). • For every object Y ∈ (T +A )(cid:48), there exists an object X ∈ T − between their images in Bim(R). A and an isomorphism 28 We start by the second item. By Lemma 4.2, it is enough to check that for every object (H, e) ∈ (T +A )(cid:48), the actions of A(I4) and A(I1) extend to an action of A(I41) on H. cut → ∆− ∪ ∆+ from the previous section. Let j : ∆ → ∆ be an orientation reversing involution that satisfies Recall that Ω := H0(∆,A), and recall the definition of ϕ∆ : S1 j(∆−) = ∆+ ∪ ∆free j(∆+) = ϕ∆(I3) j(∆free) = ϕ∆(I4), and is length-preserving in a neighbourhood of the vertex ∆−∩ ∆free of ∆. Recall from (17) that there is a unitary isomorphism v : Ω → L2(A(∆−)) that intertwines the actions of A(∆−), and satisfies v(A(j)(x)ξ) = (v(ξ))x for all x ∈ A(∆−) and ξ ∈ Ω. Using ϕ∆ to identify I34 with ∆−, we get an isomorphism f := L2(A(ϕ∆)) ◦ v : Ω → L2(A(∆−)) → L2(A(I34)) ∼= L2(R). Let us write b : S1 → S1 for the complex conjugation map z (cid:55)→ ¯z. Then the isomor- phism f : Ω → L2(R) intertwines the left actions of A(I34), and satisfies f(cid:0)A(j ◦ ϕ∆ ◦ b)(x) · ξ) = x · (f (ξ)(cid:1) for all x ∈ A(I12) and ξ ∈ Ω. action of A(I41) on H. Consider the isomorphism Recall that our goal is to show that the actions of A(I4) and A(I1) extend to an eΩ : H (cid:2) Ω → Ω (cid:2) H provided by the half-braiding. It is a homomorphism of R-R-bimodules. By the nat- urality axiom of half-braidings, it is also equivariant with respect to the actions of EndT +A (Ω) = A(∆free) on H (cid:2) Ω and on Ω (cid:2) H. We now consider the composite: F : Ω (cid:2) H (37) It is equivariant with respect to the left actions of A(I34), and intertwines the action of A(∆free) on Ω (cid:2) H with the following action on H: e−1 Ω−−−→ H (cid:2) Ω id(cid:2)f−−−→ H (cid:2) L2R ∼= H. A(∆free) A(j◦ϕ∆◦b)−1 −−−−−−−−→ A(I1) → B(H). We represent the isomorphism (37) graphically as follows: ∗ e−1 Ω−−−→ ∗ H Ω ∗ Ω H ∗ id(cid:2)f−−−→ ∗ ∗ ∼= H0 H H∗ ∗ The little stars are there to indicate that H is a priori a mere bimodule, as opposed to a soliton or a representation. Let us write : A(Ii) → B(H) ρH i : A(Ii) → B(Ω) ρΩ i and 29 for the actions of A(Ii) on H and on Ω, and let us write α for the following action of A(I1) on Ω: α : A(I1) A(j◦ϕ∆◦b) −−−−−−−−→ A(∆free) → B(Ω). By construction, the map (37) satisfies F ◦(cid:0)ρΩ F ◦(cid:0)α(x) (cid:2) idH 4 (x) (cid:2) idH (cid:1) = ρH (cid:1) = ρH 4 (x) ◦ F 1 (x) ◦ F ∀x ∈ A(I4) ∀x ∈ A(I1). (38) Since j is an isometry in a neighbourhood of ∆− ∩ ∆free, the maps ϕ∆ : I4 → ∆ and j ◦ ϕ∆ ◦ b : I1 → ∆ extend to a smooth map ϕ∆ ∪ (j ◦ ϕ∆ ◦ b) : I41 → ∆. 4 : A(I4) → B(Ω) and α : A(I1) → B(Ω) therefore extend to an action of 4 : A(I4) → B(H) : A(I1) → B(H) therefore also extend to an action of A(I14) on H. This The actions ρΩ A(I14) on Ω. By the intertwining properties (38) of F , the actions ρH and ρH 1 finishes the proof of the second item in the bullet list. We now turn our attention to the first item in the list. Let us write s : T − A → (T +A )(cid:48) for the functor (36). Let (H1, e1) and (H2, e2) be objects of (T +A )(cid:48), and let f : H1 → H2 In the first half of the proof, we be a morphism between their images in Bim(R). learned that H1 and H2 are in fact objects of T − A . By Lemma 4.1, f is a morphism in T − A . By Proposition 4.21, s(H1) = (H1, e1) and s(H2) = (H2, e2). Therefore, s(f ) : (H1, e1) → (H2, e2) is a morphism in (T +A )(cid:48). Now, by construction, s(f ) maps to f under the forgetful map (T +A )(cid:48) → Bim(R). Remark 4.23. The arguments in the proofs of Proposition 4.21 and of Theorem 4.20 never used the fact that the half-braidings are unitary (it just so happens that every half-braiding with T +A is unitary). The non-unitary version of Theorem A therefore also holds: ZBim(R)(T +A ) ∼= T − A , ZBim(R)(T − A ) ∼= T +A , Z(T +A ) ∼= Rep(A). References [BD95] [BDH14] John C. Baez and James Dolan, Higher-dimensional algebra and topological quantum field theory, J. Math. Phys. 36 (1995), no. 11, 6073 -- 6105. MR 1355899 (97f:18003) Arthur Bartels, Christopher L. Douglas, and Andr´e Henriques, Dualizabil- ity and index of subfactors, Quantum Topol. 5 (2014), no. 3, 289 -- 345. MR 3342166 30 [BDH15] , Conformal nets I: Coordinate-free nets, Int. Math. Res. Not. IMRN (2015), no. 13, 4975 -- 5052. MR 3439097 [BDH16] , Fusion of defects (formerly conformal nets III: fusion of defects), Memoirs of the AMS (2016). [BDH17] Arthur Bartels, Christopher L. Douglas, and Andr´e Henriques, Conformal Nets II: Conformal Blocks, Comm. Math. Phys. 354 (2017), no. 1, 393 -- 458. MR 3656522 [BE98] [BK01] [BSM90] [BW96] J. Bockenhauer and D. E. Evans, Modular invariants, graphs and α- induction for nets of subfactors. I, Comm. Math. Phys. 197 (1998), no. 2, 361 -- 386. MR 1652746 (2000c:46121) Bojko Bakalov and Alexander Kirillov, Jr., Lectures on tensor categories and modular functors, University Lecture Series, vol. 21, American Math- ematical Society, Providence, RI, 2001. MR 1797619 (2002d:18003) Detlev Buchholz and Hanns Schulz-Mirbach, Haag duality in conformal quantum field theory, Rev. Math. Phys. 2 (1990), no. 1, 105 -- 125. MR 1079298 (92a:81106) John W. Barrett and Bruce W. Westbury, Invariants of piecewise-linear 3-manifolds, Trans. Amer. Math. Soc. 348 (1996), no. 10, 3997 -- 4022. MR 1357878 [CJM+05] Alan L. Carey, Stuart Johnson, Michael K. Murray, Danny Stevenson, and Bai-Ling Wang, Bundle gerbes for Chern-Simons and Wess-Zumino- Witten theories, Comm. Math. Phys. 259 (2005), no. 3, 577 -- 613. MR 2174418 (2007a:58023) [CKLW15] Sebastiano Carpi, Yasuyuki Kawahigashi, Roberto Longo, and Mih´aly Weiner, From vertex operator algebras to conformal nets and back, arXiv:1503.01260, to appear in Memoirs of the AMS, 2015. [Con94] Alain Connes, Noncommutative geometry, Academic Press, Inc., San Diego, CA, 1994. MR 1303779 (95j:46063) [DLM96] Chongying Dong, Haisheng Li, and Geoffrey Mason, Simple currents and extensions of vertex operator algebras, Comm. Math. Phys. 180 (1996), no. 3, 671 -- 707. MR 1408523 (97j:17031) [DMNO13] Alexei Davydov, Michael Muger, Dmitri Nikshych, and Victor Ostrik, The Witt group of non-degenerate braided fusion categories, J. Reine Angew. Math. 677 (2013), 135 -- 177. MR 3039775 [DSPS13] Christopher L. Douglas, Christopher Schommer-Pries, and Noah Snyder, Dualizable tensor categories, arXiv:1312.7188, 2013. [DW90] Robbert Dijkgraaf and Edward Witten, Topological gauge theories and group cohomology, Comm. Math. Phys. 129 (1990), no. 2, 393 -- 429. MR 1048699 (91g:81133) [ENO05] Pavel Etingof, Dmitri Nikshych, and Viktor Ostrik, On fusion categories, Ann. of Math. (2) 162 (2005), no. 2, 581 -- 642. MR 2183279 31 [FGK88] G. Felder, K. Gawedzki, and A. Kupiainen, Spectra of Wess-Zumino- Witten models with arbitrary simple groups, Comm. Math. Phys. 117 (1988), no. 1, 127 -- 158. MR 946997 (89k:81077) [FHLT10] Daniel S. Freed, Michael J. Hopkins, Jacob Lurie, and Constantin Teleman, Topological quantum field theories from compact Lie groups, A celebration of the mathematical legacy of Raoul Bott, CRM Proc. Lecture Notes, vol. 50, Amer. Math. Soc., Providence, RI, 2010, pp. 367 -- 403. MR 2648901 (2011i:57040) [Fro76] [FRS89] [FSS96] [FSS15] [GF93] [Gui66] [Haa75] [Hen15] [Hen16] Jurg Frohlich, New super-selection sectors ("soliton-states") in two dimen- sional Bose quantum field models, Comm. Math. Phys. 47 (1976), no. 3, 269 -- 310. MR 0413868 (54 #1980) K. Fredenhagen, K.-H. Rehren, and B. Schroer, Superselection sectors with braid group statistics and exchange algebras. I. General theory, Comm. Math. Phys. 125 (1989), no. 2, 201 -- 226. MR 1016869 (91c:81047) J. Fuchs, A. N. Schellekens, and C. Schweigert, A matrix S for all simple current extensions, Nuclear Phys. B 473 (1996), no. 1-2, 323 -- 366. MR 1409292 (97i:81041) Domenico Fiorenza, Hisham Sati, and Urs Schreiber, A higher stacky per- spective on Chern-Simons theory, Mathematical aspects of quantum field theories, Math. Phys. Stud., Springer, Cham, 2015, pp. 153 -- 211. MR 3330242 Fabrizio Gabbiani and Jurg Frohlich, Operator algebras and conformal field theory, Comm. Math. Phys. 155 (1993), no. 3, 569 -- 640. MR 1231644 (94m:81090) Alain Guichardet, Sur la cat´egorie des alg`ebres de von Neumann, Bull. Sci. Math. (2) 90 (1966), 41 -- 64. MR 0201989 Uffe Haagerup, The standard form of von Neumann algebras, Math. Scand. 37 (1975), no. 2, 271 -- 283. MR 0407615 (53 #11387) Andr´e Henriques, What Chern -- Simons theory assigns to a point, arXiv:1503.06254, 2015. , The classification of chiral WZW models by H 4 +(BG, Z), Contem- porary Mathematics (2016). [Hen17a] , Conformal nets are factorization algebras, Proceedings of the String-Math 2016 conference (2017). [Hen17b] , Loop groups and diffeomorphism groups of the circle as colimits, arXiv:1706.08471, 2017. [HP17] [JS91] Andr´e Henriques and David Penneys, Bicommutant categories from fusion categories, Selecta Math. (N.S.) 23 (2017), no. 3, 1669 -- 1708. MR 3663592 Andr´e Joyal and Ross Street, Tortile Yang-Baxter operators in tensor categories, J. Pure Appl. Algebra 71 (1991), no. 1, 43 -- 51. MR 1107651 (92e:18006) 32 [Kaw02] Yasuyuki Kawahigashi, Generalized Longo-Rehren subfactors and α- induction, Comm. Math. Phys. 226 (2002), no. 2, 269 -- 287. MR 1892455 (2003i:46067) [KLM01] Yasuyuki Kawahigashi, Roberto Longo, and Michael Muger, Multi- interval subfactors and modularity of representations in conformal field theory, Comm. Math. Phys. 219 (2001), no. 3, 631 -- 669. MR 1838752 (2002g:81059) [Li01] [Lon89] [Lon08] [LR95] [Lur09] [Lur11] [LX04] [Maj91] Haisheng Li, Certain extensions of vertex operator algebras of affine type, Comm. Math. Phys. 217 (2001), no. 3, 653 -- 696. MR 1822111 (2003a:17036) Roberto Longo, Index of subfactors and statistics of quantum fields. I, Comm. Math. Phys. 126 (1989), no. 2, 217 -- 247. MR 1027496 R. Longo, Lectures on conformal nets II http://www.mat.uniroma2.it/∼longo/Lecture% 20Notes.html, 2008. R. Longo and K.-H. Rehren, Nets of subfactors, Rev. Math. Phys. 7 (1995), no. 4, 567 -- 597, Workshop on Algebraic Quantum Field Theory and Jones Theory (Berlin, 1994). MR 1332979 (96g:81151) , Jacob Lurie, On the classification of topological field theories, Current developments in mathematics, 2008, Int. Press, Somerville, MA, 2009, pp. 129 -- 280. MR 2555928 (2010k:57064) J. Lurie, Lectures notes on von Neumann algebras http://www.math.harvard.edu/∼lurie/261y.html, 2011. Roberto Longo and Feng Xu, Topological sectors and a dichotomy in con- formal field theory, Comm. Math. Phys. 251 (2004), no. 2, 321 -- 364. MR 2100058 (2005i:81087) , Shahn Majid, Representations, duals and quantum doubles of monoidal categories, Proceedings of the Winter School on Geometry and Physics (Srn´ı, 1990), no. 26, 1991, pp. 197 -- 206. MR 1151906 [MS89] Gregory Moore and Nathan Seiberg, Taming the conformal zoo, Phys. Lett. B 220 (1989), no. 3, 422 -- 430. MR 992362 (90e:81217) [MTW16] V. Morinelli, Y Tanimoto, and M. Weiner, Conformal covariance and the split property, arXiv:1609.02169, 2016. [Mug03] Michael Muger, From subfactors to categories and topology. II. The quan- tum double of tensor categories and subfactors, J. Pure Appl. Algebra 180 (2003), no. 1-2, 159 -- 219. MR 1966525 (2004f:18014) [RT91] N. Reshetikhin and V. G. Turaev, Invariants of 3-manifolds via link poly- nomials and quantum groups, Invent. Math. 103 (1991), no. 3, 547 -- 597. MR 1091619 [Sau83] Jean-Luc Sauvageot, Sur le produit tensoriel relatif d'espaces de Hilbert, J. Operator Theory 9 (1983), no. 2, 237 -- 252. MR 703809 (85a:46034) 33 [Tak03] [Tho11] [TV92] [Was98] [Wit89] [Wra10] [Xu00] [Yam92] M. Takesaki, Theory of operator algebras. II, Encyclopaedia of Mathemat- ical Sciences, vol. 125, Springer-Verlag, Berlin, 2003, Operator Algebras and Non-commutative Geometry, 6. MR 1943006 (2004g:46079) Andreas Thom, A remark about the Connes fusion tensor product, Theory Appl. Categ. 25 (2011), No. 2, 38 -- 50. MR 2771095 (2012g:46085) V. G. Turaev and O. Ya. Viro, State sum invariants of 3-manifolds and quantum 6j-symbols, Topology 31 (1992), no. 4, 865 -- 902. MR 1191386 Antony Wassermann, Operator algebras and conformal field theory. III. Fu- sion of positive energy representations of LSU(N ) using bounded operators, Invent. Math. 133 (1998), no. 3, 467 -- 538. MR 1645078 (99j:81101) Edward Witten, Quantum field theory and the Jones polynomial, Comm. Math. Phys. 121 (1989), no. 3, 351 -- 399. MR 990772 (90h:57009) Kevin Wray, Extended topological gauge theories in codimension zero and higher, Master's Thesis, Universiteit van Amsterdam, 2010. Feng Xu, Jones-Wassermann subfactors for disconnected intervals, Com- mun. Contemp. Math. 2 (2000), no. 3, 307 -- 347. MR 1776984 (2001f:46094) Shigeru Yamagami, Algebraic aspects in modular theory, Publ. Res. Inst. Math. Sci. 28 (1992), no. 6, 1075 -- 1106. MR 1203761 (94c:46123) 34
1607.04472
3
1607
2017-07-27T12:12:33
Representations of *-algebras by unbounded operators: C*-hulls, local-global principle, and induction
[ "math.OA" ]
We define a C*-hull for a *-algebra, given a notion of integrability for its representations on Hilbert modules. We establish a local-global principle which, in many cases, characterises integrable representations on Hilbert modules through the integrable representations on Hilbert spaces. The induction theorem constructs a C*-hull for a certain class of integrable representations of a graded *-algebra, given a C*-hull for its unit fibre.
math.OA
math
REPRESENTATIONS OF *-ALGEBRAS BY UNBOUNDED OPERATORS: C*-HULLS, LOCAL–GLOBAL PRINCIPLE, AND INDUCTION 7 1 0 2 l u J 7 2 ] . A O h t a m [ 3 v 2 7 4 4 0 . 7 0 6 1 : v i X r a RALF MEYER Abstract. We define a C∗-hull for a ∗-algebra, given a notion of integrability for its representations on Hilbert modules. We establish a local–global principle which, in many cases, characterises integrable representations on Hilbert mod- ules through the integrable representations on Hilbert spaces. The induction theorem constructs a C∗-hull for a certain class of integrable representations of a graded ∗-algebra, given a C∗-hull for its unit fibre. Contents Introduction Integrable representations and C∗-hulls 1. 2. Representations by unbounded operators on Hilbert modules 3. 4. Polynomials in one variable I 5. Local–Global principles 6. Polynomials in one variable II 7. Bounded and locally bounded representations 8. Commutative C∗-hulls 9. From graded ∗-1––------algebras to Fell bundles 10. Locally bounded unit fibre representations 11. Fell bundles with commutative unit fibre 12. Rieffel deformation 13. Twisted Weyl algebras References 1 5 10 15 19 27 30 37 43 53 57 64 66 71 1. Introduction Savchuk and Schmüdgen [26] have introduced a method to define and classify the integrable representations of certain ∗-algebras by an inductive construction. The original goal of this article was to clarify this method and thus make it apply to more situations. This has led me to reconsider some foundational aspects of the theory of representations of ∗-algebras by unbounded operators. This is best explained by formulating an induction theorem that is inspired by [26]. g∈G Ag be a G-graded unital ∗-algebra. That is, Ag · Ah ⊆ Agh, A∗ g = Ag−1, and 1 ∈ Ae. In particular, the unit fibre Ae is a unital ∗-algebra. Many interesting examples of this situation are studied in [7,26]. A Fell bundle over G is a family of subspaces (Bg)g∈G Let G be a discrete group with unit element e ∈ G. Let A = L 2010 Mathematics Subject Classification. Primary 47L60; Secondary 46L55. Key words and phrases. unbounded operator; regular Hilbert module operator; integrable representation; induction of representations; graded ∗-algebra; Fell bundle; C∗-algebra generated by unbounded operators; C∗-envelope; C∗-hull; host algebra; Weyl algebra; canonical commutation relations; Local–Global Principle; Rieffel deformation. 1 RALF MEYER 2 of a C∗-algebra B (which is not part of the data) such that Bg · Bh ⊆ Bgh and g = Bg−1. The universal choice for B is the section C∗-algebra of the Fell bundle. B∗ Briefly, our main result says the following. Let Be be a C∗-algebra such that "integrable" "representations" of Ae are "equivalent" to "representations" of Be. g )g∈G over G such Under some technical conditions, we construct a Fell bundle (B+ that "integrable" "representations" of A are "equivalent" to "representations" of its section C∗-algebra. Here the words in quotation marks must be interpreted carefully to make this true. A representation of a ∗-algebra A on a Hilbert D-module E is an algebra ho- momorphism π from A to the algebra of D-module endomorphisms of a dense D-submodule E ⊆ E with hξ, π(a)ηi = hπ(a∗)ξ, ηi for all ξ, η ∈ E, a ∈ A. The representation induces a graph topology on E. We restrict to closed representations most of the time, that is, we require E to be complete in the graph topology. The difference from usual practice is that we consider representations on Hilbert modules over C∗-algebras. A representation of a C∗-algebra B on a Hilbert module E is a nondegenerate ∗-homomorphism B → B(E), where B(E) denotes the C∗-algebra of adjointable operators on E. The notion of "integrability" for representations is a choice. The class of all Hilbert space representations of a ∗-algebra may be quite wild. Hence it is customary to limit the study to some class of "nice" or "integrable" representations. For instance, for the universal enveloping algebra of the Lie algebra of a Lie group G, we may call those representations "integrable" that come from a unitary representation of G. This example suggests the name "integrable" representations. In our theorem, a notion of integrability for representations of Ae ⊆ A on all Hilbert modules over all C∗-algebras is fixed. A representation of A is called integrable if its restriction to Ae is integrable. The induction theorem describes the integrable representations of A in terms of integrable representations of Ae. For instance, if Ae is finitely generated and commutative, then we may call a representation π on a Hilbert module integrable if the closure π(a) is a regular, self-adjoint operator for each a ∈ Ae with a = a∗; all examples in [7,26] are of this type. An "equivalence" between the integrable representations of a unital ∗-algebra A and the representations of a C∗-algebra B is a family of bijections – one for each Hilbert module E over each C∗-algebra D – between the sets of integrable represen- tations of A and of representations of B on E; these bijections must be compatible with isometric intertwiners and interior tensor products. These properties require some more definitions. First, an isometric intertwiner between two representations is a Hilbert module isometry – not necessarily adjointable – between the underlying Hilbert modules that restricts to a left module map between the domains of the representations. For an equivalence between integrable representations of A and representations of B we require an isometry to intertwine two representations of B if and only if it intertwines the corresponding integrable representations of A. Secondly, a C∗-correspondence from D1 to D2 is a Hilbert D2-module F with a representation of D1. Given such a correspondence and a Hilbert D1-module E, the interior tensor product E ⊗D1 F is a Hilbert D2-module. A representation of A or B on E induces a representation on E ⊗D1 F. We require our bijections between integrable representations of A and representations of B to be compatible with this interior tensor product construction on representations. We call B a C∗-hull for the integrable representations of A if the integrable representations of A are equivalent to the representations of B as explained above, that is, through a family of bijections compatible with isometric intertwiners and REPRESENTATIONS BY UNBOUNDED OPERATORS 3 interior tensor products. The Induction Theorem builds a C∗-hull for the integrable representations of A using a C∗-hull for the integrable representations of Ae and assuming a further mild technical condition, which we explain below. Many results of the general theory remain true if we only require the equivalence of representations to be compatible with interior tensor products and unitary ∗-intertwiners, that is, isomorphisms of representations; we speak of a weak C∗-hull in this case. The Induction Theorem, however, fails for weak C∗-hulls. We show this by a counterexample. Some results only need the class of integrable representations to have some properties that are clearly necessary for the existence of a C∗-hull or weak C∗-hull, but they do not need the (weak) C∗-hull itself. This is formalised in our notions of admissible and weakly admissible classes of representations. For example, let A be commutative. Let A be the space of characters of A with the topology of pointwise convergence. If A is locally compact and A is countably generated, then C0( A) is a C∗-hull for the integrable representations of A as defined above, that is, those representations where each π(a) for a ∈ A with a = a∗ is regular and self-adjoint. If, say, A = C[x] with x = x∗, then the C∗-hull is C0(R). Here the equivalence of representations maps an integrable representation π of C[x] to the functional calculus homomorphism for the regular, self-adjoint operator π(x). If A is not locally compact, then the integrable representations of A defined above still form an admissible class, but they have no C∗-hull. If, say, A is the algebra of polynomials in countably many variables, then A = R∞, which is not locally compact. The problem of associating C∗-algebras to this ∗-algebra has recently been studied by Grundling and Neeb [12]. From our point of view, this amounts to choosing a smaller class of "integrable" representations that does admit a C∗-hull. We have now explained the terms in quotation marks in our Induction Theorem and how we approach the representation theory of ∗-algebras. Most previous work focused either on representations on Hilbert spaces or on single unbounded operators on Hilbert modules. Hilbert module representations occur both in the assumptions and in the conclusions of the Induction Theorem, and hence we cannot prove it without considering representations on Hilbert modules throughout. In addition, taking into account Hilbert modules makes our C∗-hulls unique. Besides the Induction Theorem, the other main strand of this article are Local– Global Principles, which aim at reducing the study of integrability for representations on general Hilbert modules to representations on Hilbert space. We may use a state ω on the coefficient C∗-algebra D of a Hilbert module E to complete E to a Hilbert space. Thus a representation of A on E induces Hilbert space representations for all states on D. The Local–Global Principle says that a representation of A on E is integrable if and only if these induced Hilbert space representations are integrable for all states; the Strong Local–Global Principle says the same with all states replaced by all pure states. We took these names from [14]. Earlier results of Pierrot [20] show that the Strong Local–Global Principle holds for any class of integrable representations that is defined by certain types of conditions, such as the regularity and self-adjointness of π(a) for certain a ∈ A with a = a∗. For instance, this covers the integrable representations of commutative ∗-algebras and of universal enveloping algebras of Lie algebras. In all examples that we treat, the regularity of π(a) for certain a ∈ A is part of the definition of an integrable representation. Other elements of A may, however, act by irregular operators in some integrable representations. Thus affiliation and regularity are important to study the integrable representations in concrete examples, but cannot play a foundational role for the general representation theory of ∗-algebras. If B is generated in the sense of Woronowicz [31] by some self-adjoint, affiliated multipliers that belong to A, then it is a C∗-hull and the Strong Local–Global 4 RALF MEYER g = B+ g )∗ = B+ To prove B+ gh Principle holds (see Theorem 5.19). A counterexample shows that this theorem breaks down if the generating affiliated multipliers are not self-adjoint: both the Local–Global Principle and compatibility with isometric intertwiners fail in the counterexample. So regularity without self-adjointness seems to be too weak for many purposes. The combination of regularity and self-adjointness is an easier notion than regularity alone. A closed operator T is regular and self-adjoint if and only if T − λ is surjective for all λ ∈ C \ R, if and only if the Cayley transform of T is unitary, if and only if T has a functional calculus homomorphism on C0(R). Our input data is a graded ∗-algebra A =L Now we describe the Fell bundle in the Induction Theorem and, along the way, the further condition besides compatibility with isometric intertwiners that it needs. g∈G Ag and a C∗-hull Be for Ae. A representation of A is integrable if its restriction to Ae is integrable. We seek a C∗-hull for the integrable representations of A. g−1, so that the subspaces B+ e · B+ g = B+ As in [26], we induce representations from Ae to A, and this requires a positivity condition. We call representations of Ae that may be induced to A inducible. We describe a quotient C∗-algebra B+ e of Be that is a C∗-hull for the inducible, integrable representations of A. It is the unit fibre of our Fell bundle. If a representation π of A is integrable, then its restriction to Ae is integrable e . The identity e corresponds to a particular ("universal") inducible, integrable e ⊆ B+ e . The e are adjointable. Their closed linear e is faithful. The most difficult h ⊆ B+ and inducible. Thus it corresponds to a representation ¯π+ correspondence on B+ representation of Ae on B+ operators π(a)¯π+ span is the fibre B+ point is to prove B+ and (B+ e of B+ e . Its domain is a dense right ideal B+ e (b) on E for a ∈ Ag, b ∈ B+ g of our Fell bundle at g provided π+ e · B+ g for all g ∈ G; this easily implies B+ g · B+ g ⊆ B(E) form a Fell bundle. g , we need compatibility with isometric intertwiners and that induction maps inducible, integrable representations of Ae to integrable representations of A. Two counterexamples show that both assumptions are necessary for the Induction Theorem. g )g∈G over G is equivalent to an action of G on B+ Fell bundles are noncommutative partial dynamical systems. More precisely, a Fell bundle (B+ e by partial Morita–Rieffel equivalences; this is made precise in [4]. In the examples in [7,26], the group G is almost always Z; the C∗-algebras Be and hence B+ e are commutative; and the resulting Fell bundle comes from a partial action of G on the spectrum e . In these examples, the section C∗-algebra is a partial crossed product. This of B+ may also be viewed as the groupoid C∗-algebra of the transformation groupoid e . We show that the C∗-hull B for the partial action of G on the spectrum of B+ for the integrable representations of A is a twisted groupoid C∗-algebra of this transformation groupoid whenever Be is commutative. We give some criteria when the twist is absent, and examples where the twist occurs. One way to insert such twists is by Rieffel deformation, using a 2-cocycle on the group G. We show that Rieffel deformation is compatible with the construction of C∗-hulls. We describe commutative and noncommutative C∗-hulls for the polynomial algebra C[x] in §4 and §6; the noncommutative C∗-hulls for C[x] make very good counterexamples. We classify and study commutative C∗-hulls in §8. Many results about them generalise easily to locally bounded representations. Roughly speaking, these are representations where the vectors on which the representation acts by bounded operators form a core. The only ∗-algebras for which we treat locally bounded representations in some detail are the commutative ones. Through the Induction Theorem, the representation theory of commutative ∗-algebras is important even for noncommutative algebras because they may admit a grading by some group with commutative unit fibre. Many examples of this REPRESENTATIONS BY UNBOUNDED OPERATORS 5 are treated in detail in [7, 26]. We discuss untwisted and twisted Weyl algebras in finitely and infinitely many generators in §13. The twists involved are Rieffel deformations. Since these examples have commutative unit fibres, the resulting C∗-hulls are twisted groupoid C∗-algebras. As it turns out, all twists of the relevant groupoids are trivial, so that the twists do not change the representation theory of the Weyl algebras up to equivalence. I am grateful to Yuriy Savchuk for several discussions, which led me to pursue this project and eliminated mistakes from early versions of this article. And I am grateful to the referee as well for several useful suggestions. 2. Representations by unbounded operators on Hilbert modules Let A be a unital ∗-algebra, D a C∗-algebra, and E a Hilbert D-module. Our convention is that inner products on Hilbert spaces and Hilbert modules are linear in the second and conjugate-linear in the first variable. Definition 2.1. A representation of A on E is a pair (E, π), where E ⊆ E is a dense D-submodule and π : A → EndD(E) is a unital algebra homomorphism to the algebra of D-module endomorphisms of E, such that hπ(a)ξ, ηiD = hξ, π(a∗)ηiD for all a ∈ A, ξ, η ∈ E. We call E the domain of the representation. We may drop π from our notation by saying that E is an A, D-bimodule with the right module structure inherited from E, or we may drop E because it is the common domain of the partial linear maps π(a) on E for all a ∈ A. We equip E with the graph topology, which is generated by the graph norms kξka := k(ξ, π(a)ξ)k := khξ, ξi + hπ(a)ξ, π(a)ξik1/2 = khξ, π(1 + a∗a)ξik1/2 for a ∈ A. The representation is closed if E is complete in this topology. A core for (E, π) is an A, D-subbimodule of E that is dense in E in the graph topology. Definition 2.1 for D = C is the usual definition of a representation of a ∗-algebra on a Hilbert space by unbounded operators. This situation has been studied extensively (see, for instance, [27]). For E = D with the canonical Hilbert D-module structure, we get representations of A by densely defined unbounded multipliers. The domain of such a representation is a dense right ideal D ⊆ D. This situation is a special case of the "compatible pairs" defined by Schmüdgen [28]. Given two norms p, q, we write p (cid:22) q if there is a scalar c > 0 with p ≤ cq. Lemma 2.2. The set of graph norms partially ordered by (cid:22) is directed: all a1, . . . , an ∈ A there are b ∈ A and c ∈ R>0 so that kξkai representation (E, π), any ξ ∈ E, and i = 1, . . . , n. for ≤ ckξkb for any Proof. Let b =Pn j=1 a∗ 0 ≤ hξ, π(1 + a∗ ≤ hξ, π(1 + a∗ i ai)ξi +X j aj. The following computation implies kξkai i ai)ξi ≤ 5/4kξkb: i6=j hπ(aj)ξ, π(aj)ξi + hπ(b − 1/2)ξ, π(b − 1/2)ξi = hξ, π(1 + b + (b − 1/2)2)ξi = hξ, π(5/4 + b2)ξi ≤ 5/4hξ, π(1 + b∗b)ξi. (cid:3) Definition 2.3 ([19], [16, Chapter 9]). A densely defined operator t on a Hilbert module E is semiregular if its adjoint is also densely defined; it is regular if it is closed, semiregular and 1 + t∗t has dense range. An affiliated multiplier of a C∗-algebra D is a regular operator on D viewed as a Hilbert D-module. The closability assumption in [19, Definition 2.1.(ii)] is redundant by [14, Lemma 2.1]. Regularity was introduced by Baaj and Julg [1], affiliation by Woronowicz [30]. RALF MEYER 6 Remark 2.4. Let (E, π) be a representation of A on E and let a ∈ A. The opera- tor π(a) is automatically semiregular because π(a)∗ is defined on E. The closure π(a) of π(a) need not be regular. The regularity of π(a) for some a ∈ A is often assumed in the definition of integrable representations. For non-commutative A, we should expect that π(a) is irregular for some a ∈ A even if π is integrable. For instance, a remark after Corollaire 1.27 in [20] says that this happens for certain symmetric ele- ments in the universal enveloping algebra U(g) for a simply connected Lie group G: they act by irregular operators in certain representations that integrate to unitary representations of G. a∈A dom π(a). The usual norm on E is the graph norm for 0 ∈ A. Hence the inclusion map E ,→ E is continuous for the graph topology on E and extends continuously to the completion E of E in the graph topology. Proposition 2.5. The canonical map E → E is injective, and its image is (2.6) E = \ Thus (E, π) is closed if and only if E =T a∈A dom π(a). Each π(a) extends uniquely to a continuous operator π(a) on E. This defines a closed representation (E, π) of A, called the closure of (E, π). Proof. The operator π(a) for a ∈ A is semiregular and hence closable by [14, Lemma 2.1]. Equivalently, the canonical map from the completion of E in the graph norm for a to E is injective. Its image is dom π(a), the domain of the closure of π(a). The graph norms for a ∈ A form a directed set that defines the graph topology on E. So the completion of E in the graph topology is the projective limit of the graph norm completions for a ∈ A. Since each of these graph norm completions embeds into E, the projective limit in question is just an intersection in E, giving (2.6). For Hilbert space representations, this is [27, Proposition 2.2.12]. The operators π(a) ∈ EndD(E) for a ∈ A are continuous in the graph topology. Thus they extend uniquely to continuous linear operators π(a) ∈ EndD(E). These are again D-linear and the map π is linear and multiplicative because extending operators to a completion is additive and functorial. The set of (ξ, η) ∈ E × E with hξ, π(a)ηi = hπ(a∗)ξ, ηi for all a ∈ A is closed in the graph topology and contains E × E, which is dense in E × E. Hence this equation holds for all ξ, η ∈ E. So (E, π) is a representation of A on E. The graph topology on E for π extends the graph topology on E for π and hence is complete. So (E, π) is a closed representation. (cid:3) We shall need a generalisation of (2.6) that replaces A by a sufficiently large subset. Definition 2.7. A subset S ⊆ A is called a strong generating set if it generates A as an algebra and the graph norms for a ∈ S generate the graph topology in any representation. That is, for any representation on a Hilbert module, any vector ξ in its domain and any a ∈ A, there are c ≥ 1 in R and b1, . . . , bn ∈ S with kξka ≤ cPn An estimate kξka ≤ cPn j dj = c ·Pn a∗a +Pm i=1kξkbi j=1 d∗ . i=1kξkbi i=1 b∗ is usually shown by finding d1, . . . , dm ∈ A with i bi, compare the proof of Lemma 2.2. Example 2.8. Let Ah := {a ∈ A a = a∗} be the set of symmetric elements. Call an element of A positive if it is a sum of elements of the form a∗a. The positive elements and, a fortiori, the symmetric elements form strong generating sets for A. REPRESENTATIONS BY UNBOUNDED OPERATORS 7 Any element is of the form a1 + ia2 with a1, a2 ∈ Ah, and (cid:18) a + 1 (cid:19)2 2 a = (cid:18) a − 1 (cid:19)2 2 − for a ∈ Ah. Thus the positive elements generate A as an algebra. The graph norms for positive elements generate the graph topology by the proof of Lemma 2.2. Proposition 2.9. Let S ⊆ A be a strong generating set. Two closed representations (E1, π1) and (E2, π2) of A on the same Hilbert module E are equal if and only if π1(a) = π2(a) for all a ∈ S. Proof. One direction is trivial. To prove the non-trivial direction, assume π1(a) = π2(a) for all a ∈ S. Let (E, π) = (Ei, πi) for i = 1, 2. The completion of E for the completion of E in the sum of graph norms Pn Tn the graph norm of a is dom π(a), compare the proof of Proposition 2.5. Hence for b1, . . . , bn ∈ S is k=1 dom π(bk). These sums of graph norms for b1, . . . , bn ∈ S form a directed set that generates the graph topology on E. Hence (2.10) E = \ k=1kξkbk dom π(a), a∈S compare the proof of (2.6). So E1 = E2. Moreover, π1(a) = π1(a)E1 = π2(a)E2 = π2(a) for all a ∈ S. Since S generates A as an algebra and πi(A)Ei ⊆ Ei, this (cid:3) implies π1 = π2. Proposition 2.9 may fail for generating sets that are not strong, see Example 4.2. Corollary 2.11. Let S be a strong generating set of A and let (E, π) be a closed representation of A with dom π(a) = E for each a ∈ S. Then E = E and π is a ∗-homomorphism to the C∗-algebra B(E) of adjointable operators on E. Proof. Equation (2.10) gives E = E. Since π(a∗) ⊆ π(a)∗ and π(a∗) is defined everywhere, it is adjoint to π(a). So π(a) ∈ B(E) and π is a ∗-homomorphism to B(E). (cid:3) Lemma 2.12. Let A be a unital C∗-algebra. Any closed representation of A on E has domain E = E and is a unital ∗-homomorphism to B(E). Proof. Let a ∈ A. There are a positive scalar C > 0 and b ∈ A with a∗a + b∗b = C; say, take C = kak2 and b = C − a∗a. Then √ hπ(a)ξ, π(a)ξi ≤ hπ(a)ξ, π(a)ξi + hπ(b)ξ, π(b)ξi = hξ, π(a∗a + b∗b)ξi = Chξ, ξi for all ξ ∈ E. Thus the graph topology on E is equivalent to the norm topology on E. Hence E = E for any closed representation. (cid:3) An isometry I : E1 ,→ E2 between two Hilbert D-modules E1 and E2 is a right D-module map with hIξ1, Iξ2i = hξ1, ξ2i for all ξ1, ξ2 ∈ E1. Definition 2.13. Let (E1, π1) and (E2, π2) be representations on Hilbert D-modules E1 and E2, respectively. An isometric intertwiner between them is an isometry I : E1 ,→ E2 with I(E1) ⊆ E2 and I ◦ π1(a)(ξ) = π2(a) ◦ I(ξ) for all a ∈ A, ξ ∈ E1; equivalently, I◦π1(a) ⊆ π2(a)◦I for all a ∈ A, that is, the graph of π2(a)◦I contains the graph of I ◦ π1(a). We neither ask I to be adjointable nor I(E1) = E2. Let Rep(A, D) be the category with closed representations of A on Hilbert D-modules as objects, isometric intertwiners as arrows, and the usual composition. The unit arrow on (E, π) is the identity operator on E. RALF MEYER 8 Lemma 2.14. Let (E1, π1) and (E2, π2) be representations on Hilbert D-modules E1 and E2, respectively, and let I : E1 ,→ E2 be an isometric intertwiner. Then I is also an intertwiner between the closures of (E1, π1) and (E2, π2). Proof. Since I intertwines the representations, it is continuous for the graph topolo- gies on E1 and E2. Hence I maps the domain of the closure π1 into the domain of π2. This extension is still an intertwiner because it is an intertwiner on a dense (cid:3) subspace. Proposition 2.15. Let (E1, π1) and (E2, π2) be closed representations of A on Hilbert D-modules E1 and E2, respectively. Let S ⊆ A be a strong generating set. An isometry I : E1 ,→ E2 is an intertwiner from (E1, π1) to (E2, π2) if and only if I ◦ π1(a) ⊆ π2(a) ◦ I for all a ∈ S. Proof. First let I satisfy I ◦ π1(a) ⊆ π2(a) ◦ I for all a ∈ S. Then I maps the domain of π1(a) into the domain of π2(a) for each a ∈ S. Now (2.10) implies I(E1) ⊆ E2. Since πi(a) = πi(a)Ei, we get I(π1(a)(ξ)) = π2(a)(I(ξ)) for all a ∈ S, ξ ∈ E1. Since S generates A as an algebra and πi(A)Ei ⊆ Ei, this implies I ◦ π1(a) = π2(a) ◦ I for all a ∈ A, that is, I is an intertwiner. Conversely, assume that I is an intertwiner from (E1, π1) to (E2, π2). Equivalently, I ◦ π1(a) ⊆ π2(a) ◦ I for all a ∈ A. We have I ◦ π1(a) = I ◦ π1(a) because I is an isometry, and π2(a) ◦ I ⊆ π2(a) ◦ I. Thus I ◦ π1(a) ⊆ π2(a) ◦ I for all a ∈ A. (cid:3) Now we relate the categories Rep(A, D) for different C∗-algebras D. Definition 2.16. Let D1 and D2 be two C∗-algebras. A C∗-correspondence from D1 to D2 is a Hilbert D2-module with a representation of D1 by adjointable operators (representations of C∗-algebras are tacitly assumed nondegenerate). An isometric intertwiner between two correspondences from D1 to D2 is an isometric map on the underlying Hilbert D2-modules that intertwines the left D1-actions. Let Rep(D1, D2) denote the category of correspondences from D1 to D2 with isometric intertwiners as arrows and the usual composition. By Lemma 2.12, our two definitions of Rep(A, D) for unital ∗-algebras and C∗-algebras coincide if A is a unital C∗-algebra. So our notation is not ambiguous. There is no need to define representations of a non-unital ∗-algebra A because we may adjoin a unit formally. A representation of A extends uniquely to a representation of the unitisation A. Thus the nondegenerate representations of A are contained in Rep( A). To get rid of degenerate representations, we may require nondegeneracy on A when defining the integrable representations of A, compare Example 5.13. Let E be a Hilbert D1-module and F a correspondence from D1 to D2. The interior tensor product E ⊗D1 F is the (Hausdorff) completion of the algebraic tensor product E (cid:12) F to a Hilbert D2-module, using the inner product (2.17) see the discussion around [16, Proposition 4.5] for more details. We may use the balanced tensor product E (cid:12)D1 F instead of E (cid:12) F because the inner product (2.17) descends to this quotient. If we want to emphasise the left action ϕ: D1 → B(F) in the C∗-correspondence F, we write E ⊗ϕ F for E ⊗D1 F. In addition, let (E, π) be a closed representation of A on E. We are going to build a closed representation (E⊗D1 F, π ⊗D1 1) of A on E ⊗D1 F. First let X ⊆ E ⊗D1 F be the image of E (cid:12) F or E (cid:12)D1 F under the canonical map to E ⊗D1 F. Lemma 2.18. For a ∈ A, there is a unique linear operator π(a) ⊗ 1: X → X with (π(a) ⊗ 1)(ξ ⊗ η) = π(a)(ξ) ⊗ η for all ξ ∈ E, η ∈ F. The map a 7→ π(a) ⊗ 1 is a representation of A with domain X. hξ1 ⊗ η1, ξ2 ⊗ η2i = hη1,hξ1, ξ2iD1 · η2iD2, REPRESENTATIONS BY UNBOUNDED OPERATORS 9 Proof. Write ω, ζ ∈ X as images of elements of E (cid:12) F: αj ⊗ βj with ξi, αj ∈ E, ηi, βj ∈ F for 1 ≤ i ≤ n, 1 ≤ j ≤ m. Then ξi ⊗ ηi, ω = ζ = i=1 mX j=1 nX * ζ, nX i=1 + * mX j=1 (2.19) π(a)ξi ⊗ ηi = π(a∗)αj ⊗ βj, ω + . not depend on how we decomposed ω. Hence (π(a) ⊗ 1)ω := Pn An element ω0 ∈ E ⊗D1 F is determined uniquely by its inner products hζ, ω0i = 0 for all ζ ∈ X because X is dense in E ⊗D1 F. The right hand side in (2.19) does i=1 π(a)ξi ⊗ ηi well-defines an operator π(a) ⊗ 1: X → X. This is a right D2-module map, and a 7→ π(a) ⊗ 1 is linear and multiplicative because π is. Equation (2.19) says that hζ, (π(a)⊗1)ωi = h(π(a∗)⊗1)ζ, ωi for all ω, ζ ∈ X. Thus π⊗1 is a representation. (cid:3) Definition 2.20. Let (E ⊗D1 F, π ⊗D1 1) be the closure of the representation on E ⊗D1 F defined in Lemma 2.18. Lemma 2.21. Let I : E1 ,→ E2 be an isometric intertwiner between two repre- sentations (E1, π1) and (E2, π2), and let J : F1 ,→ F2 be an isometric intertwiner of C∗-correspondences. Then I ⊗D1 J : E1 ⊗D1 F1 ,→ E2 ⊗D1 F2 is an isometric intertwiner between (E1 ⊗D1 F1, π1 ⊗ 1) and (E2 ⊗D1 F2, π2 ⊗ 1). Proof. The isometry I ⊗D1 J maps the image X1 of E1 (cid:12) F1 to the image X2 of E2 (cid:12) F2 and intertwines the operators π1(a) ⊗ 1 on X1 and π2(a) ⊗ 1 on X2 for all a ∈ A. That is, it intertwines the representations defined in Lemma 2.18. It also (cid:3) intertwines their closures by Lemma 2.14. The lemma gives a bifunctor (2.22) The corresponding bifunctor ⊗D1 : Rep(A, D1) × Rep(D1, D2) → Rep(A, D2). ⊗D1 : Rep(B, D1) × Rep(D1, D2) → Rep(B, D2) for a C∗-algebra B is the usual composition of C∗-correspondences. This composition is associative up to canonical unitaries E ⊗D1 (F ⊗D2 G) ∼−→ (E ⊗D1 F) ⊗D2 G, (2.23) for all triples of composable C∗-correspondences. Lemma 2.24. If E carries a representation (E, π) of a ∗-algebra A, then the unitary in (2.23) is an intertwiner (E, π) ⊗D1 (F ⊗D2 G) ∼−→(cid:0)(E, π) ⊗D1 F(cid:1) ⊗D2 G. ξ ⊗ (η ⊗ ζ) 7→ (ξ ⊗ η) ⊗ ζ, (cid:0)(E, π)⊗D1 F(cid:1)⊗D2 G. The unitary in (2.23) intertwines between these cores. Hence Proof. The bilinear map from E × F to E ⊗D1 F is separately continuous with respect to the graph topologies on E and E ⊗D1 F and the norm topology on F. Since the image of F (cid:12) G in the Hilbert module F ⊗D2 G is dense in the norm topology, the image of E(cid:12)F (cid:12)G in E ⊗D1 (F ⊗D2 G) is a core for the representation (E, π) ⊗D1 (F ⊗D2 G). Since the image of E (cid:12) F in E ⊗D1 F is dense in the graph topology, the image of E(cid:12)F (cid:12)G in (E ⊗D1 F)⊗D2 G is a core for the representation it also intertwines between the resulting closed representations by Lemma 2.14. (cid:3) Definition 2.25. Let (E1, π1) and (E2, π2) be two representations of A on Hilbert D-modules E1 and E2. An adjointable operator x: E1 → E2 is an intertwiner if x(E1) ⊆ E2 and xπ1(a)ξ = π2(a)xξ for all a ∈ A, ξ ∈ E1. It is a ∗-intertwiner if both x and x∗ are intertwiners. 10 RALF MEYER that T Any adjointable intertwiner between two representations of a C∗-algebra B is a ∗-intertwiner. In contrast, for a general ∗-algebra, even the adjoint of a unitary intertwiner u fails to be an intertwiner if u(E1) (cid:40) E2. Example 2.26. Let t be a positive symmetric operator on a Hilbert space H. Assume n∈N dom tn is dense in H, so that t generates a representation π of the polynomial algebra C[x] on H. The Friedrichs extension of t is a positive self-adjoint operator t0 on H. It generates another representation π0 of C[x] on H. The identity map on H is a unitary intertwiner π ,→ π0. It is not a ∗-intertwiner unless t = t0. The following proposition characterises when an adjointable isometry I : E1 ,→ E between two representations on Hilbert D-modules is a ∗-intertwiner. Since E ∼= E1 ⊕ E⊥ 1 if I is adjointable, we may as well assume that I is the inclusion of a direct summand. Proposition 2.27. Let E1 and E2 be Hilbert modules over a C∗-algebra D and let (E1, π1) and (E, π) be representations of A on E1 and E1 ⊕ E2, respectively. The following are equivalent: (1) the canonical inclusion I : E1 ,→ E1 ⊕ E2 is a ∗-intertwiner from π1 to π; (2) the canonical inclusion I : E1 ,→ E1 ⊕ E2 is an intertwiner from π1 to π (3) there is a representation (E2, π2) on E2 such that π = π1 ⊕ π2. and E = E1 + (E ∩ E2); Proof. We view E1 and E2 as subspaces of E1 ⊕ E2, so we may drop the isometry I from our notation. The implication (3)⇒(1) is trivial. We are going to prove (1)⇒(2)⇒(3). First assume that I is a ∗-intertwiner. Then I is an intertwiner. In particular, E1 ⊆ E. Write ξ ∈ E as ξ = ξ1 + ξ2 with ξ1 ∈ E1, ξ2 ∈ E2. Since I∗ is an intertwiner, ξ1 = I∗(ξ) ∈ E1. Hence ξ2 = ξ − ξ1 ∈ E ∩ E2. Thus (1) implies (2). If (2) holds, then E1 ⊆ E is π-invariant and πE1 = π1 because I is an intertwiner. We claim that E2 := E ∩ E2 is π-invariant as well. Let ξ ∈ E2 and η ∈ E1. Then hη, π(a)ξi = hπ(a∗)η, ξi = hπ1(a∗)η, ξi = 0 because π1(a∗)η ∈ E1 is orthogonal to E2. Since E1 is dense in E1, this implies π(a)ξ ∈ E⊥ 1 = E2, and this implies our claim. The condition (2) implies E = E1 ⊕ E2 as a vector space with E2 = E2 ∩ E because E1 ∩E2 = {0}. Then E2 is dense in E2 because E is dense in E1 ⊕E2. Thus (E2, πE2) is a representation of A on E2. And (E, π) is the direct sum of (E1, π1) and (E2, πE2) because E = E1 ⊕ E2 and π1 = πE1. Thus (2) implies (3). (cid:3) 3. Integrable representations and C∗-hulls From now on, we tacitly assume representations to be closed. Proposition 2.5 shows that this is no serious loss of generality. Let A be a unital ∗-algebra. We assume that a class of "integrable" (closed) representations of A on Hilbert modules is chosen. Let Repint(A, D) ⊆ Rep(A, D) be the full subcategory with integrable representations on Hilbert D-modules as objects. Being full means that the set of arrows between two integrable representations of A is still the set of all isometric intertwiners. We sometimes write Repint(A) and Rep(A) for the collection of all the categories Repint(A, D) and Rep(A, D) for all C∗-algebras D. A C∗-hull is a C∗-algebra B with natural isomorphisms Rep(B, D) ∼= Repint(A, D) for all C∗-algebras D. More precisely: Definition 3.1. A C∗-hull for the integrable representations of A is a C∗-algebra B with a family of bijections Φ = ΦE from the set of representations of B on E to the set of integrable representations of A on E for all Hilbert modules E over all C∗-algebras D with the following properties: REPRESENTATIONS BY UNBOUNDED OPERATORS 11 • compatibility with isometric intertwiners: an isometry E1 ,→ E2 (not neces- sarily adjointable) is an intertwiner between two representations 1 and 2 of B if and only if it is an intertwiner between Φ(1) and Φ(2); • compatibility with interior tensor products: if F is a correspondence from D1 to D2, E is a Hilbert D1-module, and  is a representation of B on E, then Φ( ⊗D1 1F) = Φ() ⊗D1 1F as representations of A on E ⊗D1 F. The compatibility with isometric intertwiners means that the bijections Φ for all E with fixed D form an isomorphism of categories Rep(B, D) ∼= Repint(A, D) which, in addition, does not change the underlying Hilbert D-modules. The compatibility with interior tensor products expresses that these isomorphisms of categories for different D are natural with respect to C∗-correspondences. Definition 3.2. A weak C∗-hull for the integrable representations of A is a C∗-algebra B with a family of bijections Φ between representations of B and integrable representations of A on Hilbert modules that is compatible with unitary ∗-intertwiners and interior tensor products. Much of the general theory also works for weak C∗-hulls. But the Induction operator (cid:0) 0 x0 0 (cid:1), we may assume without loss of generality that E1 = E2 = E, Theorem 9.26 fails for weak C∗-hulls, as shown by a counterexample in §9.6. Proposition 3.3. Let a class of integrable representations of A have a weak C∗-hull B. Let (E1, π1) and (E2, π2) be integrable representations of A on Hil- bert D-modules E1 and E2, and let i be the corresponding representations of B on Ei for i = 1, 2. An adjointable operator x: E1 → E2 is a ∗-intertwiner from (E1, π1) to (E2, π2) if and only if it is an intertwiner from 1 to 2. Proof. Working with the direct sum representations on E1 ⊕ E2 and the adjointable E1 = E2 = E, π1 = π2 = π, and 1 = 2 = . The adjointable intertwiners for the representation  of B form a C∗-algebra B0: the commutant of B in B(E). We claim that the ∗-intertwiners for the representation π of A also form a C∗-algebra A0. Intertwiners and hence ∗-intertwiners form an algebra. Thus A0 is a ∗-algebra. We show that it is closed. Let (xi)i∈N be a sequence of adjointable intertwiners for (E, π) that converges in norm to x ∈ B(E). Let ξ ∈ E. Then xi(ξ) ∈ E because each xi is an intertwiner. Since π(a)(xiξ) = xiπ(a)ξ is norm-convergent for each a ∈ A, the sequence xi(ξ) is a Cauchy sequence for the graph topology on E. Since representations are tacitly assumed to be closed, this Cauchy sequence converges in E, so that x(E) ⊆ E. Moreover, x(π(a)ξ) = π(a)x(ξ) for all a ∈ A, ξ ∈ E, so x is again an intertwiner. Thus the algebra of intertwiners is norm-closed. This implies that A0 is a C∗-algebra. Since the family of bijections Repint(A) ∼= Rep(B) is compatible with unitary ∗-intertwiners, a unitary operator on E is a ∗-intertwiner for A if and only if it is an intertwiner for B. That is, the unital C∗-subalgebras A0, B0 ⊆ B(E) contain the same unitaries. A unital C∗-algebra is the linear span of its unitaries because any self-adjoint element t of norm at most 1 may be written as t = 1/2(cid:0)t + ip1 − t2(cid:1) + 1/2(cid:0)t − ip1 − t2(cid:1) 1 − t2 are unitary. Thus A0 = B0. This is what we had to prove. √ and t ± i (cid:3) Corollary 3.4. Let Repint(A) have a weak C∗-hull B. Direct sums and sum- mands of integrable representations remain integrable, and the family of bijec- tions Repint(A) ∼= Rep(B) preserves direct sums. Proof. Let π1, π2 be representations of A on Hilbert D-modules E1,E2. Let Si : Ei ,→ E1 ⊕E2 for i = 1, 2 be the inclusion maps. First we assume that π1, π2 are integrable. RALF MEYER i to π1 ⊕ π2 by Proposition 3.3. This implies π0 12 Let i be the representation of B on Ei corresponding to πi, and let π be the integrable representation of A on E1 ⊕E2 corresponding to the representation 1 ⊕ 2 of B. The isometries Si are intertwiners from i to 1 ⊕ 2. By Proposition 3.3, they are ∗-intertwiners from πi to π. Hence π = π1 ⊕ π2 by Proposition 2.27. Thus π1 ⊕ π2 is integrable and the family of bijections Repint(A) ∼= Rep(B) preserves direct sums. The same argument works for infinite direct sums. Now we assume instead that π1⊕π2 is integrable. Let  be the representation of B corresponding to π1⊕ π2. The orthogonal projection onto E1 is a ∗-intertwiner on the representation π1 ⊕ π2 by Proposition 2.27, and hence also on  by Proposition 3.3. Thus  = 1 ⊕ 2 for some representations i of B on Ei. Let π0 i be the integrable representation of A corresponding to i. The isometry Si is a ∗-intertwiner from i to 1 ⊕ 2 and hence from π0 i = πi, (cid:3) so that πi is integrable for i = 1, 2. Definition 3.5. Let B be a weak C∗-hull for A. The universal integrable represen- tation of A is the integrable representation (B, µ) of A on B that corresponds to the identity representation of B on itself. Proposition 3.6. Let B with a family of bijections Φ between representations of B and integrable representations of A on Hilbert modules be a weak C∗-hull for the integrable representations of A. Let (B, µ) be the universal integrable representation of A. Then Φ(E) ∼= (B, µ) ⊗B E for any C∗-correspondence E from B to D. (The proof makes this isomorphism more precise.) Proof. Let : B → B(E) be a representation of B on a Hilbert module E. Then u: B ⊗ E ∼−→ E, b ⊗ ξ 7→ (b)ξ, is a unitary ∗-intertwiner between the interior tensor product of the identity representation of B with E and the representation  on E. As Φ is compatible with interior tensor products and unitary ∗-intertwiners, u is a unitary ∗-intertwiner between (B, µ) ⊗B E and Φ(). Therefore, the image u(B(cid:12)E) = (B)E is a core for Φ(), and a ∈ A acts on this core by a 7→ u(µ(a)⊗1)u∗ or, explicitly, a · ((b)ξ) = (µ(a)b)ξ for all a ∈ A, b ∈ B, ξ ∈ E. (cid:3) Put in a nutshell, the whole isomorphism between integrable representations of A and representations of B is encoded in the single representation (B, µ) of A on B. This is similar to Schmüdgen's approach in [28]. In the following, we disregard the canonical unitary u in the proof of Proposition 3.6 and write Φ() = (B, µ) ⊗B E. A (weak) C∗-hull B does not solve the problem of describing the integrable representations of A. It only reduces it to the study of the representations of the C∗-algebra B. This reduction is useful because it gets rid of unbounded operators. If B is of type I, then any Hilbert space representation of B is a direct integral of irreducible representations, and irreducible representations may, in principle, be classified. Thus integrable Hilbert space representations of A are direct integrals of irreducible integrable representations, and the latter may, in principle, be classified. But if B is not of type I, then the integrable Hilbert space representations of A are exactly as complicated as the Hilbert space representations of B, and giving the C∗-algebra B may well be the best one can say about them. Proposition 3.7. A class of integrable representations has at most one weak C∗-hull. Proof. Let B1 and B2 be weak C∗-hulls for the same class of integrable repre- sentations of A. The identity map on B1, viewed as a representation of B1 on itself, corresponds first to an integrable representation of A on B1 and further to a representation of B2 on B1. This is a "morphism" from B2 to B1, that is, a nondegenerate ∗-homomorphism B2 → M(B1). Similarly, we get a morphism from B1 to B2. These morphisms B1 ↔ B2 are inverse to each other with respect to REPRESENTATIONS BY UNBOUNDED OPERATORS 13 the composition of morphisms because the maps they induce on representations of B1 and B2 on B1 and B2 are inverse to each other. An isomorphism in the category of morphisms is an isomorphism of C∗-algebras in the usual sense by [6, Proposition (cid:3) 2.10]. Now take any representation (B, µ) of A on B. When is this the universal integrable representation of a (weak) C∗-hull? Let D be a C∗-algebra and E a Hilbert D-module. For a representation : B → B(E), let Φ() = (B, µ)⊗ E be the induced representation of A on E as in the proof of Proposition 3.6. A representation of A is called B-integrable if it is in the image of Φ. Proposition 3.8. The C∗-algebra B is a weak C∗-hull for the B-integrable repre- sentations of A if and only if (1) if two representations 1, 2 : B ⇒ B(H) on the same Hilbert space H satisfy µ ⊗B 1 = µ ⊗B 2 as closed representations of A, then 1 = 2. It is a C∗-hull if and only if (1) and the following equivalent conditions hold: (2) Let (H, π) be a representation of A on a Hilbert space H and let (H0, π0) be a subrepresentation on a closed subspace H0 ⊆ H; that is, H0 ⊆ H and π0(a) = π(a)H0 for all a ∈ A. If both π0 and π are B-integrable, then H = H0 ⊕ (H ∩ H⊥ (3) Isometric intertwiners between B-integrable Hilbert space representations of A are ∗-intertwiners. 0 ) as vector spaces. (4) B-integrable subrepresentations of B-integrable Hilbert space representations of A are direct summands. The conditions (1)–(4) together are equivalent to (5) let : B → B(H) be a Hilbert space representation and let (H, π) be the associated representation of A on H. If (H0, πH0) is a B-integrable sub- representation of (H, π) on a closed subspace H0 ⊆ H, then the projection onto H0 commutes with (B). Proof. The map Φ is compatible with interior tensor products by Lemma 2.24. The condition (1) says that Φ is injective on Hilbert space representations. We claim that this implies injectivity also for representations on a Hilbert module E over a C∗-algebra D. Let 1, 2 be representations of B on E with µ ⊗B 1 = µ ⊗B 2. Let D → B(H) be a faithful representation. Then the representations 1⊗D1 and 2⊗D1 on the Hilbert space E⊗DH satisfy µ⊗B 1⊗D1 = µ⊗B 2⊗D1 by Lemma 2.24. Then condition (1) implies 1⊗D 1 = 2⊗D 1. Since the representation B(E) → B(E ⊗D H) is faithful, this implies 1 = 2. So Φ is injective also for representations on E. The image of Φ consists exactly of the B-integrable representations of A by definition. A unitary operator u ∗-intertwines two representations (E1, π1) and (E2, π2) of A if and only if π2 = uπ1u∗, where uπ1u∗ denotes the representation with domain u(E1) and (uπ1u∗)(a) = uπ1(a)u∗. Similarly, u intertwines two rep- resentations 1 and 2 of B if and only if 2 = u1u∗. Hence (1) implies that a unitary that ∗-intertwines two B-integrable representations of A also intertwines the corresponding representations of B. The converse is clear. So B is a weak C∗-hull for the B-integrable representations if and only if (1) holds. The equivalence between (2), (3) and (4) follows from Proposition 2.27 by writing H = H0 ⊕ H⊥ 0 . Assume that B is a C∗-hull. An isometric intertwiner for A is also one for B. Then it is a ∗-intertwiner for A and its range projection is an intertwiner for B by Proposition 3.3. Thus both (3) and (5) follow if B is a C∗-hull. Conversely, assume (1) and (3). We are going to prove that B is a C∗-hull for the B-integrable representations of A. We have already seen that B is a weak C∗-hull. We must check compatibility with isometric intertwiners. 14 RALF MEYER Let D be a C∗-algebra and let E1,E2 be Hilbert D-modules with representations 1, 2 of B. The corresponding representations (Ei, πi) of A for i = 1, 2 are the closures of the representations on i(B)Ei given by πi(a)(i(b)ξ) := i(µ(a)b)(ξ) for a ∈ A, b ∈ B, ξ ∈ Ei. Hence an isometric intertwiner for B is also one for A. Conversely, let I : E1 ,→ E2 be a Hilbert module isometry with I(E1) ⊆ E2 and π2(a)(Iξ) = I(π1(a)ξ) for all a ∈ A, ξ ∈ E1. We must prove 2(b)I = I1(b) for all b ∈ B. Let ϕ: D ,→ B(K) be a faithful representation on a Hilbert space K. Equip Hi := Ei ⊗ϕ K with the induced representations i of B and πi of A for i = 1, 2. Since the family of bijections Φ: Rep(B) ∼−→ Repint(A) is compatible with interior tensor products, it maps i to πi. The operator I induces an isometric intertwiner I from π1 to π2 by Lemma 2.21. Since π1 and π2 are B-integrable, we are in the situation of (3). So I is a ∗-intertwiner from π1 to π2. Thus I is an intertwiner from 1 to 2 by Proposition 3.3. That is, I 1(b) = 2(b)I for all b ∈ B. Equivalently, (I1(b)ξ) ⊗ η = (2(b)Iξ) ⊗ η in E2 ⊗ϕ H for all b ∈ B, ξ ∈ E, η ∈ H. Since the representation ϕ is faithful, this implies I1(b)ξ = 2(b)Iξ for all b, ξ, so that I1(b) = 2(b)I for all b, that is, I intertwines 1 and 2. Thus Φ is compatible with isometric intertwiners. Since (5) holds for C∗-hulls, we have proved along the way that (1) and (3) imply (5). It remains to show, conversely, that (5) implies (3) and (1). In the situation of (3), the projection P onto H0 commutes with B by (5). Thus the representation of B on H is a direct sum of representations on H0 and H⊥ 0 . This is inherited by the induced representation of A and its domain. So (5) implies (3). In the situation of (1), form the direct sum representation 1 ⊕ 2 on H ⊕ H and let H0 := {(ξ, ξ) ξ ∈ H}. The representation of A corresponding to 1 ⊕ 2 is µ ⊗1 H ⊕ µ ⊗2 H. Since µ ⊗1 H = µ ⊗2 H by assumption, the domain of µ⊗1H⊕µ⊗2H is H⊕H for some dense subspace H ⊆ H, and H0 := {(ξ, ξ) ξ ∈ H} is a dense subspace in H0 that is invariant for the representation µ⊗1 H ⊕ µ⊗2 H. The restricted representation on this subspace is B-integrable because it is unitarily equivalent to µ ⊗1 H = µ ⊗2 H. Therefore, the projection onto H0 commutes with (cid:3) the representation of B by (5). Thus 1 = 2. So (5) implies (1). The equivalent conditions (2)–(4) may be easier to check than (5) because they do not involve the C∗-hull. Corollary 3.9. Let A be a ∗-algebra and let Bi be C∗-algebras with represen- tations (Bi, µi) of A for i = 1, 2. Assume that for each Hilbert space H, the maps Φi : Rep(Bi,H) → Rep(A,H), i 7→ (Bi, µi) ⊗i H, are injective and have the same image. Then there is a unique isomorphism B1 ∼= B2 intertwining the representations (Bi, µi) of A for i = 1, 2. Hence a C∗-envelope as defined in [7] is unique if it exists. Proof. Both B1 and B2 are weak C∗-hulls for the same class of representations of A by Proposition 3.8. Proposition 3.7 gives the isomorphism B1 ∼= B2. (cid:3) Remark 3.10. The Hilbert space representations of a C∗-algebra only determine its bidual W∗-algebra, not the C∗-algebra itself. Hence it is remarkable that the condi- tions in Proposition 3.8 and Corollary 3.9 only need Hilbert space representations. For Corollary 3.9, this works because the bijection between the representations is of a particular form, induced by representations of A. The condition (1) in Proposition 3.8 is required in several other theories that associate a C∗-algebra to a ∗-algebra, such as the host algebras of Grundling [10,11], the C∗-envelopes of Dowerk and Savchuk [7], or the notion of a C∗-algebra generated REPRESENTATIONS BY UNBOUNDED OPERATORS 15 by affiliated multipliers by Woronowicz [31], see [31, Theorem 3.3] or the proof of Theorem 5.19 below. Definition 3.11. Let A be a ∗-algebra. A class of "integrable" representations of A on Hilbert modules over C∗-algebras is admissible if it satisfies the conditions (1)–(4) below, and weakly admissible if it satisfies (1)–(3). (1) If there is a unitary ∗-intertwiner from an integrable representation to another representation, then the latter is integrable. (2) If D and D0 are C∗-algebras, F is a correspondence from D to D0, and (E, π) is an integrable representation of A on a Hilbert D-module E, then the representation (E, π) ⊗D F on E ⊗D F is integrable. (3) Direct sums and summands of integrable representations are integrable. (4) Any integrable subrepresentation of an integrable representation of A on a Hilbert space is a direct summand. Lemma 3.12. Any class of integrable representations with a (weak) C∗-hull is (weakly) admissible. Proof. If there is a C∗-hull, Proposition 3.8 implies (4) in Definition 3.11. If there is a weak C∗-hull, then (1) and (2) in Definition 3.11 follow from the compatibility with unitary ∗-intertwiners and interior tensor products in the definition of a C∗-hull, (cid:3) and (3) follows from Corollary 3.4. Proposition 3.13. Let A be a unital ∗-algebra and let E be a Hilbert module over a C∗-algebra D. There is a natural bijection between the sets of representations of A on E and K(E). It preserves integrability if the class of integrable representations of A is weakly admissible or, in particular, if it has a weak C∗-hull. Proof. We may view E as an imprimitivity bimodule between K(E) and the ideal I in D that is spanned by the inner products hξ, ηiD for ξ, η ∈ E. Let E∗ be the inverse imprimitivity bimodule, which is a Hilbert module over K(E) with K(E∗) ∼= I. Then K(E) ∼= E ⊗D E∗ and E∗ ⊗K(E) E = I. If (π, E) is a representation of A on E, then (π, E)⊗DE∗ is a representation of A on E ⊗D E∗ = K(E). This maps Rep(A,E) to Rep(A, K(E)). If (, K) is a representation of A on K(E), then (, K) ⊗K(E) E is a representation of A on K(E) ⊗K(E) E ∼= E. This maps Rep(A, K(E)) to Rep(A,E). We claim that these two maps are inverse to each other. Both preserve integrability by (2) in Definition 3.11. The map Rep(A,E) → Rep(A, K(E)) → Rep(A,E) sends a representation (π, E) of A on E to the representation (π, E) ⊗D (E∗ ⊗K(E) E) = (π, E) ⊗D I of A on E ∼= E ⊗D I by Lemma 2.24. This is the restriction of π to E · I ⊆ E. Since E is also a Hilbert module over I, it is nondegenerate as a right I-module. Therefore, if (ui) is an approximate unit in I, then lim ξui = ξ for all ξ ∈ E. Then also lim π(a)ξui = π(a)ξ for all ξ ∈ E, a ∈ A, so lim ξui = ξ in the graph topology for all ξ ∈ E. Thus E · I = E, and we get the identity map on Rep(A,E). A similar, easier argument shows that we also get the identity map on Rep(A, K(E)). (cid:3) 4. Polynomials in one variable I Let A = C[x] with x = x∗. A (not necessarily closed) representation of A on a Hilbert D-module E is determined by a dense D-submodule E ⊆ E and a single symmetric operator π(x): E → E, that is, π(x) ⊆ π(x)∗. Then π(xn) = π(x)n. Lemma 4.1. The graph topology on E is generated by the increasing sequence of norms kξkn := khξ, (1 + π(x2n))ξik for n ∈ N. Proof. We must show that for any a ∈ C[x] there are C > 0 and n ∈ N with kξka ≤ Ckξkn. We choose n so that a has degree at most n. Then there is j j=1 RALF MEYER (cid:1) =P2n (cid:0)(x − λj)2 + µ2 k=1 b2 k. Thus kξka ≤ Ckξkn. Algebra. Then b =Qn 16 C > 0 so that C(1 + t2n) > 1 + a(t)2 for all t ∈ R. Thus the polynomial b := C(1 + x2n)− (1 + a∗a) is positive on R. So the zeros of b are complex and come in pairs λj ± iµj for j = 1, . . . , n with λj, µj ∈ R by the Fundamental Theorem of k, where each bk is a product of either x − λj or µj for j = 1, . . . , k, so bk = b∗ (cid:3) Thus the monomials {xn n ∈ N} form a strong generating set for C[x]. A representation of C[x] is determined by the closed operators π(xn) for n ∈ N by Proposition 2.9. In contrast, it is not yet determined by the single closed operator π(x) because {x} is not a strong generating set: Example 4.2. We construct a closed representation of C[x] on a Hilbert space with Let H0 := C∞(T) and let π0 : C[x] → End(H0) be the polynomial functional calculus dt. The graph topology generated by this representation of C[x] is for the operator i d the usual Fréchet topology on C∞(T). So the representation of C[x] on C∞(T) is closed. Now for some λ ∈ T, let π(x2) (cid:40)(cid:0)π(x)(cid:1)2. Let H := L2(T), viewed as the space of Z-periodic functions on R. H := {f ∈ C∞(T) f (n)(λ) = 0 for all n ≥ 1}. This is a closed, C[x]-invariant subspace in H0. Let π be the restriction of π0 to H0. This is also a closed representation of C[x]. Its domain H is dense in H0 in the graph norm of x, but not in the graph norm of x2. So π(x) = π0(x) and π(x2) (cid:40) π0(x2) =(cid:0)π(x)(cid:1)2. All notions of integrability for representations of C[x] that we shall consider imply π(xn) = π(x)n. Under this assumption, an integrable representation of C[x] is determined by the single closed operator π(x). Let B := C0(R). Let X be the identity function on R, viewed as an unbounded multiplier of B. We define a closed representation (B, µ) of A on C0(R) by (4.3) B := {f ∈ B ∀n: X n·f ∈ B} for f ∈ B, n ∈ N. Theorem 4.4. Let (E, π) be a representation of A = C[x] on a Hilbert module E over a C∗-algebra D. The following are equivalent: and µ(xn)f := X n·f (1) π = µ ⊗ 1E for a representation : B → B(E); (2) π(a) is regular and self-adjoint for each a ∈ Ah := {a ∈ A a = a∗}; (3) π(xn) is regular and self-adjoint for each n ∈ N; (4) π(x) is regular and self-adjoint and π(xn) = π(x)n for all n ∈ N; (5) π(x) is regular and self-adjoint and E =T∞ n=1 dom π(x)n. Call representations with these equivalent properties integrable. The C∗-algebra C0(R) is a C∗-hull for the integrable representations of A with (B, µ) as the universal integrable representation. Proof. If a ∈ Ah, then µ(a) is a self-adjoint, affiliated multiplier of B. Hence µ(a) ⊗D 1 is a regular, self-adjoint operator on B ⊗ E ∼= E for any representation  of B on E by [16, Proposition 9.10]. Thus (1) implies (2). The implication (2)⇒(3) is trivial. The operator π(xn) is always contained in the n-fold power π(x)n. The for C[x] by Lemma 4.1. Equation (2.10) gives E =T∞ latter is symmetric, and a proper suboperator of a symmetric operator cannot be self-adjoint. Thus (3) implies (4). The set {xn n ∈ N} is a strong generating set n=1 dom π(xn) for any (closed) representation. Thus (4) implies (5). Assume (5) and abbreviate t = π(x). The functional calculus for t is a nonde- generate ∗-homomorphism : C0(R) → B(E) (see [16, Theorem 10.9]). Let π0 be the representation µ ⊗ 1 of A on E associated to . We claim that π = π0. The REPRESENTATIONS BY UNBOUNDED OPERATORS 17 domain of π0 is T dom π0(xn) = E by condition (5) and (2.10). On this domain, functional calculus extends to affiliated multipliers and maps the identity function on R to the regular, self-adjoint operator t. This means that π0(x) = t. Then π0(xn) ⊆ tn. This implies π0(xn) = tn because π0(xn) is self-adjoint and tn is symmetric. Since the set {xn n ∈ N} is a strong generating set for C[x], the π(x) and π0(x) act by the same operator because they have the same closure. Thus π = π0 and (5) implies (1). So all five conditions in the theorem are equivalent. To show that B is a C∗-hull for the class of representations described in (1), we check (5) in Proposition 3.8. An integrable representation of A on a Hilbert space H corresponds to a self-adjoint operator t on H by (5). An integrable subrepresentation is a closed subspace H0 of H with a self-adjoint operator t0 on H0 whose graph is contained in that of t. Since t0 is self-adjoint, the subspaces (t0 ± i)(dom(t0)) = (t ± i)(dom(t0)) are equal to H0. The Cayley transform u of t maps (t+i)(dom(t0)) onto (t−i)(dom(t0)). Thus it maps H0 onto itself. Since u−1 generates the image of B = C0(R) under the functional calculus, the projection onto H0 is B-invariant. (cid:3) Example 4.5. Regularity and self-adjointness are independent properties of a sym- metric operator. Examples of regular symmetric operators that are not self-adjoint are easy to find, see §6. We are going to construct a representation π of C[x] on a Hilbert module for which π(a) is self-adjoint for each a ∈ C[x] with a = a∗, but π(x) is not regular. We follow the example after Théorème 1.3 in [20], which Pierrot attributes to Hilsum. dx on H with the following domains. For T1, we take 1-periodic smooth functions; for T2, we take the restrictions to [0, 1] of smooth functions on R satisfying f(x + 1) = −f(x). Both T1 and T2 are essentially self-adjoint. Let D := C([−1, 1]) and E := C([−1, 1],H). Let E ⊆ E be the dense subspace of all functions f : [−1, 1]×[0, 1] → C such that ∂n Let H be the Hilbert space L2([0, 1]) and let T1 and T2 be the operators i d ∂nx f(t, x) is continuous for each n ∈ N, f(t, 1) = sign(t) · ∂n ∂nx ∂n ∂nx (4.6) for all t ∈ [−1, 1], x ∈ R, t 6= 0, and f(t, 0) ∂n ∂nx f(0, 0) = ∂n ∂nx f(0, 1) = 0. (4.7) Equivalently, f(t, ॷ) belongs to the domain of T1n = T n1 for all n ∈ N, t ≤ 0 and to the domain of T2n = T n2 for all n ∈ N, t ≥ 0; indeed, this forces ∂n ∂nx f to be continuous on [−1, 1] × [0, 1] and to satisfy the boundary conditions (4.6). These imply (4.7) by continuity. Let xn ∈ C[x] act on E by . This defines a closed (cid:19)n (cid:18) ∗-representation of C[x] on E with E =T i d dx n∈N dom π(x)n. and T2 are self-adjoint and E =T The closure π(x) is the irregular self-adjoint operator described in [20]. Let a ∈ C[x] with a = a∗. Then (a) is (regular and) self-adjoint for any integrable representation  of C[x] by Theorem 4.4. Therefore, the restriction of π(a) to a single fibre of E at some t ∈ [−1, 1] \ {0} is a self-adjoint operator on L2([0, 1]) because T1 n∈N dom π(x)n. The restriction of π(a)∗ at t = 0 is contained in the self-adjoint operators a(T1) and a(T2) by continuity. We claim that a(T1) ∩ a(T2) = π(a)t=0. This claim implies that π(a)∗ is contained in π(a), that is, π(a) is self-adjoint. Let a ∈ C[x] have degree n. Then the graph norms for a and xn are equivalent in i have the same any representation by the proof of Lemma 4.1. Hence a(Ti) and T n 18 RALF MEYER i consists of functions [0, 1] → C whose nth derivative domain. The domain of T n lies in L2 and whose derivatives of order strictly less than n satisfy the boundary condition for Ti. Hence the domain of T n1 ∩ T n2 consists of those functions [0, 1] → C whose nth derivative lies in L2 and whose derivatives of order strictly less than n vanish at the boundary points 0 and 1. This is exactly the domain of the closure of (T1 ∩ T2)n = π(xn)t=0. On this domain the operators a(T1) ∩ a(T2) and π(a)t=0 both act by the differential operator a(i d dx). The algebra A = C[x] has many Hilbert space representations coming from closed symmetric operators that are not self-adjoint. There is, however, no larger admissible class of integrable representations: Proposition 4.8. Assume that an admissible class of integrable representations of A = C[x] contains all representations coming from self-adjoint Hilbert space operators. Then any integrable representation of A on a Hilbert module comes from a regular, self-adjoint operator. Proof. We first prove that there can be no more integrable Hilbert space represen- tations than those coming from self-adjoint operators. Let (H, π) be an integrable representation on a Hilbert space H. We may extend the closed symmetric operator t := π(x) on H to a self-adjoint operator t2 on a larger Hilbert space H2. This gives a representation π2 of A on H2 as in Theorem 4.4, which is integrable by assumption. The inclusion map H ,→ H2 is an isometric intertwiner from π to π2. Hence π is a direct summand of π2 by (4) in Definition 3.11. Thus π(xn) is self-adjoint for each n ∈ N, and π is the representation induced by t. Now let (E, π) be an integrable representation of A on a Hilbert D-module E. For any Hilbert space representation ϕ: D → B(H), the induced representation of A on the Hilbert space E ⊗ϕ H is also integrable by (2) in Definition 3.11. Thus π(xn) ⊗ϕ 1H is self-adjoint for any Hilbert space representation ϕ: D → B(H). A closed, densely defined, symmetric operator T on a Hilbert D-module E is self-adjoint and regular if and only if, for any state ω on D, the closure of T ⊗D 1 on the Hilbert spaces E ⊗D Hω is self-adjoint; here Hω means the GNS-representation for ω. This is called the Local–Global Principle by Kaad and Lesch ([14, Theorem 1.1]); the result was first proved by Pierrot ([20, Théorème 1.18]). We will take up Local–Global Principles more systematically in §5. Thus π(xn) is regular and self- adjoint for each n ∈ N. So π is obtained from the regular self-adjoint operator π(x) (cid:3) as in Theorem 4.4. Example 4.9. There are many admissible classes of representations of C[x] that are smaller than the class in Theorem 4.4. There are even many such classes that contain the same Hilbert space representations. For instance, let B := C0((−∞, 0))⊕ C0([0,∞)) with the representation of polynomials by pointwise multiplication. This is a C∗-hull for a class of representations of C[x] by Theorem 8.2 below. Since the standard topologies on R and (−∞, 0) t [0,∞) have the same Borel sets, both C∗-hulls C0((−∞, 0))⊕C0([0,∞)) and C0(R) give the same integrable Hilbert space representations because of the Borel functional calculus. But there are regular, self- adjoint operators on Hilbert modules that do not give a B-integrable representation. The obvious example is the multiplier X of C0(R) that generates the universal integrable representation of C[x]. Can there be an admissible class of representations of C[x] that contains some representation on a Hilbert space that does not come from a self-adjoint operator? We cannot rule this out completely. But such a class would have to be rather strange. By Proposition 4.8, it cannot contain all self-adjoint operators. By Example 2.26, it cannot contain all representations coming from positive symmetric operators REPRESENTATIONS BY UNBOUNDED OPERATORS 19 indices (0, n) for some n ∈ [1,∞]. Then dom∞(t) :=T∞ because then there would be isometric intertwiners among integrable representations that are not ∗-intertwiners. The following example rules out symmetric operators with one deficiency index 0: Example 4.10. Let t be a closed symmetric operator on a Hilbert space H of deficiency n=1 dom(tn) is a core for each power tk by [27, Proposition 1.6.1]. Thus there is a closed representation π of C[x] with domain dom∞(t) and π(xk) = tk for all k ∈ N. By assumption, the operator t + i is surjective, but t − i is not. That is, the Cayley transform c := (t − i)(t + i)−1 is a non-unitary isometry. The operator t may be reconstructed from c as in [16, Equation (10.11)]. Here c∗ is surjective, so this simplifies to dom(t) = (1 − c)c∗H = (1 − c)H, and t(1 − c)ξ = i(1 + c)ξ for all ξ ∈ H. Thus c(dom t) ⊆ dom t and ct ⊆ tc because ct(cid:0)(1 − c)ξ(cid:1) = ic(1 + c)ξ = i(1 + c)(cξ) = t(1 − c)(cξ) = (tc)(cid:0)(1 − c)ξ(cid:1). Then ctn ⊆ tnc for all n ∈ N. Thus c is an isometric intertwiner from π to itself by Proposition 2.15. If c∗ were an intertwiner as well, then c∗(dom t) ⊆ dom t and c∗(t ± i)ξ = (t ± i)c∗ξ for all ξ ∈ dom(t). So c∗c(t + i)ξ = c∗(t − i)ξ = (t − i)c∗ξ = c(t + i)c∗ξ = cc∗(t + i)ξ. This is impossible because c∗c 6= cc∗ and t + i is surjective. So the isometry c is an intertwiner, but not a ∗-intertwiner. This is forbidden for admissible classes of integrable representations. If t has deficiency indices (n, 0) instead, then −t has deficiency indices (0, n) and its Cayley transform is an isometric intertwiner that is not a ∗-intertwiner by the argument above. 5. Local–Global principles Definition 5.1. Let A be a ∗-algebra with a weakly admissible class of integrable representations (Definition 3.11). The Local–Global Principle says that a representation π of A on a Hilbert D-module E is integrable if (and only if) the representations π ⊗ 1 are integrable for all Hilbert space representations : D → B(H). The Strong Local–Global Principle says that a representation π of A on a Hilbert D-module E is integrable if (and only if) the representations π ⊗ 1 are integrable for all irreducible Hilbert space representations : D → B(H). Roughly speaking, the Local–Global Principle says that the class of integrable representations on Hilbert modules is determined by the class of integrable rep- resentations on Hilbert spaces. Examples where the Local–Global Principle fails are constructed in §6 and §8. We do not know an example with the Local–Global Principle for which the Strong Local–Global Principle fails. An irreducible representation : D → B(H) is unitarily equivalent to the GNS- representation for a pure state ψ on D. The tensor product E ⊗ H is canonically isomorphic to the completion Eψ of E to a Hilbert space for the scalar-valued inner product hx, yiC := ψ(hx, yiD). The induced representation π ⊗D 1 of A on Hψ is the closure of the representation π with domain E ⊆ E ⊆ Eψ. Any representation : D → B(H) is a direct sum of cyclic representations, and these are GNS-representations of states. Since any weakly admissible class of integrable representations is closed under direct sums, the Local–Global Principle holds if and only if integrability of π ⊗ 1 for all GNS-representations  of states on D implies integrability of π. 20 RALF MEYER Example 5.2. Define integrable representations of the polynomial algebra C[x] as in Theorem 4.4. Thus they correspond to regular, self-adjoint operators on Hilbert modules. The main result in [14] says that the integrable representations of C[x] satisfy the Local–Global Principle. This is where our notation comes from. We already used this to prove Proposition 4.8. The Strong Local–Global Principle for integrable representations of C[x] is only conjectured in [14]. This conjecture had already been proved by Pierrot in [20, Théorème 1.18] before [14] was written. It is based on the following Hahn–Banach type theorem for Hilbert submodules: Theorem 5.3 ([20, Proposition 1.16]). Let D be a C∗-algebra and let E be a Hilbert D-module. Let F (cid:40) E be a proper, closed Hilbert submodule. There is an irreducible Hilbert space representation : D → B(H) with F ⊗ H (cid:40) E ⊗ H. Corollary 5.4 ([20, Corollaire 1.17]). Let E be a Hilbert module over a C∗-algebra D. Let F1,F2 (cid:40) E be two closed Hilbert submodules. If F1 6= F2, then there is an irreducible Hilbert space representation : D → B(H) with F1 ⊗ H 6= F2 ⊗ H as closed subspaces in E ⊗ H. Corollary 5.5 ([20, Théorème 1.18]). Let T be a closed, semiregular operator on a Hilbert D-module E. The operator T is regular if and only if, for each irreducible representation : D → B(H) on a Hilbert space H, the closures of T ⊗1 and T ∗⊗1 on E ⊗ H are adjoints of each other. Hence T is regular and self-adjoint if and only if T ⊗ 1 is a self-adjoint operator on E ⊗ H for each irreducible Hilbert space representation : D → B(H). We now apply the above results of Pierrot. First we deduce a criterion for representations to be equal. Then we prove that certain definitions of integrability automatically satisfy the Strong Local–Global Principle. Theorem 5.6. Let A be a ∗-algebra and let πi for i = 1, 2 be (closed) representations of A on a Hilbert module E over a C∗-algebra D. The following are equivalent: (1) π1 = π2; (2) π1 ⊗ H = π2 ⊗ H for each irreducible Hilbert space representation  of D; (3) π1(a) = π2(a) for each a ∈ A. Proof. The equivalence (3) ⇐⇒ (1) is Proposition 2.9, and (1) clearly implies (2). Thus we only have to prove that not (3) implies not (2). Assume that there is a ∈ A with π1(a) 6= π2(a). The graphs Γ1 and Γ2 of π1(a) and π2(a) are different Hilbert submodules of E ⊕ E. Corollary 5.4 gives an irreducible representation  of D with Γ1 ⊗ H 6= Γ2 ⊗ H. This says that π1(a) ⊗ 1H 6= π2(a) ⊗ 1H because Γi ⊗ H is the graph of πi(a) ⊗ 1H. (cid:3) How do we specify which representations π of a ∗-algebra A are integrable? There are two basically different ways. The "universal way" specifies the universal integrable representation. That is, it starts with a representation (B, µ) on a C∗-algebra B that satisfies (1) in Proposition 3.8 and takes the class of B-integrable representations. The "operator way" imposes conditions on the operators π(a), such as regularity and self-adjointness of π(a) or strong commutation relations. In good cases, the same class of integrable representations may be specified in both ways. For instance, Theorem 4.4 shows that several classes of representations of C[x] are equal. The first is defined by the universal representation on C0(R). The second asks π(a) to be regular and self-adjoint for all a ∈ Ah. We are going to make the "operator way" more precise so that all classes of representations defined in this way satisfy the Strong Local–Global Principle. This is a powerful method to prove Local–Global Principles. REPRESENTATIONS BY UNBOUNDED OPERATORS 21 Definition 5.7. Let A be a ∗-algebra and Rep0(A) some weakly admissible class of representations of A on Hilbert modules over C∗-algebras. A natural construction of Hilbert submodules (of rank n ∈ N≥1) associates to each representation π on a Hilbert module E that belongs to Rep0(A) a Hilbert submodule F(π) ⊆ E n, such that ∼−→ E2 is a unitary ∗-intertwiner between two representations π1 (1) if u: E1 and π2 in Rep0(A), then u⊕n : E n1 → E n2 maps F(π1) onto F(π2); (2) let D1 and D2 be C∗-algebras and let G be a D1, D2-correspondence; let π be a representation in Rep0(A) on a Hilbert D1-module E; then the canonical isomorphism E n ⊗D1 G ∼−→ (E ⊗D1 G)n maps F(π) ⊗D1 G onto F(π ⊗D1 G); (3) if πi for i in a set I are representations in Rep0(A) on Hilbert D-modules Ei over the same C∗-algebra D, then the canonical isomorphism(cid:0)LEi i maps F(cid:0)L πi LE n (cid:1) ontoLF(πi). (cid:1)n ∼−→ In brief, F(π) ⊆ E n is compatible with unitary ∗-interwiners, interior tensor products, and direct sums. A smaller class of representations Rep00(A) ⊆ Rep0(A) is defined by a submodule condition if there are two natural constructions of Hilbert submodules Fi(π), i = 1, 2, of the same rank n, such that a representation π in Rep0(A) belongs to Rep00(A) if and only if F1(π) = F2(π). A class of representations Repint(A) ⊆ Rep(A) is defined by submodule conditions if it is defined by transfinite recursion by repeating the step in the previous paragraph. More precisely, there are a well-ordered set I with a greatest element M and least element 0 and subclasses Repi(A) ⊆ Rep(A) for i ∈ I such that (1) Rep0(A) = Rep(A) and RepM(A) = Repint(A); (2) Repi+1(A) ⊆ Repi(A) is defined by a submodule condition for each i ∈ I; i0<i Repi0(A) if i 6= 0 and i 6= i0 + 1 for all i0 ∈ I. (3) Repi(A) =T The following lemma makes this definition meaningful, the following theorem (Repi(A))i∈I is a set of weakly admissible subclasses, then T makes it interesting. Lemma 5.8. If Rep0(A) ⊆ Rep(A) is weakly admissible and Rep00(A) ⊆ Rep0(A) is defined by a submodule condition, then Rep00(A) is also weakly admissible. If i∈I Repi(A) is weakly admissible. Any class of representations defined by submodule conditions is weakly admissible. Theorem 5.9. If Repint(A) ⊆ Rep(A) is defined by submodule conditions, then it satisfies the Strong Local–Global Principle. Before we prove these two results, we give examples of classes of representations defined by one or more submodule conditions, and a few counterexamples. These show that a class of integrable representations defined in the operator way is often but not always defined by submodule conditions. Example 5.10. The regularity condition for a ∈ Ah requires π(a) to be regular and self-adjoint. Equivalently, the closures of (π(a) ± i)(E) for both signs are dense in E; this is equivalent to π(a) having a unitary Cayley transform. Sending π to the image of π(a) + i or π(a) − i is a natural construction of a Hilbert submodule. Hence the condition that π(a) is regular and self-adjoint is equivalent to the combination of two submodule conditions of rank 1. Alternatively, we may proceed as in the definition of regularity for non-self-adjoint operators. Let Γ(T) denote the closure of the graph of an operator T. A closed operator T is regular if and only if the direct sum of Γ(T) and U0(Γ(T ∗)) is E ⊕ E, 22 where U0(ξ1, ξ2) := (ξ2,−ξ1). If a ∈ Ah, then regularity and self-adjointness of π(a) together are equivalent to the equality of RALF MEYER F1(π) := Γ(π(a)) ⊕ U0(Γ(π(a∗))) and F2(π) := E ⊕ E. We claim that F1 and F2 are natural constructions of Hilbert submodules of rank 2. This is trivial for F2. That F1 is compatible with unitary intertwiners and direct sums is an easy exercise. The construction F1 is compatible with interior tensor products because the graph of (π ⊗D1 1G)(a) is Γ(π(a)) ⊗D1 G. For instance, (2) in Theorem 4.4 defines integrable representations of C[x] by regularity conditions. We generalise this in Theorem 5.17 below. Example 5.11. The class of representations where π(a) is regular for some a ∈ A is always weakly admissible by [16, Proposition 9.10]. The first example in §6 shows a class of representations defined by such a condition that does not satisfy the Local–Global Principle, in contrast to Theorem 5.9. Hence asking for π(a) to be regular for some a ∈ A cannot be a submodule condition. The problem is that the inclusion Γ(π(a)∗) ⊗D 1G ⊆ Γ(cid:0)(π(a) ⊗D 1G)∗(cid:1) for a correspondence G may be strict. Example 5.12. Let a1, a2 ∈ Ah and suppose that t1 := π(a1) and t2 := π(a2) are self-adjoint, regular operators for all representations in Rep0(A); we may achieve this by submodule conditions as in Example 5.10 in previous steps of a recursive definition. We say that t1 and t2 strongly commute if their Cayley transforms u1 and u2 commute. Equivalently, u1 commutes with t2, that is, u1t2u∗ 1 = t2. The graphs of t2 and u1t2u∗ 1 are natural constructions of Hilbert submodules of rank 2. Therefore, strong commutation of π(a1) and π(a2) is a submodule condition. Example 5.13. Let I / A be an ideal. A nondegeneracy condition for I asks the closed linear span of π(I)E to be all of E; here E is the domain of π. This means that F1(π), the closed linear span of π(a)ξ for a ∈ I, ξ ∈ E, is equal to F2(π) = E. These are natural constructions of Hilbert submodules of rank 1. So a nondegeneracy condition is a submodule condition. For instance, let I be a non-unital ∗-algebra and let A = I be its unitisation. Any representation of I extends uniquely to a unital representation of A. The class of nondegenerate representations of I inside the class of all representations of A is defined by a submodule condition. More generally, let V1, V2 ⊆ A be vector subspaces and ask the closed linear spans of π(a)ξ for a ∈ Vj, ξ ∈ E to be equal for j = 1, 2. This is a submodule condition as well. For instance, the condition π(a + i)E = E for a ∈ Ah is of this form. It holds if and only if the Cayley transform of π(a) is an isometry (possibly without adjoint). Often we need a mild generalisation of the above construction, see Example 5.14 be- low. Suppose that we have constructed a representation ϕ(π) of a unital ∗-algebra A0 on E for any representation π in Rep0(A), such that π 7→ ϕ(π) is compatible with unitary ∗-intertwiners, direct sums, and interior tensor products; the last property means that ϕ(π ⊗D1 1G) = ϕ(π) ⊗D1 1G as representations on E ⊗D1 G. Then we may ask the nondegeneracy condition for an ideal in A0 instead. In particular, A0 may be a weak C∗-hull for some class of representations containing Rep0(A). Example 5.14. Let a1, . . . , an ∈ Ah be commuting, symmetric elements and suppose that π(aj) for j = 1, . . . , n are strongly commuting, self-adjoint, regular operators for all representations in Rep0(A); we may achieve all this by previous submodule conditions as in Examples 5.10 and 5.12. A closed spectral condition asks the joint spectrum of π(a1), . . . , π(an) to be contained in a closed subset X ⊆ Rn. We claim that this is a submodule condition. Under our assumptions, the functional calculus Φ: C0(Rn) → B(E) exists. Our spectral condition means that REPRESENTATIONS BY UNBOUNDED OPERATORS 23 Φ(C0(Rn \ X))E = 0. The construction of Φ is clearly compatible with unitary ∗-intertwiners and direct sums. It is also compatible with interior tensor products, that is, the functional calculus for π ⊗ 1(a1), . . . , π ⊗ 1(an) maps f 7→ Φ(f) ⊗ 1. Hence Φ(C0(Rn \ X))E is a naturally constructed Hilbert submodule of E. So our spectral condition for closed X ⊆ Rn is a submodule condition. More generally, let X ⊆ Rn be locally closed, that is, X is relatively open in its closure X. Suppose that the spectral condition for X holds for all representations in Rep0(A), say, by previous recursion steps. Then the functional calculus homomor- phism for π(a1), . . . , π(an) exists and descends to C0(X). The spectral condition for X asks the restriction of this homomorphism to the ideal C0(X) / C0(X) to be nondegenerate. This is a submodule condition by Example 5.13. Example 5.15 (see [30, §3]). Let Aµ for some µ ∈ R \ {0} be the unital ∗-algebra generated by two elements v, n with the relations v∗v = vv∗ = 1, n∗n = nn∗, v∗nv = µn. This is the algebra of polynomial functions on the quantum group Eµ(2). The relations allow to write any element as a linear combination of vk · g(n, n∗) for k ∈ Z and a polynomial g. It follows that the graph topology of a representation of Aµ is generated by the graph norms of (n∗n)k for k ∈ N. Thus a representation k=0 π((n∗n)k), compare the proof of (2.10). The C∗-algebra of Eµ(2) is a C∗-hull for a certain class of integrable representations of Aµ that is defined by submodule conditions. First, we require π(n) to be a regular, normal operator; equivalently, π(n + n∗) and −iπ(n − n∗) are regular and self- adjoint, and they strongly commute; these are submodule conditions by Examples 5.10 and 5.12. Secondly, we require the spectrum of π(n) (or the joint spectrum of Z} ∪ {0}; this its real and imaginary part) to be contained in Xµ := {z ∈ C z ∈ µ is a submodule condition by Example 5.14. Finally, we require π(n∗n)k to be regular and self-adjoint for all k ≥ 1. These are submodule conditions by Example 5.10. We claim that an integrable representation on E is equivalent to a pair (V, N) consisting of a unitary operator V and a regular, normal operator N on E with spectrum contained in Xµ, subject to the relation V ∗N V = µN. First, any such k=0 dom(N k). Conversely, if π is an integrable representation, then let N := π(n), V := π(v). These have the properties required above. Since π((n∗n)k) is self-adjoint and contained in the symmetric operator (N∗N)k, we must have π((n∗n)k) = (N∗N)k. So the pair (V, N) gives an integrable representation of Aµ with domainT∞ domain of the representation of Aµ isT∞ is closed if and only if its domain isT∞ k=0 dom(N k). The regular, normal operator N with spectrum in Xµ defines a functional calculus  on C0(Xµ). The commutation relation v∗nv = µn is equivalent to V ∗(f)V = (α(f)) for the automorphism α(f)(x) := f(µx) on C0(Xµ). As a consequence, the crossed product C∗-algebra C0(Xµ) (cid:111)α Z is a C∗-hull for our class of integrable representations. By the way, this also follows from our Induction Theorem. For this, we give Aµ the unique Z-grading where v has degree 1 and n has degree 0. Then (Aµ)0 = C[n, n∗], and we call a representation of C[n, n∗] integrable if n is regular and normal with spectrum contained in Xµ. The C∗-hull for this class of integrable representations of C[n, n∗] is C0(Xµ). In this case, all representations of C[n, n∗] are inducible to Aµ, and the induced C∗-hull for Aµ is C0(Xµ) (cid:111)α Z. Interesting classes of representations defined by submodule conditions occur in Theorems 5.21 and 8.6. The examples in [7, 26] are also defined by submodule conditions, compare Proposition 9.4. Example 5.16. If the algebra A carries a topology, then we may restrict attention to representations of A that are continuous in some sense. For instance, if G is a 24 RALF MEYER topological group and A = C[G] is the group ring of the underlying discrete group, then representations of A are unitary representations of G, possibly discontinuous. Among them, we may restrict to the continuous representations (compare the definition of a host algebra for G in [11]). If G is an infinite-dimensional Lie group, we may restrict further to representations of C[G] that are smooth in the sense that the smooth vectors are dense. I do not expect continuity or smoothness to be a submodule condition, and I do not know when the classes of continuous or smooth representations satisfy the Local–Global Principle or its strong variant. F1(π1) = F2(π1), then F1(π2) = u⊕n(cid:0)F1(π1)(cid:1) = u⊕n(cid:0)F2(π1)(cid:1) = F2(π2). Thus π2 Semiboundedness conditions ask for certain (regular) self-adjoint operators to be bounded above, see [18]. If we specify the upper bound on the spectrum, this is a spectral condition as in Example 5.14. When we let the upper bound go to ∞, however, then direct sums no longer preserve semiboundedness. Therefore, semiboundedness conditions seem close enough to submodule conditions to be tractable, but the details require further thought. Proof of Lemma 5.8. Let F1 and F2 be natural constructions of Hilbert submodules of rank n that define Rep00(A) inside Rep0(A), and let Rep0(A) be weakly admissible. Let πi for i = 1, 2 be representations on Hilbert D-modules Ei for a C∗-algebra D ∼−→ E2 be a unitary ∗-intertwiner from π1 to π2. If that belong to Rep0(A). Let u: E1 belongs to Rep00(A) if π1 does. This verifies (1) in Definition 3.11 using (1) in Definition 5.7. Similarly, (2) and (3) in Definition 5.7 show that Rep00(A) inherits (2) and (3) in Definition 3.11 from Rep0(A). Thus Rep00(A) is again weakly admissible. It is trivial that weak admissibility is hereditary for intersections. By transfinite induction, it follows that any class of representations defined by submodule conditions (cid:3) is weakly admissible. Proof of Theorem 5.9. Let F1 and F2 be natural constructions of Hilbert sub- modules of rank n that define Rep00(A) inside Rep0(A), and assume that Rep0(A) satisfies the Strong Local–Global Principle. Let π be a representation on a Hilbert D-module E that does not belong to Rep00(A). We must find an irreducible represen- tation  of D on a Hilbert space H such that π ⊗ H does not belong to Rep00(A). If the representation does not even belong to Rep0(A), this is possible because Rep0(A) satisfies the Strong Local–Global Principle by assumption. So we may assume that π belongs to Rep0(A) but not to Rep00(A). Thus F1(π) and F2(π) are well defined and different Hilbert submodules of E n. Corollary 5.4 gives an irreducible representation  of D on a Hilbert space H such that F1(π) ⊗ H 6= F2(π) ⊗ H as closed subspaces of E n ⊗ H. Identify these with subspaces of (E ⊗ H)n. The condition (2) in Definition 5.7 gives F1(π ⊗ H) = F1(π) ⊗ H 6= F2(π) ⊗ H = F2(π ⊗ H). That is, π ⊗ H does not belong to Rep00(A). Thus Rep00(A) inherits the Strong Local–Global Principle from Rep0(A). The Strong Local–Global Principle is easily seen to be hereditary for intersections. Hence any class of representations defined by submodule conditions satisfies the (cid:3) Strong Local–Global Principle by transfinite induction. Theorem 5.17. Let A be a ∗-algebra and let S ⊆ Ah. Let RepS(A) be the class of all representations where the elements of S act by regular, self-adjoint operators. This class is defined by submodule conditions and hence satisfies the Strong Local– Global Principle. It is admissible if S is a strong generating set for A. Proof. Asking π(a) to be regular and self-adjoint for a single a ∈ S is a submodule condition by Example 5.10. In order to ask this simultaneously for a set S, let ≺ be a well-ordering on S, and add an element M with a ≺ M for all a ∈ S. Let REPRESENTATIONS BY UNBOUNDED OPERATORS 0 ) and I is a ∗-intertwiner by Proposition 2.27. generating set, (2.10) gives E0 =T 25 Repa(A) ⊆ Rep(A) for a ∈ S ∪ {M} be the class of all representations π where π(b) is regular and self-adjoint for all b ∈ S with b ≺ a. These subclasses form a recursive definition of RepS(A) by submodule conditions as in Definition 5.7. Thus RepS(A) is defined by submodule conditions. Then it is weakly admissible and satisfies the Strong Local–Global Principle by Lemma 5.8 and Theorem 5.9. From now on, we assume that S is a strong generating set. For RepS(A) to be admissible, we must prove that any isometric intertwiner I : (E0, π0) ,→ (E, π) between two Hilbert space representations in RepS(A) is a ∗-intertwiner. If a ∈ S, then π(a) and π0(a) are regular, self-adjoint operators. Hence they generate integrable representations of C[x] as in Theorem 4.4. The isometry I intertwines these representations of C[x]. Hence it is a ∗-intertwiner by Theorem 4.4. In particular, I∗ maps dom π(a) to dom π0(a) for each a ∈ S. Since S is a strong a∈S dom π0(a) and similarly for π. So I∗(E) ⊆ E0. Then E = E0 + (E ∩ E⊥ (cid:3) Corollary 5.18. Let S ⊆ Ah be a strong generating set for a ∗-algebra A and let B with universal representation µ be a weak C∗-hull. If the closed multipliers µ(a) for a ∈ S are self-adjoint and affiliated with B, then B is a C∗-hull. Proof. All B-integrable representations belong to RepS(A) because the latter is weakly admissible and contains the universal B-integrable representation. Since RepS(A) is admissible by Theorem 5.17, any smaller class of integrable representa- tions inherits the equivalent conditions (2)–(4) in Proposition 3.8, which characterise C∗-hulls among weak C∗-hulls. (cid:3) Theorem 5.19. Let A be a ∗-algebra, B a C∗-algebra, (B, µ) a representation of A on B, and T1, . . . , Tn ∈ A. Assume that µ(T1), . . . , µ(Tn) are self-adjoint and affiliated with B and generate B in the sense of Woronowicz, see [31, Definition 3.1]. Then B is a C∗-hull for the B-integrable representations of A defined by (B, µ), and these satisfy the Strong Local–Global Principle. Proof. To show that B is a C∗-hull, we check the condition (5) in Proposition 3.8. Let : B → B(H) be a representation of B on a Hilbert space H and let (E, π) be the corresponding B-integrable representation of A. Let (E0, πE0) be a B-integrable representation on a closed subspace E0 ⊆ E and let P ∈ B(E) be the projection onto E0. We must show that (B) is contained in the commutant of P. Equivalently,  is a morphism in the notation of [31] to the algebra K = K(E0) ⊕ K(E⊥ 0 ) of all compact operators on E that commute with P. Let 1 ≤ i ≤ n. Since Ti is self-adjoint and regular as an adjointable operator on the Hilbert B-module B, it generates an integrable representation of the polynomial algebra C[x] on B as in Theorem 4.4. These integrable representations form an admissible class. Therefore, a B-integrable representation of A gives an integrable representation of C[x] when we compose with the canonical map ji : C[x] → A, x 7→ Ti, and take the closure. And since π and πE0 are both B-integrable, πE0 ◦ ji is a direct summand in π ◦ ji. Equivalently, the unbounded operator π(Ti) is affiliated with K. The extension of  to affiliated multipliers maps µ(Ti) to π(Ti), which is affiliated with K. Hence  is a morphism to K because these affiliated multipliers generate B. Thus B is a C∗-hull for the B-integrable representations by Proposition 3.8. Now we check the Strong Local–Global Principle. Let (E, π) be a representation of A on a Hilbert D-module E. Assume that the representation (E, π) ⊗ω Hω is integrable for each irreducible representation ω of D on a Hilbert space Hω in the sense that it comes from a representation of B. We must show that the representation (E, π) is integrable. 26 RALF MEYER The condition that π(Ti) be self-adjoint and regular is a submodule condition by Example 5.10. Hence the class of representations with this property satisfies the Strong Local–Global Principle by Theorem 5.9. Therefore, π(Ti) is a regular, self-adjoint operator on E for i = 1, . . . , n. Let ω be the direct sum of all irreducible representations of D; this is a faithful representation of D on some Hilbert space H. The induced representation j of K(E) on K := E ⊗D H is faithful as well. By assumption, the representation π ⊗D 1 of A on K is integrable, so it comes from a representation σ of B. The extension of σ to affiliated multipliers maps µ(Ti) η B to (π ⊗D 1)(Ti). Since π(Ti) is a regular operator on E, it is an affiliated multiplier of K(E), see [19] or Proposition 3.13. Thus (π ⊗D 1)(Ti) is affiliated with the image of K(E) on K by [16, Proposition 9.10]. Thus σ(µ(Ti)) η K(E) for i = 1, . . . , n. Since the affiliated multipliers µ(Ti) generate B in the sense of Woronowicz, σ factors through a morphism τ : B → K(E). This is the same as a representation of B on E. Let π0 be the representation of A on E associated to τ. If  is an irreducible Hilbert space representation of D, then π ⊗ H = π0 ⊗ H by construction of τ. Hence Theorem 5.6 gives π = π0. Since π0 (cid:3) is integrable by construction, so is π. The first counterexample in §6 exhibits a symmetric affiliated multiplier that generates a C∗-algebra, such that the Local–Global Principle fails and B is not a C∗-hull. Without self-adjointness, we only get the following much weaker statement: Lemma 5.20. Let A be a ∗-algebra, B a C∗-algebra, (B, µ) a representation of A on B, and T1, . . . , Tn ∈ A. Assume that µ(T1), . . . , µ(Tn) are affiliated with B and generate B in the sense of Woronowicz. Then B is a weak C∗-hull for the B-integrable representations of A. Proof. To show that B is a weak C∗-hull, we check (1) in Proposition 3.8. Let 1, 2 be representations of B on a Hilbert space H with (B, µ) ⊗1 H = (B, µ) ⊗2 H. We claim that 1 ⊕ 2 : B → B(H2) = M2(B(H)) maps B into the multiplier algebra of the diagonally embedded copy K of K(H). This is equivalent to 1 = 2. Since (B, µ) ⊗1 H = (B, µ) ⊗2 H, the extension of 1 ⊕ 2 to affiliated multipliers maps µ(Ti) η B to an operator of the form (Xi, Xi) for i = 1, . . . , n; these are affiliated with K. Since these affiliated multipliers generate B, 1 ⊕ 2 is a morphism from B to K. Thus B is a weak C∗-hull for A. (cid:3) 5.1. Universal enveloping algebras. We illustrate our theory by an example. Let g be a finite-dimensional Lie algebra over R and let A = U(g) be its universal enveloping algebra with the usual involution, where elements of g are skew-symmetric. A representation of A on E, possibly not closed, is equivalent to a dense submodule E ⊆ E with a Lie algebra representation π : g → EndD(E) satisfying hξ, π(X)(η)i = −hπ(X)(ξ), ηi for all X ∈ g, ξ, η ∈ E. Let G be a simply connected Lie group with Lie algebra g and let B := C∗(G). A representation of C∗(G) on a Hilbert module E is equivalent to a strongly continuous, unitary representation of G on E. Given such a representation, let E∞ ⊆ E be its subspace of smooth vectors. This is the domain of a closed representation of U(g). We call a representation of U(g) integrable if it comes from a unitary representation of G in this way. In particular, G acts continuously on C∗(G) by left multiplication with unitary multipliers. Let B = C∗(G)∞ be the right ideal of smooth elements for this G-action, equipped with the canonical U(g)-module structure µ. By the universal property of C∗(G), the pair (B, µ) is the universal integrable representation. That is, a representation of U(g) is integrable if and only if it is of the form (B, µ) ⊗ E for a representation  of C∗(G). Let X1, . . . , Xd form a basis of g. The Laplacian is L := −Pd REPRESENTATIONS BY UNBOUNDED OPERATORS i=1 X2 i ∈ U(g). 27 regular. His proof shows that all elements ofT∞ regular and E =T∞ then the domain of π isT∞ Theorem 5.21 ([20, Théorème 2.12]). A representation (π, E) of U(g) is integrable if and only if π(Ln) is regular and self-adjoint for all n ∈ N. Proof. Since Pierrot does not require representations to be closed, his statement is slightly different from ours. Pierrot shows that there is a continuous representation  of G with differential X 7→ π(X) if and only if T := π(L) is self-adjoint and n=1 dom T n are smooth vectors for . Conversely, all smooth vectors must belong to this intersection. A representation of U(g) is determined by its domain and the closed operators π(X) for X ∈ g. So a closed representation (E, π) of U(g) is integrable if and only if T is self-adjoint and n=1 dom T n. Moreover, the proof shows that the graph topology for a representation with regular self-adjoint T is determined by the graph norms of Ln for all n ∈ N. If π(Ln) is self-adjoint, then it must be equal to T n because π(Ln) ⊆ T n and T n is symmetric. Therefore, if π(L) is regular and self-adjoint, n=1 dom T n if and only if π(Ln) is regular and self-adjoint also for all n ≥ 2. (cid:3) Theorem 5.22. The class of integrable representations of U(g) has C∗(G) as a C∗-hull and is defined by submodule conditions. So it satisfies the Strong Local–Global Principle. Proof. By Theorem 5.21, a representation is integrable if and only if all elements of the set {Ln n ∈ N} act by a regular and self-adjoint operator. Hence the assertion follows from Theorem 5.17. Alternatively, the closed multipliers of C∗(G) associated to iX1, . . . , iXd are regular and affiliated with C∗(G) and generate C∗(G) by [31, Example 3 in §3]. Hence C∗(G) is a C∗-hull and the Strong Local–Global Principle holds by Theorem 5.19. (cid:3) The results of Vassout [29] get close to proving an analogue of Theorem 5.22 for an s-simply connected Lie groupoid G with compact base. This analogue would replace g by the space of smooth sections of the Lie algebroid A(G), and U(g) by the ∗-algebra of G-equivariant differential operators on G, a subalgebra of the ∗-algebra of G-pseudodifferential operators. Any symmetric, elliptic element of U(g) should be a possible replacement for the Laplacian in Theorem 5.22. 6. Polynomials in one variable II We discuss two classes of "integrable" representations of the ∗-algebra C[x] with x = x∗ which are weakly admissible, but not admissible, and which violate the Local–Global Principle. Both examples have a weak C∗-hull, on which all powers of the generator x act by an affiliated multiplier. In the first example, these affiliated multipliers generate the weak C∗-hull, but not in the second. Neither Theorem 5.17 nor Theorem 5.19 apply because the generating affiliated multipliers are not self- adjoint. The first example shows that a C∗-algebra generated by affiliated multipliers in the sense of Woronowicz need not be a C∗-hull, though it is always a weak C∗-hull by Lemma 5.20. The second example shows that a weak C∗-hull need not be generated by affiliated multipliers. Let S ∈ B('2N) be the unilateral shift. Let Q be the closed symmetric operator on '2N with Cayley transform S. Thus Q has deficiency index (0, 1). The domain of Q is (1−S)'2N, and Q(1−S)ξ := i(1+S)ξ for all ξ ∈ '2N (see also Example 4.10). We may identify '2N with the Hardy space H2. Then Q becomes the Toeplitz operator with the unbounded symbol i(1 + z)(1 − z)−1. Let T be the Toeplitz C∗-algebra, that is, the C∗-subalgebra of B('2N) generated by S. Every element in T is of the form Tϕ + K, where Tϕ is the Toeplitz operator 2i(1−S) belongs to T0. Hence (Q+i)∗ = Q∗−i is the inverse of 1 28 with symbol ϕ ∈ C(S1) and K is a compact operator. Let T0 / T be the kernel of the unique ∗-homomorphism T → C that maps S to 1. Proposition 6.1. There is a symmetric, affiliated multiplier Q of T0 with do- main (1 − S) · T0 and Q · (1 − S) · t := i(1 + S) · t for all t ∈ T0. It generates T0 in the sense of Woronowicz. Proof. We claim that the right ideal (1−S)S∗T0 ⊆ T0 is dense. This would fail for T because the continuous ∗-homomorphism T → C, S 7→ 1, annihilates this right ideal. First, (1− S)S∗K('2N) is dense in K('2N) because (1− S)S∗ has dense range on '2N. So the closure of (1 − S)S∗T0 contains K('2N). Secondly, (1 − S)S∗T0/K('2N) is dense in T0/K('2N) ∼= C0(S1 \ {1}) because the function (1 − z)z on S1 vanishes only at 1. An affiliated multiplier of T0 is the same as a regular operator on T0, viewed as a Hilbert module over itself. Since (1 − S)S∗T0 is dense in T0, there is a regular, symmetric operator Q0 on T0 that has S as its Cayley transform, see [16, Chapter 10]. The operator Q0 has the domain (1−S)S∗T0 and acts by Q0·(1−S)S∗t := i(1+S)S∗t. Rewriting any t ∈ T0 as t = S∗St, we may replace S∗t by t here. Thus Q0 = Q. Since Q + i maps (1 − S)t to i(1 + S)t + i(1 − S)t = 2it, it is surjective, and −2i(1−S∗). (Q+i)−1 = 1 So Q∗ has domain (1 − S∗)T0 and maps (1 − S∗)t 7→ i(1 − S∗)t − 2it = −i(1 + S∗)t. As expected, Q∗ contains Q: we may write (1 − S)t = S∗St − St = (1 − S∗)(−St), and Q∗ maps this to −i(1 + S∗)(−St) = i(S + 1)t. 4(1 − S)(1 − S∗) ∈ T0. We compute Next we show that Q∗Q + 1 is the inverse of 1 Q∗Q(1 − S)(1 − S∗)t = iQ∗(1 + S)(1 − S∗)t = iQ∗(1 + S − S∗ − SS∗)t = iQ∗(1 − S∗)(2 + S − SS∗)t = (1 + S∗)(2 + S − SS∗)t = (4 − (1 − S)(1 − S∗))t. This implies (Q∗Q + 1)(1 − S)(1 − S∗)t = 4t. Since this is already surjective and Q∗Q + 1 is injective, the domain of Q∗Q + 1 is exactly (1 − S)(1 − S∗)T0, and Q∗Q + 1 is the inverse of 1 Let 1 and 2 be two Hilbert space representations of T0 whose extension to affiliated multipliers maps Q to the same unbounded operator. Then they also map the Cayley transform S of Q to the same partial isometry. So 1(S) = 2(S), which gives 1 = 2. Thus Q separates the representations of T0. Since (Q∗Q + 1)−1 ∈ T0 as well, [31, Theorem 3.3] shows that the affiliated multiplier Q generates T0. (cid:3) The domain of Qn is the right ideal (1− S)n·T0, which is dense in T0 for the same reason as (1−S)·T0. Even more, the right ideal (1−S)n+1·T0 is dense in (1−S)n·T0 in the graph norm of Qn. Thus the intersection T of this decreasing chain of dense right ideals (1 − S)nT0 is still dense in T0 by [27, Lemma 1.1.2]. This intersection is the domain of a closed representation µ of C[x] on T0 with µ(xn) = Qn. We call a representation of C[x] on a Hilbert module E Toeplitz integrable if it is of the form (T, µ) ⊗ E for some representation : T0 → B(E). Proposition 6.2. The class of Toeplitz integrable representations of C[x] is weakly admissible with the weak C∗-hull T0. It is not admissible, so T0 is not a C∗-hull. The Toeplitz integrable representations violate the Local–Global Principle. A representation (E, π) of C[x] on a Hilbert module E over a C∗-algebra D is 4(1 − S)(1 − S∗) ∈ T0 as asserted. RALF MEYER Toeplitz integrable if and only if it has the following properties: (1) π(x + i)nE = E for all n ∈ N≥1; (2) π(x) is regular. Toeplitz integrable representations on E are in bijection with regular, symmetric operators T on E for which T + i is surjective. REPRESENTATIONS BY UNBOUNDED OPERATORS 29 Proof. We checked condition (1) in Proposition 3.8 in the proof of Proposition 6.1. Thus T0 is a weak C∗-hull for the Toeplitz integrable representations, and this class is weakly admissible. Any self-adjoint operator on a Hilbert space generates a Toeplitz integrable representation of C[x] because T0/K('2N) ∼= C0(R); so does Q itself. Thus both Example 4.10 and Proposition 4.8 show that the class of Toeplitz integrable representations is not admissible. So T0 is not a C∗-hull. We claim that the representation (T, µ) of C[x] on T0 has the properties (1) domain T :=T∞ and (2) in the proposition. First, (µ(x) + i)n acts by (2i)n(1 − S)−n on its dense k=1(1 − S)kT0. Since (1 − S)k+1T0 is norm dense in (1 − S)kT0, the closure of (µ(x) + i)n is equal to (2i)n(1 − S)−n with its natural domain (1 − S)nT0, and this operator is surjective. Secondly, µ(x) = Q is regular. The property (1) is a sequence of submodule conditions, see Example 5.13. Hence it is inherited by interior tensor products by Lemma 5.8. So is the property (2) by [16, Proposition 9.10]. Hence both (1) and (2) are necessary for a representation (E, π) to be Toeplitz integrable. Conversely, let (E, π) be a representation of C[x] on E that satisfies (1) and (2). Then the closed, symmetric operator T := π(x) on E is regular by (2). So its Cayley transform s is an adjointable partial isometry such that (1 − s)s∗ has dense range (see [16, Chapter 10]). Even more, s is an isometry because (T + i)E = E. Thus s generates a unital representation  of T . The restriction of  to T0 is nondegenerate because (1 − s)s∗ has dense range. Let π0 := µ ⊗ 1 be the representation of C[x] associated to . Then π0((x + i)n) = (2i)n(1 − s)−n ⊇ π((x + i)n). Assumption (1) implies that E is dense in the domain of (2i)n(1 − s)−n in the graph norm of (2i)n(1 − s)−n. Hence even π0((x + i)n) = (2i)n(1 − s)−n = π((x + i)n). Since the domains of π(x)n form a decreasing sequence, induction on n now shows that π0(xn) = π(xn). The set {xn} is a strong generating set for C[x] by Lemma 4.1. Thus π = π0 by Proposition 2.9. This finishes the proof that Toeplitz integrable representations of C[x] are characterised by the properties (1) and (2) and that they are in bijection with regular, symmetric operators T for which T + i is surjective. For a counterexample to the Local–Global Principle, let ¯N = N ∪ {∞} be the one-point compactification of N and D = C(¯N). Let E ⊆ C(¯N, '2N) consist of all continuous functions f : ¯N → '2N with f(∞)⊥δ0. The unilateral shift S on C(¯N, '2N) restricts to a non-adjointable isometry s on this subspace. Let T be the inverse Cayley transform of s. This is a closed, symmetric operator on E that is irregular because its Cayley transform is not adjointable. If : D → B(H) is a Hilbert space representation, then the induced representation of C[x] is associated to the closed operator T ⊗1. The operator (T ⊗1)+i remains surjective, and T ⊗1 is regular because it acts on a Hilbert space. So T ⊗ 1 generates a Toeplitz integrable representation for all representations  of D. Since T itself does not generate a (cid:3) Toeplitz integrable representation, the Local–Global Principle is violated. Condition (1) in Proposition 6.2 is a submodule condition. If regularity without self-adjointness were a submodule condition as well, then the Toeplitz integrable representations of C[x] would be defined by submodule conditions; so the failure of the Local–Global Principle for them would contradict Theorem 5.19. The identical inclusion T0 ,→ M(K('2N)) is a representation of the weak C∗-hull T0 on K('2N) and thus corresponds to a Toeplitz integrable representation of C[x] on K('2N). This is simply the restriction of (T, µ) to the Hilbert T0-submodule K('2N) ⊂ T0, with domain T ∩ K('2N) and the same action µ of C[x]. Call a representation purely Toeplitz integrable if it is of the form (T ∩ K('2N), µ) ⊗ E for some representation : K('2N) → B(E). 30 RALF MEYER Proposition 6.3. The purely Toeplitz integrable representations of C[x] form a weakly admissible class that is not admissible, and K('2N) is a weak C∗-hull for it, but not a C∗-hull. This class violates the Local–Global Principle. The closed multiplier Q = µ(x) of T0 is affiliated with K('2N) but does not generate K('2N). A representation (E, π) of C[x] on a Hilbert module E over a C∗-algebra D is purely Toeplitz integrable if and only if it has the following property in addition to those in Proposition 6.2: (3) the closure of S∞ n=1(π(x − i)nE)⊥ is E. Proof. Since K('2N) has fewer representations than T0, the condition (1) in Propo- sition 3.8 for K('2N) follows from the corresponding property for T0, which we have already checked in the proof of Proposition 6.1. Hence K('2N) is a weak C∗-hull for the purely Toeplitz representations of C[x]. Since Q gives a purely Toeplitz representation of C[x] on '2(N), the class of purely Toeplitz integrable representations is not admissible by Example 4.10. Therefore, its weak C∗-hull is not a C∗-hull. The same counterexample as in the proof of Proposition 6.2 shows that the Local–Global Principle fails for the purely Toeplitz representations. Any closed operator on '2N is affiliated with K('2N). In particular, so is Q. In the identical representation of K('2N) on the Hilbert space '2N, the image of Q is affiliated with T0 by Proposition 6.1. But the representation of K('2N) is not by a morphism to T0 because the inclusion map K('2N) ,→ T0 is degenerate. Hence Q does not generate K('2N) in the sense of Woronowicz. The element Pn := 1 − Sn(S∗)n ∈ K('2N) ⊆ T0 is the orthogonal projection onto the span of δ0, . . . , δn−1. A representation of T0 maps Pn to an orthogonal projection whose image is the orthogonal complement of the image of Sn. This is also the orthogonal complement of the image of π(x − i)n. These orthogonal complements form an increasing chain of complementable submodules, and π is purely Toeplitz if and only if their union is all of E. This proves our characterisation (cid:3) of purely Toeplitz representations. 7. Bounded and locally bounded representations Let A be a ∗-algebra. A bounded representation of A on a Hilbert module E is a ∗-homomorphism π : A → B(E). Corollary 2.11 says that a closed representation is bounded once π(a) is globally defined for a in a strong generating set of A. Finite-dimensional representations are always bounded. In particular, characters are bounded. Thus commutative ∗-algebras have many bounded representations. Many other ∗-algebras, such as the Weyl algebra, have no bounded representations. In this section, we are going to study C∗-hulls related to bounded representations. These are only relevant if A has many bounded representations. Any bounded representation π of A is bounded in some C∗-seminorm q on A, that is, kπ(a)k ≤ q(a) for all a ∈ A. Then π extends to the (Hausdorff) completion Aq of A in the seminorm q, which is a unital C∗-algebra. If p, q are two C∗-seminorms on A, then max{p, q} is a C∗-seminorm as well. Thus the set N (A) of C∗-seminorms on A is directed. For q, q0 ∈ N (A) with q ≤ q0, let ϕq,q0 : Aq0 → Aq be the ∗-homomorphism induced by the identity map on A. The C∗-algebras Aq and the ∗-homomorphisms ϕq,q0 for q ≤ q0 in N (A) form a projective system of C∗-algebras. Each ∗-homomorphism ϕq,q0 is unital and surjective because its image contains A, which is unital and dense in Aq0. The C∗-seminorms in N (A) define a locally convex topology on A, where a net converges if and only if it converges in any C∗-seminorm. Let A with the canonical map j : A → A be the completion of A in this topology. This is a C∗-algebra if and REPRESENTATIONS BY UNBOUNDED OPERATORS 31 only if there is a largest C∗-seminorm on A. In general, A is the projective limit of the diagram of unital C∗-algebras (Aq, ϕq,q0) described above. Thus A is a unital pro-C∗-algebra, see [21]. As a concrete example, we describe A for a commutative ∗-algebra A. Definition 7.1. Let A be the set of ∗-homomorphisms A → C, which we briefly call characters. Each a ∈ A gives a function a: A → C, a(χ) := χ(a). We equip A with the coarsest topology making these functions continuous. That is, a net (χi)i∈I in A converges to χ ∈ A if and only if lim χi(a) = χ(a) for all a ∈ A. Let τc be the compactly generated topology associated to this topology, that is, a subset in A is closed in τc if and only if its intersection with any compact subset in A is closed. If a ∈ A, then its Gelfand transform a is a continuous function on A. This defines a ∗-homomorphism A → C( A). If the usual topology on A is locally compact or metrisable, then it is already compactly generated and hence equal to τc. The topology τc may have more closed subsets and hence more continuous functions to C. So C( A) ⊆ C( A, τc). Proposition 7.2. Let A be a commutative ∗-algebra. The directed set N (A) of C∗-seminorms on A is isomorphic to the directed set of compact subsets of A, where K ⊆ A corresponds to the C∗-seminorm kakK := sup{a(χ) χ ∈ K}. The C∗-completion of A in this C∗-seminorm is C(K). And A ∼= C( A, τc), where the inclusion map j : A → A is the Gelfand transform A → C( A, τc), a 7→ a. Proof. Let q be a C∗-seminorm on A. Let Aq ⊆ A be the subspace of all q-bounded characters, that is, χ ∈ Aq if and only if χ(a) ≤ q(a) for all a ∈ A. These are precisely the characters that extend to characters on the C∗-completion Aq. Conversely, since A is dense in Aq, any character on Aq is the unique continuous extension of a q-bounded character on A. And the subspace topology on Aq ⊆ A is equal to the canonical topology on the spectrum of Aq: a net of q-bounded characters that converges uniformly on A also converges uniformly on Aq. Thus Aq ∼= C( Aq) by the Gelfand–Naimark Theorem, and so Aq ⊆ A is compact for each q ∈ N (A). If q ≤ q0, then Aq ⊆ Aq0 and ϕqq0 : Aq0 (cid:16) Aq is the restriction map for the subspace Aq ⊆ Aq0. The pro-C∗-algebra A is the limit of this diagram of commutative C∗-algebras. Since all the maps Aq ⊆ Aq0 are injective, A is the algebra of continuous q∈N (A) Aq with the inductive limit topology. That is, a subset of q∈N (A) Aq is closed if and only if its intersection with each Aq is closed, where Aq carries the (compact) subspace topology from A. Any character χ on A is bounded with respect to some C∗-seminorm; for instance, q∈N (A) Aq = A as a set. If K ⊆ A is compact, then a ∈ C( A) for a ∈ A must be uniformly bounded on K, so that functions on S S kakχ := χ(a). ThusS limit topology onS is a C∗-seminorm on A. Thus K ⊆ Aq for some q ∈ N (A). Hence the inductive (cid:3) q∈N (A) Aq is τc. We return to the general noncommutative case. The class of q-bounded represen- tations for a fixed q ∈ N (A) is easily seen to be weakly admissible. The class of bounded representations with variable q is not weakly admissible unless A has a kakK := sup{a(χ) χ ∈ K} RALF MEYER 32 largest C∗-seminorm because it is not closed under direct sums. We are going to define the larger class of "locally bounded" representations to rectify this. Roughly speaking, a representation is locally bounded if and only if it comes from a represen- tation of the pro-C∗-algebra A. Before we define locally bounded representations, we characterise q-bounded representations by some slightly weaker estimates. Proposition 7.3. Let A be a ∗-algebra and let q be a C∗-seminorm on A. Let (E, π) be a representation of A on a Hilbert module E over some C∗-algebra D and let ξ ∈ E. The following are equivalent: (1) there is C > 0 with khξ, π(a)ξik ≤ Cq(a) for all a ∈ A; (2) there is C > 0 with kπ(a)ξk ≤ Cq(a) for all a ∈ A; (3) kπ(a)ξk ≤ kξkq(a) for all a ∈ A. The set of vectors ξ with these equivalent properties is a norm-closed A, D-submodule of E. The representation of A on this submodule extends to the C∗-completion Aq. Proof. The implications (3)⇒(2)⇒(1) are trivial. Conversely, assume (1) and let a ∈ A. Let (bn)n∈N be a sequence in A that converges in Aq towards the positive square-root of q(a)2 − a∗a. Then the sequence (a∗a + b∗ nbn) in A converges in the norm q to q(a)2 ∈ A. If (1) holds, then n→∞hξ, π(a∗a + b∗ lim nbn)ξi = q(a)2hξ, ξi. Since 0 ≤ hπ(a)ξ, π(a)ξi ≤ hπ(a)ξ, π(a)ξi + hπ(bn)ξ, π(bn)ξi = hξ, π(a∗a + b∗ for all n, this implies kπ(a)ξk ≤ limkhξ, π(a∗a + b∗ implies (3). nbn)ξi nbn)ξik = q(a)2kξk2. Thus (1) The set Eq of vectors ξ ∈ E satisfying (2) is a vector subspace and closed under left multiplication by elements of A and right multiplication by elements of D. On this subspace, the graph and norm topologies coincide because of (3). The subspace Eq is closed in the norm topology by the Principle of Uniform Boundedness. The ∗-representation of A on this submodule is globally defined and bounded by the C∗-seminorm q. Hence it extends to a representation of Aq. (cid:3) Definition 7.4. Let (E, π) be a representation of A on a Hilbert module E. A vector ξ ∈ E is bounded if it satisfies the equivalent conditions in Proposition 7.3 for some q ∈ N (A). The representation is locally bounded if the bounded vectors are dense in E in the graph topology. By Proposition 7.3, the q-bounded vectors in E for a fixed q ∈ N (A) form a closed A, D-submodule Eq ⊆ E, on which the representation of A extends to the C∗-completion Aq and hence to a representation of A. Since N (A) is directed and Eq ⊆ Eq0 if q ≤ q0, the family of sub-bimodules Eq ⊆ E is directed. The set of bounded vectors is the increasing union Eb := [ q∈N (A) Eq. Since πEq extends to A for each q, there is a representation ¯π of the pro-C∗-algebra A on Eb ⊆ E. The representation (E, π) is locally bounded if and only if (Eb, ¯π ◦ j) is a core for it. Thus (E, π) is the closure of the "restriction" ¯π ◦ j of ¯π to A. We do not claim that ¯π is closed, and neither do we claim that ¯π ◦ j is locally bounded for any representation of A: we need the representation of A to be locally bounded as well: Definition 7.5. A representation (π, E) of a pro-C∗-algebra A is locally bounded if the vectors ξ ∈ E for which A → E, a 7→ π(a)ξ, is continuous form a core. REPRESENTATIONS BY UNBOUNDED OPERATORS 33 Proposition 7.6. Composition with j : A → A induces an equivalence between the categories of locally bounded representations of A and A which is compatible with isometric intertwiners and interior tensor products. Proof. The ∗-homomorphism j induces an isomorphism between the directed sets of C∗-seminorms on A and A. Therefore, a representation ¯π of A is locally bounded if and only if the vectors ξ with k¯π(a)ξk ≤ q(a)kξk for all a ∈ A, for some q ∈ N (A), form a core. Since j(A) is dense in A, this is equivalent to kπ(a)ξk ≤ q(a)kξk for all a ∈ A. Thus the closure of ¯π ◦ j is locally bounded if and only if ¯π is. An isometric intertwiner ¯π1 ,→ ¯π2 also intertwines the closures of ¯π1 ◦ j and ¯π2 ◦ j by Lemma 2.14. Conversely, an isometric intertwiner between two locally bounded representations of A must map q-bounded vectors to q-bounded vectors for any q ∈ N (A). Thus it remains an isometric intertwiner between the canonical extensions of the representations to A. So the equivalence between the locally bounded representations of A and A is compatible with isometric intertwiners. It is also compatible with interior tensor products, that is, the closure of (¯π ⊗D 1G) ◦ j is ¯π ◦ j ⊗D 1G. (cid:3) Proposition 7.7. All irreducible, locally bounded Hilbert space representations are bounded. Proof. If π is irreducible, then the closed A-submodule Eq for a C∗-seminorm q is either {0} or E. The latter must happen for some q if π is locally bounded. (cid:3) Thus local boundedness is not an interesting notion for irreducible representations. If A has no C∗-seminorms, then A = {0} and A has no locally bounded represen- tations, so that the following discussion will be empty. Even if the map j : A → A is not injective, there are examples where all "integrable" representations of A come from A. An important case is the unit fibre for the canonical Z-grading on the Virasoro algebra studied in [26, §9.3]. In this case, A is not commutative, but all irreducible, integrable representations are characters and hence locally bounded. Proposition 7.8. If π is a locally bounded representation, then π(a) is regular and self-adjoint for each a ∈ Ah. Proof. The map of left multiplication by j(a)±i on A is invertible because j(a) ∈ A is symmetric and A is a pro-C∗-algebra. Therefore, ¯π(j(a)) ± i ⊆ π(a) ± i has dense range on E. Thus π(a) is regular and self-adjoint, see [16, Chapter 10]. (cid:3) Corollary 7.9. Let A be a ∗-algebra. The class Repb(A) of locally bounded repre- sentations of A is admissible. Proof. Being locally bounded is clearly invariant under unitary ∗-intertwiners and direct sums. It is also invariant under direct summands because a ∗-intertwiner maps bounded vectors to bounded vectors. If ξ ∈ E is bounded, then ξ ⊗ η ∈ E ⊗D F is bounded for any C∗-correspondence F. Thus a locally bounded representation on E induces one on E ⊗D F. Since Ah is a strong generating set for A by Example 2.8, the class of rep- resentations for which all a ∈ Ah act by a regular and self-adjoint operator is admissible by Theorem 5.17. This class contains the locally bounded representations (cid:3) by Proposition 7.8. Hence this subclass is also admissible. Any pro-C∗-algebra A contains a dense unital C∗-subalgebra Ab of bounded elements, see [21, Proposition 1.11]. For instance, if A is commutative, so that A ∼= C( A, τc) by Proposition 7.2, then Ab = Cb( A, τc) consists of the bounded continuous functions. 34 RALF MEYER Let (E, π) be a locally bounded representation of A. This comes from a locally bounded representation (Eb, ¯π) of A by Proposition 7.6. The closure of the restric- tion of ¯π to Ab is a representation of a unital C∗-algebra. Hence it is a unital ∗-homomorphism : Ab → B(E) by Lemma 2.12. Proposition 7.10. Two locally bounded representations π1 and π2 of A on a Hilbert module E are equal if and only if they induce the same representation of Ab. Proof. Of course, π1 and π2 induce the same representation of Ab if π1 = π2. Conversely, assume that π1 and π2 induce the same representation  of Ab. If a ∈ Ah, then the Cayley transform ca of j(a) ∈ A is a unitary element of Ab. The Cayley transforms of π1(a) and π2(a) are both equal to (ca). Hence π1(a) = π2(a). Since this holds for all a ∈ Ah, Proposition 2.9 gives π1 = π2. (cid:3) The C∗-algebra Ab usually has many representations that do not come from locally bounded representations of A. Hence it is not a C∗-hull. It is, however, a useful tool to decide when a representation µ of A on a C∗-algebra B is a weak C∗-hull, that is, when A separates the Hilbert space representations of B: Proposition 7.11. Let µ be a locally bounded representation of A on a C∗-algebra B and let : Ab → M(B) = B(B) be the associated representation of Ab. The image of  is dense in M(B) in the strict topology if and only if B is a weak C∗-hull for the class of B-integrable representations of A defined by µ. Proof. Combine Proposition 7.10 and the following proposition for D = Ab. (cid:3) Proposition 7.12. Let µ be a representation of A on a C∗-algebra B. Let D be a C∗-algebra and ϕ: D → M(B) a ∗-homomorphism. Assume that two repre- sentations 1, 2 of B on a Hilbert space H satisfy µ ⊗1 1H = µ ⊗2 1H if and only if ¯1 ◦ ϕ = ¯2 ◦ ϕ, where ¯1 and ¯2 denote the unique strictly continuous extensions of 1, 2 to M(B). Then B is a weak C∗-hull for a class of integrable representations of A if and only if ϕ(D) is dense in M(B) in the strict topology. Proof. We use the criterion for weak C∗-hulls in (1) in Proposition 3.8. Assume first that ϕ(D) is strictly dense in M(B). Let 1, 2 be two Hilbert space representations of B that satisfy µ ⊗1 1H = µ ⊗2 1H. Extend 1, 2 to strictly continuous representations ¯1, ¯2 of M(B). By assumption, ¯1 ◦ ϕ = ¯2 ◦ ϕ, that is, ¯1 and ¯2 are equal on ϕ(D). Since they are strictly continuous and ϕ(D) is strictly dense, we get ¯1 = ¯2 and hence 1 = 2. Thus the condition (1) in Proposition 3.8 is satisfied, making B a weak C∗-hull of A. Conversely, assume that ϕ(D) is not strictly dense in M(B). We claim that the image of D is not weakly dense in the bidual W∗-algebra B∗∗. Any positive linear functional on B extends to a strictly continuous, positive linear functional on M(B) by extending its GNS-representation to a strictly continuous representation of M(B). By the Jordan decomposition, the same remains true for self-adjoint linear functionals and hence for all bounded linear functionals on B. Furthermore, such extensions are unique because B is strictly dense in M(B). Hence restriction to B maps the space of strictly continuous linear functionals on M(B) isomorphically onto the dual space B∗ of B, which is also the space of weakly continuous linear functionals on B∗∗. If ϕ(D) is not strictly dense in M(B), then the Hahn–Banach Theorem gives a non-zero functional in B∗ that vanishes on the image of D. When viewed as a weakly continuous functional on B∗∗, it witnesses that ϕ(D) is not weakly dense in B∗∗. Let : B → B(H) be the direct sum of all cyclic representations of B. Then  extends to an isomorphism of W∗-algebras from B∗∗ onto the double commu- tant (B)00 of B in B(H). The extension of  to M(B) restricts to a representation REPRESENTATIONS BY UNBOUNDED OPERATORS 35 ¯ ◦ ϕ: D → B(H). Since we assume that the image of D is not strictly dense in M(B), our claim shows that ¯ ◦ ϕ(D) is not weakly dense in (B)00. By the bicommutant theorem, this is equivalent to ¯ ◦ ϕ(D)0 6= (B)0. Since these commutants are C∗-algebras, they are the linear spans of the unitaries that they contain. So there is a unitary operator U in ¯◦ ϕ(D)0 that is not contained in (B)0. So 2 := U U∗ 6= , but ¯2 ◦ ϕ = ¯ ◦ ϕ. By assumption, the latter implies µ ⊗ 1H = µ ⊗2 1H. So A fails to separate the representations , 2 of B although they are not equal. Hence B is not a weak C∗-hull of A. (cid:3) Remark 7.13. Proposition 7.12 applies whenever we can somehow produce enough bounded operators from a representation of A so that these bounded operators and the original representation have the same unitary ∗-intertwiners. For instance, it applies if the elements of a strong generating set for A act by regular operators, so that we may take their bounded transforms. Prim(Aq0) with a closed subspace of Prim(Aq). Let Prim A :=S The quotient map Aq (cid:16) Aq0 for q ≥ q0 in N (A) identifies the primitive ideal space q∈N (A) Prim(Aq). Let a ∈ A and p ∈ Prim(A). Then the norm kakp of the image of a in Aq/p is the same for all q ∈ N (A) with p ∈ Prim(Aq). Hence the function p 7→ kakp on Prim(A) is well defined. Definition 7.14. An element a ∈ A vanishes at ∞ if for every ε > 0 there is q ∈ N (A) such that kakp < ε for p ∈ Prim(A) \ Prim(Aq). An element a ∈ A is compactly supported if there is q ∈ N (A) with a ∈ p for all p ∈ Prim(A) \ Prim(Aq). Let C0(A) and Cc(A) be the subsets of elements that vanish at ∞ and have compact support, respectively. It may happen that C0(A) = {0}. In the following, we are interested in the case where C0(A) is dense in A. For instance, C0(R) is dense in C(R). Lemma 7.15. The subset C0(A) is a closed ideal in Ab. The subspace Cc(A) is a two-sided ∗-ideal in A. It is norm-dense in C0(A). More generally, if D is a C∗-algebra and ϕ: D → A is a ∗-homomorphism, then ϕ−1(Cc(A)) is dense in ϕ−1(C0(A)). Proof. The quotient maps A (cid:16) Aq (cid:16) Aq/p for p ∈ Prim(Aq) are ∗-homomorphisms. Thus C0(A) is a ∗-subalgebra of A. An element a ∈ A is bounded if and only if the norms of its images in Aq for q ∈ N (A) are uniformly bounded. The norm of a in Aq is the maximum of kakp for p ∈ Prim(Aq). Hence a is bounded if and only if the function kakp on Prim(A) is bounded. Thus C0(A) consists of bounded elements, and it is an ideal in Ab. We claim that the limit a of a norm-convergent sequence (an)n∈N in C0(A) again vanishes at ∞. Given ε > 0, there is n0 ∈ N so that ka − ankp ≤ ka − ank < ε/2 for all n ≥ n0 and all p ∈ Prim(A). Since an vanishes at ∞, there is q ∈ N (A) with kankp < ε/2 for p /∈ Prim(Aq). Thus kakp < ε for p /∈ Prim(Aq). Thus C0(A) is a closed ideal in Ab. The condition a ∈ p for fixed p ∈ Prim(A) defines a closed two-sided ∗-ideal in A. Hence Cc(A) is a two-sided ∗-ideal in A. Let a ∈ C0(A) and ε > 0. Let fε ∈ Cb([0,∞)) be increasing and satisfy fε(t) = 0 for 0 ≤ t < ε and fε(t) = 1 for 2ε ≤ t. Then ka− afε(a∗a)k ≤ 2ε, and fε(a∗a) ∈ p if ka∗akp ≤ ε. Hence afε(a∗a) ∈ Cc(A) for all ε > 0. Thus Cc(A) is dense in C0(A). Similarly, if ϕ: D → A is a ∗-homomorphism, x ∈ D, and ϕ(x) ∈ C0(A), then ϕ(xfε(x∗x)) ∈ Cc(A) and (cid:3) limε→0 xfε(x∗x) = x in the norm topology on D. Theorem 7.16. Let A be a ∗-algebra and let A be its pro-C∗-algebra completion. If C0(A) is dense in A, then C0(A) is a C∗-hull for the class of locally bounded representations of A. 36 RALF MEYER We shall prove a more general theorem that still applies if C0(A) is not dense in A. Then probably there is no C∗-hull for the class of all locally bounded representations. We may, however, find C∗-hulls for smaller classes of representations. We describe such classes of representations by a generalisation of the spectral conditions in Example 5.14. The spectral condition for a locally closed subset in Rn implicitly uses a subquotient of C0(Rn). We are going to describe subquotients of the pro-C∗- algebra A. We then associate a class Repb(A,K) of locally bounded representations of A to a subquotient K. If C0(K) is dense in K, then C0(K) is a C∗-hull for Repb(A,K). Theorem 7.16 is the special case K = A. Let J / A be a closed, two-sided ∗-ideal in the pro-C∗-algebra A. Being closed, the ideal J is complete in the subspace topology, so it is also a pro-C∗-algebra. Thus J = lim←−Jq, where Jq / Aq is the image of J in the quotient Aq. The quotient A/J is complete if A is metrisable, that is, its topology is defined by a sequence of C∗-seminorms. It need not be complete in general, however. Therefore, we replace the quotient A/J by its completion B, which is a pro-C∗-algebra as well. It is the projective limit of the quotients Aq/Jq for all q ∈ N (A). A subquotient of A is a closed, two-sided ∗-ideal K / B with B as above. Let Repb(A,K) consist of all representations π of A with the following properties: (1) π is locally bounded, so it comes from a locally bounded representation π0 (2) the representation π0 annihilates J ; (3) the representation ¯π of B induced by π0 is nondegenerate on K, that is, of A; ¯π(K)(E) is a core for ¯π. Define the C∗-algebra C0(K) and its dense ideal Cc(K) by replacing A by K in Definition 7.14. Equivalently, C0(K) = C0(B) ∩ K. We may choose J = 0 and K = A. Then Repb(A,A) = Repb(A) simply consists of all locally bounded representations of A. Hence Theorem 7.16 is the special case K = A of the following theorem: Theorem 7.17. If C0(K) is dense in K, then C0(K) is a C∗-hull for Repb(A,K). Proof. First we claim that Repb(A,K) is equivalent to the class of nondegenerate, locally bounded representations of the pro-C∗-algebra K as in Definition 7.5. If K = A, this is Proposition 7.6. A locally bounded representation π0 of A descends to a representation π00 of the quotient A/J if and only if it annihilates J ; the induced representation of A/J remains locally bounded with respect to the family of C∗-seminorms from the quotient mappings A/J (cid:16) Aq/Jq. Hence it extends uniquely to a locally bounded representation ¯π00 of the completion B. Thus locally bounded representations of A for which the corresponding representation of A annihilates J are equivalent to locally bounded representations of B. We claim that a nondegenerate, locally bounded representation  of K extends uniquely to B. Let q be a continuous seminorm on B, also write q for its restriction to K. The q-bounded vectors for  form a nondegenerate Kq-module. The module structure extends uniquely to the multiplier algebra of Kq, and Bq maps to this multiplier algebra because Kq / Bq. Letting q vary gives a locally bounded represen- tation of B that remains nondegenerate on K. Conversely, any such representation of B is obtained in this way from its restriction to K. Thus representations of A that belong to Repb(A,K) are equivalent to nondegenerate, locally bounded representa- tions of the pro-C∗-algebra K. The equivalence above is compatible with isometric intertwiners, direct sums and interior tensor products, compare Proposition 7.6. Lemma 7.15 shows that Cc(K) := Cc(B) ∩ K is dense in K. This is an ideal in B as an intersection of two ideals. Hence left multiplication defines a representation of B on K with core Cc(K), which is locally bounded by construction. Through REPRESENTATIONS BY UNBOUNDED OPERATORS on Eb =S 37 the canonical homomorphisms A → A → B, this becomes a representation of A. This representation clearly belongs to Repb(A,K). We claim that it is the universal representation for the class Repb(A,K). So let π be any representation in Repb(A,K). Then π comes from a unique nondegenerate, locally bounded representation ¯π of K. We must show that it comes from a unique nondegenerate representation  of C0(K). Let  be the restriction of ¯π to C0(K). Then (Cc(K))E ⊆ Eb ⊆ E. We are going to prove that this is a core. The bilinear map K (cid:12) Eb → E is separately continuous with respect to the pro-C∗-algebra topology on K and the inductive limit topology q∈N (A) Eq. We have assumed that it has dense range. Since Cc(K) is dense in K, the image of Cc(K) (cid:12) Eb is a core. Thus (Cc(K))E is dense in E in the graph topology. The representation  is nondegenerate, and the associated representation of A is π. So π comes from a representation of C0(K). The uniqueness of  means that C0(K) is a weak C∗-hull for some class of integrable representations of A. We check this using Proposition 7.11. For q ∈ N (A), the image of Ab in M(Kq) contains Kq and hence is strictly dense. This implies that the image of Ab in M(C0(K)) is strictly dense. So C0(K) is a weak C∗-hull for a class of representations of A by Proposition 7.11 It is even a C∗-hull because the (cid:3) class of locally bounded representations is admissible by Corollary 7.9. 8. Commutative C∗-hulls Let A be a commutative ∗-algebra. We are going to describe all commutative weak C∗-hulls for A. Actually, we describe all locally bounded weak C∗-hulls, and these turn out to be the same as the commutative ones. We study when a C∗-hull satisfies the (Strong) Local–Global Principle and when the class of all locally bounded representations has a C∗-hull. We compare the class of locally bounded representations with the class of representations defined by requiring all a ∈ Ah to act by a regular, self-adjoint operator. Proposition 8.1. Let A be a ∗-algebra and let B = C0(X) be a commuta- tive C∗-algebra. Any representation of A on B has Cc(X) as a core and is lo- cally bounded. There is a natural bijection between representations of A on B, ∗-homomorphisms A → C(X), and continuous maps A → X. Proof. Let (B, µ) be a representation. Since B is dense in B, for any x ∈ X there is f ∈ B with f(x) 6= 0. Then there is an open neighbourhood of x on which f is non-zero. A compact subset K of X may be covered by finitely many such open neighbourhoods. This gives finitely many functions f1, . . . , fn ∈ B so that hence B contains Cc(X). There is an approximate unit (ui)i∈I for C0(X) that If b ∈ B, then lim µ(a)bui = µ(a)b for all a ∈ A. That is, belongs to Cc(X). lim bui = b in the graph topology. Since bui ∈ Cc(X), Cc(X) is a core for (B, µ). Given a ∈ A, we define a function fa : X → C by fa(x) := (µ(a)b)(x) · b(x)−1 for any b ∈ Cc(X) with b(x) 6= 0. This does not depend on the choice of b, and fa is continuous in the open subset where b 6= 0. Thus fa ∈ C(X). The map A → C(X), a 7→ fa, is a ∗-homomorphism. Conversely, any ∗-homomorphism A → C(X) gives a representation of A on C0(X) with core Cc(X) by µ(a)b = fa · b for all a ∈ A, b ∈ Cc(X). The maps that go back and forth between representations on C0(X) and ∗-homomorphisms A → C(X) are inverse to each other. A ∗-homomorphism f : A → C(X) gives a continuous map X → A by mapping x ∈ X to the character a 7→ f(a)(x). Conversely, a continuous map g : X → A induces a ∗-homomorphism g∗ : A → C(X), g∗(a)(x) := g(x)(a), and these two constructions are inverse to each other. P fi · fi(x) > 0 for all x ∈ K. This sum again belongs to the right ideal B, and 38 RALF MEYER Let f : X → A be a continuous map. Then f maps compact subsets in X to compact subsets of A. If K ⊆ X is compact, then any element in C0(K \ ∂K) ⊆ C0(X) is kॷkf(K)-bounded for the C∗-seminorm on A associated to the compact subset f(K) ⊆ A. Thus all elements in Cc(X) are bounded. Since Cc(X) is a core for the representation of A associated to f, this representation is locally bounded. (cid:3) Theorem 8.2. Let A be a commutative ∗-algebra, let B = C0(X) be a commutative C∗-algebra, let f : X → A be a continuous map, and let (B, µ) be the corresponding representation of A on B. Call a representation of A on a Hilbert module E X-integrable if it is isomorphic to (B, µ) ⊗ E for a representation  of B on E. The following are equivalent: (1) f : X → A is injective; (2) B is a weak C∗-hull for the X-integrable representations; (3) B is a C∗-hull for the X-integrable representations. Furthermore, any locally bounded weak C∗-hull of A is commutative. Proof. If f is not injective, then there are x 6= y in X with f(x) = f(y). The evaluation maps at x and y are different 1-dimensional representations of B that induce the same representation of A. Hence the condition (1) in Proposition 3.8 is violated and so B is not a weak C∗-hull. Conversely, assume that f is injective. The representation of A on B associated to f is locally bounded by Proposition 8.1 and hence induces a representation of the unital C∗algebra Cb( A, τc) of bounded elements in A ∼= C( A, τc), see Proposition 7.2. Explicitly, this representation composes functions with f. Since f is injective, D := f∗(Cb( A, τc)) ⊆ Cb(X) separates the points of X. We show that D is strictly dense in Cb(X) ∼= M(B). If K ⊆ X is compact, then the image of f∗(Cb( A, τc))K in the quotient C(K) of C0(X) separates the points of K. Since this image is again a C∗-algebra, it is equal to C(K) by the Stone–Weierstrass Theorem. Let f ∈ Cb(X). For any compact subset K ⊆ X, there is dK ∈ D with dKK = f. By functional calculus, we may arrange that kdKk∞ ≤ kfk. The net (dK) indexed by the directed set of compact subsets K ⊆ X is uniformly bounded and converges towards f in the topology of uniform convergence on compact subsets. Hence it converges towards f in the strict topology (compare [11, Lemma A.1]). This finishes the proof that f∗(Ab) is strictly dense in M(C0(X)). Proposition 7.11 shows that B is a weak C∗-hull for the B-integrable representations of A. Any X-integrable representation of A is locally bounded. The class Repb(A) of locally bounded representations of A is admissible by Corollary 7.9. Hence the smaller class of X-integrable representations inherits the equivalent conditions (2)–(4) in Proposition 3.8. Thus C0(X) is even a C∗-hull. Let B with the universal representation (B, µ) be a locally bounded weak C∗-hull. Then the image of Cb( A, τ0) in the multiplier algebra of B is strictly dense by Proposition 7.11. Thus M(B) is commutative, and then so is B. Thus a locally bounded weak C∗-hull is commutative. (cid:3) Theorem 8.3. Let A be a commutative ∗-algebra, let B = C0(X) be a commutative C∗-algebra, and let f : X → A be an injective continuous map. Let Repint(A, X) be the class of X-integrable representations. The following statements are equivalent if A is metrisable: (1) f : X → A is a homeomorphism onto its image; (2) Repint(A, X) is defined by submodule conditions; (3) Repint(A, X) satisfies the Strong Local–Global Principle; (4) Repint(A, X) satisfies the Local–Global Principle; REPRESENTATIONS BY UNBOUNDED OPERATORS 39 lim f(xn) = f(x) for a sequence (xn)n∈N in X and x ∈ X, then al- (5) if ready lim xn = x. The implications (1)⇒(2)⇒(3)⇒(4)⇒(5) hold without assumptions on A. I do not know whether (1)–(4) are equivalent in general. The condition (5) is there to allow to go back from (4) to (1) at least for metrisable A. a∈Ah Proof. First we check (5)⇒(1) if A is metrisable. If f is not a homeomorphism onto its image, then there is a subset U ⊆ X that is open, such that f(U) is not open in the subspace topology on f(X) ⊆ A. Since A is metrisable, there is x ∈ U and a sequence in f(X) \ f(U) that converges towards f(x). This lifts to a sequence (xn)n∈N in X\U such that lim f(xn) = f(x). We cannot have lim xn = x because xn never enters the open neighbourhood U of x. The implication (2)⇒(3) is Theorem 5.9, and (3)⇒(4) is trivial. We are going to verify (1)⇒(2) and (4)⇒(5). This will finish the proof of the theorem. Assume (1). Let π be a representation in Repint(A, X). Then π is locally bounded, and the operators π(a) for a ∈ Ah are regular and self-adjoint by Proposition 7.8. Furthermore, their Cayley transforms belong to the image of Ab ∼= Cb( A, τc), which is commutative. Hence the operators π(a) for a ∈ Ah strongly commute with each other. The class Repint(A) of representations of A with the property that all π(a), a ∈ Ah, are regular and self-adjoint and strongly commute with each other is defined Let Y :=Q by submodule conditions by Examples 5.10 and Example 5.12. S1. Given a representation in Repint(A), there is a unique rep- resentation : C(Y ) → B(E) that maps the ath coordinate projection to the Cayley transform of π(a). We map A to Y by sending χ ∈ A to the point (cχ(a))a∈Ah ∈ Y . Here cχ(a) is the Cayley transform of the number χ(a) ∈ R or, equivalently, the value of the Cayley transform of the unbounded function a ∈ C( A) at χ. This is a homeomorphism onto its image because for a net of characters (χi) and a character χ on A, we have lim χi(a) = χ(a) if and only if lim cχi(a) = cχ(a). Thus the composite map X → A → Y is a homeomorphism onto its image as well. This forces the image to be locally closed because Y is compact and X locally compact, and a subspace of a locally compact space is locally compact if and only if its underlying subset is locally closed (see [2, I.9.7, Propositions 12 and 13]). Let X ⊆ Y be the closure of the image of X in Y . Then X is open in X. All representations in Repint(A) carry a unital ∗-homomorphism C(Y ) → B(E). Asking for this to factor through the quotient C(X) of C(Y ) is a submodule condition as in Example 5.14. Asking for the induced ∗-homomorphism C(X) → B(E) to remain nondegenerate on C0(X) is another submodule condition as in Example 5.14. The class Rep0(A) defined by these two more submodule conditions is weakly admissible by Lemma 5.8. The universal X-integrable representation belongs to Rep0(A); by weak admissibility, this is inherited by all X-integrable representations. Conversely, we claim that any representation in Rep0(A) is X-integrable. If π ∈ Rep0(A), then the unital ∗-homomorphism C(Y ) → B(E) descends to a nondegenerate ∗-homomorphism : C0(X) → B(E). By construction, the extension of  to multipliers maps the Cayley transform of f∗(a) ∈ C(X) for a ∈ Ah to the Cayley transform of π(a). Let π0 be the X-integrable representation of A associated to . The regular, self-adjoint operators π0(a) and π(a) have the same Cayley transform for all a ∈ Ah. Hence π0(a) = π(a) for all a ∈ Ah. The subset Ah is a strong generating set for A by Example 2.8. Hence Proposition 2.9 gives π0 = π. Thus Repint(A, X) is the class of representations defined by the submodule conditions above. This finishes the proof that (1)⇒(2). 40 RALF MEYER Now we prove (4)⇒(5) by contradiction. Let (xn)n∈N and x be as in (5). Let ¯N = N ∪ {∞} be the one-point compactification of N and view the sequence (xn) and x as a map ξ : ¯N → X. This map is not continuous, but composition with f gives a continuous map ¯N → A. Hence Proposition 8.1 gives a representation (D, µ) of A on C(¯N). This is not X-integrable because the map ¯N → X is not continuous. We claim, however, that the representation (D, µ) ⊗ H is X-integrable for any GNS-representation  on a Hilbert space H. A state on C(¯N) is the same as a Radon measure on ¯N. Since ¯N is countable, any Radon measure is atomic. Thus the resulting GNS-representation is a direct sum of irreducible representations associated to characters. Each character on C(¯N) gives an X-integrable representation because ξ(¯N) ⊆ f(X). Hence (D, µ) is a counterexample to the Local–Global Principle. (cid:3) So (4) cannot hold if (5) fails. Example 8.4. Let A = C[x], so that A = R. Let X be R with the discrete topology, and let f : X → R be the identity map. This is a continuous bijection, but not open. Hence the class of X-integrable representations violates the Local–Global Principle by Theorem 8.3. Nevertheless, C0(X) is a C∗-hull for the class of X-integrable representations of A by Theorem 8.2. An X-integrable representation of A is integrable as in Theorem 4.4, and so it comes from a single regular, self-adjoint operator T := π(x). The representation of C[x] associated to T is X-integrable λ∈R Eλ, where Eλ := {ξ ∈ E T ξ = λξ} for λ ∈ R is the if and only if E = L λ-eigenspace of T. Another example of a C∗-hull for C[x] where X → R is bijective but not a homeomorphism onto its image is discussed in Example 4.9. Theorem 8.5. There is a C∗-hull for Repb(A) if and only if the compactly generated topology τc on A is locally compact, and then the C∗-hull is C0( A, τc). Proof. Assume first that ( A, τc) is locally compact. The pro-C∗-algebra completion A that acts on locally bounded representations of A is C( A, τc) by Proposition 7.2. The primitive ideal space of C(K) for a compact subspace K ⊆ A is simply K, and kakp = a(p) for a ∈ C( A, τc) and p ∈ Prim C(K) ∼= K. Therefore, a function f ∈ C( A, τc) vanishes at ∞ in the sense of Definition 7.14 if and only if it vanishes at ∞ in the usual sense. The subalgebra C0(A) = C0( A, τc) is dense in A because τc is locally compact. Now Theorem 7.16 shows that C0(A) = C0( A, τc) is a C∗-hull for the class of locally bounded representations of A. Conversely, let B be a (weak) C∗-hull for the locally bounded representations of A. Then B is commutative by Theorem 8.2. Let Y be the spectrum of B. The representation of A on B ∼= C0(Y ) corresponds to a continuous map f : Y → A by Proposition 8.1. Let D = C0(X) be a commutative C∗-algebra. Any representation of A on D is locally bounded. So the bijection Repb(A, D) ∼= Rep(B, D) is a bijection between the spaces of continuous maps X → A and X → Y . More precisely, this bijection is composition with f. For the one-point space X, this bijection says that f : Y → A is bijective. The bijection for all compact X means that f becomes a homeomorphism if we replace the topologies on Y and A by the associated compactly generated ones. The topology on Y is already compactly generated because Y is locally compact. Hence f is a homeomorphism from Y to ( A, τc). So τc is locally compact. (cid:3) Let Repint(A) be the class of all representations with the property that π(a) is regular and self-adjoint for all a ∈ Ah. We are going to compare Repint(A) and Repb(A). Proposition 7.8 gives Repb(A) ⊆ Repint(A). REPRESENTATIONS BY UNBOUNDED OPERATORS 41 Theorem 8.6. The class Repint(A) is admissible and defined by submodule con- ditions. Hence it satisfies the Strong Local–Global Principle. The operators π(a) for a ∈ Ah strongly commute for all π ∈ Repint(A). Let S ⊆ Ah be a strong generating set for A. If π(a) is regular and self-adjoint for all a ∈ S, then already π ∈ Repint(A). Proof. The class RepS(A) of representations defined by requiring π(a) to be regular and self-adjoint for all a ∈ S for a strong generating set S is admissible and defined by submodule conditions by Theorem 5.17. The class Repint(A) is defined by submodule conditions as well by Example 5.10. So is the subclass Rep0(A) of all representations in Repint(A) for which the operators π(a) for all a ∈ Ah strongly commute (Example 5.12). Hence our three classes of representations satisfy the Strong Local–Global Principle by Theorem 5.9. The classes Repint(A) and Rep0(A) have the same Hilbert space representa- tions by [27, Theorem 9.1.2]. Since S is a strong generating set, the domain a∈S dom(π(a)) by (2.10). This contains a∈S,n∈N dom(π(a)n). Now [27, Theorem 9.1.3] shows that RepS(A) and Repint(A) contain the same Hilbert space representations. Since our three classes of represen- tations satisfy the (Strong) Local–Global Principle and have the same Hilbert space (cid:3) representations, they are equal. Theorem 8.7. If A is commutative and countably generated, then of any representation π in RepS(A) is T T Repint(A) = Repb(A). Proof. Proposition 7.8 gives Repb(A) ⊆ Repint(A). Conversely, let (E, π) be a representation on a Hilbert module E in Repint(A); that is, π(a) is regular and self-adjoint for each a ∈ A. Let (ai)i∈N be a countable generating set for A. We may i for all i ∈ N and that (ai) is a basis assume without loss of generality that ai = a∗ for A and hence a strong generating set. Let ξ ∈ E. We are going to approximate ξ by bounded vectors for π. This will show that π is locally bounded. For each i ∈ N, there is a canonical homomorphism αi : C[x] → A mapping x 7→ ai. The closure of π ◦ αi is an integrable representation of C[x] as in condition (2) in Theorem 4.4. Hence it corresponds to a representation i : C0(R) → B(E), the functional calculus of π(ai). The operators π(a) for a ∈ Ah strongly commute by Theorem 8.6. Thus the Cayley transform of ai commutes with π(a) and, in particular, maps the domain of π(a) to itself. The same remains true for i(f) for all f ∈ C0(R) because we get them by the (bounded) functional calculus for the Cayley transform of π(ai). So i(f)(E) ⊆ E by (2.6) and i(f)π(a) = π(a)i(f) for all f ∈ C0(R), a ∈ A as operators on E. Now we show that π ◦ αi is locally bounded. If f ∈ Cc(R) is supported in a compact subset K ⊆ R, then kπ(h(ai))i(f)ξk = ki(h · f)ξk ≤ C sup{h(x) x ∈ K} for all h ∈ C[x]; thus i(f)ξ is bounded for the representation π ◦ αi. There is an approximate unit (fn) for C0(R) that lies in Cc(R). Then lim i(fn)ξ = ξ for all ξ ∈ E, even in the graph topology for π because π(a)i(fn)ξ = i(fn)π(a)ξ for all a ∈ A, fn ∈ C0(R), ξ ∈ E. Therefore, the bounded vectors of the form (f)ξ with f ∈ Cc(R), ξ ∈ E form a core for π ◦ αi. So π ◦ αi is locally bounded. We now refine this construction to approximate ξ by bounded vectors for the whole representation π. We construct i as above. Fix i, k ∈ N and let ξ0 := 0 ≤ fi,k ≤ 1 and ki(fi,k)ξ0 − ξ0k < 2−k. Thus ki(fi,k)ξ − ξkaj < 2−k in the graph norm for aj for 0 ≤ j ≤ k. For k, l ∈ N, let ξk,l := 0(f0,k)1(f1,k+1)··· l(fl,k+l)ξ. k)(cid:1)ξ ∈ E. The argument above gives fi,k ∈ Cc(R) with (cid:0)1 + π(a2 0) + ··· + π(a2 42 RALF MEYER The operators i(fi,j) are norm-contracting, map E into itself, and commute with each other and with the unbounded operators π(a) for all a ∈ A. Hence ≤ dX i=1 ≤ dX i=1 kξk,l − ξk,l+dkaj kξk,l+i−1 − ξk,l+ikaj kl+i(fl+i,k+l+i)ξ − ξkaj ≤ dX 2−k−l−i = 2−k−l i=1 for all k, l, d ∈ N, 0 ≤ j ≤ k + l + 1. Since we assumed (aj) to be a strong generating set, the graph norms for aj generate the graph topology. So the estimate above shows that (ξk,l)l∈N with fixed k is a Cauchy sequence in E in the graph topology. Thus it converges to some ξk ∈ E. Letting ξk,−1 := ξ, the above estimate remains true for l = −1 and gives kξk,l − ξkaj ≤ 2−k+1 for all j ≤ k, uniformly in l ∈ N. This implies kξk − ξkaj ≤ 2−k+1 for j ≤ k, so that lim ξk = ξ in the graph topology. It remains to show that each ξk is a bounded vector. Fix k, i ∈ N and let b ∈ A. Choose Ri > 0 so that fi,k+i is supported in [−Ri, Ri]. If l ≥ i, then π(b)ξk,l ∈ i(C0(−Ri, Ri))E because i(fi,k+i) occurs in the definition of ξk,l. As above, this implies kπ(ai)π(b)ξkk ≤ Rikπ(b)ξkk for all b ∈ A. Thus q(a) := sup b∈A kπ(a)π(b)ξkk kπ(b)ξkk is finite for a = ai. Since ai is a basis for A and q is subadditive, we get q(a) < ∞ for all a ∈ A. Since q(a) is the operator norm of π(a)π(A)ξk, it is a C∗-seminorm on A. By construction, kπ(a)ξkk ≤ q(a) for all a ∈ A, that is, ξk is bounded. (cid:3) Proposition 8.8. If Repint(A) has a weak C∗-hull, then Repint(A) = Repb(A). Proof. Let B with the universal representation (B, µ) be a weak C∗-hull for Repint(A). First we claim that B is commutative. Let ω : B ,→ B(H) be a faithful representation. This corresponds to an integrable representation π of A. Since the equivalence Repint(A,H) ∼= Rep(B,H) is compatible with unitary ∗-intertwiners, the commutant of ω(B) is the C∗-algebra of ∗-intertwiners of π by Proposition 3.3. The commutant of this is a commutative von Neumann algebra by [27, Theorem 9.1.7]. So the bicommutant of ω(B) is commutative. This forces B to be commutative. Any representation of A on a commutative C∗-algebra is locally bounded by Theorem 8.2. If the universal representation for Repint(A) is locally bounded, then all representations in Repint(A) are locally bounded, so that Repint(A) = Repb(A). Thus Repint(A) only has a weak C∗-hull if Repint(A) = Repb(A). (cid:3) Example 8.9. Let A be the ∗-algebra C[(xi)i∈N] of polynomials in countably many symmetric generators. Then A ∼=QN R with the product topology. This is metris- able. So τc is the usual product topology. Since this is not locally compact, Repb(A) has no C∗-hull, not even a weak one (Theorem 8.5). Since A is countably generated, Repint(A) = Repb(A) by Theorem 8.7. A commutative (weak) C∗-hull for some class of representations of A is equivalent to an injective, continuous map X → A for a locally compact space X by Theorem 8.2. Let G be a topological group. A host algebra for a G is defined in [12] as a C∗-algebra B with a continuous representation λ of G by unitary multipliers, such that for each Hilbert space H, the map that sends a representation : B → B(H) to a unitary representation  ◦ λ of G is injective. We claim that commutative C∗-hulls for the polynomial algebra C[(xi)i∈N] are equivalent to host algebras of the topological group R(N) :=LN R. REPRESENTATIONS BY UNBOUNDED OPERATORS 43 Let C∗(Gd) be the C∗-algebra of G viewed as a discrete group. Representations of C∗(Gd) are equivalent to representations of the discrete group underlying G by unitary multipliers. Since any representation of C∗(Gd) is bounded, any weakly admissible class of representations of C∗(Gd) is admissible by Corollary 7.9. Call a representation of C∗(Gd) continuous if the corresponding representation of G is continuous. This class is easily seen to be weakly admissible, hence admissible. The unital ∗-homomorphism C∗(Gd) → M(B) associated to the unitary representation λ for a host algebra B is continuous by assumption. Thus B-integrable representations of C∗(Gd) are continuous. The injectivity requirement in the definition of a host algebra is exactly the condition (1) in Proposition 3.8, and this is equivalent to B being a C∗-hull. Thus a host algebra for G is the same as a C∗-hull or weak C∗-hull for a class of continuous representations of C∗(Gd). In applications, we would rather study continuous representations of G through the Lie algebra of G instead of through the inseparable C∗-algebra C∗(Gd). The Lie algebra of G = R(N) is the Abelian Lie algebra R(N), and its universal enveloping algebra is the polynomial algebra A = C[(xi)i∈N]. Call a representation of A integrable if it belongs to Repint(A) = Repb(A). Let E be a Hilbert module. We claim that an integrable representation of A on E is equivalent to a strictly continuous, unitary representation of the group R(N) on E. Indeed, a unitary representation of R is equivalent to a representation of C∗(R) ∼= C0(R), and these are equivalent to integrable representations of C[x] as in Theorem 4.4. In an integrable representation of C[(xi)i∈N], the operators π(xi) for i ∈ N strongly commute by Theorem 8.6. Hence the resulting representations of C0(R) commute. Equivalently, the resulting continuous representations of R commute, so that we may combine them to a representation of the Abelian group R(N). Conversely, a continuous unitary representation of R(N) provides nondegenerate representations of C0(Rm) for all m ∈ N by restricting the representation to Rm ⊆ R(N). These correspond to a compatible family of representations of the polynomial algebras C[x1, . . . , xm] for m ∈ N. The intersection of their domains is dense by [27, Lemma 1.1.2]. So these representations combine to a representation of A = C[(xi)i∈N]. Hence an integrable representation of A on a Hilbert module as in Theorem 8.6 is equivalent to a continuous representation of R(N). 9. From graded ∗-algebras to Fell bundles Let G be a discrete group with unit element e. sum decomposition A =L Definition 9.1. A G-graded ∗-algebra is a unital algebra A with a linear direct g = Ag−1, and 1 ∈ Ae for all g, h ∈ G. Thus Ae ⊆ A is a unital ∗-subalgebra. g∈G Ag with Ag · Ah ⊆ Agh, A∗ The articles [7,26] study many examples of G-graded ∗-algebras. We fix some notation used throughout this section. Let E be a Hilbert module over a C∗-algebra D. Let (E, π) be a representation of A on E. Let πg : Ag → EndD(E) g∈G πg. Since π is a ∗-homomorphism, for g ∈ G be the restrictions of π, so π =L πg(ag)πh(ah) = πgh(ag · ah), πg−1(a∗ g) ⊆ πg(ag)∗ for all ag ∈ Ag, ah ∈ Ah. The last condition means that hξ, πg(ag)ηi = hπg−1(a∗ for all ξ, η ∈ E. In particular, πe : Ae → End(E) is a representation of Ae. Lemma 9.2 (compare [26, Lemma 12]). The families of norms kξka := kπ(a)ξk for a ∈ A and for a ∈ Ae generate equivalent topologies on E. Hence the represen- tation πe : Ae → EndD(E) is closed if and only if π is closed. g)ξ, ηi 44 RALF MEYER Proof. Any element of A is a sum a = P many non-zero terms. We estimate kξka ≤P g∈G ag with ag ∈ Ag and only finitely by gag ∈ Ae, the graph topologies for πe and π are (cid:3) the proof of Lemma 2.2. Since a∗ equivalent. 9.1. Integrability by restriction. Definition 9.3. Let a weakly admissible class of integrable representations of Ae on Hilbert modules be given. We call a representation of A on a Hilbert module integrable if its restriction to Ae is integrable. , and kξkag g∈Gkξkag ≤ 5 4kξka∗ gag Here "restriction of π" means the representation πe with the same domain E as π. This is closed by Lemma 9.2. Proposition 9.4. If integrability for representations of Ae is defined by submodule conditions, then the same holds for A. If the Local–Global Principle holds for the integrable representations of Ae, it also holds for the integrable representations of A. If the class of integrable representations of Ae is admissible or weakly admissible, the same holds for A. Proof. The first two statements and the claim about weak admissibility are trivial because integrability for a representation of A only involves its restriction to Ae. Lemma 9.2 shows that restriction from A to Ae does not change the domain. Hence (2) in Proposition 3.8 is inherited by A if it holds for Ae. That is, admissibility (cid:3) of the integrable representations passes from Ae to A. It is unclear whether A also inherits the Strong Local–Global Principle from Ae. This may often be bypassed using Theorem 5.9. 9.2. Inducible representations and induction. Let F be a Hilbert D-module and let F ⊆ F and ϕe : Ae → EndD(F) be a representation of Ae on F. We try to induce ϕe to a representation of A as in [26]. Thus we consider the algebraic tensor product A (cid:12) F and equip it with the obvious right D-module structure and the unique sesquilinear map that satisfies ha1 ⊗ ξ1, a2 ⊗ ξ2i = δg,hhξ1, ϕe(a∗ 1a2)ξ2i for all g, h ∈ G, a1 ∈ Ag, a2 ∈ Ah, ξ1, ξ2 ∈ F. This map is sesquilinear and descends to the quotient space A (cid:12)Ae F. It is symmetric and D-linear in the sense that hx, yi = hy, xi∗ and hx, ydi = hx, yid. Let π be the action of A on A (cid:12)Ae F by left multiplication. This representation is formally a ∗-homomorphism in the sense that hx, π(a)yi = hπ(a∗)x, yi for all a ∈ A, x, y ∈ A(cid:12)Ae F. The only thing that is missing to get a representation of A on a Hilbert D-module is positivity of the inner product. This requires a subtle extra condition. Proposition 9.5. The following are equivalent: (1) the sesquilinear map on A (cid:12)Ae F defined above is positive semidefinite; Pn (2) for all g ∈ G, n ∈ N and all a1, . . . , an ∈ Ag, ξ1, . . . , ξn ∈ F, the element k,l=1hξk, ϕe(a∗ (cid:0)hξk, ϕe(a∗ (3) for all g ∈ G, n ∈ N and all a1, . . . , an ∈ Ag, ξ1, . . . , ξn ∈ F, the matrix kal)ξli ∈ D is positive; k,l ∈ Mn(D) is positive. kal)ξli(cid:1) Proof. The condition (2) for fixed g ∈ G says that the sesquilinear map on Ag (cid:12)Ae F is positive semidefinite. Since the subspaces Ag (cid:12)Ae F for different g are orthogonal, this is equivalent to positive semidefiniteness on A (cid:12)Ae F. Thus (1) ⇐⇒ (2). We prove (2) ⇐⇒ (3). Fix g ∈ G, n ∈ N, a1, . . . , an ∈ Ag and ξ1, . . . , ξn ∈ F. Let y = (ykl) ∈ Mn(D) be the matrix in (3). By [16, Lemma 4.1], y ≥ 0 in Mn(D) ⊆ B(Dn) if and only if hd, ydi ≥ 0 for all d = (d1, . . . , dn) ∈ Dn. That is, Pn k,l=1 d∗ REPRESENTATIONS BY UNBOUNDED OPERATORS 45 kykldl ≥ 0 for all d1, . . . , dn ∈ D. Since F is a right D-module, this (cid:3) condition for all ξi ∈ F, di ∈ D is equivalent to (2). Definition 9.6. A representation ϕe of Ae is inducible (to A) if it satisfies the equivalent conditions in Proposition 9.5. If Ae were a C∗-algebra, it would be enough to assume hξ, ϕe(a∗a)ξi ≥ 0 for all g ∈ G, a ∈ Ag, ξ ∈ F, which amounts to the condition a∗a ≥ 0 in Ae for all g ∈ G, a ∈ Ag. This is part of the definition of a Fell bundle over a group. For more general ∗-algebras, the positivity conditions for different n ∈ N in Proposition 9.5 may differ, compare [9]. Let A ⊗Ae F be the Hilbert module completion of A (cid:12)Ae F for the inner product above. The ∗-algebra A acts on A(cid:12)Ae F by left multiplication, a1·(a2⊗ξ) := (a1a2)⊗ξ for a1, a2 ∈ A, ξ ∈ F. As in the proof of Lemma 2.18, this module structure descends to the image of A (cid:12)Ae F in A ⊗Ae F and gives a well defined representation π of A on A⊗Ae F. Its closure is called the induced representation from ϕe, and its domain is denoted by A ⊗Ae F. g∈G Ag (cid:12)Ae F is Ae-invariant and orthogonal for the above inner product. Hence (9.7) The decomposition A (cid:12)Ae F =L A ⊗Ae F ∼=M Ag ⊗Ae F, g∈G where Ag ⊗Ae F is the closure of the image of Ag (cid:12)Ae F or, equivalently, the Hilbert D-module completion of Ag(cid:12)Ae F with respect to the restriction of the inner product. Each summand Ag⊗Ae F carries a closed representation of Ae with domain Ag⊗Ae F, and πAe is the direct sum of these representations. Lemma 9.8. Let π be any representation of A. Then πAe is inducible. Proof. For g ∈ G, a1, . . . , an ∈ Ag, ξ1, . . . , ξn ∈ E, let y :=Pn k=1 π(ak)ξk. Then hξk, πAe(a∗ kal)ξli = hπ(ak)ξk, π(al)ξli = hy, yi ≥ 0. (cid:3) nX k,l=1 nX k,l=1 Lemma 2.24 about the associativity of ⊗ has a variant for induction: Lemma 9.9. Let D1, D2 be C∗-algebras, let E be a Hilbert D1-module and let F be a C∗-correspondence between D1, D2. Let (ϕe, E) be an inducible representation of A on E. Then the representation ϕe ⊗D1 F on E ⊗D1 F is inducible and there is a canonical unitary ∗-intertwiner of representations of A, (A ⊗Ae E) ⊗D1 F ∼= A ⊗Ae (E ⊗D1 F). a1, . . . , an ∈ Agi, and ω1, . . . , ωn ∈ E⊗D1F. Let ζ :=Pn Proof. Let E ⊗D1 F ⊆ E ⊗D1 F be the domain of ϕe ⊗D1 F. Let g1, . . . , gn ∈ G, k=1 ak⊗ωk ∈ A(cid:12)(E⊗D1F). To show that ϕe ⊗D1 F is inducible, we must prove that hζ, ζi ∈ D2 is positive. Vectors in E(cid:12)F form a core for ϕe⊗D1 F by construction. Hence there is a sequence of vectors of the form 'jX ωj,τ := ξτ,j,i ⊗ ητ,j,i, ξτ,j,i ∈ E, ητ,j,i ∈ F, which, for τ → ∞, converges to ωj in the graph norms of the elements δgm,gk a∗ mak ∈ i=1 Ae for all m, k = 1, . . . , n. Let ζτ :=Pn τ→∞hζτ , ζτi = lim lim j=1 aj ⊗ ωj,τ. Then τ→∞hζτ , ζi = lim τ→∞hζ, ζτi = hζ, ζi 46 in norm and hζτ , ζτi = *X i,j RALF MEYER + X δgj ,gkhητ,j,i,hξτ,j,i, ϕe(a∗ ak ⊗ ξτ,k,m ⊗ ητ,k,m m,k aj ⊗ ξτ,j,i ⊗ ητ,j,i, = X i,j,k,m j ak)ξτ,k,miD1 · ητ,k,miD2. 0 ≤ nX k,l=1 This is also the inner product of ζτ with itself in the tensor product (A ⊗Ae E) ⊗ F. This is positive because ϕe is inducible and the usual tensor product of the Hilbert D1-module A ⊗Ae E with the D1, D2-correspondence F is a Hilbert D2-module. Hence hζτ , ζτi ≥ 0 for all τ. Since the positive elements in D2 form a closed subset, this implies hζ, ζi ≥ 0. Thus ϕe ⊗D1 F is inducible. The argument above also shows that the linear span of vectors of the form a ⊗ ξ ⊗ η with a ∈ A, ξ ∈ E, η ∈ F is a core for the representation of Ae on A ⊗Ae (E ⊗D1 F). Such vectors also form a core for the representation of A on (A ⊗Ae E) ⊗D1 F. The left actions of A and the D2-valued inner products coincide on such vectors. Hence there is a unique unitary ∗-intertwiner that maps the image of a ⊗ ξ ⊗ η in (A ⊗Ae E) ⊗D1 F to its image in A ⊗Ae (E ⊗D1 F). (cid:3) 9.3. C∗-Hulls for the unit fibre. We assume that the chosen class of integrable representations of Ae has a (weak) C∗-hull Be. We want to construct a Fell bundle whose section C∗-algebra is a (weak) C∗-hull for the integrable representations of A. At some point, we need Be to be a full C∗-hull (compatible with isometric intertwiners) and one more extra condition. But we may begin the construction without these assumptions. First we build the unit fibre B+ e of the Fell bundle. It is a (weak) C∗-hull for the inducible, integrable representations of Ae. Let (Be, µe) be the universal integrable representation of Ae on Be. Let x− for a self-adjoint element x in a C∗-algebra denote its negative part. Definition 9.10. Let B+ ideal generated by elements of the form e be the quotient C∗-algebra of Be by the closed two-sided (9.11) k · µe(a∗ b∗ kal) · bl for g ∈ G, a1, . . . , an ∈ Ag, b1, . . . , bn ∈ Be. (cid:18) nX k,l=1 e be the image of Be in B+ Let B+ representation of Ae on this quotient. e and let µ+ e : Ae → EndB (B+ e ) be the induced + e The following proposition shows that the representation (B+ e , µ+ e ) of Ae on B+ e is the universal inducible, integrable representation of Ae. Proposition 9.12. Let (F, ϕe) be an integrable representation of Ae on a Hilbert module F. Let ¯ϕe : Be → B(F) be the corresponding representation of Be. Then ϕe is inducible if and only if ¯ϕe factors through the quotient map Be (cid:16) B+ e . Thus B+ is a C∗-hull for the inducible, integrable representations of Ae. Proof. Assume first that ϕe is inducible. Let ξ ∈ F and let g ∈ G, a1, . . . , an ∈ Ag and b1, . . . , bn ∈ Be be as in (9.11). Let ξk := ϕe(bk)ξ. Since ϕe is inducible, Proposition 9.5 implies e (cid:19) − nX hξk, ϕe(a∗ kal)ξli = hξ, ¯ϕe(bk)∗ϕe(a∗ k,l=1 = kal) ¯ϕe(bl)ξi *  nX ξ, ¯ϕe +  ξ . b∗ kµe(a∗ kal)bl k,l=1 (cid:16)Pn Equivalently, ¯ϕe annihilates the negative part ofPn Since ξ ∈ F is arbitrary, this means that ¯ϕe to a homomorphism on the quotient B+ is inducible by Proposition 9.5. If ¯ϕ+ e : B+ 1F ∼= ϕe on B+ e ⊗B representation µ+ is, ϕe is inducible if ¯ϕe factors through the quotient map Be (cid:16) B+ e . e ⊗ ¯ϕ + e + e k,l=1 bkµe(a∗ k,l=1 bkµe(a∗ kal)bl kal)bl. So ¯ϕe descends e ) e , µ+ e → B(F) is a representation, then the F ∼= F is inducible by Lemma 9.9. That e . Conversely, the representation (B+ (cid:17) ≥ 0 in B(F). 47 REPRESENTATIONS BY UNBOUNDED OPERATORS Summing up, the representation ¯ϕe associated to an integrable representation ϕe of Ae descends to B+ e if and only if ϕe is inducible. The quotient map induces a fully e , D) ,→ Rep(Be, D). The argument above shows that faithful embedding Rep(B+ its image consists of those representations of Be that correspond to inducible, inte- grable representations of Ae under the correspondence Rep(Be, D) ∼= Repint(Ae, D). e is a (weak) C∗-hull for the class of inducible, integrable representations Hence B+ (cid:3) of Ae. g := Ag ⊗Ae B+ Definition 9.13. Let B+ e -module because the representation (B+ e,g) be g , µ+ the induced representation of Ae on B+ e as a core, e,g(ae)(ag ⊗ b) := (aeag) ⊗ b for ae ∈ Ae, ag ∈ Ag, b ∈ B+ with the representation µ+ e . g are the unique extensions of the following pre-Hilbert module structure on Ag (cid:12)Ae B+ e : (ag ⊗ b1) · b2 := ag ⊗ (b1 · b2) for all ag ∈ Ag, b1 ∈ B+ e (a∗ (9.14) for a1, a2 ∈ Ag, b1, b2 ∈ B+ definition, B+ completion of Ag (cid:12)Ae B+ e . This is positive definite by Proposition 9.12. By g is the e,g of Ae. e . This is a well defined Hilbert B+ e is inducible. Let (B+ g . It has the image of Ag (cid:12)Ae B+ e in the graph topology for the representation µ+ e -module structure and the inner product on B+ g is the norm completion of this pre-Hilbert B+ ha1 ⊗ b1, a2 ⊗ b2i := b∗ e , b2 ∈ B+ 1a2)b2 By definition, the right B+ e -module, and B+ e ) of Ae on B+ e , and e , µ+ 1µ+ The Hilbert B+ The Fell bundle structure on (B+ e -modules B+ g are the fibres of our Fell bundle. g )g∈G only exists under extra assumptions. Before we turn to these, we construct representations of the Hilbert B+ g from an integrable representation π of A on E. Let πg := πAg and let ¯πe : Be → B(E) be the representation of the C∗-hull corresponding to πe. Since πe is inducible by e → B(E) by Proposition 9.12. Lemma 9.8, ¯πe descends to a representation ¯π+ e (b) is defined on all of E e (b) maps E into the domain E of πe, which is also the domain of πg(a) e (b∗)πg−1(a∗), e (a∗a) · b) because ¯π+ by Lemma 9.2. Its adjoint contains the densely defined operator ¯π+ and the operator Let a ∈ Ag and b ∈ B+ e . The operator πg(a)¯π+ e (b∗)πg−1(a∗)πg(a)¯π+ ¯π+ e -modules B+ e (b∗ · µ+ e (b) = ¯π+ e (b) = ¯π+ e : B+ e (b∗)πg−1(a∗) extends to a bounded operator on E, which is e (b∗)πe(a∗a)¯π+ e (b) ∈ B(E). Define is bounded. Hence ¯π+ e (b). Thus πg(a)¯π+ adjoint to πg(a)¯π+ g : Ag (cid:12) B+ ¯π+ e → B(E), a ⊗ b 7→ πg(a)¯π+ e (b). As above, we check that g (x1)∗¯π+ ¯π+ (9.15) for all x1, x2, x ∈ Ag (cid:12) B+ defines B+ g . Thus ¯π+ g (x2) = ¯π+ e (hx1, x2i), g (x · b) = ¯π+ ¯π+ g (x)¯π+ e (b) e , b ∈ B+ e , where the inner product is the one that g extends uniquely to a bounded linear map ¯π+ g : B+ g → B(E), which still satisfies (9.15). That is, it is a representation of the Hilbert module B+ with respect to ¯π+ e . Lemma 9.16. If ¯π+ e ,→ B(E) is faithful (hence isometric), then so is ¯π+ g . e : B+ g 48 Proof. Let ξ ∈ B+ kξk = khξ, ξiB g . Then RALF MEYER e (hξ, ξiB + e )k1/2 = k¯π+ g (ξ)∗¯π+ g (ξ)k1/2 = k¯π+ g (ξ)k. (cid:3) + e k1/2 = k¯π+ Next we want to prove that h (B+ g ) · ¯π+ ¯π+ g (B+ h ) ⊆ ¯π+ g )∗ = ¯π+ g−1) g (B+ g (B+ g (B+ gh(B+ g−1(B+ gh) and ¯π+ g ))g∈G a Fell bundle structure, which would lift to (B+ e ) · ¯π+ (9.17) for all g, h ∈ G and for all integrable representations π of A. This would give (¯π+ g (B+ g )g∈G itself if ¯π+ is faithful. Lemma 9.23 below gives (9.17) provided the closed linear span of g ) for all g ∈ G. But this only holds if we impose two ¯π+ e (B+ g ) is ¯π+ extra assumptions. First, compatibility of integrability and induction gives B+ g a e -module structure. Secondly, compatibility of the weak C∗-hull B+ canonical left B+ with isometric intertwiners ensures that the representation ¯π+ g is compatible with this left B+ 9.4. Integrability and induction. Definition 9.18. We say that integrability is compatible with induction if induction of inducible representations preserves integrability; that is, if ϕe is an inducible, integrable representation of Ae on E and π is the representation of A on A ⊗Ae E induced by ϕe, then the representation πe := πAe of Ae is again integrable. e -module structure. e e We shall use this assumption in Section 9.5 to prove (9.17). But first, we study some sufficient conditions for integrability to be compatible with induction. g , µ+ g , µ+ e,g) of Ae on B+ g are integrable for all g ∈ G. A direct sum of representations is integrable if and only if each summand is integrable by Corollary 3.4. Hence integrability is compatible with induction if and only if an inducible, integrable representation ϕe on F induces integrable representations of Ae on Ag ⊗Ae F for all g ∈ G. Proposition 9.19. Integrability is compatible with induction if and only if the representations (B+ g are integrable for all g ∈ G Proof. The representations (B+ if and only if their direct sum is integrable. Denote this by (A ⊗Ae B+ e , µ+). If integrability is compatible with induction, then (A ⊗Ae B+ e , µ+) must be integrable because it is the induced representation of the universal integrable (inducible) representation (B+ e . Conversely, by Lemma 9.9, induction maps e ) of Ae on B+ e , µ+ e → B(F) the representation (B+ e , µ+ to the representation (A ⊗Ae B+ e , µ+) (cid:3) is, see Definition 3.11.(2). The (Strong) Local–Global Principle is useful to check that integrability is e )⊗F of Ae associated to a representation : B+ e , µ+) ⊗ F, which is integrable if (A ⊗Ae B+ e,g) of Ae on B+ compatible with induction: Proposition 9.20. Assume that the integrable representations of Ae satisfy the Strong Local–Global Principle and that induction maps irreducible, inducible, inte- grable Hilbert space representations of Ae to integrable Hilbert space representations of A. Then integrability is compatible with induction. The same conclusion holds if the integrable representations of Ae satisfy the Local–Global Principle and induction maps all inducible, integrable Hilbert space representations of Ae to integrable Hilbert space representations of A. e ) of Ae be the C∗-hull for the Proof. Let B+ inducible, integrable representations of Ae. By Proposition 9.19, it suffices to prove that the canonical representation of Ae on A ⊗Ae B+ of Ae on (A ⊗Ae B+ By the Strong Local–Global Principle, this follows if the induced representation π e ) ⊗ H is integrable for each irreducible representation  of B+ e with the representation (B+ e is integrable. e , µ+ e REPRESENTATIONS BY UNBOUNDED OPERATORS 49 on a Hilbert space H. The representation  is equivalent to an irreducible, inducible, integrable representation π of Ae on H, and π is the representation induced by π. By assumption, π is integrable. This finishes the proof in the case of the Strong Local–Global Principle. The argument in the other case is the same without the (cid:3) word "irreducible." Proposition 9.21. Assume the following. First, the integrable representations of Ae satisfy the Strong Local–Global Principle. Secondly, all irreducible, inte- grable Hilbert space representations of Ae are finite-dimensional. Third, all finite- dimensional inducible representations of Ae are integrable. And fourth, each Ag is finitely generated as a right Ae-module. Then integrability is compatible with induction. e is a quotient of Be, its irreducible representations form a Proof. First, since B+ subset of the irreducible representations of Be. Thus the irreducible, inducible, integrable Hilbert space representations of Ae are finite-dimensional as well. By Proposition 9.20, it suffices to check that the induced representation of Ae on Ag ⊗Ae H is integrable when H is a Hilbert space with an irreducible, inducible, integrable representation. By our assumptions, H is finite-dimensional and Ag is finitely generated as an Ae-module. Hence Ag ⊗Ae H is finite-dimensional. This representation is a direct summand in a representation of A on A ⊗Ae H and hence inducible by Lemma 9.8. By assumption, the induced representation of Ae on Ag ⊗Ae H is integrable. (cid:3) 9.5. The Fell bundle structure. If integrability is compatible with induction, the g is integrable. It is inducible as well by Lemma 9.8 representation µ+ because it is a direct summand in a representation of A. Hence there is a unique g is a core (nondegenerate) representation ¯µ+ e , x ∈ B+ for µ+ g . g → Our next goal is to show that the representations ¯π+ g : B+ B(E) constructed using (9.15) are compatible in the sense that for all be ∈ B+ (9.22) This is not automatic. The following lemma is the most subtle point in the proof of the Induction Theorem. Lemma 9.23. Equation (9.22) holds if B+ e ) · ¯π+ Then also ¯π+ g (B+ Proof. Let F := B+ g ⊗B b ∈ B+ well defined isometry I : F ,→ E. The representation ¯µ+ representation ¯µ+ the representations ¯µ+ correspond to the integrable representations µ+ respectively. Since B+ representations of Ae. We identify E ∼= B+ e ⊗¯π Then Lemma 9.9 gives a canonical unitary ∗-intertwiner e ⊗B g (b)ξ, for g , ξ ∈ E, preserves the inner products by (9.15). Hence it extends to a g induces a e on F. The meaning of (9.22) is that I intertwines e on F and E. These representations e,g ⊗ 1 and ¯π+ e,g ⊗ 1 and πe of Ae on F and E, e is a C∗-hull, it suffices to prove that I intertwines these e ⊗¯π 1E as in Proposition 3.6. E) ∼= Ag ⊗Ae E F := (Ag ⊗Ae B+ e on B+ e )B+ e (ae)b)x for all a ∈ Ae, b ∈ B+ e → B(E) and ¯π+ E and describe πe as µ+ E ∼= Ag ⊗Ae (B+ g ) = ¯π+ E. The linear map B+ e is a C∗-hull, not just a weak C∗-hull. g (cid:12) E → E, b ⊗ ξ 7→ ¯π+ g ) for all g ∈ G. e,g ⊗ 1E of B+ e,g of B+ e on B+ e,g of B+ e,g(µ+ e,g, and µ+ e,g(ae)(¯µ+ e,g(b)x) = ¯µ+ g such that ¯µ+ e,g(B+ g (bg) = ¯π+ g (¯µ+ e,g(be)bg) e,g of Ae on B+ e , bg ∈ B+ g . + e e ) ⊗B + e e (be) · ¯π+ ¯π+ e (B+ g (B+ + e of representations of Ae. An inspection of the proof shows that I corresponds to the isometry I0 : Ag ⊗Ae E ,→ E defined by I0(a ⊗ ξ) := πg(a)ξ for all a ∈ Ag, ξ ∈ E. Since I0 is an Ae-intertwiner, so is I. This finishes the proof of (9.22). Then (cid:3) ¯π+ e (B+ g ) follows because ¯µ+ e,g is nondegenerate. g ) = ¯π+ e ) · ¯π+ g (B+ g (B+ e : B+ + e + e e of B+ 50 RALF MEYER g (B+ e (B+ e ) · ¯π+ g ) = ¯π+ Lemma 9.24. Assume ¯π+ holds. Proof. We write .= to denote that two sets of operators have the same closed e (B+ linear span. By definition, ¯π+ e ) because B+ g ) implies g ) .= ¯π+ ¯π+ g (B+ e (B+ e ) · ¯π+ e ). We have seen above (9.15) that g ) .= πg(Ag)¯π+ e . Our assumption ¯π+ e ), and ¯π+ g ) .= ¯π+ e (B+ g ) for all g ∈ G. Then (9.17) e )∗ .= ¯π+ g (B+ e is dense in B+ e )∗πg(Ag)¯π+ e (B+ g (B+ e (B+ e (B+ g (B+ g (B+ e (b∗)πg−1(a∗) = ¯π+ ¯π+ e (b)∗πg−1(a∗) e , a ∈ Ag extends to a bounded operator on E that is adjoint to the for b ∈ B+ bounded operator πg(a)¯π+ (cid:0)¯π+ e (b). Therefore, e (b2)(cid:1)∗ = ¯π+ gh). g (B+ e (b1) g (B+ ⊆ ¯π+ e (B+ e (B+ g : B+ gh(B+ e (B+ e ) ¯π+ g (B+ g−1(B+ g−1(B+ h ) .= ¯π+ e (b2)∗πg−1(a∗)¯π+ e (b1)∗πg(a)¯π+ e , a ∈ Ag; both sides are globally defined bounded operators for all b1, b2 ∈ B+ e ) maps E into E. The closed linear spans on the two sides of this because ¯π+ e (B+ g )∗ and ¯π+ equality are ¯π+ g−1). g (B+ e )∗, g ∈ G, a ∈ Ag are bounded and As above, the operators ¯π+ generate (B+ g ) · ¯π+ g−1), respectively. Thus ¯π+ e )∗)πg(Ag) · πh(Ah)¯π+ g−1)∗ = B+ h (B+ e (b)πg(a) for b ∈ (B+ g . Hence e ((B+ g )∗ = ¯π+ e )∗ · πgh(Agh)¯π+ e -module A ⊗Ae B+ e because A ⊗Ae B+ The induced representation λ of A on the Hilbert B+ g ) ⊆ B(E) for g ∈ G, such that the inclusions ¯π+ e ) .= ¯π+ We used here that π is a homomorphism on A and that Ag · Ah ⊆ Agh. (cid:3) Lemma 9.25. Assume that Be is a C∗-hull and that integrability is compatible with induction. There is a unique Fell bundle structure on (B+ g )g∈G such that g → B(E) form a Fell bundle representation for any integrable the maps ¯π+ representation π of A on a Hilbert module E. Proof. Lemmas 9.23 and 9.24 show that (9.17) holds under our assumptions. Hence the multiplication and involution in B(E) restrict to a Fell bundle structure on the g ) ,→ B(E) give subspaces ¯π+ a Fell bundle representation. e gives e contains a faithful representation of B+ e = B+ the identity representation. Hence the resulting representations ¯λg of B+ g are also faithful, even isometric, by Lemma 9.16. So the Fell bundle structure on ¯λg(B+ g ) g → B(E) form a Fell bundle representation. lifts to B+ Let π be any integrable representation of A. The exterior direct sum π ⊕ λ on the Hilbert D ⊕ B+ e ) is still integrable. The resulting g (b) ⊕ ¯λg(b) for b ∈ B+ g . The maps from B+ compressions to the direct summands E and A ⊗Ae B+ e therefore restrict to Fell g ⊕ ¯λg)(B+ bundle representations with respect to the Fell bundle structure on (¯π+ g ) g ) ∼= B+ defined above. Since λ is faithful, the projection (¯π+ g ) is is a Fell bundle isomorphism. Hence the map B+ g (B+ (cid:3) a Fell bundle representation. Let (βg)g∈G be a Fell bundle over a discrete group G (see [8]). Then β :=L g∈G βg is a G-graded ∗-algebra using the given multiplications and involutions among the subspaces βg. The section C∗-algebra C∗(β) of the Fell bundle is defined as the completion of β in the maximal C∗-seminorm. By construction, a representation of C∗(β) is equivalent to a representation of the Fell bundle. This holds also for representations on Hilbert modules. g to B(E0) simply give block matrices ¯π+ e -module E0 := E ⊕ (A ⊗Ae B+ g , so that the maps ¯λg : B+ g ⊕ ¯λg)(B+ ∼−→ (¯π+ g ) → ¯λg(B+ g ) → ¯π+ e ⊇ Ae ⊗Ae B+ g ⊕ ¯λg)(B+ g (B+ g g g∈G B+ e , µ+ e g g∈G B+ g )g∈G of (B+ e of B+ e on B+ g , µ+ e,g) of Ae on B+ e =L The representations ¯µ+ REPRESENTATIONS BY UNBOUNDED OPERATORS the ∗-algebraL says that any integrable representation π =L 51 Theorem 9.26. Let A be a graded ∗-algebra for which Ae has a C∗-hull. Assume that integrability is compatible with induction as in Definition 9.18. The section C∗-algebra B of the Fell bundle (B+ g )g∈G constructed above is a C∗-hull for the integrable representations of A. Proof. Representations of B are in natural bijection with Fell bundle representa- tions: restricting a representation of B to the subspaces B+ g gives a Fell bundle representation, and conversely a Fell bundle representation gives a representation of g , which extends uniquely to the C∗-completion. Lemma 9.25 g∈G πg of A induces a Fell bundle representation (¯π+ g )g∈G and thus a representation of B. By construction, this family of maps Repint(A) → Rep(B) is compatible with interior tensor products and unitary ∗-intertwiners. We are going to show that this is a family of bijections. First we describe an integrable representation (B, µ) of A on B. By construction, A ⊗Ae B+ g is dense in B. This subspace carries a representation of A by left multiplication. We extend this to the right ideal in B generated by A⊗Ae B+ to get a representation of A on B. Let (B, µ) be its closure. e,g of B+ g is another e )B+ e · B is a core for core for the representation (B+ the restriction of the representation (B, µ) to Ae. This core shows that (B, µAe) = e ) ⊗Be B, where the interior tensor product is with respect to the canonical (B+ e ,→ B. Therefore, the restriction of (B, µ) to Ae is integrable and embedding B+ e ,→ B. the corresponding representation ¯µ+ Thus the representation (B, µ) of A on B is also integrable. The integrable representation (B, µ) of A on B yields a representation ¯µ+ g )g∈G in M(B) = B(B). By construction, the image of of the Fell bundle (B+ e (b) = µ(ag) · b. That is, B+ ag ⊗ b ∈ Ag (cid:12)Ae B+ in B+ is g acts by µ(ag)¯µ+ g ,→ B. The representation of B represented by the canonical inclusion map B+ associated to this Fell bundle representation is the identity map on B. Interior tensor product with (B, µ) gives a family of maps Rep(B) → Repint(A) that is compatible with unitary ∗-intertwiners and interior tensor products. Since the composite family of maps Rep(B) → Repint(A) → Rep(B) is compatible with interior tensor products and maps the identity representation of B to itself, the composite map on Rep(B) is the identity. Let (E, π) be an integrable representation of A on a Hilbert D-module E for some C∗-algebra D. This yields a representation (¯π+ g )g∈G and an associated representation ¯π of B. We claim that the integrable representation (E0, π0) := (B, µ) ⊗¯π E is equal to (E, π). Both representations have the same restriction to Ae because e is simply the inclusion map B+ e,g(B+ g . Therefore, B+ g )g∈G of the Fell bundle (B+ g are defined so that ¯µ+ e , µ+ e , µ+ e (B+ e ) ⊗¯πBe (B, µAe) ⊗¯π E ∼= (B+ e ) ⊗Be B ⊗¯π E ∼= (B+ e )E, ag ∈ Ag acts by mapping ¯π+ E ∼= (E, π). e )E is e (B+ Hence both representations have the same domain by Lemma 9.2. And ¯π+ e (be)ξ = a core for both. On ¯π+ e (be)ξ to πg(ag)¯π+ g ⊆ B. Since (E, π) ¯π(ag ⊗ be)ξ in both representations, where we view ag ⊗ be ∈ B+ and (E0, π0) have a common core, they are equal. This finishes the proof that our two families of maps Repint(A) ↔ Rep(B) are inverse to each other. Thus B is a weak C∗-hull for the integrable representations of A. Since Ae is a C∗-hull, the integrable representations of Ae are admissible. So are the integrable representations of A by Proposition 9.4. Thus B is a C∗-hull. (cid:3) Remark 9.27. The fibres B+ e of the Fell bundle in Theorem 9.26 are described in Definitions 9.10 and 9.13, including the right Hilbert B+ e -module structure on B+ g . The rest of the Fell bundle structure needs technical extra assumptions. The simplest e g 52 RALF MEYER e g acts on A⊗Ae B+ e by ag⊗b 7→ πg(ag)·¯π+ e (b), where ¯π+ e to an integrable representation of A on the Hilbert B+ g )g∈G is represented faithfully in B(A ⊗Ae B+ way to get it is by inducing the universal inducible, integrable representation of A e -module A⊗Ae B+ e . The on B+ Fell bundle (B+ e ) by Lemma 9.16. The multiplication, involution, and norm in our Fell bundle are simply the multiplication, involution and norm in the C∗-algebra B(A⊗Ae B+ e ). The dense image of Ag(cid:12)Ae B+ in B+ e (b) is the representation e associated to the induced representation of Ae on A ⊗Ae B+ of the C∗-hull B+ e , which is integrable by assumption. 9.6. Two counterexamples. Two assumptions limit the generality of the Induction Theorem 9.26. First, integrability must be compatible with induction. Secondly, Be should be a C∗-hull and not a weak C∗-hull. Equivalently, all isometric intertwiners between integrable Hilbert space representations of Ae are ∗-intertwiners. We show by two simple counterexamples that both assumptions are needed. In particular, there is no version of the Induction Theorem for weak C∗-hulls. Both counterexamples involve the group G = Z/2 = {0, 1}. A G-graded ∗-algebra is a ∗-superalgebra, that is, a ∗-algebra with a decomposition A = A0 ⊕ A1 such that A0 · A0 + A1 · A1 ⊆ A0, A0 · A1 + A1 · A0 ⊆ A1, A∗ 1 ∈ A0. In both examples, A0 = C[x] with x = x∗. through the involution x 7→ −x. That is, In the first example, A is the crossed product for the action of Z/2 on A0 = C[x] 0 = A0, A∗ 1 = A1, A = Chx, ε ε2 = 1, xε = −εx, x = x∗, ε = ε∗i, x ∈ A0, ε ∈ A1. Since A1 = εA0 ∼= A0 as a right A0-module, any representation of A0 is inducible. (0,∞) ,→ R = cA0 (see Proposition 8.1). This gives a C∗-hull for a class of repre- Let B0 = C0((0,∞)) with the representation of A0 from the inclusion map sentations of A0 that is defined by submodule conditions and satisfies the Strong Local–Global Principle by Theorems 8.2 and 8.3. The class of (0,∞)-integrable representations consists of those representations of C[x] that are generated by a regular, self-adjoint, strictly positive operator. In a representation of A, the element ε ∈ A acts by a unitary involution that conjugates π(x) to −π(x). Hence π(x) cannot be strictly positive. Thus the zero- dimensional representation is the only representation of A whose restriction to A0 is C0((0,∞))-integrable. The C∗-hull for this class is {0}. Theorem 9.26 does not apply here because induced representations of inducible, integrable representations of A0 are never integrable when they are non-zero. The second example is the commutative ∗-superalgebra A = Chx, ε ε2 = 1 + x2, xε = εx, x = x∗, ε = ε∗i, x ∈ A0, ε ∈ A1. Thus A1 = εC[x] ∼= A0 with the usual A0-bimodule structure and the inner product hεa1, εa2i = (1 + x2) · a1 · a2. Since (1 + x2)a2 is positive in C[x] for any a ∈ C[x], any representation of A0 is inducible. Let (E, π) be a representation of A0 on a Hilbert module E over a C∗-algebra D. The induced representation A1⊗A0(E, π) lives on the Hilbert D-module completion E1 of E for the inner product hξ1, ξ2i1 := hξ1, π(1 + x2)ξ2i. Its domain is E, viewed as a dense D-submodule in E1, and the representation of A0 is π again. The operator π(x + i) on E extends to an isometry I : E1 ,→ E because hπ(x + i)ξ1, π(x + i)ξ2i = hξ1, π(x − i)π(x + i)ξ2i = hξ1, π(1 + x2)ξ2i = hξ1, ξ2i1 for all ξ1, ξ2 ∈ E. This isometry commutes with π(a) for all a ∈ A, so it is an isometric intertwiner A1 ⊗A0 (E, π) ,→ (E, π). Now let B0 with the universal representation (B0, µ0) be one of the two non- commutative weak C∗-hulls T0 or K('2N) of C[x] described in §6. In a Toeplitz REPRESENTATIONS BY UNBOUNDED OPERATORS 53 integrable representation, π(x + i) has dense range. Even more, π(x + i)E is dense in E in the graph topology. Thus I is a unitary ∗-intertwiner A1 ⊗A0 (E, π) ∼−→ (E, π) for any integrable representation (E, π) of A0. Since all representations of A0 are inducible, the unit fibre of the Fell bundle should be B0. The other fibre B1 is A1 ⊗A0 B0, which we have identified with B0. The unitary A1 ⊗A0 B0 ∼= B0 is a ∗-intertwiner between the representations of A0 as well. Therefore, integrability is compatible with induction. And the left B0-module structure ¯µ0,1 on B1 in (9.22) is simply left multiplication. Next we describe the induced representation of A on the Hilbert B0-module The representations of A and A0 on A⊗A0 B0 have the same domain by Lemma 9.2, and for A0 the domain is B0 ⊕ B0. We claim that A acts on this domain by A ⊗A0 B0 = A0 ⊗A0 B0 ⊕ A1 ⊗A0 B0 ∼= B0 ⊕ B0. (cid:18)µ0(x) µ0(x − i) (cid:18) (cid:19) (cid:19) 0 0 0 µ0(x) , µ0(x + i) 0 . We have already seen this for x ∈ A0. Left multiplication by ε maps b ∈ B0 ⊆ B0 first to ε⊗b ∈ A1⊗A0 B0, which is mapped by the isometry I to µ0(x+i)b ∈ B0 ⊆ B0. And it maps the element µ0(x + i)b ∈ B0 for b ∈ B0, which corresponds to ε ⊗ b in the odd fibre, to ε2 ⊗ b = µ0(x2 + 1)b = µ0(x − i)µ0(x + i)b ∈ B0. This proves the formula for the action of ε. The representation ¯µ0 of B0 on A⊗A0 B0 is the representation of the weak C∗-hull that corresponds to the representation of A0 ⊆ A described above. This is x 7→ ε 7→ (cid:18)b (cid:19) 0 b . 0 Hence ε ⊗ b ∈ A1 ⊗A0 B0 for b ∈ B0 acts by the matrix ¯µ0 : B0 → M2(B0), b 7→ 0 µ0(x + i) µ0(x − i) 0 0 µ0(x + i)b µ0(x − i)b 0 (cid:19)(cid:18)b 0 (cid:18) (cid:19) 0 b = (cid:18) (cid:19) . The map µ0(x + i)b 7→ µ0(x − i)b is the Cayley transform of µ0(x). For our two weak C∗-hulls, this is the unilateral shift S ∈ M(B0) by construction. Thus the odd fibre B1 ∼= B0 of our Fell bundle should act by b 7→ The map ¯µ0 is a ∗-representation, and (9.15) gives (cid:18)0 Sb (cid:19) ¯µ1 : B0 → M2(B0), 0 b . ¯µ1(b1)∗¯µ1(b2) = ¯µ0(b∗ 1b2), ¯µ1(b1)¯µ0(b2) = ¯µ1(b1b2) for all b1, b2 ∈ B0. This is also obvious from our explicit formulas. But Sb1b2 b1Sb2 and ¯µ0(b1b2) = ¯µ0(b1)¯µ1(b2) = (cid:19) 0 (cid:18) 0 b1b2 (cid:18) 0 b1b2 (cid:19) 0 differ if, say b1 = S∗, b2 = 1. In fact, ¯µ0(B0) · ¯µ1(B0) is not contained in ¯µ1(B0). Hence there is no Fell bundle structure on (Bg)g∈Z/2 for which (¯µg)g∈Z/2 would be a Fell bundle representation. 10. Locally bounded unit fibre representations We now specialise the Induction Theorem 9.26 to the case where the universal integrable representation of the unit fibre Ae is locally bounded. In this case, we may first construct a pro-C∗-algebraic Fell bundle whose unit fibre is the pro-C∗-algebra completion of Ae. This is relevant because pro-C∗-algebras are much closer to ordinary C∗-algebras than general ∗-algebras. We will see the importance of this in RALF MEYER e of positive characters. As before, let G be a group and let A =L 54 the commutative case, where the pro-C∗-algebraic Fell bundle gives us a twisted partial group action on the space A+ g∈G Ag be a G-graded ∗-algebra. We are interested in the locally bounded representations of Ae, and representations of A that restrict to locally bounded representations on Ae. The class Repb(Ae) of locally bounded representations of Ae is admissible by Corollary 7.9. So any weak C∗-hull for some smaller class of representations will be an ordinary C∗-hull. Let Ae be the pro-C∗-algebra completion of the unit fibre Ae, that is, the com- pletion of Ae in the topology defined by the directed set N (Ae) of all C∗-seminorms on Ae. Locally bounded representations of Ae are equivalent to locally bounded representations of Ae by Proposition 7.6. When is a locally bounded representation inducible? Proposition 10.1. A locally bounded representation (E, ϕ) of Ae on a Hilbert module E is inducible if and only if ϕ(a∗a) ≥ 0 for all a ∈ Ag, g ∈ G. The difference to the general criterion for inducibility in Proposition 9.5 is that finite linear combination Pn we do not consider matrices. Proof. The subspace Eb ⊆ E of bounded vectors is a core for ϕ. As in the proof of Proposition 9.12, it suffices to prove the positivity of the inner product for a k=1 ak ⊗ ξk with ak ∈ Ag, ξk ∈ Eb for a fixed g ∈ G. Since there are only finitely many ξk, there is a C∗-seminorm q on Ae so that all ξk are q-bounded. Thus we may replace E by the Hilbert submodule Eq of q-bounded vectors, where the representation of Ae extends to the C∗-completion D := (Ae)q for q. Since we assume ϕ(a∗a) ≥ 0 for all a ∈ Ag, this representation factors through the quotient of D by the closed ideal I generated by the negative parts (a∗a)− for all a ∈ Ag, g ∈ G. The D/I-valued inner product ha1, a2i := a∗ 1a2 mod I on Ag is positive definite by construction; since D/I is a C∗-algebra, we may use the usual notion of positivity here, which does not involve matrices. Then the inner product on the tensor product Ag ⊗D/I Eq is also positive definite. This is what we had to (cid:3) prove. A pro-C∗-algebra has a functional calculus for self-adjoint elements. Hence we may construct the negative parts (a∗a)− ∈ Ae for a ∈ Ag, g ∈ G. We let A+ e be the completed quotient of Ae by the closed two-sided ideal generated by these elements. This is another pro-C∗-algebra, and it is the largest quotient in which a∗a ≥ 0 for all a ∈ Ag, g ∈ G. By Proposition 10.1, a locally bounded representation of Ae is inducible if and only if the corresponding locally bounded representation of Ae factors through A+ e . Corollary 10.2. There is an equivalence between the inducible, locally bounded rep- resentations of Ae and the locally bounded representations of the pro-C∗-algebra A+ e , which is compatible with isometric intertwiners and interior tensor products. Proof. Proposition 10.1 says that the equivalence in Proposition 7.6 maps the subclass in Repb(Ae) of inducible, locally bounded representations of Ae onto the subclass Repb(A+ (cid:3) e . This is isomorphic to the subset of N (Ae) consisting of all C∗-seminorms q on Ae for which a∗a ≥ 0 holds in the C∗-completion (Ae)q for all a ∈ Ag, g ∈ G. We would like to complete A g is a Hilbert bimodule over A+ Example 10.3. It can happen that the class of locally bounded representations of Ae is not compatible with induction. Let End∗(C[N]) be the ∗-algebra of all e . But such a construction does not work in the following example. Let N (Ae)+ be the directed set of C∗-seminorms on A+ to a ∗-algebraL e ) in Repb(Ae). g∈G A+ g with unit fibre A+ e , where each A+ REPRESENTATIONS BY UNBOUNDED OPERATORS 55 ∞ × ∞-matrix with only finitely many entries in each row and each column, with the usual matrix multiplication and involution. Let A be the Z/2-graded ∗-algebra of block 2 × 2-matrices , a ∈ C, b ∈ C[N], c ∈ C[N], d ∈ End∗(C[N]), (cid:18)a b (cid:19) c d with the grading where a, d are even and b, c are odd. Here b and c are infinite column and row vectors with only finitely many non-zero entries, respectively. Thus A ∼= End∗(C[N]) with the grading induced by the grading on C[N] where C · δ0 is the even part and the span of δi for i > 0 is the odd part. The character (a, d) 7→ a is a bounded representation of the unit fibre A0. Induction gives the standard representation of A on the Hilbert space C ⊕ '2(N) ∼= '2(N) by matrix-vector multiplication. This representation is irreducible because already the ideal of finite matrices M∞(C) in A acts irreducibly. It is not bounded, that is, some elements in End∗(C[N]) act by unbounded operators on '2(N). Hence it is not locally bounded by Proposition 7.7. To rule out this problem, we now assume that induction from Ae to A and restric- tion back to Ae maps bounded representations of Ae again to bounded representations of Ae, briefly, that boundedness is compatible with induction. This implies that local boundedness is compatible with induction because a locally bounded representation contains bounded subrepresentations whose union is a core for it. Our assumption is equivalent to the boundedness of the induced representations of Ae on the Hilbert e )q for all g ∈ G and q ∈ N (Ae)+. That is, there is (A+ another norm q0 ∈ N (Ae)+ such that q(a∗b∗ba) = kbak2 e )q-modules Ag ⊗Ae (A+ q = q0(b)2 · q(a∗a) q ≤ kbk2 q0kak2 g be the completion of Ag in the topology generated for all a ∈ Ag, b ∈ Ae. Let A+ by the family of norms q(a∗a) for q ∈ N (Ae)+. Lemma 10.4. The multiplication maps and the involutions in (Ag)g∈G extend to continuous maps A+ Proof. Given q ∈ N (Ae)+, let q0 ∈ N (Ae)+ be such that q(a∗b∗ba) ≤ q0(b)2 · q(a∗a) for all a ∈ Ah, b ∈ Ae. If b ∈ Ag, a ∈ Ah, then g−1 for g, h ∈ G. gh and A+ h → A+ g → A+ g × A+ q kbak2 q0kak2 g × A+ q ≤ kak2 h → A+ gh. for all a ∈ Ah. That is, ka∗k2 continuous as well. The completion A+ :=L q := q(a∗b∗ba) = q(a∗(b∗b)1/2(b∗b)1/2a) ≤ q0((b∗b)1/2)2 · q(a∗a) = kbk2 g )g∈G and hence extends to a jointly continuous map A+ That is, the multiplication is jointly continuous with respect to the topology defin- ing (A+ Furthermore, q(aa∗)2 = q(aa∗aa∗) ≤ q0(a∗a) · q(aa∗) and hence q(aa∗) ≤ q0(a∗a) q0 for all a ∈ Ah. Thus the involution is (cid:3) g of A is again a ∗-algebra by Lemma 10.4. By e are positive for a ∈ Ag, g ∈ G; construction of A+ this remains so for a ∈ A+ g because the subset of positive elements in A+ e is closed. Thus (A+ g )g∈G has the usual properties of a Fell bundle over G, except that the fibres are only Hilbert bimodules over a pro-C∗-algebra. We interpret (A+ g )g∈G as a partial action of G on A+ Usually, the norms q(a∗a) and q(aa∗) on Ag are not equivalent for a fixed q ∈ N (A)+. This prevents us from completing A+ to a pro-C∗-algebra. It also means that the integrable representations of A are not locally bounded on A, but only on Ae. This happens in interesting examples such as the Weyl algebra discussed in §13. This phenomenon for Fell bundles is related to the known problem that e , the inner products a∗a ∈ A+ e by Hilbert bimodules as in [4]. g∈G A+ g RALF MEYER sentations on Ae are equivalent to representations of the ∗-algebra A+ =L 56 crossed products for group actions on pro-C∗-algebras only work well if the action is strongly bounded, that is, the invariant continuous C∗-seminorms are cofinal in the set of all continuous C∗-seminorms, see [13]. Proposition 10.5. Suppose that boundedness for representations of Ae is compat- ible with induction to A. Representations of A that restrict to locally bounded repre- g∈G A+ that restrict to locally bounded representations on A+ e ; this equivalence is compatible with isometric intertwiners and interior tensor products. Proof. Let π be a representation of A for which πe is a locally bounded representation of Ae. The representation πe is inducible by Lemma 9.8. Hence πe is the closure of the restriction of a locally bounded representation ¯π+ e by Corollary 10.2. The representation πg of Ag for g ∈ G is continuous with respect to the topology defin- ing A+ g and the graph topology on the domain of πg because πg(a)∗πg(a) = πe(a∗a). Hence it extends uniquely to A+ g such that π is the closure of ¯π+ ◦ j. It is easy to see that this equivalence between the g that are locally bounded on A+ e is compatible with isometric intertwiners and interior tensor (cid:3) products. We will explore the consequences of this in the case of commutative Ae in §11. In that case, boundedness is automatically compatible with induction, and the pro- C∗-algebraic Fell bundle A+ e gives rise to a twisted groupoid with object space A+ e . Thus the C∗-hull produced by the Induction Theorem 9.26 is a twisted groupoid C∗-algebra when Ae is commutative and the integrable representations of Ae are locally bounded. Here we briefly consider the situation of Theorem 7.16 where C0(A+ e ) is dense in A+ e and provides a C∗-hull for the class of locally bounded representations. Then we define locally bounded representations of A and the representations ofL g , and this gives a representation ¯π+ ofLA+ e of A+ g∈G A+ C0(A+ g ) := {a ∈ Ag a∗a ∈ C0(A+ e )}. e ) · A+ g · Cc(A+ e ) \ Prim(A+ e . Then C0(A+ e ) is dense in A+ g are dense in C0(A+ g ). e ). The proof of Lemma 7.15 shows that A+ g ) if and only if for all ε > 0 there is q ∈ N (Ae)+ such that That is, a ∈ C0(A+ ka∗akp < ε for all p ∈ Prim(A+ e )q. Since the involutions Ag → Ag−1 and Ag−1 → Ag are both continuous, they are homeomorphisms. Thus a ∈ C0(A+ g ) if and only if aa∗ ∈ C0(A+ e ) and Cc(A+ Theorem 10.6. Assume that boundedness is compatible with induction from Ae to A and that C0(A+ g )g∈G is a Fell bundle over G whose section C∗-algebra is a C∗-hull for the class of all representations of A that restrict to a locally bounded representation of Ae. Proof. The assumption that boundedness is compatible with induction allows us to build the pro-C∗-algebraic Fell bundle (A+ g )g∈G. Call a representation of A = g integrable if the restriction to the unit fibre Ae or A+ g∈G A+ is locally bounded, respectively. These classes of integrable representations are equivalent by Proposition 10.5. e ) is dense in A+ Since C0(A+ e , it is a C∗-hull for the locally bounded representations of A+ e by Theorem 7.17. Equivalently, it is a C∗-hull for the inducible, locally g ). Representations of C0(A+) are equivalent to representations of the Fell bundle C0(A+ g ). Thus we must prove that the class of all representations of C0(A+) is equivalent to the class of integrable representations of A+. More precisely, the equivalence maps a representation  of C0(A+) on a Hilbert module E to the representation π of A+ g∈G Ag or A+ :=L bounded representations of Ae. Let C0(A+) :=L g∈G C0(A+ L e e REPRESENTATIONS BY UNBOUNDED OPERATORS 57 e ))E and π(a)(b)ξ := (a · b)ξ for all a ∈ A+, b ∈ Cc(A+ with the core (Cc(A+ e ), ξ ∈ E; here a · b is the product in A+, which belongs to C0(A+) if b ∈ Cc(A+ e ). In the converse direction, we may simply restrict a locally bounded representation of A+ to the ∗-subalgebra C0(A+). This restriction is nondegenerate because e ) ⊆ C0(A+) acts nondegenerately in any integrable representation of A+: C0(A+ this is part of the equivalence between representations of C0(A+ e ) and locally bounded representations of A+ e in Theorem 7.17. We claim that the maps from representations of C0(A+) to integrable representations of A+ and back are inverse to each other. Let π be an integrable representation of A+ on a Hilbert module E. The representations π and πA+ have the same domain by Lemma 9.2. Since πA+ is locally bounded, π(Cc(A+ g = A+ e ))E is a core for πA+ g · e ) for all g ∈ G, this subspace is π(A+)-invariant and thus a core for π. The Cc(A+ representation  of C0(A+) is the closure of the restriction of π to C0(A+) ⊆ A+. By definition, the representation of A+ has the core (Cc(A+ e ))E and acts there by π0(a)(b)ξ = (a · b)ξ. The subspace (Cc(A+ e ))E is a core for this representation because the map ξ 7→ π0(a)(b)ξ is continuous in the norm topology on E and E is dense in E. If ξ ∈ E, then (b)ξ = π(b)ξ and hence π0(a)π(b)ξ = π(a)π(b)ξ for all a ∈ A+, b ∈ Cc(A+ Now start with a representation  of C0(A+). Let π be the associated integrable e ))E and acts there by π(a)(b)ξ = representation of A+. It has the core (Cc(A+ e ), ξ ∈ E. In particular, if a ∈ C0(A+), then (a · b)ξ for all a ∈ A+, b ∈ Cc(A+ π(a)(b)ξ = (a · b)ξ = (a)(b)ξ. Since C0(A+ e ) · C0(A+ g ) for all g ∈ G, the restriction of  to C0(A+ e ) remains nondegenerate. Therefore, the e ), ξ ∈ E is dense in E. Hence  is the restriction of π to set of (b)ξ for b ∈ Cc(A+ C0(A+) ⊆ A+, as desired. (cid:3) e ), ξ ∈ E. This implies π = π0, as desired. g ) is dense in C0(A+ . Since Cc(A+ e ) · A+ e e The proof of Theorem 10.6 does not use the constructions in Section 9 and so provides an alternative proof of the Induction Theorem in case the chosen class of integrable representations of Ae is the class of all locally bounded representations. 11. Fell bundles with commutative unit fibre So let G be a discrete group and A = L In this section, we apply the Induction Theorem in the case where Ae and the chosen C∗-hull Be are commutative. This is the only case considered in [26]. Extra assumptions in [26] ensure that the C∗-hull for the integrable representations of A e ⊆ Ae of positive is the crossed product for a partial action of G on the space A+ characters. Without these assumptions, we shall get a "twisted" crossed product for a partial action. g∈G Ag a G-graded ∗-algebra such that Ae is commutative. We have already classified the possible commutative C∗-hulls for Ae in §8. In particular, all commutative weak C∗-hulls are already C∗-hulls by Theorem 8.2, and they correspond to injective, continuous maps from locally compact spaces to the spectrum Ae of Ae. Explicitly, let X be a locally compact space and let j : X → Ae be an injective, continuous map. Let Be = C0(X) and define a representation of Ae on Be with domain Cc(X) by (a · f)(x) = a(j(x)) · f(x) for all a ∈ Ae, f ∈ Cc(X), x ∈ X, where a(χ) = χ(a) for χ ∈ Ae. Let µe be the closure of this representation of Ae on Be. The C∗-algebra Be with the universal representation µe is a C∗-hull for a class Repint(Ae, X) of representations of Ae by Theorem 8.2, and any commutative C∗-hull is of this form. Let Repint(A, X) be the class of representations of A that restrict to a represen- tation in Repint(Ae, X) on Ae, as in Definition 9.3. If Repint(Ae, X) is compatible with induction to A as in Definition 9.18, then Theorem 9.26 gives a Fell bundle e = C0 RALF MEYER e of Ae, and B+ (cid:0)j−1( A+ e )(cid:1). 58 whose section C∗-algebra is a C∗-hull for Repint(A, X). We are going to characterise exactly when this happens and describe the C∗-hull for Repint(A, X) as a twisted groupoid C∗-algebra. Any representation of Ae on a commutative C∗-algebra is locally bounded by Proposition 8.1. Hence the constructions in §10 specialise to our commutative case. Actually, we shall make these results more explicit through independent proofs. First we describe the C∗-hull B+ e for the inducible representations in Repint(Ae, X) as in Proposition 10.1: Lemma 11.1. Call a character χ ∈ Ae positive if χ(a∗a) ≥ 0 for all a ∈ Ag and all g ∈ G. These form a closed subset A+ Proof. The positive characters form a closed subset in Ae by definition of the topology on Ae. We have constructed B+ e in Proposition 9.12 as a quotient of Be, such that a representation is inducible if and only if it factors through B+ e . Thus B+ corresponds to a certain closed subset of Ae. Its points are the inducible characters of Ae. Let χ be a character. Any vector in Ag ⊗Ae,χ C is of the form a ⊗ 1 for some a ∈ Ag, that is, there is no need to take linear combinations. Hence the sesquilinear form on Ag ⊗Ae,χ C for all g ∈ G is positive semidefinite if and only if χ(a∗a) ≥ 0 for all a ∈ Ag and all g ∈ G, that is, χ is positive. Thus B+ e is the quotient corresponding to those x ∈ Ae for which j(x) ∈ Ae is positive. (cid:3) Theorem 11.2. Let g ∈ G and χ ∈ A+ Dg−1 := {χ ∈ A+ e . The left Ae-module structure on Ag ⊗Ae,χ C ∼= C for χ ∈ is relatively open in A+ Dg−1 is by a character ϑg(χ) that belongs to Dg. The map ϑg is a homeomorphism from Dg−1 onto Dg, and these maps form a partial action of G on A+ e , that is, ϑe = id A Proof. As in the proof of Lemma 11.1, Ag ⊗Ae,χ C is the Hausdorff completion of Ag in the norm coming from the inner product ha1, a2i := χ(a∗ 1a2). We write λ · a for a ⊗ λ for a ∈ Ag, λ ∈ C throughout this proof, and we write a ≡ b if a, b ∈ A have the same image in Ag ⊗Ae,χ C. Let a, b ∈ Ag satisfy χ(a∗a) 6= 0 and χ(b∗b) 6= 0. We must show that a and b are parallel in Ag ⊗Ae,χ C. e . Then dim Ag ⊗Ae,χ C ≤ 1. The set e dim Ag ⊗Ae,χ C = 1} and ϑg ◦ ϑh ⊆ ϑgh for all g, h ∈ G. + e e The following computation makes [7, Footnote 3] explicit: (a∗ab∗b)2 = a∗ab∗(ba∗)(ab∗)b = a∗a(b∗a)(b∗ba∗b) = a∗ab∗ba∗bb∗a because Ae is commutative and the terms in parentheses belong to Ae. Hence χ(a∗a)2χ(b∗b)2 = χ(a∗a)χ(b∗b)χ(a∗b)χ(b∗a). χ(a∗a)χ(b∗b) = χ(a∗b)χ(b∗a) = χ(a∗b)2 6= 0. Since χ(a∗a) 6= 0 and χ(b∗b) 6= 0, this implies (11.3) The inner product on Ag ⊗Ae,χ C annihilates a · c ⊗ 1 − a ⊗ χ(c), which we write as a · c − χ(c)a, for all a ∈ Ag, c ∈ Ae. Hence χ(a∗a)χ(b∗b) a ≡ aa∗bb∗a (11.4) a = χ(a∗b)χ(b∗a) χ(a∗a)χ(b∗b) = bb∗aa∗a χ(a∗a)χ(b∗b) Thus all non-zero a, b ∈ Ag ⊗Ae,χ C are parallel, that is, dim Ag ⊗Ae,χ C ≤ 1. The space Ag ⊗Ae,χ C is non-zero if and only if there is a ∈ Ag with χ(a∗a) 6= 0. Thus (11.5) e χ(a∗a) 6= 0 for some a ∈ Ag}. Dg−1 = {χ ∈ A+ = χ(b∗a)χ(a∗a) χ(a∗a)χ(b∗b) b = χ(b∗a) χ(b∗b) b. REPRESENTATIONS BY UNBOUNDED OPERATORS 59 The latter set is relatively open in A+ e . Let χ ∈ Dg−1. Then dim Ag ⊗Ae,χ C = 1. Hence the representation of Ae on it is by a character, which we denote by ϑg(χ). This character is an inducible representation by Lemma 9.8, and hence positive by Lemma 11.1. There is b ∈ Ag with χ(b∗b) > 0. If a ∈ Ae, then (11.4) implies ab ≡ χ(b∗ab) χ(b∗b) b. Thus ϑg(χ)(a) = χ(b∗ab) χ(b∗b) (11.6) for all a ∈ Ae. Hence ϑg(χ)(cid:0)(b∗)∗b∗(cid:1) 6= 0, so that ϑg(χ) ∈ Dg by (11.5). Thus ϑg maps Dg−1 to Dg. Equation (11.6) also implies that the map ϑg is continuous on e with χ(b∗b) > 0. Since these open sets for different the open set of characters in A+ b ∈ Ag cover Dg−1, the map ϑg is continuous on all of Dg−1. Let g, h ∈ G and let χ ∈ Dh−1 and ϑh(χ) ∈ Dg−1. Then there is bh ∈ Ah with hbh) · hbh) > 0, and bg ∈ Ag with ϑh(χ)(b∗ χ(b∗ ϑh(χ)(b∗ gbg) > 0, and so (11.6) for b = bgbh ∈ Agh describes ϑgh. Hence gbg) > 0. Thus χ(b∗ gbgbh) = χ(b∗ hb∗ hb∗ ϑgh(χ)(a) = χ(b∗ gbgbh) = ϑh(χ)(b∗ gabgbh) χ(b∗ ϑh(χ)(b∗ hb∗ gabg) gbg) = ϑg (cid:0)ϑh(χ)(cid:1)(a). Thus ϑgh ⊆ ϑgϑh for all g, h ∈ G. In addition, ϑe = id A partial action of G on A+ onto Dg with inverse ϑg−1. . So the maps ϑg form a e , see [8]. In particular, ϑg is a homeomorphism from Dg−1 (cid:3) + e In the examples considered in [7,26], the space A+ e whose kernel is the group bundle A+ e , arrow spaceF e is locally compact and the C∗-hull for the integrable representations of A is the crossed product for the partial action of G on A+ e described above. In general, however, certain twists are possible. The partial action of G on A+ e may be encoded in a transformation groupoid G(cid:110) A+ e , g∈G Dg−1 with the disjoint union topology, which has object space A+ range and source maps s(g, χ) := χ, r(g, χ) := ϑg(χ) for g ∈ G, χ ∈ Dg−1, and h (Dg−1). multiplication (g, ϑh(χ)) · (h, χ) := (g · h, χ) for all g, h ∈ G, χ ∈ Dh−1 ∩ ϑ−1 The unit arrow on χ is (1, χ), and the inverse of (g, χ) is (g−1, ϑg(χ)). This is an étale topological groupoid because r and s restrict to homeomorphisms on the open subsets Dg−1 of the arrow space. The object space A+ e need not be locally compact. We are going to construct another topological groupoid Σ that is a central extension of G (cid:110) A+ e by the circle group T. That is, Σ comes with a canonical e ×T. Such an extension is also functor to G(cid:110) A+ called a twisted groupoid in [24, Section 4], following a definition by Kumjian [15]. A twisted groupoid with locally compact object space has a twisted groupoid C∗-algebra. For a suitable injective continuous map X → A+ e , we are going to identify the C∗-hull of the X-integrable representations of A with the twisted groupoid C∗-algebra of the restriction of Σ to j(X+) ⊆ A+ e . A point in Σ is a triple (g, χ, [a]), where g ∈ G, χ ∈ Dg−1, and [a] is a unit vector in the 1-dimensional Hilbert space Ag ⊗Ae,χ C. We represent unit vectors in Ag ⊗Ae,χ C by elements a ∈ Ag with χ(a∗a) = 1; two elements a, b ∈ Ag with χ(a∗a) = χ(b∗b) = 1 represent the same unit vector [a] = [b] if and only if χ(a∗b) = 1. We get the same set of equivalence classes if we allow a ∈ A with χ(a∗a) > 0 and set [a] = [b] if χ(a∗b) > 0: then a1 := χ(a∗a)−1/2a and b1 := χ(b∗b)−1/2b satisfy [a] = [a1], [b] = [b1], and [a] = [b] if and only if χ(a∗ 1b1) = 1 by (11.3). The circle group T acts on Σ by multiplication: λ · (g, χ, [a]) := (g, χ, [λa]). The orbit space projection for this circle action is the coordinate projection F : Σ (cid:16) G (cid:110) A+ e , (g, χ, [a]) 7→ (g, χ). Next we equip Σ with a topology so that this coordinate projection is a locally trivial principal T-bundle. 60 For a ∈ A, let Ua := {χ ∈ A+ RALF MEYER e χ(a∗a) 6= 0}. This is an open subset in A+ e , and χ(a∗a) > 0 if χ ∈ Ua because χ is positive. The map σa : {g} × Ua → Σ, (g, χ) 7→ (g, χ, [a]), for a ∈ Ag is a local section for the coordinate projection F. If (cid:20) a, b ∈ Ag, and χ ∈ Ua ∩ Ub, then [a] = χ(b∗a) (cid:21) χ(a∗a)1/2χ(b∗b)1/2 b , by (11.4). Since the functions sending χ to χ(b∗a), χ(a∗a) and χ(b∗b) are continuous on Ae, the two trivialisations induce the same topology on the restriction of Σ to {g} × (Ua ∩ Ub). For any χ ∈ Dg−1, there is a ∈ Ag with χ(a∗a) > 0. Thus the open subsets Ua cover Dg−1. Consequently, there is a unique topology on Σ that makes the local sections σa for all a ∈ Ag continuous, and this topology turns Σ into a locally trivial T-bundle over G (cid:110) A+ e . We define a groupoid with object space A+ e , arrow space Σ, and r(g, χ, [a]) := ϑg(χ), we must show that this multiplication is well defined. We have ab ∈ Agh and (g, [ϑh(χ)], [a])·(h, χ, [b]) := (g·h, χ, [a·b]); s(g, χ, [a]) := χ, χ(b∗a∗ab) = ϑh(χ)(a∗a) · χ(b∗b) 6= 0 by (11.6), so (g · h, χ, [a · b]) ∈ Σ. If χ(b∗b1) > 0 and ϑh(χ)(a∗a1) > 0, then χ(b∗a∗a1b1) > 0 by computations as in the proof of Theorem 11.2. Hence the multiplication is well defined. It is clearly associative. The unit arrow on χ is 1χ := (1, χ, [1]), and (g, χ, [a])−1 = (g−1, ϑg(χ), [a∗]). The multiplication, unit map and inversion are continuous and the range and source maps are open surjections (even locally trivial). So Σ is a topological groupoid. The identity map on objects and the coordinate projection F : Σ → G (cid:110) A+ e on arrows form a functor, which is a locally trivial, open surjection on arrows. The kernel of F consists of those (g, χ, [a]) ∈ Σ for which F(g, χ, [a]) is a unit arrow e . Then g = 1, and a ∈ Ae is equivalent to [a] = [χ(a) · 1] because in G (cid:110) A+ χ(a∗χ(a)1) > 0. The map (g, χ, [a]) 7→ (χ, χ(a)) is an isomorphism of topological e × T. Thus we have groupoids from the kernel of F onto the trivial group bundle A+ an extension of topological groupoids e × T (cid:26) Σ (cid:16) G (cid:110) A+ A+ e . e ) ⊆ X. e = C(X+) with X+ := j−1( A+ The three groupoids above are clearly Hausdorff. To construct C∗-algebras, we need groupoids with a locally compact object space. Therefore, we replace Ae by a locally compact space X with an injective, continuous map j : X → Ae. Then Be = C0(X) is a C∗-hull for a class Repint(Ae, X) of representations of Ae. By Lemma 11.1, the C∗-hull for the class of X-integrable, inducible representations of Ae is B+ Proposition 11.7. Let j : X → Ae be an injective, continuous map. The class Repint(A, X) is compatible with induction if and only if j(X+) ⊆ A+ e is invariant under the partial maps ϑg in Theorem 11.2 and the resulting partial maps on X+ are continuous in the topology of X+. We briefly say that the partial action of G on A+ Proof. By Proposition 9.19, it suffices to check that the induced representation of Ae on Ag ⊗Ae B+ is X-integrable for g ∈ G if and only if the partial map ϑg ◦ j on X factors through j and the resulting partial map j−1 ◦ ϑg ◦ j on X is again continuous. View the Hilbert module Ag ⊗Ae B+ e as a continuous field of Hilbert spaces over X+. The fibres of this field have dimension at most 1 by Theorem 11.2, and the set where the fibre is non-zero is the open subset j−1(Dg−1). e ) ∼= C0(j−1(Dg−1)). The representation of Ae on Ag ⊗Ae B+ Hence K(Ag ⊗Ae B+ e restricts to X+. e e REPRESENTATIONS BY UNBOUNDED OPERATORS 61 is equivalent to a representation on K(Ag ⊗Ae B+ e ) by Proposition 3.13. This is equivalent to a continuous map j−1(Dg−1) → Ae by Proposition 8.1. This map is ϑg ◦ j by a fibrewise computation. Hence the induced representation of Ae on Ag ⊗Ae B+ e is X-integrable if and only if ϑg ◦ j has values in j(X) and the partial maps j−1 ◦ ϑg ◦ j on X are continuous. (cid:3) From now on, we assume that the partial action of G on A+ e restricts to X+. By Proposition 11.7, this assumption is necessary and sufficient for X-integrability to be compatible with induction. The "restriction" of the partial action on A+ e to X+ is a partial action of G on X+ by partial homeomorphisms. Its transformation groupoid G (cid:110) X+ is constructed like G (cid:110) A+ e of arrows with range and/or source in j(X+), and the topology on the arrow space is the unique one that makes the inclusion G (cid:110) X+ → G (cid:110) A+ e and the range and source maps G (cid:110) X+ → X+ continuous. There is also a unique topology on the restriction ΣX of Σ to j(X+) so that there is an extension of topological groupoids e . Its set of arrows is the subset of G (cid:110) A+ X+ × T (cid:26) ΣX (cid:16) G (cid:110) X+. Since X+ is locally compact, the groupoids in this extension are locally compact, Hausdorff groupoids. Since G (cid:110) X+ is étale, it carries a canonical Haar system, namely, the family of counting measures. There is also a unique normalised Haar system on X+ ×T. These produce a unique Haar system on ΣX by [5, Theorem 5.1], so that the groupoid C∗-algebra C∗(ΣX) is defined. The twisted groupoid C∗-algebra C∗(G (cid:110) X+, ΣX) of G (cid:110) X+ with respect to the twist ΣX is defined in [23]. It is related to the groupoid C∗-algebra of ΣX in [5, Corollary 7.2]. Theorem 11.8. Let G be a discrete group and let A be a G-graded ∗-algebra with commutative Ae. Let j : X → Ae be an injective, continuous map, such that the partial action of G on A+ e in Theorem 11.2 restricts to X+ as in Proposition 11.7. Then C∗(G (cid:110) X+, ΣX) is a C∗-hull for Repint(A, X). Proof. The C∗-algebra C∗(G(cid:110)X+, ΣX) may be defined as the full section C∗-algebra of a certain Fell line bundle over the étale, locally compact groupoid G (cid:110) X+. The Fell line bundle involves the space of sections of the Hermitian complex line bundle L := ΣX ×T C associated to the principal T-bundle ΣX (cid:16) G (cid:110) X+ and the multiplication maps Lg × Lh → Lgh induced by the multiplication of ΣX (see [5]). By construction, the Hilbert B+ e is isomorphic to the continuous sections of this line bundle L over the subset {g} × DX g−1 of {g} × X+: an element a ⊗ b is mapped to the continuous section that sends (g, x) for x ∈ X with j(x) ∈ Dg−1 to b(x) · χ(a∗a)1/2[a]. The multiplication in ΣX is defined so that g ⊗B h → B+ B+ the multiplication maps B+ gh are exactly the multiplication maps in the Fell line bundle associated to ΣX. g )g∈G constructed in Theorem 9.26 is isomorphic to the Fell bundle (βg)g∈G, where βg is the space of C0-sections of L over {g} × Dg−1 and the multiplication and involution come from the Fell line bundle structure on L over the groupoid G (cid:110) X+. The full section C∗-algebra of this Fell bundle is canonically isomorphic to the section C∗-algebra of the corresponding Fell bundle over the groupoid G (cid:110) X+ by results of [3]. The small issue to check here is that it makes no difference whether we use C0-sections or compactly supported continuous sections of L over {g} × Dg−1. Both have the same C∗-completion. This is a special case of (cid:3) general results about Fell bundles over étale locally compact groupoids. g∈G Bg is a ∗-algebra, to which we may apply our machinery although all its representations are bounded. Thus any Fell bundle over G may come up for some choice of the G-graded ∗-algebra A. If (Bg)g∈G is any Fell bundle over G, then L Thus the Fell bundle (B+ g = Ag ⊗Ae B+ e -module B+ + e If ΣX RALF MEYER 62 Thus the section C∗-algebra of a Fell bundle (Bg)g∈G with commutative unit fibre is always a twisted groupoid C∗-algebras of a twist of an étale groupoid, namely, the transformation groupoid of a certain partial action on the spectrum of the unit fibre associated to the Fell bundle. This result is already known, even for Fell bundles over inverse semigroups with commutative unit fibre, see [3]. ∼= (G (cid:110) X+) × T as a groupoid, then C∗(G (cid:110) X+, ΣX) ∼= C∗(G (cid:110) X+). This is the same as the crossed product for the partial action of G on X+. This happens in all the examples in [7, 26]. The possible twists have two levels. First, ΣX may be non-trivial as a principal circle bundle over G (cid:110) X+. Secondly, if it is trivial as a principal circle bundle, the multiplication may create a non-trivial twist. The circle bundle ΣX (cid:16) G(cid:110)X+ is trivial if and only if its restriction to {g}×Dg−1 is trivial for each g ∈ G. For a circle bundle, this means that there is a nowhere vanishing section. For instance, if there is a ∈ Ag that generates Ag as a right Ae-module, then Ua = Dg−1 and σa is a global trivialisation of ΣX{g}×Dg−1 . The complex line bundles over a space X are classified by the second cohomology group H2(X, Z). If L is a line bundle, then the spaces of C0-sections of L⊗n for n ∈ Z form a Fell bundle over Z, and the direct sum of these spaces of sections is a Z-graded ∗-algebra such that the given line bundle L appears in the resulting twisted groupoid. If H2(X, Z) 6= 0, the space X is at least 2-dimensional. There are indeed non-trivial complex line bundles over all compact oriented 2-dimensional manifolds. The resulting ∗-algebra, however, has only ∗-representations by bounded operators if X is compact. Examples where unbounded operators appear must involve a non-trivial line bundle over a noncompact space. These first appear in dimension 3. e is, say, S2 × R and B+ It is easy to write down a Z-graded ∗-algebra A where B+ involves the Bott line bundle over S2. These examples seem artificial, however. Now assume that ΣX is trivial as a principal circle bundle over (G (cid:110) X+)1, that ∼= (G (cid:110) X+)1 × T as a T-space. We may choose this homeomorphism to be is, ΣX the obvious one on the open subset (1 (cid:110) X+) × T corresponding to 1 ∈ G. The multiplication must be of the form g (g1, ϑg2(x), λ1) · (g2, x, λ2) = (g1 · g2, x, ϕ(g1, g2, x) · λ1 · λ2) −1 1 ϕ(g1, g2 · g3, x) · ϕ(g2, g3, x) = ϕ(g1 · g2, g3, x) · ϕ(g1, g2, ϑg3(x)) for some continuous T-valued function ϕ with ϕ(1, g, x) = 1 = ϕ(g, 1, x) for all g, x; here ϕ is defined on the space of all triples (g1, g2, x2) ∈ G× G× X+ with x2 ∈ Dg −1 and ϑg2(x2) ∈ Dg 2 ; this space is homeomorphic to the space (G (cid:110) X+)2 of pairs of composable arrows in G (cid:110) X+. The associativity of the multiplication in ΣX is equivalent to the cocycle condition (11.9) for all g1, g2, g3 ∈ G, x ∈ X+ for which ϑg3(x), ϑg2 ◦ ϑg3(x), and ϑg1 ◦ ϑg2 ◦ ϑg3(x) are defined. A different trivialisation of the circle bundle ΣX (cid:16) (G (cid:110) X+)1 modifies ϕ by the coboundary (11.10) of a continuous function ψ : (G (cid:110) X+)1 → T normalised by ψ(1, x) = 0 for all x ∈ X+. Thus isomorphism classes of twists of G (cid:110) X+ are in bijection with the groupoid cohomology H2(G (cid:110) X, T), that is, the quotient of the group of continuous maps ϕ: (G (cid:110) X+)2 → T satisfying (11.9) by the group of 2-coboundaries ∂ψ of continuous 1-cochains ψ : (G (cid:110) X+)1 → T, where ∂ψ is defined in (11.10). In the easiest case, the function ϕ above does not depend on x. Then ϕ: G×G → T is a normalised 2-cocycle on G in the usual sense. These cocycles appear, for instance, in the classification of projective representations of the group G. This is related to the twists above because the Hilbert space representations of the twisted group ∂ψ(g1, g2, x) := ψ(g2, x)ψ(g1 · g2, x)−1ψ(g1, ϑg2(x)) REPRESENTATIONS BY UNBOUNDED OPERATORS 63 algebra for a 2-cocycle ϕ: G × G → T are exactly the projective representations π : G → U(H) with π(g)π(h) = ϕ(g, h)π(gh) for all g, h ∈ G. The group Z has no nontrivial 2-cocycles. They do appear, however, for the group Z2. A well known example is the noncommutative torus. Its usual gauge action corresponds to a Z2-grading, where U nV m for the canonical generators U, V has degree (n, m) ∈ Z2. In this case, Ae = C = Be = B+ e has only one point. The transformation groupoid G (cid:110) A+ e is simply G = Z2. This is discrete, so Σ is always trivial as a principal circle bundle. Thus the only non-trivial aspect of Σ is a 2-cocycle ϕ: Z × Z → T. The cohomology group H2(Z2, T) is isomorphic to T, and the resulting twisted group algebras of Z2 are exactly the noncommutative tori. Proposition 11.11. If there are subsets Sg ⊆ Ag such that Sg generates Ag as a right Ae-module, Sg · Sh ⊆ Sgh, and χ(a∗b) ≥ 0 for all a, b ∈ Sg, g ∈ G, χ ∈ j(X) ⊆ Ae, then the twist ΣX is trivial and so the C∗-hull of A is C∗(G (cid:110) X+). Proof. If χ ∈ Dg−1, then there is b ∈ Ag with χ(b∗b) 6= 0. Since Sg generates Ag as e , and A+ a right Ae-module, we may write b =Pn nX i=1 ai · ci with ai ∈ Sg, ci ∈ Ae. Then χ(c∗ i )χ(cj)χ(a∗ i aj). χ(b∗b) = i,j=1 If A+ i aj) 6= 0. Then χ(a∗ e itself is locally compact, then we may take X+ = X = A+ S i ai) 6= 0 by (11.3). This shows that Hence there are i, j with χ(a∗ Ua = Dg−1. We have (g, χ, [a]) = (g, χ, [b]) for all a, b ∈ Sg, χ ∈ j(X)∩Ua∩Ub a∈Sg because χ(a∗b) ≥ 0 for all χ ∈ j(X). Hence the local sections σa of ΣX{g}×Dg−1 for a ∈ Sg coincide on the intersections of their domains and thus combine to a (cid:3) global trivialisation. This trivialisation is multiplicative as well. e with the e is closed in Ae, this happens if Ae is locally compact. inclusion map j. Since A+ Theorem 11.12. Assume that A+ e is locally compact in the topology τc. Call a representation of A integrable if its restriction to Ae is locally bounded. Let Σ be e , τc), Σ) is a C∗-hull for the twisted groupoid constructed above. Then C∗(G (cid:110) ( A+ the integrable representations of A. Proof. If X+ = ( A+ e , τc), then integrability is compatible with induction by Propo- sition 11.7 because the construction of the topology τc is natural and compatible with restriction to open subsets. Theorem 11.8 shows that C∗(G (cid:110) ( A+ e , τc), Σ) is a C∗-hull for the class of representations of A whose restriction to Ae is A+ e -integrable. The locally bounded representations of Ae are equivalent to the locally bounded representations of the pro-C∗-algebra C( Ae, τc) by Propositions 7.2 and 7.6. Re- strictions of representations of A to Ae are automatically inducible by Lemma 9.8. By a pro-C∗-algebraic variant of Lemma 11.1, the inducible, locally bounded rep- resentations of Ae are equivalent to those representations of C( Ae, τc) that factor through the quotient C( A+ e , τc) is dense e , τc) is a C∗-hull for the inducible, in the pro-C∗-algebra C( A+ (cid:3) locally bounded representations of Ae by Theorem 7.17. Assume that Ae is countably generated. Then the usual topology on Ae is metrisable and hence compactly generated, so that τc is the standard topology on Ae. A representation of Ae is locally bounded if and only if all symmetric elements of Ae act by regular, self-adjoint operators by Theorem 8.7. Thus a representation π of A is integrable as in Theorem 11.12 if and only if π(a) is regular and self-adjoint for all a ∈ Ae with a = a∗. This class of integrable representations has the C∗-hull C∗(G (cid:110) A+ e , τc). Since A+ e , τc). Hence C0( A+ e is locally compact, C0( A+ e , Σ) if A+ e is locally compact. 64 RALF MEYER In particular, if Ae is finitely generated, then Ae is mapped homeomorphically onto a closed subset of Rn for some n ∈ N by evaluating characters on a finite set of symmetric generators. Thus Ae is locally compact. The discussion above gives: Corollary 11.13. Assume that Ae is finitely generated. Call a representation of A integrable if its restriction to Ae is locally bounded. Then A+ e is locally compact and C∗(G (cid:110) A+ e , Σ) for the twisted groupoid Σ constructed above is a C∗-hull for the integrable representations of A. Moreover, a representation π of A is integrable if and only if π(a) is regular and self-adjoint for all a ∈ Ae with a = a∗. Corollary 11.13 covers all the examples considered in [7, 26], except for the enveloping algebra W of the Virasoro algebra that is studied in [26, §9.3]. The ∗-algebra W is Z-graded. Its unit fibre W0 is noncommutative. The first step in the study of its representations in [26, §9.3] is to replace W by a certain Z-graded quotient A := W/I, whose unit fibre A0 = W0/(I ∩ W0) is commutative by construction. The motivation is that all "integrable" representations of W factor through A. The main result in [26, §9.3] shows that the partial action of Z on A+ e is free and that the disjoint union Y := X1t X2t X3 of the three families of characters described in (61)–(63) of [26] is a fundamental domain, that is, it meets each orbit of the partial action exactly once. Each subset Xi is closed in Ae and locally compact and second countable in the subspace topology. Hence so is Y . Since Z acts by partial homeomorphisms and Y is a fundamental domain, there is a continuous bijection (D−n ∩ Y ) → A+ e , (n, y) 7→ ϑn(y). X := G n∈Z Each D−n ∩ Y is an open subset of Y , so that X is locally compact. I have not checked whether this continuous bijection is a homeomorphism. If so, then A+ would be locally compact and the results in [26] for the Virasoro algebra would be contained in Theorem 11.12 after passing to the quotient W/I. If not, we would use the locally compact space X. The partial action of Z on A+ e is clearly continuous on X as well, so that Theorem 11.8 applies. e 12. Rieffel deformation Let G be a discrete group. Given a normalised 2-cocycle on G, Rieffel deformation is a deformation functor that modifies the multiplication on a G-graded ∗-algebra by the 2-cocycle. There is a similar process for Fell bundles over G, which we may transfer to section C∗-algebras. This is how Rieffel deformation is usually considered. The setting of graded algebras or Fell bundles is easier. We now define Rieffel deformation more precisely and show that it is compatible with the construction of C∗-hulls in Theorem 9.26. This deformation process has also recently been treated in [22]. A normalised 2-cocycle on a group G is a function Λ: G × G → U(1) with Λ(e, g) = 1 = Λ(g, e) for all g ∈ G and (12.1) Λ(g, h · k)Λ(h, k) = Λ(g · h, k)Λ(g, h) G-graded vector space with the deformed multiplication and involution g∈G Ag be a G-graded algebra. Let AΛ be the same for all g, h, k ∈ G. Let A =L bh := X ag ∗X X g∈G h∈G g,h∈G (cid:0)X (cid:1)† :=X Λ(g−1, g)a∗ g, ag Λ(g, h)agbh, where ag, bg ∈ Ag for all g ∈ G. We call AΛ the Rieffel deformation of A with respect to Λ. REPRESENTATIONS BY UNBOUNDED OPERATORS 65 Lemma 12.2. The deformed multiplication and involution on AΛ give a G-graded ∗-algebra with a∗b = ab if a ∈ Ae or b ∈ Ae, and a†∗b = a∗b for all g ∈ G, a, b ∈ Ag. Proof. The multiplication remains associative by the 2-cocycle condition (12.1). The normalisation of Λ and (12.1) for g, g−1, g give Λ(g, g−1) = Λ(g−1, g) for all g ∈ G. Thus g)† = Λ(g, g−1) · Λ(g−1, g)(a∗ (a† g)∗ = ag for ag ∈ Ag. The normalisation condition and (12.1) for g, h, h−1 and gh, h−1, g−1 for g, h ∈ G give Λ(gh, h−1)Λ(g, h) = Λ(h, h−1), Λ(g, g−1)Λ(gh, h−1) = Λ(gh, h−1g−1)Λ(h−1, g−1). Hence Λ(g, g−1)Λ(h, h−1) = Λ(g, h)Λ(gh, h−1g−1)Λ(h−1, g−1). This implies the condition (ag ∗ bh)† = b g for ag ∈ Ag, bh ∈ Ag: † h ∗ a† (ag ∗ bh)† = Λ(g, h) · Λ(gh, (gh)−1) · (agbh)∗ = Λ(g, g−1) · Λ(h, h−1) · Λ(h−1, g−1) · b∗ ha∗ † h ∗ a† g = b g. Thus the deformed multiplication and involution give a ∗-algebra. The formula a† ∗ b = a∗b for g ∈ G, a, b ∈ Ag is trivial, and a ∗ b = ab if a ∈ Ae or b ∈ Ae follows (cid:3) from the normalisation of Λ. g = Λ(g−1, g)a∗ The same formulas work if (Bg)g∈G is a Fell bundle over G. Let (BΛ g )g∈G be the same Banach space bundle as Bg with the multiplication and involution g for g, h ∈ G, ag ∈ Bg, bh ∈ Bh. By ag ∗ bh := Λ(g, h)agbh and a† Lemma 12.2, the deformation does not change ab for a ∈ Be or b ∈ Be and a∗b and ab∗ for a, b ∈ Bg. Hence BΛ g = Bg as Hilbert Be-bimodules, so that the positivity and completeness conditions for a Fell bundle are not affected by the deformation. We call (BΛ g )g∈G the Rieffel deformation of the Fell bundle (Bg)g∈G with respect to Λ. For a C∗-algebra of the form B = C∗(Bg) for a Fell bundle (Bg)g∈G over G, we define its Rieffel deformation with respect to Λ as BΛ := C∗(BΛ g ) for the deformed Fell bundle. If G is an Abelian group, then C∗(Bg) for a Fell bundle over G carries a canonical continuous action of G, called the dual action. Conversely, any C∗-algebra with a continuous G-action β is of the form B = C∗(Bg), where (Bg)g∈G is the spectral decomposition of the action, Bg = {b ∈ B βχ(b) = χ(g) · b for all χ ∈ G}. Thus Rieffel deformation takes a C∗-algebra with a continuous G-action to another C∗-algebra with a continuous G-action. This is how it is usually formulated. Since G is compact, there are no analytic difficulties with oscillatory integrals as in [25]. g∈G Ag be a G-graded ∗-algebra and let Be be a C∗-hull for a class of integrable representations of Ae. Assume that integrability is compatible with induction for A. Let Λ be a normalised 2-cocycle on G. Then integrability is also compatible with induction for AΛ, and the C∗-hull for the integrable representations of AΛ is the Rieffel deformation with respect to Λ of the C∗-hull for the integrable representations of A. Proof. The compatibility condition in Definition 9.18 is equivalent to the integrability of Ag ⊗Ae (E, π) for all g ∈ G, which only involves a single Ag with its Ae-bimodule structure and the Ae-valued inner product ha, bi = a∗b for a, b ∈ Ag. This is not Theorem 12.3. Let A =L 66 RALF MEYER changed by Rieffel deformation by Lemma 12.2. Hence AΛ inherits the compatibility condition from A, and Theorem 9.26 applies to both A and AΛ. The Hilbert B+ e -bimodule B+ g depends only on Ag with the extra structure above and the universal inducible, integrable representation (B+ e ) of Ae by Remark 9.27. Since none of this is changed by Rieffel deformation, the Fell bundle g . Rieffel deformation changes obtained from AΛ has the same fibres (B+ the multiplication maps Ag × Ah → Agh and the involution Ag → Ag−1 for fixed g, h ∈ G only by a scalar. Inspecting the construction above, we see that the h → B+ multiplication maps B+ g−1 in the Fell bundle are changed by exactly the same scalars. Hence the Fell bundle for AΛ (cid:3) is (B+ g )Λ. Now the assertion follows from Theorem 9.26. gh and the involution B+ e , µ+ g → B+ g )Λ as B+ g × B+ 13. Twisted Weyl algebras 0 = N associated to the A0-bimodule A1 acts on A+ and the relation aa∗ = a∗a + 1. There is a unique Z-grading A =L We illustrate our theory by studying C∗-hulls of twisted n-dimensional Weyl algebras for 1 ≤ n ≤ ∞. We begin with the case n = 1, where no twists occur. Then we consider the case of finite n without twists and with twists. Finally, we consider the case n = ∞ with and without twists. The (1-dimensional) Weyl algebra A is the universal ∗-algebra with one generator a n∈Z An with a ∈ A1. The ∗-subalgebra A0 is isomorphic to the polynomial algebra C[N] with N = a∗a, which is commutative. The other subspaces Ak ⊆ A for k ∈ N are isomorphic to A0 as left or right A0-modules because Ak = A0 · ak = ak · A0 and A−k = (a∗)k · A0 = A0 · (a∗)k for all k ≥ 0. The spectrum A0 of A0 is R, where the character C[N] → C for t ∈ R evaluates a polynomial at t. A character is positive if and only if it is positive on (a∗)kak and ak(a∗)k for all k ≥ 1. This happens if and only if t ∈ N by [26, Example 10]. Since N a = a∗aa = (aa∗ − 1)a = a · (a∗a − 1) = a · (N − 1), the partial automorphism ϑ1 of A+ e by the automorphism N 7→ N − 1, which corresponds to translation by −1. By induction, we get N · ak = a · (N − 1) · ak−1 = ··· = ak · (N − k). The domain of ϑk is as big as it could possibly be, that is, it contains all n ∈ N with n ≥ k by (11.5) (see also [26, Example 16]). For any k, l ∈ N there is a unique n ∈ Z with k − n = l. Thus the transformation groupoid Z (cid:110)ϑ N is simply the pair groupoid on N. There can be no twist in this case. First, the pair groupoid simply has no non-trivial twists. And secondly, the generators ak, (a∗)k for k ≥ 0 satisfy the positivity condition in Proposition 11.11, which also rules out a twist. Since no proper non-empty subset of N is invariant under the partial action ϑ of Z, a commutative C∗-hull for A0 for which integrability is compatible with 0 = {0}. In the second case, A has no induction gives either B+ non-zero integrable representations. In the first case, the C∗-hull for the integrable representations of A is the groupoid C∗-algebra K('2N) of the pair groupoid N × N. The universal representation of A on K('2N) is equivalent to a representation π of A on '2N by Proposition 3.13. The domain of this representation is the space S(N) kδk−1 for k ∈ N, so π(a∗)(δk) = of rapidly decreasing sequences, with π(a)(δk) = √ k + 1δk+1, π(N)(δk) = kδk. By Theorem 4.4, a representation π of A0 on a Hilbert module E is integrable if and only π(N k) is regular and self-adjoint for each k ∈ N or, equivalently, π(N) is regular and self-adjoint and π(N k) = π(N)k for all k ∈ N. By definition, a representation of A is integrable if and only if its restriction to A0 is integrable. The Z-grading on the C∗-hull K('2N) is "inner": it is induced by the Z-grading on '2N where δk has degree k. Equivalently, the dual action of T on K('2N) associated 0 = C0(N) or B+ √ j j REPRESENTATIONS BY UNBOUNDED OPERATORS 67 to the Z-grading is the inner action associated to the unitary representation U : T → U('2N), where Uz(δk) := zkδk for all z ∈ T, k ∈ N. j = a∗ j ak = λ−1 a∗ j aj + 1 for 1 ≤ j ≤ m and jk aka∗ Now let m ∈ N and let Θ = (Θjk) be an antisymmetric m × m-matrix. Let λjk = exp(2πiΘjk). Let Am,Θ be the ∗-algebra with generators a1, . . . , am and the commutation relations aja∗ (13.1) ajak = λjkakaj, for 1 ≤ j 6= k ≤ m. Since λjk = λ−1 kj , the relations (13.1) for (j, k) and (k, j) are equivalent; so it suffices to require (13.1) for 1 ≤ j < k ≤ m. The ∗-algebra Am,Θ is Zm-graded by giving aj degree ej ∈ Zm, where e1, . . . , em is the standard basis of Zm. We first consider the case Θ = 0 and write Am := Am,0. This is the m-dimensional Weyl algebra, which is the tensor product of m copies of the 1-dimensional Weyl for 0 ∈ Zm is algebra, with the induced Zm-grading. Thus the zero fibre Am0 isomorphic to the polynomial algebra C[N1, . . . , Nm] in the m generators Nj = a∗ j aj. k for k ∈ Zm is isomorphic to Am0 both as a left and Its spectrum is Rm. Each Am j)−kj for a right Am0 -module; the generator is the product of akj kj < 0 from j = 1, . . . , m. Here the order of the factors does not matter because k with the exterior tensor product of the A1-bimodules Θ = 0. We may identify Am k1 ⊗ A1 . Hence the space of positive characters on Am is Nm, and the A1 partial action of Zm on Nm is the exterior product of the partial actions of Z on N for the 1-dimensional Weyl algebras. That is, k ∈ Zm acts on Nm by translation by −k with the maximal possible domain. Thus the transformation groupoid Zm (cid:110) A+ 0 is isomorphic to the pair groupoid of the discrete set Nm. Once again, the only Zm-invariant subsets of A+ 0 are the empty set and Nm, so that the only inducible commutative C∗-hulls of A0 for which integrability is compatible with induction are {0} and C0(Nm). The first case is boring, and the second case leads to the C∗-hull K('2Nm) of the m-dimensional Weyl algebra. As for m = 1, the universal representation of Am is equivalent to a representation on '2Nm. This has the domain S(Nm), and the representation is determined by for kj ≥ 0 or (a∗ k2 ⊗···⊗ A1 π(aj)(δ(k1,...,km)) =pkjδ(k1,...,kj−1,...,km) for (k1, . . . , km) ∈ Nm and j = 1, . . . , m. Hence π(Nj)(δ(k1,...,km)) = kjδ(k1,...,km). A representation of A is integrable if and only if its restriction to A0 is integrable in the sense that it integrates to a representation of C0(Rm). This automatically descends to a representation of C0(Nm) by Lemma 9.8. There are several ways to characterise when a representation of C[N1, . . . , Nm] integrates to a representation of C0(Rm). One is that π(Nj) for j = 1, . . . , m are strongly commuting, regular, j ) = π(Nj)k for all j = 1, . . . , m, k ∈ N, compare self-adjoint operators and π(N k [27, Theorem 9.1.13]. The groups Zm for m ≥ 2 have non-trivial 2-cocycles, and Am,Θ is, by definition, km a Rieffel deformation of Am,0 for the normalised 2-cocycle (13.2) Λ(cid:0)(x1, . . . , xm), (y1, . . . , ym)(cid:1) := mY mY mY xkyj−xj yk = pλjk mY mY j=1 k=j+1 λxkyj jk . exp(−πiΘjkxjyk). We could also use the cohomologous antisymmetric 2-cocycle j=1 k=j+1 j,k=1 Theorem 12.3 says that the C∗-hull Bm,Θ of Am,Θ is the Rieffel deformation of the C∗-hull Bm,0 of Am,0 for the same 2-cocycle. 68 RALF MEYER In the classification of Fell bundles with commutative unit fibre, the important cohomology is that of the transformation groupoid G (cid:110) X+, not of G itself. Here G (cid:110) X+ is the pair groupoid of Nm. Lemma 13.3 (compare [17, Lemma 2.9]). The cohomology of the pair groupoid of a discrete set X with coefficients in an Abelian group H vanishes in all positive degrees. Proof. The set of composable n-tuples in the pair groupoid of X is X n+1. The groupoid cohomology with coefficients H is the cohomology of the chain complex with cochains Cn := H X n+1, the space of all maps X n+1 → H, and with the boundary map ∂ : Cn → Cn+1, ∂ϕ(x0, . . . , xn) := nX (−1)iϕ(x0, . . . ,bxi, . . . , xn); i=0 here the hat means that the entry xi is deleted. Pick some point x0 ∈ X and let hϕ(x1, . . . , xn) := ϕ(x0, x1, . . . , xn) for all n ∈ N, x1, . . . , xn ∈ X, ϕ ∈ Cn+1. Then ∂ ◦ h(ϕ) + h ◦ ∂(ϕ) = ϕ for all ϕ ∈ Cn, n ≥ 1. Thus the cohomology vanishes in (cid:3) positive degrees. Any twist of the pair groupoid on Nm is trivial by Lemma 13.3. Therefore, the C∗-hull Bm,Θ is isomorphic to K('2Nm), the untwisted groupoid C∗-algebra of the pair groupoid. The proof of Lemma 13.3 is explicit and so allows to construct this isomorphism. We explain another way to construct it, using properties of Rieffel deformation. Since the Z-grading on the C∗-hull K('2N) is inner or, equivalently, the corresponding action of T is inner, the same holds for the Zm-grading on the C∗-hull K('2Nm) and the corresponding Tm-action on K('2Nm). Explicitly, the Tm-action is induced by the unitary representation of Tm on '2Nm defined by U(z1,...,zm)δ(k1,...,km) := z −k1 1 ··· z−km m δ(k1,...,km). Rieffel deformation of C∗-algebras for inner actions does not change the isomorphism type of the C∗-algebra. Hence the C∗-hull for the integrable representations of Am,Θ is also isomorphic to K('2Nm). We make this more explicit in our Fell bundle language. Let U : Tm → UM(B) be a strictly continuous homomorphism to the group of unitary multipliers of a z for z ∈ Tm, b ∈ B be the resulting inner C∗-algebra B and let αz(b) := UzbU∗ action. Let (Bk)k∈Zm be the spectral decomposition of this action, that is, b ∈ Bk if and only if αz(b) = zk · b for all z ∈ Tm. In particular, U ∈ UM(B0) because Tm is commutative. Assume for simplicity that the 2-cocycle Λ is a bicharacter as above. For fixed k ∈ Zm, we view Λ(k, ॷ): Zm → T as an element Λ(k) of the dual group Tm. The map Λ: Zm → Tm is a group homomorphism. The maps ψk : Bk → Bk, b 7→ U∗ Λ(k)·b, for k ∈ N are Banach space isomorphisms that modify the multiplication as follows: ψk(b1)ψl(b2) = U∗ Λ(k)b1U∗ Λ(l)b2 = U∗ Λ(k+l)αΛ(l)(b1)b2 = ψk+l(Λ(k, l) · b1b2). They keep the involution unchanged. This is exactly what Rieffel deformation does. Hence the maps ψk form an isomorphism between the undeformed and Rieffel deformed Fell bundles. This finishes the proof that the Rieffel deformed algebra for an inner action is canonically isomorphic to the original algebra. The universal representation of Am,Θ on K('2Nm) again corresponds to a repre- sentation of Am,Θ on '2(Nm). We may construct it by carrying over the isomorphism Bm,Θ ∼= Bm,0 between the C∗-hulls. This is the inverse of the isomorphism above, so it multiplies on the left by the unitary UΛ(k) of degree 0 on elements of degree k. for k ∈ Zm act We do exactly the same on elements of Am,Θ and so let x ∈ Am,Θ k REPRESENTATIONS BY UNBOUNDED OPERATORS 69 on '2Nm by the operator UΛ(k)πm,0(x), where πm,0 is the universal representation of the untwisted Weyl algebra Am,0 on '2Nm described above. The same computa- tion as above shows that this defines a ∗-representation of Am,Θ. We compute it explicitly. First, the action of elements of Am,Θ on '2Nm is not changed. The domain of a representation of Am,Θ is equal to the domain of its restriction to Am,Θ . Hence the domain of our representation is the Schwartz space S(Nm), as in the untwisted case. Let 1 ≤ j ≤ m. The generator aj has degree ej ∈ Zm, and 0 0 Λ(ej) = (λ1,j, . . . , λj−1,j, 1, . . . , 1) ∈ Tm for our first definition of Λ in (13.2). Thus (cid:18)j−1Y (cid:19)pkjδ(k1,...,kj−1,...,km). πm,Θ(aj)δ(k1,...,km) = UΛ(ej)πm,0(aj)δ(k1,...,km) = These operators on S(Nm) satisfy the defining relations of Am,Θ. λkl j,l l=1 j aj = aja∗ j ak = aka∗ j + 1, ajak = akaj, and a∗ The infinite-dimensional Weyl algebra A∞ is the universal ∗-algebra with genera- tors aj for j ∈ N and relations a∗ j for 0 ≤ j < k. Let Z[N] be the free Abelian group on countably many generators. The Weyl algebra A∞ is Z[N]-graded, where aj has degree ej ∈ Z[N], the jth generator of Z[N]. The ∗-algebra A∞ is a tensor product of infinitely many 1-dimensional Weyl algebras. The zero fibre A∞ 0 is the polynomial algebra in the generators Nj = a∗ j aj for j ∈ N. Hence its spectrum is the infinite product A∞ 0 = RN, which is not locally compact. The tensor product structure of A∞ shows that a character is positive 0 ∼= NN is a product of countably if and only if each component is. That is, (A∞)+ many copies of the discrete space N. Since N is not compact, this is not locally compact either. Hence to build a commutative C∗-hull for A∞ 0 , we must choose some locally compact space X with a continuous, injective map f : X → NN. Here we have simplified notation by assuming that already f(X) ⊆ NN; otherwise, the first step in our construction would replace X by X+ := f−1(NN). For X-integrability to be compatible with induction, we also need f(X) to be invariant under the partial action of Z[N], and we need the restricted partial action to lift to a continuous partial action on X. The partial action of the group Z[N] on NN is the obvious one by translations. It is free, and two points (nk) and (n0 k) in NN belong to the same orbit if and only k for all k ≥ k0. Briefly, such points are called tail if there is k0 such that nk = n0 equivalent. This partial action is minimal, that is, an open, Z[N]-invariant subset is either empty or the whole space. Hence NN has no Z[N]-invariant, locally closed subsets. Thus NN has no Z[N]-invariant, locally compact subspaces. Let K be any compact subset of NN. Then the projection pj : NN → N to the jth coordinate must map K to a compact subset of N. So there is an upper bound Mj j∈N[0, Mj]N, and the right hand with pj(K) ⊆ [0, Mj]N := [0, Mj] ∩ N. Then K ⊆Q side is compact. The closure ofQ Nj ×Y X(Mk) := [ j∈N[0, Mj]N under tail equivalence is [0, Mk]N [0, Mj]N. There is a unique topology on X(Mk) where each subset Nj ×Q the restricted product of copies of N with respect to the compact-open subsets k>j[0, Mk]N is open and carries the product topology. This topology is locally compact, and the partial action of Z[N] on X(Mk) by translation is continuous. k>j (cid:18) (cid:19) , j∈N 70 RALF MEYER 0 by Theorem 8.3. Applying induction from A∞ Lemma 13.4. The Local–Global Principle fails for the X(Mk)-integrable represen- tations of A∞. Proof. Since the map X → A∞ is not a homeomorphism onto its image and A∞ 0 0 is metrisable, the Local–Global Principle fails for the X-integrable representations of A∞ 0 to A∞ to a counterexample for the Local–Global Principle for A∞ 0 produces such a counterexample also for A∞. Explicitly, choose a sequence (nk) such that nk > Mk for infinitely many k. Let xk := nkδk ∈ NN, that is, xk ∈ NN has only one non-zero entry, which is nk in the kth place. This sequence belongs to X(Mk) and converges to 0 in the product topology on A∞ 0 , but not in the topology of X(Mk). The resulting representation 0 on C(¯N) is not X(Mk)-integrable, but it becomes integrable when we tensor of A∞ with any Hilbert space representation of C(¯N), see the proof of Theorem 8.3. Now to a representation of A∞ on C(¯N). induce this (inducible) representation of A∞ 0 (cid:3) This gives a counterexample for the Local–Global Principle for A∞. I do not know a class of integrable representations of A∞ with a C∗-hull for which the Local–Global Principle holds. j. If χ ∈ NN is a positive character Let S be the set of all words in the letters aj, a∗ and x, y ∈ S ∩ A∞ k for some k ∈ Z[N], then χ(x∗y) ≥ 0. Hence Proposition 11.11 shows that there is no twist in our case, that is, the C∗-hull of the X(Mk)-integrable representations of A∞ is the groupoid C∗-algebra of the transformation groupoid Z[N] (cid:110) X(Mk). This C∗-hull is canonically isomorphic to one of the host algebras for A∞ constructed in [11], namely, to the one that is denoted L[n] in [11] with nk = Mk + 1 for all k ∈ N. We remark in passing that the construction of a full host algebra from these host algebras in [11] is wrong: the resulting C∗-algebra has too many Hilbert space representations, so it is not a host algebra any more. An erratum to [11] is currently being written. The compact subset T := Q UHF-algebra forQ algebrasN sentations of the (stabilisation of the) UHF-algebra of typeQ k∈N[0, Mk]N that we started with is a complete transversal in Z[N] (cid:110) X(Mk), that is, the range map in Z[N] (cid:110) X(Mk) restricted to s−1(T) is an open surjection onto X(Mk). Hence the groupoid Z[N] (cid:110) X(Mk) is Morita equivalent to its restriction to the compact subset T. This restriction is the tail equivalence relation on T. Its groupoid C∗-algebra is well known: it is the k∈N(Mk + 1), that is, the infinite tensor product of the matrix k∈N MMk+1. The C∗-algebra of Z[N](cid:110) X(Mk) itself is the C∗-stabilisation Thus the X(Mk)-integrable representations of A∞ are equivalent to the repre- k∈N(Mk + 1). This depends very subtly on the choice of the sequence (Mk). There is no canoni- cal ∗-homomorphism between these UHF-algebras if we increase (Mk): for some (Mk) ≤ (M0 k), there is not even a non-zero map between their K-theory groups. Instead, there are canonical morphisms, that is, there is a canonical nondegenerate k)) → M(C∗(Z[N] (cid:110) X(Mk))) if (Mk) ≤ (M0 ∗-homomorphism C∗(Z[N] (cid:110) X(M0 k). They are constructed as follows. The inclusion map X(Mk) ,→ X(M0 k) is contin- uous with dense range, but not proper. Thus it induces an injective, nondegen- k)) → Cb(X(Mk)). Therefore, if a representation erate ∗-homomorphism C0(X(M0 of A∞ k)-integrable. If a representation is X(Mk)-integrable, then it is also X(M0 0 of A∞ has X(Mk)-integrable restriction to A∞ is also k)-integrable. When we apply this to the universal representation of A∞ X(M0 on the C∗-hull C∗(Z[N] (cid:110) X(Mk)), this gives the desired canonical morphism C∗(Z[N] (cid:110) X(M0 It is injective, say, because C∗(Z[N] (cid:110) X(M0 of this UHF-algebra. 0 , then its restriction to A∞ 0 k)) → C∗(Z[N] (cid:110) X(Mk)) if (Mk) ≤ (M0 k). k)) is simple. REPRESENTATIONS BY UNBOUNDED OPERATORS 71 j = a∗ jk aka∗ Now let Θ = (Θjk)j,k∈N be a skew-symmetric matrix. It corresponds first to a matrix λjk := exp(2πiΘjk) and then to a 2-cocycle Λ on Z[N] as in (13.2). The Rieffel deformation of A∞ by Θ is the universal ∗-algebra A∞,Θ with the same generators (aj)j∈N and the relations aja∗ j aj + 1, ajak = λjkakaj, and j for 0 ≤ j < k. We define X(Mk) for a sequence (Mk) and the j ak = λ−1 a∗ X(Mk)-integrable representations of A∞,Θ as above. By Theorem 12.3, this has a C∗-hull, namely, the Rieffel deformation of the C∗-hull for the X(Mk)-integrable representations of the undeformed Weyl algebra. The Rieffel deformation gives a twist of the groupoid Z[N] (cid:110) X(Mk), and the C∗-hull is the twisted groupoid C∗-algebra of Z[N] (cid:110) X(Mk) for this twist. Proposition 13.5. Let (Mk) ∈ NN. The C∗-hulls for the X(Mk)-integrable repre- sentations of the twisted Weyl algebras A∞,Θ are isomorphic for all Θ. Proof. The C∗-hull of A∞,Θ is a twisted groupoid algebra of the transformation groupoid Z[N] (cid:110) X(Mk), which is isomorphic to the tail equivalence relation R on X(Mk). We are going to prove that any twist X(Mk) × T (cid:26) Σ (cid:16) R is trivial. Hence the C∗-hull of A∞,Θ is isomorphic to the untwisted groupoid C∗-algebra of R for all Θ. subsets Rd ⊆ R with R =S Rd, and each Rd is also an equivalence relation. The The arrow space of R is totally disconnected because X(Mk) is totally disconnected and R is étale. Hence any locally trivial principal bundle over R is trivial. Thus Σ ∼= R×T as a topological space, and the twist is described by a continuous 2-cocycle ϕ: R(2) := R ×s,X,r R → T. We must show that ϕ is a coboundary. Let Rd for d ∈ N be the equivalence relation on X(Mk) defined by (nk) Rd (n0 k) k for all k ≥ d. This is an increasing sequence of open if and only if nk = n0 equivalence relation Rd is isomorphic to the product of the pair groupoid on Nd and the space X(Mk+d) for the shifted sequence (Mk+d)k∈N. Thus the cohomology of Rd with coefficients in T is isomorphic to the cohomology of the pair groupoid on Nd with values in the Abelian group of continuous map X(Mk+d) → T. This cohomology vanishes in positive degrees by Lemma 13.3. Therefore, for each d ∈ N there is ψd : Rd → T such that ϕRd : Rd×s,r Rd → T is the coboundary ∂ψd. The restriction of ψd+1 to Rd and ψd both have coboundary ϕRd. Hence ψ−1 d+1Rd · ψd is a 1-cocycle on Rd. Again by Lemma 13.3, there is χ: X → T with ψ−1 d+1Rd · ψd = ∂Rd χ. We d+1 := ψd+1· ∂Rd+1χ, where ∂Rd+1 χ means the coboundary of χ on replace ψd+1 by ψ0 d+1Rd = ψd. the groupoid Rd+1. This still satisfies ∂ψ0 d : Rd → T for all d ∈ N that satisfy Proceeding like this, we get continuous maps ψ0 d+1Rd = ψ0 d = ϕRd for all d ∈ N. These combine to a continuous map ψ0 ψ0 : R → T with ∂ψ0 = ϕ. (cid:3) d+1 = ∂ψd+1 = ϕRd+1, and ψ0 d and ∂ψ0 References [1] Saad Baaj and Pierre Julg, Théorie bivariante de Kasparov et opérateurs non bornés dans les C∗-modules hilbertiens, C. R. Acad. Sci. Paris Sér. I Math. 296 (1983), no. 21, 875–878, available at http://gallica.bnf.fr/ark:/12148/bpt6k55329549/f21.item. MR 715325 [2] Nicolas Bourbaki, Topologie générale. Chapitres 1 à 4, Éléments de mathématique, Hermann, Paris, 1971. MR 0358652 [3] Alcides Buss and Ruy Exel, Fell bundles over inverse semigroups and twisted étale groupoids, J. Operator Theory 67 (2012), no. 1, 153–205, available at http://www.theta.ro/jot/archive/ 2012-067-001/2012-067-001-007.html. MR 2881538 [4] Alcides Buss and Ralf Meyer, Inverse semigroup actions on groupoids, Rocky Mountain J. Math. 47 (2017), no. 1, 53–159, doi: 10.1216/RMJ-2017-47-1-53. MR 3619758 [5] , Iterated crossed products for groupoid fibrations (2016), eprint. arXiv: 1604.02015. [6] Alcides Buss, Ralf Meyer, and Chenchang Zhu, A higher category approach to twisted actions on C∗-algebras, Proc. Edinb. Math. Soc. (2) 56 (2013), no. 2, 387–426, doi: 10.1017/S0013091512000259. MR 3056650 72 RALF MEYER [7] Philip A. Dowerk and Yuriı Savchuk, Induced ∗-representations and C∗-envelopes of some quantum ∗-algebras, J. Lie Theory 23 (2013), no. 1, 229–250, available at http: //www.heldermann.de/JLT/JLT23/JLT231/jlt23013.htm. MR 3060775 [8] Ruy Exel, Partial dynamical systems, Fell bundles and applications, 2015, eprint. arXiv: [9] Jürgen Friedrich and Konrad Schmüdgen, n-positivity of unbounded ∗-representations, Math. 1511.04565. Nachr. 141 (1989), 233–250, doi: 10.1002/mana.19891410122. MR 1014429 [10] Hendrik Grundling, A group algebra for inductive limit groups. Continuity problems of the canonical commutation relations, Acta Appl. Math. 46 (1997), no. 2, 107–145, doi: 10.1023/A:1017988601883. MR 1440014 [11] Hendrik Grundling and Karl-Hermann Neeb, Full regularity for a C∗-algebra of the canonical commutation relations, Rev. Math. Phys. 21 (2009), no. 5, 587–613, doi: 10.1142/S0129055X09003670. MR 2533429 , Infinite tensor products of C0(R): towards a group algebra for R(N), J. Operator [12] [13] Maria Joiţa, A new look at the crossed products of pro-C∗-algebras, Ann. Funct. Anal. 6 Theory 70 (2013), no. 2, 311–353, doi: 10.7900/jot.2011aug22.1930. MR 3138360 (2015), no. 2, 184–203, doi: 10.15352/afa/06-2-16. MR 3292525 [14] Jens Kaad and Matthias Lesch, A local global principle for regular operators in Hilbert C∗-modules, J. Funct. Anal. 262 (2012), no. 10, 4540–4569, doi: 10.1016/j.jfa.2012.03.002.MR 2900477 [15] Alexander Kumjian, On C∗-diagonals, Canad. J. Math. 38 (1986), no. 4, 969–1008, doi: [16] E. Christopher Lance, Hilbert C∗-modules, London Mathematical Society Lecture Note Series, vol. 210, Cambridge University Press, Cambridge, 1995. doi: 10.1017/CBO9780511526206 MR 1325694 10.4153/CJM-1986-048-0. MR 854149 [17] Marius Măntoiu, Radu Purice, and Serge Richard, Twisted crossed products and magnetic pseudodifferential operators, Advances in operator algebras and mathematical physics, Theta Ser. Adv. Math., vol. 5, Theta, Bucharest, 2005, pp. 137–172. arXiv: math-ph/0403016. MR 2238287 [18] Karl-Hermann Neeb, Semibounded unitary representations of double extensions of Hilbert-loop groups, Ann. Inst. Fourier (Grenoble) 64 (2014), no. 5, 1823–1892, doi: 10.5802/aif.2898. MR 3330925 [19] Arupkumar Pal, Regular operators on Hilbert C∗-modules, J. Operator Theory 42 (1999), no. 2, 331–350, available at http://www.theta.ro/jot/archive/1999-042-002/1999-042-002-005. html. MR 1716957 [20] François Pierrot, Opérateurs réguliers dans les C∗-modules et structure des C∗-algèbres de groupes de Lie semisimples complexes simplement connexes, J. Lie Theory 16 (2006), no. 4, 651–689, available at http://www.heldermann.de/JLT/JLT16/JLT164/jlt16037.htm.MR 2270655 [21] N. Christopher Phillips, Inverse limits of C∗-algebras, J. Operator Theory 19 (1988), no. 1, 159–195, available at http://www.theta.ro/jot/archive/1988-019-001/1988-019-001-010. html. MR 950831 [22] Iain Raeburn, Deformations of Fell bundles and twisted graph algebras, Math. Proc. Cambridge Philos. Soc. 161 (2016), no. 3, 535–558, doi: 10.1017/S0305004116000359. MR 3569160 [23] Jean Renault, Représentation des produits croisés d'algèbres de groupoïdes, J. Operator The- ory 18 (1987), no. 1, 67–97, available at http://www.theta.ro/jot/archive/1987-018-001/ 1987-018-001-005.html. MR 912813 , Cartan subalgebras in C∗-algebras, Irish Math. Soc. Bull. 61 (2008), 29–63, available [24] at http://www.maths.tcd.ie/pub/ims/bull61/S6101.pdf. MR 2460017 (1993), no. 506, x+93, doi: 10.1090/memo/0506. MR 1184061 [25] Marc A. Rieffel, Deformation quantization for actions of Rd, Mem. Amer. Math. Soc. 106 [26] Yuriı Savchuk and Konrad Schmüdgen, Unbounded induced representations of ∗-algebras, Algebr. Represent. Theory 16 (2013), no. 2, 309–376, doi: 10.1007/s10468-011-9310-6. MR 3035996 [27] Konrad Schmüdgen, Unbounded operator algebras and representation theory, Operator Theory: Advances and Applications, vol. 37, Birkhäuser Verlag, Basel, 1990. doi: 10.1007/978-3-0348- 7469-4 MR 1056697 , On well-behaved unbounded representations of ∗-algebras, J. Operator Theory 48 (2002), no. 3, suppl., 487–502, available at http://www.theta.ro/jot/archive/2002-048-003/ 2002-048-003-002.html. MR 1962467 [29] Stéphane Vassout, Unbounded pseudodifferential calculus on Lie groupoids, J. Funct. Anal. [28] 236 (2006), no. 1, 161–200, doi: 10.1016/j.jfa.2005.12.027. MR 2227132 REPRESENTATIONS BY UNBOUNDED OPERATORS 73 [30] Stanisław Lech Woronowicz, Unbounded elements affiliated with C∗-algebras and noncompact quantum groups, Comm. Math. Phys. 136 (1991), no. 2, 399–432, available at http:// projecteuclid.org/euclid.cmp/1104202358. MR 1096123 , C∗-algebras generated by unbounded elements, Rev. Math. Phys. 7 (1995), no. 3, [31] 481–521, doi: 10.1142/S0129055X95000207. MR 1326143 E-mail address: [email protected] Mathematisches Institut, Georg-August Universität Göttingen, Bunsenstrasse 3–5, 37073 Göttingen, Germany
1807.10402
1
1807
2018-07-27T00:02:13
Unbounded Derivations in Bunce-Deddens-Toeplitz Algebras
[ "math.OA" ]
In this paper we study decompositions and classification problems for unbounded derivations in Bunce-Deddens-Toeplitz and Bunce-Deddens algebras. We also look at implementations of these derivations on associated GNS Hilbert spaces.
math.OA
math
UNBOUNDED DERIVATIONS IN BUNCE-DEDDENS-TOEPLITZ ALGEBRAS SLAWOMIR KLIMEK, MATT MCBRIDE, SUMEDHA RATHNAYAKE, KAORU SAKAI, AND HONGLIN WANG Abstract. In this paper we study decompositions and classification problems for un- bounded derivations in Bunce-Deddens-Toeplitz and Bunce-Deddens algebras. We also look at implementations of these derivations on associated GNS Hilbert spaces. 1. Introduction The study of derivations on C∗-algebras, which was started in 1953 by Kaplansky, had undergone several stages during its course: theory of bounded derivations, unbounded deriva- tions and noncommutative vector-fields, according to [1]. Originally motivated by research on dynamics in statistical mechanics, development of the theory of unbounded derivations in C∗-algebras began much later than its bounded counterpart; see [16]. The focus was on closability, generator properties and classification of closed derivations. More recently, classification and generator properties of derivations which are well behaved with respect to the action of a locally compact group were some of the major concerns [2]. Additionally, derivations feature in the theory of noncommutative vector fields [9], which was inspired by Connes work on noncommutative geometry [6]. In this paper we study classification and decompositions of unbounded derivations in Bunce-Deddens-Toeplitz and Bunce-Deddens algebras [3], [4]. Given an increasing sequence {lk}∞ k=0 of nonnegative integers such that lk divides lk+1 for k ≥ 0, the Bunce-Deddens- Toeplitz algebra is defined as the C∗-algebra of operators on ℓ2(Z≥0) generated by all lk- periodic weighted shifts for all k ≥ 0. Different sequences {lk} may lead to the same algebras, with the classifying invariant being the supernatural number N =Qp−prime pǫp, where ǫp := sup{j : ∃k pjlk}. In this paper we adopt a slightly different definition of the Bunce-Deddens- Toeplitz algebra A(N) associated with the supernatural number N that uses N more directly. We consider both finite and infinite N. The algebra K of compact operators on ℓ2(Z≥0) is contained in A(N) and the quotient A(N)/K := B(N) is known as the Bunce-Deddens algebra. The structure of all those algebras is quite different depending on whether N is finite or infinite. The main objects of study in this paper are densely defined derivations d : A(N) → A(N) in the Bunce-Deddens- Toeplitz algebras, where A(N) is the subalgebra of polynomials of lk-periodic weighted shifts, as well as derivations δ : B(N) → B(N) in the Bunce-Deddens algebras, where B(N) is the image of A(N) under the quotient map A(N) → A(N)/K = B(N). Intriguingly, if d : A(N) → A(N) is any derivation then d preserves the ideal of compact operators K, and consequently [d] : B(N) → B(N) defined by [d](a + K) = d(a) + K Date: July 30, 2018. 1 2 KLIMEK, MCBRIDE, RATHNAYAKE, SAKAI, AND WANG It is a non-trivial problem to describe properties of the map is a derivation in B(N). In general, on any C∗-algebra, bounded derivations preserve closed ideals and d 7→ [d]. so define derivations on quotients. It was proven in [13] that for bounded derivations and separable C∗-algebras the above map is onto, i.e., derivations can be lifted from quotients in separable cases but not in general. We prove here that lifting unbounded derivations from Bunce-Deddens to Bunce-Deddens-Toeplitz algebras is always possible when N is finite and conjecture that it is true for any supernatural number N. The main results of this paper are that any derivation in Bunce-Deddens or Bunce- Deddens-Toeplitz algebras can be uniquely decomposed into a sum of a certain special deriva- tion and an approximately inner derivation. The special derivations are not approximately inner, are explicitly described, and depend on whether N is finite or infinite. The algebra A(N) has a natural S1 action given by scalar multiplication of the generators, see formula (3.1), which also quotients to B(N). The key technique, like in [2], is to use Fourier series decomposition with respect to this action. The Fourier components of a derivation d satisfy a covariance property with respect to the S1 action. It turns out that such n-covariant derivations can be completely classified and their properties explicitly analyzed. We then use Ces`aro convergence of Fourier series to infer properties of d. Additionally, we describe implementations of derivations in various GNS Hilbert spaces associated with the algebras. Some of those implementations can be used to construct spectral triples on Bunce-Deddens-Toeplitz and Bunce-Deddens algebras, similarly to what was done in [11],[12]. 2. Definitions, Notations and preliminary results. In this section we introduce notation and terminology used in the paper. 2.1. Z/NZ rings. A supernatural number N is defined as the formal product: N = Yp−prime pǫp, ǫp ∈ {0, 1, · · · , ∞}. pǫp+ǫ′ p. their product is given by: If P ǫp < ∞ then N is said to be a finite supernatural number (a regular natural number), otherwise it is said to be infinite. If N ′ =Qp−prime pǫ′ NN ′ = Yp−prime p is another supernatural number, then A supernatural number N is said to divide M if M = NN ′ for some supernatural number N ′, or equivalently, if ǫp(N) ≤ ǫp(M) for every prime p. For the remainder of the paper we work with a fixed N. We let JN = {j : jN, j < ∞} be the set of finite divisors of N. Notice that (JN , ≤) is a directed set where j1 ≤ j2 if and only if j1j2N. Consider the collection of rings {Z/jZ}j∈JN and the family of ring homomorphisms πij : Z/jZ → Z/iZ, j ≥ i πij(x) = x (mod i) UNBOUNDED DERIVATIONS IN BUNCE-DEDDENS-TOEPLITZ ALGEBRAS 3 satisfying Then the inverse limit of the system can be denoted as: πik = πij ◦ πjk for all i ≤ j ≤ k. Z/NZ := lim ←− j∈JN Z/jZ =({xj} ∈ Yj∈JN Z/jZ : πij(xj) = xi) , and let πj : Z/NZ ∋ {xj} 7→ xj ∈ Z/jZ be the corresponding homomorphisms. In particular, if N is finite the above definition coincides with the usual meaning of the symbol Z/NZ, while if N = p∞ for a prime p, then the above limit is equal to Zp, the ring of p-adic integers, dee for example [15]. In general we have the following simple consequence of the Chinese Reminder Theorem. Proposition 2.1. If N = Qp−prime ǫp6=0 pǫp, then Z/NZ ∼= Qp−prime ǫp6=0 Z/pǫpZ. When the ring Z/NZ is equipped with the Tychonoff topology it forms a compact, abelian topological group. Thus it has a unique normalized Haar measure dHx. Also, if N is an infinite supernatural number then Z/NZ is a Cantor set [18]. Let qj : Z → Z/jZ be the quotient maps and let q : Z → Z/NZ be defined by: q(x) = {x (mod i)}. (2.1) We have the following simple property: πj ◦ q = qj. As a consequence of this and the structure of cylinder sets, we obtain the following observa- tion, needed later in the description of Bunce-Deddens algebras. Proposition 2.2. The range of q is dense in Z/NZ. We denote by E(Z/NZ) the space of locally constant functions on Z/NZ. This is a dense subspace of the space of continuous functions on Z/NZ. For f ∈ E(Z/NZ), consider the sequence: Then we have the following observation: af (k) = f (q(k)), k ∈ Z≥0. Proposition 2.3. If f ∈ E(Z/NZ), then there exists j ∈ JN such that af (k + j) = af (k) for every k ∈ Z≥0. Conversely, if a(k) is a j-periodic sequence for some j ∈ JN , then there is a unique f ∈ E(Z/NZ) such that a(k) = af (k). Proof. The result follows from an observation that any locally constant function on Z/NZ is a pullback via πj of a function on Z/jZ for some jN, see [14]. (cid:3) 2.2. BD algebras. Consider the Hilbert space ℓ2(Z≥0) equipped with the canonical basis {Ek}∞ k=0. Let U : ℓ2(Z≥0) → ℓ2(Z≥0) be the unilateral shift given by UEk = Ek+1. The adjoint of U is given by: U ∗Ek =(Ek−1 0 if k ≥ 1 if k = 0, 4 KLIMEK, MCBRIDE, RATHNAYAKE, SAKAI, AND WANG and we have the relation: U ∗U = I. We also use the following diagonal label operator: KEk = kEk. If {a(k)}∞ k=0 is a bounded sequence, then a(K) is a bounded operator given by: a(K)Ek = a(k)Ek. In numerous formulas below we use convention a(−1) = 0, so that, for example, we have: a(K − I)Ek =(a(k − 1)Ek 0 if k ≥ 1 if k = 0. Given a supernatural number N, we define the following algebra of diagonal operators: Adiag, per(N) = {a(K) : a(k) is j-periodic for some jN} . The norm closure of Adiag, per(N), denoted by Adiag, per(N), is a commutative unital C∗- algebra which, by Proposition 2.3, is canonically isomorphic to the C∗-algebra of continuous functions on Z/NZ: Adiag, per(N) =: Adiag, per(N) ∼= C(Z/NZ). (2.2) Definition 2.1. Given a supernatural number N, the Bunce-Deddens-Toeplitz algebra, de- noted by A(N), is the C∗-algebra of operators in ℓ2(Z≥0) generated by U and Adiag, per(N): A(N) = C∗(U, Adiag, per(N)). It is easy to see that for infinite N this definition coincides with the original definition [3], [4] given in the introduction. Let Adiag(N) be the commutative ∗-subalgebra of A(N) consisting of operators diagonal with respect to the canonical basis {Ek} of ℓ2(Z≥0). If the space of sequences which are eventually zero is denoted by c00, we define: Adiag(N) := {a(K) : a(k) = a0(k) + aper(k), a0(k) ∈ c00 and aper(K) ∈ Adiag, per(N)} which is a separable unital ∗-algebra. Some useful properties of this algebra are described in the following statement. Proposition 2.4. Adiag(N) is a dense ∗-subalgebra of Adiag(N). If the space of sequences converging to zero is denoted by c0, then we have the identification: Adiag(N) = Adiag(N) = {a(K) : a(k) = a0(k) + aper(k), a0(k) ∈ c0, aper(k) ∈ C(Z/NZ)}. Proof. Other than Adiag, per(N), the algebra Adiag(N) also contains additional diagonal op- erators that are in the algebra generated by the unilateral shift U. Those are precisely the compact diagonal operators: {a0(K) : a0(k) ∈ c0}, see [11]. The additive decomposition a(k) = a0(k) + aper(k) in Adiag(N) persists in completion Adiag(N) because compact diagonal operators form an ideal in Adiag(N), with the quotient isomorphic to C(Z/NZ). In fact, we have the following easy estimate: which implies directly the decomposition when passing to limits. (cid:3) ka(K)k = ka0(K) + aper(K)k ≥ kaper(K)k, UNBOUNDED DERIVATIONS IN BUNCE-DEDDENS-TOEPLITZ ALGEBRAS 5 Let A(N) denote the ∗-algebra generated algebraically by U, U ∗ and Adiag, per(N). We have the following description of A(N). Proposition 2.5. A(N) is a dense ∗-subalgebra of A(N). Moreover, we have the following description: A(N) :=na ∈ A(N) : a =Xn≥0 U na+ n,0(K) +Xn≥1 (U ∗)na− n,per(K), a± n,0(k) ∈ c00, a± +Xn≥1 a− U na+ n,per(K) n,0(K)(U ∗)n +Xn≥0 n,per(K) ∈ Adiag, per(N), finite sumso. Proof. By Proposition 3.1 of [11] the polynomials in U and U ∗ which are compact operators are precisely the finite sums of the form: Xn≥0 U na+ n,0(K) +Xn≥1 n,0(K)(U ∗)n, a− where a± n,0(k) ∈ c00. They form an ideal in A(N) so that, using additionally the commutation relation (2.3) below, all the remaining polynomials in U, U ∗ and Adiag, per(N) can be written as the last two terms in the statement of the proposition. (cid:3) If a(K) ∈ Adiag(N), then we have the commutation relation: a(K)U = Ua(K + I). (2.3) In fact, A(N) is the partial crossed product of Adiag(N) with Z≥0 where the action of Z≥0 on Adiag(N) is translation by one [8], [17]. In the trivial case of N = 1, the algebra A(1) is the Toeplitz algebra, i.e., the C∗-algebra generated by U. If N is finite, we can also identify A(N) as the tensor product of the Toeplitz algebra with matrices of size N × N (see [7] and also Section 4): A(N) ∼= A(1) ⊗ MN (C). If K are the compact operators in ℓ2(Z≥0), then K is an ideal in A(N), and we have the short exact sequence: 0 → K → A(N) ξ −→ B(N) → 0 where B(N) := A(N)/K and ξ : A(N) → A(N)/K is the quotient map. For any supernatural number N, we will call B(N) the Bunce-Deddens algebra. The Bunce-Deddens algebras are simple for infinite N, mutually non-isomorphic and have unique tracial state [3],[4],[7]. 2.3. Structure of BD algebras. We now proceed to a more detailed description of the Bunce-Deddens algebras B(N). Suppose {El}l∈Z is the canonical basis of ℓ2(Z), we let V : ℓ2(Z) → ℓ2(Z) be the bilateral shift given by: let L be the diagonal label operator: and let Bdiag(N) be defined as: V El = El+1, LEl = lEl, Bdiag(N) := {b(L) : b(l + j) = b(l) for some j N}. Notice that Bdiag(N) := Bdiag(N) is naturally isomorphic to C(Z/NZ), just like in (2.2). 6 KLIMEK, MCBRIDE, RATHNAYAKE, SAKAI, AND WANG Similarly to (2.3) we have the commutation relation: (2.4) For any N we introduce the Toeplitz-like operator T : B(ℓ2(Z)) → B(ℓ2(Z≥0)) given by b(L)V = V b(L + I). the formula: (2.5) where f ∈ ℓ2(Z≥0), and P : ℓ2(Z) → ℓ2(Z) is the orthogonal projection onto the subspace S = span{El : l ≥ 0}, which is naturally isomorphic with ℓ2(Z≥0). It is clear that we have: T (b)f = P bf, T (Iℓ2(Z)) = Iℓ2(Z≥0). The operator T is a linear, continuous, and ∗-preserving map between the spaces of bounded operators on ℓ2(Z) and ℓ2(Z≥0), and moreover it has the following properties: Lemma 2.6. For every a, b ∈ B(ℓ2(Z)) and any bounded diagonal operator b(L): (i) T (b V n) = T (b)U n and T (V −nb) = (U ∗)nT (b) for n ≥ 0 (ii) T (a b(L)) = T (a)b(K) (iii) T (b(L) a) = b(K)T (a). Proof. Those statements are obtained via direct calculations. For example, we have: T (bV n)f = P bV nf = P bP V nf = T (b)U nf because for n ≥ 0 the operator V n preserves S. Other calculations are similar. (cid:3) Since any element in C∗(V, Bdiag(N)) can be approximated by a finite sum of the form: with bn(L) ∈ Bdiag(N), it is clear that T maps C∗(V, Bdiag(N)) into A(N). V nbn(L), (2.6) Xn∈Z Proposition 2.7. For any supernatural number N the algebras B(N) and C ∗(V, Bdiag(N)) are isomorphic. Proof. For any b1, b2 ∈ C∗(V, Bdiag(N)), it can be shown just like for regular Toeplitz opera- tors, that: for some compact operator K ∈ K. Now, the map T (b1b2) = T (b1)T (b2) + K defined by: [T ] : C∗(V, Bdiag(N)) → A(N)/K [T ](b) = T (b) + K gives the required isomorphism. (cid:3) Let B(N) be the ∗-algebra generated algebraically by V, V −1 and Bdiag(N). Notice that we have: B(N) = A(N)/(A(N) ∩ K), i.e. B(N) is the image of A(N) under the quotient map ξ. Also, because of the commutation relation (2.4), the elements of B(N) are precisely the finite sums of the form given in (2.6). We have the following further identification of B(N), see [8]. UNBOUNDED DERIVATIONS IN BUNCE-DEDDENS-TOEPLITZ ALGEBRAS 7 Proposition 2.8. For infinite N the algebra B(N) can be identified with the crossed product of C(Z/NZ) with Z, acting on C(Z/NZ) via shifts. i.e., B(N) ∼= C(Z/NZ) ⋊σ Z where for f ∈ C(Z/NZ), σf (x) = f (x + 1). For finite N one can identify B(N) with C(S1) ⊗ MN (C). This is useful for the purpose of classifying derivations in A(N) and B(N) in the next section. We describe this identification in detail below. Proposition 2.9. For a finite supernatural number N there is an isomorphism: C∗(V, Bdiag(N)) ∼= C(S1) ⊗ MN (C). Proof. We first relabel the basis elements of ℓ2(Z) as follows: {EkN +j k ∈ Z, 0 ≤ j < N}. Consider the following sequence: eN (l) =(1 if N l 0 otherwise. (2.7) Then clearly we have periodicity and the following formula: eN (l + N) = eN (l), eN (L)EkN +j = δj,0EkN +j. For 0 ≤ s, r < N, we define the operators: Psr := V seN (L)V −r. It is easy to verify using the above formulas that Psr have the following properties: sr = Prs (i) P ∗ (ii) PsrPtq = δtrPsq. As a consequence, if Esr are the standard basis elements of MN (C), then the map Psr 7→ Esr induces the isomorphism C∗(Psr) ∼= MN (C). Moreover, any element of Bdiag(N) can be written as a linear combination of Prr, 0 ≤ r < N. We also have the relation: V = P10 + P21 + · · · + P(N −1)(N −2) + V N P0(N −1), which can be verified by a direct calculation on basis elements. Therefore, we obtain: C∗(V, Bdiag(N)) ∼= C∗(Psr, V N ). Consequently, because V N commutes with the operators Psr for all 0 ≤ s, r < N, we have: C∗(V, Bdiag(N)) ∼= C∗(V N ) ⊗ C∗(Psr) ∼= C(S1) ⊗ MN (C). Here C∗(V N ) is isomorphic with C(S1) because V N is equivalent to the usual bilateral shift. (cid:3) 8 KLIMEK, MCBRIDE, RATHNAYAKE, SAKAI, AND WANG 3. Covariant Derivations 3.1. Derivations. A derivation d in A(N) with domain A(N) is a linear map d : A(N) → A(N) which satisfies the Leibniz rule: d(ab) = d(a)b + ad(b) In this paper we only study derivations d with domain A(N), and for all a, b ∈ A(N). derivations δ in B(N) with domain B(N), so we will not explicitly mention domains below. A derivation d is called approximately inner if there are an ∈ A(N) such that d(a) = lim n→∞ [an, a] for a ∈ A(N). The first important observation is that any derivation in A(N) preserves compact opera- tors. Theorem 3.1. If d : A(N) → A(N) is a derivation, then d : A(N) ∩ K → K. Proof. It is enough to prove that d(P0) is compact, where P0 is the orthogonal projection onto the one-dimensional subspace spanned by E0, because A(N) ∩ K is comprised of linear combinations of expressions of the form: U rP0(U ∗)s. The result then follows immediately from the Leibniz property. To see that d(P0) is compact, simply apply d to both sides of the relation P0 = P 2 0 to obtain: d(P0) = d(P0)P0 + P0d(P0) ∈ K, which completes the proof. (cid:3) As a consequence of the above theorem, if d : A(N) → A(N) is a derivation in A(N), then [d] : B(N) → B(N) defined by gives a derivation in B(N) where, as before, B(N) = A(N) + K. [d](a + K) := da + K 3.2. Classification of covariant derivations. For each θ ∈ [0, 2π), let ρK be defined by: θ : A(N) → A(N) Then we have: ρK θ (a) = eiθKae−iθK. ρK θ (U) = eiθU, ρK θ (U ∗) = e−iθU ∗ and ρK θ is a well-defined automorphism of A(N), and ρK θ (a(K)) = a(K). θ preserves A(N). Thus, ρK (3.1) Definition 3.1. Given n ∈ Z, a derivation d in A(N) is said to be a n-covariant derivation if the relation (ρK θ )−1d(ρK θ (a)) = e−inθd(a) holds. Similarly, for θ ∈ [0, 2π), we let ρL θ be the automorphism of B(N), preserving B(N), defined by: ρL θ (b) = eiθLbe−iθL. UNBOUNDED DERIVATIONS IN BUNCE-DEDDENS-TOEPLITZ ALGEBRAS 9 Definition 3.2. Given n ∈ Z, a derivation δ in B(N) is said to be a n-covariant derivation if the relation (ρL θ )−1δ(ρL θ (b)) = e−inθδ(b) holds. An important step in classifying derivations on Bunce-Deddens-Toeplitz algebras is the classification the n-covariant derivations in A(N) since they arise as Fourier coefficients of general derivations. First we establish the following useful description of covariant subspaces in A(N). Proposition 3.2. We have the following equality: Adiag(N) = {a ∈ A(N) : ρK θ (a) = a}. Proof. Clearly if a ∈ Adiag(N), then ρK satisfies ρK θ (a) = a then the equation: θ (a) = a by formula (3.1). Conversely, if a ∈ A(N) implies that have: (Ek, eiθKae−iθKEl) = (Ek, aEl) eiθ(k−l)(Ek, aEl) = (Ek, aEl) for every θ ∈ [0, 2π) and every k, l, from which it follows that a is a diagonal operator. (cid:3) Proposition 3.3. Denote by An(N) the n-th spectral subspace of ρK θ : An(N) := {a ∈ A(N) : ρK θ (a) = einθa}. Then we have: An(N) =({U na(K) : a(K) ∈ Adiag(N)} {a(K)(U ∗)−n : a(K) ∈ Adiag(N)} if n ≥ 0 if n < 0. Proof. We will give the proof for n > 0; the proof for n < 0 works similarly. Since we have: θ (U na(K)) = einθU na(K), ρK one containment clearly follows. Conversely, if a ∈ An(N) then a(U ∗)n ∈ Adiag(N), hence is of the form a(U ∗)n = a(K) by the previous proposition. Consequently, we have: a = a(K)U n = U na(K + nI), which shows the other containment. (cid:3) It turns out that n-covariant derivations in A(N) can be described explicitly. Theorem 3.4. If d is an n-covariant derivation in A(N), then there exists a diagonal operator βn(K) such that d can be written as: d(a) =([U nβn(K), a] [βn(K)(U ∗)−n, a] if n ≥ 0 if n < 0, (3.2) where the operator βn(K) satisfies the following conditions: if N is infinite and n 6= 0 or N is finite but N ∤ n, then βn(K) ∈ Adiag(N), 10 KLIMEK, MCBRIDE, RATHNAYAKE, SAKAI, AND WANG (so in particular it is bounded); otherwise: βn(K) − βn(K − I) ∈ Adiag(N). The operator βn(K) is unique except when n = 0 where β0(K) is unique up to an additive constant. Conversely, given any βn(K) satisfying those properties, the formulas above define n-covariant derivations in A(N). Proof. Suppose n > 0 and d is a n-covariant derivation in A(N). It follows that we have d(a(K)) ∈ An(N), and hence the formula: d(a(K)) = U n d(a(K)) for some d(a(K)) ∈ Adiag(N) by Proposition 3.3. Similarly, there exists αn(K) ∈ Adiag(N) such that: d(U ∗) = −U n−1αn(K) d(U) = U n+1αn(K + I), and (3.3) (3.4) where the last equation follows from the relation d(U ∗)U + U ∗d(U) = 0. From formula (3.3) for every a(K), b(K) ∈ Adiag(N) we have the following: d(a(K)b(K)) = (U ∗)nd(a(K)b(K)) = (U ∗)nd(a(K))b(K) + (U ∗)na(K)d(b(K)) = d(a(K))b(K) + a(K + nI) d(b(K)). Since d(a(K)b(K)) = d(b(K)a(K)), it follows that: d(a(K))[b(K) − b(K + nI)] = d(b(K))[a(K) − a(K + nI)]. (3.5) For given n we can always choose a(K) such that a(k) − a(k + n) 6= 0 for every k. Using such a(K) we define: which is independent of the choice of a by the formula (3.5). It follows that: βn(K) = d(a(K))(a(K) − a(K + nI))−1, d(a(K)) = U nβn(K)[a(K) − a(K + nI)] (3.6) for any a(K) because, when a(k + n) − a(k) = 0 for some k, then both sides of the above equation are zero, as implied again by the formula (3.5). Next, applying d to the commutation relation U ∗a(K) = a(K + I)U ∗ we obtain: (βn(K) − βn(K − I) − αn(K))[a(K) − a(K + nI)] = 0. It follows that we must have: This leads to formulas: αn(K) = βn(K) − βn(K − I). d(U ∗) = −U n−1(βn(K) − βn(K − I)) d(U) = U n+1(βn(K + I) − βn(K)), and UNBOUNDED DERIVATIONS IN BUNCE-DEDDENS-TOEPLITZ ALGEBRAS 11 and so we have that d(a) = [U nβn(K), a] holds true for all the generators and hence for every a ∈ A(N) and n > 0. Notice also that we can compute βn in terms of αn by the formula: The proof for n < 0 works similarly. βn(k) = αn(i), k Xi=0 If n = 0, the formulas for d(U) and d(U ∗) are: d(U) = Uα0(K), d(U ∗) = −α0(K)U ∗. We claim that in this case we have: This is because for an invariant derivation we have: d(a(K)) = 0. d : Adiag(N) → Adiag(N), and elements of Adiag(N) are finite linear combinations of diagonal orthogonal projections. If P ∈ Adiag(N) is such a projection, by applying d to P 2 = P we obtain: (I − 2P )d(P ) = 0, which implies d(P ) = 0. To obtain d(a) = [β0(K), a], we define the operator β0(K) as the solution of the equation: and so it is determined only up to additive constant. We usually make a particular choice: α0(K) = β0(K + I) − β0(K), with β0(0) = 0. β0(k) = α0(i), k−1 Xi=0 There are additional restrictions on βn(K). For N infinite and n 6= 0, we must have d(a(K)) = βn(K)[a(K) − a(K + nI)] ∈ Adiag(N) for every a(K) ∈ Adiag(N). By choosing for example a(k) = e2πik/l, an l-periodic sequence, with l N but l ∤ n, it is clear that a(k + n) − a(k) 6= 0 for every k and so the operator a(K + n) − a(K) is invertible. It follows that βn(K) must belong to Adiag(N). If N is finite and N ∤ n then we can choose an N-periodic sequence and argue as above to show that βn(K) ∈ Adiag(N). Conversely, given any βn(K) satisfying the properties and derivation d given by the com- mutator formula (3.2), the expressions for d on generators (3.4), (3.6) imply that d is a well defined derivation in A(N). In particular, if N is finite, N n, and a(k) is any l-periodic sequence where l N is also n-periodic, then a(k) − a(k + n) = 0. Moreover, if a(k) a sequence that is eventually zero, then for large enough k we have a(k) − a(k + n) = 0. Thus d(a(K)) is eventually zero for every k, and there are no additional restrictions on βn(K) in this case. (cid:3) 12 KLIMEK, MCBRIDE, RATHNAYAKE, SAKAI, AND WANG From this theorem it is clear that if N is infinite and n 6= 0 or N is finite but N ∤ n, then the n-covariant derivation d in A(N) is inner. Otherwise such an n-covariant derivation d is in general not inner. We simultaneously state here without a detailed proof a similar classification of n-covariant derivations on Bunce-Deddens algebras. In fact, all of the arguments in the unilateral shift case of Theorem 3.4 work the same (if not simpler) in the bilateral case needed for Bunce- Deddens algebras. Theorem 3.5. If δ is an n-covariant derivation in B(N), then there exists ηn(L) such that δ(a) = [V nηn(L), a] for every a in B. If N is infinite and n 6= 0 or N is finite but N ∤ n then: otherwise we have: ηn(L) ∈ Bdiag(N), ηn(L + I) − ηn(L) ∈ Bdiag(N). Conversely, given any ηn(L) satisfying those properties, the formulas above define n-covariant derivations in B(N). 3.3. Properties of covariant derivations. In general, if an n-covariant derivation is ap- proximately inner then it can also be approximated by inner n-covariant derivations. Proposition 3.6. Suppose d is an n-covariant derivation in A(N). If d is approximately inner, then there exists a sequence of n-covariant derivations {dM } in A(N) such that, for every a ∈ A(N) we have: d(a) = lim M →∞ dM (a). Proof. Given an element a ∈ A(N), define its ρK θ n-th Fourier component by: (a)n = e−inθρK θ (a)dθ. 1 2πZ 2π 0 If d is approximately inner then there is a sequence {zM }, with zM ∈ A(N) such that: d(a) = lim M →∞ [a, zM ] for every a ∈ A(N). It can be easily checked that the Fourier component (zM )n is in An(N). So, it is sufficient to show that: d(a) = lim M →∞ [a, (zM )n] on the generators U, U ∗ and a(K) as the result then follows from Proposition 3.3. Since UzM − zM U → d(U), we equivalently have: e−inθ(zM − U ∗zM U) → e−inθU ∗d(U). So, given ǫ > 0, there is an integer m such that for every M ≥ n, we can estimate: ke−inθ(zM − U ∗zM U − U ∗d(U))k < ǫ. UNBOUNDED DERIVATIONS IN BUNCE-DEDDENS-TOEPLITZ ALGEBRAS 13 Since ρK θ (U ∗d(U)) = einθU ∗d(U), the estimate can be written as: ke−inθρK θ (zM ) − e−inθU ∗ρK θ (zM )U − U ∗d(U)k = ke−inθρK θ (zM − U ∗zM U − U ∗d(U))k < ǫ. Consequently, we have: k(zM )n − U ∗(zM )nU − U ∗d(U)k ≤ 1 2πZ 2π 0 thus proving the convergence: ≤ ke−inθρK θ (zM ) − e−inθU ∗ρK θ (zM )U − U ∗d(U)kdθ < ǫ, d(U) = lim M →∞ [U, (zM )n]. A similar proof works for d acting on U ∗ and a(K), so the result follows. (cid:3) The explicit formulas of Theorem 3.4 allow us to discuss when an n-covariant derivation is approximately inner. There are several cases to consider. When N is infinite we separately consider the case when n = 0, while when N is finite, there are differences depending on whether n is a multiple or N or not. First consider an invariant derivation d in A(N) given by d(U) = Uα0(K), d(U ∗) = −α0(K)U ∗, d(a(K)) = 0. Lemma 3.7. Suppose d is an invariant derivation in A(N) with N infinite. If α0(k) ∈ c0 then d is approximately inner. Proof. Like in [11] we define αM 0 (K) ∈ Adiag(N) by: αM 0 (k) =(α0(k) 0 if k ≤ M otherwise. Then we see that αM 0 (K) converges to α0(K) in norm as M tends to infinity because: kα0(K) − αM 0 (K)k = sup k α0(k) − αM 0 (k) = sup k>M α0(k) M →∞−−−−→ 0. The sequence βM 0 (k) defined by: k−1 is eventually constant and in particular, it is bounded. Therefore, we have that βM 0 (k) := αM 0 (j) Xj=0 dM (a) := [βM 0 (K), a] is an inner derivation. To prove that dM (a) → d(a) as M → ∞ for every a ∈ A(N), it is enough to check that on the generators U and U ∗. But this follows easily since we have: d(U) − dM (U) = U(α0(K) − αM 0 (K)), and similarly for U ∗. Thus d is an approximately inner derivation. (cid:3) The second case of invariant approximately inner derivations is described next. 14 KLIMEK, MCBRIDE, RATHNAYAKE, SAKAI, AND WANG Lemma 3.8. Suppose d is an invariant derivation in A(N) with an infinite supernatural number N. If α0(k) = f (q(k)) where f ∈ C(Z/NZ) with RZ/N Z f (x)dHx = 0 and q is the quotient map introduced in (2.1), then d is approximately inner. Proof. If f ∈ C(Z/NZ), then f (x) = lim M →∞ f M (x) uniformly for some sequence of locally constant functions f M (x) on Z/NZ. By Proposition 2.3 there is a sequence of numbers {jM }, such that jM N and for every M the sequence f M (q(k)) is jM -periodic. Moreover, by subtracting constants: if necessary, we can choose f M so that: f M (x)dHx M →∞−−−−→ 0, ZZ/N Z ZZ/N Z Now consider the sequence {αM 0 (k)} := {f M (q(k))}. A simple calculation shows that the f M (x)dHx = 0. (3.7) equation (3.7) is equivalent to the following condition: Furthermore, defining jM −1 Xi=0 αM 0 (i) = 0. we have that βM 0 is also jM -periodic because: βM 0 (k) = αM 0 (i), k−1 Xi=0 k−1 k+jM −1 k−1 βM 0 (k + jM ) = αM 0 (i) + Xi=0 Xi=k αM 0 (i) = αM 0 (i) + Xi=0 jM −1 Xi=0 αM 0 (i) = βM 0 (k). Let dM : A(N) → A(N) be the derivation defined by: dM (U) = UαM 0 (K), dM (U ∗) = −αM 0 (K)U ∗, dM (a(K)) = 0. Thus, we have: dM (a) = [a, βM 0 (K)] for every a ∈ A(N) and, since βM Moreover, the sequence {dM } approximates d because: 0 (K) ∈ Adiag(N), it follows that dM is an inner derivation. kd(U) − dM (U)k = sup k α0(k) − αM 0 (k) = sup k f (q(k)) − f M (q(k)) → 0 as M → ∞. Similarly, we obtain that: lim M →∞ and therefore, d is approximately inner. dM (U ∗) = d(U ∗), (cid:3) Now we consider the case of finite N and N n. Further examples of approximately inner n-covariant derivations are described by the following lemma. UNBOUNDED DERIVATIONS IN BUNCE-DEDDENS-TOEPLITZ ALGEBRAS 15 Lemma 3.9. Suppose d is an n-covariant derivation in A(N) where N is finite and N n. Then d is approximately inner if αn(k) ∈ c0. Proof. The proof is essentially the same as that of Lemma 3.7. (cid:3) To complete the classification of n-covariant derivations we introduce special derivations dn,K in A(N) given by: dn,K(a) =([U n(K + I), a] [(K + I)(U ∗)n, a] if n ≥ 0 if n < 0. Notice that by Theorem 3.4, derivations dn,K(a) are well-defined in A(N) when N is infinite and n = 0 or when N is finite and N n, because we have the following relation for the diagonal operator coefficients: and there are no other restrictions on the coefficients for those cases. (K + I) − K = I ∈ Adiag(N), Theorem 3.10. If d is an n-covariant derivation in A(N) where N is infinite and n = 0 or when N is finite and N n, then there exists a unique constant Cn such that for every a ∈ A(N), where d is an approximately inner derivation. d(a) = Cndn,K(a) + d(a) Proof. Consider the case n > 0 and finite N. We then have the formula: d(a) = [U nβn(K), a], and the condition: αn(K) = βn(K + I) − βn(K) ∈ Adiag(N). We apply Proposition 2.4 to αn(k), and refine it in the following way: where Cn is a constant, αn,0(k) ∈ c0, αn,per(k + N) = αn,per(k) and αn(k) = αn,0(k) + Cn + αn,per(k) N −1 We then decompose βn using the following: αn,per(k) = 0. Xk=0 k k βn,0(k) := αn,0(j), βn,per(k) := Xj=0 αn,per(j). Xj=0 It is easy to verify that βn,per(k) is N-periodic, just as in Lemma 3.8. We then obtain: βn(k) = βn,0(k) + Cn(k + 1) + βn,per(k). So, for n > 0, the derivation d decomposes as follows: d(a) = [U nβn,0(K), a] + Cndn,K(a) + [U nβn,per(K), a]. We know that [U nβn,0(K), a] is approximately inner by Lemma 3.9. Moreover, since βn,per(K) ∈ A(N), [U nβn,per(K), a] is an inner derivation. To conclude the theorem for n > 0, and verify the uniqueness, it only remains to show that dn,K(a) is not approximately inner. This easily follows from the methods of Theorem 4.4 in [11], in the following way. 16 KLIMEK, MCBRIDE, RATHNAYAKE, SAKAI, AND WANG Assume to the contrary that dn,K is approximately inner. By Proposition 3.6 there exists a sequence µM (K) ∈ Adiag(N), M = 1, 2, . . ., such that: for all a ∈ A. In particular, we must have: dn,K(a) = lim M →∞ [U nµM (K), a] dn,K(U) = U n+1 = lim M →∞ U n+1(µM (K + I) − µM (K)). Without loss of generality assume µM (k) are real, or else in the argument below simply consider the real part of µM (k). The above equation implies that: lim M →∞ sup (µM (k + 1) − µM (k)) − 1) = 0. k Therefore for any small ε > 0 there are k and m large enough so that we have: 1 − ε ≤ µM (k + 1) − µM (k) ≤ 1 + ε. By telescoping µM (k), we get: µM (k) = (µM (k) − µM (k − 1)) + · · · + (µM (k0 + 1) − µM (k0)) + µM (k0) for some fixed k0. Together the last two formulas imply that: µM (k) ≥ (1 − ε)(k − k0) + µM (k0), which goes to infinity as k goes to infinity. This contradicts the fact that µM (K) ∈ Adiag(N) which completes the proof for n > 0 and finite N. Cases n = 0 and n < 0 can be proved very similarly. (cid:3) We summarize the remaining cases of our classification of n-covariant derivations in A(N) in the next theorem. Theorem 3.11. Suppose d is an n-covariant derivation in A(N). If N is infinite with n 6= 0 or if N is finite with N ∤ n, then d is an inner derivation. Proof. From Theorem 3.4 we already know that βn(K) ∈ Adiag(N) when N is infinite with n 6= 0 or when N is finite with N ∤ n. Thus d is an inner derivation. (cid:3) This concludes the classification of n-covariant derivations in A(N). Classification of n- covariant derivations in B(N) is somewhat simpler and can be obtained by applying the same methods as used in the classification of n-covariant derivations in A(N). Theorem 3.12. If δ is a n-covariant derivation in B(N) where N is infinite and n 6= 0 or N is finite but N ∤ n then δ is an inner derivation. Otherwise there exists a unique constant Cn such that δ(a) = Cn[V nL, a] + δ(a) for every a ∈ B(N), where if N is finite and N n, then δ is an inner derivation, and if N is infinite and n = 0 then δ is an approximately inner derivation. UNBOUNDED DERIVATIONS IN BUNCE-DEDDENS-TOEPLITZ ALGEBRAS 17 4. General unbounded derivations This chapter contains our main results: the classification of derivations in A(N) and B(N). The structure of such derivations differs depending on whether N is finite or infinite, and is interesting even for the simplest case of N = 1, when A(1) is the Toeplitz algebra. The main technique is the use of Fourier series with respect to the S1 action ρK θ on A(N), and ρL θ on B(N). The Fourier coefficients of derivations are defined in the following way. Definition 4.1. If d is a derivation in A(N), the n-th Fourier component of d is defined as: Definition 4.2. If δ is a derivation in B(N), the n-th Fourier component of δ is defined as: dn(a) = einθ(ρK θ )−1dρK θ (a) dθ. δn(b) = einθ(ρL θ )−1δρL θ (b) dθ. 0 1 2πZ 2π 2πZ 2π 1 0 We have the following simple observation. Proposition 4.1. If d is a derivation in A(N), then dn is an n-covariant derivation and well-defined on A(N). Proof. It is straightforward to see that dn is a derivation and is well-defined on A(N). The following computation verifies that dn is n-covariant: (ρK θ )−1dnρK θ (a) = = 0 1 2π Z 2π 2π Z 2π 1 0 einφ(ρK θ )−1ρ−1 φ dρφρK θ (a) dφ einφρ−1 θ+φdρθ+φ(a) dφ. Changing to new variable θ + φ, and using the translation invariance of the measure, it now follows that (ρK (cid:3) θ (a) = e−inθdn(a). θ )−1dnρK We have the following key Ces`aro mean convergence result for Fourier components of d, which is more generally valid for unbounded derivations in any Banach algebra with the continuous circle action preserving the domain of the derivation. Lemma 4.2. If d is a derivation in A(N) then: d(a) = lim M →∞ for every a ∈ A(N). Proof. We need to show that: 1 M + 1 M Xj=0 j Xn=−j dn(a)! , (4.1) dn(a) − d(a)! M →∞−−−−→ 0 M 1 M + 1 Xj=0 j Xn=−j dn(a) − d(a)! = for all a ∈ A(N). Using the standard Fourier analysis [10] we can write: 1 M + 1 M Xj=0 j Xn=−j 1 2πZ 2π 0 θ )−1dnρK FM (θ)(cid:0)(ρK θ (a) − d(a)(cid:1) dθ, is the Fej´er kernel, which is manifestly positive and satisfies: FM (θ) = 1 M + 1 sin(cid:0) M +1 2 (cid:1) θ sin(cid:0) θ 2(cid:1) !2 1 2π Z 2π 0 FM (θ)dθ = 1. 18 where: KLIMEK, MCBRIDE, RATHNAYAKE, SAKAI, AND WANG Since (ρK θ )−1dnρK we have estimates: θ (a) − d(a) is continuous in θ, given ǫ > 0 we can find small ω > 0 so that 0 1 2π Z ω 2π Z 2π 2π−ω 1 FM (θ)k(ρK θ )−1dnρK θ (a) − d(a)kdθ ≤ FM (θ)k(ρK θ )−1dnρK θ (a) − d(a)kdθ ≤ and ǫ 3 ǫ 3 . Moreover, on the remaining interval we can estimate as follows: 1 2π Z 2π−ω ω FM (θ)k(ρK θ )−1dnρK θ (a) − d(a)kdθ ≤ const (M + 1) sin2(ω/2) for some constant in the numerator. Consequently, we can choose M large enough so that we get: which completes the proof of (4.1). 1 M + 1 M Xj=0 j Xn=−j (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) dn(a) − d(a)!(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≤ ǫ, (cid:3) The first case we consider is a description of derivations for infinite N. Theorem 4.3. Suppose d is a derivation in A(N) with N infinite. Then there exists a unique constant C such that d(a) = C[K, a] + d(a) where d is approximately inner. Proof. Let d0 be the 0-th Fourier component of d. It is an invariant derivation, so by Theorem 3.10 we have the unique decomposition: d0(a) = Cd0,K(a) + d0(a) = C[K, a] + d0(a), for every a ∈ A(N), where d0 is an approximately inner derivation. From Theorem 3.11 we have that the Fourier components dn, n 6= 0 are inner derivations. It follows from (4.1), by extracting d0, that we have: d(a) = d0(a) + lim M →∞ 1 M + 1 M Xj=1   Xn≤j, n6=0 dn(a)  . The terms under the limit sign are all finite linear combinations of n-covariant derivations and so they are inner derivations themselves, meaning that the limit is approximately inner, which ends the proof. (cid:3) UNBOUNDED DERIVATIONS IN BUNCE-DEDDENS-TOEPLITZ ALGEBRAS 19 In exactly the same way we obtain the corresponding classification result for unbounded derivations in Bunce-Deddens algebras for infinite N. Theorem 4.4. Suppose δ is a derivation in B(N) with N infinite. Then there exists a unique constant C such that δ(b) = C[L, b] + δ(b) where δ is approximately inner. We now turn to the classification of derivations in Bunce-Deddens and Bunce-Deddens- Toeplitz algebras for finite N. We start with the following simple observation. Lemma 4.5. If δ : B(N) → B(N) is a derivation, then δ(V N ) ∈ C∗(V N ). Proof. Applying δ to the relation V N Psr = PsrV N , we see that δ(V N ) commutes with Psr for every r, s and so it must be in C∗(V N ). (cid:3) By Propositions 2.7 and 2.9, we know that for finite N we have an isomorphism of C∗- algebras: B(N) ∼= C(S1) ⊗ MN (C) and B(N) can be identified with the set of N by N matrix-valued trigonometric polynomials F (t) on S1. For any f ∈ C(S1) we define the following special derivation δf in B(N): δf (F (t)) = f (t) 1 i d dt F (t). Derivations δf are used in the following theorem which gives very concrete and explicit classification of derivations in B(N). Theorem 4.6. Suppose N is finite and δ is a derivation in B(N). Then there exists a unique f ∈ C(S1) such that δ = δf + δ where δ is inner. Proof. If p(t) is a trigonometric polynomial and A ∈ MN (C), then we have: δ(p(t)A) = δ(p(t))A + p(t)δ(A). Here C(S1) ∼= C∗(V N ) and by Lemma 4.5, there is f ∈ C(S1) such that δ(eit) = f (t)eit and hence we have: δ(p(t)) = f (t) p(t). 1 i d dt Moreover, given any derivation δ : MN (C) → C(S1, MN (C)), the following continuous matrix-valued function H(t) ∈ C(S1, MN (C)) given by: H(t) = satisfies the easily verifiable relation: 1 N N Xr,s=1 δ(Prs)(t)Prs δ(A)(t) = [H(t), A]. Consequently, we have: δ(p(t)) =(cid:18)f (t) 1 i d dt p(t)(cid:19) A + p(t)[H(t), A] = δf (p(t)A) + [H(t), p(t)A], 20 KLIMEK, MCBRIDE, RATHNAYAKE, SAKAI, AND WANG which completes the proof. (cid:3) It remains to classify derivations in A(N) for finite N. For any f ∈ C(S1) we define a special derivation df in A(N) to be the unique derivation such that: 1 N UT (f (V N )), df (U ∗) = − df (aper(K)) = 0, df (U) = 1 N T (f (V N ))U ∗, (4.2) where aper(K) is any element of Adiag, per(N). Here T (f (V N )) is the Toeplitz operator of formula (2.5), where f (V N ) is an operator in ℓ2(Z) defined by the functional calculus. The derivation df is given on generators of A(N), hence, if it exists it is unique; to see that it is unambiguously defined on all of A(N) we need an additional argument. Let f (t) =Pn∈Z fneint be a trigonometric polynomial which we decompose as: where f +(t) =Pn≥0 fneint and f −(t) has a similar formula, then we claim that we have the following formula for df : f (t) = f +(t) + f −(t), To verify (4.2) we calculate using Lemma 2.6: N (cid:2)T (f +(V N ))(K + I) + (K + I)T (f −(V N )), a(cid:3) . 1 df (a) = (4.3) Ndf (aper(K)) = T(cid:0)(cid:2)f +(V N ), aper(L)(cid:3)(cid:1) (K + I) + (K + I)T(cid:0)(cid:2)f −(V N ), aper(L)(cid:3)(cid:1) . Since aper(L) is N periodic, it commutes with V N , thus the above commutators are zero and hence df (aper(K) = 0. Next, notice that UT (f +(V N )) = T (f +(V N ))U since f +(V N ) only contains nonnegative powers of V . Using this fact, the commutation relation (2.3), and Lemma 2.6 we have Ndf (U) = T (f +(V N )) [(K + I)U − U(K + I)] + [(K + I) − KUU ∗] T (f −(V N )V ) =(cid:0)T (f +(V N )) + T (f −(V N ))(cid:1) U = T (f (V N ))U. For similar reasons as above we have U ∗T (f −(V N )) = T (f −(V N ))U ∗. Using this, the commutation relation (2.3), and again Lemma 2.6, we obtain the last part of formula (4.2). Thus this completes the proof of existence of df for polynomial f . It is clear from those formulas that df is a well-defined derivation A(N) → A(N). For a general f ∈ C(S1) we use an approximation argument to construct df . Namely if {f M } is a sequence of trigonometric polynomials converging uniformly to f then, by formulas (4.2), the sequence of derivations {df M } converges on generators of A(N), and hence it converges for every a ∈ A(N). The limit, which must be a derivation in A(N), gives a construction of df . Derivations df are used in the theorem below. Compared to the proof of Theorem 4.3, the classification of derivations in A(N) for finite N gets more complicated since in this case a derivation may have infinitely many non-inner Fourier components. To handle those difficulties we need the following lemma which is more generally valid for unbounded derivations in any algebra if the domain is finitely generated. Lemma 4.7. If N is finite, d is a derivation in A(N), and there is a sequence {dM } of approximately inner derivations such that for every a ∈ A(N): d(a) = lim M →∞ dM (a), UNBOUNDED DERIVATIONS IN BUNCE-DEDDENS-TOEPLITZ ALGEBRAS 21 then d is also approximately inner. Proof. For finite N the algebra A(N) is finitely generated; for example we can choose the following set of generators: G := {U, U ∗, eN (K)}, where the sequence eN (k) was defined in (2.7). Also, dM are approximately inner which means that there is a sequence {zM,W } of elements of A(N) such that for every a ∈ A(N): For every positive integer j we can choose Mj such that for every a ∈ G we have: dM (a) = lim W →∞ [zM,W , a]. which can be done because the generating set G is finite. Then choose Wj such that for every a ∈ G we have: 1 2j , 1 2j . 1 j , (cid:13)(cid:13)d(a) − dMj (a)(cid:13)(cid:13) ≤ (cid:13)(cid:13)dMj (a) − [zMj ,Wj , a](cid:13)(cid:13) ≤ (cid:13)(cid:13)d(a) − [zMj ,Wj, a](cid:13)(cid:13) ≤ d(a) = lim j→∞ [zMj ,Wj , a] By the triangle inequality we obtain: which means that we have: for every a ∈ G, which by the Leibniz identity implies the above convergence for every a ∈ A(N). Consequently, d is approximately inner, finishing the proof. (cid:3) With this preparation we are now ready to state our classification result for derivations in A(N) with finite N. Theorem 4.8. Suppose N is finite and d is a derivation in A(N). Then there exists unique function f ∈ C(S1) such that: d = df + d, where d is approximately inner and df is defined by formula (4.2). Proof. Consider the derivation [d] : B(N) → B(N) given by: [d](a + K) = d(a) + K. It is easy to see from Definitions 4.1 and 4.2 that we have the following equality for the Fourier components: [d]n = [dn]. To construct the function f in the statement of the theorem we notice that Lemma 4.5 states that [d](V N ) ∈ C∗(V N ) and hence we can write: [d](V N ) = f (V N )V N for some f ∈ C(S1). It follows that we have: [d]n(V N ) =(fjV jN +N 0 if n = jN otherwise, (4.4) 22 KLIMEK, MCBRIDE, RATHNAYAKE, SAKAI, AND WANG where fj are the Fourier coefficients of f . On the other hand, from Theorem 3.4 we know that dn(U N ) = [U nβn(K), U N ] = U n+N (αn(K + (N − 1)I) + · · · + αn(K)) . Next, for N b, we decompose αn(k) as in the proof of Theorem 3.10: where Cn is a constant, αn,0(k) ∈ c0, αn,per(k + N) = αn,per(k) and αn(k) = αn,0(k) + Cn + αn,per(k), N −1 αn,per(k) = 0. Xk=0 This decomposition is also valid for N ∤ n but with Cn = 0 by Theorem 3.11. It follows that we have: dn(U N ) = U n+N (αn,0(K + (N − 1)I) + · · · + αn,0(K) + NCn) , and consequently, we obtain: [dn](V N ) = NCnV n+N =(NCjN V jN +N 0 if n = jN otherwise. (4.5) Comparing equation (4.4) and (4.4) implies the following formulas for constants Cn: Cn =( 1 N fj 0 if n = jN otherwise. It then follows from the formula (4.3) that we have: (df )n(a) = 1 N fjdn,K(a) =([U nCn(K + I), a] [Cn(K + I)(U ∗)−n, a] if n ≥ 0 if n < 0. (4.6) As in the proof of Theorem 3.10 we decompose βn using: k k βn,0(k) := αn,0(j), βn,per(k) := αn,per(j). Xj=0 Xj=0 This gives the following formulas for the Fourier components of the difference between d and df : dn(a) := (d − df )n(a) =([U n (βn,0(K) + βn,per(K)) , a] [(βn,0(K) + βn,per(K)) (U ∗)−n, a] if n ≥ 0 if n < 0. From Theorem 3.11 we know that dn is an inner derivation if N ∤ n. If we denote by ( d0)n and ( dper)n the following derivations on A(N): [βn,0(K)(U ∗)−n, a] ( d0)n(a) =([U nβn,0(K), a] ( dper)n(a) =([U nβn,per(K), a] [βn,per(K)(U ∗)−n, a] if n ≥ 0 if n < 0, if n ≥ 0 if n < 0 UNBOUNDED DERIVATIONS IN BUNCE-DEDDENS-TOEPLITZ ALGEBRAS 23 then, when N n, we know that ( dper)n is inner while ( d0)n is approximately inner from Theorem 3.10. To conclude that d is approximately inner we first use formula (4.1) which says, in view of the above discussion, that d is a limit of approximately inner derivations. Consequently, using Lemma 4.7, we see that d is approximately inner. To show the uniqueness of this decomposition, it is sufficient to prove that df is approx- imately inner if and only if f = 0. If f = 0, it is clear that df is approximately inner. To prove the converse statement, notice that if df is approximately inner then so are the Fourier components (df )n, which by formula (4.6) are proportional to derivations dn,K, which in turn were proved in Theorem 3.10 not to be approximately inner. This gives a contradiction and finishes the proof of the theorem. (cid:3) 5. Implementations The purpose of this section is to investigate implementations of unbounded derivations in Bunce-Deddens algebras B(N) as operators in Hilbert spaces. This study is inspired by the following noncommutative geometry concept of first order elliptic operator with respect to a C∗-algebra: an unbounded operator D acting in a Hilbert space H which carries a representation π of a C∗-algebra A is called a first order elliptic operator with respect to A if it satisfies two properties: (1) [D, π(a)] is bounded for all a in some dense ∗-subalgebra A of A. (2) D has a compact parametrices, which by the appendix of [11] is equivalent to the two operators (I + D∗D)−1/2 and (I + DD∗)−1/2 being compact operators. Such first a order elliptic operator with respect to A is a key component of the notion of a spectral triple in noncommutative geometry; see [5]. If a first order elliptic operator D is an implementation of a densely-defined unbounded derivation in A then the first condition of the above definition is automatically satisfied. Hence we are mainly interested in establishing when implementations of derivations in B(N) have compact parametrices. We only consider here the representations of B(N) in Hilbert spaces obtained from the GNS construction as those are the most geometrical representations of those algebras. In [11] and [12] implementations of 0-covariant, that is invariant, and 1-covariant deriva- tions in the quantum disk (Toeplitz algebra) and the quantum annulus were studied to see if it was possible to construct spectral triples on those quantum domains. Here we continue this analysis for n-covariant derivations in B(N). A state τ : B(N) → C is called a ρL θ -invariant state on B(N) if for all a ∈ B(N) it satisfies the following: τ (ρL θ (a)) = τ (a). It is not difficult to describe the ρL θ -invariant states on B(N). To do this we use the iden- tification B(N) ∼= C(Z/NZ) ⋊σ Z; see Proposition 2.8. There is a natural expectation E : B(N) → C(Z/NZ), a positive, linear map such that E2 = E. For an element V nbn(x) ∈ B(N), b =Xn∈Z 24 KLIMEK, MCBRIDE, RATHNAYAKE, SAKAI, AND WANG see (2.6), it is given by: E(b) = 1 2πZ 2π 0 ρL θ (b) dθ = b0(x) ∈ C(Z/NZ). Since C(Z/NZ) is the fixed point algebra for ρL servation: t : C(Z/NZ) → C such that: if τ : B(N) → C is a ρL θ , we immediately obtain the following ob- θ −invariant state on B(N) then there exists a state Conversely given a state t : C(Z/NZ) → C, then τ (b) = t(E(b)) defines a ρθ−invariant state on B(N). Therefore the invariant states are given by probabilistic measures on Z/NZ. τ (b) = t(E(b)). We will concentrate below on the following two most interesting and natural ρL θ -invariant states on B(N), namely τ0 and τHaar defined by: τ0(b) = E(b)(0) and τHaar(b) =ZZ/N Z E(b)(x) dHx, where dHx is the unique normalized Haar measure. Denote by H0 and HHaar the GNS Hilbert spaces corresponding to τ0 and τHaar respectively and let π0, πHaar be the corresponding representations. We have the following concrete description of those Hilbert spaces and representations. Proposition 5.1. The GNS Hilbert spaces H0 and HHaar are naturally isomorphic to the following: H0 ∼= ℓ2(Z) and HHaar ∼= L2(Z × Z/NZ). The representation π0 : B(N) → B(ℓ2(Z)) is the defining representation of B(N), i.e. π0(a) = a for all a ∈ B(N). The representation πHaar : B(N) → B(L2(Z × Z/NZ)) is completely described by: 1. πHaar(V )f (m, x) = f (m − 1, x) 2. πHaar(a(q(L)))f (m, x) = a(x + m)f (m, x), where f (m, x) ∈ L2(Z × Z/NZ) and a(x) ∈ C(Z/NZ). Proof. To properly identify the Hilbert space H0 = B(N)/{b ∈ B(N) : τ0(b∗b) = 0} we must study τ0(b∗b) = 0 for b ∈ B(N). In fact, due to the continuity of τ0, we only need to work on the dense subalgebra B(N). For any b ∈ B(N) given by equation (2.6), a straightforward calculation yields: Therefore, if τ0(b∗b) = 0, it follows that bn(0) = 0 for all n. Then the formula: τ0(b∗b) =Xn∈Z bn(0)2. H0 ∋ [b] 7→ {bn(0)} ∈ ℓ2(Z) ∼= ℓ2(Z), similar to the proof of Proposition 5.4 in [11]. Notice that gives an isomorphism H0 the class [I] in the completion of the quotient B(N)/{b ∈ B(N) : τ0(b∗b) = 0} corresponds UNBOUNDED DERIVATIONS IN BUNCE-DEDDENS-TOEPLITZ ALGEBRAS 25 to the basis element E0 in ℓ2(Z). From the formula: it follows that we have: An analogous calculation shows: V nbn−1(L), V b =Xn∈Z π0(V )[b] = {bn−1(0)}n∈Z. π0(a(L))[b] = {a(n)bn(0)}n∈Z for a(L) ∈ Bdiag(N). This proves the first part of the proposition. In the second example we have τHaar(b∗b) = 0 if and only if b = 0. If b ∈ B is given by: then the corresponding function in L2(Z × Z/NZ) is given by: b =Xn∈Z V nbn(L) =Xn∈Z V nfn(q(L)) [b](m, x) = fm(x). Otherwise calculations with τHaar are very similar. (cid:3) We remark here briefly that because B(N) is defined as the quotient of A(N) with the ideal of compact operators, an invariant state on B(N) lifts to an invariant state on A(N). The corresponding GNS Hilbert spaces are the same as for B(N), with compact operators represented trivially. In general, for a GNS Hilbert space Hτ of B(N) with respect to a state τ we have that B(N) ⊆ Hτ is dense in Hτ and [I] ∈ Hτ is cyclic. Consequently, the subspace is dense in Hτ . Define Vτ,θ : Hτ → Hτ via the equation: Dτ := πτ (B(N)) · [I] Vτ,θ[b] = [ρL θ (b)]. Notice that for every θ, the operator Vτ,θ extends to a unitary operator in Hτ . Moreover by a direct calculation we get: Vτ,θπτ (b)V −1 meaning that Vτ,θ is an implementation of ρL the following inclusions: τ,θ = πτ (ρL θ (b)), θ . It follows from the definitions that we have Vτ,θ(Dτ ) ⊆ Dτ and πτ (B(N))(Dτ ) ⊆ Dτ . Let δ be an n-covariant derivation in B(N) and let τ be a ρL θ −invariant state. Implemen- tations of δ in the GNS Hilbert space Hτ are defined in the following way. Definition 5.1. An operator Dτ : Dτ → Hτ is called a covariant implementation of an n-covariant derivation δ if and [Dτ , πτ (b)] = πτ (δ(b)) Vτ,θDτ V −1 τ,θ = einθDτ . 26 KLIMEK, MCBRIDE, RATHNAYAKE, SAKAI, AND WANG Below we find all covariant implementations of n-covariant derivations on the two GNS Hilbert spaces H0 and HHaar of Proposition 5.1, and establish when they have compact parametrices. We start by recapping Theorem 3.5 with additional details needed for the formulation of the implementation results. Any n-covariant derivation δ in B(N) is of the form: δ(b) = [V nηn(L), b], where for N infinite, n 6= 0 and N finite and N ∤ n the operator ηn(L) is in Bdiag(N); hence it comes from a function hn in C(Z/NZ), so that we have: In other cases the increment ηn(L) = hn(q(L)). γn(L) := ηn(L) − ηn(L − I) is in Bdiag(N), so it can be written as: γn(L) = gn(q(L)) for some gn(x) ∈ C(Z/NZ). decompositions: It follows that there is a constant Cn such that we have gn(x) = Cn + gn(x) and ηn(L) = CnL + ηn(L), where the function gn(x) ∈ C(Z/NZ) satisfies the property: and we have: ZZ/N Z gn(x) dHx = 0, gn(q(L)) = ηn(L) − ηn(L − I). When N is infinite and n = 0, in general, it is possible for η0(l) to be unbounded. However, when N is finite and N n then ηn(l) must be in the finite dimensional vector space C(Z/NZ), and so we have: ηn(L) = hn(q(L)) for some hn(x) ∈ C(Z/NZ). All of this notation is used in the following implementation statements. Theorem 5.2. Any covariant implementation Dτ0 : Dτ0 → ℓ2(Z) of an n-covariant deriva- tion δ in B(N) is of the form: Dτ0 =(cid:26) V nηn(L) η0(L) + c · I for n 6= 0 for n = 0, with arbitrary constant c for n = 0. If N is infinite and n 6= 0 or if N is finite and N ∤ n, then the operator Dτ0 is bounded, so it does not have compact parametrices. In all other cases, ηn(l) → ∞ as l → ∞ is a necessary and sufficient condition for Dτ0 to have compact parametrices. UNBOUNDED DERIVATIONS IN BUNCE-DEDDENS-TOEPLITZ ALGEBRAS 27 Proof. It is easy to see that Dτ0 coincides with c00 ⊆ ℓ2(Z). The formulas for Dτ0 follow from simple calculations, just like in [11]. See also the next theorem for more details of similar calculations in the Haar measure state case. From the appendix of [11], Dτ0 has compact parametrices if and only if (I + D∗ τ0Dτ0)−1/2 τ0)−1/2 are compact operators. A direct calculation yields the following formula: and (I +Dτ0D∗ I + D∗ τ0Dτ0 =( I + ηn2(L) (1 + c2) · I + 2Re η0(L) + η02(L) for n 6= 0 for n = 0, τ0Dτ0)−1/2 is compact which is a diagonal operator for all n. Therefore, it follows that (I + D∗ if and only if ηn(l) goes to infinity, in particular when Cn 6= 0. An analogous computation works for (I + Dτ0D∗ (cid:3) τ0)−1/2, thus completing the proof. Similar analysis can also be performed for implementations of n-covariant derivations in the GNS Hilbert space corresponding to the invariant state on B(N) determined by the Haar measure on Z/NZ. Theorem 5.3. There exists a function ψ(x) ∈ L2(Z/NZ, dHx) such that any implementation DτHaar : DτHaar → L2(Z × Z/NZ) of δ is of the form: (DτHaarf ) (m, x) = hn(x + m − n)f (m − n, x) + (ψ(x) − hn(x))f (m − n, x + n), if N is infinite, n 6= 0, or if N is finite, N ∤ n, and (DτHaarf ) (m, x) = (C0m + (g0(x + m − 1) + · · · + g0(x)) + ψ(x)) f (m, x), if N is infinite, n = 0, or (DτHaarf ) (m, x) =(cid:16)Cn · (m − n) + hn(x + m) − hn(x) + ψ(x)(cid:17) f (m − n, x), if N is finite, N n. For N finite and N n a necessary and sufficient condition for DτHaar to have compact parametrices is Cn 6= 0. In all other cases DτHaar does not have compact parametrices. Proof. First notice that [I] = χ0(m, x) where χ0(m, x) = 1 when m = 0 and zero for all other values of m. Given b ∈ B(N) we compute as follows: DτHaar[b] = DτHaarπHaar(b)[I] = [DτHaar, πHaar(b)][I] + πHaar(b)DτHaar [I] = πHaar(δ(b))[I] + πHaar(b)DτHaar[I]. Applying the covariance condition VτHaar,θDτHaarV −1 that there exists a function ψ(x) ∈ L2(Z/NZ, dH x) such that τHaar,θ = einθDτHaar to [I] = χ0(m, x) shows DτHaarχ0(m, x) = ψ(x)(πHaar(V n)χ0)(m, x). It follows that we have the formula: πHaar(b)DτHaar[I](m, x) = ψ(x)(πHaar(bV n)χ0)(m, x) = ψ(x)[b](m − n, x + n), because of the following calculation with Fourier components of b: bV n = Xm∈Z V m+nfm(q(L) + n · I) = Xm∈Z V mfm−n(q(L) + n · I). 28 KLIMEK, MCBRIDE, RATHNAYAKE, SAKAI, AND WANG This implies the following general expression of the operator DτHaar: (DτHaar[b])(m, x) ="Xm∈Z V mηn(L + (m − n) · I)bm−n(L) − ηn(L)bm−n(L + n · I)# (m, x) + ψ(x)[b](m − n, x + n). If N is infinite and n 6= 0 or if N is finite and N ∤ n then ηn(L) is in Bdiag(N) hence it comes from a function hn(x) in C(Z/NZ). Consequently, we have the formula: (DτHaar[b]) (m, x) = hn(x + m − n)[b](m, x) + (ψ(x) − hn(x))[b](m, x + n). The first and last terms of the above expression (those containing hn(x)) are bounded oper- ators and hence DτHaar has compact parametrices if and only if the middle term has compact parametrices by the results in the appendix of [11]. That term is unitarily equivalent to the operator: f (m, x) 7→ ψ(x)f (m, x), which for every m is the multiplication operator by an L2-function in L2(Z/NZ, dHx) and therefore DτHaar can not have compact parametrices. In the case when N is infinite and n = 0, there is in general no function h0(x) such that η0(L) = h0(q(L)), and so we write the difference η0(L + m · I) − η0(L) in terms of ηn(L) − ηn(L − I) = γn(L) = g0(q(L)) = Cn · I + g0(q(L)) to obtain the following expression: (DτHaar[b]) (m, x) = (C0m + (g0(x + m − 1) + · · · + g0(x)) + ψ(x)) [b](m, x). As in the first case, since for each fixed m the above formula is a diagonal operator that is a multiplication by a L2-function, it is therefore impossible for DτHaar to have compact parametrices. Finally in the last case, when N is finite and N n, the Hilbert space L2(Z/NZ, dHx) is now a finite dimensional Hilbert space. Hence, we can decompose ηn(L) as follows: Using N-periodicity in x we arrive at the expression: ηn(L) = CnL + gn(q(L)). (DτHaar[b]) (m, x) =(cid:16)Cn · (m − n) + hn(x + m) − hn(x) + ψ(x)(cid:17) [b](m − n, x). Notice that since hn(x) is N-periodic then hn(x + m) is uniformly bounded in m and x. It now follows that DτHaar have compact parametrices if and only if Cn 6= 0. This completes the proof. (cid:3) References [1] Bratteli, O., Derivations, Dissipations and Group Actions on C∗ -algebras, Lecture Notes in Math. 1229, Springer, 1986. [2] Bratteli, O., Elliott, G. A., and Jorgensen, P. E. T., Decomposition of unbounded derivations into invariant and approximately inner parts. Jour. Reine Ang. Math., 346, 166 - 193, 1984. [3] Bunce, J.W., and Deddens, J. A., C∗ -algebras generated by weighted shifts, Indiana Univ. Math. J., 23, 257 - 271, 1973. [4] Bunce, J.W., and Deddens, J. A., A family of simple C∗ -algebras related to weighted shift operators, J. Funct. Analysis, 19, 13 - 24, 1975. UNBOUNDED DERIVATIONS IN BUNCE-DEDDENS-TOEPLITZ ALGEBRAS 29 [5] Carey, A. L., Phillips, J., Rennie, A. C., Spectral triples: examples and index theory in Noncommutative Geometry and Physics: Renormalisation, Motives, Index Theory, pp 175-265, European Mathematical Society, 2011. [6] Connes, A., Noncommutative Geometry, Academic Press, 1994. [7] Davidson, K.R., C∗-algebras by Example. American Mathematical Soc., 1996. [8] Exel, R., The Bunce-Deddens Algebras as Crossed Products by Partial Automorphisms, Bol. Soc. Bras. Mat., 25, 173 - 179, 1994. [9] Jorgensen, P., Approximately inner derivations, decompositions and vector fields of simple C∗-algebras, in Mappings of operator algebras: Proceedings of the Japan-U.S. Joint Seminar (Philadelphia, 1988), pp. 15 - 113, (H. Araki and R.V. Kadison, eds.), Progr. Math., vol. 84, Birkhauser, Boston, 1990. [10] Katznelson, Y., An Introduction to Harmonic Analysis, Cambridge University Press, 2004. [11] Klimek, S., McBride, M., Rathnayake, S., Sakai, K., Wang, H., Derivations and Spectral Triples on Quantum Domains I: Quantum Disk, SIMGA, 013, 1 - 26, 2017. [12] Klimek, S., McBride, M., Rathnayake, S., Derivations and Spectral Triples on Quantum Domains II: Quantum Annulus, to appear in Sci. Chi. Math., arXiv:1710.06257. [13] Pedersen, G. K., Lifting Derivations from Quotients of Separable C ∗-algebras, Proc. Natl. Acad. Sci., 73, 1414 - 1415, 1976 [14] Ramakrishnan, D., Valenza, R., Fourier Analysis on Number Fields, Springer 1999. [15] Robert, A., A Course in p-adic Analysis, Springer, 2000. [16] Sakai, S. Operator Algebras in Dynamical Systems, Cambridge University Press, 1991. [17] Stacey, P. J., Crossed products of C ∗-algebras by ∗-endomorphisms, J. Austral. Math. Soc. Ser. A, 54, 204 - 212, 1993 [18] Willard, S., General Topology, Addison-Wesley Publishing, 1970 Department of Mathematical Sciences, Indiana University-Purdue University Indianapo- lis, 402 N. Blackford St., Indianapolis, IN 46202, U.S.A. E-mail address: [email protected] Department of Mathematics and Statistics, Mississippi State University, 175 President's Cir., Mississippi State, MS 39762, U.S.A. E-mail address: [email protected] Department of Mathematics, University of Michigan, 530 Church St., Ann arbor, MI 48109, U.S.A. E-mail address: [email protected] Department of Mathematical Sciences, ndiana University-Purdue University Indianapo- lis, 402 N. Blackford St., Indianapolis, IN 46202, U.S.A. E-mail address: [email protected] Department of Mathematical Sciences, Indiana University-Purdue University Indianapo- lis, 402 N. Blackford St., Indianapolis, IN 46202, U.S.A. E-mail address: [email protected]
1001.1012
1
1001
2010-01-07T01:19:56
Infinite Tensor Products of C_0(R): Towards a Group Algebra for R^\infty
[ "math.OA", "math.FA" ]
The construction of an infinite tensor product of the C*-algebra C_0(R) is not obvious, because it is nonunital, and it has no nonzero projection. Based on a choice of an approximate identity, we construct here an infinite tensor product of C_0(R), denoted L_V. We use this to construct (partial) group algebras for the full continuous unitary representation theory of the group R^(N) = the infinite sequences with real entries, of which only finitely many entries are nonzero. We obtain an interpretation of the Bochner-Minlos theorem in R^(N) as the pure state space decomposition of the partial group algebras which generate L_V. We analyze the representation theory of L_V, and show that there is a bijection between a natural set of representations of L_V and the continuous unitary representations of R^(N), but that there is an extra part which essentially consists of the representation theory of a multiplicative semigroup which depends on the initial choice of approximate identity.
math.OA
math
Infinite Tensor Products of C0(R) : Towards a Group Algebra for R(N). Hendrik Grundling Karl -- Hermann Neeb Department of Mathematics, Fachbereich Mathematik, University of New South Wales, Technische Universitat Darmstadt, Sydney, NSW 2052, Australia. Schlossgartenstrasse 7, [email protected] D -- 64289 Darmstadt Germany. FAX: +61-2-93857123 [email protected] Running title: Towards a Group Algebra for R(N). Abstract The construction of an infinite tensor product of the C*-algebra C0(R) is not obvious, because it is nonunital, and it has no nonzero projection. Based on a choice of an approximate identity, we construct here an infinite tensor product of C0(R), denoted LV , and use it to find (partial) group algebras for the full continuous representation theory of R(N). We obtain an interpretation of the Bochner -- Minlos theorem in R(N) as the pure state space decomposition of the partial group algebras which generate LV . We analyze the representation theory of LV , and show that there is a bijection between a natural set of representations of LV and Rep(cid:0)R(N), H(cid:1) , but that there is an extra part which essentially consists of the representation theory of a multiplicative semigroup Q which depends on the initial choice of approximate identity. Keywords: C*-algebra, group algebra, infinite tensor product, topological group, Bochner -- Minlos theorem, state space decomposition, continuous representation. Mathematics Classification: 22D25, 46L06, 43A35. 1 Introduction The class of locally compact groups has a rich structure theory with a great many tools developed to analyze the representation theory of such groups, e.g., group C*-algebras, induction, integral decompo- sitions etc. Unfortunately there are many non-locally compact groups which naturally arise in analysis or physics applications, e.g. mapping groups or inductive limit groups, and for such groups these tools fail, and one has to do the analysis on a case-by-case basis, with no systematic theory to draw on. Here we want to consider the question of how to generalize the notion of a (twisted) group algebra to topological groups which are not locally compact (hence have no Haar measure). Such a generalization, called a full host algebra, has been proposed in [Gr05]. Briefly, it is a C∗ -algebra A whose multiplier algebra M (A) admits a homomorphism η : G → U (M (A)), such that the (unique) extension of the 1 representation theory of A to M (A) pulls back via η to the continuous unitary representation theory of G . There is also an analogous concept for unitary σ -- representations, where σ is a continuous T -valued 2 -cocycle on G . Thus, given a full host algebra A, the continuous unitary representation theory of G can be analyzed on A with a large arsenal of C∗ -algebraic tools. Such a host algebra need not exist for a general topological group because there exist topological groups with faithful unitary representations but without non-trivial irreducible ones (cf. [GN01]). One example of a full host algebra for a group which is not locally compact, has been constructed explicitly for the σ -- representations of an infinite dimensional topological linear space S, considered as a group cf. [GrN09]. Probably the simplest infinite dimensional group is R(N) (the set of real-valued sequences with the natural inclusions only finitely many nonzero entries) with the inductive limit topology w.r.t. Rn ⊂ R(N). This group is well -- studied in stochastic analysis, and will be the main object of study also in this paper. Our aim here is to construct explicitly C*-algebras which have useful host algebra properties for R(N). Recall that for the group C*-algebras we have: C∗(Rn) ⊗ C∗(Rm) ∼= C∗(Rn+m) and this suggests that for a host algebra of R(N) we should try an infinite tensor product of C∗(R). This is difficult to do, for two reasons: • C∗(R) ∼= C0(R) is nonunital, and the standard infinite tensor products of C*-algebras require unital algebras. • There is a definition for an infinite tensor product of nonunital algebras developed by Blackadar cf. [Bl77], but this requires the algebras to have nonzero projections, and the construction depends on the choice of projections. (We used this construction in [GrN09] to construct an infinite tensor product to produce a host algebra.) However, C∗(R) ∼= C0(R) has no nonzero projections, so this method will not work. In the light of these difficulties, we will develop here an infinite tensor product of C0(R) relative to a choice of approximate identity in each entry, to replace the choice of projections in Blackadar's approach. As expected, the construction will depend on the choice of approximate identities, though it still produces for each choice an algebra with strong host algebra properties. The construction of ("semi-")host algebras for R(N) will aid our understanding of the Bochner -- Minlos theorem. We first recall: 1.1 Theorem (Bochner -- Minlos Theorem for R(N) ) There is a bijection between continuous normalized positive definite functions (states) ω of R(N) and regular Borel probability measures µ on RN (with product topology) given by the Fourier transform: ω(x) =ZRN eix·ydµ(y) , x ∈ R(N) where x · y := ∞Pn=1 xnyn , x ∈ R(N), y ∈ RN. If we replace both R(N) and RN by Rn, this is the classical Bochner theorem, which we can obtain immediately from the state space integral decomposition of any state of C∗(Rn) ∼= C0(Rn) in terms of 2 pure states. This suggests that if we have a host algebra of R(N), we can obtain the Bochner -- Minlos theorem from state space decompositions of states on the host algebra in terms of pure states. We will see below that we can already obtain the Bochner -- Minlos theorem from the weaker "semi -- host" algebras which we will construct. The structure of this paper is as follows. In Section 2 we collect the basic definitions and notation for host algebras, in Section 3 we give a detailed treatment of the aspects of infinite tensor products which we will need for this paper. In Section 4 we start in a concrete setting on L2(RN, µ), where µ is a product measure of probability measures, each absolutely continuous w.r.t. the Lebesgue measure, and we construct an infinite tensor product of C0(R) w.r.t. a choice (compatible with µ) of approximate identity in each entry. This concrete C*-algebra can already produce Bochner -- Minlos decompositions for the limited class of positive definite functions on R(N) associated with it. In Section 5 we develop abstractly the infinite tensor product of C0(R) w.r.t. an arbitrary choice of elements of a fixed approximate identity, we analyze its representation theory and through the unitary embedding of R(N) in its multiplier algebra, we consider the relation of its representation theory to that of R(N). We find that it can adequately model a subset of the representation theory of R(N), but there is a small additional part. We show that the Bochner -- Minlos decompositions for any continuous positive definite function on R(N) can be obtained from the pure state space decomposition of these algebras. Finally, in Section 6, we collect these algebras together in one large C*-algebra, which we show, can model the full continuous representation theory of R(N). However, the representation theory of this algebra also has an additional part which essentially consists of the representation theory of a multiplicative semigroup Q which depends on the initial fixed choice of approximate identity. 2 Definitions and notation We will need the following notation and concepts for our main results. • In the following, we write M (A) for the multiplier algebra of a C∗ -algebra A and, if A has a unit, U (A) for its unitary group. We have an injective morphism of C∗ -algebras ιA : A → M (A) and will just denote A for its image in M (A) . Then A is dense in M (A) with respect to the strict topology, which is the locally convex topology defined by the seminorms pa(m) := km · ak + ka · mk, a ∈ A, m ∈ M (A) (cf. [Wo95]). • For a complex Hilbert space H , we write Rep(A, H) for the set of non-degenerate represen- tations of A on H . Note that the collection Rep A of all non-degenerate representations of A is not a set, but a (proper) class in the sense of von Neumann -- Bernays -- Godel set theory, cf. [Tak75], and in this framework we can consistently manipulate the object Rep A. However, to avoid set -- theoretical subtleties, we will express our results below concretely, i.e., in terms of Rep(A, H) for given Hilbert spaces H. We have an injection Rep(A, H) ֒→ Rep(M (A), H), which identifies the non-degenerate representation π of A with that representation eπ of its multiplier algebra which extends π and is continuous with respect to the strict topology on π 7→ eπ with eπ ◦ ιA = π, 3 M (A) and the topology of pointwise convergence on B(H) . We will refer to eπ as the strict extension of π, and it is easily obtained by where {Eλ}λ∈Λ ⊂ A is any approximate identity of A. eπ(M ) = s-lim λ→∞ π(M Eλ) ∀ M ∈ M (A) • For topological groups G and H we write Hom(G, H) for the set of continuous group homo- morphisms G → H . We also write Rep(G, H) for the set of all (strong operator) continuous unitary representations of G on H . Endowing U (H) with the strong operator topology turns it into a topological group, denoted U (H)s , so that Rep(G, H) = Hom(G, U (H)s) . The set of continuous normalized positive definite functions on G (also called states) and denoted by is in bijection with the state space of the group C*-algebra C∗(G) when G is locally compact. If G is not locally compact, S(G) is in bijection with a subset of the state space of C∗(Gd) , where Gd denotes G with the discrete topology, and the question arises as to whether there is a C*-algebra which can play the role of C∗(G). We clarify first what is meant by this: S(G), 2.1 Definition Let G be a topological group. A host algebra for G is a pair (L, η) where L is a C∗ -algebra and η : G → U (M (L)) homomorphism such that for each complex Hilbert space H the corresponding map is a η∗ : Rep(L, H) → Rep(G, H), π 7→ eπ ◦ η is injective. We then write Rep(G, H)η ⊆ Rep(G, H) for the range of η∗ . We say that (L, η) is a full host algebra of G if η∗ is surjective for each Hilbert space H . If the map η∗ is not injective, we will call the pair (L, η) a semi-host algebra for G. Note that by the universal property of group algebras, the homomorphism η : G → U (M (L)) extends uniquely to the discrete group C*-algebra C∗(Gd), η : C∗(Gd) → U (M (L)) (still denoted by η ). i.e. we have a *-homomorphism A similar notion can also be defined for projective representations (cf. [GrN09]). 2.2 Remark (1) It is well known that for each locally compact group G , the group C∗ -algebra C∗(G) , and the natural map ηG : G → M (C∗(G)) provide a full host algebra ( [Dix77, Sect. 13.9]). The map ηG : G → M (C∗(G)) the strict topology of M (C∗(G)) (this is an easy consequence of the fact that im(ηG) is bounded and that the action on the corresponding L1 -algebra is continuous). is continuous w.r.t. (2) Note that for a host algebra (L, η) the map η∗ preserves direct sums, unitary conjugation, subrepresentations, and for full host algebras, irreducibility (cf. [Gr05]). (3) When (L, η) is merely a semi-host algebra for G, then the map η∗ still preserves direct sums, unitary conjugation, subrepresentations, but in general, not irreducibility. However, in the case that G is Abelian (as it will be in this paper), since irreducible representations are just charac- ters, and the map η∗ takes one -- dimensional representations to one -- dimensional ones, here it will preserve irreducibility. So for Abelian groups, semi -- hosts are useful to carry representation struc- ture (e.g. integral decompositions) from the representation theory of L to the representation theory of G, and we will use that in this paper to analyze the Bochner -- Minlos theorem. 4 3 Basic Theory of Infinite Tensor Products Since we need to develop the concept of infinite tensor products of non-unital algebras, it is neces- sary to collect first some basic material on infinite tensor products, and to fix notation. We follow Bourbaki [Bou89] and Wegge -- Olsen [WO93]. There are several different concepts of infinite tensor products of unital algebras. See Bourbaki [Bou89], Guichardet [Gu67], Araki [AW66], though infinite tensor products of algebras without identity are only done in Blackadar [Bl77]. 3.1 Algebraic tensor products of arbitrary many factors. 3.1 Definition Let (Xt)t∈T be an indexed set of non-zero complex vector spaces, where T can Xt. A map have any cardinality. We write x = (xt)t∈T for the elements of the product space Qt∈T f : Qt∈T each t0 ∈ T and x ∈ Qt∈T \{t0} Xt → V to a vector space V is said to be multilinear if it is linear in each entry. That is, for Xt, the map Xt0 → V, yt0 7→ f (x × yt0 ) Xt is that element for which zt = xt if t 6= t0 and zt0 = yt0. is linear, where x × yt0 =: z ∈ Qt∈T A pair (ι, V ) consisting of a vector space V and a multilinear map ι : Qt∈T (algebraic) tensor product of (Xt)t∈T if it has the following universal property Xt → V is called an (UP) For each multilinear map ϕ : Qt∈T Xt → W , there exists a unique linear map eϕ : V → W with The usual arguments (cf. Proposition T.2.1 [WO93]) show that the universal property determines a tensor product up to linear isomorphism (factoring through the maps ι). We may thus denote V eϕ ◦ ι = ϕ . by Nt∈T Xt and denote the elementary tensors by ⊗ t∈T xt := ι(x) ∈Ot∈T Xt, for x ∈ Yt∈T Xt . To simplify notation, we write X := Qt∈T Xt in the following. Observe that no order in T appears in this definition, so e.g. X1 ⊗ X2 and X2 ⊗ X1 (in the usual notation) will be identified. 3.2 Lemma For each indexed set (Xt)t∈T of complex vector spaces, a tensor product exists. (ι, Nt∈T Xt) Proof: (cf. [Bou89, Ch. II,§3.9] for a more general construction) We consider the free complex vector space C(X) :=(cid:8)f : X → C supp(f ) is finite(cid:9) = Span(cid:8)δx x ∈ X(cid:9) where δx(y) = 1 if x = y and zero otherwise. Note that (cid:8)δx x ∈ X(cid:9) is a basis for C(X) . Define the sets Na := (cid:8)δx + δy − δz(cid:12)(cid:12) ∃ r ∈ T such that xr + yr = zr , and xt = yt = zt ∀ t 6= r(cid:9) Nm := (cid:8)δx − µδy(cid:12)(cid:12) µ ∈ C, and ∃ r ∈ T such that xr = µyr , and xt = yt ∀ t 6= r(cid:9) N := Span(cid:0)Na ∪ Nm(cid:1) ⊂ C(X) . 5 We now consider the quotient space V := C(X)(cid:14)N and write ι : X → V, x 7→ δx + N for the induced map. The definition of N immediately implies that ι is multilinear and we only have to verify the universal property. Let ϕ : X → M be a multilinear map. We extend ϕ to a linear map ϕ : C(X) → M by f (x) ϕ(x) . The multilinearity of ϕ now implies that its linear extension annihilates the ϕ(f ) := Px∈X subspace N , hence it factors through a linear map eϕ : V → M satisfying eϕ ◦ ι = ϕ. That eϕ is uniquely determined by this property follows from the fact that im(ι) spans V. 3.3 Theorem (Associativity) Let (cid:8) Ts ⊂ T (cid:12)(cid:12) s ∈ S(cid:9) be a partition of T such that ψ : Yt∈T Xt → Ns∈S(cid:0) Nts∈Ts Xts(cid:1), ψ((xt)t∈T ) := Ns∈S(cid:0) Nts∈Ts S < ∞ . Then the map is multilinear and factors through a linear isomorphism eψ : Nt∈T Proof. It is clear from the definition that ψ is multilinear, so we obtain a unique linear map eψ : Xt → Ns∈S(cid:0) Nt∈Ts Nt∈T To see that eψ is a linear isomorphism, it suffices to observe that the multilinear map ψ has the universal property (UP). So let ϕ : X → V be a multilinear map. With Ys := Qt∈Ts X = Qs∈S Ys . Then for each s0 ∈ S and for each y ∈ Qs∈S\s0 Xt(cid:1) with eψ ◦ ι = ψ . Ys we obtain a unique map Xt , we have Xts(cid:1). xts(cid:1) Xt → Ns∈S(cid:0) Nts∈Ts ϕs0 y : Ys0 = Yt∈Ts0 Xt → V, ϕs0 y (ys0 ) := ϕ(y × ys0 ) , which is clearly multilinear w.r.t. the factors Qt∈Ts0 Xt = Ys0 hence induces a linear map on Nt∈Ts0 Xt. Xt, we can apply the argument Since y 7→ ϕs0 y (v) is multilinear in y ∈ Qs∈S\s0 Ys for fixed v ∈ Nt∈Ts0 again to an s1 6= s0 ∈ S for this map, and then continue the process until we have exhausted S. This produces a multilinear map Xt(cid:17) → V bϕ : Ys∈S(cid:16)Ot∈Ts eϕ(cid:0) ⊗ s∈S(cid:0) ⊗ with which factors through a linear map eϕ : Os∈S(cid:16)Ot∈Ts Xt(cid:17) → V i.e., eϕ ◦ ψ = ϕ . Moreover, since Ns∈S(cid:16) Nt∈Ts xts(cid:1) it follows that eϕ is uniquely determined by the last equation. Thus ψ has the universal property (UP), hence eψ is a linear isomorphism. Xt(cid:17) is spanned by elements of the form ⊗ xts(cid:1)(cid:1) = ϕ((xt)t∈T ), s∈S(cid:0) ⊗ ts∈Ts ts∈Ts 3.4 Remark Associativity does not seem to hold for a partition of T into infinitely many sets (i.e., x(ts) is spanned by elementary tensors, and Ns∈S(cid:16) nsPts=1 Nrs∈Ts for S = ∞ ). This is because Nt∈T rs (cid:17) Xt 6 cannot be written as a finite linear combination of elementary tensors if there are infinitely many s ∈ S with ns > 1 . 3.5 Definition (a) Assume that (Xt)t∈T is a family of complex algebras. We now construct an algebra structure on their tensor product. For each fixed x ∈ X = Qt∈T Xt by µx(y) := Nt∈T µx : X → Nt∈T Xt , define a map xtyt = ι(x · y) where x · y ∈ X is given by (x · y)t := xtyt (xn)t := (xt)n for all t ∈ T and n ∈ N. Since µx is multilinear, it induces a linear map for all t ∈ T , and we will also let xn ∈ X denote This defines a multilinear map µx :Ot∈T Xt →Ot∈T Xt. µ : X → End(cid:0) Nt∈T Xt → End(cid:0) Nt∈T Xt(cid:1) by µ(x) := µx Xt(cid:1) . Explicitly we have for a = Pi ι(xi) and b = and thus a linear map µ : Nt∈T ι(yj ) ∈ Nt∈T Pj Xt that µ(a)(b) =Xi µxi(cid:16)Xj ι(yj )(cid:17) =Xi Xj µxi(cid:0)ι(yj )(cid:1) =Xi Xj ι(xi · yj) where the sums are finite. We denote the multiplication as usual by a b := µ(a)(b) for a, b ∈ Nt∈T Associativity for this multiplication follows from componentwise associativity, and hence Nt∈T is a ∗ -algebra. We want to turn Nt∈T (b) Next, we assume, in addition, that each Xt algebra over C . Xt Xt Xt . is an into a ∗ -algebra. Given any vector space V over C , each t ∈ T , the involution ∗ : Xt → X c Xt ). Define a map let V c denote the conjugate vector space. Thus, for t becomes a C -- linear map (instead of conjugate linear on Xt(cid:1)c by γ(x) := Nt∈T x∗ t = ι(x∗) γ : X →(cid:0) Nt∈T Xt(cid:1)c where x∗ ∈ X is given by (x∗)t := x∗ t map γ : Nt∈T componentwise properties. As usual, we write a∗ := γ(a) for a ∈ Nt∈T Xt → (cid:0) Nt∈T a ∗ -algebra over C . for all t ∈ T . Since γ is multilinear, it defines a linear . Its intertwining properties with multiplication then follow from the Xt , and hence Nt∈T Xt becomes This defines the basic objects which we will work with. 3.2 Stabilized spaces. We will also need the following structures. 3.6 Definition We define an equivalence relation on X by x ∼ y whenever the set {t ∈ T xt 6= yt} is finite. Denote the equivalence class of x ∈ X by [x]∼ and define JxK := Span(cid:8) ⊗ t∈T yt y ∼ x(cid:9) ⊂ Nt∈T Xt . 7 3.7 Proposition The following assertions hold: (i) For any pair (x, F ) such that x ∈ X and F ⊆ T a finite subset with xt 6= 0 for t 6∈ F , there exists a linear map ϕF : Ot∈T Xt →Ot∈F Xt satisfying JyK ⊆ Ker ϕF for y 6∼ x and ϕF(cid:0)( ⊗ t∈F yt) ⊗ ( ⊗ t6∈F xt)(cid:1) = ⊗ t∈F yt for yt ∈ Xt, t ∈ F. (ii) JxK 6= {0} if and only if at most finitely many components of x vanish. (iii) The subspace JxK is isomorphic to the direct limit of the finite tensor products ⊗ t∈J Xt , J ⊆ T finite, with respect to the connecting maps ϕK,J : Nt∈J Xt → Nt∈K Xt with ϕK,J(cid:0) ⊗ t∈J yt(cid:1) :=(cid:0) ⊗ t∈J yt(cid:1) ⊗(cid:0) ⊗ s∈K\J xs(cid:1). (iv) Nt∈T Xt is the direct sum of the subspaces JxK , x ∈ X . Proof: (i) For t 6∈ F we pick linear functionals λt ∈ X ∗ t with λt(xt) = 1 and define a map bϕF : X → Of ∈F Xf , bϕF (y) :=  Qt∈T \F λt(yt) ·(cid:0) ⊗ s∈F ys(cid:1) 0 for y ∼ x for y 6∼ x. ys = y′ s for s 6= t . Then either both are equivalent to x or none is. In either case, the definition of We claim that bϕF is multilinear. To see that bϕF is linear in the t -component, let y, y′ ∈ X with bϕF implies the linearity of the map zt 7→ bϕF (y × zt). Therefore bϕF is multilinear, hence induces a linear map ϕF : O Xt →Ot∈F Xt satisfying all requirements. (ii) If the set {t ∈ T xt = 0} is finite, then (i) implies that JxK 6= {0} since none of the spaces Xt vanishes by our initial assumption. We also note that, if infinitely many xt vanish, then JxK is spanned by elements ι(y) , where y has at least one zero entry. Then ι(y) = 0 , and consequently JxK = {0} . (iii) Let J ⊂ K ⊂ T such that K < ∞ . Then we obtain linear maps Since ϕL,K ◦ ϕK,J = ϕL,J for J ⊂ K ⊂ L , and L < ∞ , this is an inductive system. We write lim yt(cid:1) :=(cid:0) ⊗ t∈J yt(cid:1) ⊗(cid:0) ⊗ s∈K\J xs(cid:1). ϕK,J : Nt∈J Xt → Nt∈K Xt with ϕK,J(cid:0) ⊗ Xt, ϕK,J(cid:1) for its limit. We also have linear maps yt(cid:1) :=(cid:0) ⊗ Xt → JxK by ϕJ(cid:0) ⊗ ϕJ : Nt∈J t∈J t∈J t∈J −→(cid:0)Nt∈J satisfying ϕK ◦ ϕK,J = ϕJ , so that they induce a linear map ϕ : lim s∈T \J yt(cid:1) ⊗(cid:0) ⊗ −→(cid:0)Nt∈J xs(cid:1) ∈ JxK Xt, ϕK,J(cid:1) → JxK . As every element of JxK lies in the image of some map ϕJ , and by (i) this map is injective if J ⊇ {t ∈ T xt = 0} , ϕ is a linear isomorphism. 8 (iv) Since ι(x) is contained JxK , it suffices to show that the sum of the non-zero sub- JxK spaces is direct. Suppose that the elements x1, . . . , xn are pairwise non-equivalent with JxiK 6= {0} , and that vi ∈ JxiK satisfy Pi vi = 0 . From (i) we know that there exists for each i and each finite subset F ⊇ {t ∈ T xi,t = 0} a linear map ϕ(i) F : Ot∈T Xt →Ot∈F Xt with ϕ(i) F (cid:0)( ⊗ t∈F yt) ⊗ ( ⊗ t6∈F xi,t)(cid:1) = ⊗ t∈F yt and vanishing on Jxj K for j 6= i . We conclude that ϕ(i) chosen arbitrarily large, the definition of JxiK now implies that vi = 0 . F (vi) = 0 for each F . Since F can be is an algebra and x2 3.8 Remark If each Xt linear space JxK is a subalgebra. If each Xt many t ∈ T , then JxK is a ∗ -subalgebra. In the literature (on topological tensor products), suitable t = xt holds for all but finitely many t ∈ T , then the for all but finitely is a ∗ -algebra and x∗ t = xt = x2 t closures of JxK are often called stabilized infinite tensor products (stabilized by x ). 3.9 Remark In particular, for x, y ∈ X with JxK 6= {0} 6= JyK , we have that JxK ∩ JyK = {0} if and only if x 6∼ y . So, if yt = λtxt where λt 6= 1 for infinitely many t ∈ T , then x 6∼ y and hence ⊗ xt. This is different in Guichardet's version [Gu67] of continuous t∈T tensor products. is not a multiple of ⊗ t∈T λtxt When the Xt are algebras, we have the following algebraic relations for the spaces JxK in the algebra Nt∈T Xt . 3.10 Theorem If each Xt is a complex algebra, then (i) JxK · JyK ⊆ Jx · yK for all x, y ∈ X . If Xt · Xt = Xt Span(cid:0)JxK · JyK(cid:1) = Jx · yK . (ii) JxK∗ = Jx∗K for all x ∈ X if all Xt are ∗ -algebras. for all t , then we have the equality: (iii) If ∅ 6= Gt ⊂ Xt \ {0} is a nonzero multiplicative semigroup for each t ∈ T , then M := Xa∈Qt∈T Gt JaK (finite sums) Xt . If in addition, each Xt is a ∗ -algebra and each Gt is ∗ -invariant, is a subalgebra of Nt∈T then M is a ∗ -subalgebra. Proof. (i) Since JxK is spanned by elements of the form ι(a) , a ∼ x and JyK likewise by elements ι(b) with b ∼ y , and we have a·b ∼ x·y, the first assertion follows from ι(a)ι(b) = ι(a·b) ∈ Jx·yK . To show that we have equality when Xt · Xt = Xt elements of the form ι(a) =(cid:16) ⊗ s∈S as(cid:17) ⊗(cid:16) ⊗ t∈T \S for all t , note that Jx · yK is spanned by xtyt(cid:17) , where S is finite. Since each as ∈ XsXs by assumption, it follows that ι(a) ∈ JxKJyK , which proves the required equality. (ii) Since ∗ is involutive, it suffices to show that JxK∗ ⊆ Jx∗K . As JxK∗ is spanned by elements of the form ι(a)∗ , a ∼ x , the assertion follows from ι(a)∗ = ι(a∗) with a∗ ∼ x∗ . 9 (iii) Since the set {x ∈ X xt ∈ Gt ∀ t ∈ T } is a semigroup w.r.t. the componentwise multiplication, the first statement regarding M follows from (i). The second statement likewise follows from (ii). 3.11 Remark (a) Regarding the condition Xt · Xt = Xt Theorem 5.2.2 in [Pa94], we know that if A is a Banach algebra with a bounded left approximate in part (i), this is easily fulfilled, since by identity and T : A → B(X) is a continuous representation of A on the Banach space X , then for each y ∈ Span(T (A)X) there are elements a ∈ A and x ∈ X with y = T (a)x . Thus, if X = A (b) In regard to the choice of semigroup Gt and T : A → B(X) is defined by T (A)B := AB , then since A has an approximate identity, we have A = Span(T (A)X) and hence A · A = A . In particular, A · A = A for any C∗ -algebra A . in (iii) above, when one has unital algebras, the conventional choice is to set all Gt = {1} . If the ∗ -algebras Xt are nonunital but have projections, then one can take each Gt to be a projection (cf. Blackadar [Bl77]) though the final tensor product algebra depends on this choice of projections. If the ∗ -algebras Xt have no nonzero projections, e.g. C0(R) below, then we will choose each Gt to be a small ∗ -closed semigroup generated by one element (which will be positive, of norm 1 ). 3.3 Tensor products of representations. Below, we will need to complete some ∗ -subalgebras of the algebraic tensor product in the operator norm of a suitable representation, hence need to make explicit the structures involved with infinite tensor products of Hilbert space representations. Let (Ht)t∈T be a family of Hilbert spaces. We want to equip selected subspaces of Nt∈T Ht with (xt, yt)t " whenever the right hand side makes sense. There the inner product (ι(x), ι(y)) := " Qt∈T are many possibilities, but here we recall the tensor product constructions of von Neumann [vN61]. Let L :=nx ∈ Yt∈T Ht(cid:12)(cid:12)(cid:12)Xt∈T(cid:12)(cid:12)kxtkt − 1(cid:12)(cid:12) < ∞o that only countably many summands {αtn n ∈ N} are non-zero and that S = where we interpret the convergence of a sum (resp. product) over an uncountable set T as convergence αt, αt ∈ C, this implies of the net of finite partial sums, resp., products. For sums such as S := Pt∈T ∞Pn=1 converges absolutely (cf. Lemmas 2.3.2 and 2.3.3 in [vN61]). Moreover, we have that P = Qt∈T if and only if either αt = 0 for some t (in which case P = 0 ), or else Pt∈T(cid:12)(cid:12)αt − 1(cid:12)(cid:12) < ∞ (cf. [vN61, Lemma 2.4.1]). We will not need to use general products P = Qt∈T αt, αt ∈ C, for which the αtn , and it αt < ∞ convergence is a more difficult notion (cf. Lemma 2.4.2 and Definition 2.5.1 in [vN61]). Thus x ∈ L implies that kxtkt = 1 for all t ∈ T \R where R is at most countable, and that the kxtkt converges. Obviously, any x such that kxtkt = 1 for all t ∈ T is in L . Note product Qt∈T that if x ∈ L then [x]∼ ⊂ L also. For x, y ∈ L , we define x ≈ y if Xt∈T(cid:12)(cid:12)(xt, yt)t − 1(cid:12)(cid:12) < ∞ . 10 (1) Then ≈ is an equivalence relation by Lemma 3.3.3 in [vN61], and we denote its equivalence classes by [x]≈ . Observe that if x ∈ L then [x]∼ ⊂ [x]≈ , and moreover, each ≈-equivalence class contains an a ∈ L such that katkt = 1 for all t ∈ T (cf. Lemma 3.3.7 in [vN61]). 3.12 Definition Given such an a ∈ [x]≈ ⊂ L , we can define an inner product on JaK by sesqui-linear extension of (cid:0)ι(x), ι(y)(cid:1) := Yt∈T (xt, yt)t for x ∼ a ∼ y . (Note that the infinite products occurring here have only finitely many entries different from 1 hence [a]Ht . Then this is are unproblematic). Denote the closure of JaK w.r.t. this Hilbert norm by Nt∈T von Neumann's "incomplete direct product," and it contains Span(cid:0)ι([a]≈)(cid:1) as a dense subspace (cf. Lemma 4.1.2 in [vN61]). The direct sum of the spaces Nt∈T [a]Ht where we take one representative a from each ≈-equivalence class, is von Neumann's "complete direct product" (cf. Lemma 4.1.1 in [vN61]). An analogous associativity theorem to Theorem 3.3 holds for this complete direct product (cf. Theorem VII in [vN61]). Next, consider the case where (At)t∈T is a family of ∗ -algebras, each equipped with a bounded At we can define a linear map Hilbert space ∗ -representation πt : At → B(Ht) . For any A ∈ Qt∈T π(cid:0)ιA(A)(cid:1) on ⊗ Ht by t∈T π(cid:0)ιA(A)(cid:1)ι(x) = ⊗ t∈T πt(At)xt = ι(cid:0)π(A)x(cid:1) for all x ∈ Yt∈T Ht for all t ∈ T . Then π is a representation, because it is one for each entry. To obtain Hilbert space ∗ -representations from π , we need to restrict it to suitable pre-Hilbert Ht hence need to restrict to those A such that π(cid:0)ιA(A)(cid:1) preserves the Hilbert space involved (and produces a bounded operator). where (cid:0)π(A)x(cid:1)t := πt(At)xt subspaces of Nt∈T 3.13 Definition Consider the Hilbert space completion Nt∈T At are all unital, then J1K ⊂ Nt∈T π(A)x ∈ [a]∼ for all x ∈ [a]∼ ⊂ Qt∈T identity operator. Thus it extends to a bounded operator on Nt∈T of the ∗ -algebra J1K on the (stabilized) tensor product Nt∈T At definition of a tensor representation. [a]Ht of JaK , as above. When the algebras is a ∗ -subalgebra, where (1)t = 1t ∈ At for all t ∈ T . Then Ht and A ∼ 1 . In particular, π(cid:0)ιA(A)(cid:1) preserves JaK and it is bounded, since it is a tensor product of a finite tensor product (of bounded operators) with the [a]Ht . This defines a ∗ -representation [a]Ht , and it is the most commonly used When the ∗ -algebras At are not unital, consider the case where they contain nontrivial hermitian At At , the subspace JPK ⊂ Nt∈T for all t ∈ T , we can now define a tensor projections Pt ∈ At . Then, for any choice of such projections P ∈ Qt∈T is a ∗ -subalgebra. For any a ∈ Qt∈T product representation of JPK on Nt∈T Ht with πt(Pt)at = at representations. [a]Ht . Below we will consider more general tensor product 11 4 Semi -- host algebras for Gaussians ∞Pi=1 In this section, µ will be a fixed Gaussian product measure on RN and µn denotes its projection on the nth component. For x ∈ RN and y ∈ R(N) , we write hx, yi := for the standard xiyi pairing. Recall that from µ one constructs a unitary representation πµ : R(N) → U(cid:0)L2(RN, µ)(cid:1) by (cid:0)πµ(x)f(cid:1)(y) := exp(cid:0)ihx, yi(cid:1) f (y), x ∈ R(N), y ∈ RN. Then there is a unitary map U : [e]Hn → L2(RN, µ) , where Hn := L2(R, µn) . The sequence e = (e1, e2, . . .) of stabilizing vectors en ∈ Hn is given by the constant functions en(x) = 1 for all x ∈ R . Explicitly, U is given by ∞Nn=1 U (f1 ⊗ f2 ⊗ · · · ⊗ fk ⊗ ek+1 ⊗ ek+2 ⊗ · · · )(x1, x2, . . .) = f1(x1) · f2(x2) · · · fk(xk) which is clearly a cylinder function on RN . Then πµ = U(cid:16) ∞Nn=1 πµn(cid:17)U −1 , where each πµn : R → U(cid:0)L2(R, µn)(cid:1) is (cid:0)πµn (x)f(cid:1)(y) := eixy f (y) for x, y ∈ R . The stabilizing sequence defines a cyclic vector Ω := the corresponding positive definite function satisfies: ∞ ⊗ n=1 en . Immediate calculation establishes that ωµ(t) :=(cid:0)Ω, πµ(t)Ω(cid:1) =ZRN exp(cid:0)iht, yi(cid:1) dµ(y) for t ∈ R(N) , (2) which is part of the Bochner -- Minlos Theorem (cf. [GV64]). We will show that it expresses the decom- position of a state into the pure states of a (semi-) host algebra for R(N) , and that there is a similar expression for other states (which is also part of the Bochner -- Minlos theorem). Specialize the notation of the last section by setting: T = N and Xt = C0(R) ∼= C∗(R) for all t . We first try to define an appropriate infinite tensor product C∗ -algebra of all the C0(R)'s , which seems to be a problem because C0(R) is nonunital, and has no nontrivial projection. By the last section we always have the algebraic tensor product C0(R) , but this is too large. We want to look at its ∗ -subalgebras of the type defined in Theorem 3.10(iii), and will consider the following multiplicative semigroups in C0(R) . For each n ∈ N , define Vn :=(cid:8)f ∈ C0(R)(cid:12)(cid:12) f (R) ⊆ [0, 1], f ↾ [−n, n] = 1, supp(f ) ⊆ [−n − 1, n + 1](cid:9) and observe that it is a semigroup, that kf k = 1 for all f ∈ Vn and that any sequence {un ∈ Vn n ∈ N} is an approximate identity for C0(R) . Moreover Vn · Vm = Vn if m > n ∞Nk=1 and hence Vn is a semigroup. For each f ∈ Vn we have the subsemigroup ∞Sn=1 and for these we also have that Vn(f ) · Vm(g) = Vn(f ) if m > n . Vn(f ) := {f k k ∈ N} ⊂ Vn, For any sequence f = (f1, f2, . . .) ∈ C0(R)N with fn ∈ Vkn for all n , we consider the ∗ -algebra generated in C0(R) by Jf K , and note that ∞Nk=1 ∗ -alg(cid:0)Jf K(cid:1) = Span(cid:8)Jf kK k ∈ N} ⊂ ∞Oi=1 C0(R), where (cid:0)f k(cid:1)n := f k n ∀ n (3) 12 and for the equality we needed the fact that C0(R) · C0(R) = C0(R) (Remark 3.11), and Theo- rem 3.10(i). to close it in. We will show that there are f Next, we want to define a convenient representation of ∗ -alg(cid:0)Jf K(cid:1) to provide us with a C∗ -norm for which we can define a representation of ∗ -alg(cid:0)Jf K(cid:1) ∞Nn=1 [e]Hn in a natural way. on 4.1 Proposition We now have: (i) Let Pk denote multiplication of functions on R by χ[−k,k] . Then there exists a sequence (ki)i∈N such that ∞Pn=1(cid:12)(cid:12)(Pkn en, en)n − 1(cid:12)(cid:12) < ∞ . (ii) Fix a sequence (ki)i∈N as in (i) as well as f ∈ πe : ∗ -alg(cid:0)Jf K(cid:1) → B(cid:16) ∞Nn=1 [e]Hn(cid:17) such that gn(cid:17) ∞ πe(cid:16) ∞ ⊗ n=1 ⊗ k=1 ck = ∞ ⊗ n=1 gncn ∈ [e]Hn ∞Nn=1 ∞Qj=1 Vkj . Then there is a ∗ -representation for all g ∼ f ℓ, c ∼ e and ℓ ∈ N , and where gncn is the usual pointwise product of functions on R . Proof. (i) For any ε > 0 , there is a k ∈ N such that (cid:12)(cid:12)(Pken, en)n − 1(cid:12)(cid:12) < ε by the Monotone Convergence Theorem. Thus there is a sequence (ki)i∈N such that ∞Pn=1(cid:12)(cid:12)(Pkn en, en)n − 1(cid:12)(cid:12) < ∞ . [e]Hn of JeK , where (ii) Recall from Definition 3.12 that Span(cid:0)ι([e]≈)(cid:1) is dense in the closure ∞Nn=1 ∞Xn=1(cid:12)(cid:12)(en, vn)n − 1(cid:12)(cid:12) < ∞o . fn(x) dµn(x) = (fnen, en)n ≤ 1 With the given choice of (ki)i∈N and f we have ∞Yn=1 Hn (cid:12)(cid:12)(cid:12) [e]≈ =nv ∈ ∞Xn=1(cid:12)(cid:12)kvnkn − 1(cid:12)(cid:12) < ∞ and (Pkn en, en)n = µn(cid:0)[−kn, kn](cid:1) ≤ Z kn+1 −kn−1 so that (cid:12)(cid:12)(fnen, en)n − 1(cid:12)(cid:12) ≤ (cid:12)(cid:12)(Pkn en, en)n − 1(cid:12)(cid:12), ∞Pn=1(cid:12)(cid:12)kf ℓ nen, en)n − 1(cid:12)(cid:12) < ∞ for all ℓ ∈ N . This implies that ∞Pn=1(cid:12)(cid:12)(fnen, en)n − 1(cid:12)(cid:12) < ∞ . As (fj)ℓ ∈ Vkj and hence ∞Pn=1(cid:12)(cid:12)(f ℓ via Lemma 3.3.2 in [vN61] that for all ℓ ∈ N , we have in fact that nenk2 n − 1(cid:12)(cid:12) < ∞ which implies ∞Pn=1(cid:12)(cid:12)kf ℓ nenkn − 1(cid:12)(cid:12) < ∞ . Hence f ℓ · e ∈ [e]≈ and so n(cid:17)(cid:16) ∞ (cid:16) ∞ ek(cid:17) = f ℓ nen ∈ ∞ ⊗ n=1 [e]Hn . ⊗ n=1 ⊗ k=1 f ℓ ∞Nn=1 Since any c ∼ e only differs from e in finitely many entries, the convergence arguments above will still hold if we replace e by c . Likewise, we can replace f ℓ by any g ∼ f ℓ , i.e., we have shown [e]Hn for all g ∼ f ℓ and c ∼ e . Since the multiplication map that ∞ ⊗ n=1 gncn ∈ ∞Nn=1 (cid:16)[ℓ∈N [f ℓ]∼(cid:17) × [e]∼ → ∞On=1 Hn, (g, c) 7→ ∞ ⊗ n=1 gncn 13 is multilinear, it defines a bilinear map on Span(cid:16) Sℓ∈N Jf ℓK(cid:17) × JeK , denoted by (a, b) 7→ πe(a)b , thus ∗ -alg(cid:0)Jf K(cid:1) follows obtaining the formula for πe in the theorem. That πe is a representation of from the explicit formula, and the ∗ -property is also clear. It remains to show that each πe(a) is [e]Hn ). It suffices to check this for the generating bounded (hence extends as a bounded operator to ∞Nn=1 elements of ∗ -alg(cid:0)Jf K(cid:1) . Let a ∈ Jf K with a ∼ f : a = (a1 ⊗ · · · ⊗ ap) ⊗ fp+1 ⊗ fp+2 ⊗ · · · for some p < ∞ . Moreover any b ∈ JeK can also be written in the form: b = bp ⊗ ep+1 ⊗ ep+2 ⊗ · · · with bp ∈ pOj=1 Hj, where we may take the same p as in the preceding expression (e.g. by adjusting the initial part). Then kπe(a)bk = kApbpk · ∞Yk=p+1 kfkekk, where Apv = (a1 ⊗ · · · ⊗ ap)v. Since Ap is bounded on the completion cNj=1,...,p as kfkekk ≤ kekk = 1 , we see that Hj of pNj=1 Hj, we have kApbpk ≤ kApk · kbpk , and kπe(a)bk2 ≤ kApk2kbpk2 · ∞Yk=p+1 kekk2 = kApk2 · kbk2 and hence πe(a) is a bounded operator on JeK so extends to a bounded operator on [e]Hn . ∞Nn=1 4.2 Definition Thus for any f ∈ ∞Qj=1 Vkj , we can define Lµ[f ] := C∗(cid:0)πe(cid:0) ∗ -alg(cid:0)Jf K(cid:1)(cid:1)(cid:1) ⊂ B(cid:16) ∞On=1 4.3 Remark Recall that we also have the unitaries πµ(RN) ⊂ U(cid:0) ∞Nn=1 [e]Hn(cid:17) . [e]Hn(cid:1), where πµ(x) ∞ ⊗ k=1 ck = ∞ ⊗ n=1(cid:0) expxn·cn(cid:1) ∈ JeK, x ∈ R(N) , c ∼ e, expxn (t) := eixnt . Then πµ(x) · πe(cid:0)ι(g)(cid:1) = πe(cid:0)ι(g)(cid:1) · πµ(x) = πe(cid:16) ∞ n=1(cid:0) expxn·gn(cid:1)(cid:17) ∈ Lµ[f ], ⊗ for all x ∈ R(N), g ∼ f ℓ and ℓ ∈ N . The inclusion needed the fact that x has only finitely many nonzero entries, and that expxn·C0(R) ⊂ C0(R). Thus πµ(R(N)) · Lµ[f ] ⊂ Lµ[f ] . Since for each x ∈ R(N) we can find a sequence (cid:0)ι(gn)(cid:1)n∈Z ⊂ Jf K such that πe(cid:0)ι(gn)(cid:1) · πµ(x) converges in norm to πµ(x), we have a faithful embedding of R(N) as unitaries into the multiplier algebra M(cid:0)Lµ[f ](cid:1) denoted η : R(N) → M(cid:0)Lµ[f ](cid:1). In the next section we will investigate to what extent Lµ[f ] is a host algebra of R(N) . 14 4.4 Lemma With f as in Proposition 4.1(ii), we have (i) The C∗ -algebra Lµ[f ] is separable. (ii) Let ω be a pure state on Lµ[f ], and let eω be its strict extension to the unitaries η(R(N)) ⊂ M(cid:0)Lµ[f ](cid:1) . Then eω ◦ η is a character and there exists an element a ∈ RN with eω(η(x)) = exp(cid:0)ihx, ai(cid:1) for all x ∈ R(N) . Proof. (i) Since πe(cid:0) ∗ -alg(cid:0)Jf K(cid:1)(cid:1) is dense in Lµ[f ] , where ∗ -alg(cid:0)Jf K(cid:1) = Span(cid:8)Jf kK k ∈ N(cid:9) C0(R)(cid:17) ⊗ f k m+2 ⊗ · · ·o, ∞[m=1n(cid:16) mNℓ=1 m+1 ⊗ f k Jf kK = and (i) follows immediately from the separability of C0(R) . (ii) As Lµ[f ] is commutative, any pure state ω of it is a point evaluation, hence a ∗ -homomorphism. continuous (since it is determined by the factor Thus the strict extension eω to η(R(N)) ⊂ M(cid:0)Lµ[f ](cid:1) is also a ∗ -homomorphism, hence eω ◦ η is a character. The restriction of eω ◦ η to the subgroup Rn ⊂ R(N) Rn ) hence of the form eω ◦ η(x) = exp(ix · a(n)) for some a(n) ∈ Rn . Since eω ◦ η is a character on n ∈ N . Then eω ◦ η(x) = exp(cid:0)ihx, ai(cid:1) since for any x ∈ Rn ⊂ R(N) this restricts to the previous formula for eω ◦ η . all of R(N), the family {a(n) ∈ Rn n ∈ N} is a consistent family, i.e., if n < m then a(n) first n entries of a(m). Thus there is an a ∈ RN such that a(n) is still a character, and it is C0(R) in Lµ[f ] which is the group algebra of is the first n entries of a for any n ⊗ j=1 is the Since Lµ[f ] is separable and commutative, it follows from Theorem II.2.2 in [Da96] that all its cyclic representations are multiplicity free, and hence by Theorem 4.9.4 in [Ped89], for any state ω on Lµ[f ] , there is a regular Borel probability measure ν on the states S(Lµ[f ]) concentrated on the pure states Sp(Lµ[f ]) such that ω(A) =ZSp(Lµ[f ]) ϕ(A) dν(ϕ) ∀ A ∈ Lµ[f ] . (4) We will show that this decomposition produces similar decompositions to the one in (2) for other continuous positive definite functions than ωµ . Since Lµ[f ] is separable, it has a countable approximate identity {En}n∈N ⊂ Lµ[f ] (cf. Re- mark 3.1.1 [Mu90]). For a state ω on Lµ[f ], let eω be its strict extension to the unitaries η(R(N)) ⊂ M(cid:0)Lµ[f ](cid:1) , then we have for any countable approximate identity {En}n∈N ⊂ Lµ[f ] that where we used the Lebesgue dominated convergence theorem in the second line, since (cid:12)(cid:12)ϕ(η(x)En)(cid:12)(cid:12) ≤ 1 and the constant function 1 is integrable. By Lemma 4.4(ii) we can define a map ξ : Sp(Lµ[f ]) → RN by eϕ ◦ η(x) = exp(cid:0)ihx, ξ(ϕ)i(cid:1) 15 for x ∈ R(N) , eω ◦ η(x) = lim n→∞ ω(η(x)En) = lim = ZSp(Lµ[f ]) lim n→∞ ϕ(η(x)En) dν(ϕ) n→∞ZSp(Lµ[f ]) ϕ(η(x)En) dν(ϕ) =ZSp(Lµ[f ]) eϕ ◦ η(x) dν(ϕ) so using ξ we define a probability measure eν on RN by eν := ξ∗ν, and so: for x ∈ R(N) , eω ◦ η(x) =ZRN exp(cid:0)ihx, yi(cid:1) deν(y) which generalises the integral representation (2) to those positive definite functions eω which are strict extensions of states of Lµ[f ] (and this includes ωµ). We will obtain the full Bochner -- Minlos theorem for R(N) in a C∗ -algebraic context, if we can show that every continuous normalized positive definite function is of this type for some µ and some f . This is what we will do in the next section. (5) 5 Semi-host algebras for R(N) Inspired by the good properties which we found for Lµ[f ] above, we now examine more general versions of these algebras. The semi-host algebras which we obtain will be the building blocks for the algebra hosting the full representation theory of R(N), which will be constructed in the next section. For the rest of this section we fix a sequence (kn)n∈N ∈ NN and f ∈ Then we have that ∞Qn=1 Vkn such that Jf K 6= 0 . ∗-alg(cid:0)Jf K(cid:1) = Span(cid:8)Jf kK k ∈ N(cid:9) = lim := Span(cid:8)A1 ⊗ · · · ⊗ Am ⊗ f k Am[f ] Am[f ], −→ m+1 ⊗ f k where (6) m+2 ⊗ · · · Ai ∈ C0(R) ∀ i ∈ N, k ∈ N(cid:9) and the inductive limit is w.r.t. set inclusion of the ∗ -algebras Am[f ] ⊂ Aℓ[f ] Associativity Theorem 3.3, we can write if m < ℓ . By the Am[f ] =(cid:16) mNk=1 C0(R)(cid:17) ⊗(cid:16) ∗ -alg(cid:0) ∞Nj=m+1 fj(cid:1)(cid:17). The natural C∗ -norm on on the first factor is clear, but not on the second factor. So we next investigate possible bounded ∗ -representations to provide ∗ -algebra is given by specifying the single operator π(E) . Since E is positive, we require π(E) ≥ 0 , fj(cid:17) is generated by the single element E := ∗ -alg(cid:16) ∞Nj=m+1 and as we want a tensor norm on the larger ∗ -alg(cid:0)Jf K(cid:1), we need that kπ(E)k ≤ 5.1 Lemma Let f ∈ Qn∈N Vkn and let (cid:8)πk : C0(R) → B(H) k ∈ N(cid:9) be a set of ∗ -representations ∗ -alg(cid:0)Jf K(cid:1) with a C∗ -- norm. Since ∞Nj=m+1 fj , any representation π of this ∞Qj=m+1 kfjk = 1 . on the same space with commuting ranges. Then (i) The strong limit F (ℓ) k := s-lim n→∞ πk(f ℓ k) · · · πn(f ℓ n) ∈ B(H) exists, and 0 ≤ F (ℓ) k ≤ 1 for k, ℓ ∈ N . := s-lim (ii) P [f ] k→∞ F (ℓ) k P [f ] = F (ℓ) k F (ℓ) k . (an increasing limit) is a projection independent of ℓ ∈ N satisfying (iii) Let Q ∈ B(H) be such that 0 ≤ Q ≤ 1 , and such that it commutes with πk(C0(R)) for each k ∈ N. Let A := A1 ⊗ · · · ⊗ Am ⊗ f ℓ m+1 ⊗ f ℓ m+2 ⊗ · · · ∈ ∗ -alg(cid:0)Jf K(cid:1) and define πQ(A) := π1(A1) π2(A2) · · · πm(Am) F (ℓ) m+1Qℓ . Then πQ defines a ∗ -representation πQ : ∗ -alg(cid:0)Jf K(cid:1) → B(H) . 16 F (ℓ) k ≤ 1 . it follows from kCnk = (cid:13)(cid:13)πk(f ℓ operators (cid:0)F (ℓ) (ii) By definition, F (ℓ) (iv) The representation πQ is non-degenerate if and only if all πi are non-degenerate, P [f ] = 1 and Ker Q = {0} . If πQ is degenerate, Ker Q = 0, and all πj are non-degenerate, then P [f ] is the projection onto the essential subspace of πQ. (i) Since the operators πk(f ℓ k), πj(f ℓ Proof. joint spectral theory that their product πk(f ℓ for all k, ℓ ∈ N, we derive that πk(f ℓ Cn := πk(f ℓ Thm 4.1.1, p. 113], Cn converges in the strong operator topology to some limit F (ℓ) F (ℓ) k j ) ∈ B(H) commute and are positive, it follows from k) · πj(f ℓ k) ≤ 1 k) and hence, for a fixed k , the operators n) form a decreasing sequence of commuting positive operators. Thus, by [Mu90, . It is clear that j ) is also a positive operator. From πk(f ℓ is positive, and using k) · · · πn(f ℓ j ) ≤ πk(f ℓ k) · πj(f ℓ k kT k = sup(cid:8)(cid:12)(cid:12)(ψ, T ψ)(cid:12)(cid:12) ψ ∈ H, kψk = 1(cid:9) k) · · · πn(f ℓ n)(cid:13)(cid:13) ≤ 1 for all n that kF (ℓ) whenever T = T ∗ , k k ≤ 1 and hence that 0 ≤ k = πk(f ℓ k)F (ℓ) k+1 and 0 ≤ πk(f ℓ k) ≤ 1 and so the commuting sequence of k (cid:1)k∈N is increasing, and bounded above by 1 . Thus it follows again from Theorem 4.1.1 exists, is positive and bounded above by 1 . Since in [Mu90] that the strong limit P (ℓ)[f ] := s-lim k→∞ the operator product is jointly strong operator continuous on bounded sets, we get F (ℓ) k F (ℓ) k P (ℓ)[f ] = s-lim n→∞ = s-lim n→∞ πk(f ℓ k) · · · πn−1(f ℓ n−1) · s-lim n→∞ F (ℓ) n πk(f ℓ k) · · · πn−1(f ℓ n−1) F (ℓ) n = s-lim n→∞ F (ℓ) k = F (ℓ) k . Thus by P (ℓ)[f ] = s-lim k→∞ F (ℓ) k = s-lim k→∞ conclude that it is a projection. To see that P (ℓ)[f ] have: F (ℓ) k P (ℓ)[f ] =(cid:0)P (ℓ)[f ](cid:1)2 and the fact that P (ℓ)[f ] is positive we is independent of ℓ , note that for k ≤ m we F (ℓ) k F (j) m = s-lim n→∞ = s-lim n→∞ πk(f ℓ k) · · · πn(f ℓ n) · s-lim p→∞ πm(f j m) · · · πp(f j p ) πk(f ℓ k) · · · πm−1(f ℓ m−1) πm(f ℓ+j m ) · · · πn(f ℓ+j n ) = πk(f ℓ k) · · · πm−1(f ℓ m−1) F (ℓ+j) m . (7) This leads to P (ℓ)[f ] · P (j)[f ] = s-lim k→∞ F (ℓ) k s-lim m→∞ F (j) m = s-lim n→∞ n F (j) F (ℓ) n = s-lim n→∞ F (ℓ+j) n = P (ℓ+j)[f ]. is idempotent, i.e., P (ℓ)[f ] = P (2ℓ)[f ] for all ℓ ∈ N , hence P (ℓ)[f ] is inde- However, each P (ℓ)[f ] pendent of ℓ . (iii) Since ∗ -alg(cid:0)Jf K(cid:1) = lim −→ Am[f ] = Sm∈N representation on each ∗ -algebra Am[f ] , and that πQ restricts to its correct values on any Ak[f ] ⊂ Am[f ] for k < m . Recall that Am[f ] , it suffices to show that πQ defines a ∗ - Am[f ] =(cid:16) mNk=0 C0(R)(cid:17) ⊗(cid:16) ∗ -alg(cid:0) ∞Nj=m+1 fj(cid:1)(cid:17) . Now π(m) a : mOk=0 C0(R) → B(H), π(m) a (A1 ⊗ · · · ⊗ Am) := π1(A1) · · · πm(Am) 17 is a well-defined ∗ -representation obtained by the universal property of the tensor product. Moreover, since ∗ -alg(cid:16) ∞Nj=m+1 assignment π(m) b (cid:16) ∞Nj=m+1 fj(cid:17) is generated by a single element not satisfying any polynomial relation, the fj(cid:17) → m+1Q ≥ 0 defines a ∗ -representation π(m) fj(cid:17) := F (1) b : ∗ -alg(cid:16) ∞Nj=m+1 fj(cid:1)ℓ(cid:17) . b (cid:16)(cid:0) ∞Nj=m+1 B(H) . Note from Equation (7) that F (k) m+1 · F (ℓ) m+1 = F (k+ℓ) m+1 , which leads to the factorization πQ(cid:0)A1 ⊗ · · · ⊗ Am ⊗ f ℓ m+1 ⊗ f ℓ m+2 ⊗ · · ·(cid:1) = π(m) a (A1 ⊗ · · · ⊗ Am) · π(m) Thus, since it is multilinear, we obtain a linear map πQ on Am[f ] , and as the ranges of the ∗ - representations πa and πb commute, πQ is a ∗ -representation on Am[f ] . For k < m we have from the definition that π(k) b (cid:16) ∞Oj=k+1 and hence fj(cid:17) = F (1) k+1Q = πk+1(fk+1) · · · πm(fm)F (1) m+1Q π(m) a b (cid:16) ∞Nj=m+1 (A1 ⊗ · · · ⊗ Ak ⊗ fk+1 ⊗ · · · fm) · π(m) so it is clear that the value of πQ on Ak[f ] ⊂ Am[f ] b (cid:16) ∞Nj=k+1 defined on Am[f ]. Hence πQ is consistently defined as a ∗ -representation of ∗ -alg(cid:0)Jf K(cid:1) . , we have πQ(A)P [f ] = πQ(A) for all A ∈ ∗ -alg(cid:0)Jf K(cid:1) , hence, if P [f ] 6= 1 , then πQ(cid:0)*-alg(cid:0)Jf K(cid:1)(cid:1) has null spaces, i.e., πQ is degenerate. Likewise, if Ker Q 6= {0} is the same as the restriction of the map πQ a (A1 ⊗ · · · ⊗ Ak) · π(k) fj(cid:17) = π(k) (iv) Note that by F (ℓ) k P [f ] = F (ℓ) fj(cid:17) is degenerate, then since by commutativity: then πQ is degenerate. Moreover, if any πi k πQ(cid:0)A1 ⊗ · · · ⊗ Am ⊗ f ℓ m+1 ⊗ f ℓ m+2 ⊗ · · ·(cid:1) = π1(A1) · · · \πi(Ai) · · · πm(Am) F (ℓ) m+1Qℓπi(Ai) , where the hat means omission, it follows that πQ is also degenerate. Conversely, let πQ be degenerate, i.e., there is a nonzero ψ ∈ H such that πQ(A)ψ = 0 for all A , hence πQ(cid:0)A1 ⊗ · · · ⊗ Am ⊗ f ℓ m+1 ⊗ f ℓ m+2 ⊗ · · ·(cid:1)ψ = π1(A1) · · · πm(Am) F (ℓ) m+1Qℓψ = 0 for all Ai ∈ C0(R) and m, ℓ ∈ N . If all πj are non-degenerate, then it follows inductively that F (ℓ) m Qℓψ = 0 for all m and ℓ . If Ker Q = 0 , then F (ℓ) m ψ = 0 for all m , hence P [f ]ψ = 0 , i.e., P [f ] 6= 1 . non-degenerate, then P [f ] By the last step we also see that when πQ is degenerate, Ker Q = 0, and all πj are by (ii) it is zero on the null space of πQ. Since F (ℓ) follows from the definition of πQ that πQ(A)P [f ] = πQ(A) for all A ∈ *-alg(cid:0)Jf K(cid:1) . Thus P [f ] k P [f ] = F (ℓ) i.e. it is the projection onto this essential subspace. is the identity on the essential subspace of πQ, k 5.2 Definition Using this lemma, we can now investigate natural representations of ∗ -alg(cid:0)Jf K(cid:1) . Start with the universal representation of R(N) denoted πu : R(N) → U(Hu) which we recall, is the direct sum of the cyclic strong -- operator continuous unitary representations of R(N), one from each unitary equivalence class. Since for the kth component we have an inclusion R ⊂ R(N) by x −→ 18 (0, . . . , 0, x, 0, 0, . . .) ( kth entry), πu restricts to a representation on the kth component, denoted u : R → U(Hu) . By the host algebra property of C∗(R) ∼= C0(R), this produces a unique by πk representation πk u : C0(R) → B(Hu) , which is non-degenerate. Since the set of representations (cid:8)πk u : C0(R) → B(Hu) k ∈ N(cid:9) have commuting ranges, we can apply Lemma 5.1, with Q = 1 , to define a representation πu : ∗ -alg(cid:0)Jf K(cid:1) → B(Hu) by an abuse of notation. Below we will use the notation F (ℓ) u,k for the operator F (ℓ) k of πu. 5.3 Definition The C∗ -algebra L[f ] is the C∗ -completion of πu(cid:0) ∗ -alg(cid:0)Jf K(cid:1)(cid:1) in B(Hu) . 5.4 Remark (1) We see directly from (6) and the separability of C0(R) that L[f ] is separable. (2) Observe that the representation πu of ∗-alg(Jf K) may be degenerate. Although all πu non-degenerate, it is possible that P [f ] 6= 1 . By Lemma 5.1(iv) it then follows that P [f ] projection onto the essential subspace of πu. k are is the (3) Since L[f ] ⊂ B(Hu) is given as a concrete C*-algebra, this selects the class of those represen- tations of L[f ] which are normal maps w.r.t. the σ -- strong topology of B(Hu) on L[f ]. We will say that such a representation π is normal w.r.t. the defining representation πu. This will be the case if the vector states of π (L[f ]) are normal states for πu (L[f ]) (cf. Proposi- tion 7.1.15 [KR86]). (4) From Fell's Theorem [Fe60, Thm. 1.2] we know that any state of L[f ] is in the weak-*-closure of the convex hull of the vector states of πu . We will need the following proposition. 5.5 Proposition If S ⊂ N is a finite subset, then (i) there is a C∗ -algebra BS[f ] ⊂ B(Hu) and a copy of the C*-complete tensor product LS := C0(R) in B(Hu) such that cNs∈S L[f ] = C∗(cid:0)LS · BS[f ](cid:1) ∼= LSb⊗BS[f ] . (ii) the natural embeddings ζS : M (LS) → M(cid:0)L[f ](cid:1) = M(cid:0)LSb⊗BS[f ](cid:1) by ζS(M )(A ⊗ B) := (M · A) ⊗ B for all A ∈ LS and B ∈ BS[f ] are topological embeddings w.r.t. the strict topology on each bounded subset of M (LS) . Moreover, LS is dense in M (LS) w.r.t. the relative strict topology of M(cid:0)L[f ](cid:1) . (iii) The group homomorphism η : R(N) → M(cid:0)L[f ](cid:1) is strictly continuous. Proof. (i) By associativity (Theorem 3.3): C0(R)(cid:17) , and so, applying this to ∗ -alg(cid:0)Jf K(cid:1), and using the fact that it is the span of elementary tensors of the type C0(R)(cid:17) ⊗ (cid:16) Nt∈N\S C0(R) = (cid:16) Ns∈S ∞Nk=1 A1 ⊗ · · · ⊗ Am ⊗ f ℓ m+1 ⊗ f ℓ m+2 ⊗ · · · with Ai ∈ C0(R) and m, ℓ ∈ N, we get ∗ -alg(cid:0)Jf K(cid:1) =(cid:16)Ns∈S C0(R)(cid:17) ⊗(cid:16) ∗ -alg(cid:0)JfN\SK(cid:1)(cid:17) , 19 where (fN\S)t = ft for t ∈ N\S and ∗ -alg(cid:0)JfN\SK(cid:1) denotes the ∗ -algebra generated in Nt∈N\S C0(R) by n ⊗ t∈N\S gt(cid:12)(cid:12)(cid:12) g ∈ Yt∈N\S C0(R), g ∼ fN\So . Below, we need unital algebras, so adjoin identities, and define which contains essential space Hess ⊂ Hu, πu : C0 → B(Hu) , where ∞Nk=1(cid:0)C1 + C0(R)(cid:1) C0(R)(cid:17) ⊗(cid:16)C1 + ∗ -alg(cid:0)JfN\SK(cid:1)(cid:17) ⊂ C0 :=(cid:16)C1 + Ns∈S ∗ -alg(cid:0)Jf K(cid:1) as a ∗ -ideal. Since πu(cid:0) ∗ -alg(cid:0)Jf K(cid:1)(cid:1) acts non-degenerately on its ess. Define C := C∗(cid:0)πu(C0)(cid:1) = C∗(cid:0)A · B(cid:1) B := C∗(cid:16)πu(cid:16)1 ⊗(cid:0)C1 + ∗ -alg(cid:0)JfN\SK(cid:1)(cid:17)(cid:17) . if we let the null space of πu be H⊥ it determines a unique extension of πu to a representation and A := C∗(cid:16)πu(cid:16)(cid:0)C1 + Ns∈S C0(R)(cid:1) ⊗ 1(cid:17)(cid:17) Thus the unital C∗ -algebra C is generated by the two commmuting unital C∗ -algebras A and B. Moreover, since πu contains tensor representations (w.r.t. the two factors of it follows that if AB = 0 for an A ∈ A and a B ∈ B, then either A = 0 or B = 0 . Thus ∗ -alg(cid:0)Jf K(cid:1) above), by [Tak79, Ex. 2, p. 220], it follows that C ∼= Ab⊗B , where the tensor C∗ -norm is unique, since both A and B are commutative, hence nuclear. We conclude that the original C∗ -norm defined on C is in fact a cross -- norm. Since its restriction to is still a cross -- norm, and the latter is unique by commutativity of the algebras (given the norms on the factors), it follows from C∗(cid:2)πu(cid:0) Ns∈S ∗ -alg(cid:0)Jf K(cid:1) =(cid:16)Ns∈S C0(R)(cid:17) ⊗(cid:16) ∗ -alg(cid:0)JfN\SK(cid:1)(cid:17) ⊂ C0 C0(R)(cid:1)(cid:3) = cNs∈SC0(R) that L[f ] = (cid:16)Ns∈S C0(R)(cid:17) ⊗ C∗hπu(cid:16)1 ⊗(cid:0)∗-alg(cid:0)JfN\SK(cid:1)(cid:17)i = LSb⊗BS[f ] C0(R)(cid:1) ⊗ 1(cid:17) · πu(cid:16)1 ⊗(cid:0)∗-alg(cid:0)JfN\SK(cid:1)(cid:17)i, = C∗hπu(cid:16)(cid:0)Ns∈S where BS[f ] := C∗hπu(cid:16)1 ⊗(cid:0) ∗ -alg(cid:0)JfN\SK(cid:1)(cid:1)(cid:17)i . (ii) This follows from (i) and Lemma A.2 in [GrN09]. (iii) Since η(R(N)) consists of unitary multipliers, it suffices to verify that the set of all elements A ∈ L[f ] for which the map ηA : R(N) → L[f ], x → η(x)A is continuous span a dense subalgebra. To establish this, let A = ι(y) for some y ∼ f k for some k ∈ N . Now R(N) is a topological direct limit, so that it suffices to verify continuity on the finite dimensional subgroups Rn. For these, it follows from the strict continuity of the action of the group Rn on its C∗ -algebra C∗(Rn) ∼= C0(Rn) and the fact that by part (i) we have for a C∗ -algebra A , where Rn acts by unitary multipliers on the first tensor factor and the identity on the second factor. L[f ] ∼= C0(Rn)b⊗A, 20 Note that for S = {1, 2, . . . , n} , Rn ⊂ R(N) ⊂ U M (L[f ]) . Below we will abbreviate the notation to L(n) := L{1,2,...,n} = cNn the map ζS identifies Rn ⊂ U M (LS) with the unitaries k=1C0(R) . For ease of notation, we sometimes also omit explicit indication of the embeddings ζS, using inclusions instead. Next, let π : L[f ] → B(Hπ) be a given fixed non-degenerate ∗ -representation. Let eπ denote the strict extension of π to M (L[f ]) , so that πk := eπ ↾ L{k} and π(n) := eπ ↾ L(n) are the respectively. Then (cid:8)πk k ∈ N(cid:9) is a set of non-degenerate representations with commuting ranges ζ{k} ֒−−→ M (L[f ]) and L(n) ⊂ M (L(n)) strict extensions of π to L{k} ⊂ M (L{k}) ζ{1,...,n} ֒−−−−−→ M (L[f ]) as in Lemma 5.1, hence we specialize its notation to: F (ℓ) π,k := s-lim n→∞ πk(f ℓ k) · · · πn(f ℓ n) ∈ B(Hπ) and Pπ[f ] := s-lim k→∞ F (ℓ) π,k ∈ B(Hπ) . Since the commuting sequence of operators (cid:0)F (ℓ) π,k(cid:1)∞ a nonzero ψ ∈ Hπ such that F (ℓ) π,kψ = 0 for all k and ℓ . k=1 is increasing, Pπ[f ] 6= 1 implies that there is We will show in the next proposition that, for a certain choice of Q , there is a representation πQ constructed as in Lemma 5.1 from the set (cid:8)πk k ∈ N(cid:9) which coincides with π . 5.6 Proposition Fix a non-degenerate ∗ -representation π : L[f ] → B(Hπ) with Hπ 6= {0}. (i) Let Bn := eπ(cid:0) satisfies 0 < Q ≤ 1. n−1 factors z 1 ⊗ · · · ⊗ 1 ⊗fn ⊗ fn+1 ⊗ · · ·(cid:1) . Then the strong limit Q := s-lim } n→∞ { Bn exists and (ii) If A := A1 ⊗ · · · ⊗ Am ⊗ f ℓ m+1 ⊗ f ℓ m+2 ⊗ · · · ∈ ∗ -alg(cid:0)Jf K(cid:1), then π(A) = π1(A1) π2(A2) · · · πm(Am) F (ℓ) π,m+1 Qℓ = πQ(A), i.e., πQ = π ↾ ∗-alg(Jf K) . Moreover Pπ[f ] = 1 and Ker Q = {0} . (iii) Let π(n) : L(n) → B(Hπ) denote the strict extension of π to L(n) ⊆ M (L(Jf K)) . Then π(L1 ⊗ L2 ⊗ · · · ) = s-lim n→∞ π(n)(cid:0)L1 ⊗ L2 ⊗ · · · ⊗ Ln(cid:1)Qℓ for all elementary tensors L1 ⊗ L2 ⊗ · · · ∈ Jf ℓK ⊂ ∗ -alg(cid:0)Jf K(cid:1). Proof. (i) We need to prove this claim in greater generality than stated above, for use in the subsequent part. By definition, we have for A := A1 ⊗ · · · ⊗ Am ⊗ f ℓ m+1 ⊗ f ℓ m+2 ⊗ · · · ∈ ∗ -alg(cid:0)Jf K(cid:1), that πu(A) = π1 u(A1) π2 u(A2) · · · πm u (Am) F (ℓ) u,m+1 ∈ L[f ], where F (ℓ) that F (ℓ) u,k := s-lim n→∞ πk u(f ℓ k) · · · πn u (f ℓ u,n ∈ M(cid:0)L[f ](cid:1) . Thus the operator } B(ℓ) n :=eπ(cid:0) n) = eπu(cid:0)1 ⊗ · · · ⊗ 1 ⊗ f ℓ z n−1 factors 1 ⊗ · · · ⊗ 1 ⊗f ℓ n ⊗ f ℓ { satisfies 0 ≤ B(ℓ) operator commuting with B(ℓ) n ≤ 1 since 0 ≤ F (ℓ) u,n(cid:1) n+1 ⊗ · · ·(cid:1) =eπ(cid:0)F (ℓ) n ⊗ f ℓ n+1 ⊗ · · ·(cid:1) ∈ B(Hu) . Hence we have u,n ≤ 1 . As B(ℓ) n = πn(f ℓ n)B(ℓ) n+1 and πn(f ℓ n+1, we see that B(ℓ) n ≤ B(ℓ) n+1. Thus the strong limit Q(ℓ) := s-lim n→∞ n) ≤ 1 is a positive B(ℓ) n 21 exists by Theorem 4.1.1 in [Mu90], and satisfies 0 < Q(ℓ) ≤ 1 (note that Q(ℓ) 6= 0 since π is non-degenerate and Hπ 6= {0}). Since the operator product is jointly strongly continuous on bounded sets we have: Q(ℓ)Q(m) = s-lim = s-lim = s-lim n ⊗ f ℓ n ⊗ f ℓ n→∞eπ(cid:0)1 ⊗ · · · ⊗ 1 ⊗ f ℓ n→∞eπ(cid:0)1 ⊗ · · · ⊗ 1 ⊗ f ℓ n→∞eπ(cid:0)1 ⊗ · · · ⊗ 1 ⊗ f ℓ+m z n−1 factors 1 ⊗ · · · ⊗ 1 ⊗f ℓ n ⊗ f ℓ { Thus Q(ℓ) = Qℓ where Q := Q(1) . (ii) Now B(ℓ) k→∞eπ(cid:0)1 ⊗ · · · ⊗ 1 ⊗ f m n+1 ⊗ · · ·(cid:1) s-lim n+1 ⊗ · · ·(cid:1)eπ(cid:0)1 ⊗ · · · ⊗ 1 ⊗ f m n+1 ⊗ · · ·(cid:1) = Q(ℓ+m) . n ⊗ f m k ⊗ f m k+1 ⊗ · · ·(cid:1) n+1 ⊗ · · ·(cid:1) n ⊗ f ℓ+m n = eπ(cid:0) = eπ(cid:0)1 ⊗ · · · ⊗ 1 ⊗ f ℓ = πn(f ℓ n+1 ⊗ · · ·(cid:1) } n ⊗ 1 ⊗ · · ·(cid:1) ·eπ(cid:0)1 ⊗ · · · ⊗ 1 ⊗ f ℓ n)eπ(cid:0)1 ⊗ · · · ⊗ 1 ⊗ f ℓ z k)eπ(cid:0) m→∞eπ(cid:0)1 ⊗ · · · ⊗ 1 ⊗ f ℓ n+2 ⊗ · · ·(cid:1) { 1 ⊗ · · · ⊗ 1 ⊗f ℓ n) · · · πk(f ℓ n) · · · πk(f ℓ n+1 ⊗ f ℓ k+1 ⊗ f ℓ = s-lim k→∞ πn(f ℓ = s-lim k→∞ π,nQ(ℓ) = F (ℓ) = F (ℓ) k) s-lim π,nQℓ πn(f ℓ } k factors n+1 ⊗ f ℓ n+2 ⊗ · · ·(cid:1) k+2 ⊗ · · ·(cid:1) m+1 ⊗ f ℓ m+2 ⊗ · · ·(cid:1) (8) (9) where we used again the joint strong operator continuity of the product on bounded sets. Let A := A1 ⊗ · · · ⊗ Am ⊗ f ℓ m+1 ⊗ f ℓ m+2 ⊗ · · · ∈ ∗ -alg(cid:0)Jf K(cid:1). Then π(A) = π1(A1) ·eπ(cid:0)1 ⊗ A2 ⊗ A3 ⊗ · · · ⊗ Am ⊗ f ℓ = π1(A1) π2(A2) · · · πm(Am) ·eπ(cid:0)1 ⊗ · · · ⊗ 1 ⊗ f ℓ = π1(A1) π2(A2) · · · πm(Am) · F (ℓ) π,m+1Qℓ = πQ(A) m+1 ⊗ f ℓ m+2 ⊗ · · ·(cid:1) = · · · m+2 ⊗ · · ·(cid:1) m+1 ⊗ f ℓ making use of (8) above. Since π is non-degenerate, it follows from Lemma 5.1(iii) that Pπ[f ] = 1 and Ker Q = {0} . (iii) Note first that from Proposition 5.5(ii) above and [Tak79, Lemma 4.1 on p.203] that π1(A1) π2(A2) · · · πn(An) = π(n)(A1 ⊗ · · · ⊗ An) for all Ai ∈ C0(R) . Thus, if we continue equa- tion (9) above π(A) = π1(A1) π2(A2) · · · πm(Am) · F (ℓ) π,m+1Qℓ = π1(A1) π2(A2) · · · πm(Am) s-lim n→∞ πm+1(f ℓ m+1) · · · πn(f ℓ n) Qℓ = s-lim n→∞ which establishes the claim. π(n)(A1 ⊗ · · · ⊗ Am ⊗ f ℓ m+1 ⊗ · · · ⊗ f ℓ n) Qℓ 5.7 Definition Given a representation π of L[f ], we will call its associated operator Q its excess. This proposition creates a difficulty for the host algebra project, because by part (iii) we can see that to construct its representations, we need more information than what is contained in the representations 22 of R(N), i.e., we need the excess operators Q . It is therefore very important to establish whether there are representations πQ with Q 6= 1 (below we will see such πQ will not be normal w.r.t. πu ). f 5.8 Proposition Let representations on the same space with commuting ranges. Then for any positive operator Q ∈ B(H) be as before and let (cid:8)πk : C0(R) → B(H) k ∈ N(cid:9) be a set of ∗ - with Q ≤ 1 which commutes with the ranges of all πk, we have that πQ : ∗ -alg(cid:0)Jf K(cid:1) → B(H) extends to a ∗ -representation of L[f ] . Proof. We show first that σ (Fu,k) = [0, 1]. Let ω be a character of R(N). Then since it is a one -- dimensional subrepresentation of πu there is a vector ψω ∈ Hu such that (cid:0)ψω, πu(x)ψω(cid:1) = ω(x) u(h)ψω(cid:1) for all h ∈ L{k} = C0(R) is also a character, hence for all x ∈ R(N). Then ωk(h) = (cid:0)ψω, πk k ∈ R , and in fact we obtain all point evaluations of L{k} = C0(R) a point evaluation at a point xω this way. Thus Fω,k := s-lim n→∞ ωk(fk) · · · ωn(fn) = lim n→∞ fk(xω k ) · · · fn(xω n) = ∞Yn=k fn(xω n) ∈ [0, 1] , and as we can choose our ω, hence points xω k ∈ R arbitrarily, it is clear that we can find ω to set Fω,k equal to any value in [0, 1]. Since Fω,k := lim n→∞ ωk(fk) · · · ωn(fn) = eω(cid:0) z k−1 factors 1 ⊗ · · · ⊗ 1 ⊗f ℓ k ⊗ f ℓ } { k+1 ⊗ · · ·(cid:1) =(cid:0)ψω, Fu,kψω) for all k. defines a character on C∗(Fu,k) we see that σ (Fu,k) = [0, 1] . Since for {πk k ∈ N} and Q as in Next, note that in a diagonalization of Fu,k ≥ 0 we can write it as Fu,k(x) = x for x ∈ σ(Fu,k), the initial hypotheses we always have that 0 ≤ Fπ,kQ ≤ 1, it follows that σ(cid:0)Fπ,kQ(cid:1) ⊆ [0, 1] = σ (Fu,k) and hence kp(Fu,k)k = sup(cid:8)(cid:12)(cid:12)p(x)(cid:12)(cid:12) x ∈ σ(Fu,k)(cid:9) . From this it is immediate that σ (Fπ,kQ) ⊆ σ (Fu,k) implies kp(Fπ,kQ)k ≤(cid:13)(cid:13)p(cid:0)Fu,k(cid:1)(cid:13)(cid:13) for all polynomials p . Finally, recall that ∗ -alg(cid:0)Jf K(cid:1) = lim Am[f ] where −→ Am[f ] := Span(cid:8)A1 ⊗ · · · ⊗ Am ⊗ f k m+1 ⊗ f k m+2 ⊗ · · · Ai ∈ C0(R) ∀ i ∈ N, k ∈ N(cid:9) and the inductive limit is w.r.t. to the inclusion Am[f ] ⊂ Aℓ[f ] . Thus L[f ] is the inductive limit of the C∗ -closures Lm of πu(cid:0)Am[f ](cid:1) w.r.t. set inclusion. Since C0(R)(cid:17) ⊗(cid:16) ∗ -alg(cid:0) ∞Nj=m+1 Am[f ] =(cid:16) mNk=0 fj(cid:1)(cid:17) , and the norm of L[f ] Next we define (as in the proof of Lemma 5.1(iii)) two ∗ -representations π(m) is a product norm by Proposition 5.5(i), we have that Lm ∼= L(m)b⊗C∗(Fu,m+1) . C0(R) → B(H) a : and π(m) b fj(cid:17) → B(H) as follows. First, we have that : ∗ -alg(cid:16) ∞Nj=m+1 mOk=0 fj(cid:17) is generated by a single element not satisfying any polynomial relation, the (A1 ⊗ · · · ⊗ Am) := π1(A1) · · · πm(Am) C0(R) → B(H), π(m) a π(m) a : defines a well-defined ∗ -representation by the universal property of the tensor product. Moreover, since ∗ -alg(cid:16) ∞Nj=m+1 mNk=0 23 b (cid:16) ∞Nj=m+1 fj(cid:17) := F (1) m+1Q ≥ 0 defines a ∗ -representation π(m) b : ∗ -alg(cid:16) ∞Nj=m+1 fj(cid:17) → assignment π(m) B(H) . Note from Equation (7) that F (k) m+1 · F (ℓ) m+1 = F (k+ℓ) m+1 , which leads to the factorization πQ(cid:0)A1 ⊗ · · · ⊗ Am ⊗ f ℓ m+1 ⊗ f ℓ m+2 ⊗ · · ·(cid:1) = π(m) a (A1 ⊗ · · · ⊗ Am) · π(m) b (cid:16)(cid:0) ∞Nj=m+1 fj(cid:1)ℓ(cid:17) . Now π(m) a has a unique extension to L(m) , and as π(m) ∗ -alg(cid:16) ∞Nj=m+1 that (cid:13)(cid:13)π(m) fj(cid:17) =(cid:8)p(cid:0)Fu,k(cid:1) p a polynomial(cid:9) on which it is continuous by the fact proven above, (cid:0)p(Fu,k)(cid:1)(cid:13)(cid:13) = kp(Fπ,kQ)k ≤ (cid:13)(cid:13)p(cid:0)Fu,k(cid:1)(cid:13)(cid:13) . Thus it extends uniquely to C∗(Fu,m+1), hence ∗ -alg(cid:0)Jf K(cid:1) and is continuous on all Am ) it follows that πQ πQ has a unique continuous extension to Lm . Since πQ respects the inductive limit structure (since it does so on the dense subalgebra extends uniquely to a continuous ∗ -representation of L[f ] . b b is defined on the dense ∗ -algebra We conclude that there is an abundance of representations π of L[f ] with Q 6= 1. Having investigated the representations of L[f ], we next consider its host algebra properties. First label the unitary embedding η : R(N) → M(cid:0)L[f ](cid:1) where η(x1, . . . , xn, 0, 0, . . .)(L1 ⊗ L2 ⊗ · · · ) = η1(x1)L1 ⊗ · · · ⊗ ηn(xn)Ln ⊗ Ln+1 ⊗ Ln+2 ⊗ · · · = ζ{1,...,n}(x1, . . . , xn)(cid:0)L1 ⊗ L2 ⊗ · · ·(cid:1) for all (x1, . . . , xn) ∈ Rn, n ∈ N, Li ∈ L{i} = C0(R), and where ηi : R → M (C∗(R)) is the usual unitary embedding. Then the map η∗ : Rep(cid:0)L[f ], H(cid:1) → Rep(cid:0)R(N), H(cid:1) consists of the strict extension of (non-degenerate) representations of L[f ] to η(R(N)), i.e. η∗(π)(x) := s-lim α→∞ π(cid:0)η(x)Eα(cid:1) for x ∈ R(N) and any approximate identity {Eα}α∈Λ in L[f ]. Since L[f ] and R(N) are commutative, their irreducible representations are all one-dimensional, hence η∗ takes irreducible representations to irre- ducible representations. 5.9 Theorem Given the preceding notation, we have that (i) η : R(N) → M(cid:0)L[f ](cid:1) is continuous w.r.t. the strict topology of M(cid:0)L[f ](cid:1) . (ii) Let Rep0(cid:0)L[f ], H(cid:1) denote those non-degenerate ∗ -representations of L[f ] with excess operators Q = 1 (cf. Proposition 5.6). Then η∗ is injective on Rep0(cid:0)L[f ], H(cid:1) . (iii) The range η∗(cid:0)Rep(cid:0)L[f ], H(cid:1)(cid:1) is the same as η∗(cid:0)Rep0(cid:0)L[f ], H(cid:1)(cid:1) and consists of those π ∈ Rep(cid:0)R(N), H(cid:1) such that 1 = s-lim representation in Rep(cid:0)L{k}, H(cid:1) such that η∗ πk(fk) · · · πn(fn) with πk the unique is the kth basis k→∞ eFk where eFk := s-lim k(πk) = π ↾ Rek , where ek ∈ R(N) n→∞ vector. (iv) For a state ω ∈ S(cid:0)L[f ](cid:1) , its GNS -- representation πω is in Rep0(cid:0)L[f ], Hω(cid:1) if and only if ω ∈ S0(cid:0)L[f ](cid:1) :=(cid:8)ϕ ∈ S(cid:0)L[f ](cid:1) lim n→∞ eϕ(cid:0) 24 n−1 factors z 1 ⊗ · · · ⊗ 1 ⊗fn ⊗ fn+1 ⊗ · · ·(cid:1) = 1(cid:9) . } { Moreover, the restriction η∗ : S0(cid:0)L[f ](cid:1) → S(cid:0)R(N)(cid:1) ≡ states of R(N), consisting of is injective, with range ω ∈ S(cid:0)R(N)(cid:1) such that lim k→∞ lim n→∞ (Ωω, πω k (fk) · · · πω n (fn)Ωω) = 1 with πω j as in (iii), and Ωω is the cyclic GNS -- vector. (v) π is normal w.r.t. the defining representation πu of L[f ] if and only if Q = 1 . Proof. (i) This is proven already in Proposition 5.5(iii). (ii) Let π ∈ Rep0(cid:0)L[f ], H(cid:1) and let eπ be its strict extension to M(cid:0)L[f ](cid:1). As eπ is strictly continuous, (i) implies that the unitary representation η∗(π) = eπ ◦ η : R(N) → U(H) operator continuous. We need to show that η∗ is injective on Rep0(cid:0)L[f ], H(cid:1). for two representations π, π′ ∈ Rep0(cid:0)L[f ], H(cid:1), then η∗ all n ∈ N. But Span(cid:0)η(n)(Rn)(cid:1) ⊂ M (L(n)) is strictly dense, and by Proposition 5.5(ii) this is still true for the strict topology of M (L[f ]) ⊃ ζ{1,...,n}(cid:0)M (L(n))(cid:1). Thus eπ ↾ ζ{1,...,n}(cid:0)L(n)(cid:1) = π(n) = eπ′ ↾ ζ{1,...,n}(cid:0)L(n)(cid:1), i.e., π and π′ produce the same representation π(n) : L(n) → B(H). Thus by is strong If η∗(π) = η∗(π′) for Proposition 5.6(iii) (using Q = 1 ) we find {1,...,n}(π′) on Rn ⊂ R(N) {1,...,n}(π) = η∗ π(L1 ⊗ L2 ⊗ · · · ) = s-lim n→∞ for the elementary tensors in ∗ -alg(cid:0)Jf K(cid:1) , π(n)(cid:0)L1 ⊗ L2 ⊗ · · · ⊗ Ln(cid:1) = π′(L1 ⊗ L2 ⊗ · · · ) i.e., π = π′. Thus η∗ is injective on Rep0(cid:0)L[f ], H(cid:1). (iii) To see that η∗(cid:0)Rep(cid:0)L[f ], H(cid:1)(cid:1) = η∗(cid:0)Rep0(cid:0)L[f ], H(cid:1)(cid:1) , note that for πQ as in Lemma 5.1: πQ(cid:0)η(x1, . . . , xn, 0, 0, . . .)(L1 ⊗ L2 ⊗ · · · )(cid:1) = πQ(cid:0)η1(x1)L1 ⊗ · · · ⊗ ηn(xn)Ln ⊗ Ln+1 ⊗ Ln+2 ⊗ · · ·(cid:1) by Prop. 5.6(iii) π(k)(cid:0)η1(x1)L1 ⊗ · · · ⊗ ηn(xn)Ln ⊗ Ln+1 ⊗ Ln+2 ⊗ · · · ⊗ Lk(cid:1)Qℓ π1(cid:0)η1(x1)L1(cid:1) · · · πn(cid:0)ηn(xn)Ln(cid:1)πn+1(cid:0)Ln+1(cid:1)πn+2(cid:0)Ln+2(cid:1) · · · πk(cid:0)Lk(cid:1)Qℓ = s-lim k→∞ = s-lim k→∞ 1π1(x1) · · · η∗ = η∗ nπn(xn) s-lim k→∞ π1(cid:0)L1(cid:1) · · · πk(cid:0)Lk(cid:1)Qℓ = η∗ 1π1(x1) · · · η∗ nπn(xn) πQ(L1 ⊗ L2 ⊗ · · · ) for L1 ⊗ L2 ⊗ · · · ∈ Jf ℓK ⊂ ∗ -alg(cid:0)Jf K(cid:1), which shows that η∗(πQ)(x1, . . . , xn, 0, 0, . . .) = η∗(π1)(x1, . . . , xn, 0, 0, . . .), and establishes the claim. To characterize the range of η∗, have from Lemma 5.1 that let π ∈ Rep0(cid:0)L[f ], H(cid:1) and note that as it is non-degenerate, we 1 = Pπ[f ] := s-lim k→∞ F (ℓ) π,k where F (ℓ) π,k := s-lim n→∞ πk(f ℓ k) · · · πn(f ℓ n) ∈ B(Hπ) , and πk = eπ ↾ L{k}. From the uniqueness of the strict extension eπ on M (L[f ]) and the fact that the strict topology of M (L{k}) ⊂ M (L[f ]) coincides with that of M (L[f ]) on bounded subsets, we see that η∗ k(πk) = η∗π ↾ Rek and hence eFk = F (1) k→∞ eFk . let π ∈ Rep(cid:0)R(N), Hπ(cid:1) be such that 1 = s-lim k→∞ eFk . We want to define πL ∈ Rep0(cid:0)L[f ], Hπ(cid:1) such that η∗(πL) = π . Consider first the case that π is cyclic. Recall that L[f ] π,k. Thus 1 = s-lim Conversely, 25 is the norm closure of πu(cid:0) ∗ -alg(cid:0)Jf K(cid:1)(cid:1) . By definition of πu, Hπ is a direct summand of Hu such that π(x) = Pππu(x) ↾ Hπ for all x ∈ R(N). and there is a projection Pπ ∈ πu(R(N))′ Then πk(A) = Pππk u,k ↾ Hπ . We define πL : L[f ] → B(Hπ) by πL(A) := Pππu(A) ↾ Hπ which is obviously a ∗ -representation, satisfy- u(A) ↾ Hπ for all A ∈ L{k}, and hence eFk = PπF (1) ing FπL,k = eFk, with excess 1 (as it is normal w.r.t. πu ), and as k→∞ eFk = 1 PπL [f ] = s-lim k→∞ FπL,k = s-lim by hypothesis, πL is non-degenerate. Next, relax the requirement that π be cyclic. Then π is a direct sum of cyclic representations. Let (πc, Hc) be a cyclic subrepresentation of π, and denote the projection onto Hc by Pc. Since π ↾ Rek also preserves Hc , it follows that πc u(A) ↾ Hc for all A ∈ L{k}. Now, recalling that 1 = s-lim πk(fk) · · · πn(fn), we have that k→∞ eFk where eFk := s-lim k(A) = Pcπk n→∞ 1Hc = Pc ↾ Hc = Pc s-lim s-lim n→∞ Pcπk(fk) · · · πn(fn) ↾ Hc k→∞ eFk ↾ Hc = s-lim k→∞ = s-lim k→∞ s-lim n→∞ πc k(fk) · · · πc n(fn) = s-lim k→∞ eF c k k(fk) · · · πc πc k := s-lim n→∞ representation πc πL : L[f ] → B(Hπ) as the direct sum of all the πc where eF c we have that πL ∈ Rep0(cid:0)L[f ], Hπ(cid:1). L : L[f ] → B(Hc) by πc n(fn). Thus, by the previous part we can construct a nondegenerate L(A) := Pπc πu(A) ↾ Hπc which is normal w.r.t. πu. Define L. Since this is normal w.r.t. πu and nondegenerate, Since the strict extension of πL produces the same representations πk on L{k} than obtained from π ↾ Rek , the strict extension of πL must coincide on R(N) with π, (iv) It is immediate from the definitions that if πω ∈ Rep0(cid:0)L[f ], Hω(cid:1) , is commutative, we know L[f ] ∼= C0(X), with X its spectrum. Then there is a probability measure µ on X and a unitary U : Hω → L2(X, µ) such Conversely, let ω ∈ S0(cid:0)L[f ](cid:1). Then, as L[f ] that (U πω(h)ψ)(x) = h(x)(cid:0)U ψ(cid:1)(x) for all h ∈ C0(X), ψ ∈ Hω, x ∈ X , and moreover U Ωω = 1 . then ω ∈ S0(cid:0)L[f ](cid:1) . i.e. η∗(πL) = π . Then n−1 factors z 1 ⊗ · · · ⊗ 1 ⊗fn ⊗ fn+1 ⊗ · · ·(cid:1) = (Ωω, QΩω) } { and as 0 < Q ≤ 1 we have: 1 = lim n→∞eω(cid:0) = ZX 0 = ZX(cid:12)(cid:12)1 − (U QU −1)(x)(cid:12)(cid:12) dµ(x). (U QU −1)(x) dµ(x) i.e., Q = 1 and thus πω ∈ Rep0(cid:0)L[f ], Hω(cid:1) . Hence (U QU −1)(x) = 1 µ -- a.e., The last part of the claim now follows from this, (iii), and the observation that η∗ω(g) = (Ωω, η∗πω(g)Ωω) for all g ∈ R(N). Note that the state condition on the range of η∗ implies the operator condition in (iii) by a similar argument than the one above for Q . (v) Let π be normal w.r.t. πu(L[f ]) . Then it is continuous on bounded sets w.r.t. the strong operator topologies of both sides, hence Q = s-lim n→∞ F (1) u,n is the projection onto the essential subspace of is the identity for πu(L[f ]), is in fact defined in πu, by Lemma 5.1(iv) we have that Pu[f ] = s-lim n→∞ πu(L[f ]). Thus, since L[f ] it follows that Pu[f ] n = s-lim n→∞eπ(cid:0)F (1) u,n(cid:1) = eπ(cid:0) s-lim u,n(cid:1) . However, B(1) F (1) n→∞ hence Q =eπ(cid:0) s-lim n→∞ F (1) u,n(cid:1) = 1 . 26 Conversely, let Q = 1, then by part (iii) η∗π is a continuous representation of R(N), and by Proposition 5.6(iii) (with Q = 1) we have that π(L1 ⊗ L2 ⊗ · · · ) = s-lim n→∞ π(n)(cid:0)L1 ⊗ L2 ⊗ · · · ⊗ Ln(cid:1) = s-lim n→∞ π1(L1)π2(L2) · · · πn(Ln) for all elementary tensors L1 ⊗ L2 ⊗ · · · ∈ Jf ℓK ⊂ ∗ -alg(cid:0)Jf K(cid:1). This is precisely the formula ∗ -alg(cid:0)Jf K(cid:1) which we used to define πu. Now in which Lemma 5.1 defined representations on πu(R(N))′′ = {π(n) cyclic components of π are contained in the direct summands of πu, ϕ : πu(L[f ])′′ → B(Hπ) such that ϕ ◦ πu = π. Thus π is normal to πu. u (L(n)) n ∈ N}′′ = πu(L[f ])′′ and a similar equation holds for π. Since the there is a normal map Thus, though L[f ] is not actually a host algebra for R(N), it does have good properties, e.g., η∗ is bijective between two large sets of representations, and it takes irreducible representations to irreducibles. In fact, using the algebras L[f ] , we can now give a full C∗ -algebraic interpretation of the Bochner -- Minlos Theorem. Our aim is not to re -- prove the Bochner -- Minlos Theorem in the C*- context, but just to identify the measures and decompositions of it with the appropriate measures and decompositions arising from the current C* -- context. First, we transcribe Lemma 4.4 for the current context: 5.10 Lemma As before, let f ∈ ∞Qn=1 Vkn such that Jf K 6= 0 . Let ω be a pure state on L[f ], and let Proof. As L[f ] is determined by the factor is commutative, any pure state ω of it is a ∗ -homomorphism. Thus the strict eω be its strict extension to the unitaries η(R(N)) ⊂ M(cid:0)L[f ](cid:1) . Then eω ◦ η is a character and there exists an element a ∈ RN with eω(η(x)) = exp(cid:0)ihx, ai(cid:1) for all x ∈ R(N) . extension eω to η(R(N)) ⊂ M(cid:0)L[f ](cid:1) is also a ∗ -homomorphism, hence eω ◦ η is a character. The restriction of eω ◦ η to the subgroup Rn ⊂ R(N) form eω ◦ η(x) = exp(ix · a(n)) for some a(n) ∈ Rn . Since eω ◦ η is a character on all of R(N), the Then eω ◦ η(x) = exp(cid:0)ihx, ai(cid:1) since for any x ∈ Rn ⊂ R(N) this restricts to the previous formula for eω ◦ η . family {a(n) ∈ Rn n ∈ N} is a consistent family, i.e., if n < m then a(n) of a(m). Thus there is an a ∈ RN such that a(n) is still a character, and it is continuous (since it C0(R) in L[f ] which is the group algebra of Rn ) hence of the is the first n entries of a for any n ∈ N . Thus there is a map from the pure states SP (L[f ]) to RN denoted by is the first n entries n ⊗ j=1 ξ : SP (L[f ]) → RN satisfying eϕ(η(x)) = exp(cid:0)ihx, ξ(ϕ)i(cid:1) ∀ x ∈ R(N) , ϕ ∈ SP (L[f ]) . 5.11 Theorem For each state ω of R(N) there is an f ∈ Vkn where kn ∈ N and a unique state ω0 ∈ S0(cid:0)L[f ](cid:1) such that η∗(ω0) = ω . Then (i) there is a regular Borel probability measure ν on S(cid:0)L[f ](cid:1) concentrated on the pure states ∞Qn=1 SP(cid:0)L[f ](cid:1) such that ω0(A) =ZSP (L[f ]) ϕ(A) dν(ϕ) ∀ A ∈ L[f ] . 27 (ii) The probability measure eν on RN given by eν := ξ∗ν is (up to sets of measure zero) the Bochner -- Minlos measure for ω , i.e., ω(x) =ZRN exp(cid:0)ihx, yi(cid:1) deν(y) ∀ x ∈ R(N) . Proof. Fix an ω ∈ S(cid:0)R(N)(cid:1). Then by Theorem 5.9(iv) it suffices to show that there is an f ∈ (Ωω, πω k (fk) · · · πω n (fn)Ωω) = 1 . However, since there is an approximate Vkn such that lim k→∞ lim n→∞ ∞Qn=1 identity {En}n∈N of C0(R) in and we do this as follows. Since Vn , it is possible to choose an f satisfying this limit condition, πω k (En)Ωω = Ωω, choose for each n ∈ N an fn := Ekn such ∞ ∪ n=1 lim n→∞ that (cid:13)(cid:13)πω n (Ekn )Ωω − Ωω(cid:13)(cid:13) ≤ 1/n2 . Then for 1 < k < n we have: πω k (fk) · · · πω n (fn)Ωω − Ωω = πω k (fk) · · · πω Hence: k (fk) · · · πω + πω k (fk) · · · πω n−2(fn−2)(cid:0)πω n (fn)Ωω − Ωω(cid:13)(cid:13) ≤ (cid:13)(cid:13)πω from which we see that condition. lim k→∞ lim n→∞(cid:13)(cid:13)πω k (fk) · · · πω n−1(fn−1)(cid:0)πω n (fn) − 1(cid:1)Ωω 1 1 k2 k (fk) − 1(cid:1)Ωω . n−1(fn−1) − 1(cid:1)Ωω + · · · +(cid:0)πω 1 n2 + <Z n+1 n (fn)Ωω − Ωω(cid:13)(cid:13) = 0, and this implies the required limit (n − 1)2 + · · · + 1 x2 dx = k − 1 n + 1 k−1 − 1 1 (i) Since L[f ] is separable and commutative, it follows from Theorem II.2.2 in [Da96] that all its GNS -- representations are multiplicity free, and hence by Theorem 4.9.4 in [Ped89], for any state ω0 on L[f ] there is a regular Borel probability measure ν on S(cid:0)L[f ](cid:1) concentrated on the pure states SP(cid:0)L[f ](cid:1) such that ϕ(A) dν(ϕ) ∀ A ∈ L[f ] . ω0(A) =ZSP (L[f ]) (ii) For the state ω0 on L[f ], then we have for any countable approximate identity {En}n∈N ⊂ Lµ[f ] that let eω0 be its strict extension to the unitaries η(cid:0)R(N)(cid:1) ⊂ M(cid:0)L[f ](cid:1) , eω0 ◦ η(x) = lim n→∞ ω0(η(x)En) = lim ϕ(η(x)En) dν(ϕ) n→∞ZSp(L[f ]) = ZSp(L[f ]) lim n→∞ ϕ(η(x)En) dν(ϕ) =ZSp(L[f ]) eϕ ◦ η(x) dν(ϕ) 1 and the constant function 1 is integrable. where we used the Lebesgue Dominated Convergence Theorem in the second line, since (cid:12)(cid:12)ϕ(η(x)En)(cid:12)(cid:12) ≤ If we define a probability measure eν on RN by eν := ξ∗ν , where the map ξ : Sp(L[f ]) → RN given by eϕ ◦ η(x) = exp(cid:0)ihx, ξ(ϕ)i(cid:1) for x ∈ R(N) was exp(cid:0)ihx, yi(cid:1) deν(y) ∀ x ∈ R(N). ω(x) = eω0 ◦ η(x) =ZRN mentioned above, we obtain Hence eν coincides (up to sets of measure zero) with the usual Bochner -- Minlos measure on RN by uniqueness of the measure on RN producing this decomposition (cf. Lemma 7.13.5 in [Bo07]). 28 Thus we can interpret the Bochner -- Minlos Theorem as an expression of the pure state space decompositions of the C∗ -- algebras L[f ] . We will not consider the uniqueness of the measures in the decompositions of the Bochner -- Minlos Theorem, as that is easy to prove. To understand L[f ] at a more concrete level, we consider its spectrum X . Since L[f ] is commu- tative, we know L[f ] ∼= C0(X), and as each ω ∈ X is a character, we obtain from Propositions 5.6 and 5.8 that ω(L1 ⊗ L2 ⊗ · · · ) = lim n→∞ ω1(L1)ω2(L2) · · · ωn(Ln)qℓ is a character of L{i} = C0(R) hence a point evaluation ωi(f ) = f (xi) . Since ω is uniquely determined for all elementary tensors L1 ⊗ L2 ⊗ · · · ∈ Jf ℓK ⊂ ∗ -alg(cid:0)Jf K(cid:1) , where q ∈ (0, 1] and each ωi by its values on ∗ -alg(cid:0)Jf K(cid:1) , this defines (via Proposition 5.8) a surjective map γ : RN × (0, 1] → X ∪ {0} by γ(x, q)(L1 ⊗ L2 ⊗ · · · ) := lim n→∞ L1(x1)L2(x2) · · · Ln(xn)qℓ for L1 ⊗ L2 ⊗ · · · ∈ Jf ℓK . To obtain a bijection with X from γ, note that if A := L1 ⊗ L2 ⊗ · · · = A1 ⊗ · · · ⊗ Am ⊗ f ℓ m+1 ⊗ f ℓ m+2 ⊗ · · · ∈ ∗ -alg(cid:0)Jf K(cid:1), then ∞Yk=1 ωk(Lk) = A1(x1) A2(x2) · · · Am(xm) fk(xk)ℓ = 0 ∀Ai, m, ℓ ∞Yk=m+1 if and only if lim m→∞ ∞Qk=m fk(xk) = 0 . Thus we define Nf :=nx ∈ RN lim m→∞ ∞Yk=m fk(xk) = 0o and hence the restriction γ : (cid:0)RN\Nf(cid:1) × (0, 1] → X is a surjection. That γ is bijective, is clear since each γ(x, q) is nonzero (as x 6∈ Nf ), and in each factor in the product, a component of L[f ] will separate the characters, and in the last entry, by definition all elementary tensors will separate different values of q . Thus we may identify (as sets) X with (cid:0)RN\Nf(cid:1) × (0, 1] . Note that Nf contains the set (cid:8)x ∈ RN(cid:12)(cid:12) xn ∈ f −1 n (0) for infinitely many n(cid:9) , hence since the fn are of compact Sn ⊂ RN where only finitely many of the Sn is contained in the union of sets support, RN\Nf ∞Qn=1 are not relatively compact. The w ∗ -topology of X w.r.t. L[f ] is not clear. The most important subset in X is X0 := X ∩ Rep0(cid:0)L[f ], C(cid:1) which corresponds to (cid:0)RN\Nf(cid:1) × {1} . We prove that it is a Gδ -- set. To see this, note ωk(fk) = 1 . This is an increasing limit. By using approximate that ω ∈ X0 if and only if identities in each factor L{k} , we can find for each n a net {A(n) α } ⊂ L[f ], 0 < A(n) α < 1 , such lim n→∞ ∞Qk=n ω(A(n) α ) for all ω ∈ X . Define a function qf : X → [0, 1] n−1 factors that ω(cid:0) z 1 ⊗ · · · ⊗ 1 ⊗fn ⊗ fn+1 ⊗ · · ·(cid:1) = sup } { α α ) then X0 = q−1 ({1}) . Since qf ω(A(n) f by qf (ω) := sup α, n on X it is lower semicontinous (cf. 6.3 in [Ko69]), i.e., q−1 X0 = q−1 it follows that X0 is a Gδ -- set. q−1 ({1}) = Tn∈N f (cid:0)( n−1 n , ∞)(cid:1), f To make a host algebra out of L[f ], i.e., to make η∗ injective, we need to reduce its spectrum to X0. However, since we do not know whether X0 is a locally compact subset of X this is not easy. From the fact that it is a Gδ -- set, we can identify X0 as the common characters of the decreasing is the supremum of continuous functions f (cid:0)(t, ∞)(cid:1) is open for all t ∈ R . Since 29 sequence of C*-algebras C0(cid:0)q−1 f (cid:0)( n−1 n , ∞)(cid:1)(cid:1) ⊂ L[f ], where of course η(R(N)) still acts on these as multipliers (i.e., as elements of Cb(X), with pointwise multiplication). 6 Hosting the full representation theory of R(N) We first want to extend the semi-host algebra L[f ] η∗(cid:0)Rep(cid:0)LV , H(cid:1)(cid:1) = Rep(cid:0)R(N), H(cid:1) . Recall that for above to an algebra LV , such that Vn :=(cid:8)f ∈ C0(R)(cid:12)(cid:12) f (R) ⊆ [0, 1], f ↾ [−n, n] = 1, supp(f ) ⊆ [−n − 1, n + 1](cid:9) . we obtain a multiplicative subsemigroup V := Vn in C0(R) . Thus, by Theorem 3.10(iii), V = V ∗, ∞Sn=1 implies that is a ∗ -subalgebra of A(V) := Span(cid:8)b ∈ Jf K f ∈ V N(cid:9) = Span(cid:8) ∞ ∞Nn=1 C0(R) . ⊗ n=1 gn g ∼ f ∈ V N(cid:9) 6.1 Proposition There is a ∗ -representation πu : A(V) → B(Hu) such that πu(L1 ⊗ L2 ⊗ · · · ) = s-lim n→∞ π1 u(L1) π2 u(L2) · · · πn u (Ln) for all elementary tensors L1 ⊗ L2 ⊗ · · · ∈ A(V), where πk above Definition 5.3). u : C0(R) → B(Hu) are as before (cf. text Proof. By Proposition 5.6(iii), πu is already a ∗ -representation on each ∗ -alg(cid:0)Jf K(cid:1) for f ∈ V N , hence it is a linear map on each Jf K for f ∈ V N. However, by Proposition 3.7(iv) we know that for f , g ∈ V N with Jf K 6= {0} 6= JgK we have Jf K ∩ JgK = {0} if and only if Jf K 6∼ JgK. Thus the set the sum of the subspaces Jf K is direct. Thus, since πu is defined as a linear map on each Jf K, of spaces (cid:8)Jf K f ∈ V N(cid:9) is labelled by the equivalence classes [f ] ⊂ V N, and by Proposition 3.7(iv), extends uniquely to a linear map πu on A(V) = Span(cid:8)b ∈ Jf K f ∈ V N(cid:9). To show that πu is a ∗ -homomorphism, it suffices to check this on the elementary tensors gn it ∞ ⊗ n=1 with g ∼ f ∈ V N. For f , g ∈ V N, let A = A1 ⊗ · · · ⊗ Ak−1 ⊗ fk ⊗ fk+1 ⊗ · · · ∈ Jf K and B = B1 ⊗ · · · ⊗ Bk−1 ⊗ gk ⊗ gk+1 ⊗ · · · ∈ JgK where we can choose the same k for both. Then by Proposition 5.6(ii) we have πu(A) = π1 u(A1) · · · πk−1 u (Ak−1)Fu,k[f ] and πu(B) = π1 u(B1) · · · πk−1 u (Bk−1) Fu,k[g] and where πu(AB) = π1 u(A1B1) · · · πk−1 u (Ak−1Bk−1)Fu,k[f · g] Fu,k[f ] := s-lim n→∞ πk u(fk) · · · πn u (fn) . Since πj u is a representation for all j, we only need to show that Fu,k[f · g] = Fu,k[f ] Fu,k[g] to establish that πu(AB) = πu(A) πu(B) . We have Fu,k[f · g] = s-lim n→∞ = s-lim n→∞ u (fngn) = s-lim n→∞ πk u(gk) · · · πm u (fn) · s-lim m→∞ πk u(fkgk) · · · πn πk u(fk) · · · πn πk u(fk) · · · πn u (fn) πk u(gk) · · · πn u (gn) u (gm) = Fu,k[f ] Fu,k[g] 30 since the operator product is jointly continuous in the strong operator topology on bounded subsets. Thus πu is a homomorphism. To see that it is a ∗ -homomorphism, note that πu(A)∗ = π1 u(A∗ 1) · · · πk−1 u (A∗ k−1)Fu,k[f ] = πu(A∗) since all πj ∗ -homomorphism of A(V). u are ∗ -homomorphisms with commuting ranges, and Jf K∗ = Jf ∗K = Jf K. Thus πu is a As in Section 5, we define 6.2 Definition The C∗ -algebra LV is the C∗ -completion of πu(cid:0)A(V)(cid:1) in B(Hu) . Note that LV = C∗(cid:8)L[f ] f ∈ V N(cid:9) ⊂ B(Hu) . η : R(N) → M(cid:0)LV(cid:1) , where We extend the unitary embeddings η : R(N) → U M(cid:0)L[f ](cid:1) from above to LV as follows. Define η(x1, . . . , xn, 0, 0, . . .)(L1 ⊗ L2 ⊗ · · · ) = η1(x1)L1 ⊗ · · · ⊗ ηn(xn)Ln ⊗ Ln+1 ⊗ Ln+2 ⊗ · · · = ζ{1,...,n}(x1, . . . , xn)(cid:0)L1 ⊗ L2 ⊗ · · ·(cid:1) for all (x1, . . . , xn) ∈ Rn, n ∈ N, Li ∈ L{i} = C0(R), and where ηi : R → M (C∗(R)) is the usual unitary embedding. Clearly, η restricts to the previous definition of it on each L[f ] ⊂ LV . Then the map η∗ : Rep(cid:0)LV , H(cid:1) → Rep(cid:0)R(N), H(cid:1) consists of the strict extension of (non-degenerate) representations of LV to η(R(N)), i.e., (η∗π)(x) := s-lim α→∞ π(cid:0)η(x)Eα(cid:1) ∀ x ∈ R(N) and any approximate identity {Eα}α∈Λ ⊂ LV , and η∗ obviously takes irreducibles to irreducibles by commutativity. 6.3 Definition Let Rep0(cid:0)LV , H(cid:1) denote those non-degenerate ∗ -representations π : LV → B(H) for which π ↾ L[f ] ∈ Rep0(cid:0)L[f ], Hf(cid:1) for all f , where Hf := π(cid:0)L[f ](cid:1)H . That is, each restriction of π to L[f ] has excess operator Qf = 1 on its essential subspace Hf . By Proposition 5.6, this means that all Qf (π) := s-lim n→∞ Bn[f ] are projections, where: Bn[f ] :=eπ(cid:0) n−1 factors { } z 1 ⊗ · · · ⊗ 1 ⊗fn ⊗ fn+1 ⊗ · · ·(cid:1). F (1) π,k where F (1) In fact, the projections Qf (π) must be the range projections Pπ[f ] = s-lim k→∞ πk(fk) · · · πn(fn) . Note that a direct sum of representations πi ∈ Rep0(cid:0)LV , Hi(cid:1), s-lim n→∞ π,k := i ∈ I (an index set) is again of the same type, i.e., Li∈I πi ∈ Rep0(cid:0)LV ,Li∈I Hi(cid:1). 6.4 Theorem Given the preceding notation, we have that (i) η : R(N) → M(cid:0)LV(cid:1) is continuous w.r.t. the strict topology of M(cid:0)LV(cid:1) . (ii) The map η∗ is injective on Rep0(cid:0)LV , H(cid:1) . (iii) The range η∗(cid:0)Rep(cid:0)LV , H(cid:1)(cid:1) is the same as η∗(cid:0)Rep0(cid:0)LV , H(cid:1)(cid:1) and is all of Rep(cid:0)R(N), H(cid:1) . (iv) π ∈ Rep0(cid:0)LV , H(cid:1) if and only if π is normal w.r.t. πu. 31 Proof. (i) Since η : R(N) → M(cid:0)LV(cid:1) is bounded, it suffices to show that the space is norm continuous(cid:9) (cid:8)L ∈ LV the map R(N) ∋ x 7→ η(x)L ∈ LV is dense in LV . But this follows from the fact that by Theorem 5.9(i), this space contains all Jf K ⊂ L[f ], and these spaces span A(V) which is dense in LV . (ii) Consider π, π′ ∈ Rep0(cid:0)LV , H(cid:1) such that η∗π = η∗π′. Then for the restrictions to Rn ⊂ R(N) we have eπ(n) := η∗π ↾ Rn = η∗π′ ↾ Rn := eπ′(n) . Moreover, L(n) embeds in M (LV ) as L(n) ⊗ 1 (acting on the elementary tensors), hence π also extends to it to define a non-degenerate π(n) : L(n) → B(Hπ) . Since η is defined via the natural actions, we have η(x)L(n) ⊆ L(n) for all x ∈ Rn . Since eπ(n)(x) π(n)(L) π(A) = η∗π(x) π(LA) = π(cid:0)η(x)LA(cid:1) = π(n)(η(x)L) π(A) for all x ∈ Rn, L ∈ L(n), A ∈ LV , we see by nondegeneracy of π that eπ(n)(x) π(n)(L) = π(n)(η(x)L) for all L ∈ L(n), and hence since eπ(n) and π(n) are non-degenerate and L(n) for Rn, this relation gives a bijection between eπ(n) and π(n). We conclude from eπ(n) = eπ′(n) k for all k . Now for each elementary tensor L1 ⊗ L2 ⊗ · · · ∈ ∗ -alg(cid:0)Jf K(cid:1) ⊂ L[f ] we know by for all n . A similar argument for the kth component alone, also shows that that π(n) = π′(n) πk = π′ Proposition 5.6(iii) that is a host algebra π(L1 ⊗ L2 ⊗ · · · ) = s-lim n→∞ π(n)(cid:0)L1 ⊗ L2 ⊗ · · · ⊗ Ln(cid:1)Qf (π) . (10) Recall that by hypothesis Qf (π) = Pπ[f ] = s-lim k→∞ ogous expressions hold for π′, thus since πk = π′ hence from Equation (10) it follows from π(n) = π′(n) L[f ] hence on all of LV , which proves the claim. F (1) π,k , where F (1) k for all k , πk(fk) · · · πn(fn) . Anal- π,k := s-lim n→∞ it follows that Qf (π) = Qf (π′) and for all n , that π and π′ coincides on all (iii) Let π ∈ Rep0(cid:0)LV , H(cid:1) and let eπ be its strict extension to M(cid:0)LV(cid:1). As eπ is strictly con- tinuous, (i) implies that the unitary representation η∗(π) = eπ ◦ η : R(N) → U(H) is strong operator continuous, i.e., η∗(cid:0)Rep(cid:0)LV , H(cid:1)(cid:1) ⊆ Rep(cid:0)R(N), H(cid:1) . To prove the claim of this theorem, we need to show that for each π ∈ Rep(cid:0)R(N), Hπ(cid:1), there is a π(0) ∈ Rep0(cid:0)LV , Hπ(cid:1) such that η∗π(0) = π . Since each π ∈ Rep(cid:0)R(N), Hπ(cid:1) is a direct sum of cyclic representations, and η∗ preserves direct sums, it suffices to show that for each cyclic π ∈ Rep(cid:0)R(N), Hπ(cid:1), there is a π(0) ∈ Rep0(cid:0)LV , Hπ(cid:1) such that η∗π(0) = π . Fix a cyclic π ∈ Rep(cid:0)R(N), Hπ(cid:1), then there is a projection Pπ ∈ πu(R(N))′ such that π = (Pππu) ↾ Hπ where Hπ = PπHu . Recall the inclusion R → R(N), x 7→ xek , so let πk : R → U(Hπ) be πk(x) := π(xek) . By the host algebra property of C∗(R) ∼= C0(R), this produces a unique non-degenerate representation πk : C0(R) → B(Hu) , which is characterized by πk(x) πk(L) = πk(ηk(x)L) = π(cid:0)η(0, . . . , 0, x, 0, 0, . . .)(1 ⊗ · · · 1 ⊗ L ⊗ 1 ⊗ · · · )(cid:1) ( x and L in the kth entries) for all x ∈ R and L ∈ C0(R). Since π(cid:0)η(0, . . . , 0, x, 0, . . .)(1 ⊗ · · · 1 ⊗ L ⊗ 1 ⊗ · · · )(cid:1) = Pππu(cid:0)η(0, . . . , 0, x, 0, . . .)(1 ⊗ · · · 1 ⊗ L ⊗ 1 ⊗ · · · )(cid:1) ↾ Hπ = Pππk u(x) πk u(L) ↾ Hπ = πk(x) Pππk u(L) ↾ Hπ we get that πk(L) = Pππk u(L) ↾ Hπ for all L ∈ C0(R). Since the set of representations (cid:8)πk : C0(R) → B(Hπ) k ∈ N(cid:9) have commuting ranges, we can apply Lemma 5.1 (with the choice 32 Q = 1 ) to define a representation π(0) : ∗ -alg(cid:0)Jf K(cid:1) → B(Hπ), for all f , and we need to show that π(0) extends to a representation of LV . Now Pπ commutes with the images of all πk u (since it commutes with πu(R(N)) ) hence all πk u(L) preserve Hπ and so by its definition πu(LV ) preserves Hπ. Thus the map A ∈ LV → Pππu(A) ↾ Hπ is a ∗ -representation of LV and it coincides with π(0) on each ∗ -alg(cid:0)Jf K(cid:1) because Pππu(L1 ⊗ L2 ⊗ · · · ) ↾ Hπ = s-lim n→∞ Pππ1 u(L1) π2 u(L2) · · · πn u (Ln) ↾ Hπ = s-lim n→∞ π1(L1) π2(L2) · · · πn(Ln) = π(0)(L1 ⊗ L2 ⊗ · · · ) for all elementary tensors L1 ⊗ L2 ⊗ · · · ∈ A(V). This defines a π(0) degenerate, note that its restriction to any L[f ] ⊂ LV has essential projection Pπ[f ] = s-lim : LV → B(Hπ) by π(0)(A) = Pππu(A) ↾ Hπ for all A ∈ LV . To see that it is non- ∗ -representation n→∞ where eFk := s-lim πk(fk) · · · πn(fn) by Theorem 5.9(iii) and Lemma 5.1(ii). It is suffices to show that for each nonzero ψ ∈ Hπ there is a sequence f ∈ V N such that Pπ[f ]ψ 6= 0. Fix a nonzero ψ ∈ Hπ. Since there is an approximate identity of C0(R) in V, it is possible to choose for each n ∈ N a fn ∈ V such that kψ − πn(fn)ψk < 1/n2, hence we may write πn(fn)ψ = ψ + ξn/n2 where kξnk ≤ 1. Then k→∞ eFk πk(fk) · · · πn(fn)ψ = ψ + 1 n2 πk(fk) · · · πn−1(fn−1) ξn + 1 (n − 1)2 πk(fk) · · · πn−2(fn−2) ξn−1 + · · · + 1 k2 ξk . Thus eFkψ = ψ + ∞Xj=k 1 j2 j−1Yℓ=k πℓ(fℓ) ξj, where j−1Yℓ=k (cid:13)(cid:13) πℓ(fℓ) ξj(cid:13)(cid:13) ≤ 1 and hence Pπ[f ]ψ = ψ as the series converges. Thus π(0) is non-degenerate. Since π(0) is obviously normal to πu, it follows that the excess operator is Q = 1 for the restriction of π(0) to any L[f ], and hence π(0) ∈ Rep0(cid:0)LV , Hπ(cid:1) . To see that η∗π(0) = π , note that for x ∈ Rk ⊂ R(N) we have η∗π(0)(x) π(0)(L1 ⊗ L2 ⊗ · · · ) = π(0)(cid:0)η(x)(L1 ⊗ L2 ⊗ · · · )(cid:1) = π(0)(cid:0)η1(x1)L1 ⊗ · · · ⊗ ηk(xk)Lk ⊗ Lk+1 ⊗ Lk+2 ⊗ · · ·(cid:1) = Pππu(cid:0)η1(x1)L1 ⊗ · · · ⊗ ηk(xk)Lk ⊗ Lk+1 ⊗ Lk+2 ⊗ · · ·(cid:1) ↾ Hπ = Pπ s-lim n→∞ π1 u(η1(x1)L1) π2 u(η2(x2)L2) · · · πk u(ηk(xk)Lk) πk+1 u (Lk+1) · · · πn u (Ln) ↾ Hπ = Pπ s-lim n→∞ π1 u(x1)π1 u(L1) π2 u(x2)π2 u(L2) · · · πk u(xk)πk u(Lk) πk+1 u (Lk+1) · · · πn u (Ln) ↾ Hπ = π1(x1) · · · πk(xk) s-lim n→∞ π1(L1) π2(L2) · · · πn(Ln) = π(x)π0(L1 ⊗ L2 ⊗ · · · ) for all elementary tensors L1⊗L2⊗· · · ∈ A(V), hence η∗π0(x) π(0)(A) = π(x) π(0)(A) for all A ∈ LV . Since π(0) is non-degenerate, it follows that η∗π(0) = π as required. (iv) By Theorem 5.9(v) we have that π ∈ Rep0(cid:0)LV , H(cid:1) if and only if π ↾ L[f ] (on its essential f ∈ V N. Let π ∈ Rep(cid:0)LV , H(cid:1) be normal w.r.t. subspace Hf ) is normal w.r.t. πu(cid:0)L[f ](cid:1) for all 33 πu(LV ) . Then it is continuous on bounded sets w.r.t. the strong operator topologies of both sides, and it follows that this is true for its restrictions to each πu(L[f ]), and hence that each restriction is normal w.r.t. πu. Thus π ∈ Rep0(cid:0)LV , H(cid:1). Conversely, given π ∈ Rep0(cid:0)LV , H(cid:1) then by part (iii) η∗π is a continuous representation of R(N), and by Proposition 5.6(iii) (with Q = 1) we have that on each Hf π(L1 ⊗ L2 ⊗ · · · ) = s-lim n→∞ π(n)(cid:0)L1 ⊗ L2 ⊗ · · · ⊗ Ln(cid:1) = s-lim n→∞ π1(L1)π2(L2) · · · πn(Ln) for all elementary tensors L1 ⊗ L2 ⊗ · · · ∈ Jf ℓK ⊂ ∗ -alg(cid:0)Jf K(cid:1). Now πu(R(N))′′ = u (L(n)) n ∈ N}′′ = πu(cid:0)(cid:8)L[f ] f ∈ V N(cid:9)(cid:1)′′ {π(n) = πu(LV )′′ and a similar equation holds for π. Since the cyclic components of π are contained in the direct summands of πu, there is a normal map ϕ : πu(LV )′′ → B(H) such that ϕ ◦ πu = π. Thus π is normal to πu. Thus LV is a semi-host for the full representation theory of R(N), i.e., η∗ : Rep(cid:0)LV , H(cid:1) → Rep(cid:0)R(N), H(cid:1) is surjective, but not necessarily injective. We want to examine the remaining representations of LV outside of Rep0(cid:0)LV , H(cid:1) . Denote the universal representation of LV by πU : LV → B(HU ) (not to be confused with the defining representation πu) . Let Q :=(cid:8)Qf (πU ) f ∈ V N(cid:9) ⊂ L′′ V := πU (LV )′′ , i.e., the set of all excess operators w.r.t. πU . Since Q is in the positive part of the unit ball of L′′ V it has a natural partial order, and in a moment we will see that Q is a multiplicative semigroup. Let , Rep(cid:0)Q, H(cid:1) :=(cid:8)γ : Q → B(H) 0 ≤ γ(Q1) ≤ 1, γ(Q1Q2) = γ(Q1)γ(Q2) ∀ Qi ∈ Q(cid:9) . 6.5 Proposition With notation above, we have (i) Qf1(πU ) · Qf2(πU ) = Qf1·f2(πU ) for all fi ∈ V N . Thus Q is a multiplicative semigroup, and the map [f ]∼ → Qf (πU ) is a surjective homomorphism V∞ → Q of multiplicative semigroups where V∞ := {[f ]∼ f ∈ V N} . (ii) Fix a non-degenerate ∗ -representation π : LV → B(Hπ) . Then the map [f ]∼ → Qf (π) defines a representation of V∞ as well as of Q . Thus every π ∈ Rep(cid:0)LV , H(cid:1) is of the form: π(A) := π0(A)γ(f ) for A ∈ Jf K for some π0 ∈ Rep0(cid:0)LV , H(cid:1) and γ ∈ Rep(cid:0)Q, H(cid:1) with γ(Q) ⊂ π(LV )′ . { Proof. (i) Recall that Qf (π) := s-lim n→∞ the operator product is jointly continuous on bounded subsets we have: n−1 factors } z 1 ⊗ · · · ⊗ 1 ⊗fn ⊗ fn+1 ⊗ · · ·(cid:1). Since Bn[f ] , where Bn[f ] := eπ(cid:0) n→∞ fπU(cid:0)1 ⊗ · · · ⊗ 1 ⊗ fn ⊗ fn+1 ⊗ · · ·(cid:1) s-lim k→∞ fπU(cid:0)1 ⊗ · · · ⊗ 1 ⊗ gk ⊗ gk+1 ⊗ · · ·(cid:1) n→∞ fπU(cid:0)1 ⊗ · · · ⊗ 1 ⊗ fn ⊗ fn+1 ⊗ · · ·(cid:1)fπU(cid:0)1 ⊗ · · · ⊗ 1 ⊗ gn ⊗ gn+1 ⊗ · · ·(cid:1) n→∞ fπU(cid:0)1 ⊗ · · · ⊗ 1 ⊗ fngn ⊗ fn+1gn+1 ⊗ · · ·(cid:1) = s-lim = s-lim = Qf ·g(πU ) . Qf (πU ) · Qg(πU ) = s-lim 34 It will suffice for this part to show that the map [f ] → Qf (πU ) is well-defined, i.e., that Qf (πU ) only depends on the equivalence class [f ] not on any particular representative which is chosen. However, this is immediate from the definition of Qf (πU ) . (ii) By the universal property of πU (cf. Theorem 10.1.12 in [KR86]) there is a central projection that π(A) = α(PππU (A)) for all A ∈ LV . The map α is normal in both directions (cf. Proposi- Pπ ∈ Z(cid:0)πU (LV )′′(cid:1) and a *-isomorphism of Von Neumann algebras α : PππU (LV )′′ → π(LV )′′ such tion 2.5.2 in [Ped89]). It is also true that eπ(A) = α(PπeπU (A)) for all A ∈ M (LV ). So it follows n→∞ fπU(cid:0)1 ⊗ · · · ⊗ 1 ⊗ fn ⊗ fn+1 ⊗ · · ·(cid:1)(cid:17) n→∞eπ(cid:0)1 ⊗ · · · ⊗ 1 ⊗ fn ⊗ fn+1 ⊗ · · ·(cid:1) = α(cid:16)Pπ s-lim Qf (π) = s-lim from = α(PπQf (πU )) and part (i) that Qf (π) · Qg(π) = Qf ·g(π) hence the map [f ]∼ → Qf (π) defines a representation of V∞ as well as of Q . The second claim is immediate. Thus the the additional part of Rep(cid:0)LV , H(cid:1) to Rep(cid:0)R(N), H(cid:1) is in Rep(cid:0)Q, H(cid:1) . so it is a lower semi-continuous function on the spectrum of LV . closure Lm V By definition, each Q ∈ Q is the strong operator limit of increasing positive elements in πU (LV ), In fact, Q is in the monotone (cf. [Tak79, Thm. 6.8 and above, p. 182]). Let X be the spectrum of LV , and let it has to be 0 X0 := X ∩ Rep0(cid:0)LV , C(cid:1). Then since ω(Q) must be idempotent for ω ∈ X0, Q ∈ Q, or 1. Thus X0 ⊂ Q−1({0}) ∪ Q−1({1}), and by the definition of Rep0(cid:0)LV , C(cid:1) we get that X0 = \Q∈Q(cid:0)Q−1({0}) ∪ Q−1({1})(cid:1) . This suggests that to obtain a full host algebra for R(N) we only need to apply the homomorphism Q−1((0, 1)) , but this is not possible, because we do not know whether the which factors out by SQ∈Q last set is open, as the Q are only lower semi-continuous. 7 Discussion Here we constructed an infinite tensor product of the algebras C0(R), denoted LV , and used it to find semi -- hosts for the full continuous representation theory of R(N). Due to commutativity, these were as useful as host algebras, because η∗ preserves irreducibility in this context. We also interpreted the Bochner -- Minlos theorem in R(N) as the pure state space decomposition of the partial hosts which LV comprises of. We analyzed the representation theory of LV , and showed that η∗ is a bijection between Rep0(cid:0)LV , H(cid:1) and Rep(cid:0)R(N), H(cid:1) , but that there is an extra part which essentially consists of the representation theory of a multiplicative semigroup Q. Much further analysis remains, e.g. the topological structure of the spectrum X of LV , especially the important question as to whether X0 is locally compact with the relative topology. Moreover, one can easily apply the methods developed here to construct infinite tensor products of other C*-algebras without nontrivial projections. A very important issue, is to extend the C*-algebraic interpretation of the Bochner -- Minlos theorem developed here, to general nuclear spaces. 35 Acknowledgements HG wishes to thank the Mathematics Department of the Technische Universitat Darmstadt who gener- ously supported his visit to Darmstadt in 2009, as well as the Deutscher Akademischer Austauschdienst (DAAD) who funded his trip to Germany. References [AW66] Araki, H., Woods, E.J., Complete Boolean algebras of type I factors, Publ. Res. Inst. Math. Sci. Kyoto Univ. 2 (1966), 157 -- 242 [Bl77] Blackadar, B., Infinite tensor products of C∗ -algebras, Pac. J. Math. 77 (1977), 313 -- 334 [Bo07] Bogachev, V.I., "Measure Theory. Volume II." Springer-Verlag, Berlin, Heidelberg, 2007 [Bou89] Bourbaki, N., "Algebra, Chapters 1 -- 3," 2nd printing, Springer -- Verlag, Berlin, 1989 [BR79] Bratteli, O., Robinson, D.W., "Operator Algebras and Quantum Statistical Mechanics I. C∗ and W ∗ -Algebras, Symmetry Groups, Decomposition of States." Springer -- Verlag, New York, 1979 [Da96] Davidson, K. R., " C∗ -algebras by example," American Mathematical Society, Providence, Rhode Island, 1996 [Dix77] Dixmier, J., " C∗ -algebras," North Holland Publishing Company, Amsterdam, New York, Oxford, 1977 [Fe60] Fell, J. M. G., The Dual Spaces of C∗ -algebras, Trans. Amer. Math. Soc. 94 (1960), 365 -- 403 [GV64] Gel'fand, I. M., and N. Ya Vilenkin, "Generalized Functions. Vol. 4: Applications of Harmonic Analysis," Translated by Amiel Feinstein, Academic Press, New York, London, 1964 [GN01] H. Glockner, K.-H. Neeb, Minimally almost periodic abelian groups and commutative W ∗ - algebras, in "Nuclear Groups and Lie Groups", Eds. E. Martin Peinador et al., Research and Exposition in Math. 24, Heldermann Verlag, 2001, 163 -- 186 [Gu67] Guichardet, A., Produits tensoriels continus d'espaces et d'alg`ebres de Banach, Commun. Math. Phys. 5 (1967), 262 -- 287 [Gr97] Grundling, H., A group algebra for inductive limit groups. Continuity problems of the canonical commutation relations, Acta Applicandae Mathematicae 46 (1997), 107 -- 145 [Gr05] -- , Generalizing group algebras, J. London Math. Soc. 72 (2005), 742 -- 762. An erratum is in J. London Math. Soc. 77 (2008), 270 -- 271 [GrN09] Grundling, H., Neeb, K-H., Full regularity for a C*-algebra of the Canonical Commutation Relations, Rev. Math. Phys. 21 (2009), 587 -- 613 [Ko69] Kothe, G., "Topological Vector Spaces I," Berlin: Springer-Verlag, 1969 36 [KR83] Kadison, R. V., Ringrose, J. R., "Fundamentals of the theory of operator algebras. Vol. 1: Elementary theory," Academic Press, New York, London, Sydney, 1983 [KR86] -- , "Fundamentals of the theory of operator algebras. Vol. 2: Advanced theory," Academic Press, New York, London, Sydney, 1986 [Mu90] Murphy, G. J., "C ∗ -- Algebras and Operator Theory," Boston, Academic Press, 1990 [Pa94] Palmer, T. W., "Banach Algebras and the General Theory of C∗ -algebras. Volume I; Alge- bras and Banach Algebras," Cambridge Univ. Press, 1994 [Ped89] Pedersen, G. K., "C ∗ -- Algebras and their Automorphism Groups," London: Academic Press, 1989 [Tak79] Takesaki, M., "Theory of operator algebras I," New York, Springer-Verlag, 1979 [Tak75] Takeuti, G., Zaring, W. M., "Introduction to Axiomatic Set Theory," Berlin, Springer-Verlag, 1975 [vN61] von Neumann, J., On Infinite Direct Products, Comp. Math. 6, 1 -- 77. Collected Works, Vol. 3, Chapter 6. (ed. A.H. Taub), Pergamon Press, Oxford, New York, Paris 1961. [WO93] Wegge -- Olsen, N. E., "K -- theory and C∗ -- algebras," Oxford University Press, Oxford, 1993 [Wo95] Woronowicz, S., C∗ -algebras generated by unbounded elements, Rev. Math. Phys. 7 (1995), 481 -- 521 37
1802.01297
1
1802
2018-02-05T08:22:49
On a gauge action on sigma model solitons
[ "math.OA" ]
In this paper we consider a gauge action on sigma model solitons over noncommutative tori as source spaces, with a target space made of two points introduced in \cite{DKL:Sigma}. Using new classes of solitons from Gabor frames, we quantify the condition about how to gauge a Gaussian to a prescribed Gabor frame.
math.OA
math
ON A GAUGE ACTION ON SIGMA MODEL SOLITONS HYUN HO LEE Abstract. In this paper we consider a gauge action on sigma model solitons over noncommutative tori as source spaces, with a target space made of two points introduced in [4]. Using new classes of solitons from Gabor frames, we quantify the condition about how to gauge a Gaussian to a prescribed Gabor frame. 1. Introduction Noncommutative analogues of non-linear σ-models appeared in [4, 5] for the first time. Later other examples including noncommutative spacetimes were considered by [12, 15]. Among them there is a continuous analog of the Ising model which consists of field maps from a noncommutative torus to a two-point space. Cosidering only an energy term in the action or excluding a gravity related term in the action, the stable maps are called noncommutative harmonic maps and in this particular case such maps correspond to some smooth projections in the noncommutative torus. It turned out that enegy minimizing ones carry a nontrivial topological charge and satisfy a Belavin-Polyakov bound [11]. The construction of such maps depends on a geometric picture or a strong Morita equivalence. Since we can view a noncommutative torus Aθ as an endomorphism algebra of a suitable finitely generated projective bundle, we can think of a projection in Aθ as an operator on the bundle. The bundle is in fact a bimodule over two different noncommutative tori with operator valued Hermitian structures compatible with each other. By choosing suitable vectors ξ in the module, we consider Rieffel- type projections (see the paragraph after Proposition 2.1) and lift the field equation on the noncommutative torus to an equation of ξ on the module. The vectors in the module both inducing Rieffel-type projections and satisfying a (anti) self duality equation are called noncommutative instantons or solitons following G. Landi. Using the idea of a natural transformation on the bundle a gauge action on non- commutative solitons is defined by the right multiplication of invertible elements g of a different noncommutative torus Aα. This gauge action is well behaved with Rieffel-type projections so that the vector ξ · g generates a Rieffel-type projection again, and satisfy a (anti) self duality equation whenever ξ does. An important class of ξ's, that are Gaussians, is already known to be solutions for the self duality equa- tion with a constant parameter and the condition when two Gaussians to be gauged each other is characterized in [4, 5]. However, for the generic case it is not true that 2000 Mathematics Subject Classification. Primary:58B20, 35C08. Secondary:58B16, 58J05, 42B35. Key words and phrases. Nonlinear sigma model, Gauge action, Gabor frames, Noncommutative torus, Noncommutative solitons. This research was supported by Basic Science Research Program through the National Research Foundation of Korea(NRF) funded by the Ministry of Education(NRF-2015R1D1A1A01057489). 1 2 HYUN HO LEE any solution vector ξ could be gauged away to a Gaussian solution since there is an obstruction in the form of ∂-equation (see Corollary 4.10). Nonetheless, there is a good chance that a class of vectors ξ could be gauged to a Gaussian and it is our purpose to provide an affirmative example for this question. In this direction it is necessary to know noncommutative solitons other than Gaussians and recently new classes of sigma-model solitons are discovered by Dabrowski, Landi, and Luef [6] un- der the observation that a problem in a time-frequency analysis and Gabor analysis is equivalent to find Rieffel-type projections in a noncommutative torus. This paper is organized as follows; In section 2, we explain a nonlinear σ-model on noncommutative tori introduced by Dabrowski, Krajewski, and Landi and define Rieffel-type projections using Hilbert module frames. In section 3, we introduce a class of functions called Gabor frames whose name is originated from Gabor analysis and clarify the condition for a vector ξ to be a Gabor frame in terms of Hermitian structures on the module. Then, in Section 4 we show that a Gaussian could be gauged to a hyperbolic secant based on Theorem 4.11 which provides a useful tool to check whether a concrete Gabor frame is gauged to a Gaussian with Corollary 4.10 in one hand. 2. σ-model on noncommutative torus Let Aθ be a ∗-algebra consisting of power series of the form a = X(m,n)∈Z2 amnU m 1 U n 2 with amn a complex-valued Schwarz function on Z2, or decreasing rapidly. Two unitaries U1, U2 have a commutation relation (1) U2U1 = e2πiθU1U2. For θ irrational, there is a unique faithful trace Tr on A given by Tr(Xm,n amnU m 1 U n 2 ) = a00. One can equip Aθ with a norm kak = sup(m,n)∈Z2 amn < ∞ and the closure of Aθ with respect to this norm is the universal C ∗-algebra Aθ generated by two unitaries satisfying the relation (1): Aθ is dense in Aθ and is a pre-C ∗-algebra. Also, it is well-known that Aθ is the smooth subalgebra of Aθ, and closed under the holomor- phic functional calculus [1]. Throughout the article, we are interested in the case θ irrational and call both Aθ and Aθ noncommutative torus without confusion. To define the noncommutative action functional for morphisms from a pre-C ∗ alge- bra B to a pre-C ∗-algebra A, note that there is a formal prescription due to Mathai and Rosenberg [15]; recall that a spectral triple (A, H, D) is given by an involutive ∗-algebra represented as bounded operators on a Hilbert space H and a self-adjoint (unbounded) operator D with a compact resolvant such that commutators [D, a] are bounded for all a ∈ A. A spectral triple (A, H, D) is said to be even if the Hilbert space H is endowed with a super-grading γ that commutes with all a ∈ A and anti- commutes with D. In addition, we say a spectral triple (A, H, D) is (2, ∞)-summable ON A GAUGE ACTION ON SIGMA MODEL SOLITONS 3 if Trω(aD−2) < ∞ where Trω is the Dixmier trace. With a (2, ∞)-summable even spectral triple (A, H, D) one can define a positive Hochschild 2-cocycle ψ2 given by ψ2(a0, a1, a2) = i 2π Trω((1 + γ)a0[a1, D][a2, D]D−2). We can compose it with a field map φ : B → A where B is a target space and A represents a string worldsheet in a noncommutative formalism of the classical σ- model. To assign a number to any field map we evaluate the induced cocycle on a suitably chosen element of B⊗B⊗B. Such an element is taken as the noncommutative analogue of the metric on the target, and we choose a positive element of the form G =Xi b0idbi idbi 2 in the space of universal 2-forms Ω2(B). Then the quantity S(φ) = φ∗(ψ)(G) ≥ 0 is the action functional of non-linear σ-model in noncommutative geometry. There is a well-known even spectral triple (Aθ, H, D) for the noncommutative torus Aθ with γ =(cid:18)1 0 −1(cid:19) , D = ∂1σ1 + ∂2σ2 0 where σ1, σ2 are Pauli matrices (see Section 4 for derivations ∂i's). When the target space is a two-point space, then B is just two dimensional complex vector space C2. Since a morphism φ : C2 → Aθ is determined by the image of e the characteristic function on a point, we denote it by a projection p ∈ Aθ. Taking G = dede ∈ Ω2(B) the action functional can be written as (2) S(p) = Tr(∂p∂p), where ∂ or ∂ are the derivations coming from the complex structure on noncommu- tative torus(see Section 4). It is known from [4, 5] or [12] that the Euler-Lagrange equation for this functional is (3) p(∆p) − (∆p)p = 0 where ∆ is the Laplacian. It is well known that there exist a lot of projections in Aθ, which is of real rank zero [7], contrary to the fact that a noncommutative torus is a deformation quantization of commutative two torus. But it is unclear whether there are smooth projections in Aθ. Thus it was a remarkable discovery of M. Rieffel to construct a projection in Aθ [17], so that a morphism φ : C2 → Aθ is well-defined. In fact, there is a systematic way to construct projections in a ∗-algebra with a left action module and the dual action algebra. Accordingly we call such projections Rieffel-type projections; for the moment, A is a ∗-algebra. Suppose that there is a ∗-algebra B that is strongly Morita equivalent to A via the bimodule Ξ (see Section 3 for the definition). If we denote the A(B)-valued hermitian inner product by Ah , i(h , iB), then Ahξ, ξi is a projection in A provided that hξ, ξiB = 1B. More generally, if we have ξ1, ξ2, . . . , ξn in Ξ, then the matrix, whose i, j entry is Ahξi, ξji, is a projection in Mn(A) provided k=1hξk, ξkiB = 1B. We call the set {ξ1, . . . , ξn} a module frame for Ξ. More that Pn 4 HYUN HO LEE precisely, it is called a (Parseval) standard module frame {ξ1, . . . , ξn} for Ξ [9]. In general, a standard module frame for Ξ is a set {ξ1, . . . , ξn} such that (4) Ahξ, ξiiAhξi, ξi ≤ c2Ahξ, ξi, for all ξ ∈ Ξ, c1Ahξ, ξi ≤Xi for positive constants c1 and c2. Since we are interested in a projection in A rather in Mn(A) the matrix algebra of A, we restrict ourselves to the case of a single frame {η} for Ξ. The following is one of strategies to find a single standard module frame which was used by many experts. Proposition 2.1. Let η be an element such that hη, ηiB is invertible. Then {η} is a standard module frame. Proof. Suppose that a positive element hη, ηi is invertible, then its spectrum is bounded below for a positive number c1 > 0 and bounded above by the norm of it, say c2. Thus, by the functional calculus, c11B ≤ hη, ηi ≤ c21B. Since Ahξ, ηiAhη, ξi = AhAhξ, ηi · η, ξi = Ahξ · hη, ηiB, ξi, (4) is satisfied. Therefore the invertibility of hη, ηiB is a sufficient condition for η to be a standard module frame. (cid:3) Once we have a frame η as in Proposition 2.1, we get a Parseval one η by the normalization and obtain a projection Ahη, ηi in A. We call such a projection Rieffel- type projection and the first example of Rieffel-type projections in Aθ was found by M. Rieffel using a compactly supported smooth function as η [17], but later F. Boca discovered another one using η, a Gaussian (A hard computation involving a quantum theta function was needed to show that {η} is a standard frame)[2]. Recently, F. Luef noticed that a fundamental duality principle in Gabor analysis is linked to the invertibility of hη, ηiB in the case of noncommutative torus and found a large class of standard module frames which include previous examples of Rieffel and Boca [13]. Surprisingly this class of standard module frames gives rise to minimizing solutions of (3) as claimed in [6]. We are going to explain this fact more carefully and give a detailed proof in Section 4. 3. Gabor frames and Noncommutative tori In this section, we summarize the strong Morita equivalence of Aθ with its dual B in terms of Gabor analysis from [13],[14], and explain a class of functions generates Rieffel-type projections in Aθ. We say that two pre C ∗-algebras A and B are strongly Morita equivalent if there is a bimodule Ξ, on which both A and B act left and right respectively, equipped with A-valued inner product Ah , i and B-valued inner product h , iB which satisfy the following conditions; for any f, g ∈ Ξ, and a ∈ A, b ∈ B Ahf, gi∗ = Ahg, f i, Aha · f, gi = a · Ahf, gi, B = hg, f iB, hf, gi∗ hf, g · biB = hf, giB · b, f · hg, hiB = Ahf, gi · h, (a · f ) · b = a · (f · b). ON A GAUGE ACTION ON SIGMA MODEL SOLITONS 5 Let π : (x, ω) ∈ R2 → B(L2(R)) be a (projective) representation defined by , or (π(x, ω)ξ)(t) = e2πitωξ(t − x) π(x, ω) = MωTx where (Mωξ)(t) = e2πitωξ(t) and (Txξ)(t) = ξ(t−x). Then the canonical commutation relation for Mω and Tx holds, (5) It follows that MωTx = e2πixωTxMω. π(z)π(z′) = e−2πix·ηπ(z + z′) for z = (x, ω), z′ = (y, η). We can easily check c : R × R → T defined by c(z, z′) = e−2πixη for z = (x, ω), z′ = (y, η) is a 2-cocycle. Let Λ be a lattice of R2( for our purpose, we may assume that Λ is of the form θZ × Z). Then G(g, Λ) = {π(λ)g λ ∈ Λ} in L2(R) is said to be a Gabor system. A Gabor system is a Gabor frame for L2(R) if there exist α, β > 0 such that (6) αkf k2 2 ≤Xλ∈Λ hf, π(λ)gi2 ≤ βkf k2 2. In this case, g is called a Gabor atom or window in time-frequency analysis, but we call it a Gabor frame abusing notation since g will give rise to a module frame in our setting. Recall that l1(Λ, c) is a l1(Λ) with a twisted convolution of a and b defined by a♮b(λ) =Xµ∈Λ a(µ)b(λ − µ)c(µ, λ − µ), and involution a∗ = (a∗(λ)) of a is given by a∗(λ) = c(λ, λ)a(−λ) for λ ∈ Λ. Then C ∗(Λ, c) is the completion of l1(Λ, c) under π. More precisely, C ∗(Λ, c) is the completion of the involutive representation of a's, π(a) =Xλ∈Λ a(λ)π(λ) for a = (a(λ))λ∈Λ with the product and the involution π(a)π(b) = π(a♮b), π(a)∗ = π(a∗). Weighted analogues of the twisted group algebra are needed to obtain A in terms of Gabor analysis ; for s ≥ 0 let l1 s(Λ) be the space of all sequences a with kaks = s(Λ), ♮,) and the involutive representation of Pλ∈Λ a(λ)(1 + λ2)s/2. We consider (l1 it, so A1 s(Λ, c) = {T ∈ B(L2(R)) T =Xλ∈Λ a(λ)π(λ), kaks < ∞} 6 HYUN HO LEE is an involutive algebra with respect to the norm kT k = Xλ∈Λ turns out that A∞(Λ, c) = Ts≥0 A1 torus A. A dual lattice to Λ is defined by a(λ)(1 + λ2)s/2. It s(Λ, c) is equal to the smooth noncommutative Λ◦ = {z ∈ R2 π(λ)π(z) = π(z)π(λ) for all λ ∈ Λ}. Then we have C ∗(Λ◦, ¯c), A1 space in time-frequency analysis. More explicitly, s(Λ◦, ¯c), A∞(Λ◦, ¯c) similarly. Let M 1 s (R) be the modulation M 1 s (R) = {f ∈ L2(R) kf kM 1 Vφf (x, ω)(1 + x2 + ω2)s/2dxdω < ∞} s :=ZR where Vφf is the short-time Fourier transform of a function f with respect to the window φ, which is defined by hf, π(x, ω)φiL2(R). We can characterize the Schwartz space in terms of modulation spaces: S (R) = \s≥0 M 1 s (R). Theorem 3.1. [13, Theorem 2.3] For any s ≥ 0 M 1 tween A1 and A∞(Λ◦, ¯c). s (R) is an equivalence bimodule be- s(Λ◦, ¯c) and S (R) is an equivalence bimodule between A∞(Λ, c) s(Λ, c) and A1 Although we do not need here, the strong Morita equivalence between C ∗(Λ, c) and C ∗(Λ◦, ¯c) can be obtained using the above theorem. From now on, B denotes A∞(Λ◦, ¯c), which is also a smooth noncommutative torus for −1/θ [17]. We note that for f, g ∈ S (R) Ahf, gi =Xλ hf, giB =Xλ◦ hf, π(λ)giπ(λ), hπ(λ◦)g, f iπ∗(λ◦), a(λ)π(λ)f for a ∈ l1(Λ), π(a) · f =Xλ f · π(b) = vol(Λ)−1Xλ◦ b(λ◦)π∗(λ◦)f for b ∈ l1(Λ◦). The following theorem shows that Rieffel-type projections are linked to a hard prob- lem in Gabor analysis. Theorem 3.2. Suppose g ∈ S (R). Then G(g, Λ) is a Gabor frame if and only if hg, giB is invertible. Proof. Given an equivalence bimodule Ξ between A and B, we denote by EndA(Ξ) the algebra of module endomorphisms with respect to the action of A on Ξ. It is well known that the equivalence between EndA(Ξ) and B via b 7→ φb where φb(ξ) = ξ ·b for ξ ∈ Ξ. We note that G(g, Λ) is a Gabor frame when ΘΛ g,g ∈ EndA(S (R)) is invertible where ΘΛ g,g(f ) = f ·hg, giB for f ∈ S (R) since hΘΛ Pλ hf, π(λ)gi2 for f ∈ S (R). Thus the invertibility of ΘΛ ibility of hg, giB, and vice versa. g,g(f ), f i =Pλhf, π(λ)gihπ(λ)g, f i = g,g implies that the invert- (cid:3) ON A GAUGE ACTION ON SIGMA MODEL SOLITONS 7 Theorem 3.3. [13, Theorem 3.3] Let G(g, Λ) be a Gabor system on L2(R) with g in S (R). Then the following conditions are equivalent. (i) G(g, Λ) is a tight Gabor frame for L2(R). (ii) G(g, Λ◦) is an orthogonal system. (iii) hg, giB = 1B. (iv) hg, π(λ◦)giL2(R) = vol(Λ)δλ◦,0 for all λ◦ ∈ Λ◦. An important fact for us is that if G(g, Λ) is a Gabor frame for L2(R), then {π(λ◦)g λ◦ ∈ Λ◦} is a Riesz basis for the closed linear span of the set {π(λ◦)g λ◦ ∈ Λ◦}. Moreover, if we take eg = ghg, gi−1/2 interpret Wexler-Raz duality in Gabor analysis in terms of module relations as it appeared in [6] without a proof. , then G(eg, Λ) becomes a tight Gabor frame. We B Theorem 3.4. (7) for f ∈ S (R). f =egheg, f iB Proof. Since the system {π(λ◦)eg λ◦ ∈ Λ◦} is dual to itself, [8, Theorem 1.2.2 p.40] implies that f, h ∈ L2(R). hf, hi =Xλ◦ hf, π(λ◦)egihπ(λ◦)eg, hi, Since π(λ◦)∗ = c(λ◦, λ◦)π(−λ◦) and c(λ◦, λ◦)2 = 1, Xλ◦ hf, π(λ◦)egihπ(λ◦)eg, hi =Xλ◦ =Xλ◦ =Xλ◦ =Xλ◦ = hXλ◦ = heg · heg, f iB, hi. c(λ◦, λ◦)2hf, π(−λ◦)egihπ(−λ◦)eg, hi hf, c(λ◦, λ◦)π(−λ◦)egihc(λ◦, λ◦)π(−λ◦)eg, hi hf, π∗(λ◦)egihπ∗(λ◦)eg, hi hπ(λ◦)f,egiheg, π(λ◦)hi hπ(λ◦)f,egiπ∗(λ◦)eg, hi (cid:3) Recently a large class of functions are proven to be Gabor frames [10]. Recall that a function η is totally positive if for every two sets of increasing real numbers x1 < · · · < xN and y1 < · · · < yN the determinant of the matrix (η(xj − yk)1≤j,k≤N ) is non-negative. A totally positive function η is of finite type M, M ∈ N with M ≥ 2, if its Laplace transform bη has the form: bη(ω) = e−δω2 for real non-zero parameters δj, δ > 0. e−δ0ω MYj=1 1 1 + 2πiδjω 8 HYUN HO LEE Corollary 3.5. [6, Lemma 6.2] Let η be a totally positive function of finite type greater than 2. Then η is a standard module frame for S (R). Passing to a Parseval frame η = η · hη, ηi−1/2 , pη = Ahη, ηi is a Rieffel-type projection in A for 0 < θ < 1. B We remark that such examples include Gaussians, and hyperbolic secants. 4. A group action on solitons A commutative torus T2-action α is defined on Aθ by α(z1,z2)(U1) = z1U1, α(z1,z2)(U2) = z2U2 for (z1, z2) ∈ T2. We denote by ∂1 and ∂2 the infinitesimal generators of each factor of T2 under α [1]. These are unbounded derivations on Aθ, but well defined on Aθ. For ν, µ = 1, 2 ∂ν(Uµ) = 2πiδν,µUµ. Similarly, we have such derivations on B, and use same notations without confusion. Equipped with a B-valued hermitian structure h , iB on Ξ = S (R), we can lift derivations to covariant derivatives ∇1, ∇2 on Ξ given by (∇1ξ)(t) = 2πit θ ξ(t) and (∇2ξ)(t) = ξ′(t). Then as proved in [3] we have the (right) Leibnitz rule for both covariant derivatives: (8) ∇ν(ξ · b) = (∇νξ) · b + ξ · (∂νb) for ν = 1, 2, and compatibility with the hermitian structure: (9) ∂ν(hξ1, ξ2iB) = h∇νξ1, ξ2iB + hξ1, ∇νξ2iB for ν = 1, 2. We introduce complex derivations ∂ = ∂1 + i∂2 and ∂ = ∂1 − i∂2. Accordingly, we introduce the anti-holomorphic connection ∇ = ∇1 + i∇2 and the holomorphic one ∇ = ∂1 − i∂2. Then using linearity (8) and (9) hold for ∇(∇) and ∂(∂) respectively. Lemma 4.1. Let p be a projection in Aθ. Then p(∂νp)p = 0. Proof. Note that ∂ν(p) = (∂νp)p + p(∂νp). By multiplying p both sides, the conclusion follows. Lemma 4.2. Tr(∂νp∂νp) = 2 Tr(p∂νp∂νp) Proof. Tr(∂νp∂νp) = Tr([(∂νp)p + p(∂νp)][(∂νp)p + p(∂νp)]) = Tr((∂νp)p(∂νp)p + p(∂νp∂νp)p + (∂νp)p(∂νp) + (∂νp)p(∂νp)p) =2 Tr(p∂νp∂νp) using Lemma 4.1 and cyclicity of Tr. (cid:3) (cid:3) ON A GAUGE ACTION ON SIGMA MODEL SOLITONS 9 Let us recall a characterization of the minimizing solitons from [4, 5]; let Q(p) be the topological charge or the first Chern number defined by 1 2πi Tr(p[∂1p∂2p − ∂2p∂1p]), which is an integer [1]. Since Tr((∂(p)p)∗(∂(p)p)) ≥ 0 or Tr((∂(p)p)∗(∂(p)p)) ≥ 0, we have, combining (2) with Lemma 4.2, where the equality occurs exactly when the self duality equation S(p) ≥ ±4πQ(p) holds or the anti-self duality equation ∂(p)p = 0 ∂(p)p = 0 holds since Tr is faithful. An important result of [4] is the following observation which lifts a self-duality equation to a linear equation involving the anti-holomorphic connection on the mod- ule. Theorem 4.3. [5, Section 5.3] Let ξ be a standard Parseval frame or hξ, ξiB = 1B. Let pξ = Ahξ, ξi be a projection. Then (∂pξ)pξ = 0 if and only if ∇ξ = ξ · b for some b ∈ B Remark 4.4. In fact, if ∇ξ = ξ · b holds, then b must be hξ, ∇ξiB. Whenever we have a standard module frame η, then by passing from a standard module frame to a Parseval one η one gets a noncommutative soliton. With Theorem 4.11 in mind, we need a slightly stronger form of a linear equation of η than η. From now on, we write A instead of Aθ for the subscript in the operator valued inner product. Theorem 4.5. Suppose that hη, ηiB is invertible. Then let η = ηhη, ηi−1/2 hη, ηi. Then and pη =A B (∂pη)pη = 0 if and only if ∇η = η · b for some b ∈ B Proof. In view of Theorem 4.3, the easier way is to show that the equivalence between ∇η = η · b′ for some b′ ∈ B and ∇η = η · b for some b ∈ B. In the following we give a direct proof. For this, it is better to view pη as Ahη · hη, ηi−1 B i. We abbreviate pη as p without confusion. First note the following cancellation property; B , ηi or Ahη, η · hη, ηi−1 (10) Then Since Ahζ, ηip = Ahζ, ηi ∀ζ. ∂ν(p)p =Ah∇ν(η · hη, ηi−1 =Ah∇ν(η · hη, ηi−1 B ), ηip + Ahη · hη, ηi−1 B ), ηi + Ahηhη, ηi−1 B , ∇νηip B h∇νη, ηiBhη, ηi−1 B , ηi. ∇ν(η · hη, ηi−1 B ) = (∇νη) · hη, ηi−1 B + η∂ν(hη, ηi−1 B ), 10 and HYUN HO LEE ∂ν(hη, ηi−1 B ) = −hη, ηi−1 B (h∇νη, ηiB + hη, ∇νηiB) hη, ηi−1 B , ∂ν(p)p =Ah∇ν(η · hη, ηi−1 =Ah∇νη · hη, ηi−1 =(1 − p)Ah∇νη, η · hη, ηi−1 B ), ηi + Ahηhη, ηi−1 B , ηi − pAh∇νη · hη, ηi−1 B i. B , ηi B h∇νη, ηiBhη, ηi−1 B , ηi Using linearity, it follows that ∂(p)p = 0 if and only if (1 − p)Ah∇η, η · hη, ηi−1 B i = 0. Then if ∇η = η · b for some b ∈ B, (1 − p)Ah∇νη, η · hη, ηi−1 B i =Ahη · b, η · hη, ηi−1 =Ahη · b, η · hη, ηi−1 =0. B i − Ahη · hη, ηi−1, ηiAhη · b, η · hη, ηi−1 B i B i − Ahhη · hη, ηi−1 B hη, η · biB, η · hη, ηi−1 B i Conversely, if ∂(p)p = 0, then (1 − p)Ah∇η, η · hη, ηi−1 0 = ((1 − p)Ah∇η, η · hη, ηi−1 = (1 − p)(Ah∇η, η · hη, ηi−1 = (1 − p)(∇η · hη · hη, ηi−1 B i = 0. It follows that B i) · η B i · η) B , ηiB) Therefore, = (1 − p)∇η. ∇η = p∇η = Ahη · hη, ηi−1 = η · (hη, ηi−1 B , ηi∇η B hη, ∇ηiB) (cid:3) Remark 4.6. If ∇η = η·b, then b must be hη, ηi−1 0 = ∂(hη, ηi−1 B hη, ηiB). From the first equality, B hη, ∇ηiB; note that ∂(hη, ηiBhη, ηi−1 B ) = (11) Thus we have (12) (h∇η, ηiB + hη, ∇ηiB)hη, ηi−1 B + hη, ηiB∂(hη, ηi−1 B ) = 0. ∂(hη, ηi−1 B ) = −hη, ηi−1 B (h∇η, ηiB + hη, ∇ηiB)hη, ηi−1 B . From the second equality, ∂(hη, ηi−1 (13) B )hη, ηiB + hη, ηi−1 B (h∇η, ηiB + hη, ∇ηiB) = 0. In (13), substitute ∂(hη, ηi−1 B ) using (12) and ∇η by η · b in the last term B (h∇η, ηiB + hη, ∇ηiB)hη, ηi−1 B hη, ηiB + hη, ηi−1 B (h∇η, ηiB + hη, η · biB) − hη, ηi−1 = −hη, ηi−1 B hη, ∇ηiB + b = 0. Corollary 4.7. [6, Proposition 6.3] Let η be a Gabor frame in Ξ. Then pη is a solution of the self duality equation. ON A GAUGE ACTION ON SIGMA MODEL SOLITONS 11 Proof. By Wexler-Raz duality (7) for a tight Gabor frame, we have ζ = η · hη, ζiB ∀ζ ∈ Ξ. Since η in Ξ, so is ∇(η). Thus taking ζ as ∇η ∇(η) = ηhη, ηi−1 B hη, ∇ηiB. Then the conclusion follows from Theorem 4.5. (cid:3) Let GL(B) be the set of invertible elements in B. An action of GL(B) on noncom- mutative solitons was introduced in [4] by the right multiplication; for U ∈ GL(B) ξ → ξ · U = ξU . Indeed, if ξ satisfies a self-duality equation, or ∇(ξ) = ξ · b for some b ∈ B, then by the Leibniz rule for the connection, one finds that ξU is the solution of an equation of the form ∇ξU = ξU · bU where (14) bU = U −1bU + U −1∂U. Note that this action preserves the invertibility of hξ, ξiB, thus preserves Gabor frames. Moreover, it is not difficult to check that the Rieffel-type projections are invariant under the action (see [11, p. 232]). When λ is a scalar, i.e., λ ∈ C, ∇η = η · λ has the solutions, the Gaussians of the form Ce−πθt2−2iλt. In [4, 5] it is analyzed when two Gaussian solitons are gauge to each other. Proposition 4.8. Let ξ be a solution of equation with λ ∈ C; and let U ∈ GL(B). Then the transformed λU will be again constant if and only if there exists a pair of integers (m, n) such that Furthermore, U = CmnU m 1 U n 2 . λU − λ = πi(m + ni). In [4, 5] Dabrowski, Krajewski, and Landi suggest the following question; Q: is it possible to gauge a Gaussian soliton to any solution of the self duality equation? In view of Theorem 4.5, this question is equivalent to the following statement: Choose a λ ∈ C, then for any b ∈ B is there an element U ∈ GL(B) such that b = λ + U −1∂U ? This is related to solving inhomogeneous Cauchy-Riemann equation of the form (15) ∂U = U(b − λ). Based on the following Polishchuk's observation, this question is reduced to com- pute the trace of b(see Corollary 4.10). Theorem 4.9. [16, Theorem 3.6], [18, Theorem 6.2] Let τ be the unique trace of the noncommutative torus B. Then for b ∈ B, U −1∂U = b has a nontrivial solution if and only if τ (b) ∈ πi(Z + iZ). 12 HYUN HO LEE Proof. This follows from (1) of [16, Theorem 3.6]. Because of a slightly different notation for ∂ up to the factor 2, the range of τ (U −1∂U) is changed to πi(Z + iZ). (cid:3) Corollary 4.10. Let η(∈ Ξ) be a solution of ∇η = η ·b for some b ∈ B. Then there is a Gaussian ξ and U ∈ GL(B) such that η = ξ · U if and only if λ − τ (b) ∈ πi(Z + iZ) where ∇ξ = ξ · λ for λ ∈ C. It is already mentioned in [11] that the question Q is not true in general as we know a constraint exists. However, in some good cases it is possible to find a non- trivial solution of (15) for some λ ∈ C. In other words, we can gauge a class of noncommutative solitons to Gaussian solitons. Theorem 4.11. Let η be a Gabor frame for Gabor system G(η, Λ) and satisfy ∇η = η · b for some b ∈ B. Then τ (b) = h∇η, ηiL2(R) where h·, ·iL2(R) is the inner product on the Hilbert space L2(R). Proof. By Wexler-Raz biorthogonality the fact that G(η, Λ) is a tight Gabor frame for L2(R) implies that G(η, Λ◦) is an orthogonal system, a Riesz basis for L2(R). Therefore we can expand ∇η in terms of {π(λ◦)η λ◦ ∈ Λ◦}. Write i.e. b =Pλ◦∈Λ◦ b(λ◦)π(λ◦) where b(λ◦)'s are rapidly decreasing. Then b(λ◦)π(λ◦)∗η η · b = Xλ◦∈Λ◦ = Xλ◦∈Λ◦ b(−λ◦)c(λ◦, λ◦)π(λ◦)η Thus the condition ∇η = η · b implies that Xλ◦∈Λ◦ h∇η, π(λ◦)ηiL2(R)π(λ◦)η = Xλ◦∈Λ◦ b(−λ◦)c(λ◦, λ◦)π(λ◦)η. It follows that from the orthogonality of {π(λ◦)η λ◦ ∈ Λ◦} h∇η, π(λ◦)ηiL2(R) = b(−λ◦)c(λ◦, λ◦). Hence τ (b) = b(0, 0) = h∇η, ηiL2(R). Proposition 4.12. Let η be the hyperbolic secant of the form (cid:16)π 2(cid:17) 1 2 τ (b) = 0 where ∇η = η · b, so that the Gaussians ξ associated with ∇ξ = ξ · λ for λ ∈ πi(Z + iZ) are gauged to η. Moreover, if the invertible W such that η = ξ · W cannot be of the form U mV n (modulo T ) where U and V are generators of B. (cid:3) 1 cosh(πt) . Then Proof. η is a Gabor frame due to Janssen and Strohmer and pη belongs to B by [13, Theorem 3.6]. Note that ∇η(t) = iπ(cid:18)2t θ − tanh(t)(cid:19) η(t) up to a constant. Therefore h∇η, ηiL2(R) = 2πiZ ∞ −∞ tη2(t)dt − iπZ ∞ −∞ tanh(t)η2(t)dt. Since η is an even function and both t and tanh(t) are odd functions, two terms vanish by the definition of the Lebesgue integral. Thus τ (b) = 0. Since η is a Gabor frame in Ξ, then it satisfies ∇η = η · b for some b. Moreover, if η = ξ · W , then ON A GAUGE ACTION ON SIGMA MODEL SOLITONS 13 τ (b) = λ + τ (W −1∂W ) by (14). Thus we must have λ = −τ (W −1∂W ) ∈ πi(Z + iZ). The last statement follows from the fact that two Gaussians are gauge equivalent if and only if one is gauged to the other via U mV n(for some m, n) only up to constants [4]. Remark 4.13. The second statement in Proposition 4.12 is related to the author's question in [12] if any two solutions of (3) are equivalent under a Z2-action which is defined by an inner automorphisms Ad W where W is a unitary of the form U mV n modulo T. The answer is no as we see from the above example. (cid:3) 5. Acknowledgements The author would like to thank F. Luef for a series of lectures of his work during his visit to Korea. He also would like to express his gratitude to Hun Hee Lee for a final tip in proving Theorem 4.11. I References [1] A. Connes, C ∗-alg`ebres et g´eometrie diff´erentille, C.R. Acad. Sci. Paris S´er. A 290 (1980), no. 13, 599 -- 604.MR1690050(81c:46053) [2] F. Boca, Projections in rotation algebras and theta functions, Comm. Math. Phys. 202 (1999), no. 2, 325 -- 357.MR1690050(2000j:46101) [3] A. Connes and M. Rieffel, Yang-Mills for noncommutative two-tori, Contemp. Math. 62 (1987), 335 -- 348. MR454645 (56#:12894) [4] L. Dabrowski, T. Krajewski, and G. Landi, Some properties of Non-linear σ-models in noncom- mutative geometry, Int. J. Mod. Phys. B14 (2000), 2367 -- 2382. MR0470685 (57 #10431) [5] , Non-linear σ-models in noncommutative geometry: fields with values in finite spaces, Mod. Physics Lett. A 18 (2003), 2371 -- 2379. [6] L Dabrowski, G. Landi, and F Luef, Sigma-model solitons on noncommutative spaces, Lett. Math. Phys. 105 (2015), no. 12, 1633 -- 1688, DOI 10.1007/s11005-015-0790-x. MR3420593 [7] G. Elliott and D. Evans, The structure of the irrational rotation C ∗-algebra, Ann. of Math. (2)138 (1993), no. 3, 477-501. MR1247990 (94j:46066) [8] H Feichtinger and Strohmer, Gabor analysis and Algorithms, Springer Science+Business Media, LLC, 1998. [9] M. Frank and D. Larson, Frames in Hilbert C ∗-modules and C ∗-algebras, J. Operator Theory (2002), 273 -- 314. [10] K. Grochenig and Y. Lyubarskii, Gabor (super)frames and totally positive functions, Duke Math. J 162 (2013), no. 5, 1003 -- 1031. [11] G. Landi, On harmonic maps in noncommutative geometry, Non-commutative Geometry and Number Theory Springer (2006), 217 -- 234. [12] H. Lee, A note on nonlinear σ-models in noncommutative geometry, IDAQP 19 (2016), no. 1, DOI 10.1142/S0239025716500065. MR2733573 (2011k:46079) [13] F. Luef, Projections in noncommutative tori and Gabor frames, Proc. A.M.S. 139 (2010), no. 2, 571 -- 582. [14] F. Luef, Projective modules over noncommutative tori are multi-window Gabor frames for mod- ulation spaces, J. Funct. Anal. 257 (2009), no. 6, 1921 -- 1946. MR2540994 [15] V. Mathai and J. Rosenberg, A noncommutative sigma-model, J. Noncommut. Geom. 5 (2011), 265 -- 294. [16] P. Polishchuck, Analogues of the exponential map associated with complex structures on non- commutative two-tori, Pacific J. Math. 226 (2006), no. 1, 153 -- 178. [17] M. Rieffel, C ∗-algebras associated with irrational rotations, Pacific. J. Math. 93 (1981), no. 1, 415 -- 429. MR623572(83b:46087) 14 HYUN HO LEE [18] J. Rosenberg, Noncommutative variations on Laplace equation, Anal. PDE. 1 (2008), no. 1. MR2444094 (2009f:58041) Department of Mathematics, University of Ulsan, Ulsan, South Korea 44610 E-mail address: [email protected]
1808.09647
1
1808
2018-08-29T05:45:10
The Toeplitz algebra has nuclear dimension one
[ "math.OA" ]
We prove the title by constructing 2-colourable completely positive approximations for the Toeplitz algebra. Besides results about nuclear dimension and completely positive contractive order zero maps, our argument involves projectivity of the cone over a finite dimensional C*-algebra and Lin's theorem on almost normal matrices.
math.OA
math
THE TOEPLITZ ALGEBRA HAS NUCLEAR DIMENSION ONE LAURA BRAKE AND WILHELM WINTER Abstract. We prove the title by constructing 2-colourable completely pos- itive approximations for the Toeplitz algebra. Besides results about nuclear dimension and completely positive contractive order zero maps, our argument involves projectivity of the cone over a finite dimensional C∗-algebra and Lin's theorem on almost normal matrices. Nuclear dimension is a notion of covering dimension for amenable C∗-algebras; it was introduced in [12] and, in contrast to the decomposition rank of [5], may take finite values also for infinite C∗-algebras. The most basic example of an infinite C∗-algebra is the Toeplitz algebra T , which may be thought of as the C∗-algebra generated by the unilateral shift on ℓ2(N); it is an extension of the continuous functions on the circle by the compact operators on ℓ2(N): 0 −→ K −→ T −→ C(S1) −→ 0 Finiteness of nuclear dimension is preserved under taking extensions, and it was ob- served already in [12, Example 6.3] that the Toeplitz algebra has nuclear dimension either 1 or 2. The exact value remained unknown; cf. [12, Problem 9.2]. Before settling the matter, let us explain why we even care. The arguably most striking appearance of nuclear dimension was in Elliott's classification programme for simple nuclear C∗-algebras; see [10] for an overview. In this context, there now is a result as complete as can be, and the crucial hypothesis is just finiteness of nuclear dimension, with no reference to the actual value. In hindsight the latter is not too surprising, since we now know that for separable, simple, unital C∗-algebras the only possible values for the nuclear dimension are 0, 1, and infinity. On the other hand, in the nonsimple case the nuclear dimension potentially carries more information (after all, for continuous trace C∗-algebras it coincides on the nose with covering dimension of the spectrum). But even then one could argue why the exact value is relevant for an example which is otherwise as well understood as the Toeplitz algebra. Indeed, we do not so much expect new insights in the structure of said algebra, but our construction is a new instance of dimension reduction phenomena in noncommutative topology. Dimension reduction occurs in its most simple form in the commutative setting, where dimension is additive, and not multiplicative, under taking products: for example, an interval may be approximated by open covers with two colours (where members of the cover with the same colour are not allowed to intersect), and hence has dimension at most one. (Showing that it is not zero, i.e. finding a lower bound, is another matter.) But then the square, being a product of two intervals, can obviously be approximated by open covers with four colours, hence has dimension at most three. To improve this and find approximations with three colours (and Supported by the Deutsche Forschungsgemeinschaft through SFB 878. 1 2 LAURA BRAKE AND WILHELM WINTER to arrive at dimension two, as it should be) one has to employ (an ever so small amount of) geometric insight and come up, for instance, with a bricklaying scheme. Nuclear dimension expresses all this in terms of functions, with open covers replaced by partitions of unity and Cartesian products by tensor products. For commuta- tive C∗-algebras nothing changes, but the noncommutative situation becomes more complicated. On the one hand, we do not know in general whether nuclear dimen- sion is subadditive with respect to tensor products. On the other hand, there are several situations where dimension reduces when there is enough 'noncommutative space' available, in particular [3, 4, 9, 7, 8, 1]. In all of these, dimension reduction is quite dramatic, and eventually relies (more or less explicitly) on the existence of simple -- in most cases, strongly self-absorbing -- local tensor factors, like the Cuntz algebras O2 and O∞, UHF algebras, or the Jiang -- Su algebra Z. It is not at all obvious why dimension reduction should be possible for the Toeplitz algebra, which has no local tensor product decomposition. The argument of [12, Proposi- tion 2.9] yields an approximation of T by noncommutative partitions of unity with three colours, hence showing that the nuclear dimension is at most two. (Of these three colours, two come from approximations of the circle; the third comes from the compacts.) In this paper we show that these approximations can be modified and repackaged so that only two colours suffice. One can think of our method as a noncommutative (and 2-coloured) bricklaying scheme; for this to work we will have to use the little available noncommutative space most efficiently. The key will be Lin's theorem on almost normal elements of matrix algebras; cf. [6, The Main Theorem]. The remainder of the paper is organised as follows: We first recall the definition of nuclear dimension from [12]. Then we apply the definition to the circle, both to make it explicit and to settle notation. Afterwards we isolate a lemma which encapsulates the additional geometric insight that makes everything work. Finally, we prove our result by exhibiting a system of approximations for the Toeplitz alge- bra that shows its nuclear dimension to be one. Definition. The nuclear dimension of a C∗-algebra A is the smallest natural num- ber d so that the following holds: There exists a net (Fλ, ψλ, ϕλ)λ∈Λ such that the Fλ are finite-dimensional C∗-algebras, and such that ψλ : A → Fλ and ϕλ : Fλ → A are completely positive maps satisfying (i) kϕλ ◦ ψλ(a) − ak −→ 0 for all a ∈ A, (ii) kψλk ≤ 1, (iii) for each λ, Fλ decomposes into d + 1 ideals Fλ = F (0) λ := ϕλF (l) that ϕ(l) for l = 0, ..., d, where order zero means that ϕ(l) elements to pairs of orthogonal elements. λ such is a completely positive contractive order zero map λ ⊕ . . . ⊕ F (d) λ λ sends pairs of orthogonal Example. The nuclear dimension of A = C(S1) is one, and can be realised by completely positive approximations coming from '2-coloured' partitions of unity. We set these up explicitly, as we will need them later on in the paper. THE TOEPLITZ ALGEBRA HAS NUCLEAR DIMENSION ONE 3 For 0 6= k ∈ N, j = 1, . . . , k and l = 0, 1 set δ(l) ψk : C(S1) → Ck ⊕ Ck by a sum of point evaluations, j,k := e2πi 2j+l 2k ∈ S1. Define ψk := ψ(0) k ⊕ ψ(1) k := (cid:0)Lk j=1 evδ(0) j,k(cid:1) ⊕ (cid:0)Lk j,k(cid:1). j=1 evδ(1) Again for 0 6= k ∈ N, j = 1, . . . , k and l = 0, 1 take a partition of unity for S1 by sawtooth functions {¯e(l) j,k}j∈{1,...k}, l∈{0,1} as depicted below: ¯e(1) j−1,k ¯e(0) j,k ¯e(1) j,k δ(0) j−1,k δ(1) j−1,k δ(0) j,k δ(1) j,k δ(0) j+1,k Define completely positive contractions ϕ(0) ϕ(l) k (a1, . . . , ak) := Pk and ϕk : Ck ⊕ Ck → C(S1) by ϕk := ϕ(0) k are order zero, i.e., preserve orthogonality. With this setup the sequence (ϕkψk)k converges to the identity map on C(S1) in point-norm topology. k ; note that both maps ϕ(l) k , ϕ(1) j=1 aj ¯e(l) k : Ck → C(S1) by j,k for l = 0, 1, k + ϕ(1) The lemma below is key for the proof of our main result. It relies on two nontrivial facts and a beautiful theorem. The first fact is [11, Corollary 3.1]: order zero contractions out of a C∗-algebra A are in bijection with ∗-homomorphisms out of C0((0, 1]) ⊗ A, the cone over A. The second fact is that the cone over C(S1), which is isomorphic to C0(D\{0}) (with D the unit disc), is canonically isomorphic to the universal C∗-algebra generated by a normal contraction. The theorem we will be using is Lin's result on almost normal matrices; it says that for every η > 0 there is a δ > 0 such that, whenever x is a square matrix with kxx∗ −x∗xk < δ, there is a square matrix y such that yy∗ −y∗y = 0 and kx−yk < η. The point is of course that δ depends only on η, and not on the matrix size. This was shown by Lin in [6]; an alternative proof was given by Friis and Rørdam in [2]. Lemma. Let (Bn)n∈N be a sequence of finite-dimensional C∗-algebras and let β : C(S1) −→ Qn∈N Bn.Pn∈N Bn be a completely positive contractive order zero map. Then β has a completely posi- tive contractive order zero lift ¯β : C(S1) −→ Qn∈N Bn. Proof. Let u ∈ C(S1) be the identity function. By [11, Corollary 3.1], β induces a ∗-homomorphism π : C0((0, 1]) ⊗ C(S1) −→ Qn∈N Bn.Pn∈N Bn with π(id(0,1] ⊗u) = β(u). In particular we see that β(u) is normal. Let (vn)n∈N ∈ Qn∈N Bn be a contractive lift of β(u). Note that each vn is a direct sum of matrices and that kvnv∗ nvnk converges to 0 as i goes to infinity. But then it follows from n − v∗ 4 LAURA BRAKE AND WILHELM WINTER Lin's theorem that there is a sequence (¯vn)n∈N ∈ Qn∈N Bn of normal contractions such that k¯vn − vnk → 0, i.e., (¯vn)n∈N is a normal contractive lift of β(u). Upon identifying C0((0, 1]) ⊗ C(S1) with C0(D \ {0}) and regarding the latter as the universal C∗-algebra generated by a normal contraction, we obtain a ∗- homomorphism with ¯π(id(0,1] ⊗u) = (¯vn)n∈N. Note that this implies that ¯π lifts π. Now define ¯π : C0((0, 1]) ⊗ C(S1) −→ Qn∈N Bn ¯β( . ) := ¯π(id(0,1] ⊗ . ) : C(S1) −→ Qn∈N Bn, then ¯β is the composition of a ∗-homomorphism with an order zero map, hence itself an order zero map. Moreover, since ¯π lifts π, ¯β lifts β. (cid:3) Theorem. The Toeplitz algebra has nuclear dimension one. Proof. First, we choose a quasicentral idempotent approximate unit h = (hn)n∈N for K ⊳ T so that every hn has finite rank (here, 'idempotent' means that hn+1hn = hn for all n ∈ N). Let h = (hn)n∈N be essentially the same approximate unit as h, but with index shifted by one (and with h0 dropped). Upon regarding h and h as elements in Qn∈N T , we have h · h = h. To find approximations for T , we work with the following diagram: ι qC Q Cn Q Cn(cid:30)P Cn ι4 ι3 Ck ⊕ Ck ψk ϕk γ T π µ C(S1) ψ(0) k ¯β Ck ρk q Q T Q T(cid:30)P T ι2 β α qB Q Bn Q Bn(cid:30)P Bn ι1 Q An Here π is the quotient map from the short exact sequence, and µ is a completely positive contractive lift for π (for example map u ∈ C(S1) to the bilateral shift on ℓ2(Z) and compress to ℓ2(N)). Moreover, we let An = hnT hn, Bn = (cid:0)hn − hn(cid:1)T (cid:0)hn − hn(cid:1), Cn = (cid:0)1 − hn(cid:1)T (cid:0)1 − hn(cid:1) THE TOEPLITZ ALGEBRA HAS NUCLEAR DIMENSION ONE 5 for all n ∈ N, and n(cid:1)n∈N, 1 n xh 2 1 2 α : x 7−→ (cid:0)h β : f 7−→ (cid:0)(hn − hn) γ : f 7−→ (cid:0)(1 − hn) 1 1 2 µ(f )(hn − hn) 1 2(cid:1)n∈N, 2 µ(f )(1 − hn) 1 2(cid:1)n∈N. Let q, qB, qC be the canonical quotient maps, let ι1, ι2, ι3, ι4 be the natural inclusion maps, and let ι be the canonical constant sequence embedding. The map ¯β is a completely positive contractive order zero lift of the order zero map qB ◦ β; this is where our lemma above (and hence Lin's theorem) enters. Before explaining the dotted arrows, note that with the setup so far the map ι is precisely the sum of the lower three paths from left to right through the diagram, (1) ι = q ◦ ι1 ◦ α + ι2 ◦ qB ◦ ¯β ◦ π + ι3 ◦ qC ◦ γ ◦ π. For each 0 6= k ∈ N, let C(S1) ψk−→ Ck ⊕ Ck ϕk−→ C(S1) k + ϕ(1) k ⊕ ψ(1) k and ϕk = ϕ(0) with ψk = ψ(0) tions described in our example above. The definition of the maps ρk in the diagram above is more complicated, but it is at the heart of the matter. We define them component-wise, working with the diagram below: k be the completely positive approxima- ¯βn ´βn Bn βn C0(D \ {0}) ηn C0(σ(¯vn) \ {0}) C(S1) τ ψ(0) k ϕk,n ρk,n Ck Here, τ is the canonical order zero map which sends the identity function u ∈ C(S1) to the identity function on D (this is just the map (f 7→ id(0,1] ⊗f ) followed by the identification C0((0, 1]) ⊗ C(S1) ∼= C0(D \ {0}) given by (id(0,1] ⊗ u 7−→ idD). We write ¯βn for the (order zero) components of ¯β. With the notation of our lemma, we have normal elements ¯βn(u) = ¯vn, n ∈ N. Upon identifying C0(D \ {0}) with the universal C∗-algebra generated by a normal contraction, we obtain the canonically induced ∗-homomorphisms ´βn, which canonically factorise through continuous functions on the spectrum of ¯vn via the ∗-homomorphisms ηn and βn. 6 LAURA BRAKE AND WILHELM WINTER For the definition of the maps ϕk,n and ρk,n we write the pointed disc D \ {0} as the disjoint union of 'pizza slices' Sj,k := (cid:8)re2πit 2j−1 2k ; 0 < r ≤ 1(cid:9) for j = 1, . . . , k. 2k < t ≤ 2j+1 Let χSj,k be the characteristic functions of these sets and define Borel functions ´χj,k : D −→ C by ´χj,k(z) := χSj,k (z) · z. Now since for each n the spectrum of the element ¯vn ∈ Bn is discrete, the ´χj,k restrict to continuous functions on σ(¯vn) and we may define completely positive contractive order zero maps ϕk,n : Ck −→ C0(σ(¯vn) \ {0}) by ϕk,n : (a1, . . . , ak) 7−→ Pk j=1 aj · ´χj,kσ(¯vn)\{0}. ρk,n := βn ◦ ϕk,n We then define and compute ρk,n ◦ ψ(0) k (f ) = βn(cid:0)Pk j=1 f (δ(0) j,k ) · ´χj,kσ(¯vn)\{0}(cid:1) = βn ◦ ηn(cid:0)Pk j=1 f (δ(0) j,k ) · ´χj,k(cid:1) for f ∈ C(S1). Here, the δ(0) j,k are as in our example above, and we have we have tacitly extended the restriction ∗-homomorphism ηn to all bounded Borel functions on D. Now observe that for each f ∈ C(S1) we have whence (cid:13)(cid:13)ρk,n ◦ ψ(0) k (f ) − ¯βn(f )(cid:13)(cid:13) = (cid:13)(cid:13) j=1 f (δ(0) (cid:13)(cid:13)Pk k→∞−→ 0, j,k ) · ´χj,k − τ (f )(cid:13)(cid:13)∞,D j=1 f (δ(0) βn ◦ ηn(cid:0)Pk j=1 f (δ(0) j,k ) · ´χj,k(cid:1) − τ (f )(cid:13)(cid:13)∞,D ≤ (cid:13)(cid:13)(cid:0)Pk k→∞−→ 0 j,k ) · ´χj,k(cid:1) − βn ◦ ηn ◦ τ (f )(cid:13)(cid:13) for each f ∈ C(S1). Note that convergence here does not depend on n. As a consequence, we see that the maps ρk ◦ ψ(0) converge to ¯β in point-norm topology, k (cid:13)(cid:13)ρk ◦ ψ(0) k (f ) − ¯β(f )(cid:13)(cid:13) k→∞−→ 0 for every f ∈ C(S1). Upon plugging this into (1) we obtain for every x ∈ T k→∞−→ 0. (2) (cid:13)(cid:13)ι(x) − q ι1 α(x) + ι2 qB ρk ψ(0) Next, we check that for each k the completely positive contractive map k π(x) + ι3 qC γ ϕk ψk π(x)(cid:13)(cid:13) k : Ck −→ Q T (cid:14)P T ι2 ◦ qB ◦ ρk + ι3 ◦ qC ◦ γ ◦ ϕ(0) is order zero. Since the two summands are each order zero, we only need to examine the mixed terms, i.e., we have to confirm that (3) ι2 ◦ qB ◦ ρk(ej) · ι3 ◦ qC ◦ γ ◦ ϕ(0) k (ej ′ ) = 0 for 1 ≤ j 6= j′ ≤ k, where we write ej for the canonical jth generator of Ck. The reason why (2) holds is that our 'pizza slices' ´χj,k and the τ (¯e(0) j ′ ,k) are orthogonal functions over D whenever j 6= j′. To be more precise, let (dm)m∈N ⊂ C(S1) be normalised functions such THE TOEPLITZ ALGEBRA HAS NUCLEAR DIMENSION ONE 7 that dm(e2πit) = 1 for all 2j−1 uniformly as m goes to infinity. We then have ´χj,k ≤ τ (dm), whence 2k < t ≤ 2j+1 and such that ¯e(0) 2k j ′,k · dm goes to zero ρk,n(ej) = βn ◦ ϕk,n(ej) = βn( ´χj,kσ(¯vn)\{0}) ≤ βn ◦ ηn ◦ τ (dm) = ¯βn(dm) for each m and n. It follows that qB ◦ ρk(ej) ≤ qB ◦ ¯β(dm) = qB ◦ β(dm) for each m, and therefore 2 1 1 k (ej ′ ) · ι2 ◦ qB ◦ ρk(ej)(cid:13)(cid:13) 2 · (h − h) j ′ ,k)) · ι(µ(dm))(cid:13)(cid:13) (cid:13)(cid:13)ι3 ◦ qC ◦ γ ◦ ϕ(0) ≤ (cid:13)(cid:13)ι3 ◦ qC ◦ γ ◦ ϕ(0) 2 ι(µ(¯e(0) ≤ (cid:13)(cid:13)(1 − h) = (cid:13)(cid:13)(1 − h)(h − h)ι(µ(¯e(0) ≤ (cid:13)(cid:13)ι(µ(¯e(0) m→∞−→ 0, j ′ ,k · dm))(cid:13)(cid:13) k (ej ′ ) · ι2 ◦ qB ◦ β(dm) · ι3 ◦ qC ◦ γ ◦ ϕ(0) k (ej ′ )(cid:13)(cid:13) 2 ι(µ(dm))(h − h) j ′ ,k))(1 − h) 2(cid:13)(cid:13) 1 1 which entails (3). Next, recall that by [5, Remark 2.4] order zero maps out of finite-dimensional C∗- algebras into quotient C∗-algebras always lift to order zero maps, and so there are completely positive contractive order zero lifts ϕ(1) k : Ck −→ Q Cn of qC ◦ γ ◦ ϕ(1) k , and ϕ(0) k , fitting into the following diagram: k : Ck −→ Q T of ι2 ◦ qB ◦ ρk + ι3 ◦ qC ◦ γ ◦ ϕ(0) Ck ψ(1) k π T C(S1) ι ϕ(1) k Q Cn ι4 ϕ(0) k q Q T Q T(cid:30)P T ψ(0) k Ck α ι1 We already know from (2) that for every x ∈ T Q An k ◦ ψ(0) (cid:13)(cid:13)ι(x) − q(cid:0)ι1 ◦ α(x) + ϕ(0) n→∞(cid:13)(cid:13)x − ι1,n ◦ αn(x) + ϕ(0) lim lim k→∞ expressing the quotient norm as a limit we obtain for every x ∈ T k,n ◦ ψ(0) k ◦ π(x) + ι4,n ◦ ϕ(1) k ◦ π(x) + ι4 ◦ ϕ(1) k ◦ ψ(1) k→∞−→ 0; k ◦ π(x)(cid:1)(cid:13)(cid:13) k,n ◦ ψ(1) k ◦ π(x)(cid:13)(cid:13) = 0. 8 LAURA BRAKE AND WILHELM WINTER This remains true for finite sets of elements x, and using separability of T we run a diagonal sequence argument to find, for every index k, some nk such that for x ∈ T lim k→∞(cid:13)(cid:13)x − ι1,nk ◦ αnk (x) + ϕ(0) Now for each 0 6= k ∈ N we set k,nk ◦ ψ(0) k ◦ π(x) + ι4,nk ◦ ϕ(1) k,nk ◦ ψ(1) k ◦ π(x))(cid:13)(cid:13) = 0. k k := Ck, k = Ck ⊕ Ck ⊕ Ank , := Ck ⊕ Ank , k ⊕ F (1) k ◦ π + ψ(1) nk : Ck −→ T , F (0) F (1) Fk := F (0) ψk := ψ(0) ϕ(0) := ϕ(0) k ϕ(1) := ι4,nk ◦ ϕ(1) C,nk k k ⊕ ϕ(1) ϕk := ϕ(0) k : Ck ⊕ Ck ⊕ Ank −→ T ; k ◦ π + αnk : T −→ Ck ⊕ Ck ⊕ Ank , ⊕ ι1,nk : Ck ⊕ Ank −→ T , is a completely positive order zero contraction since each of the two note that ϕ(1) k summands is, and since they map into the orthogonal algebras Cnk and Ank . The system ( Fk, ψk, ϕk)k∈N\{0} indeed establishes that the Toeplitz algebra has nuclear dimension one. (cid:3) References 1. Joan Bosa, Nathanial P. Brown, Yasuhiko Sato, Aaron Tikuisis, Stuart White, and Wilhelm Winter, Covering dimension of C∗-algebras and 2-coloured classification, arXiv:1506.03974v2; to appear in Mem. Amer. Math. Soc., 2015. 2. Peter Friis and Mikael Rørdam, Almost commuting self-adjoint matrices -- a short proof of Huaxin Lin's theorem, J. Reine Angew. Math. 479 (1996), 121 -- 131. 3. Guihua Gong, On the classification of simple inductive limit C∗-algebras. I. The reduction theorem., Doc. Math. 7 (2002), 255 -- 461. 4. Eberhard Kirchberg and Mikael Rørdam, Purely infinite C∗-algebras: ideal-preserving zero homotopies, Geom. Funct. Anal. 15 (2005), no. 2, 377 -- 415. 5. Eberhard Kirchberg and Wilhelm Winter, Covering dimension and quasidiagonality, Internat. J. Math. 15 (2004), no. 1, 63 -- 85. 6. Huaxin Lin, Almost commuting selfadjoint matrices and applications, Operator algebras and their applications (Waterloo, ON, 1994/1995), Fields Inst. Commun., vol. 13, Amer. Math. Soc., Providence, RI, 1997, pp. 193 -- 233. 7. Hiroki Matui and Yasuhiko Sato, Decomposition rank of UHF-absorbing C∗-algebras, Duke Math. J. 163 (2014), no. 14, 2687 -- 2708. 8. Yasuhiko Sato, Stuart White, and Wilhelm Winter, Nuclear dimension and Z-stability, Invent. Math. 202 (2015), 893 -- 921. 9. Aaron Tikuisis and Wilhelm Winter, Decomposition rank of Z-stable C∗-algebras, Analysis & PDE 7 (2014), 673 -- 700. 10. Wilhelm Winter, From quasidiagonality to classification, and back again, arXiv preprint 1712.00247; to appear in Proceedings of the ICM 2018, 2017. 11. Wilhelm Winter and Joachim Zacharias, Completely positive maps of order zero, Munster J. Math. 2 (2009), 311 -- 324. 12. , The nuclear dimension of C∗-algebras, Adv. Math. 224 (2010), no. 2, 461 -- 498. E-mail address: [email protected] E-mail address: [email protected] Mathematisches Institut der WWU Munster, Einsteinstr. 62, 48149 Munster, Germany
1605.08932
3
1605
2018-08-08T14:51:59
Ueda's peak set theorem for general von Neumann algebras
[ "math.OA", "math-ph", "math.FA", "math.LO", "math-ph" ]
We extend Ueda's peak set theorem for subdiagonal subalgebras of tracial finite von Neumann algebras, to sigma-finite von Neumann algebras (that is, von Neumann algebras with a faithful state; which includes those on a separable Hilbert space, or with separable predual.) To achieve this extension completely new strategies had to be invented at certain key points, ultimately resulting in a more operator algebraic proof of the result. Ueda showed in the case of finite von Neumann algebras that his peak set theorem is the fountainhead of many other very elegant results, like the uniqueness of the predual of such subalgebras, a highly refined F and M Riesz type theorem, and a Gleason-Whitney theorem. The same is true in our more general setting, and indeed we obtain a quite strong variant of the last mentioned theorem. We also show that set theoretic issues dash hopes for extending the theorem to some other large general classes of von Neumann algebras, for example finite or semi-finite ones. Indeed certain cases of Ueda's peak set theorem, for a von Neumann algebra M, may be seen as `set theoretic statements' about M that require the sets to not be `too large'.
math.OA
math
UEDA'S PEAK SET THEOREM FOR GENERAL VON NEUMANN ALGEBRAS DAVID P. BLECHER AND LOUIS LABUSCHAGNE Abstract. We extend Ueda's peak set theorem for subdiagonal subalgebras of tracial finite von Neumann algebras, to σ-finite von Neumann algebras (that is, von Neumann algebras with a faithful state; which includes those on a separable Hilbert space, or with separable predual.) To achieve this extension completely new strategies had to be invented at certain key points, ultimately resulting in a more operator algebraic proof of the result. Ueda showed in the case of finite von Neumann algebras that his peak set theorem is the fountainhead of many other very elegant results, like the uniqueness of the predual of such subalgebras, a highly refined F & M Riesz type theorem, and a Gleason-Whitney theorem. The same is true in our more general setting, and indeed we obtain a quite strong variant of the last mentioned theorem. We also show that set theoretic issues dash hopes for extending the theorem to some other large general classes of von Neumann algebras, for example finite or semi-finite ones. Indeed certain cases of Ueda's peak set theorem, for a von Neumann algebra M , may be seen as 'set theoretic statements' about M that require the sets to not be 'too large'. 1. Introduction In a series of papers the authors extended most of the theory of generalized H p spaces for function algebras from the 1960s to the setting of Arveson's (finite maximal) subdiagonal algebras. Most of this is summarized in the survey [8]. We worked in the setting where the subdiagonal algebra A was a unital weak* closed subalgebra of a von Neumann algebra M , and where M possesses a faithful normal tracial state. Ueda followed this work in [41] by removing a hypothesis involv- ing a dimensional restriction on A ∩ A∗ in four or five of our theorems, and also establishing several other beautiful results such as the fact that such an A has a unique predual, all of which followed from his very impressive noncommutative (Amar-Lederer) peak set type theorem. (We will say more about peak sets and peak projections later in this introduction when we describe notation and technical Date: June 28, 2021. 2010 Mathematics Subject Classification. 46L51, 46L52, 47L75, 47L80 (primary), 03E10, 03E35, 03E55, 46J15, 46K50, 47L45 (secondary). Key words and phrases. Subdiagonal operator algebra, peak projection, noncommutative Lebesgue decomposition, noncommutative Hardy space, sigma-finite von Neumann algebra, Ka- plansky density theorem, F & M Riesz theorem. This work is based on research supported by the National Research Foundation (IPRR Grant 96128), and the National Science Foundation. Any opinion, findings and conclusions or recom- mendations expressed in this material, are those of the author, and therefore the NRF do not accept any liability in regard thereto. 1 2 DAVID P. BLECHER AND LOUIS LABUSCHAGNE background. Also Section 2 of the paper is devoted to general results about peak projections, for example giving some useful characterizations of peak projections in C∗-algebras, von Neumann algebras, and general operator algebras that do not seem to appear explicitly in the literature.) Ueda's peak set result may be phrased as saying that the support projection in M ∗∗ of a singular state ϕ on M is dom- inated by a peak projection p for A (so ϕ(p) = 1) with p in the 'singular part' of M ∗∗ (that is, p annihilates all normal functionals on M ). With the theory of subdiagonal subalgebras of von Neumann algebras with a faithful normal tracial state reaching a level of maturity, several authors turned their attention to the more general σ-finite von Neumann algebras. Important structural results were obtained by Ji, Ohwada, Saito, Bekjan, and Xu [28, 29, 43, 26, 27, 4]. Recently in [34] the second author used Haagerup's reduction theory [21] to make several significant advances in generalizing aspects of the earlier theory to the σ- finite case, most notably the Beurling invariant subspace theory. The present work flowed out of this, being a direct continuation of the line of attack in [34]. Here we extend Ueda's peak set theorem, and its corollaries, to maximal subdiagonal algebras A in more general von Neumann algebras M , thereby demonstrating that such algebras too for example have a unique predual, admit a highly refined F & M. Riesz type theorem, have a powerful Gleason-Whitney theorem (in particular, every normal functional on A has a unique Hahn-Banach extension to M , and this extension is also normal), etc. We remark that special cases of two of these results were obtained under an additional semi-finite hypothesis in [42]. The technically difficult extension of Ueda's theorem to the general σ-finite case is found in Section 4, while the applications mentioned a few lines back are discussed in Section 5. In Section 3 we establish some Kaplansky density type results for operator spaces and subdiagonal algebras which we need. In Section 6 we discuss the extent to which Ueda's theorem might be generalized beyond the σ-finite case. There is some limited good news: our results will have variants valid for any von Neumann algebra under an appropriate condition on its center, since it is known that any von Neumann algebra is a direct sum of algebras of the form Mi = Ri ¯⊗B(Ki) for σ-finite von Neumann algebras Ri. The central projections ei corresponding to this direct sum will sometimes allow a decompo- sition of a maximal subdiagonal algebra A of M as a direct sum of subalgebras Ai of the Ri, and it is easy to see that then these are maximal subdiagonal sub- algebras of the σ-finite algebras Ri. The 'bad news' is that there is little hope of proving Ueda's theorem in ZFC for all von Neumann algebras, or even for com- mutative (and hence finite) or semi-finite von Neumann algebras. Indeed we show that the validity Ueda's theorem for commutative atomic von Neumann algebras is a stronger statement than (it would imply) a ZFC proof of the nonexistence of uncountable measurable cardinals, a famous problem in set theory which nobody today seems to believe is solvable. Indeed certain cases of Ueda's peak set theorem, for a von Neumann algebra M , may be seen as 'set theoretic statements' about M that require the sets to not be 'too large'. These issues are discussed in Section 6, and this also led to a sequel paper with Nik Weaver [13]. Some of the ramifica- tions of [13] are described at the end of the present paper, for example that that work indicates that one cannot generalize Ueda's peak set theorem in ZFC much beyond the σ-finite case (not even to l∞(R)). Thus the main result of our paper is somewhat sharp. UEDA'S THEOREM 3 We now turn to our set-up, background, and notation. We recall that a σ-finite von Neumann algebra M is one with the property that every collection of mutually orthogonal projections is at most countable. Equivalently, M has a faithful normal state (or even just a faithful state); or has a faithful normal representation possess- ing a cyclic separating vector. We often write 1l for the identity of M , which may be viewed as the identity operator on the underlying Hilbert spaces M is acting on. A projection p ∈ M is called finite if it is not Murray von Neumann equivalent to any proper subprojection; M is said to be finite if 1l is finite. Beware: σ-finite von Neumann algebras are not sums of finite ones, nor is every finite von Neumann algebra σ-finite. However a von Neumann algebra M possesses a faithful normal tracial state (the setting of [8] and most of [3]) if and only if it is both finite and σ-finite. (For the difficult direction of this one may compose the center valued trace on a finite von Neumann algebra, with a faithful normal state on the center, which in this case is σ-finite. From this it follows easily from e.g. [35, Proposition 2.2.5]) that any finite von Neumann algebra is a direct sum of algebras with faithful normal tracial states.) Any von Neumann algebra which is separably acting, or equivalently has separable predual M∗, is σ-finite. We will sometimes mention semi-finite von Neumann algebras; that is 1 is a sum of mutually orthogonal finite projections, or equivalently that every nonzero projection has a nonzero finite subprojection. For a subalgebra A of C(K), the continuous scalar functions on a compact Haus- dorff space K, a peak set is a set of form f −1({1}) for f ∈ A, kf k = 1. By replacing f by (1 + f )/2 we may assume also that f = 1 only on E. A noncommutative version of this called peak projections was considered in [24] and developed there and in a series of papers e.g. [6, 5, 10, 9, 11, 12]. There are various useful equivalent definitions of peak projections in the latter papers. They may be defined to be the weak* limits q = limn an in the bidual for a ∈ Ball(A) in the case such limit exists [11, Lemma 1.3]. We will say much more about peak projections in Section 2 below. Let M be a σ-finite von Neumann algebra, and let ν be a fixed faithful normal state on M . We write N for the crossed product M ⋊ν R of M with the modular group (σν t ) induced by ν. If M acts on the Hilbert space H, this crossed product is constructed by canonically representing the elements a of M as operators on L2(R, H) by means of the prescription π(a)ξ(t) = σν −t(a)(ξ(t)), and then generating a "larger" von Neumann algebra by means of the elements π(a) and the shift operators λs(ξ)(t) = ξ(t − s). The crossed product is known to admit a dual action of R in the form of an automorphism group (θs) indexed by R, and a normal faithful semifinite trace τ characterised by the property that τ ◦ θs = e−sτ . (See [40].) The identification a → π(a) above turns out to be a *-isomorphic embedding of M into N , and we will for the sake of simplicity identify M with π(M ). For sim- plicity of notation the canonical Hilbert space on which N acts will be denoted by K rather than L2(R, H). We will work in the space N of all τ -measurable operators on K affiliated to N . We remind the reader that the τ -measurable operators are those closed densely defined affiliated operators f which are "almost" bounded in the sense that for any ǫ > 0 we may find a projection e ∈ N with τ (1l − e) < ǫ, and with f e ∈ N . This space turns out to be a very well-behaved complete *-algebra large enough to admit all the noncommutative function spaces of interest. Using 4 DAVID P. BLECHER AND LOUIS LABUSCHAGNE the fact that τ ◦ θs = e−sτ , it is a simple matter to show that the group of *- automorphisms (θs) above admits an extension to continuous ∗-automorphisms on N (see for example either of [19, bottom p. 42] and [33, Proposition 4.7]). We will retain the notation θs for these extensions. Within this framework, the Haagerup Lp-spaces (0 < p < ∞) are defined by Lp(M ) = {a ∈ N : θs(a) = e−s/pa, s ∈ R}. Having thus being identified as subspaces of N , the Haagerup Lp-spaces in a very natural way inherit both a conjugation and an order structure from N . In fact even the natural (quasi-)norm topology on Lp(M ) is inherited from N -- the (quasi- )norm topology on each Lp(M ) agrees with the subspace topology inherited from N [40, Proposition II.26]. Via the canonical embedding of L1(M ) in N a functional ψ ∈ M∗ is positive (resp. selfadjoint, i.e. a difference of two positive functionals) if and only if its image in N is positive (resp. selfadjoint) [40]. We remind the reader that the crossed product admits an operator valued weight from the extended positive part of N to that of M . Using this operator valued weight, any normal weight ω on M may be extended to a dual weight eω on N by means of the simple prescription eω = ω ◦ T . In our analysis h will denote the Radon-Nikodym derivative of the (faithful normal semifinite) dual weight ν of our fixed faithful normal state ν above with respect to τ . It is known that in the σ- finite case, h belongs to the positive cone of the Haagerup space L1(M ). Using this operator it is also known that for each 0 ≤ c ≤ 1, a → hc/2ah(1−c)/p defines a dense embedding of M into Lp(M ) (1 ≤ p < ∞) [32]. Inspired by this fact, the Hardy spaces H p(A) (1 ≤ p < ∞) are defined to be the closure in Lp(M ) of the subspace hc/pAh(1−c)/p where 0 ≤ c ≤ 1. (We remind the reader that the closures for the various values of c all agree [27]). Given such a von Neumann algebra M and E a faithful normal conditional ex- pectation from M onto a von Neumann subalgebra D, a subdiagonal algebra A in M (with respect to E) is defined to be a weak* closed subalgebra of M containing 1l such that A + A∗ is weak* dense in M , and for which the action of the condi- tional expectation E : M → D = A ∩ A∗ is multiplicative on A. We say that A is maximal if it is not properly contained in any larger proper subdiagonal algebra in M with respect to E. Maximality of such unital weak* closed subdiagonal al- gebras satisfying the aforementioned weak* density condition, is characterised by the requirement that A be invariant under the modular automorphism group (σν t ) introduced a few paragraphs earlier (ν is as above). (see [34, Theorem 1.1], or equivalently [43, Theorem 1.1] & [28, Theorem 2.4]). Since we will have occasion to use the Haagerup reduction theorem [21], we pause to explain the essentials of that theory. From the von Neumann algebra M one constructs a larger algebra R by computing the crossed product with the diadic rationals QD (not R). So in this case one uses only the *-automorphisms σν t with symbols t in QD to similarly construct a copy πQD (M ) of M inside B(L2(QD, H)), with R = M ⋊ν QD then being the algebra generated by the elements belonging to this copy of M , and the shift operators λs with symbol s in QD. The discreteness of the group ensures that in this case the associated operator valued weight from the extended positive part of R to that of M , is in fact a faithful normal conditional expectation Φ : R → M . Inside R one may then construct an increasing net Rn of finite von Neumann algebras and a concomitant net of expectations Φn : R → Rn for which Φn ◦ Φm = Φm ◦ Φn = Φn when n ≥ m. (In the present setting this net UEDA'S THEOREM 5 actually turns out to be a sequence.) Each Rn comes equipped with a faithful state νn = ν ◦ ΦRn and a faithful normal tracial state τn. The vital fact regarding this construction, is that it may be adapted to the study of maximal subdiagonal algebras. Following [43], let D be the von Neumann subalgebra of R generated by D and the shift operators λs (s ∈ QD). (This is in essence just a copy of D ⋊σν QD.) Similarly let A be the weak* closed subalgebra generated by A and the same set of shift operators. Since A is invariant under in that reference, A may be defined as the weak* closure of sums of terms of σϕ t the form λ(t)x with x ∈ A. It is shown in [43] that A ∩ A∗ = D. The canonical expectation E : M → D extends to an expectation E : R → D, and if indeed A is maximal subdiagonal with respect to E, the algebra A will be maximal subdiagonal with respect to E. Moreover the expectation Φ : R → M , maps A and D onto A and D respectively. By equations (2.5) and (3.2) in [43], and the fact which we mentioned a few lines back regarding the definition of A, we see that E ◦ Φ = Φ ◦ E on A. Hence Φ( A0) = A0 since if E(a) = 0 then E(Φ(a)) = Φ( E(a)) = 0. Taking this one step further, the subalgebras An = A ∩ Rn turn out to be finite maximal subdiagonal subalgebras of the finite von Neumann algebras Rn, with the restriction of E to Rn acting multiplicatively on An, and mapping Rn onto D ∩ Rn. The algebras An turn out to be an increasing sequence of algebras which are weak* dense in A. 2. Peak projections As we said in the introduction, peak projections with respect to an operator algebra A may be defined to be the weak* limits q = limn an in the bidual, for a ∈ Ball(A) in the case such limit exists. Historically, if A is a C∗-algebra B then peak projections are very closely related to Edwards and Ruttiman's element u(a) (see e.g. [16]), computed in B∗∗. Certainly they are the same if a ≥ 0, and in that case they also agree with the B∗∗-valued Borel functional calculus element 1l{1}(a). Also they are the same, that is q above is u(a), if k1 − 2ak ≤ 1 (see [9, Corollary 3.3]). Thus we shall sometimes simply write our peak projections as u(a). Indeed every peak projection is of the form u(x) where k1 − 2xk ≤ 1 (if A is unital and an → q in the weak* topology with a ∈ Ball(A) set x = 1 2 (1 + a) [6, Corollary 6.9], or see [9, Theorem 3.4 (3)] for the general case). We call u(x) the peak for x. There is an elementary connection with the support projection s(·) (computed in B∗∗) which is often useful: if B is a unital C∗-algebra then u(1 − x) = 1l{1}(1 − x) = 1l(0,∞)(x)⊥ = s(x)⊥, x ∈ Ball(B)+. A similar but more general result holds in a unital nonselfadjoint algebra A: in the notation of Proposition 2.22 in [10] (see also [24, Proposition 5.4]) that result says that if k1 − 2xk ≤ 1 then the peak for x is u(x) = s(1 − x)⊥, where s(·) is the support projection in A∗∗ studied in e.g. [10, Section 2]. The following fact is implicit in the noncommutative peak set theory (see e.g. [10, 5, 9]), but we could not find it stated explicitly (except in the case of two projections -- see e.g. [24]). Lemma 2.1. If A is a closed subalgebra of a C∗-algebra B then the infimum of any countable collection of peak projections for A is a peak projection for A. 6 DAVID P. BLECHER AND LOUIS LABUSCHAGNE Proof. We may assume that A is unital, for example by Proposition [11, Proposition 6.4 (1)] (see also [9, Lemma 3.1]). Suppose that qn is a peak for an ∈ A, and that k1 − 2ank ≤ 1 (which can always be arranged as we said). Let q = ∧n qn, and an 2n . We will show that q is the peak for a. By a relation above the lemma we have a =Pn u(a) = s(1 − a)⊥ = s(Xn 1 − an 2n )⊥ = (∨n s(1 − an))⊥ = ∧n s(1 − an)⊥ = ∧n qn. In the last line we have used e.g. Proposition 2.14 or Theorem 2.16 (2) in [10], and the easy and known fact that the support projection of the closure of a sum of right ideals with left contractive approximate identities is the supremum of the individual support projections [10]. (cid:3) Remark 2.2. There is also a 'facial' proof of the previous result along the lines of [9, Proposition 1.1]. Another proof follows from an appeal to the next two results. For a compact Hausdorff space K, the peak sets for C(K) can be characterized abstractly as the compact Gδ subsets. There is a similar fact for C∗-algebras using Akemann's noncommutative topology (see [2] and references therein): the next result chararacterizes the peak projections for any C∗-algebra B as the 'compact Gδ projections'. A Gδ projection is the infimum in B∗∗ of a sequence (pn) of open projections in B∗∗, where a projection in B∗∗ is said to be open if it is a weak* limit of an increasing net from B+. The orthogonal complement of an open projection is called closed. A compact projection in B∗∗ is a projection q ∈ B∗∗ which is closed and satisfies qa = q for some a ∈ Ball(B)+ (or equivalently, which is closed with respect to B1; see e.g. [2], or 2.47 in [14]). If B is unital then 'compact' is the same as 'closed'. We have not been able to find the following result in the literature except for some form (see e.g. [14]) of parts of the unital case: Proposition 2.3. If B is a C∗-algebra and q is a projection in B∗∗, the following are equivalent: • q is a peak projection with respect to B. • q is a compact Gδ projection. • q is the weak* limit of a decreasing sequence from Bsa. Proof. (1) ⇒ (2) If q = u(a) for a ∈ Ball(B)+ let pn be the B∗∗-valued spectral projection of (1− 1 n ) for a. This gives a decreasing sequence of open projections in B∗∗ whose infimum (= weak* limit) equals q by the Borel functional calculus. It is well known that peak projections are compact (e.g. since q = aq). n , 1+ 1 (1) ⇒ (3) Clearly an ց u(a) weak* if a ∈ Ball(B)+ . If B is unital then one may finish the proof using the relation u(1 − x) = s(x)⊥ discussed above, and known results about the support projection s(·). Thus (2) implies by e.g. [14, Corollary 3.34] that 1 − q is a support projection, so that q is a peak projection. Similarly if B is unital then (3) implies that 1 − q is the weak* 1 2n an, then k ≤ 1 − q. A standard argument shows that k is strictly positive in the hereditary subalgebra (HSA) determined by 1 − q (any state of that HSA annihilating k also annihilates each an, hence also 1 − q, which is impossible). Thus 1 − q is the support projection of k, so that q is the peak projection of 1 − k. limit of an increasing sequence (an) from B+. Let k =P∞ k=1 UEDA'S THEOREM 7 If B is nonunital then (2) or (3) imply similar conditions with respect to B1, so that by the unital case q is a peak for a + t1 for some t ∈ [0, 1] and a ∈ Ball(B)+. The norm of a + t1 is kak + t = 1, and so 0 ≤ t = 1 − kak < 1 (or else a = 0 which is impossible). It is then easy to see, by e.g. the functional calculus for a, that q = u(a + t1) = u(a/kak), giving (1). (cid:3) We now describe general peak projections in terms of the C∗-algebraic peak projections characterized in the last result. Lemma 2.4. If A is a closed subalgebra of a C∗-algebra B and q ∈ B∗∗ then q is a peak projection for A if and only if q ∈ A⊥⊥ and q is a peak projection for B. Proof. If q is a peak projection for A, the peak for x ∈ Ball(A), then q is the weak* limit of (xn), which is in A⊥⊥. It is also the peak for some a ∈ Ball(B)+ (e.g. for x∗x or x, this follows for example from the proof of [9, Lemma 3.1] or from the formula u(x∗x) = u(x)∗u(x)). The converse is Corollary 4.5 of [12], and has even been generalized to Jordan operator algebras in a recent paper of the first author with Neal. A sketch of a proof: Suppose that q ∈ A⊥⊥ and q is a peak projection for B. We may assume that B is unital. Let A1 be the span of A and 1B. By Proposition 2.3 (ii) and [12, Lemma 4.4], q is a peak projection for A1. [11, Proposition 6.4 (1)] (see also [9, Lemma 3.1]), q is a peak projection for A. (cid:3) The following result, which characterizes peak projections in subalgebras of von Neumann algebras, will also be used in [13]. Theorem 2.5. If A is a closed subalgebra (not necessarily with any kind of approx- imate identity) of a von Neumann algebra M and q is a projection in M ∗∗ , then q is a peak projection for A if and only if q ∈ A⊥⊥ and q = ∧n qn, the infimum in M ∗∗ of a (decreasing, if desired) sequence (qn) of projections in M . n , 1 + 1 Proof. If q is a peak projection for x ∈ Ball(A) then by the last lemma q is in A⊥⊥, and q is the peak for some a ∈ Ball(M )+, so that q = 1l{1}(a), the M ∗∗- valued spectral projection of {1}. Let qn be the M -valued spectral projection of (1 − 1 n ) for a. We claim that the decreasing sequence (qn) in M has infimum q in M ∗∗. To see this note that as in Proposition 2.3, q is the infimum of (pn) in M ∗∗ where pn is the M ∗∗-valued spectral projection of (1 − 1 n ) for a. However, qn ≤ pn ց q. This may be seen from viewing the M -valued Borel functional calculus as the M ∗∗-valued Borel functional calculus multiplied by the canonical central projection z with zM ∗∗ ∼= M (this follows in turn from the uniqueness property of the Borel functional calculus). Also q ≤ qn (as may be seen e.g. by the above functional calculi, using continuous f with 1l{1} ≤ f ≤ 1l(1− 1 n , 1 + 1 n ,1+ 1 n )). Conversely, suppose that q = ∧n qn. Note that qn is clearly a peak projection (cid:3) for M , hence so is q by Lemma 2.1. Now apply the last lemma. 3. A Kaplansky density type result The following simple principle will be useful for dealing with Kaplansky density type results in unital operator spaces. Lemma 3.1. Let M be a unital operator space or operator system. Let σ be any linear topology on M weaker than the norm topology, e.g. the weak or weak* topology (the latter if M is a dual space too). Let X be a subspace of M for which Ball(X) 8 DAVID P. BLECHER AND LOUIS LABUSCHAGNE is dense in Ball(M ) in the topology σ. Then {x ∈ X : x + x∗ ≥ 0} is dense in {x ∈ M : x + x∗ ≥ 0} in the topology σ. Proof. Suppose that x ∈ M with x + x∗ ≥ 0. Then z = x + 1 and n satisfies z + z∗ ≥ 0 z + z∗ ≥ ≥ Cz∗z 2 n for some constant C > 0. This implies that C2z∗z − C(z + z∗) + 1 = (1 − Cz)∗(1 − Cz) ≤ 1. We may then approximate 1−Cz in the topology σ by a net xt ∈ Ball(X), and so 1 t ≥ 0 we have shown that z is in the closure of {x ∈ X : x + x∗ ≥ 0} in the topology σ. Hence so is x. (cid:3) C (1 − xt) → z with respect to σ. Since 2 − xt − x∗ The following is a Kaplansky density type result generalizing the one in Corollary 4.3 in [7], and [41, Section 4] (where Ueda points out that the dimensional restriction in [7, Corollary 4.3] can be removed). Theorem 3.2. If A is a maximal subdiagonal algebra in a σ-finite von Neumann algebra M , then Ball(A + A∗) is weak* dense in Ball(M ). Hence Ball(A + A∗)sa is weak* dense in Ball(M )sa. Moreover, (A + A∗)+ is weak* dense in M+. Also, in all of these statements we can replace 'weak*' by σ-strong*. Proof. The first assertion is known in the case that M has a faithful normal tracial state (this is the case discussed immediately before the theorem). Let x ∈ Ball(M ). As stated in the introduction, one may construct a σ-finite von Neumann super- algebra R of M with M appearing as the image of a faithful normal conditional expectation Φ : R → M . This R may be constructed so that it appears as the weak* closure of an increasing sequence Rn of finite von Neumann algebras each of which is the image of a faithful normal conditional expectation Φn : R → Rn for which we have that Φn ◦ Φm = Φm ◦ Φn = Φn when n ≥ m. In fact each x ∈ R is the weak* limit of the sequence Φn(x). As shown by [43], this construction can be modified in such a way that R admits a maximal subdiagonal subalgebra A for which Φ will map A onto A, and A∩ A∗ onto A ∩ A∗. Moreover the subalgebras An ∩ Rn ⊂ Rn, are each maximal subdiagonal in Rn, with ∪∞ n) is for each n n) of Ball( A + A∗) must n=1Ball( An + A∗ weak* dense in Ball(Rn). So the subset ∪∞ be weak* dense in the weak* closure of ∪∞ n=1Ball(Rn), namely Ball(R). It therefore follows that Φ(Ball( A + A∗)) = Ball(A + A∗) is weak* dense in Φ(Ball(R)) = Ball(M ). An weak*-dense in A. By known results Ball( An + A∗ n=1 The second assertion follows from the first by taking the real part. The third follows by applying the previous Lemma to the first. The last assertion follows from the previous assertions and [38, Theorem 2.6 (iv)]. (cid:3) We give a corollary of this which we will use later. Note that any element in A + A∗ has a unique representation a∗ + d + b with a, b ∈ A0 and d ∈ D. This is because if a∗ +d+b = 0 then applying E we see that d = 0. Also A0 ∩A∗ 0 ⊂ D∩A0 = (0). Thus A + A∗ = A0 ⊕ D ⊕ A∗ 0. It follows from this that selfadjoint elements x in A + A∗ are of form a + d + a∗ for a ∈ A0, d ∈ Dsa; and d must be positive if x ≥ 0 since d = E(x). We write H for the Hilbert transform on L2(M ) with respect to A as presented in [27]. It is shown there that H is continuous on L2(M ). In this insightful paper Ji shows that for each fixed 1 < p < ∞ the operators Hθ(hθ/p(a + d + b∗)h(1−θ)/p = ihθ/p(b∗ − a)h(1−θ)/p, a, b ∈ A, d ∈ D, θ ∈ [0, 1], UEDA'S THEOREM 9 extend to a unique bounded operator on Lp(M ) (the Hilbert transform) not de- In the case p = ∞, the pendent on the parameter θ. See [27, Theorem 3.2]. Hilbert transform H is only partially defined on M by means of the formula by H(a + d + b∗) = i(b∗ − a), for a, b ∈ A, d ∈ D. We remind the reader that h is the Radon-Nikodym derivative of the (faithful normal semifinite) dual weight ν with respect to the canonical trace on N , and its role in the definition of H p(A) is discussed in the introduction. Similarly, the selfadjointness and positivity referred to below was discussed there too. Lemma 3.3. Let A be as in the previous result, and H the Hilbert transform on L2(M ) with respect to A. If x ∈ Msa, then h 2 ) is selfadjoint. Moreover H(xh 2 )∗ = H(h 2 H(xh 2 x). 1 1 1 1 1 1 1 1 1 1 2 (a∗ − a)h 2 ) = i(a∗ − a)h 2 by the definition in [27]. Hence h 2 ) converges weakly to xh 1 2 b will be in L1, whence tr(xλh Proof. It suffices to prove the claim for the case where x ∈ M+. If a ∈ A0, d ∈ Dsa 2 H((a∗ + then H((a∗ + d + a)h 2 is selfadjoint. Any x ∈ M+ is the weak* limit of a net 2 ) = ih d + a)h xλ = a∗ λ + dλ + aλ, where aλ ∈ A and dλ ∈ D+, by Theorem 3.2 and the comment 2 in L2. To see this following it. Hence the net (xλh note that for any b ∈ L2, h 2 b). Since any norm continuous operator is also weakly continuous, the L2 continuity of H 2 ) in L2. This in turn ensures ensures that (H(xλh 2 ) in L1. By the lines at the start that (h 2 ) is selfadjoint with of this paragraph, h respect to the conjugation structure inherited from N . Similarly in view of the fact 2 x, we again have that that (xλh 2 ) H(h for each λ, from which it follows that H(xh (cid:3) 2 xλ) is weakly convergent to H(h 2 xλ is weakly convergent to (xh 2 )) converges weakly to H(xh 2 x). It is obvious that H(h 2 )) converges weakly to h 2 ) is selfadjoint. Hence h 2 xλ)∗ = H(xλh 2 x), as required. 1 2 )∗ = H(h 2 b) → tr(xh 1 2 )∗ = h 1 1 2 )∗ = h 1 1 2 H(xh 1 2 H(xλh 1 2 H(xh 1 1 2 H(xλh 1 1 1 1 1 1 1 1 1 1 1 1 1 1 4. Ueda's peak set theorem for σ-finite M Theorem 4.1. Let A be a maximal subdiagonal subalgebra of a σ-finite von Neu- mann algebra M . For a nonzero singular ϕ ∈ M ∗, there exists a contraction a ∈ A and a projection p ∈ M ∗∗ with (1) an → p weak* in M ∗∗. (2) an → 0 weak* in M , or equivalently ψ(p) = 0 for all ψ ∈ M∗. (3) ϕ(p) = ϕ(1), where ϕ is the absolute value of ϕ regarded as a member of the predual of the W ∗-algebra M ∗∗. Since ϕ is known to be singular if and only if ϕ is singular [38], one may assume that ϕ is a state if one wishes. In this case as in [41], (1) -- (3) may be restated as saying that (1) p is a peak projection, (2) p is dominated by the 'singular part' projection of M ∗∗, and (3) ϕ(p) = 1. The present section is devoted to generalizing Ueda's elegant proof of the tracial state case of Theorem 4.1. As in Theorem 1 of [41] we may find a decreasing sequence (pn) of projections in M with strong limit 0 and ϕ(pn) = ϕ(p0) = ϕ(1) 6= 0 for all n, where p0 is the strong limit of (pn) in M ∗∗. We may also assume that ν(pn) < n−6 where ν is the fixed faithful normal state on M . The formal series g = Pk kpk, may then be shown to correspond to a well-defined positive unbounded closed and densely defined operator affiliated to M . Moreover 10 DAVID P. BLECHER AND LOUIS LABUSCHAGNE the formal prescription g → h1/2g yields a well-defined embedding of this operator into L2(M ). These facts are proved in the following lemma: Lemma 4.2. Let the projections pn be as in the previous paragraphs, for n ∈ N. k=1 kpk)] = P∞ selfadjoint operator affiliated to M . Moreover h1/2g is a densely defined closable operator for which the closure is a well-defined element of L2(M ) which appears k=1 kpk. k=1 k[h1/2pk].) Similarly the formal oper- ator gh1/2 may be regarded as the sum in L2(M ) of the series limn→∞ gnh1/2 = Then the formal operator g = Pk kpk corresponds to a densely defined positive as the L2-norm limit of the sequence (h1/2gn) ⊂ L2(M ), where gn = Pn (In other words [h1/2(P∞ P∞ algebra. (See [39, §IX.4].) Observe that gn =Pn Proof. We first prove the claim regarding the affiliation of g. For this we will make use of the well known theory of the extended positive part of a von Neumann k=1 kpk may be in a canonical way be regarded as an increasing sequence of elements of the extended positive part of M . It is clear from the discussions following [39, Definition IX.4.4 & IX.4.6] that the supremum of this sequence (which we identify with g) is a well-defined element of the extended positive part of M . k=1 k[pkh1/2]. Next recall that by [39, Corollary IX.4.9], the action of the canonical faithful normal state ν extends to the extended positive part of M . In terms of this action, we must have that ν(g) = supn∈N ν(gn), and hence that ν(g) = sup n∈N ν(gn) ≤ lim n→∞ nXk=1 k−5 < ∞. However by [39, Theorem IX.4.8], g has a spectral decomposition of the form g(ω) =Z ∞ 0 λ dω(eλ) + ∞ω(p) ω ∈ M + ∗ . On considering the case where ω = ν and comparing the resulting formula to the the fact that ν(g) < ∞, it is clear that we must then have that ν(p) = 0, i.e. p = 0. It is now not difficult to conclude from the discussion in [39] that this can only be the case if the "operator part" of g (see [39, Lemma IX.4.7]) is all of g. See e.g. the last paragraph of the proof of Theorem IX.4.8 there. Hence g is a densely defined affiliated operator. We proceed to verify the claim regarding h1/2g (the gh1/2 statement follow- ing e.g. by duality). Firstly note that by the choice of the pk's we have that k=1 k[pkh1/2] = limn→∞ gnh1/2 must correspond to a well defined τ -measurable element G of L2(M ) ⊂ N . Recall that pmpk = pkpm = pm whenever m ≥ k. This ensures that for any fixed m ≥ 1 and any 1 ≤ n ≤ ∞, we have that gn(pm − pm+1) = nXk=1 kpk(pm − pm+1) = ( mXk=1 k)(pm − pm+1). (Here g∞ is identified with g.) So for each m ≥ 1 then have that h1/2g(pm−pm+1) = limn→∞ h1/2gn(pm − pm+1). Taking into account that p1 = ⊕∞ m=1(pm − pm+1), Pk kkpkh1/2k2 < +∞. Indeed So the formal series P∞ kkpkh1/2k2 2 = k2tr(h1/2pkh1/2) = k2ν(pk) < 1 k4 . UEDA'S THEOREM 11 it follows that h1/2g = h1/2gp1 = limn→∞ h1/2gnp1 = limn→∞ h1/2gn = G∗ as required. (cid:3) Remark 4.3. We note that if g =Pn npn, viewed as a supremum in the extended 2 ∈ L1(M ) and the latter can be shown to be 1 positive part M+ of M , then h the supremum and limit in L1(M ) of (h 2 gh 1 1 2 gnh 1 2 ). We will not use this though. Let g (resp. gn) be the Hilbert transform of gh 1 2 (resp. gnh 1 2 ) as in [27, Section 3], and let f = gh 1 2 + ig (resp. fn = gnh 1 2 + ign). Then fn, f ∈ H 2(A). In the following result we use the notion of accretive operators (see e.g. [22, Appendix C.7]). In Lp(M ) an operator is accretive if the associated operator T ∈ N has T + T ∗ positive in N . Corollary 4.4. With g = Pk kpk as above, and f = gh selfadjoint in L1(M ), so that h 2 f = h 2 g is 2 g is accretive in the sense above. 2 + ig, we have h 2 + ih 2 gh 1 1 1 1 1 1 Proof. If gn is as defined above, then by Lemma 3.3 we have h adjoint. By Lemma 4.2 and the continuity of H from [27], it follows that h selfadjoint. Thus, and since g is positive, h 2 g is accretive. 2 ) is self- 2 g is (cid:3) 2 H(gnh 2 f = h 2 + ih 2 gh 1 1 1 1 1 1 1 A σ-finite von Neumann algebra M has a convenient 'standard form'. Indeed as we recalled in the introduction, a characterization of σ-finite algebras is the existence of a (normal faithful) Hilbert space representation H possessing a fixed cyclic and separating vector Ω. Then ν(x) = (Ω, x Ω) is a faithful normal state on M . It is known that in this context (4.1) (M, H, P, J, Ω), is a 'standard form' for M , where P and J respectively denote the naturally as- sociated cone and the modular conjugation. The modular automorphism group σt is implemented by σt(·) = ∆it · ∆−it, where ∆ is the modular operator. By the universality of the standard form (see [1, 20, 40]) and hence also of the natural cone, we may identify the context with (M, H, P, J, Ω) (M, L2(M ), L2 +(M), ∗, h 1 2 ). In what follows we choose to work with the copy of M living inside B(L2(M )) as multiplication operators. In view of the above correspondence, we may do so 2 as the fixed cyclic and separating vector for without loss of generality. We view h this action of M . 1 Lemma 4.5. For each k ∈ N, there exist nets (a(k)λ) ⊂ A0, (d(k)λ) ⊂ D+ such that (a(k)∗ λ + d(k)λ + a(k)λ) converging to gk in the σ-strong* topology. Hence for any q ∈ L2(M ), the nets (a(k)∗ λ + d(k)λ + a(k)λ)q λ + d(k)λ + a(k)λ) will respectively converge in L2-norm to gq and qg. and q(a(k)∗ λ + d(k)λ + a(k)λ) ∈ M+, with (a(k)∗ Proof. This follows from Theorem 3.2 and the observation following it. (cid:3) The Hilbert transform H in the next result is the map partially defined on M by H(a + d + b∗) = i(b∗ − a), for a, b ∈ A, d ∈ D. 12 DAVID P. BLECHER AND LOUIS LABUSCHAGNE Lemma 4.6. Given a ∈ A0, d ∈ D+ with a∗ + d + a ∈ M+, the element (a∗ + d + a + 1l) + iH(a∗ + d + a) has an inverse v belonging to A, with both v and 1 − v contractive. Proof. Observe that with a and d as in the hypothesis, H(a∗ + d + a) = i(a∗ − a) is selfadjoint. Thus x = a∗ + d + a + iH(a∗ + d + a) is accretive. By the basic theory of accretive operators (see e.g. [22, Appendix C.7]) 1l + x has a contractive inverse v. Note that v ∈ A since the numerical range and hence the spectrum of x in A is in the right half plane. Also x(1l + x)−1 = 1l − (1l + x)−1 = 1l − v is a contraction in A, being the average of 1l and the well known Cayley transform of x. We remark that the map x 7→ x(1l + x)−1 is called the F-transform in recent papers of Charles Read and the first author. (cid:3) Proposition 4.7. There exist elements (wk) and wg of A for which each of wk, wg, wk − 1l and wg − 1l are contractions, with wk[(gk + 1l)h1/2 + iH(gkh1/2)] = h1/2 = [h1/2(gk + 1l) + iH(h1/2gk)]wk and wg[(g + 1l)h1/2 + iH(gh1/2)] = h1/2 = [h1/2(g + 1l) + iH(h1/2g)]wg. Moreover there exists a subnet of (wk) which is weak* convergent to wg. λ + d(k)λ + a(k)λ + 1l) + iH(a(k)∗ Proof. Choose nets (a(k)λ) ⊂ A0, (d(k)λ) ⊂ D+ as in Lemma 4.5. By Lemma 4.6 each (a(k)∗ λ + d(k)λ + a(k)λ) has an inverse v(k)λ belonging to A, with both v(k)λ and 1 − v(k)λ contractive. By passing to a subnet if necessary, we may assume that (v(k)λ) is weak* convergent. Let wk be the weak* limit of (v(k)λ). (Since both (v(k)λ) and (v(k)λ − 1l) are contained in the weak* compact set Ball(A), it is clear that both wk and wk − 1l will also be in this set.) We wish to prove that wk[(gk + 1l)h1/2 + iH(gkh1/2)] = h1/2 = [h1/2(gk + 1l) + iH(h1/2gk)]wk. In view of the similarity of the proofs, we prove only the first equality. Notice that (v(k)λ[(gk + 1l)h1/2 + iH(gkh1/2)]) converges weakly in L2 to wk[(gk + 1l)h1/2 + λ + d(k)λ + a(k)λ)h1/2 → gkh1/2 in iH(gh1/2)]. By Lemma 4.5 we have that (a(k)∗ L2-norm, and so also H((a(k)∗ λ + d(k)λ + a(k)λ)h1/2) → H(gkh1/2) in L2-norm by the continuity of H established in [27]. Since the v(k)λ's are contractive, it easily follows that kv(k)λ[(gk + 1l)h1/2 + iH(gkh1/2)] − h1/2k2 = kv(k)λ[(gk + 1l)h1/2 + iH(gkh1/2)] −v(k)λ[((a(k)∗ λ + d(k)λ + a(k)λ + 1l) + iH(a(k)∗ λ + d(k)λ + a(k)λ))h1/2]k2 ≤ k[(gk + 1l)h1/2 + iH(gkh1/2)] −((a(k)∗ λ + d(k)λ + a(k)λ + 1l) + iH(a(k)∗ λ + d(k)λ + a(k)λ))h1/2k2. Hence (v(k)λ[(gk + 1l)h1/2 + iH(gkh1/2)]) is norm convergent to h1/2. The claim regarding the gk's now follows from the uniqueness of limits. Since (wk) is bounded, it will admit a weak* convergent subnet (wγ). Let wg be the limit of that subnet. The claim regarding wg can now be verified with an essentially similar proof, but with the roles of v(k)λ and (a(k)∗ λ + d(k)λ + a(k)λ) UEDA'S THEOREM 13 respectively being played by wγ and gγ, and with Lemma 4.2 replacing Lemma 4.5. Thus we begin by noting that wγ [(g + 1l)h1/2 + iH(gh1/2)] → wg[(g + 1l)h1/2 + iH(gh1/2)] weakly in L2. Amending the previous argument as described above, now leads to the conclusion that kwγ[(g + 1l)h1/2 + iH(gh1/2)] − h1/2k2 → 0. So again the claim regarding the g's follows from the uniqueness of limits. (cid:3) We proceed to use Proposition 4.7 to analyse the structure of [(g + 1l)h1/2 + iH(gh1/2)]. Theorem 4.8. For any n ∈ N, we have kpnwgk ≤q 2 n(n+1) . Proof. Let gk and wk be as in Proposition 4.7; we had there a weak* convergent subnet of the latter sequence with limit wg. As usual one may replace (wk) by the subnet. For ease of notation, we will assume that (wk) is weak* convergent to wg. n(n+1) for every k ≥ n. To see this recall n(n+1) is weak* closed. So if each pnwk (k ≥ n) is It then suffices to show that kpnwkk ≤q 2 that the closed ball of radius q 2 in this ball, so is pnwg. Observe that for a, d and v as in Lemma 4.6, we have v∗(a + d + a∗ + 1l)v = = 1 2 1 2 v∗[((a∗ + d + a + 1l) + iH(a∗ + d + a)) +((a∗ + d + a + 1l) − iH(a∗ + d + a))]v [v + v∗]. Since v and v∗ are both contractive, this means that v∗(a + d + a∗ + 1l)v ≤ 1l. This in turn ensures that (a + d + a∗ + 1l)vv∗(a + d + a∗ + 1l) = (v−1)∗v∗(a + d + a∗ + 1l)vv∗(a + d + a∗ + 1l)vv−1 ≤ (v−1)∗v∗(a + d + a∗ + 1l)vv−1 = (a + d + a∗ + 1l). Hence (4.2) λ(a∗ kv∗ = hw∗ ≤ hw∗ = h(a∗ λ + dλ + aλ + 1l)wkh1/2ak2 λ + dλ + aλ + 1l)vλv∗ k(a∗ λ(a∗ k(a∗ λ + dλ + aλ + 1l)wkh1/2a , h1/2ai λ + dλ + aλ + 1l)wkh1/2a , wkh1/2ai λ + dλ + a)λ + 1l)wkh1/2a , h1/2ai Now let a ∈ M be given, and let the nets (a(k)λ) ⊂ A0, (d(k)λ) ⊂ D+ be as in Lemma 4.5. Then the nets (a(k)∗ λ + d(k)λ + a(k)λ + 1l)wkh1/2a converge to (gk + 1l)wkh1/2a in L2-norm. As we saw in the proof of Proposition 4.7, on passing to a subnet if necessary, we may assume that the nets (v(k)λ)'s described by Lemma 4.6, are weak* convergent to the wk's. Since the v(k)λ's are contractive, we have that k[v(k)∗ λ(gk + 1l)wkh1/2a]k λ(a(k)∗ λ + d(k)λ + a(k)λ + 1l)wkh1/2a] − [v(k)∗ 14 DAVID P. BLECHER AND LOUIS LABUSCHAGNE ≤ k[(a(k)∗ λ + d(k)λ + a(k)λ + 1l)wkh1/2a] − [(gk + 1l)wkh1/2a]k → 0. Thus [v(k)∗ λ(a(k)∗ λ + d(k)λ + a(k)λ + 1l)wkh1/2a] − [v(k)∗ λ(gk + 1l)wkh1/2a] → 0 in norm. Since also v(k)∗ w∗ weakly convergent to w∗ k(gk + 1l)wkh1/2a, it follows that v(k)∗ k(gk + 1l)wkh1/2a. λ(gk + 1l)wkh1/2a is weakly convergent in L2(M ) to λ + d(k)λ + a(k)λ + 1l)wkh1/2a is λ(a(k)∗ We proceed to show that k(gk + 1l)1/2wkh1/2ak2 ≤ kh1/2ak2. To see this we firstly observe that hv(k)∗ λ(a(k)∗ λ + d(k)λ + a(k)λ + 1l)wkh1/2a, h1/2ai and that → hw∗ k(gk + 1l)wkh1/2a, h1/2ai = k(gk + 1l)1/2wkh1/2ak2, h(a(k)∗ λ + d(k)λ + a(k)λ + 1l)wkh1/2a , wkh1/2ai → h(gk + 1l)wkh1/2a, wkh1/2ai = k(gk + 1l)1/2wkh1/2ak2. Next observe that by inequality (4.2), we have that kv(k)∗ = h(a(k)∗ λ(a(k)∗ λ + d(k)λ + a(k)λ + 1l)wkh1/2ak2 λ + d(k)λ + a(k)λ + 1l)wkh1/2a , wkh1/2ai It follows from the above inequality that hv(k)∗ ≤ kv(k)∗ ≤ [h(a(k)∗ λ(a(k)∗ λ + d(k)λ + a(k)λ + 1l)wkh1/2a , h1/2ai λ + d(k)λ + a(k)λ + 1l)wkh1/2ak.kh1/2ak λ(a(k)∗ λ + d(k)λ + a(k)λ + 1l)wkh1/2a , wkh1/2ai]1/2.kh1/2ak. On taking limits, we have k(gk + 1l)1/2wkh1/2ak2 ≤ k(gk + 1l)1/2wkh1/2ak.kh1/2ak, or equivalently, k(gk + 1l)1/2wkh1/2ak ≤ kh1/2ak as claimed. Finally note that since the pn's are decreasing, we as before have that (n + 1)n 2 pn = nXm=1 mpn ≤ nXm=1 mpm, which is dominated by Pk m=1 mpm = gm. Hence kpnwkh1/2ak2 = hw∗ kpnwk(h 1 2 a), (h 1 2 a)i ≤ ≤ = ≤ 2 n(n + 1) 2 n(n + 1) 2 n(n + 1) 2 n(n + 1) hw∗ kgkwk(h 1 2 a), (h 1 2 a)i hw∗ k(gk + 1l)wk(h 1 2 a), (h 1 2 a)i k(gk + 1l)1/2wkh1/2ak2 kh1/2ak2. Since the subspace {h1/2a : a ∈ M } is dense in L2(M ), it follows that the operator n(n+1) . This (cid:3) of left multiplication by pnwk on L2(M ), has norm dominated by q 2 proves the claim. UEDA'S THEOREM 15 Thus with b = 1 − wg we deduce that kpk − pkbk ≤ q 2 k(k+1) as needed for the argument in [41] to proceed. Indeed the rest of that argument is identical. We obtain p0 = p0b = bp0 where p0 is the strong limit of (pn) in M ∗∗, and if a = (1 + b)/2 then (an) converges weak* to a peak projection p ≥ p0 with ϕ(p) = ϕ(p0) = ϕ(1). If kaξk2 = kξk2 for ξ ∈ L2(M ), then as in [41] we obtain bξ = ξ, so that in our notation above we have ξ ∈ Ker(wg) = 0. However Ker(wg) = (0). Indeed the projection associated with the kernel is in M ; and if e ∈ M is a projection with wge = 0 then by the last equality in Proposition 4.7 we obtain h1/2e = 0, so that e = 0. Hence as in [41] (which relies here on the noncommutative peak theory [24], see also e.g. [6, 9]) we obtain an → 0 weak* in M . This completes the proof of the generalization of Ueda's peak set theorem to σ-finite algebras. 5. Consequences of Ueda's peak set theorem for σ-finite M All the other consequences from [41] of Ueda's peak set theorem, now go through with unaltered proofs for maximal subdiagonal subalgebras A of a σ-finite von Neumann algebra M . Indeed this is true rather generally. If A is a weak* closed subalgebra of a von Neumann algebra M then we say that A is an Ueda algebra in M if Ueda's peak set theorem holds for A; that is if for every singular state on M there is a peak projection q for A with ϕ(q) = 1 and q is dominated by the 'singular part' projection of M ∗∗, as in the restatement after Theorem 4.1. The ideas in [13, Lemma 9.1] give the following restatement: Corollary 5.1. Suppose that A is a weak* closed subalgebra of a von Neumann algebra M . Then A is an Ueda algebra in M if and only if for every singular state ϕ on M , there exists a (increasing, if desired) sequence (qn) of projections in Ker(ϕ) with supremum 1 in M , and supremum in M ∗∗ lying in A⊥⊥. If (qn) is increasing then the last condition means that ψ(qn) → 0 for any ψ ∈ A⊥. Proof. By Theorem 2.5, the information about q in the lines above the present corollary is equivalent to: there is a (decreasing, if desired) sequence (qn) of projec- tions in M with infimum q in M ∗∗ lying in A⊥⊥ satisfying ϕ(q) = 1, and ψ(q) = 0 for all normal states ψ of M . As in [13, Lemma 9.1] the last condition is equivalent to the infimum in M of (qn) being 0, and ϕ(q) = 1 if and only if ϕ(qn) = 0 for all n. Finally set p = q⊥ and replace qn by q⊥ n . (cid:3) We remark that if A is an Ueda algebra then it is easy to see that so is A∗ = {x∗ : x ∈ A}. Corollary 5.2. Suppose that a weak* closed subalgebra A of a von Neumann algebra M is an Ueda algebra. If ϕ ∈ M ∗ has nonzero singular part ϕs, then there exists a contraction a ∈ A and a projection p ∈ M ∗∗ with an → p weak* in M ∗∗, an → 0 weak* in M , and ϕs = ϕ · p. Theorem 5.3. Suppose that a weak* closed subalgebra A of a von Neumann algebra M is an Ueda algebra. Write A∗ n for the set of restrictions to A of singular and normal functionals on M . Each ϕ ∈ A∗ has a unique Lebesgue decomposition relative to M : ϕ = ϕn + ϕs with ϕn ∈ A∗ s. Moreover, kϕk = kϕnk + kϕsk. Corollary 5.4. Suppose that a weak* closed subalgebra A of a von Neumann algebra M is an Ueda algebra. Then the predual A∗ of A is unique, and is an L-summand in A∗. Also, A∗ has property (V∗) and is weakly sequentially complete. n and ϕs ∈ A∗ s and A∗ 16 DAVID P. BLECHER AND LOUIS LABUSCHAGNE (See also e.g. [31] for recent similar results for a completely different class of dual operator algebras.) Theorem 5.5 (F. & M. Riesz type theorem). Suppose that a weak* closed subal- gebra A of a von Neumann algebra M is an Ueda algebra. If ϕ ∈ M ∗ annihilates A (that is, ϕ ∈ A⊥) then the normal and singular parts, ϕn and ϕs, also annihilate A. Our proofs from [7] then give the following results (suitably modified by an appeal to Theorem 5.5 instead of to the F & M type theorem in [7]), as noted in [41] and suggested by the referee of that paper. One may define an F & M Riesz algebra to be a weak* closed subalgebra A of a von Neumann algebra M , such that if ϕ ∈ A⊥ then the normal and singular parts, ϕn and ϕs, also annihilate A. Theorem 5.5 then says that any Ueda algebra is an F & M Riesz algebra. Again, it is easy to argue (by considering ψ∗(x) = ψ(x∗)) that if A is an F & M Riesz algebra then so is A∗ = {x∗ : x ∈ A}. By proofs in [7] we then have: Corollary 5.6. Suppose that A is an F & M Riesz algebra in a von Neumann algebra M such that A+A∗ is weak* dense in M . If ϕ ∈ M ∗ annihilates A+A∗ then ϕ is singular. Any normal functional on M is the unique Hahn-Banach extension of its restriction to A+ A∗, and in particular is normed by A+ A∗. In addition, any Hahn-Banach extension to M of a weak* continuous functional on A, is normal. Corollary 5.7. If A is an F & M Riesz algebra in a von Neumann algebra M such that A + A∗ is weak* dense in M then Ball(A + A∗) is weak* dense in Ball(M ). Moreover in this case we obtain all the assertions of Theorem 3.2 too. The last assertion of the Corollary 5.6 is related to the well known Gleason- Whitney theorem in function theory. A special case of the following appears in [7, Theorem 4.1] and [34, Theorem 3.4]. We can express some of the ideas in those results more abstractly and generally as follows: Lemma 5.8. Suppose A is a weak* closed subalgebra of a von Neumann algebra M . Then A + A∗ is weak* dense in M if and only if there is at most one normal Hahn-Banach extension to M of any normal weak* continuous functional on A. Proof. (⇒) Choose a ∈ Ball(A) such that ϕ(a) = 1, and let e be the left support of a, which is the support of aa∗. We may suppose that ϕ ∈ M∗ and that ψ is another normal Hahn-Banach extension of ϕA. We have 1 = ϕ(a) ≤ ϕ(aa∗) ≤ ϕ(e), so that ϕ(e⊥) = 0. Hence ϕe⊥ = 0 and ϕe⊥ = 0. Similarly ψe⊥ = 0 and (ϕ − ψ)e⊥ = 0. Next note that ϕa is contractive and unital, so positive. Similarly for ψ, and so (ϕ − ψ)a is a selfadjoint normal functional. It vanishes on A, hence also on A + A∗ and on M . From this it is easy to see that (ϕ − ψ)e = 0. So ϕ − ψ = (ϕ − ψ)e + (ϕ − ψ)e⊥ = 0. (⇐) It is enough to show that if normal ψ annihilates A + A∗ then ψ = 0. By taking real and imaginary parts we may assume that ψ = ψ∗. Suppose ψ = ψ1 − ψ2 for positive normal ψk. Then ψ1 = ψ2 + ψ and ψ2 agree on A, and are normal Hahn-Banach extensions since the norm of a positive functional is its value at 1. So ψ1 = ψ2 and ψ = 0. (cid:3) UEDA'S THEOREM 17 Corollary 5.9 (Gleason-Whitney type theorem). Suppose that A is an F & M Riesz algebra in a von Neumann algebra M . Then A + A∗ is weak* dense in M if and only if every normal functional on A has a unique Hahn-Banach extension to M , and if and only if every normal functional on A has a unique normal Hahn- Banach extension to M . Of course all of these hold when A is a maximal subdiagonal subalgebra of a σ- finite von Neumann algebra M . Conversely these properties characterize maximal subdiagonal subalgebras. The following is a partial strengthening of [34, Theorem 3.4] (the equivalence of (i) and (iv) there). Corollary 5.10 (Gleason-Whitney type theorem). Let A be a weak* closed unital subalgebra of a σ-finite von Neumann algebra M , for which • σν t (A) = A for each t ∈ R (where σν t for M described in our Introduction), and is the modular automorphism group • the canonical expectation E : M → A ∩ A∗ = D is multiplicative on A. Then A+A∗ is weak* dense in M (that is, A is maximal subdiagonal with respect to D) if and only if every normal functional on A has a unique normal Hahn-Banach extension to M . 6. The case of semi-finite and general von Neumann algebras We first briefly discuss the results of our paper in the setting of general von Neu- mann algebras. We recall from e.g. [35, Proposition 2.2.5] that any von Neumann algebra M is a direct sum of algebras Mi of the form Ri ¯⊗B(Hi) for a σ-finite von Neumann algebra Ri. If A is a maximal subdiagonal algebra in M , and if the center of M is contained in the center of A ∩ A∗, then the central projections correspond- ing to the direct sum will allow a decomposition of a maximal subdiagonal algebra A of M as a direct sum of algebras Ai ⊂ Mi, and it is easy to see that these are maximal subdiagonal subalgebras of Mi. Assuming that the B(Hi)'s appearing in the form of Mi above correspond to separable Hilbert spaces Hi, then Ri ¯⊗B(Hi) is σ-finite, and we get Ueda's theorem in this case (Theorem 4.1 but with the σ-finite M replaced by our M above). We immediately deduce that all the results in the last section (Section 5) are valid for this A and M . We now turn to investigating when Ueda's peak set theorem fails. Of course if Ueda's peak set theorem fails for a von Neumann algebra M then it also fails for every weak* closed unital subalgebra A of M . Thus henceforth in this section we shall assume that A = M . An Ulam measurable cardinal is one such that if I is a set of this cardinality, then there exists a free ultrafilter p on I such that every sequence A1 ⊇ A2 ⊇ · · · of nonempty sets in p has nonempty intersection [15, 25]. (Remark: it is a pleasant exercise that an ultrafilter allows no empty countable intersections of members if and only if it is closed under countable intersections. Also, one may always make countable intersections 'decreasing'.) The concept of measurable cardinal used in the next result will be explained a little more at the start of its proof. This result shows that there is little chance of generalizing Ueda's peak set theorem to semi- finite von Neumann algebras in the usual set theoretic universe used in most of functional analysis, since this would imply a solution to one of the famous open "problems" in mathematics. We use quotes because nowadays this problem is not believed to be solvable. 18 DAVID P. BLECHER AND LOUIS LABUSCHAGNE The strategy of our proof is simple: it is known that a bound on the size of a set I in relation to being of measurable cardinality is equivalent to being 'realcompact'. Also, I not being realcompact is known to imply that βI \ I contains points not contained in closed Gδ subsets of βI of a certain type. Finally, for C(K) spaces the closed Gδ sets are exactly the peak sets, by the strict form of the Urysohn lemma or as in Proposition 2.3. However since we lack a good reference (besides scattered pieces found in an internet search for 'realcompact discrete spaces'; see e.g. [15, p. 402 ff]), we will include short arguments for several of these points for the reader's convenience. Theorem 6.1. If Ueda's peak set theorem held for all finite von Neumann algebras then there exist no (uncountable) measurable cardinals. Proof. The existence of (uncountable) measurable cardinals is known to be equiva- lent to the existence of Ulam measurable cardinals [15, 25]. Suppose that I was an uncountable set of Ulam measurable cardinality. Clearly M = ℓ∞(I) is a finite (and semi-finite) von Neumann algebra, and A = M is a maximal subdiagonal algebra. We may view the free ultrafilter p in the definition of Ulam measurable cardinality as a (singleton) closed set in βI \I. It is the support of a Dirac probability measure, which can be viewed as a pure state on M = C(βI) (evaluation at p). This state is singular (we leave this as an exercise since there are many ways to see this). Moreover via the well known correspondences between minimal projections and pure states and their supports, the support of this state in C(βI)∗∗ is the minimal projection which is the image e of the characteristic function of F = {p} in C(βI)∗∗ (that is, it is the image of the functional µ 7→ µ(F ) on C(βI)∗, viewing the latter as a space of measures). Indeed here we are just invoking aspects of the well known noncommutative dictionary between the basic theory of probability measures and that of states. If Ueda's theorem held for M then there would exist a peak projection q ∈ C(βI)∗∗ with e ≤ q ≤ z, where z is the orthogonal complement of the canonical projection in M ∗∗ corresponding to M∗. These three projections e, z, q correspond to closed sets in βI, namely to sets F = {p}, βI \ I, and E say, respectively; and the latter is a classical peak set by the 'peaking' theory [6, 24, 5, 10, 9, 11]. (That z corresponds to βI \ I is well known, and was sketched in an earlier version of the present paper available on arXiV.) By Theorem 2.5, the characteristic function of any peak set E for M is an intersection of a decreasing sequence of projections in M = C(βI) = l∞(I). Thus by the theory of the Stone-Cech compactification, E = ∩∞ n=1 [An], where [An] is the (clopen) closure in βI of (open) An ⊂ I, where A1 ⊇ A2 ⊇ · · · . Also, ∩∞ n=1 An = ∅. To see the latter, note that n=1 [An] ⊂ βI \ I if and only if ∩∞ ∩∞ n=1 An ⊂ (∩∞ n=1 [An]) ∩ I ⊂ (βI \ I) ∩ I = ∅ if ∩∞ n=1 [An] ⊂ βI \ I. The converse follows from the inclusion I ∩ (∩∞ n=1 [An]) ⊂ I ∩ [An] = An, n ∈ N . Thus for any closed subset F of a peak set E for C(βI), with E ⊂ βI \ I, we have F ⊂ ∩∞ n=1 [An] for sets A1 ⊇ A2 ⊇ · · · in I with empty intersection. In our special case where F = {p}, the fact that p ∈ [An] implies that An ∈ p for all n ∈ N, with p regarded as an ultrafilter. This contradicts the property p has in the definition of Ulam measurable cardinality above. So there is no Ulam measurable cardinal. (cid:3) UEDA'S THEOREM 19 Remark 6.2. By the last proof Ueda's peak set theorem holding for M = A = ℓ∞(I), is equivalent to saying that every closed set F in βI \ I which is the support of a Borel probability measure, is contained in ∩∞ n=1 [An] for sets A1 ⊇ A2 ⊇ · · · in I with empty intersection. Closed sets in βI \ I have a nice characterization in the basic literature of the Stone-Cech compactification. It turns out that Ueda's peak set theorem also fails when M = A = B(H) with H of dimension an Ulam measurable cardinal, or a real valued measurable cardinal, as is discussed together with Nik Weaver in [13]. Indeed in that paper (which was written after the first distributed version of the present paper) it is shown that if M is a von Neumann algebra then Ueda's peak set theorem fails when M = A if and only if M possesses a singular state ϕ which is regular, that is, ϕ(∨n qn) = 0 for every sequence of projections (qn) in Ker(ϕ). (See [23] for other characterizations and facts about regular states; hence Ueda's peak set theorem is strongly tied to 'quantum measure theory' in the sense of that reference.) This is also equivalent to saying that there is a collection of mutually orthogonal projections in M of cardinality ≥ a fixed cardinal κ, namely the first cardinal on which there is a 'regular' singular finitely additive probability measure. (Here and below measures are assumed to be defined on all subsets of the cardinal.) The existence of regular singular states or such regular measures is generally believed to be consistent with ZFC set theory. Indeed as explained in [13] it is believed to be consistent with ZFC set theory that the latter 'first cardinal' is ≤ the cardinality of the real numbers. On the other hand, since any cardinal on which there is a singular probability measure dominates the 'first cardinal' above, it follows that if M ⊂ B(H) where dim(H) is smaller than any real-valued measurable cardinal (or if measurable cardinals do not exist), then Ueda's peak set theorem holds for M (and taking A = M ). From the assertion in the last paragraph about the cardinality of the real num- bers, it follows that one should not hope to be able to prove Ueda's peak set theorem for A = M = l∞(R) in ZFC. Indeed Ueda's theorem in this case implies by the assertion in the last paragraph about regular states, a negative solution to the famous 'Banach measure problem': Is there a probability measure defined on all subsets of [0, 1] which is zero on singletons? (It is well known that if there is, then one can find another that extends Lebesgue measure.) Banach showed that you cannot prove an affirmitive answer to this in ZFC. The existence of a negative answer is equivalent to the nonexistence of measurable cardinals in ZFC. However as we have stated earlier, it is generally believed by set theorists that the existence of measurable cardinals is consistent with ZFC. This shows that one cannot hope to be able to prove Ueda's peak set theorem in ZFC for von Neumann algebras that are much 'bigger' than σ-finite (the case of the main theorem of our paper). And indeed experts in von Neumann algebras are usually happy to only consider σ-finite von Neumann algebras in their results, because 'bigger' algebras are often pathological. On the other hand it is shown in [13] that Ueda's peak set theorem holds in ZFC for A = M = l∞(ℵ1), where ℵ1 is the first uncountable cardinal, and this von Neumann algebra is not σ-finite. Hence if we assume the continuum hypothesis then Ueda's peak set theorem does hold for A = M = l∞(R). Assuming the negation of the continuum hypothesis, a remaining question seems to be for what cardinals κ between the infinite countable cardinal and the cardinality of the reals can one prove Ueda's peak set theorem in ZFC for 20 DAVID P. BLECHER AND LOUIS LABUSCHAGNE A = M = l∞(κ). Nik Weaver has sketched to us a proof in the case of ℵ2, and this trick seems to extend to ℵn for n ∈ N. Thinking about the last paragraph in conjunction with the proof of our main theorem, suggests to us that it may possibly be interesting to study Haagerup's reduction theory, the standard form, and related topics, for von Neumann algebras possessing uncountable collections of mutually orthogonal projections of cardinality smaller than the cardinality of the reals (assuming of course the negation of the continuum hypothesis). Acknowledgments. We are grateful to Nik Weaver for a very helpful and lengthy conversation shortly before distribution of an earlier version of the present paper, which led to [13], a paper largely devoted to quantum measure theory in the sense of [23], quantum cardinals, and various continuity properties of states on von Neumann algebras. In Section 9 of that paper the set theoretic issues associated with Ueda's peak set theorem for von Neumann algebras are explored more fully, as mentioned above. We are also grateful to the referee for many helpful comments. References [1] H. Araki, Some properties of modular conjugation operator of von Neumann algebras and a non-commutative Radon-Nikodym theorem with a chain rule, Pacific J. Math. 50 (1974), 309-354. [2] C. A. Akemann, J. Anderson, and G. K. Pedersen, Approaching infinity in C ∗-algebras, J. Operator Theory 21 (1989), 255271. [3] W. B. Arveson, Analyticity in operator algebras, Amer. J. Math. 89 (1967), 578 -- 642. [4] T. Bekjan, Riesz factorization of Haagerup noncommutative Hardy spaces, preprint. [5] D. P. Blecher, Noncommutative peak interpolation revisited, Bull. London Math. Soc. 45 (2013), 1100 -- 1106. [6] D. P. Blecher, D. M. Hay, and M. Neal , Hereditary subalgebras of operator algebras, Journal of Operator Theory 59 (2008), 333-357. [7] D. P. Blecher and L. E. Labuschagne, Noncommutative function theory and unique exten- sions, Studia Math. 178 (2007), 177-195. [8] D. P. Blecher and L. E. Labuschagne, Von Neumann algebraic H p theory, Function Spaces: Fifth Conference on Function Spaces, Contemp. Math. Vol. 435, Amer. Math. Soc. (2007). [9] D. P. Blecher and M. Neal, Open projections in operator algebras II: Compact projections, Studia Math 209 (2012), 203 -- 224. [10] D. P. Blecher and C. J. R. Read, Operator algebras with contractive approximate identities, J. Functional Analysis 261 (2011), 188-217. [11] D. P. Blecher and C. J. R. Read, Operator algebras with contractive approximate identities II, J. Functional Analysis 264 (2013), 1049 -- 1067. [12] D. P. Blecher and C. J. R. Read, Order theory and interpolation in operator algebras, Studia Math. 225 (2014), 61 -- 95. [13] D. P. Blecher and N. Weaver, Quantum measurable cardinals, preprint (revised December 2016). [14] L. G. Brown, Semicontinuity and multipliers of C ∗-algebras, Canad. J. Math. 40 (1988), 865-988. [15] W. W. Comfort and S. Negrepontis, The theory of ultrafilters, Die Grundlehren der mathe- matischen Wissenschaften, Band 211, Springer-Verlag, New York-Heidelberg, 1974. [16] C. M. Edwards and G. T. Ruttimann, Compact tripotents in bi-dual JB∗-triples, Math. Proc. Camb. Philos. Soc. 120 (1996), 155 -- 173. [17] T. Fack and H. Kosaki, Generalized s-numbers of τ -measurable operators, Pacific J. Math. 123 (1986), 269-300. [18] S. Goldstein and J. L. M. Lindsay, Lp-spaces and quantum dynamical semigroups, Quantum probability (Gdansk, 1997), 211 -- 216, Banach Center Publ., 43, Polish Acad. Sci., Warsaw, 1998. UEDA'S THEOREM 21 [19] S. Goldstein and J. L. M. Lindsay, Markov semigroups KMS-symmetric for a weight, Math. Ann. 313 (1999), 39 -- 67. [20] U. Haagerup, The standard form of von Neumann algebras, Math. Scand. 37 (1975), 271 -- 283. [21] U. Haagerup, M. Junge and Q. Xu, A reduction method for noncommutative L p-spaces and applications, Trans. Amer. Math. Soc. 362 (2010), 2125 -- 2165. [22] M. Haase, The functional calculus for sectorial operators, Operator Theory: Advances and Applications, 169, Birkhauser Verlag, Basel, 2006. [23] J. Hamhalter, Quantum Measure Theory, Fundamental Theories of Physics 134, Kluwer Academic Publishers Group, 2003. [24] D. M. Hay, Closed projections and peak interpolation for operator algebras, Integral Equations Operator Theory 57 (2007), 491 -- 512. [25] T. Jech, Set theory. The third millennium edition, revised and expanded, Springer Mono- graphs in Mathematics, Springer-Verlag, Berlin, 2003. [26] G. Ji. A noncommutative version of H p and characterizations of subdiagonal algebras, Inte- gral Equations Operator Theory 72 (2012), 131 -- 149. [27] G. Ji. Analytic Toeplitz algebras and the Hilbert transform associated with a subdiagonal algebra, Sci. China Math. 57 (2014), 579 -- 588. [28] G. Ji, T. Ohwada and K-S. Saito, Certain structure of subdiagonal algebras, J Operator Theory 39 (1998), 309 -- 317. [29] G. Ji and K-S. Saito,Factorization in Subdiagonal Algebras, J. Funct. Anal. 159 (1998), 191 -- 201. [30] M. Junge, Doob's inequality for non-commutative martingales, J. Reine Angew. Math. 549 (2002), 149 -- 190. [31] M. Kennedy and D. Yang, A non-self-adjoint Lebesgue decomposition, Anal. PDE 7 (2014), 497 -- 512. [32] H. Kosaki, Applications of the complex interpolation method to a von Neumann algebra: Noncommutative Lp-spaces, J. Funct. Anal. 56 (1984), 29 -- 78. [33] L. E. Labuschagne, Composition Operators on Non-commutative Lp-spaces, Expo. Math 17 (1999), 429 -- 468. [34] L. E. Labuschagne, Invariant subspaces for H 2 spaces of σ-finite algebras, Preprint (2016) arXiv:1604.01968 [35] S. Sakai, C ∗-algebras and W ∗-algebras, Classics in Mathematics, Springer-Verlag, Berlin, 1998. [36] L. M. Schmitt, The Radon-Nikodym theorem for Lp-spaces of W ∗-algebras, Publ. RIMS. Kyoto. Univ. 22(1986), 1025 -- 1034. [37] T. P. Srinivasan and J-K. Wang, Weak*-Dirichlet algebras, In Function algebras, Ed. Frank T. Birtel, Scott Foresman and Co., 1966, 216-249. [38] M. Takesaki, Theory of Operator Algebras I, Springer, New York, 1979. [39] M. Takesaki, Theory of Operator Algebras II, Encyclopaedia of Mathematical Sciences, Vol. 125, Springer-Verlag, Berlin, 2003. [40] M. Terp, Lp spaces associated with von Neumann algebras, Notes, Math. Institute, Copen- hagen Univ. 1981. [41] Y. Ueda, On peak phenomena for non-commutative H∞, Math. Ann. 343 (2009), 421 -- 429. [42] Y. Ueda, On the predual of non-commutative H∞, Bull. Lond. Math. Soc. 43 (2011), 886 -- 896. [43] Q. Xu, On the maximality of subdiagonal algebras, J. Operator Th. 54 (2005), 137 -- 146. Department of Mathematics, University of Houston, Houston, TX 77204-3008 E-mail address, David P. Blecher: [email protected] DST-NRF CoE in Math. and Stat. Sci,, Unit for BMI,, Internal Box 209, School of Comp., Stat., & Math. Sci., NWU, PVT. BAG X6001, 2520 Potchefstroom, South Africa E-mail address: [email protected]
1512.04820
2
1512
2017-06-01T16:04:44
Strong solidity of free Araki-Woods factors
[ "math.OA" ]
We show that Shlyakhtenko's free Araki-Woods factors are strongly solid, meaning that for any diffuse amenable von Neumann subalgebra that is the range of a normal conditional expectation, the normalizer remains amenable. This provides the first class of nonamenable strongly solid type III factors.
math.OA
math
STRONG SOLIDITY OF FREE ARAKI–WOODS FACTORS R ´EMI BOUTONNET, CYRIL HOUDAYER, AND STEFAAN VAES Abstract. We show that Shlyakhtenko's free Araki–Woods factors are strongly solid, mean- ing that for any diffuse amenable von Neumann subalgebra that is the range of a normal conditional expectation, the normalizer remains amenable. This provides the first class of nonamenable strongly solid type III factors. 1. Introduction A von Neumann algebra N is amenable if there exists a norm one projection Φ : B(L2(N )) → N . By Connes's fundamental result [Co75], any amenable von Neumann algebra is approximately finite dimensional. Moreover, the class of amenable factors with separable predual is completely classified by the flow of weights [Co72, Co75, Ha85, Kr75]. In particular, there exists a unique amenable II1 factor with separable predual: it is the hyperfinite II1 factor R of Murray and von Neumann [MvN43]. Starting with [Po01], Popa's deformation/rigidity theory has lead to far reaching classification and structure theorems for nonamenable factors. Particular attention was given to several types of indecomposability results for von Neumann algebras M , like primeness (the impos- sibility to write a factor as a nontrivial tensor product), solidity (inside M , there is no room for a nonamenable subalgebra and a diffuse subalgebra to commute) and absence of Cartan subalgebras (the impossibility to write a factor as one coming from a group action or an equiva- lence relation). The strongest possible indecomposability property for a von Neumann algebra M , encompassing primeness, solidity and the absence of Cartan subalgebras was discovered in Ozawa and Popa's breakthrough article [OP07] and called strong solidity. Definition. Let M be any diffuse von Neumann algebra. Following [OP07], we say that M is strongly solid if for any diffuse amenable von Neumann subalgebra Q ⊂ M with faithful normal conditional expectation EQ : M → Q, the normalizer NM (Q)′′ generated by NM (Q) := {u ∈ U (M ) uQu∗ = Q} remains amenable. In [OP07], the free group factors L(Fn), 2 ≤ n ≤ +∞, were shown to be strongly solid. This result strengthens both Voiculescu's [Vo95] proving that the free group factors have no Cartan subalgebra and Ozawa's [Oz03] proving that the free group factors are solid. The type III counterparts of the free group factors are Shlyakhtenko's free Araki–Woods factors [Sh96], defined via Voiculescu's free Gaussian functor [Vo85, VDN92]. Although free Araki– Woods factors were shown to be solid and to have no Cartan subalgebra [HR10], strong solidity remained an open problem. So far, there were even no examples of strongly solid type III factors altogether. As we explain in detail below, the strong solidity of a type III factor M is closely 2010 Mathematics Subject Classification. 46L10, 46L54, 46L36. Key words and phrases. Free group factors; Free Araki–Woods factors; Popa's deformation/rigidity theory; Strong solidity; Type III factors. RB is supported by NSF Career Grant DMS 1253402. CH is supported by ERC Starting Grant GAN 637601. SV is supported by ERC Consolidator Grant 614195, and by long term structural funding – Methusalem grant of the Flemish Government. 1 2 R´EMI BOUTONNET, CYRIL HOUDAYER, AND STEFAAN VAES related to a relative strong solidity property of its continuous core c(M ), which is a semifinite von Neumann algebra. The main results of [OP07] apply to the finite corners pc(M )p. But, in order to be applicable to c(M ), we need to control, inside pc(M )p, not only normalizers but also so-called groupoid-normalizers or stable normalizers. That is precisely the problem that we solve in the present paper and that allows us to prove that all free Araki–Woods factors are strongly solid. Following [Sh96], to any orthogonal representation U : R y HR on a real Hilbert space, one associates the free Araki–Woods von Neumann algebra Γ(HR, U )′′, which comes equipped with the free quasi-free state ϕU (see Section 2 for a detailed construction). Free Araki–Woods factors were first studied in the framework of Voiculescu's free probability theory. A complete description of their type classification as well as fullness and computation of Connes's Sd and τ invariants was obtained in [Sh96, Sh97a, Sh97b, Sh02] (see also the survey [Va04]). We have Γ(HR, id)′′ ∼= L(Fdim(HR)) when U = 1HR and Γ(HR, U )′′ is a full type III factor when U 6= 1HR . Moreover, the free Araki–Woods factor Γ(HR, U )′′ admits a discrete decomposition in the sense of [Co74] if and only if the orthogonal representation U : R y HR is almost periodic. The class of free Araki–Woods factors is quite large. Indeed, there are uncountably many pairwise nonisomorphic type III1 free Araki–Woods factors that admit a discrete decomposition [Sh96] as well as uncountably many that do not [Sh02]. More recently, free Araki–Woods factors were studied using Popa's deformation/rigidity theory. This new approach allowed to obtain various indecomposability results in [Ho08, HR14] and to show that free Araki–Woods factors satisfy the complete metric approximation property (CMAP) [HR10, Theorem A] and have no Cartan subalgebra [HR10, Theorem B]. The following is then our main result. Main theorem. For every orthogonal representation U : R y HR such that dim HR ≥ 2, the free Araki–Woods factor Γ(HR, U )′′ is strongly solid. The main step to prove this result is to adapt the proof of Ozawa–Popa's [OP07, Theorem 3.5] so as to cover as well the groupoid-normalizer or stable normalizer of Q ⊂ M , defined as the von Neumann algebra generated by {x ∈ M xQx∗ ⊂ Q and x∗Qx ⊂ Q}. We thus prove in particular that for any diffuse amenable Q ⊂ L(Fn), the stable normalizer of Q remains amenable. To prove that a free Araki–Woods factor M = Γ(HR, U )′′ is strongly solid, we proceed as follows. Fix a diffuse amenable von Neumann subalgebra Q ⊂ M with expectation (meaning that there exists a faithful normal conditional expectation EQ : M → Q). We have to prove that P := NM (Q)′′ remains amenable. Using Connes's continuous decomposition [Co72], we have natural inclusions of the semifinite continuous cores c(Q) ⊂ c(P ) ⊂ c(M ). In general, it is not true that c(P ) is contained in the normalizer of c(Q). However, since M is solid (see e.g. [HR14, Theorem A]), we may replace Q by Q ∨ (Q′ ∩ M ) and assume that Q′ ∩ M = Z(Q). Then, c(P ) is contained in the normalizer of c(Q) (see Lemma 4.1) and by Takesaki's duality theorem [Ta03, Theorem X.2.3], it suffices to show that Nc(M )(c(Q))′′ is amenable. Cutting down by any nonzero finite projection in c(Q), we obtain the inclusions of finite (tracial) von Neumann algebras Q ⊂ P ⊂ M where Q = pc(Q)p, P = p(Nc(M )(c(Q))′′)p and M = pc(M )p. It is important to point out that P need not be contained in the normalizer of Q, but is always contained in the stable normalizer of Q. By [HR10, Theorem A], the tracial von Neumann algebra M has CMAP and it has a natural malleable deformation in the sense of [Po03]. So we are exactly in the setting of [OP07], except that we need to extend their main result on weak compactness to stable normalizers. We do this in Proposition 3.6. STRONG SOLIDITY OF FREE ARAKI–WOODS FACTORS 3 Acknowlegment. R.B. is grateful to Adrian Ioana for many stimulating discussions on this project. 2. Preliminaries 2.1. Background on σ-finite von Neumann algebras. For any von Neumann algebra M , we denote by Z(M ) the center of M , by U (M ) the group of unitaries in M and by (M, L2(M ), J, L2(M )+) the standard form of M . We say that an inclusion of von Neumann algebras P ⊂ M is with expectation if there exists a faithful normal conditional expectation EP : M → P . We say that a σ-finite von Neumann algebra M is tracial if it is endowed with a faithful normal tracial state τ . Let M be any σ-finite von Neumann algebra with predual M∗ and ϕ ∈ M∗ any faithful state. We denote by σϕ the modular automorphism group of the state ϕ. The continuous core of M with respect to ϕ, denoted by cϕ(M ), is the crossed product von Neumann algebra M ⋊σϕ R. The natural inclusion πϕ : M → cϕ(M ) and the unitary representation λϕ : R → cϕ(M ) satisfy the covariance relation λϕ(t)πϕ(x)λϕ(t)∗ = πϕ(σϕ t (x)) for all x ∈ M and all t ∈ R. Put Lϕ(R) = λϕ(R)′′. There is a unique faithful normal conditional expectation ELϕ(R) : cϕ(M ) → Lϕ(R) satisfying ELϕ(R)(πϕ(x)λϕ(t)) = ϕ(x)λϕ(t) for all x ∈ M and all t ∈ R. The faithful normal semifinite (fns) weight defined by f 7→RR exp(−s)f (s) ds on L∞(R) gives rise to a fns weight Trϕ on Lϕ(R) via the Fourier transform. The formula Trϕ = Trϕ ◦ ELϕ(R) extends it to a fns trace on cϕ(M ). Because of Connes's Radon–Nikodym cocycle theorem [Co72, Th´eor`eme 1.2.1] (see also [Ta03, Theorem VIII.3.3]), the semifinite von Neumann algebra cϕ(M ) together with its trace Trϕ does not depend on the choice of ϕ in the following precise sense. If ψ ∈ M∗ is another faithful state, there is a canonical surjective ∗-isomorphism Πϕ,ψ : cψ(M ) → cϕ(M ) such that Πϕ,ψ ◦ πψ = πϕ and Trϕ ◦ Πϕ,ψ = Trψ. Note however that Πϕ,ψ does not map the subalgebra Lψ(R) ⊂ cψ(M ) onto the subalgebra Lϕ(R) ⊂ cϕ(M ) (and hence we use the symbol Lϕ(R) instead of the usual L(R)). 2.2. Free Araki–Woods factors. Let HR be any real Hilbert space and U : R y HR any orthogonal representation. Denote by H = HR ⊗R C = HR ⊕ iHR the complexified Hilbert space, by I : H → H : ξ + iη 7→ ξ − iη the canonical anti-unitary involution on H and by A the infinitesimal generator of U : R y H, that is, Ut = Ait for all t ∈ R. Observe that A−1+1 )1/2ζ defines an isometric embedding of HR into H. Moreover, we j : HR → H : ζ 7→ ( have IAI = A−1. Put KR := j(HR). It is easy to see that KR∩ iKR = {0} and that KR + iKR is dense in H. 2 We introduce the full Fock space of H: F(H) = CΩ ⊕ ∞Mn=1 H ⊗n. The unit vector Ω is called the vacuum vector. For all ξ ∈ H, define the left creation operator ℓ(ξ) : F(H) → F(H) by (cid:26) ℓ(ξ)Ω = ξ, ℓ(ξ)(ξ1 ⊗ ··· ⊗ ξn) = ξ ⊗ ξ1 ⊗ ··· ⊗ ξn. We have kℓ(ξ)k∞ = kξk and ℓ(ξ) is an isometry if kξk = 1. For all ξ ∈ KR, put W (ξ) := ℓ(ξ) + ℓ(ξ)∗. The crucial result of Voiculescu [VDN92, Lemma 2.6.3] is that the distribution of the self-adjoint operator W (ξ) with respect to the vector state ϕU = h· Ω, Ωi is the semicircular law of Wigner supported on the interval [−2kξk, 2kξk]. 4 R´EMI BOUTONNET, CYRIL HOUDAYER, AND STEFAAN VAES Definition 2.1 (Shlyakhtenko, [Sh96]). Let U : R y HR be any orthogonal representation of R on a real Hilbert space HR. The free Araki–Woods von Neumann algebra associated with U : R y HR is defined by Γ(HR, U )′′ := {W (ξ) : ξ ∈ KR}′′. The vector state ϕU = h· Ω, Ωi is called the free quasi-free state and is faithful on Γ(HR, U )′′. The modular automorphism group σϕU satisfies the formula σϕU t (W (ξ)) = W (Utξ) for all ξ ∈ KR and all t ∈ R. We also point out that M = Γ(HR, U )′′ satisfies Ozawa's condition (AO) (see e.g. [HI15, Ap- pendix]) and hence is solid by [Oz03, VV05], that is, for any diffuse subalgebra with expectation Q ⊂ M , the relative commutant Q′ ∩ M is amenable. 2.3. Popa's intertwining-by-bimodules. Popa introduced his method of intertwining-by- bimodules in [Po01, Po03]. In the present work, we make use of these results in the context of semifinite von Neumann algebras. Let (M, Tr) be any semifinite σ-finite von Neumann algebra endowed with a fns trace. Let p ∈ M be any nonzero finite trace projection and A ⊂ pM p any von Neumann subalgebra. Let B ⊂ M be any von Neumann subalgebra such that TrB is still semifinite and denote by EB : M → B the unique trace preserving conditional expectation. We say that A embeds into B inside M (and write A ≺M B) if there exists a nonzero projection q ∈ B with Tr(q) < +∞ such that A ≺ qBq in the sense of Popa (inside the finite von Neumann algebra (p ∨ q)M (p ∨ q)). We refer to e.g. [HR10, Section 2.1] for further details. 3. Stable normalizers in II1 factors In this section, we prove stable strong solidity results for II1 factors. As pointed out in [Ho09, Proposition 5.2], strong solidity is preserved under finite amplifications. However as we explain below, there is a priori no reason for strong solidity to be preserved under infinite amplifications. Nevertheless, we prove in this section that for many known cases of strongly solid II1 factors, their infinite amplification remains strongly solid. Definition 3.1. Given a von Neumann algebra M and a von Neumann subalgebra Q ⊂ M , define sNM (Q) := {x ∈ M xQx∗ ⊂ Q and x∗Qx ⊂ Q} . We call the von Neumann algebra sNM (Q)′′ the stable normalizer of Q inside M . The terminology stable normalizer is motivated by Lemma 3.4(3) saying that the stable nor- malizer of Q is given by the normalizer of the stabilization Q ⊗ B(ℓ2(N)) of Q. Note that xy and x∗ belong to sNM (Q) for all x, y ∈ sNM (Q). Also, Q ⊂ sNM (Q) and the linear span of sNM (Q) is a ∗-algebra containing Q and Q′ ∩ M . The polar decomposition x = vx of an element x ∈ sNM (Q) satisfies the following properties: x ∈ Q, v is a partial isometry whose initial projection p = v∗v and final projection q = vv∗ belong to Q and that satisfies vQv∗ = qQq. Assume moreover that Q ⊂ M is with expectation. Then sNM (Q)′′ ⊂ M is also with expec- tation. Indeed, we may choose a faifhful state ϕ ∈ M∗ such that Q ⊂ M is globally invariant under σϕ. Then σϕ t (x) ∈ sNM (Q) for all x ∈ sNM (Q) and all t ∈ R. This implies that sNM (Q)′′ ⊂ M is globally invariant under σϕ and thus sNM (Q)′′ ⊂ M is with expectation by [Ta03, Theorem IX.4.2]. Definition 3.2. We say that a diffuse von Neumann algebra M is stably strongly solid if for every diffuse amenable von Neumann subalgebra with expectation A ⊂ M , we have that sNM (A)′′ remains amenable. STRONG SOLIDITY OF FREE ARAKI–WOODS FACTORS 5 A priori, strong solidity does not imply stable strong solidity. The reason for this is that the stable normalizer has a qualitatively more general behavior than the normalizer, as can be seen as follows. Assume that M is a II1 factor with tracial state τ . Let A ⊂ M be a von Neumann subalgebra. When u ∈ NM (A), then Ad u defines a trace preserving automorphism of A. In particular, the restriction of Ad u to the center of A defines a trace preserving automorphism of Z(A). When v ∈ M is a partial isometry with p = v∗v ∈ A, q = vv∗ ∈ A and vAv∗ = qAq, the restriction of Ad v to the center of A defines a partial automorphism of Z(A) that need not be trace preserving. Writing Z(A) = L∞(X, µ), it even happens quite naturally that the orbit equivalence relation induced by the orbits of all these Ad v is a type III equivalence relation on (X, µ) (see [MV13] for an example where this phenomenon occurs). Note that the stable normalizer of a von Neumann subalgebra Q ⊂ M is contained in the quasi-normalizer, defined as the von Neumann algebra generated by QN M (Q) =nx ∈ M (cid:12)(cid:12)(cid:12) ∃xi, yj ∈ M, xQ ⊂ nXi=1 Qxi and Qx ⊂ mXj=1 yjQo . If Q ⊂ M is with expectation, then QN M (Q)′′ ⊂ M is also with expectation. The proof is entirely analogous to the one showing that sNM (Q)′′ ⊂ M is with expectation. In general, the inclusion sNM (Q)′′ ⊂ QN M (Q)′′ is strict. Nevertheless, when Q is abelian, we have the following result. Proposition 3.3. Let M be a stably strongly solid von Neumann algebra and A ⊂ M a diffuse abelian von Neumann subalgebra with expectation. Then QN M (A)′′ is amenable. Proof. Since A is abelian, QN M (A)′′ is generated by partial isometries v ∈ M with right support p = v∗v and left support q = vv∗ belonging to A′ ∩ M and with vAv∗ = Aq. Writing Q = A ∨ (A′ ∩ M ), it follows that QN M (A)′′ ⊂ sNM (Q)′′. Since M is in particular solid, Q is diffuse and amenable. By stable strong solidity and since Q ⊂ M is with expectation, we conclude that sNM (Q)′′ is amenable and thus also QN M (A)′′ is amenable since QN M (A)′′ ⊂ sNM (Q)′′ is with expectation. 3.1. Properties of the stable normalizer. The set sNM (Q) behaves well under amplifica- tions/reductions and is itself a stable version of the normalizer NM (Q). In [JP81, Lemma 2.1] and [Po03, Lemma 3.5], a detailed analysis of normalizing unitaries versus amplifica- tions/reductions was made and several key techniques were introduced. We only need the following easy and well known lemma (see e.g. [FSW10, Lemma 3.2] for a proof of the last statement, also based on [JP81, Lemma 2.1]). Lemma 3.4. Let M be a von Neumann algebra and Q ⊂ M a von Neumann subalgebra. (cid:3) (1) For every projection q ∈ Q, we have qsNM (Q)q = sNqM q(qQq). (2) For every Hilbert space K, we have sNM ⊗B(K)(Q ⊗ B(K))′′ = sNM (Q)′′ ⊗ B(K). (3) If Q is σ-finite and K = ℓ2(N), we have sNM (Q)′′ ⊗ B(K) = NM ⊗B(K)(Q ⊗ B(K))′′ . We also need the following technical lemma providing an explicit dilation of a partial isometry in sNM (Q) to a normalizing unitary of an infinite amplification of Q. We need this explicit version to get as a conclusion that for every amenable Q and every fixed x ∈ sNM (Q), the von Neumann algebra (Q ∪ {x, x∗})′′ remains amenable. Again the method of proof is basically given by [JP81, Lemma 2.1]. Lemma 3.5. Let M be a von Neumann algebra and Q ⊂ M a σ-finite von Neumann subalgebra. Define eQ = Q ⊗ B(ℓ2(N)) ⊗ ℓ∞(N) and view eQ as a von Neumann subalgebra of fM = M ⊗ 6 R´EMI BOUTONNET, CYRIL HOUDAYER, AND STEFAAN VAES x ⊗ e = u(a ⊗ e) = (b ⊗ e)u . B(ℓ2(N × N)). Denote by e ∈ B(ℓ2(N × N)) the minimal projection given by the unit vector δ0 ⊗ δ0. For every x ∈ sNM (Q), there exist u ∈ N fM (eQ) and a, b ∈ Q such that If Q is amenable and x ∈ sNM (Q), then (Q ∪ {x, x∗})′′ remains amenable. Proof. Denote by e00 ∈ B(ℓ2(N)) the minimal projection given by the unit vector δ0. By taking the polar decomposition of x, we may assume that x = v is a partial isometry with p = v∗v and q = vv∗ belonging to Q and vQv∗ = qQq. We prove that there exists a u ∈ N fM (eQ) such that v ⊗ e = u(p⊗ e) = (q ⊗ e)u. Denote by zp the central support of p in Q and similarly define zq. Define Q1 = Q⊗ B(ℓ2(N)) and view Q1 as a von Neumann subalgebra of M1 = M ⊗ B(ℓ2(N)). In Q1, the projection zp⊗1 is equivalent with infinitely many copies of p⊗e00. So we can take a nwn = p⊗e00 for all n andPn wnw∗ sequence of partial isometries wn ∈ Q1 such that w∗ n = zp⊗1. We make this choice such that w0 = p ⊗ e00. Similarly take a sequence of partial isometries nvn = q ⊗ e00 and Pn vnv∗ vn ∈ Q1 such that v∗ n = zq ⊗ 1. Also here, we make this choice such that v0 = q ⊗ e00. Define V =Pn vn(v ⊗ e00)w∗ n. Note that V ∗V = zp ⊗ 1, V V ∗ = zq ⊗ 1 and V Q1V ∗ = Q1(zq ⊗ 1). By construction, V (p ⊗ e00) = v ⊗ e00 = (q ⊗ e00)V . Identify Z(Q) = L∞(X, µ) for some probability space (X, µ). Take U ,V ⊂ X such that zp = 1U and zq = 1V . The restriction of Ad V to Z(Q1) = Z(Q)⊗1 induces a nonsingular transformation ϕ : U → V. Denote by R the countable nonsingular equivalence relation on (X, µ) generated by the graph of ϕ. Using powers of V and V ∗, it follows that every partial transformation ψ in the full pseudogroup of R is given by the restriction of Ad W to Z(Q1) for some partial isometry W ∈ M1 such that z = W ∗W and z′ = W W ∗ belong to Z(Q1) and W Q1W ∗ = Q1z′. Define the equivalence relation eR on eX = X × N given by (x, i)eR(y, j) iff xRy. We say that two Borel subsets of eX are equivalent if they are, up to measure zero, the range (resp. domain) of an element of the full pseudogroup [[eR]]. By construction, the sets U × {0} and V × {0} are equivalent. Being properly infinite with eR-saturations equal to eX, also the sets eX \ (U × {0}) and eX \ (V ×{0}) are equivalent. This means that we can find a partial isometry W ∈ fM such that W ∗W = 1 − zp ⊗ 1 ⊗ e00, W W ∗ = 1 − zq ⊗ 1 ⊗ e00 and W eQW ∗ = eQ(1 − zq ⊗ 1 ⊗ e00). The unitary u = V ⊗ e00 + W belongs to N fM (eQ) and satisfies v ⊗ e = u(p ⊗ e) = (q ⊗ e)u. Finally assume that Q is amenable and x ∈ sNM (Q). First, replace M by (Q ∪ {x, x∗})′′. Then, take u as above. Since eQ is amenable and u normalizes eQ, also eP := (eQ ∪ {u, u∗})′′ is amenable. Since the von Neumann algebra (1 ⊗ e)eP (1 ⊗ e) contains Q and x, it must be equal to M ⊗ e and we conclude that M is amenable. 3.2. A general weak compactness argument. In [OP07, Theorem 3.5], Ozawa and Popa proved the following seminal result: if a tracial von Neumann algebra (M, τ ) has the complete metric approximation property (CMAP) and A ⊂ M is any amenable von Neumann subalgebra, then the action of the normalizer NM (A) on A given by conjugacy is weakly compact in the sense of [OP07, Definition 3.1], meaning that there exists a net of unit vectors ξn ∈ L2(A⊗ Aop) such that (cid:3) • limn kξn − (a ⊗ a)ξnk2 = 0 for all a ∈ U (A), • limn kξn − Ad(u ⊗ u)ξnk2 = 0 for all u ∈ NM (A), • limnh(a ⊗ 1)ξn, ξni = τ (a) for all a ∈ A. Here, we denote by Aop the opposite von Neumann algebra of A. We also denote a = (a∗)op for every a ∈ A. We adapt the proof of [OP07, Theorem 3.5] to also cover conjugation by elements x ∈ sNM (A). STRONG SOLIDITY OF FREE ARAKI–WOODS FACTORS 7 v = zr x. We denote by αv : Z(A)zr x as the support projection of EZ (x∗x) and we define zl Fix a tracial von Neumann algebra (M, τ ) and a von Neumann subalgebra A ⊂ M . Denote by EZ the unique trace preserving conditional expectation of A onto Z(A). For every x ∈ sNM (A), we define zr x as the support of EZ (xx∗). Write x = vx for the polar decomposition of x ∈ sNM (A) and let p = v∗v and q = vv∗. Then we have x ∈ A and vAv∗ = qAq so that v ∈ sNM (A). Note that we have zl v = zl x and zr v the unique ∗-isomorphism determined by va = αv(a)v for all a ∈ Z(A)zr v. Multiplying by x ∈ A on the right hand side, we obtain xa = vxa = vax = αv(a)vx = αv(a)x for all a ∈ Z(A)zr v. Letting αx = αv, we obtain xa = αx(a)x for all a ∈ Z(A)zr x. The main difficulty comes from the fact that αx need not be trace preserving. We denote by ∆x the Radon–Nikodym derivative between τ and τ ◦ αx, i.e. ∆x is the unique positive self-adjoint nonsingular operator affiliated with Z(A)zl x satisfying τ (∆xαx(a)) = τ (a) for all a ∈ Z(A)zr x. Note that ∆xy = ∆xαx(∆y) for all x, y ∈ sNM (A). Also note that ∆x = EZ (xx∗) αx(EZ (x∗x)−1). We need the following notation. v → Z(A)zl sN 0 M (A) = {x ∈ sNM (A) ∃δ > 0 such that EZ (x∗x) ≥ δzr x and EZ (xx∗) ≥ δzl x } . M (A) for every n. In particular, sN 0 Note that for every x ∈ sNM (A), we can choose a sequence of projections zn ∈ Z(A) such that zn → 1 strongly and xzn ∈ sN 0 M (A) generates the same von Neumann algebra as sNM (A). For every x ∈ sN 0 M (A), the Radon–Nikodym derivative ∆x is a bounded invertible operator in Z(A)zl x. Proposition 3.6 (Weak compactness). Let (M, τ ) be a tracial von Neumann algebra with the CMAP and take an amenable von Neumann subalgebra A ⊂ M . Then there exists a net of positive vectors ξn ∈ L2(A ⊗ Aop) such that (1) limn k(a ⊗ 1)ξn − (1 ⊗ aop)ξnk2 = 0, for all a ∈ A ; (2) limn k(x ⊗ 1)ξn(x∗∆1/2 (3) limnh(x ⊗ 1)ξn, ξni = τ (x), for all x ∈ M . x ⊗ 1) − (1 ⊗ xop)ξn(1 ⊗ x)k2 = 0, for all x ∈ sN 0 M (A) ; Also, for every partial isometry v ∈ sN 0 M ⊗ M op and a sequence of elements T (v, k) in the unit ball of M ⊗alg M op such that M (A), there exists an element T (v) in the unit ball of (3.1) n k(v ⊗ 1)ξn − (1 ⊗ vop)ξnT (v)k2 = 0 , lim n k(v∗ ⊗ 1)ξn − (1 ⊗ v)ξnT (v)∗k2 = 0 , lim k (cid:16)lim sup lim k (cid:16)lim sup k(v ⊗ 1)ξn − (1 ⊗ vop)ξnT (v, k)k2(cid:17) = 0 , k(v∗ ⊗ 1)ξn − (1 ⊗ v)ξnT (v, k)∗k2(cid:17) = 0 . lim n n Proof. Since M has CMAP, we can take a net of finite rank, normal, completely bounded maps ϕn : M → M such that limn kϕnkcb = 1 and limn kϕn(x)− xk2 = 0 for all x ∈ M . Exactly as in the proof of [OP07, Theorem 3.5], we then define, for every amenable von Neumann subalgebra Q ⊂ M , the normal functionals µQ n (a ⊗ bop) = τ (ϕn(a)b) for all a, b ∈ Q and satisfying limn kµQ n k = 1. We define the normal states ωQ n ∈ (Q ⊗ Qop)∗ given by n = kµQ n . Since limn µQ ωQ n (a ⊗ a) = 1 for all a ∈ U (Q), we get, still in the same way as in the proof of [OP07, Theorem 3.5], that limn kµQ n k = 0 and limn kωQ (3.2) n k = 0 for all a ∈ U (Q). This implies that n ∈ (Q ⊗ Qop)∗ given by µQ n k = 0, limn k(a ⊗ a) · ωQ n · (a ⊗ a) − ωQ n k−1 µQ n − ωQ n − ωQ n kωQ lim n · (a ⊗ 1) − ωQ n · (1 ⊗ aop)k = 0 n k(a ⊗ 1) · ωQ lim n − (1 ⊗ aop) · ωQ n k = 0 and for all a ∈ Q. 8 R´EMI BOUTONNET, CYRIL HOUDAYER, AND STEFAAN VAES n as an element of L1(Q ⊗ Qop)+. We define ξn = (ωA We view ωQ ξn ∈ L2(A ⊗ Aop)+ satisfies the conclusions of the proposition. Properties (1) and (3) hold immediately. Since we already have (1), it suffices to prove property (2) when x = v is a partial isometry in sN 0 M (A). Fix such a v and write q = vv∗, p = v∗v. Define Dr = (EZ (p))1/2 and Dl = (EZ (q))1/2. Denote by zr, zl the support projections of Dr, Dl. Then, Dr (resp. Dl) are invertible operators in Z(A)zr (resp. Z(A)zl) and ∆1/2 r ). Put Q = (A ∪ {v, v∗})′′. By Lemma 3.5, Q is amenable. It then follows from (3.2) that (3.3) n )1/2 and prove that the net v = Dl αv(D−1 n · (v∗ ⊗ 1) − (1 ⊗ vop) · ωQ The restriction of µQ n k1 = 0. Because v ∈ sNM (A), we have that EA(vyv∗) = vEA(y)v∗ for all y ∈ M . Applying EA⊗Aop to (3.3), we conclude that n . Therefore, limn kEA⊗Aop (ωQ n k(v ⊗ 1) · ωQ lim n to A⊗ Aop equals µA n · (1 ⊗ v)k1 = 0 . n )− ωA (3.4) n k(v ⊗ 1) · ωA lim n · (v∗ ⊗ 1) − (1 ⊗ vop) · ωA n · (1 ⊗ v)k1 = 0 . By Lemma 3.9 below, we can take sequences of elements ai, bj ∈ A such that ∞Xj=0 j bj = D−2 b∗ (3.5) i ai = D−2 a∗ ∞Xj=0 ∞Xi=0 ∞Xi=0 j = zr i = zl aia∗ bj b∗ q , , , l r p . We make this choice such that a0 = q and b0 = p. Consider the von Neumann algebra N = B(ℓ2(N2)⊕C)⊗M ⊗M op with its canonical semifinite trace Tr⊗τ and associated 1-norm k · k1 and 2-norm k · k2. View B(C, ℓ2(N2)) ⊂ B(ℓ2(N2)⊕C) and denote by eij ∈ B(C, ℓ2(N2)) the operator given by eij(µ) = µδij, where (δij)i,j∈N is the canonical orthonormal basis of ℓ2(N2). We will identify B(C, C)⊗ M ⊗ M op = CeC⊗ M ⊗ M op with M ⊗ M op. Observe that this identification preserves the 1-norm k · k1 and the 2-norm k · k2. Define V ∈ B(C, ℓ2(N2)) ⊗ M ⊗ 1 given by V =Xi,j Note that V is a well defined bounded operator satisfying V ∗V = zr ⊗ 1. We similarly define W ∈ B(C, ℓ2(N2)) ⊗ 1 ⊗ M op given by W =Xi,j eij ⊗ 1 ⊗ (aivb∗ eij ⊗ Dlaivb∗ j ⊗ 1 . j Dr)op and note that W ∗W = 1 ⊗ zop We claim that (3.4) together with properties (1) and (3) implies that . l r ⊗ 1) ωA n kV (D2 lim (3.6) For every finite subset F ⊂ N2, we define VF and WF in the same way as V and W by only summing over (i, j) ∈ F. Note that kVFk ≤ 1 for all F ⊂ N2 and note that kV − VFk2 can be made arbitrarily small. For all U1, U2 ∈ B(C, ℓ2(N2)) ⊗ M ⊗ 1, we find using the Cauchy–Schwarz inequality and property (3) that n V ∗ − W (1 ⊗ (D2 n W ∗k1 = 0 . l )op) ωA lim sup n kU1 ωA n U ∗ 2k1 ≤ lim sup n kU1 ξnk2 kU2 ξnk2 = kU1k2kU2k2 . The same inequality holds for all U1, U2 ∈ B(C, ℓ2(N2)) ⊗ 1 ⊗ M op. From (3.4) and property (1), we immediately get that for every finite subset F ⊂ N2, n kVF (D2 lim r ⊗ 1) ωA n V ∗ F − WF (1 ⊗ (D2 l )op) ωA n W ∗ Fk1 = 0 . By the preceding discussion, we conclude that also (3.6) holds. STRONG SOLIDITY OF FREE ARAKI–WOODS FACTORS 9 Because Dr belongs to the center of A and V ∗V = zr ⊗ 1, the element V (Dr ⊗ 1) ξn V ∗ is the positive square root of V (D2 l ) ξn W ∗ is the positive square root of W (1 ⊗ (D2 (3.7) l )op) ωA n W ∗. The Powers–Størmer inequality in N then implies that n kV (Dr ⊗ 1) ξn V ∗ − W (1 ⊗ Dop lim n V ∗. Similarly, W (1 ⊗ Dop l ) ξn W ∗k2 = 0 . r ⊗ 1) ωA Multiplying on the left with e∗ 00 ⊗ 1 ⊗ 1 and on the right with e00 ⊗ 1 ⊗ 1, we find that n k(DlvDr ⊗ 1) ξn (v∗Dl ⊗ 1) − (1 ⊗ (DlvDr)op) ξn (1 ⊗ vDr)k2 = 0 . lim Since vDr = αv(Dr)v, using (1) and multiplying on the left with D−1 right with αv(D−1 l ⊗ (D−1 r )op and on the r ) ⊗ 1, we find that n k(v ⊗ 1) ξn (v∗Dlαv(D−1 lim r ) ⊗ 1) − (1 ⊗ vop) ξn (1 ⊗ v)k2 = 0 . v = Dlαv(D−1 Since ∆1/2 the unit ball of M ⊗ M op. Multiplying (3.7) on the left with e∗ V , we find that r ), we finally obtain (2). Denote T (v) = W ∗V . Then T (v) belongs to 00 ⊗ 1 ⊗ 1 and on the right with n k(DlvDr ⊗ 1) ξn − (1 ⊗ (DlvDr)op) ξn T (v)k2 = 0 . lim Using (1) and multiplying on the left with D−1 (3.1) holds. l ⊗ (D−1 r )op, we find that the first estimate in Define Vk by the same formula as V , but only summing over i, j = 1, . . . , k. Define T (v, k) = W ∗Vk and note that T (v, k) belongs to the unit ball of M ⊗alg M op. Write V ∗Vk = dk ⊗ 1, where dk ∈ A and limk kzr − dkk2 = 0. Multiplying (3.7) on the left with e∗ 00 ⊗ 1 ⊗ 1 and on the right with Vk and reasoning as before, we find that lim sup n k(v ⊗ 1) ξn (dk ⊗ 1) − (1 ⊗ vop) ξn T (v, k)k2 = 0 . Since limn kξn(dk ⊗ 1) − ξn(zr ⊗ 1)k2 = kdk − zrk2, we conclude that also the third estimate in (3.1) holds. Replacing in the above reasoning v by v∗, in (3.5), we interchange to roles of ai and bj. We then get that Drbj v∗a∗ T (v∗) =Xi,j T (v∗, k) = (Dr ⊗ (D−1 r )op) T (v, k)∗ (D−1 l ⊗ Dop l ) . i ⊗ (Dlaivb∗ j )op = (Dr ⊗ (D−1 r )op) T (v)∗ (D−1 l ⊗ Dop l ) and r )op) − (1 ⊗ v)ξnk2 = 0 and limn k(v∗ ⊗ 1)ξn(D−1 Since limn k(1 ⊗ v)ξn(Dr ⊗ (D−1 (v∗ ⊗ 1)ξnk2 = 0, it follows that also the second and fourth estimate in (3.1) hold. l ⊗ Dop l ) − (cid:3) 3.3. Consequences of weak compactness. The approximate invariance given by weak com- pactness as in (3.1) combines very well with deformation/rigidity theory. In particular, we can apply to ξn any s-malleable deformation of M , in the sense of [Po03], i.e. a trace preserv- ing inclusion (M, τ ) ⊂ (fM , τ ) together with a strongly continuous one-parameter group of trace preserving automorphisms (αt)t∈R of fM and a trace preserving period 2 automorphism β ∈ Aut(fM , τ ) satisfying β ◦ αt = α−t ◦ β and βM = id. Following [PV11, Definition 2.3], for any tracial von Neumann algebras P ⊂ (M, τ ) and (Q, τ ), we say that an M -Q-bimodule MKQ is left P -amenable if there exists a P -central state Ω on B(K) ∩ (Qop)′ whose restriction to M equals τ . The methods of [OP07, Section 4] can be applied and give the following result. 10 R´EMI BOUTONNET, CYRIL HOUDAYER, AND STEFAAN VAES Proposition 3.7. Let (M, τ ) be a tracial von Neumann algebra with the CMAP and A ⊂ M a von Neumann subalgebra. Assume that ξn ∈ L2(A ⊗ Aop) is a net of positive vectors satisfying the conclusion of Proposition 3.6. Also assume that (αt)t∈R is an s-malleable deformation as above. Then at least one of the following statements holds. (1) We have that limt→0(cid:0)supa∈U (A) kαt(a) − ak2(cid:1) = 0. (2) Writing P = sNM (A)′′, there exists a nonzero projection p ∈ Z(P ) such that the pM p- M -bimodule pL2(fM ⊖ M ) is left P p-amenable. Proof. Denote by eM : L2(fM ) → L2(M ) the orthogonal projection and write e⊥ One of the following properties holds. M = 1 − eM . (1) For every ε > 0, there exists a t0 > 0 such that for every t ∈ R with t ≤ t0, we have (2) There exists an ε > 0 and a sequence tk ∈ R such that limk tk = 0 and such that for lim supn k(e⊥ M αt ⊗ id)(ξn)k2 ≤ ε. every k, we have lim supn k(e⊥ M αtk ⊗ id)(ξn)k2 > ε. We prove that the first (resp. second) of these properties implies the first (resp. second) con- clusion in the proposition. Assume that (1) holds. By [Po06, Lemma 2.1], the following transversality condition holds for every t ∈ R and ξ ∈ L2(M ⊗ M op). k(α2t ⊗ id)(ξ) − ξk2 ≤ 2k(e⊥ M αt ⊗ id)(ξ)k2 . Choose ε > 0. We can then take a t0 > 0 such that for every t ∈ R with t ≤ t0, we have lim supn k(αt ⊗ id)(ξn) − ξnk2 ≤ ε. Fix t ∈ R with t ≤ t0. We prove that kαt(a) − ak2 ≤ 2√ε for every a ∈ U (A), so that the first conclusion of the proposition indeed holds. Fix a ∈ U (A). Because n h(αt(a) ⊗ a) (αt ⊗ id)(ξn), (αt ⊗ id)(ξn)i = 1 lim and because lim supn k(αt ⊗ id)(ξn) − ξnk2 ≤ ε, we get that lim sup n 1 − h(αt(a) ⊗ a)ξn, ξni ≤ 2ε . But the left hand side equals lim sup n 1 − h(αt(a)a∗ ⊗ 1)ξn, ξni = 1 − τ (αt(a)a∗) . So, we have proved that 1 − τ (αt(a)a∗) ≤ 2ε. Then also, kαt(a) − ak2 2 = 2 Re(1 − τ (αt(a)a∗)) ≤ 4ε . Next assume that (2) holds with ε > 0 and (tk)k. We start by proving the following claim : for every x ∈ M and every δ > 0, we have that for small enough t ∈ R lim sup n k(x ⊗ 1)(αt ⊗ id)(ξn) − (αt ⊗ id)((x ⊗ 1)ξn)k2 < δ . Indeed, it suffices to observe that the left hand side equals lim sup n k((α−t(x) − x) ⊗ 1)ξnk2 = kα−t(x) − xk2 and that kα−t(x) − xk2 → 0 as t → 0. Then also for every T ∈ M ⊗alg M op and every δ > 0, we have that for small enough t ∈ R lim sup n k(αt ⊗ id)(ξn)T − (αt ⊗ id)(ξnT )k2 < δ . STRONG SOLIDITY OF FREE ARAKI–WOODS FACTORS 11 We construct a subnet ζi of the net ζj,n = (e⊥ and such that for every partial isometry v ∈ sN 0 i k(v ⊗ 1)ζi − (1 ⊗ vop)ζiS(v, i)k2 = 0 , lim lim sup (3.8) k(x ⊗ 1)ζik2 ≤ kxk2 , i M αtj ⊗ id)(ξn) such that ε ≤ kζik2 ≤ 1 for every i M (A) and every x ∈ M , we have i k(v∗ ⊗ 1)ζi − (1 ⊗ v)ζiS(v, i)∗k2 = 0 , lim where the S(v, i) are elements in the unit ball of M ⊗alg M op. The index set of the net ζi is given by i = (F ,G, δ) where F is a finite set of partial isometries in sN 0 M (A), G ⊂ M is a finite subset and δ > 0. Given i = (F ,G, δ) and using the notation of (3.1), we take k large enough such that for every v ∈ F, we have that k(v⊗1)ξn−(1⊗vop)ξnT (v, k)k2 < δ k(v∗⊗1)ξn−(1⊗v)ξnT (v, k)∗k2 < δ . lim sup lim sup and n n Using the claim in the previous paragraph, we then take j large enough such that lim sup n k(x ⊗ 1)(αtj ⊗ id)(ξn) − (αtj ⊗ id)((x ⊗ 1)ξn)k2 < δ for all x ∈ F ∪ G and such that lim sup n k(αtj ⊗ id)(ξn)T (v, k) − (αtj ⊗ id)(ξnT (v, k))k2 < δ for all v ∈ F. We finally take n large enough such that the vector ζi = (e⊥ with S(v, i) := T (v, k) satisfies ε ≤ kζik2 ≤ 1 and M αtj ⊗id)(ξn) together , k(v∗ ⊗ 1)ζi − (1 ⊗ v)ζiS(v, i)∗k2 < 3δ , k(v ⊗ 1)ζi − (1 ⊗ vop)ζiS(v, i)k2 < 3δ k(x ⊗ 1)ζik2 < kxk2 + 2δ for all v ∈ F and all x ∈ G. So we have found the net ζi satisfying (3.8). Denote P = sNM (A)′′ and define S as the commutant of the right M -action on L2(fM ⊖ M ). Taking a subnet of the net ζi, we may assume that the net of positive functionals S → C : T 7→ h(T ⊗ 1)ζi, ζii converges weakly∗ to a positive functional Ω on S. By (3.8), we get for all T ∈ S and all partial isometries v ∈ sN 0 M (A) that Ω(T v) = lim i h(T ⊗ vop)ζiS(v, i), ζii i h(T v ⊗ 1)ζi, ζii = lim i h(T ⊗ 1)ζi, (1 ⊗ v)ζiS(v, i)∗i = lim i h(vT ⊗ 1)ζi, ζii = Ω(vT ) . = lim = lim i h(T ⊗ 1)ζi, (v∗ ⊗ 1)ζii Since kζik2 ≥ ε for all i, also Ω(1) ≥ ε, so that Ω is nonzero. By (3.8), we get that Ω(x∗x) ≤ τ (x∗x) for all x ∈ M . The Cauchy–Schwarz inequality then implies that and Ω(vT − wT )2 ≤ kv − wk2 Ω(T v) − Ω(T w)2 ≤ Ω(T T ∗)kv − wk2 2 Ω(T ∗T ) for all T ∈ S and v, w ∈ M . Since Ω(T v) = Ω(vT ) when v is a partial isometry in sN 0 M (A), taking linear combinations and k · k2-limits, it follows that Ω is a nonzero P -central functional on S. Since ΩM ≤ τ , the restriction of Ω to M is normal. Denote by p ∈ M the support projection of ΩM . Then, p ∈ P ′ ∩ M . Since A ⊂ P and A′ ∩ M ⊂ P , it follows that p ∈ Z(P ). As in [OP07, Theorem 2.1], we conclude that the pM p-M -bimodule pL2(fM ⊖ M ) is left P p-amenable. (cid:3) 2 Propositions 3.6 and 3.7 will be key to prove that free Araki–Woods factors are strongly solid. We also have the following consequence. Theorem 3.8. Let Γ be a countable group with the CMAP that admits a proper 1-cocycle into an orthogonal representation that is weakly contained in the regular representation. Then, M = L(Γ) is stably strongly solid: for every diffuse amenable von Neumann subalgebra A ⊂ M , we have that sNM (A)′′ remains amenable. 12 R´EMI BOUTONNET, CYRIL HOUDAYER, AND STEFAAN VAES Proof. This follows immediately from Propositions 3.6 and 3.7 by using the s-malleable defor- mation associated with a 1-cocycle in [Si10]. (cid:3) We needed the following well known lemma. Lemma 3.9. Let (A, τ ) be a tracial von Neumann algebra. Denote by EZ : A → Z(A) the unique trace preserving conditional expectation. If x, y ∈ A+ satisfy EZ (x) = EZ (y), then there exists a sequence of elements ak ∈ A such that Xk a∗ kak = y . aka∗ k = x and Xk Proof. The result follows from a maximality argument and the equality EZ (a∗a) = EZ (aa∗) for all a ∈ A. (cid:3) Remark 3.10. By Proposition 3.3, for the same groups Γ as in Theorem 3.8, the group von Neumann algebra M = L(Γ) has the following property: for every diffuse abelian von Neumann subalgebra A ⊂ M , the quasi-normalizer QN M (A)′′ is amenable. However, we leave open the question whether QN M (A)′′ is amenable for every diffuse amenable subalgebra A ⊂ M . Note here that in specific case of the free group factors M = L(Fn) and using free entropy dimension, it was proved in [Vo95] that for a diffuse abelian A ⊂ M , the ∗-algebra QN M (A) cannot be dense in M . In [Ha15], this result was generalized to arbitrary diffuse amenable subalgebras A ⊂ M . 3.4. Relative stable strong solidity. In Proposition 3.6, we showed how to adapt the weak compactness of [OP07] so as to cover the stable normalizer sNM (A)′′ rather than the normalizer NM (A)′′. In exactly the same way, the methods of [PV11, Section 5.1] can be extended to the stable normalizer. As a consequence, one obtains the following improvement of [PV11, Theorem 1.6]. Theorem 3.11. Let Γ be a countable group with the CMAP that admits a proper 1-cocycle into an orthogonal representation that is weakly contained in the regular representation. Assume that Γ y (B, τ ) is any trace preserving action on the tracial von Neumann algebra (B, τ ). Put M = B ⋊ Γ and let A ⊂ M be a von Neumann subalgebra that is amenable relative to B. Then at least one of the following statements holds. (1) A ≺M B. (2) sNM (A)′′ remains amenable relative to B. As a consequence of Theorem 3.11, also the results of [Io12, Va13] on the normalizer of sub- algebras A of amalgamated free products M = M1 ∗B M2 generalize to the stable normalizer sNM (A)′′. On the other hand, it is not clear to us whether Proposition 3.6 and Theorems 3.8 and 3.11 remain valid if we replace CMAP by weak amenability because so far, we were unable to extend the methods of [Oz10, Section 4] to the stable normalizer. 4. Proof of the Main Theorem The following general lemma will be the key to deduce strong solidity results for type III factors from structural results of their continuous core. Lemma 4.1. Let Q ⊂ M be any inclusion of σ-finite von Neumann algebras with faithful normal conditional expectation EQ : M → Q. Let ϕ ∈ M∗ be any faithful state such that ϕ ◦ EQ = ϕ. Then for any u ∈ NM (Q) and any t ∈ R, we have u∗σϕ t (u) ∈ Q ∨ (Q′ ∩ M ). STRONG SOLIDITY OF FREE ARAKI–WOODS FACTORS 13 t (x) = σϕu Proof. Put ϕu := ϕ ◦ Ad(u) and note that Q is globally invariant under the modular automor- phism groups σϕ and σϕu. Put ψ := ϕQ and ψu := ϕuQ. Observe that σψu t (x) and σψ t (x) = σϕ t (x) for every t ∈ R and every x ∈ Q. By [Co72, Th´eor`eme 1.2.1], there exists a t = Ad(wt) ◦ σψ σ-strongly continuous map R → U (Q) : t 7→ wt such that σψu for every t ∈ R. t (u)) ◦ σϕ By [Co72, Lemme 1.2.3(c)], we have σϕu t (u)) σϕ t = σϕu for every x ∈ Q, it follows that w∗ Q ∨ (Q′ ∩ M ) for every t ∈ R. Proof of the main theorem. First, observe that every free Araki–Woods factor is contained with expectation in a free Araki–Woods factor of type III1. Indeed, for any orthogonal representation U : R y HR, we have the following inclusion with expectation t (u)) ∈ Q′ ∩ M . Therefore, we have u∗σϕ t = Ad(u∗σϕ t (x) = (u∗σϕ t (u∗σϕ for every t ∈ R. Since t (x) (u∗σϕ t (u))∗ wt σϕ t (x) w∗ t (u) ∈ (cid:3) t t Γ(HR, U )′′ ⊂ Γ(HR ⊕ L2 R(R), U ⊕ λ)′′, R(R) is the regular representation and Γ(HR⊕L2 where λ : R y L2 R(R), U ⊕λ)′′ is a free Araki– Woods factor of type III1. Since strong solidity is preserved under taking diffuse subalgebras with expectation, we may assume that M := Γ(HR, U )′′ is a free Araki–Woods factor of type III1. Let Q ⊂ M be any diffuse amenable von Neumann subalgebra with expectation. We want to prove that P := NM (Q)′′ is amenable. Fix a faithful state ψ ∈ M∗ such that Q is globally invariant under the modular automorphism group σψ. Observe that Q′ ∩ M , NM (Q)′′ and NM (Q ∨ (Q′ ∩ M ))′′ are all globally invariant under the modular automorphism group σψ. Since M is solid by [Oz03, VV05] (see also [HR14, Theorem A]), Q′ ∩ M is amenable and so is Q = Q ∨ (Q′ ∩ M ). Observe that Q′ ∩ M = Z(Q). Since the inclusion NM (Q)′′ ⊂ NM (Q)′′ is with expectation, it suffices to prove that NM (Q)′′ is amenable. Therefore, without loss of generality, we may further assume that Q = Q, that is, Q′ ∩ M = Z(Q). Now, assume by contradiction that P = NM (Q)′′ is not amenable. Then there exists a nonzero central projection z ∈ Z(P ) such that P z has no amenable direct summand. Since Z(P ) ⊂ Q′ ∩ M = Z(Q), we have z ∈ Z(Q). Then we have Qz ⊂ zM z, (Qz)′ ∩ zM z = Z(Qz) and NzM z(Qz)′′ = P z has no amenable direct summand. Since zM z ∼= M , we may replace Q ⊂ M by Qz ⊂ zM z and assume without loss of generality that P = NM (Q)′′ has no amenable direct summand. Claim. We have cψ(P ) ⊂ Ncψ(M )(cψ(Q))′′. Since λψ(t) ∈ U (cψ(Q)), we have λψ(t) ∈ Ncψ(M )(cψ(Q)) for every t ∈ R. Let now u ∈ NM (Q). Then we have πψ(u)πψ(Q)πψ(u)∗ = πψ(Q) ⊂ cψ(Q). Moreover since Q′ ∩ M = Z(Q), Lemma 4.1 shows that we have uσψ t (u∗) ∈ U (Q) for all t ∈ R. Therefore, we have πψ(u)λψ(t)πψ(u)∗ = πψ(uσψ t (u∗))λψ(t) ∈ πψ(Q)λψ(t) ⊂ cψ(Q) for all t ∈ R. Altogether, we have that πψ(u)cψ(Q)πψ(u)∗ ⊂ cψ(Q). Replacing u ∈ NM (Q) by u∗ ∈ NM (Q), we obtain πψ(u)∗cψ(Q)πψ(u) ⊂ cψ(Q) and hence πψ(u)cψ(Q)πψ(u)∗ = cψ(Q). Since cψ(P ) =_{πψ(u), λψ(t) u ∈ NM (Q), t ∈ R} , we finally have cψ(P ) ⊂ Ncψ(M )(cψ(Q))′′. This proves the claim. Since P has no amenable direct summand, cψ(P ) has no amenable direct summand either by [BHR12, Proposition 2.8]. The above claim further implies that Ncψ(M )(cψ(Q))′′ has no amenable direct summand. Denote by ϕ = ϕU the free quasi-free state on M and put M0 := cϕ(M ), Q0 := Πϕ,ψ(cψ(Q)) and P0 := Πϕ,ψ(Ncψ(M )(cψ(Q))′′) = NM0(Q0)′′. 14 R´EMI BOUTONNET, CYRIL HOUDAYER, AND STEFAAN VAES Since Q is diffuse, [HU15, Lemma 2.5] and the paragraph following [HU15, Theorem 2.2] imply that pQ0p ⊀M0 Lϕ(R) for any nonzero finite trace projection p ∈ Πϕ,ψ(Lψ(R)). Take such a projection p ∈ Πϕ,ψ(Lψ(R)) ⊂ Q0, so that (4.1) Since M is a type III1 factor, M0 is a type II∞ factor and hence there exists a unitary u ∈ U (M0) such that upu∗ ∈ Lϕ(R). Therefore, up to conjugating Q0 (and P0) by a unitary in U (M0), we may assume that p ∈ Lϕ(R). We still have (4.1). By [HR10, Theorem A], M has the CMAP and so does M0 by [AD93, Lemma 4.6 and The- orem 4.9]. Following [HR10, Section 4.1], we know that N := pM0p admits an s-malleable pQ0p ⊀M0 Lϕ(R). deformation in the sense of Popa such that the N -N -bimodule L2(eN ⊖ N ) is weakly contained in the coarse N -N -bimodule L2(N ) ⊗ L2(N ). Then [HR10, Theorem 4.3] and (4.1) imply that this deformation does not converge uniformly on U (pQ0p). Put P1 := sNpM0p(pQ0p)′′ and observe that pP0p ⊂ P1. Proposition 3.7 implies that there exists a nonzero projection q ∈ Z(P1) such that the qN q-N -bimodule qL2(eN ⊖ N ) is left P1q-amenable. Therefore, the coarse qN q-N -bimodule is left P1q-amenable by [PV11, Corollary 2.5] which further implies that P1q is amenable by [PV11, Proposition 2.4]. This however contradicts the fact that P0 has no amenable direct summand. (cid:3) 5. Further remarks on stable strong solidity In this section, we clarify the relationship between stable strong solidity and strong solidity. Lemma 3.4(3) shows that if the infinite amplification M ⊗ B(ℓ2(N)) of a diffuse σ-finite von Neumann algebra M is strongly solid then M is stably strongly solid. We will show that the converse is also true. First, we prove the following useful lemma. Lemma 5.1. Let (M, Tr) be any semifinite von Neumann algebra endowed with a fns trace. Assume that N ⊂ M is a subalgebra with expectation EN : M → N such that N ′ ∩ M ⊂ N . Then Tr is semifinite on N and Tr ◦ EN = Tr. Proof. Since M is semifinite and N ⊂ M is with expectation, N is a semifinite von Neumann algebra. Denote by TrN a fns trace on N and define a fns weight ϕ on M by the formula ϕ := TrN ◦ EN . Denote by h the unique positive self-adjoint operator affiliated with M such that ϕ = Tr(h·) (see [PT72, Theorem 5.12]). Note that N centralizes both weights Tr and ϕ. Then for all t ∈ R, the Radon–Nikodym derivative (Dϕ : DTr)t = hit commutes with N . This shows that hit ∈ N ′ ∩ M = Z(N ) for all t ∈ R and hence h is affiliated with the center Z(N ) of N . Since ϕ is semifinite on N and h is a nonsingular operator affiliated with Z(N ), also Tr is semifinite on N . Since ϕ◦ EN = ϕ, also Tr ◦ EN = Tr. Corollary 5.2. Let M be any diffuse σ-finite von Neumann algebra. Then the following facts are true: (cid:3) (1) M is solid if and only if its infinite amplification M ⊗ B(ℓ2(N)) is solid. (2) M is stably strongly solid if and only if its infinite amplification is strongly solid. Proof. First, observe that a direct sum of von Neumann algebras is solid (resp. (stably) strongly solid) if and only if each direct summand is solid (resp. (stably) strongly solid). Next, any solid von Neumann algebra with diffuse center is amenable. Also, the notions of stable strong solidity and strong solidity coincide for properly infinite von Neumann algebras by Lemma 3.4(2, 3). Therefore, we only need to consider the case where M is a II1 factor. Denote by M∞ := M ⊗ B(ℓ2(N)) the infinite amplification of M equipped with the canonical fns trace Tr := τ ⊗ TrB(ℓ2(N)). STRONG SOLIDITY OF FREE ARAKI–WOODS FACTORS 15 (1) If M∞ is solid then M is solid as well since M ⊂ M∞ is with expectation and solidity is preserved under taking diffuse subalgebras with expectation. Assume now that M is solid and take a diffuse subalgebra N ⊂ M∞ with expectation. Take a diffuse abelian subalgebra A ⊂ N with expectation. To prove that N ′ ∩ M∞ is amenable, it is sufficient to prove that A′ ∩ M∞ is amenable since N ′ ∩ M∞ ⊂ A′ ∩ M∞ is with expectation. Since A is abelian, we have A ⊂ A′∩M∞ := Q and hence Q′∩M∞ ⊂ Q. Then Lemma 5.1 implies that the semifinite trace Tr on M∞ remains semifinite on Q. Since Q is diffuse, TrQ is semifinite and Tr(1) = +∞, we may take a sequence of pairwise orthogonal projections pn ∈ Q such that Tr(pn) = 1 for all n and Pn pn = 1. Then for all n, we have that pnQpn = (Apn)′ ∩ pnM∞pn. Since pnM∞pn ∼= M is solid, we have that pnQpn is amenable for all n and we conclude that Q is amenable. (2) Lemma 3.4(3) shows the "if" part. Let us show the "only if" part. Assume that M is stably strongly solid and take a diffuse amenable subalgebra Q ⊂ M∞ with expectation. By (1), M∞ is solid and hence Q := Q ∨ (Q′ ∩ M∞) is also diffuse amenable and NM∞(Q)′′ is contained with expectation inside NM∞(Q)′′. Hence replacing Q by Q if necessary, we may assume that Q′ ∩ M∞ ⊂ Q. Then Lemma 5.1 implies that the semifinite trace Tr on M∞ remains semifinite on Q. Take a sequence of pairwise orthogonal projections pn ∈ Q such that Tr(pn) = 1 for all n and Pn pn = 1. Since pnM∞pn ∼= M is stably strongly solid, we have that sNpnM∞pn(pnQpn)′′ is amenable for all n. Hence pn(NM∞(Q)′′)pn ⊂ pn(sNM∞(Q)′′)pn = sNpnM∞pn(pnQpn)′′ is amenable for all n and we conclude that NM∞(Q)′′ is amenable as well. (cid:3) References [AD93] C. Anantharaman-Delaroche, Amenable correspondences and approximation properties for von Neumann algebras. Pacific J. Math. 171 (1995), 309–341. [BHR12] R. Boutonnet, C. Houdayer, S. Raum, Amalgamated free product type III factors with at most one Cartan subalgebra. Compos. Math. 150 (2014), 143–174. [Co72] A. Connes, Une classification des facteurs de type III. Ann. Sci. ´Ecole Norm. Sup. 6 (1973), 133–252. [Co74] A. Connes, Almost periodic states and factors of type III1. J. Funct. Anal. 16 (1974), 415–445. [Co75] A. Connes, Classification of injective factors. Cases II1, II∞, IIIλ, λ 6= 1. Ann. of Math. 74 (1976), 73–115. [FSW10] J. Fang, R. Smith, S. White, Groupoid normalisers of tensor products: infinite von Neumann algebras. J. Operator Theory 69 (2013), 545–570. [Ha85] U. Haagerup, Connes' bicentralizer problem and uniqueness of the injective factor of type III1. Acta [Ha15] [Ho08] [Ho09] [HI15] Math. 69 (1986), 95–148. B. Hayes, 1-bounded entropy and regularity problems in von Neumann algebras. To appear in Int. Math. Res. Not. IMRN. arXiv:1505.06682 C. Houdayer, Structural results for free Araki–Woods factors and their continuous cores. J. Inst. Math. Jussieu 9 (2010), 741–767. C. Houdayer, Strongly solid group factors which are not interpolated free group factors. Math. Ann. 346 (2010), 969–989. C. Houdayer, Y. Isono, Unique prime factorization and bicentralizer problem for a class of type III factors. Adv. Math. 305 (2017), 402–455. [HR14] C. Houdayer, S. Raum, Asymptotic structure of free Araki–Woods factors. Math. Ann. 363 (2015), 237–267. [HR10] C. Houdayer, ´E. Ricard, Approximation properties and absence of Cartan subalgebra for free Araki– Woods factors. Adv. Math. 228 (2011), 764–802. [HU15] C. Houdayer, Y. Ueda, Asymptotic structure of free product von Neumann algebras. Math. Proc. [Io12] [JP81] Cambridge Philos. Soc. 161 (2016), 489–516. A. Ioana, Cartan subalgebras of amalgamated free product II1 factors. Ann. Sci. ´Ecole Norm. Sup. 48 (2015), 71–130. V.F.R. Jones, S. Popa, Some properties of MASAs in factors. In Invariant subspaces and other topics (Timi¸soara/Herculane, 1981), Operator Theory: Adv. Appl. 6, Birkhauser, Basel-Boston, 1982, pp. 89–102. 16 R´EMI BOUTONNET, CYRIL HOUDAYER, AND STEFAAN VAES [Kr75] W. Krieger, On ergodic flows and the isomorphism of factors. Math. Ann. 223 (1976), 19–70. [MV13] N. Meesschaert, S. Vaes, Partial classification of the Baumslag-Solitar group von Neumann algebras. Doc. Math. 19 (2014), 629–645. [MvN43] F. Murray, J. von Neumann, Rings of operators. IV. Ann. of Math. 44 (1943), 716–808. [Oz03] N. Ozawa, Solid von Neumann algebras. Acta Math. 192 (2004), 111–117. [Oz10] N. Ozawa, Examples of groups which are not weakly amenable. Kyoto J. Math. 52 (2012), 333–344. [OP07] N. Ozawa, S. Popa, On a class of II1 factors with at most one Cartan subalgebra. Ann. of Math. 172 (2010), 713–749. [PT72] G. Pedersen, M. Takesaki, The Radon–Nikodym theorem for von Neumann algebras. Acta Math. [Po01] [Po03] [Po06] [PV11] 130 (1973), 53–87. S. Popa, On a class of type II1 factors with Betti numbers invariants. Ann. of Math. 163 (2006), 809–899. S. Popa, Strong rigidity of II1 factors arising from malleable actions of w-rigid groups, I. Invent. Math. 165 (2006), 369–408. S. Popa, On the superrigidity of malleable actions with spectral gap. J. Amer. Math. Soc. 21 (2008), 981–1000. S. Popa, S. Vaes, Unique Cartan decomposition for II1 factors arising from arbitrary actions of free groups. Acta Math. 212 (2014), 141–198. D. Shlyakhtenko, Free quasi-free states. Pacific J. Math. 177 (1997), 329–368. [Sh96] [Sh97a] D. Shlyakhtenko, Some applications of freeness with amalgamation. J. Reine Angew. Math. 500 (1998), 191–212. [Sh97b] D. Shlyakhtenko, A-valued semicircular systems. J. Funct. Anal. 166 (1999), 1–47. [Sh02] D. Shlyakhtenko, On the classification of full factors of type III. Trans. Amer. Math. Soc. 356 (2004), 4143–4159. T. Sinclair, Strong solidity of group factors from lattices in SO(n, 1) and SU(n, 1). J. Funct. Anal. 260 (2011), 3209–3221. [Si10] [Ta03] M. Takesaki, Theory of operator algebras. II. Encyclopaedia of Mathematical Sciences, 125. Operator [Va04] [Va13] [VV05] Algebras and Non-commutative Geometry, 6. Springer-Verlag, Berlin, 2003. xxii+518 pp. S. Vaes, ´Etats quasi-libres libres et facteurs de type III (d'apr`es D. Shlyakhtenko). S´eminaire Bourbaki, expos´e 937, Ast´erisque 299 (2005), 329–350. S. Vaes, Normalizers inside amalgamated free product von Neumann algebras. Publ. Res. Inst. Math. Sci. 50 (2014), 695–721. S. Vaes, R. Vergnioux, The boundary of universal discrete quantum groups, exactness and factori- ality. Duke Math. J. 140 (2007), 35–84. [Vo85] D.-V. Voiculescu, Symmetries of some reduced free product C∗-algebras. Operator algebras and Their Connections with Topology and Ergodic Theory, Lecture Notes in Mathematics 1132. Springer-Verlag, (1985), 556–588. [Vo95] D.-V. Voiculescu, The analogues of entropy and of Fisher's information measure in free probability theory. III. The absence of Cartan subalgebras. Geom. Funct. Anal. 6 (1996), 172–199. [VDN92] D.-V. Voiculescu, K.J. Dykema, A. Nica, Free random variables. CRM Monograph Series 1. American Mathematical Society, Providence, RI, 1992. Institut de Math´ematiques de Bordeaux, CNRS, Universit´e Bordeaux I, 33405 Talence, FRANCE E-mail address: [email protected] Laboratoire de Math´ematiques d'Orsay, Universit´e Paris-Sud, CNRS, Universit´e Paris-Saclay, 91405 Orsay, FRANCE E-mail address: [email protected] Department of Mathematics, KU Leuven, Celestijnenlaan 200B, B-3001 Leuven, BELGIUM E-mail address: [email protected]
1003.5618
2
1003
2010-07-16T08:20:28
A Note on Dirac Operators on the Quantum Punctured Disk
[ "math.OA", "math-ph", "math-ph" ]
We study quantum analogs of the Dirac type operator $-2\bar{z}\frac{\partial}{\partial\bar{z}}$ on the punctured disk, subject to the Atiyah-Patodi-Singer boundary conditions. We construct a parametrix of the quantum operator and show that it is bounded outside of the zero mode.
math.OA
math
Symmetry, Integrability and Geometry: Methods and Applications SIGMA 6 (2010), 056, 12 pages A Note on Dirac Operators on the Quantum Punctured Disk⋆ Slawomir KLIMEK and Matt MCBRIDE Department of Mathematical Sciences, Indiana University-Purdue University Indianapolis, 402 N. Blackford St., Indianapolis, IN 46202, USA E-mail: [email protected], [email protected] Received March 30, 2010, in final form July 07, 2010; Published online July 16, 2010 doi:10.3842/SIGMA.2010.056 Abstract. We study quantum analogs of the Dirac type operator −2z ∂ ∂z on the punctured disk, subject to the Atiyah -- Patodi -- Singer boundary conditions. We construct a parametrix of the quantum operator and show that it is bounded outside of the zero mode. Key words: operator theory; functional analysis; non-commutative geometry 2010 Mathematics Subject Classification: 46L99; 47B25; 81R60 1 Introduction The main technical and computational part of the Atiyah, Patodi, Singer paper [1] is the initial section containing a study of a nonlocal boundary value problem for the first order differential operators of the form Γ( ∂ ∂t + B) on the semi-infinite cylinder R+ × Y , where t ∈ R+ and B, Γ live on the boundary Y . The novelty of the paper was the boundary condition, now called the APS boundary condition, that involved a spectral projection of B. The authors explicitly compute and estimate the fundamental solutions on the cylinder. This is later used to construct a parametrix for the analogical boundary value problem on a manifold with boundary by gluing it with a contribution from the interior, see also [2]. The present paper aims, in a special case, to reproduce such results in the noncommutative setup of [5]. A similar but different study of an example of APS boundary conditions in the context of noncommutative geometry is contained in [4]. This paper is a continuation of the analysis started in [3] and [9]. The goal of those articles was to provide simple examples of Dirac type operators on noncommutative compact manifolds with boundary and then study Atiyah -- Patodi -- Singer type boundary conditions and the corresponding index problem. This was done for the noncommutative disk and the noncommutative annulus and for two somewhat different types of operators constructed by taking commutators with weighted shifts. 0 1 0 2 l u J 6 1 ] . A O h t a m [ 2 v 8 1 6 5 . 3 0 0 1 : v i X r a In this paper we consider such non-commutative analogs of the Dirac type operator ∂ ∂ ∂ϕ on the cylinder R+ × S1, which we view as a punctured disk. Using a weighted shift, which plays the role of the complex coordinate z on the disk, we construct quantum Dirac operators, and analogs of the L2 Hilbert space of functions in which they act. We then consider the boundary condition of Atiyah, Patodi, Singer. This is done in close analogy with the commutative case. The main result of this note is that a quantum operator has an inverse which, minus the zero mode, is bounded just like in Proposition 2.5 of [1]. In contrast with our previous papers the analysis here is more subtle because of the noncompactness of the cylinder. In particular the components of a parametrix are not compact operators and we use the Schur -- Young inequality ∂t + 1 i ⋆This paper is a contribution to the Special Issue "Noncommutative Spaces and Fields". The full collection is available at http://www.emis.de/journals/SIGMA/noncommutative.html 2 S. Klimek and M. McBride to estimate their norms. It is hoped that in the future such results will be needed to construct spectral triples and a noncommutative index theory of quantum manifolds with boundary. The paper is organized as follows. In Section 2 the classical APS result for the operator −2z ∂ ∂z on the cylinder is stated and re-proved using the Schur -- Young inequality. Section 3 contains the construction of the non-commutative punctured disk and the first type of noncommutative analogs of the operator from the previous section. The operators here are similar to those of [3]. Also in this section a non-commutative Fourier decomposition of the Hilbert spaces and the operators is discussed. Section 4 contains the construction and the analysis of the Fourier components of the parametrix and the proof of the main result. Finally in Section 5 we consider the "balanced" versions of the quantum Dirac operators in the spirit of [9] and show how to modify the previous arguments to estimate the parametrix. 2 Classical Dirac operator on the punctured disk In this section we revisit the analysis of Atiyah, Patodi, Singer in the simple case of semi-infinite cylinder R+ × S1, or equivalently a punctured disk. Using complex coordinates of the latter, we construct a parametrix of a version of the d-bar operator and prove norm estimates on its components by using different techniques than those in [1]. Let D∗ = {z ∈ C : 0 < z ≤ 1} be the punctured disk. Consider the following Dirac type operator on D∗: D = −2z ∂ ∂z . In polar coordinates z = reiϕ the operator D has the following representation: D = −r ∂ ∂r + 1 i ∂ ∂ϕ = −r ∂ ∂r + B, where B = 1 i ∂ ∂ϕ is the boundary operator. We wish to study D, subject to the APS boundary condition, on the Hilbert space L2(D∗, dµ) with measure µ(z) given by the following formula: dµ(z) = 1 2iz2 dz ∧ dz. (2.1) Let us recall the APS condition. Let us define P≥0 to be the spectral projection of B in L2(S1) onto the non-negative part of the spectrum of B. Equivalently, P≥0 is the orthogonal projection onto span{einϕ}n≥0. Then we say that D satisfies the APS boundary condition when its domain consists of those functions f (z) = f (r, ϕ) on D∗ which have only negative frequencies at the boundary, see [1] and [2] for more details. More precisely: dom(D) =(cid:8)f ∈ L2(D∗, dµ) : Df ∈ L2(D∗, dµ), P≥0f (1, ·) = 0(cid:9) . Notice that, by the change of variable, t = − ln r, the Dirac operator, D on L2(D∗, dµ), is (2.2) equivalent to the operator, ∂ ∂t + 1 i ∂ ∂ϕ on L2(R+ × S1), since one has: dϕ ∧ dt = 1 2iz2 dz ∧ dz. This matches the APS setup. A Note on Dirac Operators on the Quantum Punctured Disk 3 We proceed the in the same way as in [1] by considering the spectral decomposition of the boundary operator B, which in our case amounts to Fourier decomposition: fn(r)e−inϕ. f (z) =Xn∈Z This yields the following decomposition of the Hilbert space L2(D∗, dµ): L2(D∗, dµ) =Mn∈Z(cid:18)L2(cid:18)(0, 1], dr r (cid:19) ⊗(cid:2)e−inϕ(cid:3)(cid:19) ∼=Mn∈Z L2(cid:18)(0, 1], dr r (cid:19) . (2.3) (2.4) Now we consider the decomposition of D and its inverse. The theorem below is a special case of Proposition 2.5 of [1] but we supply a proof that generalizes to the noncommutative setup. Define A be f (r) := −rf ′(r) − nf (r) on the maximal domain in L2((0, 1], dr ) : f (1) = 0}. We have: but with domain {f (r) ∈ dom(A r ), and let A0 (n) the operator A (n) (n) (n) Theorem 1. Let D be the Dirac operator defined above on the domain (2.2). With respect to the decomposition (2.4) one has (n) A D ∼=Mn>0 ⊕Mn≤0 (n) . A0 Moreover, there exists an operator Q such that DQ = I = QD, and Q(n) = Q(0) + Q, Q =Mn∈Z where Q is bounded. Proof . Staring with a function g(z) ∈ L2(D∗, dµ) we want to solve the following equation Df (z) = g(z) with f (z) satisfying the APS boundary condition. The Fourier decomposition (2.3) yields Xn∈Z(cid:0)−rf ′ n(r) − nfn(r)(cid:1) e−inϕ =Xn∈Z gn(r)e−inϕ. Therefore we must solve the differential equation −rf ′ n(r) − nfn(r) = gn(r) where additionally fn(1) = 0 for n ≤ 0. This, and the requirement that f is square integrable, assures that there is a unique solution given by the following formula: fn(r) = Q(n)gn(r) =  −Z r 0 (cid:16) ρ r(cid:17)n Z 1 r (cid:16) ρ r(cid:17)n gn(ρ) dρ ρ , gn(ρ) dρ ρ , n > 0, n ≤ 0. This gives us the formula for the parametrix: Q = ⊕n∈ZQ(n). Showing QD = DQ = I is a simple computation and is omitted. We want to prove that Q = ⊕n6=0Q(n) is bounded. One has k Qk ≤ sup n∈Z\{0}(cid:13)(cid:13)Q(n)(cid:13)(cid:13). In what follows we show that the Q(n) are uniformly bounded, in fact of order O(cid:0) 1 tool is the following inequality, see [7]. n(cid:1). The main 4 S. Klimek and M. McBride Lemma 1 (Schur -- Young inequality). Let T : L2(Y ) −→ L2(X) be an integral operator: T f (x) =Z K(x, y)f (y)dy. Then one has kT k2 ≤(cid:18)sup x∈XZY K(x, y)dy(cid:19) sup y∈Y ZX K(x, y)dx! . For negative n one can rewrite Q(n) as Q(n)gn(r) =Z 1 0 K(r, ρ)gn(ρ) dρ ρ with integral kernel K(r, ρ) = χ(r/ρ)(r/ρ)n. Here the characteristic function χ(t) = 1 for t ≤ 1 and is zero otherwise. Next we estimate: sup r Z 1 r rn ρn+1 dρ = sup r Similarly one has: 1 n(cid:0)1 − rn(cid:1) ≤ 1 n . sup ρ Z ρ 0 rn−1 ρn dr = sup ρ 1 n · 1 ρn · ρn = 1 n . Thus one has by the Schur -- Young inequality that kQ(n)k ≤ 1 positive n gives kQ(n)k ≤ 1 n for all n 6= 0. Hence one has that Q is bounded. n . A similar computation for (cid:4) Since in this note we do not attempt to go beyond the analysis on the semi-infinite cylinder, we will simply ignore n = 0. In [1] this was not an issue as Q(0) is continuous when mapping into an appropriate local Sobolev space. 3 Dirac operators on the quantum punctured disk In this section we construct the non-commutative punctured disk and the quantum analog of the Dirac operator of the previous section. In particular, a non-commutative Fourier decomposition of that operator is discussed. Let us also mention that a version of a quantum punctured disk was previously considered in [8]. We start with defining several auxiliary objects needed for our construction. Let ℓ2(Z) be the Hilbert space of square summable bilateral sequences, and let {ek}k∈Z be its canonical basis. We need the following two operators: let U be the shift operator given by: U ek = ek+1 and let K be the label operator defined by the following formula: Kek = kek. By the functional calculus, if f : Z → C, then f (K) is a diagonal operator and satisfies the relation f (K)ek = f (k)ek. A Note on Dirac Operators on the Quantum Punctured Disk 5 Next assume we are given a sequence {w(k)}k∈Z of real numbers with the following properties: 1) w(k) < w(k + 1); w(k) =: w+ exists; 2) 3) lim k→∞ lim k→−∞ w(k) = 0; (3.1) 4) sup k w(k) w(k − 1) < ∞. In particular we have w(k) > 0. The function w : Z → C gives a diagonal operator w(K) as above. From this we define the weighted shift operator Uw := U w(K) which plays the role of a noncommutative complex coordinate on the punctured disk. Clearly: Uwek = w(k)ek+1, U ∗ wek = w(k − 1)ek−1. w, Uw], for which one has Sek = (w2(k) − w2(k − 1))ek. If Consider the commutator S := [U ∗ we let S(k) := w2(k) − w2(k − 1), then we can write S = S(K). Notice that S is a trace class operator and a simple computation gives tr(S) = w2 +. The quantum punctured disk C ∗(Uw) is defined to be the C ∗-algebra generated by Uw. General theory, see [6], gives us the following short exact sequence: 0 −→ K −→ C ∗(Uw) σ −→ C(cid:0)S1(cid:1) −→ 0, where K is the ideal of compact operators and σ is the noncommutative "restriction to the boundary" map. Let b ∈ C ∗(Uw) and we consider the densely defined weight on C ∗(Uw) by τ (b) = tr(cid:0)S(cid:0)U ∗ wUw(cid:1)−1b(cid:1) (compare with (2.1)). We use this weight to define the Hilbert space H on which the Dirac operator will live. This is done by the GNS construction for the algebra C ∗(Uw) with respect to τ . In other words H is obtained as a Hilbert space completion H = (C ∗(Uw), h·, ·iτ = k · k2 w), where kbk2 w = τ (bb∗). Now we are ready to define the operator that we wish to study, the quantum analog of the operator of the previous section. Define D by the following formula: Db = −S−1U ∗ w [b, Uw] . (3.2) Let, as before, P≥0 be the orthogonal L2 projection onto span{einϕ}n≥0. The APS boundary conditions on D amount to the following choice of the domain: dom(D) =(cid:8)b ∈ H : kDbk2 w < ∞, P≥0σ(b) = 0(cid:9) . There are certain subtleties in this definition which are clarified in the statement of Proposition 2 at the end of this section. The next proposition describes a (partial) Fourier series decomposition of the Hilbert space H. Define a(k) := w(k)2 S(k) , (3.3) 6 and let S. Klimek and M. McBride ℓ2 a(Z) =({g(k)}k∈Z : kgk2 a =Xk∈Z a(k)−1g(k)2 < ∞) . Now we are ready for the Fourier decomposition of H which is just like (2.4). Proposition 1. Let H be the Hilbert space defined above. Then the formula gn(K) (U ∗)n b =Xn∈Z defines an isomorphism of Hilbert spaces a(Z) ∼= H. ℓ2 Mn∈Z Proof . The proof is identical to the one in [9]. In particular we have kbk2 w = Xn∈Z tr(cid:0)S(K)w−2(K)gn(K)2(cid:1) . (3.4) (cid:4) The main reason we consider the Fourier decomposition is that it (again partially) diagonalizes the operator D. This is the subject of the next lemma. Before stating it we need some more notation. Consider the ratios: c(n)(k) := w(k + n) w(k) and notice that since {w(k)} is an increasing sequence we have: c(n)(k) = 1 c(n)(k) > 1 c(n)(k) < 1 for n = 0, for n > 0, for n < 0. The coefficients c(n) are needed to define the following operators in ℓ2 a(Z). The first is: (n) A g(k) = a(k)(cid:0)g(k) − c(n)(k)g(k + 1)(cid:1) with domain dom(A) =(cid:8)g ∈ ℓ2 a(Z) : kAgka < ∞(cid:9) . (n) Additionally consider the operator A0 which is the operator A (n) but with domain dom(cid:0)A0 (n)(cid:1) =(cid:8)g ∈ dom(cid:0)A (n)(cid:1) : g∞ := lim k→∞ g(k) = 0(cid:9). The last definition makes sense since by the analysis of [9] the limit limk→∞ g(k) exists for g ∈ dom(A). One has the following proposition, which is a quantum analog of the first part of Theorem 1. A Note on Dirac Operators on the Quantum Punctured Disk 7 Proposition 2. With respect to the decomposition (3.4) one has: (n) A D ∼=Mn>0 ⊕Mn≤0 (n) . A0 Equivalently: (n) A Db =Xn>0 gn(K)(U ∗)n +Xn≤0 (n) A0 gn(K)(U ∗)n, where gn(K) (U ∗)n b =Xn∈Z Proof . The proof is a direct calculation identical to the one in [9]. (cid:4) 4 Construction of the parametrix In this section we construct and analyze in detail the inverse (= a parametrix) Q for the opera- tor D. The construction is fairly similar to the one done in Section 4 in [9], however the norm estimates are quite different. Somewhat surprisingly the norm estimates below hold for any choice of sequence of weights {w(k)} satisfying (3.1). We start with a lemma containing estimates of sums through integrals. Recall that the sequence {w(k)} is increasing with limits at ±∞ equal, correspondingly, to w+ and 0. Lemma 2. If f (t) is a decreasing continuous function on (0, (w+)2) then Xl<k Xk≤l Xk∈Z f(cid:0)w(k)2(cid:1)S(k) =Xl<k f(cid:0)w(k)2(cid:1)S(k) ≤Z w(l)2 f(cid:0)w(k − 1)2(cid:1)S(k) ≥Z w2 0 0 + f (t)dt, f (t)dt. f(cid:0)w(k)2(cid:1)(cid:0)w(k)2 − w(k − 1)2(cid:1) ≤Z w2 w(l)2 + f (t)dt, (4.1) (4.2) (4.3) The proof of the statements of the lemma follows from a straightforward comparison of the Riemann sums of the left hand side with the integrals on the right hand side. Our presentation in this section is as follows. First we discuss the kernels of the A operators for the three cases n = 0, n > 0, n < 0. Secondly we construct the parametrices for all three cases. Thirdly we discuss the norm estimates of the parametrices, and finally we summarize the analysis in the main result of this paper. (n) Below we show that the operator D has no kernel by analyzing the terms in the decomposition of Proposition 2. Proposition 3. The operators A (n) for n ≥ 0 and A0 (n) for n < 0 have no kernel. Proof . We start with n = 0. Here c(n)(k) = 1 and it is clear that the only solution of that A R(0) = 0 is, up to a constant, R(0) = 1. But one has (0) (cid:13)(cid:13)R(0)(cid:13)(cid:13) 2 a =Xk∈Z 1 a(k) =Xk∈Z S(k) w(k)2 ≥ constXk∈Z S(k) w(k − 1)2 = ∞, 8 S. Klimek and M. McBride where we used condition 4 of (3.1) as well as (4.3) for f (t) = 1/t. Therefore R(0)(K) 6∈ ℓ2 and hence A has no kernel. (0) a(Z) Next we discuss the kernel of A when n > 0. It is not too hard to see that any element of (n) the kernel has to be proportional to R(n)(k) := c(n)(l) = ∞ Yl=k The norm calculation gives (w+)n w(k)w(k + 1) · · · w(k + n − 1) . kR(n)k2 a =Xk∈Z 1 a(k) R(n)(k)2 =Xk∈Z 1 a(k) ∞ Yl=k c(n)(l)2. Since c(n)(l) > 1 and Pk∈Z Finally we discuss the kernel of the operator A0 1 a(k) = ∞, the sum above diverges and hence R(n)(K) 6∈ ℓ2 a(Z). (n) when n < 0. Yet again the kernel is formally one dimensional and spanned by R(n)(k) = c(n)(l) = ∞ Yl=k w(k + n)w(k + n − 1) · · · w(k − 1) (w+)−n . While one can easily show that R(n) ∈ ℓ2 a(Z), one however has lim k→∞ R(n)(k) = 1 6= 0, so this means R(n) 6∈ dom(cid:0)A0 (n)(cid:1). Thus the result follows. The second portion of the discussion is the construction of the parametrices for all three cases. Since there are no kernels (and cokernels) involved we simply compute the inverses of opera- tors A f (k) = g(k) where additionally f (k) = 0 for n ≤ 0. This is done in a similar manner to the methods in [3, 9]. In we need lim k→∞ . Thus, given g(k), one needs to solve the equation A (n) (n) the case when n > 0 one arrives at the following formula: (cid:4) f (k) = −Xl<k R(n)(k) R(n)(l)a(l) g(l) = −Xl<k w(l) · · · w(l + n − 1) w(k) · · · w(k + n − 1) S(l) w(l)2 g(l). Similarly in the case n ≤ 0 one has: f (k) =Xk≤l R(n)(k) R(n)(l) g(l) a(l) =Xk≤l w(k + n) · · · w(k − 1) w(l + n) · · · w(l − 1) S(l) w(l)2 g(l). The right hand sides of the above equation give the parametrices Q(n) for all three cases. We thus have the following: Q(n)g(k) = −Xl<k Q(n)g(k) = −Xl<k Q(n)g(k) =Xk≤l S(l) w(l)2 g(l) w(l) · · · w(l + n − 1) w(k) · · · w(k + n − 1) w(k + n) · · · w(k − 1) w(l + n) · · · w(l − 1) S(l) w(l)2 g(l) S(l) w(l)2 g(l) for n = 0, for n > 0, for n < 0. (4.4) We summarize the above analysis in the following proposition. A Note on Dirac Operators on the Quantum Punctured Disk 9 Proposition 4. Let Q(n) be defined by the formulas above, then we have the following (n) A Q(n) = I and Q(n)A (n) = I (n) A0 Q(n) = I and Q(n)A0 (n) = I for n > 0; for n ≤ 0. Next we discuss the boundedness for the parametrices in the cases n > 0 and n < 0. The difficulty comes for k → ∞: while the ratios of weights are always less than 1, the series S(k) w(k)2 is not summable and we cannot replicate the estimates of [3] and [9]. In fact the integral operators Q(n) are not Hilbert -- Schmidt. The trick is to estimate most but not all weight ratios by one. The remaining sums, containing potentially divergent terms, are estimated by integrals using Lemma 2. We have the following result. Pk∈Z Proposition 5. The operators Q(n) defined above are bounded operators in ℓ2 a(Z) when n 6= 0. Proof . First consider the case that n > 0. Applying the Schur -- Young inequality and the inequalities (4.1), and (4.2) one has w(l) · · · w(l + n − 1) w(k) · · · w(k + n − 1) S(k) w(k)2! 2 a ≤ sup (cid:13)(cid:13)Q(n)(cid:13)(cid:13) S(l) S(l) S(k) k Xl<k k 1 k   k w(l) · · · w(l + n − 1) w(k) · · · w(k + n − 1) l Xl<k w(l)2! sup l w(l)Xl<k w(l)! sup w(k)3! w(k)Xl<k w(l) l w(l)Z w2 w(k − 1) Xl≤k−1  sup 2 dt! · 2 sup w(k − 1)Z w(k−1)2 l (cid:18)1 − t− 3 t− 1 S(l) w(l)2 1 1 0 + ≤ sup ≤ sup ≤ sup 2 dt! w+ (cid:19) ≤ 2 · 2 = 4. w(l) Thus Q(n) is bounded for n > 0. Next consider the case n < 0. Here one has quite similar estimates: 2 a ≤ sup (cid:13)(cid:13)Q(n)(cid:13)(cid:13) w(k + n) · · · w(k − 1) w(l + n) · · · w(l − 1) S(k) w(k)2  1 S(l) S(l) w(k + n) · · · w(k − 1) w(l + n) · · · w(l − 1) w(l)2 l  Xk≤l  sup w(k) w(l)2w(l − 1) l  w(l − 1)Xk≤l  sup   w(k) l  w(l)3! sup k w(k − 1) Xk−1<l w(l − 1)Xk≤l   k w(k − 1)Z w2 l 2 dt! sup w(l − 1)Z w(l)2 w(k−1)2 S(k) S(k) t− 3 S(l) 1 1 0 + t− 1 2 dt! ≤ sup k  Xk≤l k  w(k − 1)Xk≤l w(l − 1)(cid:19) sup ≤(cid:18)sup w(l − 1)(cid:19) sup ≤(cid:18)sup w(l − 1)(cid:19)2 ≤ 4(cid:18)sup w(l) w(l) w(l) l l l < ∞. Thus Q(n) is bounded for n < 0 and this completes the proof. (cid:4) Finally we put together the previous information about the parametrix Q of the Dirac ope- rator D defined in Section 3. We state the main result of this note. 10 S. Klimek and M. McBride Theorem 2. Let D be the operator (3.2) with domain (3.3). Then there exists an operator Q such that QD = DQ = I. Moreover, with respect to the decomposition (3.4) one has Q(n) = Q(0) + Q, Q =Mn∈Z where the operators Q(n) are given by (4.4) and Q is bounded. (4.5) Proof . By Proposition 2 one can decompose D asLn>0 A the decomposition (4.5) of Q. One has that (n) ⊕Ln≤0 A0 (n) which in turn gives k Qkw = sup n6=0(cid:13)(cid:13)Q(n)(cid:13)(cid:13)a. Then from Proposition 5, one has the following inequalities kQk2 w ≤ 4(cid:18)sup l w(l) w(l − 1)(cid:19)2 < ∞, where the last inequality follows from the assumptions in (3.1). To see that one has DQ=QD=I we use the decompositions of Q and D and Proposition 4. This completes the proof. (cid:4) 5 The balanced quantum Dirac operators In this section we study a version of the constructions of the previous sections that is more like the theory of [9]. The main objects: the Hilbert space and the Dirac operator are called balanced since in their definitions the left multiplication is not preferred over the right multiplication. Since the results for the balanced Dirac operators are completely analogous to the "unba- lanced" case and the proofs require only trivial modification, we simply state the main steps of the construction. The only significant difference between the two cases are the estimates on the components of the parametrix. In order to avoid unnecessary complications we simply recycle the old notation. As before the starting point is the choice of a sequence of weights {w(k)}k∈Z satisfying (3.1). The Hilbert space H is the space of power series: gn(K) (U ∗)n b =Xn∈Z but this time with a different, balanced norm: kbk2 w = tr(cid:16)S1/2w(K)−1bb∗w(K)−1S1/2(cid:17) = Xn∈Z tr(cid:16)pS(K)S(K + n)w−1(K)w−1(K + n)gn(K)2(cid:17) . The balanced Dirac operator is Db = −S−1/2U ∗ [b, Uw] w(K)S−1/2 with the domain: dom(D) =(cid:8)b ∈ H : kDbk2 w < ∞, P≥0σcirc(b) = 0(cid:9) . A Note on Dirac Operators on the Quantum Punctured Disk 11 As before the Dirac operator splits into Fourier components. To describe them we need to modify the coefficients of the previous sections. Actually, the coefficients c(n)(k) := w(k + n) w(k) stay the same, but we need to change: a(n)(k) := w(k)w(k + n) pS(k)S(k + n) . Those are used for the following previously defined operators in ℓ2 a(Z). The first operator is: (n) A g(k) = a(n)(g(k) − c(n)(k)g(k + 1)) with domain dom(A) =(cid:8)g ∈ ℓ2 (n) a(Z) : kAgkℓ2 a(Z) < ∞(cid:9) , (n) and the second operator A0 is the operator A but with domain dom(cid:0)A0 (n)(cid:1) =(cid:8)g ∈ dom(cid:0)A (n)(cid:1) : g∞ := lim k→∞ g(k) = 0(cid:9). With that notation, the Proposition 2 remains true. In particular one has: (n) A D ∼=Mn>0 ⊕Mn≤0 (n) . A0 The problem of inverting the operator D is tackled as in the previous section. The components of the inverse are given by formulas like (4.4) with the only modification coming from the diffe- rent a(n) coefficients. We end up with the following expressions for the parametrices: g(l) for n = 0, w(l)w(l + n) Q(n)g(k) = −Xl<k pS(l)S(l + n) Q(n)g(k) = −Xl<k Q(n)g(k) =Xk≤l w(l) · · · w(l + n − 1) w(k) · · · w(k + n − 1)pS(l)S(l + n) w(l + n) · · · w(l − 1) pS(l)S(l + n) w(k + n) · · · w(k − 1) w(l)w(l + n) w(l)w(l + n) g(l) for n > 0, g(l) for n < 0. One can verify directly that for the operator Q = Ln∈Z Q(n) we have QD = DQ = I. The following is the main result of this section. Proposition 6. The operators Q(n) defined above are bounded operators in ℓ2 a(Z) when n 6= 0. Proof . We use the Schur -- Young inequality and follows the steps of the proof of the Proposi- tion 5, with some modifications. We show the details for n < 0, the other case is completely analogous. There are two sums that we need to estimate. The first sum is: Σn 1 (k) :=Xk≤l w(k + n) · · · w(k − 1) w(l + n) · · · w(l − 1) pS(l)S(l + n) w(l)w(l + n) . 12 S. Klimek and M. McBride Using Cachy -- Schwartz inequality we estimate: w(k + n) · · · w(k − 1) w(l + n) · · · w(l − 1) w(k + n) · · · w(k − 1) w(l + n) · · · w(l − 1) Σn 1 (k) ≤ Xk≤l ≤ w(k − 1)Xk≤l S(l) 1/2 w(l)2 Xk≤l  1/2 w(k + n)Xk≤l S(l) w(l)2w(l − 1)  1/2 . S(l + n) w(l + n)3  1/2 S(l + n) w(l + n)2  Since the weights in the denominator are bigger than the corresponding weights in the numerator, their ratios were estimated by one. The first term on the rights hand side of the above was already estimated in the proof of Proposition 5. The second term is essentially the same as the first: sup k  w(k + n)Xk≤l S(l + n) w(l + n)3 k  w(k)Xk≤l  = sup S(l) w(l)3  . It follows that Σn 1 (k) is bounded uniformly in n. The second sum in the Schur -- Young inequality is Σn 2 (l) :=Xk≤l w(k + n) · · · w(k − 1) w(l + n) · · · w(l − 1) pS(k)S(k + n) w(k)w(k + n) and we bound it in the same fashion as the first sum: w(k + n) · · · w(k − 1) w(l + n) · · · w(l − 1) w(k + n) · · · w(k − 1) w(l + n) · · · w(l − 1) Σn 2 (l) ≤ Xk≤l ≤ w(l − 1)Xk≤l  1 S(k) w(k)2  1/2 Xk≤l w(k + n) w(l + n)Xk≤l  S(k + n) 1 1/2 . S(k) w(k)  1/2  1/2 S(k + n) w(k + n)2  Again the first term above was already estimated in the proof of Proposition 5, and the second term is essentially the same as the first. It follows that Σn 1 (k) is uniformly bounded. Repeating the same steps for n > 0 gives the boundedness of Q for the balanced Dirac operator. (cid:4) References [1] Atiyah M.F., Patodi V.K., Singer I.M., Spectral asymmetry and Riemannian geometry. I, Math. Proc. Cambridge Philos. Soc. 77 (1975), 43 -- 69. [2] Booss-Bavnbek B., Wojciechowski K.P., Elliptic boundary problems for Dirac operators, Mathematics: Theo- ry & Applications, Birkhauser Boston, Inc., Boston, MA, 1993. [3] Carey A.L., Klimek S., Wojciechowski K.P., Dirac operators on noncommutative manifolds with boundary, arXiv:0901.0123. [4] Carey A.L., Phillips J., Rennie A., A noncommutative Atiyah -- Patodi -- Singer index theorem in KK-theory, arXiv:0711.3028. [5] Connes A., Non-commutative differential geometry, Academic Press, 1994. [6] Conway J., Subnormal operators, Research Notes in Mathematics, Vol. 51, Pitman, Boston, Mass. -- London, 1981. [7] Halmos P.R., Sunder V.S., Bounded integral operators on L2 spaces, Results in Mathematics and Related Areas, Vol. 96, Springer-Verlag, Berlin -- New York, 1978. [8] Klimek S., Lesniewski A., Quantum Riemann surfaces. III. The exceptional cases, Lett. Math. Phys. 32 (1994), 45 -- 61. [9] Klimek S., McBride M., D-bar operators on quantum domains, arXiv:1001.2216.
0909.4099
2
0909
2011-02-23T02:46:09
Constructing the extended Haagerup planar algebra
[ "math.OA", "math.QA" ]
We construct a new subfactor planar algebra, and as a corollary a new subfactor, with the `extended Haagerup' principal graph pair. This completes the classification of irreducible amenable subfactors with index in the range $(4,3+\sqrt{3})$, which was initiated by Haagerup in 1993. We prove that the subfactor planar algebra with these principal graphs is unique. We give a skein theoretic description, and a description as a subalgebra generated by a certain element in the graph planar algebra of its principal graph. In the skein theoretic description there is an explicit algorithm for evaluating closed diagrams. This evaluation algorithm is unusual because intermediate steps may increase the number of generators in a diagram.
math.OA
math
Constructing the extended Haagerup planar algebra STEPHEN BIGELOW SCOTT MORRISON EMILY PETERS NOAH SNYDER URLs: http://www.math.ucsb.edu/~bigelow/ http://tqft.net/ http://euclid.unh.edu/~eep and http://math.berkeley.edu/~nsnyder Email: [email protected], [email protected], [email protected] and [email protected] Abstract We construct a new subfactor planar algebra, and as a corollary a new subfactor, with the 'extended Haagerup' principal graph pair. This com- pletes the classification of irreducible amenable subfactors with index in the range (4, 3 + √3), which was initiated by Haagerup in 1993. We prove that the subfactor planar algebra with these principal graphs is unique. We give a skein theoretic description, and a description as a subalgebra generated by a certain element in the graph planar algebra of its principal graph. In the skein theoretic description there is an explicit algorithm for evaluating closed diagrams. This evaluation algorithm is unusual because intermediate steps may increase the number of generators in a diagram. AMS Classification 46L37 ; 18D10 Keywords Planar Algebras, Subfactors, Skein Theory, Principal Graphs 1 Introduction A subfactor is an inclusion N ⊂ M of von Neumann algebras with trivial center. The theory of subfactors can be thought of as a nonabelian version of Galois theory, and has had many applications in operator algebras, quantum algebra, and knot theory. For example, the construction of a new finite depth subfactor, as in this paper, also yields two new fusion categories (by taking the even parts) and a new 3-dimensional TQFT (via the Ocneanu-Turaev-Viro construction [62, 48, 51]). A subfactor N ⊂ M has three key invariants. From strongest to weakest, they are: the standard invariant (which captures all information about "basic" bimodules over M and N ), the principal and dual principal graphs (which together describe the fusion rules for these basic bimodules), and the index (which is a real number measuring the "size" of the basic bimodules). We will use the axiomatization of the standard invariant as a subfactor planar algebra, which is due to Jones [26]. Other axiomatizations include Ocneanu's paragroups [47] and Popa's λ-lattices [56]. (For readers more familiar with tensor categories, these three approaches are analogous to the diagram calculus [51, 58, 35], basic 6j symbols [61, Chapter 5], and towers of endomorphism algebras [65], respectively.) The standard invariant is a complete invariant of amenable subfactors of the hyperfinite II1 factor [53, 55]. The index of a subfactor N ⊂ M must lie in the set n 4 cos2(cid:16) π n(cid:17)(cid:12)(cid:12)(cid:12) n ≥ 3o ∪ [4,∞], 1 and all numbers in this set can be realized as the index of a subfactor [27]. Early work on classifying subfactors of "small index" concentrated on the case of index less than 4. The principal graphs of these subfactors are exactly the Dynkin di- agrams An , D2n , E6 and E8 . Furthermore there is exactly one subfactor planar algebra with principal graph An or D2n and there is exactly one pair of complex conjugate subfactor planar algebras with principal graph E6 or E8 . (See [47] for the outline of this result, and [5, 22, 23, 38] for more details.) The story of the corre- sponding classification for index equal to 4 is outlined in [55, p. 231]. In this case, the principal graph must be an affine Dynkin diagram. For some principal graphs there are multiple non-conjugate subfactors with the same principal graph, which are distinguished by homological data. The classification of subfactors of "small index" greater than 4 was initiated by Haagerup [18]. His main result is a list of all possible pairs of principal graphs of ir- reducible subfactors of index larger than 4 but smaller than 3 +√3. Here we begin to see subfactors whose principal graph is different from its dual principal graph. If Γ refers to a pair of principal graphs and we need to refer to one individually, we will use the notation Γp and Γd . Any subfactor N ⊂ M has a dual given by the ba- sic construction M ⊂ M1 . Taking duals reverses the shading on the planar algebra, switches the principal and dual principal graphs, and preserves index. Haagerup's list is as follows (we list each pair once). • (A∞, A∞), • the infinite family Hn =  Bn =  , , , n∈N  ,  n∈N which has index(H0) = 5+√13 , index(H1) the largest root of x3−8x2+17x−5, and index(Hn) monotonically increasing with n, converging to the real root of x3 − 6x2 + 8x − 4, (thus index(H0) ≈ 4.30278, index(H1) ≈ 4.37720, and limn index(Hn) ≈ 4.38298), 2 • the infinite family which has index(B0) = 7+√5 , and index(Bn) monotonically increasing in n, converging to the real root of x3 − 8x2 + 19x − 16, (thus index(B0) ≈ 4.61803, and limn index(Bn) ≈ 4.65897), 2 • one more pair of graphs, AH = which has index 5+√17 2 ≈ 4.56155.  , , 2 Haagerup's paper announces this result up to index 3 + √3 ≈ 4.73205, but only proves it up to index 3 + √2 ≈ 4.41421; this includes all of the graphs Hn , but none of the graphs Bn or AH. Haagerup's proof of the full result has not yet appeared. In work in progress, Jones, Morrison, Penneys, Peters, and Snyder have independently confirmed his result (following Haagerup's outline except at one point using a result from [31]), and have extended his techniques to give a partial result up to index 5 (see [45, 42, 25, 50]). In this paper, we will only rely on the part of Haagerup's classification that has appeared in print. Haagerup's original result did not specify which of the possible principal graphs are actually realized. Considerable progress has since been made in this direction. Asaeda and Haagerup [2] proved the existence and uniqueness of a subfactor pla- nar algebra whose principal graphs are H0 (called the Haagerup subfactor), and a subfactor planar algebra for AH (called the Asaeda-Haagerup subfactor). Izumi [24] gave an alternate construction of the Haagerup subfactor. Bisch [8] showed none of the graphs Bn can be principal graphs because they give inconsistent fu- sion rules. Asaeda [1] and Asaeda-Yasuda [3] proved that Hn is not a principal graph for n ≥ 2. To do this, they showed that the index is not a cyclotomic integer, and then appealed to a result of Etingof, Nikshych and Ostrik [15], which in turn is proved by reduction to the case of modular categories, where it was proved in the context of rational conformal field theories by Coste -- Gannon [12] using a result of de Boere -- Goeree [13]. The main result of our paper is Theorem 3.10 Thereisasubfactorplanaralgebrawithprincipalgraphs H1. In addition, we prove in Theorem 3.9 that this planar algebra is the only one with these principal graphs. This result completes the classification of all subfactor pla- nar algebras up to index 3 + √3: Corollary ([18], [2], [8], [3], and Theorem 3.10) The only irreducible subfactor planaralgebraswithindexintherange (4, 3 + √3) are • thenon-amenableTemperley-Liebplanaralgebraateveryindexinthisrange, withprincipalgraphs (A∞, A∞), • theHaagerupplanaralgebrawithprincipalgraphs H0,anditsdual, • theHaagerup-AsaedaplanaralgebrawithprincipalgraphsAH,anditsdual, and • the extended Haagerup planar algebra with principal graphs H1, and its dual. By Popa's classification [53] the latter three pairs can each be realized uniquely as the standard invariant of a subfactor of the hyperfinite II1 factor. This gives a complete classification of amenable subfactors of the hyperfinite II1 factor with index between (4, 3 + √3). The non-amenable case remains open because it is un- known for which indices Temperley-Lieb can be realized as the standard invariant of the hyperfinite II1 factor, nor in how many ways it can be realized (see [54, 6] 3 for some work in this direction). Furthermore, there remain many interesting ques- tions about small index subfactors of arbitrary factors. It was already expected that the extended Haagerup subfactor should exist, thanks to approximate numerical evidence coming from computations by Ikeda [21]. We note that although our construction relies on a computation of the traces of a few large matrices, this computation consists of exact arithmetic in a number field, and is a very different calculation from the one Ikeda did numerically. The search for small index subfactors has so far produced the three pairs of "spo- radic" examples: the Haagerup, Asaeda-Haagerup and extended Haagerup subfac- tors. These are some of the very few known subfactors that do not seem to fit into the frameworks of groups, quantum groups, or conformal field theory [20]. (See also a generalization of the Haagerup subfactor due to Izumi [24, Example 7.2]). You might think of them as analogs of the exceptional simple Lie algebras, or of the sporadic finite simple groups. (Without a good extension theory, it is not yet clear what "simple" should mean in this context.) In this paper, we study the extended Haagerup planar algebra. We construct the ex- tended Haagerup planar algebra by locating it inside the graph planar algebra [29] of its principal graph. By a result of Jones -- Penneys [46] (generalized in [33]) every subfactor planar algebra occurs in this way. We find the right planar subalgebra by following a recipe outlined by Jones [29, 31] and further developed by Peters [52], who applied it to the Haagerup planar algebra. We also give a presentation of the extended Haagerup planar algebra using a single planar generator and explicit relations. We prove that the subalgebra of the graph planar algebra contains an element also satisfying these relations. This is conve- nient because different properties become more apparent in different descriptions of the planar algebra. For example, the subalgebra of the graph planar algebra is clearly non-trivial, which would be difficult to prove directly from the generators and relations. In the other direction, in §5 we prove that our relations result in a space of closed diagrams that is at most one dimensional, which would be difficult to prove in the graph planar algebra setting. In §2 we recall the definitions of planar algebras and graph planar algebras [26, 29]. We also set some notation for the graph planar algebra of Hp 1 . In §3 we prove our two main theorems, Theorems 3.9 and 3.10. Theorem 3.9, the uniqueness theorem, says that for each k there is at most one subfactor planar algebra with principal graphs Hk . Furthermore we give a skein theoretic description by generators and relations of the unique candidate planar algebra. Theorem 3.10, the existence the- orem, constructs a subfactor planar algebra with principal graphs H1 by realiz- ing the skein theoretic planar algebra as a subalgebra of the graph planar algebra. Proofs of several key results needed for the main existence and uniqueness argu- ments are deferred to §4, §5, and §6. In particular, §4 describes an evaluation algo- rithm that uses the skein theory to evaluate any closed diagram (Theorem 3.8). This is crucial to our proofs of both existence and uniqueness and may be of broader in- terest in quantum topology. This section can be read independently of the rest of the paper. Section 5 consists of calculations of inner products using generators and relations. Section 6 gives the description of the generator of our subfactor planar algebra inside the graph planar algebra and verifies its properties. Appendix A 4 gives the tensor product rules for the two fusion categories associated to the ex- tended Haagerup subfactor. Part of this work was done while Stephen Bigelow and Emily Peters were visiting the University of Melbourne. Scott Morrison was at Microsoft Station Q and the Miller Institute for Basic Research during this work. Emily Peters was supported in part by NSF Grant DMS0401734 and a fellowship from Soroptimist International and Noah Snyder was supported in part by RTG grant DMS-0354321 and in part by an NSF Postdoctoral Fellowship. We would like to thank Vaughan Jones for many useful discussions, and Yossi Farjoun for lessons on Newton's method. 2 Background 2.1 Planar algebras Planar algebras were defined in [26] and [29]. More general definitions have since appeared elsewhere, but we only need the original notion of a shaded planar algebra, which we sketch here. For further details see [29, §2], [26, §0], or [9]. Definition 2.1 A(shaded)planartanglehasanouterdisk,afinitenumberofinner disks, and a finite number of non-intersecting strings. A string can be either a closedlooporanedgewithendpointsonboundarycircles. Werequirethattherebe anevennumberofendpointsoneachboundarycircle,andacheckerboardshading of the regions in the complement of the interior disks. We further require that therebeamarkedpointontheboundaryofeachdisk,andthattheinnerdisksare ordered. Two planar tangles are considered equal if they are isotopic (not necessarily rel boundary). Hereisanexampleofaplanartangle. ⋆ 1 2 ⋆ ⋆ 3 ⋆ Planar tangles can be composed by placing one planar tangle inside an interior diskofanother,liningupthemarkedpoints,andconnectingendpointsofstrands. The numbers of endpoints and the shadings must match up appropriately. This compositionturnsthecollectionofplanartanglesintoacoloredoperad. Definition 2.2 A(shaded)planaralgebraconsistsof 5 • Afamilyofvectorspaces{V(n,±,)}n∈N,calledthepositiveandnegative n-box spaces. • For each planar tangle, a multilinear map Vn1,±1 ⊗ . . . ⊗ Vnk,±k → Vn0,±0 where ni ishalfthenumberofendpointsonthe ithinteriorboundarycircle, n0 is half the number of endpoints on the outer boundary circle, and the signs ± are positive (respectively negative) when the marked point on the correspondingboundarycircleisinanunshadedregion(respectivelyshaded region). Forexample,theplanartangleabovegivesamap V1,+ ⊗ V2,+ ⊗ V2,− → V3,+. The linear map associated to a 'radial' tangle (with one inner disc, radial strings, and matching marked points) must be the identity. We require that the action of planar tangles be compatible with composition of planar tangles. In other words, compositionofplanartanglesmustcorrespondtotheobviouscompositionofmul- tilinear maps. In operadic language this says that a planar algebra is an algebra overtheoperadofplanartangles. We will refer to an element of Vn,± (and specifically Vn,+ , unless otherwise stated) as an "n-box." We make frequent use of three families of planar tangles called multiplication, trace, and tensor product, which are shown in Figure 1. "Multiplication" gives an asso- ciative product Vn,± ⊗ Vn,± → Vn,± . "Trace" gives a map Vn,± → V0,± . "Tensor product" gives an associative product Vm,± ⊗ Vn,± → Vm+n,± if m is even, or Vm,± ⊗ Vn,∓ → Vm+n,± if m is odd. . . . ⋆ ⋆ 1 . . . , ⋆ ··· , ⋆⋆ 1 . . . . . . . . . ⋆ 2 . . . ⋆ 2 . . . Figure 1: The multiplication, trace, and tensor product tangles. The (shaded or unshaded) empty diagrams can be thought of as elements of V0,± , since the 'empty tangle' induces a map from the empty tensor product C to the space V0,± . If the space V0,± is one dimensional then we can identify it with C by sending the empty diagram to one. In many other cases, we can make do with the following. 6 Definition 2.3 Apartitionfunctionisapairoflinearmaps Z± : V0,± → C thatsendtheemptydiagramsto 1. Inaplanaralgebrawithapartitionfunction,let tr : Vn,± → C denotethecompositionofthetracetanglewith Z. Sometimes we will need to refer simply to the action of the trace tangle, which we denote tr0 : Vn,± → V0,± . Notice that the above trace tangle is the "right trace." There is also a "left trace" where all the strands are connected around the left side. Definition 2.4 Aplanaralgebrawithapartitionfunctioncanbe: • Positivedefinite:Thereisanantilinearadjointoperation∗ oneach Vn,±,com- patible with theadjointoperationonplanar tanglesgiven byreflection. The sesquilinearform hx, yi = tr(xy∗) ispositivedefinite. • Spherical: Thelefttrace andtherighttrace areequal. trl : V1,± → V0,∓ Z∓−−→ C trr : V1,± → V0,± Z±−−→ C The spherical property implies that the left and right traces are equal on every Vn,± . Since every planar algebra we consider is spherical, we will usually ignore the distinction between left and right trace. Definition 2.5 A subfactor planar algebra is a positive definite spherical planar algebrasuchthat dim V0,+ = dim V0,− = 1 and dim Vn,± < ∞. As a consequence of being spherical and having 1-dimensional 0-box spaces, sub- factor planar algebras always have a well-defined modulus, as described below. Definition 2.6 Wesaythattheplanaralgebrahasmodulus d ifthefollowingrela- tionshold. = d · , = d · . The principal graphs of a subfactor encode the fusion rules for the basic bimodules N MM and M MN . The vertices of the principal graph are the isomorphism classes of simple N -N and N -M bimodules that occur in tensor products of the basic bi- modules. The edges give the decompositions of tensor products of simple bimod- ules with the basic bimodule. The dual principal graph encodes similar informa- tion, but for M -N and M -M bimodules. This definition is due to Connes [11] and 7 Ocneanu [47], together with later work of Jones [32] that lets you replace the orig- inal Hilbert space bimodules with algebraic bimodules, for example, M L2(M )N with M MN . In the language of planar algebras, the basic bimodule is a single strand, and the isomorphism classes of simple bimodules are equivalence classes of irreducible pro- jections in the box spaces. For a more detailed description, see [43, §4.1]. Definition 2.7 Asubfactorplanaralgebra P isirreducibleif dimP1,+ = 1. A subfactor planar algebra has finite depth if it has finitely many isomorphism classes of irreducible projections, that is, finitely many vertices in the principal graphs. [47,53] The following is well-known, and combines the results of [53] and [26]. Theorem 2.8 Finitedepthfiniteindexsubfactorsofthehyperfinite II1 factorarein one-to-onecorrespondencewithfinitedepthsubfactorplanaralgebras. Irreducible subfactors(thoseforwhich M isirreducible as an N -M bimodule) correspondto irreduciblesubfactorplanaralgebras. Proof Suppose we are given a subfactor N ⊂ M . The corresponding planar al- gebra is constructed as follows. Let C be the 2-category of all bimodules that ap- pear in the decomposition of some tensor product of alternating copies of M as a N -M bimodule and M as a M -M bimodule. These are A-B bimodules for A, B ∈ {M, N}, and form the 1-morphisms of C . Composition of 1-morphisms is given by tensor product. The 2-morphisms of C are the intertwiners. We can define duals in C by taking the contragradient bimodule, which interchanges M as an N -M bimodule with M as an M -N bimodule. Now define the associated planar algebra by Pn,± = EndC(cid:18)dOn M±(cid:19) . Here M± means M or M∗ , and cNnM means M ⊗ M∗ ⊗ M ⊗ ··· ⊗ M± . The action of tangles is via the usual interpretation of string diagrams as 2-morphisms in a 2-category [35], with critical points interpreted as evaluation and coevaluation maps. The difficult direction is to recover a subfactor from a planar algebra. The proof of this result was given in [53]. However in that paper, Popa uses towers of com- mutants instead of tensor products of bimodules, and λ-lattices instead of planar algebras. See [26] to translate from λ-lattices into planar algebras. See [7] and [32] to translate from towers of commutants into tensor products of bimodules. Remark. The above theorem says that a certain kind of subfactor is completely characterized by its representation theory (that is, the bimodules). This can be thought of as a subfactor version of the Dopplicher-Roberts theorem [14], or more generally, of Tannaka-Krein type theorems [36]. 8 In general, given any extremal finite index subfactor of a II1 factor, the standard invariant is a subfactor planar algebra. Several other reconstruction results have been proved. Popa extended his results on finite depth subfactors to amenable sub- factors of the hyperfinite II1 factor in [55]. The general situation for non-amenable subfactors of the hyperfinite II1 factor is more complicated: some subfactor planar algebras cannot be realized at all (an unpublished result of Popa's, see [54]), while others can be realized by a continuous family of different subfactors [10]. Further- more, once you move beyond the hyperfinite II1 factor there are many new ques- tions. On the one hand any subfactor planar algebra comes from a (canonically constructed, but not necessarily unique) subfactor of the free group factor L(F∞) [57, 34, 39, 17], while on the other hand there exist factors for which only the trivial planar algebra can be realized as the standard invariant of a subfactor [63]. 2.2 Temperley-Lieb Everyone's favorite example of a planar algebra is Temperley-Lieb. It was defined (as an algebra) in [60], applied to subfactor theory in [27], and formulated diagram- matically in [37]. The vector space T Ln,± is spanned by non-crossing pairings of 2n points (where the ± depends on whether the marked point is in a shaded region or unshaded region). These pairings are drawn as intervals in a disc, starting from a marked point on the boundary. For example, T L3,+ = span ⋆ ⋆ , ⋆ , ⋆ , ⋆ , .  Planar tangles act on Temperley-Lieb elements "diagrammatically:" the inputs are inserted into the inner disk, strings are smoothed out, and each loop is discarded in exchange for a factor of d ∈ C. For example, ⋆ ⋆ ⋆   = ⋆ ⋆ = d2 . If d ∈ R we can introduce an antilinear involution ∗ by reflecting diagrams. If d ≥ 2 then Temperley-Lieb is a subfactor planar algebra. If d = 2 cos π n for n = 3, 4, 5, . . . then we can take a certain quotient to obtain a subfactor planar algebra. The irre- ducible projections in the Temperley-Lieb planar algebra are the Jones-Wenzl idem- potents [28, 64]. Definition 2.9 TheJones-Wenzlidempotent f (n) ∈ T Ln,± ischaracterizedby f (n) 6= 0 f (n)f (n) = f (n) eif (n) = f (n)ei = 0 for i = 1, 2, . . . , n − 1 9 where e1, . . . , en−1 aretheJonesprojections . . . , . . . , e1 = 1 d e2 = 1 d . . . , en−1 = 1 d . . . . The following gives a recursive definition of the Jones-Wenzl idempotents. We should mention that the following pictures are drawn using rectangles instead of disks, and the marked points are assumed to be on the left side of the rectangles (including the implicit bounding rectangles). The quantum integers [n] = qn−q−n q−q−1 which appear below are specialized at a value of q such that [2] = d. Lemma 2.10 ([16]) a f (k) = f (k−1) + 1 [k] k−1Xa=1 (−1)a+k[a] f (k−1) . Proof It is straightforward to check that the right hand side of this equation satis- fies the characterizing relations of Definition 2.9. We will also use the following more symmetrical version of the lemma. Lemma 2.11 a f (k) = f (k−1) + 1 [k][k − 1] k−1Xa,b=1 (−1)a+b+1[a][b] f (k−2) . b Proof First apply Lemma 2.10, and then apply the vertically reflected version of that lemma to all but the first term in the resulting expression. We sometimes consider the complete expansion of a Jones-Wenzl idempotent into a linear combination of Temperley-Lieb diagrams. Suppose β is a k -strand Tem- perley-Lieb diagram. Let Coeff f (k)(β) denote the coefficient of β in the expansion of f (k) . Thus f (k) =X Coeff f (k)(β)β, where the sum is over all k -strand Temperley-Lieb diagrams β . We will frequently state values of Coeff f (k)(β) without giving the details of how they are computed. A convenient formula for these values is given by Frenkel and Khovanov in [16]. See also [41] for a detailed exposition, including a helpful exam- ple at the end of §4 of that paper. A related formula was announced by Ocneanu [49], and special cases of this were proved by Reznikoff [59]. 10 2.3 The graph planar algebra In this section we define the graph planar algebra GP A(G) of a bipartite graph G with a chosen base point, and recall some of its basic properties [29]. Except in degenerate cases, this fails to be a subfactor planar algebra because dim GP A(G)0,+ and dim GP A(G)0,− are greater than 1. However, after specifying a certain parti- tion function, all the other axioms for a subfactor planar algebra hold. The box space GP A(G)n,± is the space of functionals on the set of loops on G that have length 2n and are based at an even vertex in the case of GP A(G)n,+ , or an odd vertex in the case of GP A(G)n,− . Suppose T is a planar tangle with k inner disks, and f1, . . . , fk are functionals in the appropriate spaces GP A(G)ni,±i . Then we will define T (f1, . . . , fk) as a certain "weighted state sum." A state on T is an assignment of vertices of G to regions of T and edges of G to strings of T , such that unshaded regions are assigned even vertices, shaded regions are assigned odd vertices, and the edge assigned to the string between two regions goes between the vertices assigned to those regions. In particular, a state for any graph is uniquely specified by giving only the assignment of edges, and a state for a simply laced graph is specified by giving only the assignment of vertices. Since all the graphs we consider are simply laced, we typically specify states by giving the assignment of vertices. The inner boundaries ∂i(σ) and outer boundary ∂0(σ) of a state are the loops obtained by reading the edges assigned to strings clockwise around the corresponding disk. We define T (f1, . . . , fk) by describing its value on a loop ℓ. This is given by the following weighted state sum. T (f1, . . . , fk)(ℓ) =Xσ c(T, σ) · Yi=1,...,k fi(∂i(σ)). Here, the sum is over all states σ on G such that ∂0(σ) = ℓ, and the weight c(T, σ) is defined below. To specify the weight c(T, σ), it helps to draw T in a certain standard form. Each disk is drawn as a rectangle, with the same number of strands meeting the top and bottom edges, no strands meeting the side edges, and the starred region on the left side. The strands are drawn smoothly, with a finite number of local maxima and minima. Then c(T, σ) = Yt∈E(T )s dσ(tconvex) dσ(tconcave) , where E(T ) is the set of local maxima and minima of the strands of T , dv is the Perron-Frobenius dimension of the vertex v , and tconvex and tconcave are respec- tively the regions on the convex and concave sides of t. The Perron-Frobenius dimension of a vertex is the corresponding entry in the Perron-Frobenius eigenvec- tor of the adjacency matrix. This is the largest-eigenvalue eigenvector, normalized so the Perron-Frobenius dimension of the base point is 1, and its entries are strictly positive. It is now easy to check that this planar algebra has modulus d, the Perron-Frobenius dimension of G. 11 Example 1 Fix G,asimplylacedgraph. Consider ρ8 = , the "two-click" rotation on 8-boxes, already drawn in standard form, and a loop γ = γ1γ2 . . . γ16γ1. Then γ1 γ2 γ3 γ4 γ5 γ6 γ7 γ8 ρ(f )(γ) = f γ16 γ15 γ14 γ13 γ12 γ11 γ10 γ9 f (γ3 . . . γ16γ1γ2γ3). dγ1dγ9 =s dγ3dγ11 v =Podd vertices v d2 It is a general fact about the Perron-Frobenius dimensions of bipartite graphs that v . Call this number I , the global index. We use Peven vertices v d2 the partition function Z :GP A0,± → C d2 v f (v) I f 7→Xv and the involution ∗ given by reversing loops: f∗(γ1γ2γ3 . . . γnγ1) := f (γ1γn . . . γ3γ2γ1). Proposition 2.12 For any bipartite graph G with base point the planar algebra with partition function and involution (GP A(G), Z,∗) is a spherical positive defi- niteplanaralgebrawhosemodulusisthePerron-Frobeniuseigenvaluefor G. Proof This is due to [29], but we recall the easy details here. The inner prod- uct is positive definite, because the basis of Kronecker-delta functionals on loops {δγ}γ∈Γ2k is an orthogonal basis and hδγ , δγi = dγ1 dγk+1 > 0. I 12 Sphericality is a straightforward computation: trl(X) = Z X X(e)   to odd vertices to odd vertices =Z Xedges e from even = Xedges e from even  =Z Xedges e from even = Xedges e from even to odd vertices to odd vertices X(e) X(e)dt(e) X(e)ds(e) dt(e) I ds(e) dt(e) · δt(e) dt(e) ds(e) · δs(e) ds(e) I , and trr(X) = Z X where s(e) and t(e) are respectively the even and odd vertices of the edge e. The main reason for interest in graph planar algebras is the following result from [46, 33]. Theorem 2.13 GivenafinitedepthsubfactorplanaralgebraP withprincipalgraph Γ thereisaninjectivemapofplanaralgebras P ֒→ GP A(Γ). This theorem assures us that if we believe in the existence of the extended Haagerup subfactor, and have enough perseverance, we will inevitably find it as a subalgebra of the graph planar algebra. Indeed, this paper is the result of such perseverance. On the other hand, nothing is this paper logically depends on the above theorem. 2.4 Notation for Hk Let dk be the Perron-Frobenius dimension of the graphs Hk . For the Haagerup subfactor, we have d0 =q 5+√13 2 ≈ 2.07431. For the extended Haagerup subfactor, d1 is the largest root of the polynomial x6 − 8x4 + 17x2 − 5, 3r 13 2 (cid:16)−5 − 3i√3(cid:17) + 3r 13 2 (cid:16)−5 + 3i√3(cid:17), d1 =s 8 1 3 1 3 + 3 approximately 2.09218. Throughout, if d is the modulus of a planar algebra, we let q be a solution to q + q−1 = d, and use the quantum integers [n] = qn − q−n q − q−1 . 13 By Hp about loops or paths on Hp k we mean the first graph in the pair of principal graphs Hk . When we talk 1 it is useful to have names for the vertices and arms. v0 w0 x0 y0 z0 a0 b0 c arm 0 → b2 a2 z2 b1 a1 z1 ← arm 2 ← arm 1 Lemma 2.14 If [2] = dk,then [3][4k + 4] = [4k + 8]. Proof The dimensions of the three vertices on an arm of Hp branch, are k , counting from the dim b1 = dim a1 = [4k + 5] 2 , [4k + 6] − [4k + 4] 2 , and dim z1 = [4k + 7] − [4k + 5] − [4k + 3] 2 . The condition [2] dim z1 = dim a1 easily gives the desired formula. 3 Uniqueness and existence 3.1 Uniqueness The goal of this section is to prove that there is at most one subfactor planar algebra with principal graphs Hk . To prove this, we will give a skein theoretic description of a planar algebra Qk/N (which is not necessarily a subfactor planar algebra). We then prove in Theorem 3.9 that any subfactor planar algebra with principal graphs Hk must be isomorphic to Qk/N . Definition 3.1 We say that a n-box S is uncappable if ǫi(S) = 0 for all i = 1, . . . , 2n where ǫ1 = ⋆⋆ ·· · , ǫ2 = ⋆⋆ ·· · , . . . , ǫ2n = ⋆ ⋆ . ·· · Wesay S isarotationaleigenvectorwitheigenvalue ω if ρ(S) = ωS where ρ = ⋆ . ·· · Notethat ω mustbea nth rootofunity. ⋆ 14 As described in [30], every subfactor planar algebra is generated by uncappable rotational eigenvectors. Definition 3.2 If S is an n-box, we use the following names and numbers for relationson S: (1) ρ(S) = −S, (2) S isuncappable, (3) S2 = f (n), (4) one-strand braiding substitute: S ⋆ 2n − 1 f (2n+2) 2n + 2 ⋆ S n − 1 ⋆ S = ip[n][n + 2] [n + 1] n + 1 n + 1 f (2n+2) 2n + 2 (5) two-strand braiding substitute: , ⋆ S ⋆ S n − 1 ⋆ S n − 1 S⋆ 2n f (2n+4) 2n + 4 = [2][2n + 4] [n + 1][n + 2] n + 1 2 n + 1 f (2n+4) 2n + 4 . We call relations (4) and (5) "braiding substitutes" because we think of them as allowing us to move a generator "through" strands, rather like an identity X∗ = X∗ (3.1) in a braided tensor category. The planar algebras we consider in this paper are not braided, and do not satisfy the Equation (3.1). Nevertheless, we found it useful to look for relations that could play a similar role. In particular, the evaluation algorithm described in §4 was inspired by the evaluation algorithms in [43] and [4] for planar algebras of types D2n , E6 , and E8 , where Equation (3.1) holds. Definition 3.3 Let Qk bethesphericalplanaralgebra ofmodulus [2] = dk,gener- atedbya (4k + 4)-box S,subjecttorelations(1)-(5) above. Definition 3.4 Anegligibleelementofa sphericalplanaralgebra P isanelement x ∈ Pn,± suchthatthediagrammatic trace tr0 (xy) iszeroforall y ∈ Pn,±. 15 The set N of all negligible elements of a planar algebra P forms a planar ideal of P . In the presence of an antilinear involution ∗, we can replace tr0 (xy) in the def- inition with tr0 (xy∗) without changing the ideal. If the planar algebra is positive definite, then N = 0. The following is well known. Proposition 3.5 Suppose P is a spherical planar algebra with non-zero modulus andN istheidealofnegligiblemorphisms. IfthespacesP0,± areone-dimensional theneverynon-trivialplanaridealiscontainedin N . Proof Suppose that a planar ideal I contains a non-negligible element x and with- out loss of generality assume x ∈ Pn,+ . Then there is some element y ∈ Pn,+ so tr0 (xy) 6= 0 ∈ P0,+ . The element tr0 (xy) is itself in the planar ideal, so since P0,+ is one-dimensional, it must be entirely contained in I , and so the unshaded empty diagram is in the ideal. Drawing a circle around this empty diagram, and using the fact that the modulus is non-zero, shows that the shaded empty diagram is also in the ideal. Now, every box space Pm,± is a module over P0,± under tensor product, with the empty diagram acting by the identity. Thus Pm,± ⊂ I for all m ∈ N. The sesquilinear pairing descends to Qk/N and is then nondegenerate. Let ρ1/2 denote the "one-click" rotation from Pn,+ to Pn,− given by . ·· · ⋆ ⋆ Definition 3.6 LettheHaagerupmomentsbeasfollows: • tr(cid:0)S2(cid:1) = [n + 1], • tr(cid:0)S3(cid:1) = 0, • tr(cid:0)S4(cid:1) = [n + 1], • tr(cid:0)ρ1/2(S)3(cid:1) = i [2n+2] √[n][n+2] . Proposition 3.7 Suppose P is a positive definite spherical planar algebra with modulus dk, and S ∈ Pn,+, where n = 4k + 4. If S is uncappable and satisfies ρ(S) = −S andtheHaagerupmomentsgiveninDefinition3.6then S satisfiesthe fiverelationsgiveninDefinition3.2. We defer the proof until §5. Theorem 3.8 IfP isaplanaralgebrathatissinglygeneratedbyan n-box S satisfy- ingthefiverelationsofDefinition3.2thenanycloseddiagramin P0,+ isamultiple oftheemptydiagram(so dim(Po,+) = 1). The proof of Theorem 3.8 is given in §4. 16 Theorem 3.9 Ifthereexistsasubfactorplanaralgebra P withprincipalgraphsHk then P isisomorphicto Qk/N . Proof The principal and dual principal graphs of P each have their first trivalent vertex at depth 4k+3. In the language of [31], P has n-excess one, where n = 4k+4. We follow [31, Section 5.1]. There exists S ∈ Pn,+ such that hS, T Ln,+i = 0, so Pn,+ = T Ln,+ ⊕ CS. Let r be the ratio of dimensions of the two vertices at depth 4k + 4 on the principal graph, chosen so that r ≥ 1. By [31], we can choose S to be self-adjoint, uncappable, and a rotational eigenvector, such that S2 = (1 − r)S + rf (n). (Our S is − R in [31].) The symmetry of Hp k implies that r = 1, so S2 = f (n) . We can use this to compute powers of S and their traces. These agree with the first three Haagerup moments, as given in Definition 3.6. In a similar fashion, one defines r ≥ 1 and S from the dual principal graph and calculates the moments of S . Since the complement of S is one-dimensional, S and ρ1/2(S) must be multiples of each other; that multiple can be calculated to be S = q r ωr ρ1/2(S). (This can be done by comparing (cid:10) S, S(cid:11) and (cid:10)ρ1/2(S), ρ1/2(S)(cid:11), as we learned from an earlier draft of [31]). Then it follows that r(r − 1)[n + 1], for some square root ω1/2 of the rotational eigenvalue of S . tr(cid:16)ρ1/2(S)3(cid:17) = ω3/2r r r (3.2) Although we will not use it here, we record the identity ρ1/2(S)2 = −ω1/2r1/2(r1/2 − r−1/2)ρ1/2(S) + ω−1rf (n), (3.3) which is equivalent to S2 = (1 − r) S + rf (n). By [31, Theorem 5.1.11], whenever P has n-excess one then r = [n+2] [n] and r + 1 r = 2 + 2 + ω + ω−1 [n][n + 2] . Since r = 1, this implies that ω = −1. Note that we could also compute r directly from the dual principal graph. Jones's proof that ω = −1 uses the converse of a result along the lines of Lemma 5.13, since there must be some linear relation of the form given in that Lemma. In the case r = 1 we also have the freedom to replace S with −S , which we use to (arbitrarily) choose the square root ω1/2 = −i. Substituting all these quantities into Equation (3.2), and using the identity [n + 1]([n + 2] − [n]) = [2n + 2], we now see that S has all of the Haagerup moments, as given in Definition 3.6. 17 By Proposition 3.7, S satisfies the relations given in Definition 3.2. Thus there is a planar algebra morphism Qk → P given by sending S to S . Since P is positive definite, this descends to the quotient to give a map Φ : Qk/N → P. By Proposition 3.5, Qk/N has no nontrivial proper ideals. Since Φ is non-zero, it must be injective. The image of Φ is the planar algebra in P generated by S . This is a subfactor planar algebra with the same modulus as P . Its principal graphs are not (A∞, A∞) because the dimension of the n-box space is too large. Haagerup's classification shows that the principal graphs of the image of Φ must be Hk . How- ever, since the principal graphs determine the dimensions of all box spaces, the image of Φ must be all of P . Thus Φ is an isomorphism of planar algebras. 3.2 Existence The subfactor planar algebra with principal graphs H0 is called the Haagerup pla- nar algebra, and is isomorphic to Q0/N . The corresponding subfactor was con- structed in [2] and [24]. The subfactor planar algebra was directly constructed in [52]. There is no subfactor planar algebra with principal graphs Hk for k > 1. In this case, by [3], Qk/N cannot be a finite depth planar algebra, let alone a subfactor planar algebra. The following theorem deals with the one remaining case. Theorem 3.10 Thereisasubfactorplanaralgebrawithprincipalgraphs H1. We prove this by finding H1 as a sub-planar algebra of the graph planar algebra of one of the extended Haagerup graphs. The following lemma simplifies the proof of irreducibility for subalgebras of graph planar algebras. Lemma 3.11 If P ⊂ GP A(G), dimP0,+ = 1, and G has an evenunivalent vertex, then P isanirreduciblesubfactorplanaralgebra. Proof To show that P is an irreducible subfactor planar algebra, we need to show that P is spherical and positive definite, and that dimP0,± = 1 and dimP1,+ = 1. By Proposition 2.12, the graph planar algebra is spherical and positive definite. The subalgebra P inherits both of these properties. We are given dimP0,+ = 1. Also, P0,− injects into P1,+ , (by tensoring with a strand on the left). It remains only to show that dimP1,+ = 1. Let v be an even univalent vertex of G. Let w be the unique vertex connected to v . Suppose X ∈ P1,+ is some functional on paths of length two in G. Now tr0 (X) is a closed diagram with unshaded exterior. (Note here we use tr0 , the diagrammatic trace, without applying a partition function, even though dim P0,+ = 1.) This is a functional defined on even vertices, via a state sum. Since v is univa- lent, the state sum for tr0 (X) (v) has only one term, giving tr0 (X) (v) = X(vw) dw dv . 18 Similarly, tr0 (X∗X) (v) = X(vw)X∗(vw) dw dv Thus if tr0 (X) (v) is zero then tr0 (X∗X) (v) is zero also. Note that tr0 (X) and tr0 (X∗X) are both scalar multiples of the empty diagram, and so tr0 (X∗X) (v) = 0 implies that tr0 (X∗X) = 0. Therefore, if tr0 (X) is zero then tr0 (X∗X) is zero. Then by positive definiteness, if tr0 (X∗X) is zero then X is zero. We conclude that the diagrammatic trace function is injective on P1,+ and thus P1,+ is one-dimensional. Recall the Haagerup moments from Definition 3.6. In the current setting, n = 8, [2] = d1 , and the Haagerup moments are as follows. • tr(cid:0)S2(cid:1) = [9] ≈ 24.66097, • tr(cid:0)S3(cid:1) = 0, • tr(cid:0)S4(cid:1) = [9], • tr(cid:0)ρ1/2(S)3(cid:1) = i Proposition 3.12 Supposethat S ∈ GP A(Hp 1)8,+ is self-adjoint, uncappable, a ro- tational eigenvector with eigenvalue −1, and has the above Haagerup moments. Let PA(S) be the subalgebra of GP A(Hp 1)8,+ generated by S. Then PA(S) is an irreduciblesubfactorplanaralgebrawithprincipalgraphs H1. [18]√[8][10] ≈ 15.29004i. Proof By Proposition 3.7, S ∈ GP A(Hp 1) satisfies all of the relations used to de- fine Q1 . Thus by Theorem 3.9, PA(S) is isomorphic to Q1/N . By Theorem 3.8, (Q1/N )0,+ is 1-dimensional. By Lemma 3.11 it follows that Q1/N is an irreducible subfactor planar algebra. By Haagerup's classification [18] it follows that the princi- pal graphs of Q1/N must be the unique possible graph pair with the correct graph norm, namely H1 . To prove Theorem 3.10, it remains to find S ∈ GP A(Hp 1)8,+ satisfying the require- ments of the above proposition. We defer this to Section 6, where we give an ex- plicit description of S and some long computations of the moments, assisted by computer algebra software. 4 The jellyfish algorithm The aim of this section is to prove that the relations of Definition 3.2 enable us to reduce any closed diagram built from copies of S to a scalar multiple of the empty diagram. 19 Questions of whether you can evaluate an arbitrary closed diagram are ubiquitous in quantum topology. The simplest such algorithms (e.g., the Kauffman bracket al- gorithm for knots) involve decreasing the number of generators (in this case, cross- ings) at each step. Slightly more complicated algorithms (e.g., HOMFLY evalua- tions) include steps that leave the number of generators constant while decreasing some other measure of complexity (such as the unknotting number). Another com- mon technique is to apply Euler characteristic arguments to find a small "face" (with generators thought of as vertices) that can then be removed. Again, the sim- plest such algorithms decrease the number of faces at every step (e.g., Kuperberg's rank 2 spiders [40]), while more difficult algorithms require steps that maintain the number of faces before removing a face (e.g., Peters' approach to H0 in [52]). In all of these algorithms, the number of generators is monotonically non-increasing as the algorithm proceeds. The algorithm we describe below is unusual in that it initially increases the number of generators in order to put them in a desirable configuration. We hope that this technique will be of wider interest in quantum topology (see [19] for a subsequent application of this technique). Therefore we have written this section to be independent of the rest of the paper, apart from references to Definitions 3.1 and 3.2. The algorithm we will describe gives a proof of Theorem 3.8, which we repeat from above: Theorem 3.8 IfP isaplanaralgebrathatissinglygeneratedbyan n-box S satisfy- ingthefiverelationsofDefinition3.2thenanycloseddiagramin P0,+ isamultiple oftheemptydiagram(so dim(Po,+) = 1). We do not actually need the full strength of the relations of Definition 3.2. The theorem is true for any planar algebra that is singly generated by an n-box S such that: • S is a rotational eigenvector: ρ(S) = ωS for some ω , • S is uncappable (see Definition 3.1), • S2 = aS + bf (n) for some scalars a and b (multiplication is defined in Figure 1), and • S satisfies one- and two-strand braiding substitutes of the form: ⋆ ⋆ S ⋆ S n − 1 S 2n − 1 n + 1 n + 1 f (2n+2) 2n + 2 = x f (2n+2) 2n + 2 S⋆ ⋆ S n − 1 ⋆ S n − 1 ⋆ S 2n n + 1 2 n + 1 f (2n+4) 2n + 4 = y 20 f (2n+4) 2n + 4 for some scalars x and y in C. Before going through the details, we briefly sketch the idea. First use the one- and two-strand braiding substitutes to pull all copies of S to the outside of the diagram. This will usually increase the number of copies of S . We can then guarantee that there is a pair of copies of S connected by at least n strands. This is a copy of S2 , which we can then express using fewer copies of S . All copies of S remain on the outside, and so we can again find a copy of S2 . Repeating this eventually gives an element of the Temperley-Lieb planar algebra, which is evaluated as usual. See Figure 2 for an example. We like to think of the copies of S as "jellyfish floating to the surface," and hence the name for the algorithm. Figure 2: The initial steps of the jellyfish algorithm. The dotted ovals represent lin- ear combinations of Temperley-Lieb diagrams. This is only a schematic illustration - to be precise, the result should be a linear combination of diagrams with various (sometimes large) numbers of copies of S . Definition 4.1 Suppose D isadiagramin P . Let S0 beafixedcopyofthegenera- torinside D. Suppose γ isanembeddedarcin D fromapointontheboundaryof S0 toapointonthetopedgeof D. Suppose γ isingeneralposition,meaningthat itintersectsthestrandsof D transversely,anddoesnottouchanygeneratorexcept at its initial point on S0. Let m be the number of points of intersection between γ and the strands of D. If m is minimal over all such arcs γ then we say γ is a geodesicand m isthedistancefrom S0 tothetopof D. Lemma 4.2 Suppose X is a diagram consisting of one copy of S with all strands pointingdown, and d parallel strandsforming a "rainbow" over S, where d ≥ 1. Then X isalinearcombinationofdiagramsthatcontainatmostthreecopiesof S, eachhavingdistancelessthan d fromthetopofthediagram. 21 Proof First consider the case d = 1. Up to some number of applications of the rotation relation ρ(S) = ωS , X is as shown in Figure 3. S ⋆ 2n − 1 Figure 3: X in the case d = 1. Recall that we have the relation ⋆ S n − 1 ⋆ S 2n − 1 n + 1 n + 1 S ⋆ f (2n+2) 2n + 2 = x f (2n+2) 2n + 2 . Consider what happens to the left side of the above relation when we write f (2n+2) as a linear combination of Temperley-Lieb diagrams β . The term in which β is the identity occurs with coefficient one, and gives the diagram X . Suppose β is not the identity. Then β contains a cup that connects two adjacent strands from X . If both ends of the cup are attached to S then the resulting diagram is zero. If not, then the cup must be at the far left or the far right of β . Such a cup converts X to a rotation of S , so gives distance zero from S to the top of the diagram. Now consider what happens to the right side of the above relation when we write f (2n+2) as a linear combination of Temperley-Lieb diagrams β . Every term in this expansion is a diagram with two copies of S , each of having distance zero from the top. By rearranging terms in the one-strand braiding substitute, we can write X as a linear combination of diagrams that contain one or two copies of S , each having distance zero from the top of the diagram. This completes the case d = 1. The case d = 2 is similar, but we use the two-strand braiding substitute. Finally, suppose d > 2. If d is odd then γ begins in a shaded region of X . Then X contains a copy of the diagram shown in Figure 3, up to the rotation relation ρ(S) = ωS . We can therefore use the one-strand braiding substitute, as we did in the case d = 1. Similarly, if d is even then we use the two-strand braiding substitute. Definition 4.3 Wesayadiagram D in P isinjellyfishformifalloccurrencesof S lieinarowatthetopof D,andallstrandsof D lieentirelybelowtheheightofthe topsofthecopiesof S. 22 Lemma 4.4 Every diagram in P is a linear combination of diagrams in jellyfish form. Proof Suppose D is a diagram in P (not necessarily closed), drawn in such a way that all endpoints lie on the bottom edge of D. If every copy of the generator in D is distance zero from the top edge of D then D is already in jellyfish form, up to isotopy. If not, we will use Lemma 4.2 to pull each copy of S to the top D. It is convenient for our proof, but not necessary for the algorithm, to move copies of our generator S along geodesics. Suppose S0 is a copy of S that has distance d from the top of D, where d ≥ 1. Let γ be a geodesic from S0 to the top edge of D. Let X be a small neighborhood of S0 ∪ γ . By applying an isotopy, we consider X to be a diagram in a rectangle, consisting of a copy of S0 with all strands pointing down, and a "rainbow" of d strands over it. By Lemma 4.2, X is a linear combination of diagrams that contain at most three copies of S , each having distance less than d from the top of the diagram. Let X′ be one of the terms in this expression for X . Let D′ be the result of replacing X by X′ in D. Suppose S1 is a copy of the generator in D′ . If S1 lies in X′ then the distance from S1 to the top of D′ is at most d − 1. Now suppose S1 does not lie in X′ . By basic properties of geodesics, there is a geodesic in D from S1 to the top of D that does not intersect γ . This geodesic is still a path in general position in D′ , and still intersects strands in the same number of points. Thus the distance from S1 to the top of D does not increase when we replace X by X′ . In summary, if we replace X by X′ , then S0 will be replaced by one, two or three copies of S that are closer to the top of D, and no other copy of S will become farther from the top of D. Although the number of copies of S may increase, it is not hard to see that this process must terminate. For example, we have decreased the sum over each generator S0 of 4 to the power of the distance from S0 to the top. We now prove Theorem 3.8, that dim(P0,+) = 1. Proof of Theorem 3.8. Suppose D is a closed diagram with unshaded exterior. We must show that D is a scalar multiple of the empty diagram. By the previ- ous lemma, we can assume D is in jellyfish form. We can also assume there are no closed loops or cups attached to generators, so that every strand must connect two different copies of the generator. We argue that there is a copy of the generator whose strands only go to one or both of its immediate neighbors. This is a simple combinatorial fact about this kind of planar graph. Think of the copies of the generator as vertices, and consider all strands that do not connect adjacent vertices. Amongst these, find one that has the smallest (positive) number of vertices between its endpoints. Any vertex between the endpoints of this strand can connect only to its two neighbors. 23 Let S0 be a copy of the generator such that the strands of S0 only go to one or both of its immediate neighbors. Then S0 is connected to some neighbor, S1 , by at least n parallel strands. See Figure 4 for an example. Recall that S2 = aS + bf (n) . Thus we can replace S0 and S1 with aS + bf (n) , giving a linear combination of diagrams that are still in jellyfish form, but contain fewer copies of the generator. By induction, D is a scalar multiple of the empty diagram. Figure 4: Jellyfish form, illustrating (with n = 8) the proof of Theorem 3.8. 5 Relations from moments In this section we prove Proposition 3.7, which says that certain conditions on an element S imply the five relations of Definition 3.2. We use this Proposition once in the proof of Theorem 3.9, and once in the proof of Theorem 3.10. These are the uniqueness and existence results. In the proof of uniqueness, we must show that a certain subfactor planar algebra P is isomorphic to Qk/N . We use Proposition 3.7 to show that an element S of P satisfies the defining relations of Qk/N . In the proof of existence, we must show that a certain subalgebra P of a graph planar algebra has a one-dimensional space of n-boxes. We use Proposition 3.7 to show that the generator S of P satisfies relations, which we then use in the algorithm of §4. Most of this section consists of computations of inner products between diagrams. Since the values of these inner products may be useful for studying other planar algebras, we strive to use weaker assumptions whenever possible. Assumption 5.1 P isasphericalplanaralgebrawithmodulus [2] = q+q−1,where q isnotarootofunity(sowecansafelydividebyquantumintegers). Furthermore, S ∈ Pn,+ isuncappableandhasrotationaleigenvalue ω. Recall that the Haagerup moments are as follows. • tr(cid:0)S2(cid:1) = [n + 1], • tr(cid:0)S3(cid:1) = 0, • tr(cid:0)S4(cid:1) = [n + 1], • tr(cid:0)ρ1/2(S)3(cid:1) = i [2n+2] √[n][n+2] . 24 Assumption 5.2 P ispositivedefiniteandhasmodulus [2] = dk,where n = 4k+4. Furthermore, S ∈ Pn,+ has rotational eigenvalue ω = −1, and has the Haagerup moments. Restatement of Proposition 3.7 Suppose P is a planar algebra, S ∈ Pn,+, and Assumptions 5.1 and 5.2 hold. Then S satisfies the five relations given in Defini- tion3.2. The proof involves some long and difficult computations, but the basic idea is very simple. We will define diagrams A, B , C and D. We must prove certain linear re- lations hold between A and B , and between C and D. Since P is positive definite, we can do this by computing certain inner products. In general, there is a linear relation between X and Y if and only if In this case, hX, XihY, Y i = hX, Y i2. hY, Y iX − hX, Y iY = 0, as can be seen by taking the inner product of this expression with itself. To compute the necessary inner products we must evaluate certain closed diagrams. Most of these closed diagrams involve a Jones-Wenzl idempotent. In principal, we could expand this idempotent into a linear combination of Temperley-Lieb di- agrams, and evaluate each resulting tangle in terms of the moments, or using rela- tions that have already been proved. In practice, we must take care to avoid dealing with an unreasonably large number of terms. 5.1 Definitions and conventions Notation We use the notation that a thick strand in a Temperley-Lieb diagram alwaysrepresents n − 1 parallelstrands. Forexample, istheidentityof T L2n+2. Definition 5.3 For m ≥ 0,let asin[31,Definition4.2.6]. Wm = qm + q−m − ω − ω−1, The diagrams A, B , C and D of Figures 5 and 6 are the terms in the "braiding" relations we wish to prove. Along the way, we will also use the diagrams Γ and B′ , as shown in Figure 7. 25 ⋆ S n − 1 ⋆ S 2n − 1 n + 1 n + 1 S ⋆ A = f (2n+2) , B = f (2n+2) . 2n + 2 2n + 2 Figure 5: A and B S⋆ ⋆ S n − 1 ⋆ S n − 1 ⋆ S 2n n + 1 2 n + 1 C = f (2n+4) 2n + 4 , D = f (2n+4) 2n + 4 . Figure 6: C and D S ⋆ 2 2 n − 1 n − 1 Γ = n − 1 S ⋆ 2 S ⋆ , B′ = n − 1 S ⋆ Figure 7: Γ and B′ ⋆ S n − 1 ⋆ S n + 1 n − 1 f (2n+2) . 2n + 2 5.2 Computing inner products We now calculate the necessary inner products. Lemma 5.4 IfAssumption5.1holdsthen hA, Ai = 1 [2n + 2] Thesameholdswiththereverseshading. W2n+2tr(cid:0)S2(cid:1) . 26 Proof We must evaluate the closed diagram hA, Ai = Z  S ⋆ 2n − 1 f (2n+2) 2n − 1 ⋆ S .  The idea is to apply Lemma 2.11 to the copy of f (2n+2) , and then evaluate each of the resulting diagrams. Consider the first term. Here, f (2n+2) is replaced by a copy of f (2n+1) together with a single vertical strand on the right. Since the planar algebra is spherical, we can drag this strand over to the left. This results in a partial trace of f (2n+1) , which is equal to [2n + 2] [2n + 1] f (2n). By noting that S · f (n) = S , we obtain [2n + 2] [2n + 1] tr(cid:0)S2(cid:1) as the value of the first term. Now consider the terms in the sum over a and b. Here, f (2n+2) is replaced by a copy of f (2n) together with a "cup" and a "cap" in positions given by a and b. In most cases, the resulting diagram is zero because S is uncappable. We only need to consider the four cases where a, b ∈ {1, 2n + 1}. Each of these gives tr(cid:0)S2(cid:1), up to some rotation of one or both copies of S . We obtain 1 [2n + 1][2n + 2] (−1 − [2n + 1]2 − [2n + 1]ω − [2n + 1]ω−1)tr(cid:0)S2(cid:1) . The result now follows by adding the above two expressions and writing the quan- tum integers in terms of q . Lemma 5.5 IfAssumption5.1holdsandeither ω = −1 or S2 = f (n) then hA, Bi = tr(cid:16)ρ1/2(S)3(cid:17) . 27 Proof We must evaluate the closed diagram hA, Bi = Z  S ⋆ 2n − 1 f (2n+2) n + 1 n + 1 S ⋆ n − 1 S ⋆ .  Consider the complete expansion of f (2n+2) into a linear combination of Temperley- Lieb diagrams β ∈ T L2n+2 . For most such β , the resulting diagram is zero because S is uncappable. There are only three values of β we need to consider. For each of these, we compute the corresponding coefficient, and easily evaluate the corre- sponding diagram. The results are shown in Table 1. β Coeff f (2n+2)(β) value of diagram 1 [2n+2] (−1)n+1 [n+1] (−1)n+1 [n+1] [2n+2] tr(cid:0)ρ1/2(S)3(cid:1) tr(cid:0)S3(cid:1) ω−1tr(cid:0)S3(cid:1) Table 1: The terms of f (2n+2) that contribute to hA, Bi. Now take the sum over all β in the table of the coefficient times the value of the diagram. Note that if S2 = f (n) then tr(cid:0)S3(cid:1) = 0. Thus the two non-identity values of β either cancel or give zero. The term where β is the identity gives the desired result. Lemma 5.6 IfAssumption5.1holdsand S2 = f (n) then hB, Bi = [n + 1][2n + 2] [n][n + 2] . Proof Consider the two diagrams hB, Bi = Z  ⋆ S S ⋆ n − 1 ⋆ S n + 1 n + 1 f (2n+2) n + 1 n + 1 n − 1 S ⋆   f ( n + 1 ) n + 1 n + 1 f (2n+2) n + 1 n + 1 S ⋆ n − 1 S ⋆ .  , Df (n+1), BE = Z 28 The second is clearly zero. We will compare what happens to each of these dia- grams when we expand the copy of f (2n+2) into a linear combination of Temperley- Lieb diagrams. Let β be a Temperley-Lieb diagram in the expansion of f (2n+2) . Suppose β con- tains a cup that connects endpoints number i and i+1 at the top for some i 6= n+1. The corresponding diagram for hB, Bi is zero because the cup connects two strands from the same copy of S . Similarly, the corresponding diagram for hf (n+1), Bi is zero because the cup connects two strands from the same side of f (n+1) . Thus both diagrams corresponding to β are zero. Now suppose β contains a cup in the middle, connecting endpoints number n + 1 and n + 2 at the top. In the corresponding diagram for hB, Bi, this cup produces a copy of S2 , which we can replace with f (n) . For hf (n+1), Bi, this cup produces a partial trace of f (n+1) , which we can replace with [n+2] [n+1] f (n) . Thus the two diagrams corresponding to β differ only by a factor of [n+2] [n+1] . Finally, suppose β is the identity diagram. The corresponding diagram for hB, Bi consists of four copies of S arranged in a rectangle. The left and right sides of this rectangle consist of n + 1 parallel strands. We can replace one of these sides with a partial trace of f (n) . Thus the diagram is equal to [n+1] [n] tr(cid:0)S2(cid:1). The corresponding diagram for hf (n+1), Bi is equal to tr(cid:0)S2(cid:1). We can now evaluate hB, Bi = hB, Bi − [n + 1] [n + 2]hf (n+1), Bi. For both terms on the right hand side, we express f (2n+2) as a linear combination of Temperley-Lieb diagrams β . For every β except the identity, these terms cancel. The identity term gives the following. [n + 1] hB, Bi = [n] tr(cid:0)S2(cid:1) − [n + 1] [n + 2] tr(cid:0)S2(cid:1) . The result now follows from a quantum integer identity, and that fact that tr(cid:0)S2(cid:1) = [n + 1]. Lemma 5.7 IfAssumption5.1holdsthen hC, Ci = [2][2n + 2] [2n + 3][2n + 4] W2n+4hA, Ai. Proof We must evaluate the closed diagram hC, Ci = Z  S⋆ 2n f (2n+4) 2n S⋆ 29 .  The proof is very similar to that of Lemma 5.4. Indeed, these are both special cases of a general recursive formula. The idea is to apply Lemma 2.11 to the copy of f (2n+4) , and then evaluate each of the resulting diagrams. Consider the first term. Here, f (2n+4) is replaced by a copy of f (2n+3) together with a single vertical strand on the right. We can use sphericality to drag this strand over to the left. This results in a partial trace of f (2n+3) , which is Using Lemma 5.4, we obtain [2n + 4] [2n + 3] f (2n+2). [2n + 4] [2n + 3]hA, Ai. Now consider the terms in the sum over a and b. Here, f (2n+4) is replaced by a copy of f (2n+2) together with a "cup" and a "cap" in positions given by a and b. In most cases, the resulting diagram is zero because S is uncappable. A cup at the leftmost position also gives zero since it connects two strands coming from the top right of f (2n+2) . Similarly, a cup in the rightmost position gives zero, as does a cap in the leftmost or rightmost position. The only cases that give a non-zero diagram are when a, b ∈ {2, 2n + 2}. We obtain 1 [2n + 3][2n + 4] (−[2]2 − [2n + 2]2 − [2][2n + 2]ω − [2][2n + 2]ω−1)hA, Ai. The result now follows by adding the above two expressions and expanding the quantum integers in terms of q . Lemma 5.8 IfAssumption5.1holdsand S2 = f (n) then hD, Di = [n + 1]2[2n + 2] [2][n]2[2n + 3] ([n + 3] − [2][n]). Proof We must evaluate the closed diagram hD, Di = Z  ⋆ S S ⋆ n − 1 ⋆ S n − 1 ⋆ S n + 1 2 n + 1 f (2n+4) n + 1 2 n + 1 n − 1 S ⋆ n − 1 S ⋆ .  Consider the expansion of f (2n+4) into a linear combination of Temperley-Lieb di- agrams β ∈ T L2n+4 . For each such β , we compute the coefficient and the value of the corresponding diagram. There are twelve values of β that give a non-zero diagram. Many of these are reflections or rotations of each other. They are shown in Table 2. 30 Coeff f (2n+4)(β) value of diagram 1 [n+1] β , , , , [n][n+1][n+2][n+3] − [n+1][n+3] [2n+4] (cid:16) [n+1] [n] (cid:17)2 tr(cid:0)S2(cid:1) [n] tr(cid:0)S4(cid:1) (cid:16) [n+1] [n] (cid:17)2 tr(cid:0)S2(cid:1) [n] (cid:17)2 (cid:16) [n+1] tr(cid:0)S2(cid:1) [n] tr(cid:0)S4(cid:1) tr(cid:0)S6(cid:1) Table 2: The terms of f (2n+4) that contribute to hD, Di. [2n+3][2n+4] [2][n+1]2[n+2]2 [2n+3][2n+4] − [n][n+1]2[n+2] [2][2n+3][2n+4] [2][2n+3][2n+4] [n]2[n+1]2 [n+1] , , Now take the sum over all β in the table of the coefficient times the value of the diagram. Since S2 = f (n) , we have tr(cid:0)S2(cid:1) = tr(cid:0)S4(cid:1) = tr(cid:0)S6(cid:1) . Lemma 5.9 IfAssumption5.1holdsand S2 = f (n) and ω = −1 then hC, Di = Z(Γ) + 2 [n] + 1 [2n + 3]hA, Ai. Proof First we prove a formula for D. ⋆ S n − 1 ⋆ S n − 1 ⋆ S n 2 n + 1 D = f (2n+3) 2n + 3 . (5.1) Apply a left to right reflection of Lemma 2.10 to the copy of f (2n+4) in D. The first term gives the desired diagram. Now consider a term in the sum over a. This con- tains a cup in a position given by a. For all but two values of a, this cup connects two strands from the same copy of S , so the resulting diagram is zero. The remain- ing two values of a are n + 1 and n + 3. For each of these, the cup connects two dif- ferent copies of S , giving rise to a copy of S2 . We can replace this with f (n) , which is a linear combination of Temperley-Lieb diagrams. But any such Temperley-Lieb diagram gives rise to a cup connected to the top edge of f (2n+3) , and thus gives zero. This completes the derivation of Equation (5.1). 31 Next we prove the following. ⋆ S n − 1 ⋆ S n − 1 ⋆ S n 2 n ⋆ S 2n D = f (2n+2) 2n + 2 + 1 [2n + 3] f (2n+2) 2n + 4 . (5.2) To prove this, apply Lemma 2.10 to the copy of f (2n+3) in Equation (5.1). The first term in the expansion gives the first term in the desired expression for D. It remains to show that the sum over a is equal to the second term in the desired expression. Consider a term for a 6∈ {n, n + 2}. The cup connects two strands from the same copy of S , giving zero. Consider the term corresponding to a = n + 2. The position of the cup is such that the right two copies of S are connected by n strands. This is a copy of S2 , which is equal to f (n) , which in turn is a linear combination of Temperley-Lieb diagrams. Any such Temperley-Lieb diagram results in a cap connected to the top edge of f (2n+2) , giving zero. Now consider the term corresponding to a = n. The coefficient of this term is (−1)n+1 [n] [2n + 3] . The left two copies of S form a copy of S2 , which is equal to a sideways copy of f (n) , which in turn we express as a linear combination of Temperley-Lieb diagrams β . Every such β gives zero except which has coefficient β = ... , (−1)n+1 1 [n] and gives the second diagram in the desired expression for D. The total coeffi- cient of this diagram is the product of the above coefficients for the term a and the diagram β . This completes the derivation of Equation (5.2). Now we return to our computation of hC, Di. We must evaluate the expression Z  .  S⋆ 2n f (2n+4) n + 1 2 n + 1 n − 1 S ⋆ n − 1 S ⋆ S ⋆ 32 Apply Equation (5.2), upside down, to the bottom half of this diagram. For the last term of this equation, apply sphericality and use Lemma 5.4 to reverse the shading. We obtain the term 1 [2n + 3]hA, Ai. The first term from the equation gives Z  S⋆ 2n f (2n+2) n 2 n S ⋆ n − 1 S ⋆ n − 1 S ⋆ .  (5.3) We expand f (2n+2) into a linear combination of Temperley-Lieb diagrams β . There are five values of β we need to consider. These are shown in Table 3. β Coeff f (2n+2)(β) value of diagram 1 [2n+2] (−1)n [n+2] (−1)n [n+2] (−1)n [n] [2n+2] [2n+2] Z(Γ) (−1)n+1 1 (−1)n+1 1 (−1)n+1 1 [n] ωtr(cid:0)S2(cid:1) [n] ω−1tr(cid:0)S2(cid:1) [n] tr(cid:0)S2(cid:1) , Table 3: The terms of f (2n+2) that contribute to (5.3). Now take the sum over all β in the table of the coefficient times the value of the diagram. The following inner products involving B′ will help us to evaluate Z(Γ). Lemma 5.10 IfAssumption5.1holdsand S2 = f (n) and ω = −1 then hA, B′i = [n − 1] [n + 1] tr(cid:16)ρ1/2(S)3(cid:17) 33 Proof We must evaluate the closed diagram Z  Z  .  .  The proof is very similar to that of Lemma 5.5, so we will omit the details. The relevant table is shown in Table 4. β Coeff f (2n+2)(β) value of diagram [2n] [2n+2] [2] [2n+2] tr(cid:0)ρ1/2(S)3(cid:1) ω−1tr(cid:0)ρ1/2(S)3(cid:1) Table 4: The terms of f (2n+2) that contribute to hA, B′i. Lemma 5.11 IfAssumption5.1holdsand S2 = f (n) and ω = −1 then hB, B′i = Z(Γ) + [2][n + 1] [n][n + 2] . Proof We must evaluate the closed diagram S ⋆ 2n − 1 f (2n+2) n + 1 n − 1 2 S ⋆ n − 1 S ⋆ ⋆ S n − 1 ⋆ S n + 1 n + 1 f (2n+2) n + 1 n − 1 2 S ⋆ n − 1 S ⋆ Inspired by the proof of Lemma 5.6, we observe that hB, B′i = hB, B′i − [n + 1] [n + 2]hf (n+1), B′i. 34 We expand the copies of f (2n+2) on the right hand side. By the same argument as for Lemma 5.6, all terms will cancel except for those coming from the identity diagram. If we replace f (2n+2) by the identity in hB, B′i then we obtain Γ. If we replace f (2n+2) by the identity in hf (n+1), B′i then we obtain a diagram containing two copies of S and one copy of f (n+1) . We must now expand f (n+1) as a linear combi- nation of Temperley-Lieb diagrams β . For all but one such diagram β , the resulting diagram is zero because S is uncappable. The only diagram we need to consider is β = ... , which has coefficient [2] [n][n+1] and gives the diagram ω−1tr(cid:0)S2(cid:1). 5.3 Proving relations We now use our inner products, together with Assumptions 5.1 and 5.2, to prove that the required relations hold. Note that the assumption ω = −1 implies W2m =(cid:18) [2m] [m](cid:19)2 . Lemma 5.12 IfAssumptions5.1and5.2holdthen S2 = f (n). Proof The relevant inner products are as follows. • hS2, S2i = tr(cid:0)S4(cid:1) = [n + 1], • hS2, f (n)i = tr(cid:0)S2(cid:1) = [n + 1], • hf (n), f (n)i = tr(cid:0)f (n)(cid:1) = [n + 1]. The inner product of S2 − f (n) with itself is zero, and the result follows from the assumption that P is positive definite. Lemma 5.13 IfAssumptions5.1and5.2holdthen A = i √[n][n+2] [n+1] B. Proof By Lemmas 5.4, 5.5, 5.6, and our values for the moments, we have the fol- lowing. • hA, Ai = [2n+2] [n+1] , [2n+2] √[n][n+2] • hA, Bi = i , • hB, Bi = [n+1][2n+2] [n][n+2] . Thus Thus A and B are linearly dependent. The precise relation is then hA, AihB, Bi = hA, Bi2. A = hA, Bi hB, Bi B. 35 Lemma 5.14 IfAssumptions5.1and5.2holdthen Z(Γ) = [n − 1][2n + 2] − [2][n + 1] [n][n + 2] . Proof By Lemmas 5.10 and 5.11, [2n+2] • hA, B′i = [n−1] √[n][n+2] [n+1] i • hB, B′i = Z(Γ) + [2][n+1] [n][n+2]. √[n][n+2] . By Lemma 5.13, A = i [n+1] B . Thus hA, B′i = ip[n][n + 2] [n + 1] hB, B′i. The result follows by solving for Z(Γ). Lemma 5.15 IfAssumptions5.1and5.2holdthen C = [2][2n + 4] [n + 1][n + 2] D. Proof By Lemmas 5.7, 5.9, 5.8 and our values for Z(Γ) and the moments, we have the following. • hC, Ci = [2][2n+2]2[2n+4] [n+1][n+2]2[2n+3] . • hC, Di = [2n+2]2 [n+2][2n+3] . • hD, Di = [n+1]2[2n+2] [2][n]2[2n+3] ([n + 3] − [2][n]). Here, we have used quantum integer identities to simplify the expression for hC, Di. For arbitrary n, m, and q , [n + m] = 1 [4] ([4 − m][n] + [m][n + 4]), and By Lemma 2.14 and the assumption [2] = dk , [2m] = [m]([m + 1] − [m − 1]). [n + 4] = [3][n]. (This is the only time we use the assumption [2] = dk .) We can now express each of our inner products in terms of [n], [2], [3], and [4]. After some computation we find that Thus C and D are linearly dependent. The precise relation is then hC, CihD, Di = hC, Di2. C = hC, Ci hC, Di D. 36 6 Properties of the generator In this section we construct an 8-box S ∈ GP A(Hp 1) that satisfies the hypotheses of Proposition 3.12. The planar algebra generated by S is the desired extended Haagerup planar algebra, thus completing the proof of Theorem 3.10. We start with a brief description of how we found S , since the definition of S is not very enlightening on its own. The goal was to find S ∈ GP A(Hp 1)8 satisfying the first three relations of Definition 3.2, which say that ρ(S) = −S , S is uncap- pable, and S2 = f (8) . The dimension of GP A(Hp 1)8 is the number of loops of length 16 based at even vertices of Hp 1 , which is equal to 148375. We found the 19-dimensional space of solutions to the first two relations, then tried to solve the equation S2 = f (8) . This one equation in the 8-box space is actually 148375 sep- arate equations over C. We expect of course that there are many redundancies amongst these equations. At this point the problem sounds quite tractable, but we were still unable to solve it by general techniques. We then used various ad hoc methods. First, we searched for quadratics that are perfect squares and solved those. This reduced the problem from 19 variables to 9. We then chose a small collection of quadratics, correspond- ing to certain 'extremal' loops in the basis, and found numerical approximations to a solution of these, using Newton's method. Approximating such a solution by al- gebraic numbers, we could then go back and check that all the quadratic equations are satisfied exactly. Remark. Our solution S need not be unique. Although the subfactor planar alge- bra with principal graphs H1 is unique, there may be more than one way to embed it in its graph planar algebra. Indeed, −S is also a solution, which corresponds to applying the graph automorphism, by Lemma 6.4. Due to the approximate nature of our search for S , we cannot say whether there are any other solutions. The above description may sound daunting, but we manage to give a definition of S that involves specifying only 21 arbitrary looking numbers, and we reduce all the conditions we need to check on this element to computing certain powers of two matrices. The definition of S is still somewhat overwhelming, and the verification of its properties is done by computer. Those intrepid readers who continue reading this section will have to read a short Mathematica program in order to fully verify some of the steps. We will use the notation for vertices of Hp 1 given in Section 2.4. Let λ =p2 − d2, C = −21516075λ4 + 8115925λ2 + 45255025. For each w ∈ {0, 1, 2}8 we will define an element pw ∈ Z[λ] below. Definition 6.1 Suppose γ is a loop of length 16 in Hp 1. Let the collapsed loopbγ bethesequencein {0, 1, 2}8 suchthat γ2i−1 isinarmnumberbγi (inthenotationof Section2.4)of Hp vertices v0, w0, zi and ai appearin γ. 1. Further,define σ(γ) tobe −1 raisedtothenumberoftimesthe S(γ) = Cσ(γ)pbγ 1 pdγ1dγ9 37 16Yi=1 , 1 pdγi (6.1) wherethe pbγ aredefinedbelow. Remark. The main reason we write S in the above form is that the pbγ have much better number theoretic properties than the S(γ). In particular, the pbγ all live in a degree 6 extension of Q while the S(γ) live in a much larger number field. As a consequence it is easier to do exact arithmetic using the pbγ . There is a general rea- son for this phenomenon: the convention that subfactor planar algebras be spher- ical is not the best convention from the point of view of number theory. A much more convenient convention is the "lopsided" one where shaded circles count for 1 while unshaded circles count for the index d2 . Furthermore, this convention is also well-motivated from the perspective of subfactor theory where N MM has as its left von Neumann dimension the index while its right von Neumann dimension is 1. These issues warrant further investigation (see [44, 46]). We make an apparently ad-hoc definition of 21 elements of Z[λ]. p00000001 = −2λ4 − λ2 + 9 p00000101 = 2λ4 + λ2 − 9 p00001001 = λ5 − λ3 − 3λ p00010001 = 2λ4 + λ2 − 9 p00010101 = λ4 − 2λ2 + 1 p00011011 = λ4 − λ2 − 3 p00100111 = λ2 + 1 p00101101 = λ5 − λ p00110111 = −λ5 − 2λ3 − 4λ2 − λ − 5 p01010111 = λ4 + λ2 p01110111 = λ4 + 6λ2 + 6 p00000011 = −λ5 − λ3 + 3λ p00000111 = 1 p00001011 = λ3 − 1 p00010011 = λ5 − λ4 + λ2 − 3λ + 4 p00010111 = −λ4 + 1 p00100101 = −2λ4 + 5 p00101011 = −λ5 − λ3 + λ + 1 p00110011 = 2λ5 + 5λ3 + 4λ p01010101 = −4λ4 + 3λ2 + 7 p01011011 = λ4 − 2λ2 − 4 These elements are also defined in a Mathematica notebook, available along with the sources for this article on the arXiv (as the file extra/code/Generator.nb), or at http://tqft.net/EH/notebook. A PDF printout of the notebook is available by following this URL, then replacing .nb with .pdf. Everything that follows in this section is paralleled in the notebook, and in particular each of the statements below that requires checking some arithmetic has a corresponding test defined in the notebook. Definition-Lemma 6.2 We can consistently extend these definitions to every pw for w ∈ {0, 1}8 bytherules and pabcdef gh = −pbcdef gha, pabcdef gh = pahgf edcb, p00000000 = 0, pabcd1111 = 0. 38 (6.2) (6.3) (6.4) (6.5) Proof For example, one can get from p00110011 to p01100110 either by rotating, or by reversing; fortunately p00110011 is purely imaginary. Under the operations im- plicit in Equations (6.2) and (6.3) each orbit in {0, 1}8 contains exactly one of the elements on which p is defined above or in Equations (6.4) and (6.5). The Mathe- matica notebook provides functions VerifyRotation and VerifyConjugation to check that these rules hold uniformly. We further extend these definitions to every pw for w ∈ {0, 1, 2}8 by the rules px0y + px1y + px2y = 0. Lemma 6.3 Forevery abcd ∈ {0, 1, 2}4 pabcd2222 = 0. (6.6) (6.7) Proof This is a direct computation of 16 cases for abcd ∈ {0, 1}4 using Equation (6.6), after which the general case of abcd ∈ {0, 1, 2}4 follows, again from (6.6). The Mathematica notebook provides a function Verify2sVanish that checks this Lemma. A final interesting note on the pw : Lemma 6.4 If w′ isobtainedfrom w byexchangingall 1sand 2s,then pw = −pw′ . Proof A direct computation which you can verify using the Mathematica function VerifyGraphSymmetry. This ends the definition of S . We now prepare to prove that it has the properties required to generate a subfactor planar algebra with principal graph H1 . Lemma 6.5 The generator S is self-adjoint, has rotational eigenvalue −1 and is uncappable: S∗ = S, ρ(S) = −S, ǫi(S) = 0 for i = 1, . . . , 2k. Lemma 6.6 The generator S and its "one-click" rotation ρ1/2(S) have the follow- ingmoments: tr0(cid:0)S2(cid:1) = [9], tr0(cid:0)S3(cid:1) = 0, tr0(cid:0)S4(cid:1) = [9], tr0(cid:16)ρ1/2(S)3(cid:17) = i Note that the scalars on the right sides of these equations actually refer to scalar multiples of the empty diagram in GP A(Hp 1)0,±. In particular, each of them is a constantfunctionontheeven(oroddinthelastcase)verticesofthegraph Hp 1. p[8][10] [18] . 39 Proof of Lemma 6.5. Self-adjointness follows immediately from Equation (6.3). To show ρ(S(γ)) = −S(γ), first note that σ(γ) andQ16 are independent of the order of vertices appearing in γ = γ1 ··· γ16γ1 . Thus, recalling Example 1 from §2.3, 1√dγi i=1 ρ(S)(γ) =s dγ3 dγ11 =s dγ3 dγ11 dγ9dγ1 dγ9dγ1 S(γ3 . . . γ16γ1γ2γ3) Cσ(γ)p \γ3...γ16γ1γ2γ3 1 16Yi=1 1 pdγi 1 pdγ11 dγ3 pdγi = Cσ(γ)(−pbγ) = −S(γ) 1 pdγ9 dγ1 16Yi=1 where we used Equation (6.2) in the second to last step. Next consider ǫi(S), the result of attaching a cap on strands i and i + 1. Since we know S is a rotational eigenvector, we only need to check ǫ1(S) = ǫ2(S) = 0. In particular, we don't need to explicitly treat the more complicated cases of ǫ8(S) and ǫ16(S), in which the cap is attached "around the side" of S , and the coefficients coming from critical points in the graph planar algebra are more complicated. Let Γk be the set of length-k loops on Hp The graph planar algebra formalism tells us that for ϕ ∈ Γ14 , 1 , and γi denote the ith vertex of γ ∈ Γk . ǫi(S)(ϕ) = Xγ ∈ Γ16 with γj≤i=ϕj,γi+2=ϕi and γj≥i+3 = ϕj−2 s dγi+1 dγi S(γ) We consider three cases, depending on whether the valence of ϕi is 1, 2 or 3. If ϕi has valence 1, that is, it is an endpoint, then there is just one term in the sum: if γi = v0 , then γi+1 = w0 , and if γi = zj , then γi+1 = aj . In the first case, the collapsed loopbγ must be 00000000, so S(γ) = 0 by Equation (6.4). In the second case, if γi = z1 thenbγ must contain at least 4 consecutive 1s, so S(γ) = 0 by Equation (6.5). If γi = z2 , then bγ must contain at least 4 consecutive 2s, so S(γ) = 0 by Lemma 6.3. If ϕi has valence 2, that is, it lies on one of the arms, then there are two terms in the sum, say γ+ and γ− . Moreover, the collapsed loops for the two terms are the 40 same, and σ(γ+) = −σ(γ−). Thus ǫi(S)(ϕ) = Cp cγ± 1 dϕj × 14Yj=1 1 9 1 dγ± vuut qdγ+ vuut dγ± i+1 9 1 qdγ± σ(γ+) qdγ± 1 1 = Cp cγ± + σ(γ−) i i+1 dγ+ dγ+ vuut dϕj  σ(γ+) qdγ+ 1 i dγ+ i+2 dγ+ i+2 14Yj=1 1 i+1 qdγ− qdγ− + i dγ− i+2 σ(γ−) dγ− i+2 i i+1 dγ− dγ− vuut   = 0. Since γ+ i = γ−i = γ+ i+2 = γ−i+2 = ϕi , the two terms in the parentheses cancel exactly. Finally, if ϕi has valence 3, then it must be the triple point, c. There are then three terms, say γ0 , γ1 and γ2 , with γj i+1 = bj . Now the collapsed paths differ; are all equal. Thus we obtain bγj = w1jw2 for some fixed words w1 and w2 . On the other hand, the signs σ(γj)  . dϕj  ǫi(S)(ϕ) = Cσ(γj) vuut 14Yj=1 pcγ0 pcγ1 pcγ2 qdγ0 qdγ1 qdγ2 i+2 = ϕi , the three terms in the parentheses cancel exactly by Equation qdγj i = γj dγj dγ0 dγ1 dγ2 + + i+2 i+2 i+2 1 1 1 9 i i i Since γj (6.6). We now verify the moments of S are the Haagerup moments. Computer-assisted proof of Lemma 6.6. We first treat the moments of S , and later describe the changes required to calculate the moments of ρ1/2(S). With multiplication given by the multiplication tangle from Figure 1, the vector space GP A(H1)8,+ becomes a finite-dimensional semisimple associative algebra, which of course must just be a multimatrix algebra. It is easy to see that the simple summands are indexed by pairs of even vertices, and that the minimal idempo- tents in the summand indexed by (s, t) are given by symmetric loops of length 16, which go from s to t in 8 steps, then return the same way. Since there are 8 even vertices (v0, x0, z0, b0, b1, b2, z1 and z2 ), there are 64 simple summands As,t , although four of these (Av0,z1,Az1,v0,Av0,z2 , and Az2,v0 ) are trivial because s and t are more than 8 edges apart. Moreover, the trace tangle from Figure 1 composed with the partition function puts a trace on each of these matrix algebras. We write As,t = (cid:16)Mk×k, dt the trace of the identity in Mk×k is dt ds(cid:17) to indicate there are k paths of length 8 from s to t, and that (GP A(H1)8,+, multiplication tangle) ∼= Ms,t k . We find that As,t. ds even vertices Now to compute the required moments, we just need to identify the image of S in this multimatrix algebra, compute the appropriate powers via matrix multiplica- tion, and take weighted traces. It turns out that the necessary calculation, namely 41 taking kth powers of the matrices for S , for k = 2, 3 and 4, is actually computation- ally difficult! First notice that some of the matrices are quite large, up to 118 × 118. Worse than this, the entries are quite complicated numbers, involving square roots of dimensions, and so the arithmetic step of simplifying matrix entries after multi- plication turns out to be extremely slow. One can presumably do these calculations directly with the help of a computer, using exact arithmetic, but our implementa- tion in Mathematica took more than a day attempting to simplify the matrix entries in S4 before we stopped it. Instead, we choose a matrix (really, a multimatrix) A so that all the entries of ASA−1 lie in the number field Q(λ); this matrix certainly has the same moments as S , but once the computer can do its arithmetic inside a fixed number field, everything happens much faster. In particular, the moments required here take less than an hour to compute, using Mathematica 7 on a 2.4Ghz Intel Core 2 Duo. See the remark following Definition 6.1 for an explanation of why this trick works: we cooked up the matrix A with the desired property by compar- ing the usual definition of the graph planar algebra with an alternative definition that produces the corresponding "lopsided" planar algebra. The matrix A is defined by (As,t)π,ǫ = δπ=ǫ (6.8) 8Yi=1pdπi recalling that the matrix entries in As,t are indexed by pairs of paths π, ǫ from s to t, so π = π1 ··· π9 and ǫ = ǫ1 ··· ǫ9 with π1 = ǫ1 = s and π9 = ǫ9 = t. Notice that the index in the product ranges from 1 to 8, leaving out the endpoint t. Lemma 6.7 Theentriesof ASA−1 liein Q(λ). (The proof appears below.) The second half of the Mathematica notebook referred to above produces the matri- ces for ASA−1 (these, and the corresponding matrices for ρ1/2(S) described below, are also available at http://tqft.net/EH/matrices in machine readable form and as a PDF typeset for an enormous sheet of paper) and actually does the moment calculation. Any reader wanting to check the details should look there. Here, we'll just indicate the schematic calculation: dt ds tr(cid:0)S2(cid:1) (s) =Xt =Xt tr(cid:0)(Ss,t)2(cid:1) tr(cid:0)(As,tSs,tA−1 s,t )2(cid:1) dt ds approximately 8 minutes later... = [9] tr(cid:0)S3(cid:1) (s) =Xt =Xt dt ds dt ds tr(cid:0)(Ss,t)3(cid:1) tr(cid:0)(As,tSs,tA−1 s,t )3(cid:1) approximately 16 minutes later.... = 0 42 tr(cid:0)S4(cid:1) (s) =Xt =Xt dt ds dt ds tr(cid:0)(Ss,t)4(cid:1) tr(cid:0)(As,tSs,tA−1 s,t )4(cid:1) approximately 24 minutes later..... = [9]. Note that in each case above we're actually computing 8 potentially different num- bers, as s ranges over the even vertices of the graph. The moments of ρ1/2(S) can be calculated by a very similar approach. The other 8-box space GP A(H1)8,− becomes a multimatrix algebra with summands indexed by pairs of odd vertices on the graph H1 . Lemma 6.8 Theentriesof As,tρ1/2(S)s,tA−1 s,t liein d · Q(λ). (Again, the proof appears below.) We thus compute tr(cid:16)ρ1/2(S)3(cid:17) = d3tr(cid:16)(d−1As,tρ1/2(S)s,tA−1 s,t )3(cid:17) . As before, this is implemented in Mathematica. The calculation takes slightly longer than in the first case. The details can be found in the notebook. Proof of Lemma 6.7 Let ⊔ denote concatenation of paths, and ¯ǫ be the reverse of the path ǫ. We readily calculate 1 √dsdt 8Yi=1 1 pdπi 9Yi=2 1 i=1pdπi pdǫiQ8 i=1pdǫi Q8 (As,tSs,tA−1 s,t )π,ǫ = Cσπ⊔¯ǫp[π⊔¯ǫ Cσπ⊔¯ǫp[π⊔¯ǫ i=1 dǫi = Q9 Q9 in d · Q[λ2]. So the productQ9 sions, lies in d4Q[d2] ⊂ Q(λ) and (ASA−1)γ,ǫ ∈ Q(λ). Most of the factors in this product are already in Q(λ); the one in question is i=1 dǫi . All even dimensions are in Q[d2] = Q[λ2], and all odd dimensions are i=1 dǫi , a product of five even and four odd dimen- Proof of Lemma 6.8 First, we have ρ1/2(S)(γ1γ2 ··· γ16γ1) =s dγ2 dγ10 =s dγ2 dγ10 dγ9 dγ1 dγ9 dγ1 S(γ2 ··· γ16γ1γ2) Cσ(γ)p \γ2···γ16γ1γ2 16Yi=1 1 pdγi 1 pdγ2dγ10 16Yi=1 pdγi 1 . = Cσ(γ)p \γ2···γ16γ1γ2 1 pdγ1dγ9 43 Be careful here: although this looks very similar to the formula in Equation (6.1) for S , the path γ here starts at an odd vertex. We now conjugate by a multimatrix A that has exactly the same formula for its definition as appears in Equation (6.8), except again the paths γ and ǫ start and finish at odd vertices. We obtain (As,tρ1/2(S)s,tA−1 s,t )π,ǫ = rσπ⊔¯ǫp \ǫ2π1···π8ǫ9···ǫ2 . One readily checks that these matrix entries are in d · Q(λ). i=1 dǫi Q9 We've finally shown the existence and uniqueness of the extended Haagerup sub- factor. Uniqueness is Theorem 3.9. By Lemma 6.5 and Lemma 6.6, S satisfies the hypotheses of Proposition 3.12. Therefore PA(S) is a subfactor planar algebra with principal graphs H1 . A Fusion categories coming from the extended Haagerup subfactor The even parts of a subfactor are the unitary tensor categories of N −N and M −M bimodules respectively. Hence every finite depth subfactor yields two unitary fu- sion categories. In terms of the planar algebra, the simple objects in these categories are the irreducible projections in the box spaces P2m,± for some m. In the case of extended Haagerup, the global dimension of each of these fusion categories is the largest real root of x3 − 585x2 + 8450x − 21125 (approximately 570.247). The fusion tables are given in Figures 8 and 9. 1 f (2) f (4) f (6) P Q A B ⊗ f (2) f (4) f (6) P Q A B f (2) f (4) 1+f (2) +f (4) f (2) +f (4)+f (6) f (2) +f (4)+f (6) f (4)+W A+W B +W P Q 1+f (2)+ f (4) +W f (2)+f (4)+ A+B +2W f (4) +B +2W f (4)+A+2W f (6) +Q f (6)+P f (6) f (4) +W f (2) +f (4)+ A+B +2W P B +W Q A+W A Q B P f (4) +A+2W f (4) +B +2W f (6) +P f (6)+Q 1+W +Z f (6) +Q+Z f (6) +P +Z f (4) +B +W f (4) +A+W f (6) +Q+Z f (6) +P +Z f (4) +B +W f (4) +A+W f (2) +B +W 1+P +Z f (6) +Z f (4) +W f (6) +Z 1+Q+Z f (2) +A+W f (4) +W f (2) +A+W f (4) +W f (4)+W f (2) +B +W f (6) 1+Q 1+P f (6) Figure 8: The simple objects and fusion rules for the N − N fusion category coming from the extended Haagerup subfactor. We use the abbreviations W = f (6) + P + Q and Z = A + B + f (2) + 2f (4) + 3f (6) + 3P + 3Q. 44 1 f (2) f (4) f (6) P ′ Q′ ⊗ f (2) f (4) f (2) 1+f (2) +f (4) f (2)+f (4) +f (6) f (6) f (4) +f (6)+P ′+Q′ P ′ Q′ f (6) +2P ′+Q′ f (6)+P ′ f (4) f (2)+f (4) +f (6) 1+f (2) +f (4) +f (6)+P ′+Q′ f (2) +f (4)2f (6) +3P ′+Q′ f (4)+3f (6) +3P ′ +2Q′ f (4) +f (6) +2P ′+Q′ f (6) f (4)+f (6) +P ′+Q′ f (2)+f (4) +2f (6) +3P ′+Q′ 1+f (2) +2f (4) +4f (6) +5P ′+3Q′ f (2) +3f (4) +5f (6) +6P ′ +3Q′ f (2)+f (4) +3f (6) +3P ′ +2Q′ P ′ f (6)+2P ′ +Q′ f (4) +3f (6) +3P ′ +2Q′ Q′ f (6) +P ′ f (4) +f (6) +2P ′+Q′ f (2) +3f (4) +5f (6) f (2)+f (4) +3f (6) +6P ′ +3Q′ 1+2f (2) +3f (4) +6f (6) +7P ′+4Q′ f (2) +2f (4) +3f (6) +4P ′ +2Q′ +3P ′ +2Q′ f (2) +2f (4) +3f (6) +4P ′ +2Q′ 1+f (4) +2f (6) +2P ′+Q′ Figure 9: The simple objects and fusion rules for the M−M fusion category coming from the extended Haagerup subfactor. References [1] Marta Asaeda, Galois groups and an obstruction to principal graphs of subfactors, Internat. J. Math. 18 (2007) 191 -- 202, MR2307421 DOI:10.1142/S0129167X07003996 arXiv:math.OA/0605318 [2] Marta Asaeda, Uffe Haagerup, Exotic subfactors of finite depth with Jones indices (5 + √13)/2 and (5 + √17)/2 , Comm. Math. Phys. 202 (1999) 1 -- 63, MR1686551 DOI:10.1007/s002200050574 arXiv:math.OA/9803044 [3] Marta Asaeda, Seidai Yasuda, On Haagerup's list of potential principal graphs of subfactors, Comm. Math. Phys. 286 (2009) 1141 -- 1157, MR2472028 DOI:10.1007/s00220-008-0588-0 arXiv:0711.4144 [4] Stephen Bigelow, Skein theory for the ADE planar algebras, J. Pure Appl. Algebra 214 (2010) 658 -- 666, http://dx.doi.org/10.1016/j.jpaa.2009.07.010, arXiv:math.QA/0903.0144 MR2577673 DOI:10.1016/j.jpaa.2009.07.010 [5] Jocelyne Bion-Nadal, An example of a subfactor of the hyperfinite II 1 factor whose principal graph invariant is the Coxeter graph E6 , from: "Current topics in operator algebras (Nara, 1990)", World Sci. Publ., River Edge, NJ (1991) 104 -- 113, MR1193933 [6] Dietmar Bisch, An example of an irreducible subfactor of the hyperfinite II 1 factor with rational, noninteger index, J. Reine Angew. Math. 455 (1994) 21 -- 34, http://dx.doi.org/10.1515/crll.1994.455.21, arXiv:MR1293872 [7] Dietmar Bisch, Bimodules, higher relative commutants and the fusion algebra associated to a subfactor, from: "Operator algebras and their applications (Water- loo, ON, 1994/1995)", Fields Inst. Commun. 13, Amer. Math. Soc., Providence, RI (1997) 13 -- 63, MR1424954 (preview at google books) [8] Dietmar Bisch, Principal graphs of subfactors with small Jones index, Math. Ann. 311 (1998) 223 -- 231, MR1625762 DOI:http://dx.doi.org/10.1007/s002080050185 [9] Dietmar Bisch, Subfactors and planar algebras, from: "Proceedings of the Interna- tional Congress of Mathematicians, Vol. II (Beijing, 2002)", Higher Ed. Press, Beijing (2002) 775 -- 785, MR1957084 arXiv:math.OA/0304340 [10] Dietmar Bisch, Remus Nicoara, Sorin Popa, Continuous families of hyperfinite subfactors with the same standard invariant, Internat. J. Math. 18 (2007) 255 -- 267, MR2314611 arXiv:math.OA/0604460 DOI:10.1142/S0129167X07004011 45 [11] Alain Connes, Noncommutative geometry, Academic Press Inc., San Diego, CA (1994), MR1303779 [12] Antoine Coste, Terry Gannon, Remarks on Galois symmetry in rational conformal field theories, Phys. Lett. B 323 (1994) 316 -- 321, MR1266785 DOI:10.1016/0370-2693(94)91226-2 [13] Jan de Boer, Jacob Goeree, Markov traces and II1 factors in conformal field theory, Comm. Math. Phys. 139 (1991) 267 -- 304, MR1120140 euclid.cmp/1104203304 [14] Sergio Doplicher, John E Roberts, A new duality theory for compact groups, In- vent. Math. 98 (1989) 157 -- 218, MR1010160 DOI:10.1007/BF01388849 [15] Pavel Etingof, Dmitri Nikshych, Viktor Ostrik, On fusion categories, Ann. of Math. (2) 162 (2005) 581 -- 642, MR2183279 DOI:10.4007/annals.2005.162.581 arXiv:math.QA/0203060 [16] Igor B Frenkel, Mikhail G Khovanov, Canonical bases in tensor products and graphical calculus for Uq(sl2) , Duke Math. J. 87 (1997) 409 -- 480, MR1446615 [17] Alice Guionnet, Vaughan F R Jones, Dimitri Shlyakhtenko, Random matrices, free probability, planar algebras and subfactors, arXiv:0712.2904 [18] Uffe Haagerup, Principal graphs of subfactors in the 4 < [M : N ] < 3 + √2, Sci. River http://tqft.net/other-papers/subfactors/haagerup.pdf "Subfactors (1994) Edge, NJ (Kyuzeso, Publ., from: 1 -- 38, MR1317352 index range 1993)", World at available [19] Richard Han, A Construction of the 2221 Planar Algebra, Ph.D. thesis, University of California, Riverside (2010), arXiv:1102.2052 [20] Seung-Moon Hong, Eric Rowell, Zhenghan Wang, On exotic modular tensor categories, Commun. Contemp. Math. 10 (2008) 1049 -- 1074, MR2468378 DOI:10.1142/S0219199708003162 arXiv:0710.5761 [21] Kei Ikeda, Numerical evidence for flatness of Haagerup's J. Math. Sci. Univ. Tokyo 5 http://tqft.net/other-papers/subfactors/ikeda.pdf (1998) 257 -- 272, MR1633929 available connections, at [22] Masaki Izumi, Application of fusion rules to classification of subfactors, Publ. Res. Inst. Math. Sci. 27 (1991) 953 -- 994, MR1145672 DOI:10.2977/prims/1195169007 [23] Masaki Izumi, On flatness of the Coxeter graph E8 , Pacific J. Math. 166 (1994) 305 -- 327, MR1313457 euclid.pjm/1102621140 [24] Masaki Izumi, The structure of sectors associated with Longo-Rehren II. Examples, Rev. Math. Phys. 13 (2001) 603 -- 674, MR1832764 inclusions. DOI:10.1142/S0129055X01000818 [25] Masaki Izumi, Vaughan F R Jones, Scott Morrison, Noah Snyder, Classification of subfactors of index less than 5, part 3: quadruple points, in preparation. [26] Vaughan F R Jones, Planar algebras, I, arXiv:math.QA/9909027 [27] Vaughan F R Jones, Index for subfactors, Invent. Math. 72 (1983) 1 -- 25, MR696688 DOI:10.1007/BF01389127 [28] Vaughan F R Jones, Braid groups, Hecke algebras and type II 1 factors, from: "Ge- ometric methods in operator algebras (Kyoto, 1983)", Pitman Res. Notes Math. Ser. 123, Longman Sci. Tech., Harlow (1986) 242 -- 273, MR866500 [29] Vaughan F R Jones, The planar algebra of a bipartite graph, from: "Knots in Hellas '98 (Delphi)", Ser. Knots Everything 24, World Sci. Publ., River Edge, NJ (2000) 94 -- 117, MR1865703 (preview at google books) 46 [30] Vaughan F R Jones, The annular structure of subfactors, from: "Essays on geome- try and related topics, Vol. 1, 2", Monogr. Enseign. Math. 38, Enseignement Math., Geneva (2001) 401 -- 463, MR1929335 [31] Vaughan F R Jones, Quadratic tangles in planar algebras (2003), arXiv:1007.1158 [32] Vaughan F R Jones, Two subfactors and the algebraic decomposition of bimodules over II1 factors. (2008), pre-print available at http://math.berkeley.edu/~vfr [33] Vaughan F R Jones, David Penneys, The embedding theorem for finite depth subfactor planar algebras, arXiv:1007.3173 [34] Vaughan F R Jones, Dimitri Shlyakhtenko, Kevin Walker, An orthogonal approach to the subfactor of a planar algebra, arXiv:0807.4146 [35] Andr´e Joyal, Ross Street, The geometry of tensor calculus. I, Adv. Math. 88 (1991) 55 -- 112, MR1113284 [36] Andr´e Joyal, Ross Street, An introduction to Tannaka duality and quantum groups, "Category theory (Como, 1990)", Lecture Notes in Math. 1488, Springer, Berlin (1991) 413 -- 492, MR1173027 available at http://www.maths.mq.edu.au/~street/CT90Como.pdf from: [37] Louis H Kauffman, State models and the Jones polynomial, Topology 26 (1987) 395 -- 407, http://dx.doi.org/10.1016/0040-9383(87)90009-7, MR899057 [38] Yasuyuki Kawahigashi, On flatness of Ocneanu's connections on the Dynkin J. Funct. Anal. 127 (1995) 63 -- 107, diagrams and classification of subfactors, MR1308617 DOI:10.1006/jfan.1995.1003 [39] Vijay Kodiyalam, Viakalathur S Sunder, From subfactor planar algebras http://dx.doi.org/10.1142/S0129167X0900573X, to arXiv:0807.3704 MR2574313 DOI:10.1142/S0129167X0900573X subfactors (2009), [40] Greg Kuperberg, Spiders for rank 2 Lie algebras, Comm. Math. Phys. 180 (1996) 109 -- 151, MR1403861 arXiv:q-alg/9712003 euclid.cmp/1104287237 [41] Scott Morrison, A formula for the Jones-Wenzl projections, unpublished, available at http://tqft.net/math/JonesWenzlProjections.pdf [42] Scott Morrison, David Penneys, Emily Peters, Noah Snyder, Classification of subfactors of index less than 5, part 2: triple points (2010), arXiv:1007.2240 [43] Scott Morrison, Emily Peters, Noah Snyder, planar D2n http://dx.doi.org/10.1016/j.jpaa.2009.04.010, MR2559686 DOI:10.1016/j.jpaa.2009.04.010 Pure Appl. Algebra algebras, J. Skein theory for the 214 117 -- 139, arXiv:math/0808.0764 (2010) [44] Scott Morrison, Noah Snyder, Non-cyclotomic fusion categories (2010), arXiv:1002.0168 [45] Scott Morrison, Noah Snyder, Subfactors of index less than 5, part 1: the principal graph odometer (2010), arXiv:1007.1730 [46] Scott Morrison, Kevin Walker, The graph planar algebra embedding theorem, preprint available at http://tqft.net/gpa [47] Adrian Ocneanu, Quantized groups, string algebras and Galois theory for algebras, from: "Operator algebras and applications, Vol. 2", London Math. Soc. Lecture Note Ser. 136, Cambridge Univ. Press, Cambridge (1988) 119 -- 172, MR996454 [48] Adrian Ocneanu, Chirality for operator algebras, from: "Subfactors (Kyuzeso, 1993)", World Sci. Publ., River Edge, NJ (1994) 39 -- 63, MR1317353 [49] Adrian Ocneanu, The classification of subgroups of quantum SU(N ) , from: "Quan- tum symmetries in theoretical physics and mathematics (Bariloche, 2000)", Con- temp. Math. 294, Amer. Math. Soc., Providence, RI (2002) 133 -- 159, MR1907188 47 [50] David Penneys, James Tener, Classification of subfactors of index less than 5, part 4: cyclotomicity (2010), arXiv:1010.3797 [51] Roger Penrose, Applications of negative dimensional tensors, from: "Combinato- rial Mathematics and its Applications (Proc. Conf., Oxford, 1969)", Academic Press, London (1971) 221 -- 244 [52] Emily Peters, A planar algebra construction of the Haagerup subfactor (2009), arXiv:0902.1294, to appear in Internat. J. Math [53] Sorin Popa, Classification of subfactors: the reduction to commuting squares, In- vent. Math. 101 (1990) 19 -- 43, MR1055708 DOI:10.1007/BF01231494 [54] Sorin Popa, Subfactors and classification in von Neumann algebras, from: "Pro- ceedings of the International Congress of Mathematicians, Vol. I, II (Kyoto, 1990)", Math. Soc. Japan, Tokyo (1991) 987 -- 996, MR1159284 [55] Sorin Popa, Classification of amenable subfactors of type II, Acta Math. 172 (1994) 163 -- 255, MR1278111 DOI:10.1007/BF02392646 [56] Sorin Popa, An axiomatization of the lattice of higher relative commutants of a subfactor, Invent. Math. 120 (1995) 427 -- 445, MR1334479 DOI:10.1007/BF01241137 [57] Sorin Popa, Dimitri Shlyakhtenko, Universal properties of L(F∞) in subfactor theory, Acta Math. 191 (2003) 225 -- 257, MR2051399 DOI:10.1007/BF02392965 [58] Nicolai Reshetikhin, Vladimir G Turaev, Invariants of 3-manifolds via link polynomials and quantum groups, Invent. Math. 103 (1991) 547 -- 597, MR1091619 euclid.cmp/1104180037 [59] Sarah A Reznikoff, Coefficients of the one- and two-gap boxes in the 3129 -- 3150, Jones-Wenzl http://dx.doi.org/10.1512/iumj.2007.56.3140, MR2375712 Indiana Univ. Math. idempotent, J. 56 (2007) [60] Harold N V Temperley, Elliott H Lieb, Relations between the "percolation" and "colouring" problem and other graph-theoretical problems associated with regular planar lattices: some exact results for the "percolation" problem, Proc. Roy. Soc. London Ser. A 322 (1971) 251 -- 280, MR0498284 [61] Vladimir G Turaev, Quantum invariants of knots and 3-manifolds, volume 18 of de Gruyter Studies in Mathematics, Walter de Gruyter & Co., Berlin (1994), MR1292673 (preview at google books) [62] Vladimir G Turaev, Oleg Ya Viro, State sum invariants of 3-manifolds and quantum 6j -symbols, Topology 31 (1992) 865 -- 902, MR1191386 [63] Stefaan Vaes, Explicit computations of all finite index bimodules for a family of II 1 factors, Ann. Sci. ´Ec. Norm. Sup´er. (4) 41 (2008) 743 -- 788, MR2504433 [64] Hans Wenzl, On sequences of projections, C. R. Math. Rep. Acad. Sci. Canada 9 (1987) 5 -- 9, MR873400 [65] Hans Wenzl, On the structure of Brauer's centralizer algebras, Ann. of Math. (2) 128 (1988) 173 -- 193, MR951511 DOI:10.2307/1971466 This paper is available online at arXiv:0909.4099, and at http://tqft.net/EH. 48
1510.07987
3
1510
2016-10-07T12:01:34
Bi-exact groups, strongly ergodic actions and group measure space type III factors with no central sequence
[ "math.OA", "math.DS", "math.GR" ]
We investigate the asymptotic structure of (possibly type III) crossed product von Neumann algebras $M = B \rtimes \Gamma$ arising from arbitrary actions $\Gamma \curvearrowright B$ of bi-exact discrete groups (e.g. free groups) on amenable von Neumann algebras. We prove a spectral gap rigidity result for the central sequence algebra $N' \cap M^\omega$ of any nonamenable von Neumann subalgebra with normal expectation $N \subset M$. We use this result to show that for any strongly ergodic essentially free nonsingular action $\Gamma \curvearrowright (X, \mu)$ of any bi-exact countable discrete group on a standard probability space, the corresponding group measure space factor ${\rm L}^\infty(X) \rtimes \Gamma$ has no nontrivial central sequence. Using recent results of Boutonnet-Ioana-Salehi Golsefidy [BISG15], we construct, for every $0 < \lambda \leq 1$, a type III$_\lambda$ strongly ergodic essentially free nonsingular action $\mathbf F_\infty \curvearrowright (X_\lambda, \mu_\lambda)$ of the free group $\mathbf F_\infty$ on a standard probability space so that the corresponding group measure space type III$_\lambda$ factor ${\rm L}^\infty(X_\lambda, \mu_\lambda) \rtimes \mathbf F_\infty$ has no nontrivial central sequence by our main result. In particular, we obtain the first examples of group measure space type III factors with no nontrivial central sequence.
math.OA
math
BI-EXACT GROUPS, STRONGLY ERGODIC ACTIONS AND GROUP MEASURE SPACE TYPE III FACTORS WITH NO CENTRAL SEQUENCE CYRIL HOUDAYER AND YUSUKE ISONO Abstract. We investigate the asymptotic structure of (possibly type III) crossed product von Neumann algebras M = B ⋊Γ arising from arbitrary actions Γ y B of bi-exact discrete groups (e.g. free groups) on amenable von Neumann algebras. We prove a spectral gap rigidity result for the central sequence algebra N ′ ∩ M ω of any nonamenable von Neumann subalgebra with normal expectation N ⊂ M . We use this result to show that for any strongly ergodic essentially free nonsingular action Γ y (X, µ) of any bi-exact countable discrete group on a standard probability space, the corresponding group measure space factor L∞(X) ⋊ Γ has no nontrivial central sequence. Using recent results of Boutonnet -- Ioana -- Salehi Golsefidy [BISG15], we construct, for every 0 < λ ≤ 1, a type IIIλ strongly ergodic essentially free nonsingular action F∞ y (Xλ, µλ) of the free group F∞ on a standard probability space so that the corresponding group measure space type IIIλ factor L∞(Xλ, µλ) ⋊ F∞ has no nontrivial central sequence by our main result. In particular, we obtain the first examples of group measure space type III factors with no nontrivial central sequence. 1. Introduction and statement of the main results The group measure space construction of Murray and von Neumann [MvN43] associates to any ergodic (essentially) free nonsingular action Γ y (X, µ) of a countable discrete group on a standard probability space a factor denoted by L∞(X)⋊Γ. A fundamental question in operator algebras is how much information does the group measure space factor L∞(X) ⋊ Γ retain from the group action Γ y (X, µ)? This question has attracted a lot of attention during the last 15 years and several important developments regarding the structure and the rigidity of group measure space factors have been made possible thanks to Popa's deformation/rigidity theory [Po06a]. We refer the reader to [Ga10, Va10, Io12b] for recent surveys on this topic. One of the questions we address in this paper is the following general problem: Under which assumptions on the countable discrete group Γ and the ergodic free nonsingular action Γ y (X, µ), the group measure space factor L∞(X) ⋊ Γ is full? Recall from [Co74] that a factor M with separable predual is full if its asymptotic centralizer Mω is trivial for some (or any) nonprincipal ultrafilter ω ∈ β(N) \ N. By [AH12, Theorem 5.2], a factor M with separable predual is full if and only if its central sequence algebra M ′ ∩ M ω is trivial for some (or any) nonprincipal ultrafilter ω ∈ β(N) \ N (see Section 2 for further details). If the group measure space factor L∞(X)⋊Γ is full then the free nonsingular action Γ y (X, µ) is necessarily strongly ergodic, that is, any Γ-asymptotically invariant sequence of measurable subsets of X is trivial. The converse is not true in general as demonstrated in the celebrated example by Connes and Jones [CJ81]. Indeed, they exhibited an example of a strongly ergodic free probability measure preserving (pmp) action such the associated group measure space II1 factor is McDuff, that is, tensorially absorbs the hyperfinite II1 factor of Murray and von Neumann. 2010 Mathematics Subject Classification. 46L10, 46L36, 46L06, 37A20. Key words and phrases. Bi-exact discrete groups; Full factors; Group measure space construction; Ozawa's condition (AO); Popa's intertwining techniques; Strongly ergodic actions; Ultraproduct von Neumann algebras. CH is supported by ERC Starting Grant GAN 637601. YI is supported by JSPS Research Fellowship. 1 2 CYRIL HOUDAYER AND YUSUKE ISONO The general problem mentioned above has nevertheless a satisfactory answer in the case when the action Γ y (X, µ) is pmp. Indeed, it was shown by Choda in [Ch81] that when the countable discrete group Γ is not inner amenable and the free pmp action Γ y (X, µ) is strongly ergodic, then the group measure space II1 factor L∞(X) ⋊ Γ is full. The facts that the group Γ is not inner amenable and the action Γ y (X, µ) is pmp imply that all the central sequences in L∞(X) ⋊ Γ must asymptotically lie in L∞(X). It follows immediately that L∞(X) ⋊ Γ is full if the action is strongly ergodic. In the above reasoning, the assumption that the action Γ y (X, µ) is pmp is crucial since nonamenable (and in particular non-inner amenable) groups always admit an amenable (in the sense of Zimmer [Zi84, Definition 4.3.1]) type III ergodic nonsingular action, namely the Poisson boundary action. Very little is known about the general problem mentioned above when the action Γ y (X, µ) is no longer pmp and is more generally nonsingular (possibly of type III). In this paper, we investigate the asymptotic structure of (possibly type III) group measure space factors L∞(X) ⋊ Γ and more generally of (possibly type III) crossed product von Neumann algebras B ⋊ Γ arising from arbitrary actions Γ y B of bi-exact discrete groups on amenable von Neumann algebras. The class of bi-exact discrete groups was introduced by Ozawa in [Oz04] (see also [BO08, Chapter 15]) and includes amenable groups, free groups, Gromov word-hyperbolic groups and discrete subgroups of connected simple Lie groups of real rank one. We refer the reader to Section 2 for a precise definition. Any bi-exact discrete group is either amenable or non-inner amenable [Oz04]. Ozawa's celebrated result [Oz03] asserts that bi-exact discrete groups Γ give rise to solid group von Neumann algebras L(Γ), that is, for any diffuse von Neumann algebra A ⊂ L(Γ), the relative commutant A′ ∩ L(Γ) is amenable. Moreover, any solid II1 factor is either amenable or full [Oz03, Proposition 7]. Recall that an inclusion of von Neumann algebras N ⊂ M is with expectation if there exists a faithful normal conditional expectation EN : M → N . Our first main result is a spectral gap rigidity result inside crossed product von Neumann algebras M = B ⋊ Γ arising from arbitrary actions Γ y B of bi-exact discrete groups on amenable σ-finite von Neumann algebras. More precisely, we prove that for any von Neumann subalgebra with expectation N ⊂ M , either N has a nonzero amenable direct summand or the central sequence algebra N ′ ∩ M ω lies in the smaller algebra Bω ⋊ Γ. Our Theorem A can be regarded as an analogue of the spectral gap rigidity results discovered by Peterson in [Pe06, Theorem 4.3] and Popa in [Po06b, Theorem 1.5] and [Po06c, Lemma 2.2]. Theorem A. Let Γ be any bi-exact discrete group, B any amenable σ-finite von Neumann algebra and Γ y B any action. Denote by M := B ⋊ Γ the corresponding crossed product von Neumann algebra. Let p ∈ M be any nonzero projection and N ⊂ pM p any von Neumann subalgebra with expectation. Let ω ∈ β(N) \ N be any nonprincipal ultrafilter. Then at least one of the following conditions holds true: • The von Neumann algebra N has a nonzero amenable direct summand. • We have N ′ ∩ pM ωp ⊂ p(Bω ⋊ Γ)p. In this case, we further obtain A (cid:22)Bω ⋊Γ Bω for any finite von Neumann subalgebra with expectation A ⊂ N ′ ∩ pM ωp. We refer the reader to Section 2 for ultraproduct von Neumann algebras and Popa's intertwining techniques inside arbitrary von Neumann algebras. The proof of Theorem A given in Section 4 (see Theorems 4.1 and 4.2) uses a combination of Ozawa's C∗-algebraic techniques [Oz03, Oz04, Is12], ultraproduct von Neumann algebraic techniques [Oc85, AH12] and the recent generalization of Popa's intertwining-by-bimodules to the framework of type III von Neumann algebras developed by the authors in [HI15]. The interesting feature of the proof of Theorem A is that it does not rely on Connes -- Tomita -- Takesaki modular theory. Indeed, unlike other instances of Popa's spectral gap rigidity results in the literature which typically rely on using amenable traces and hence require the ambient von Neumann algebra to be (semi)finite, we BI-EXACT GROUPS, STRONGLY ERGODIC ACTIONS AND TYPE III FACTORS 3 use instead unital completely positive (ucp) maps and exploit Ozawa's C∗-algebraic techniques [Oz03, Oz04] to prove the existence of norm one projections. The main advantage of this approach is that it allows us to work directly inside the (possibly type III) crossed product von Neumann algebra M = B ⋊ Γ without appealing to the continuous core decomposition. In this respect, our approach is similar to the one we developed in our previous paper [HI15]. We refer the reader to [HR14, HU15, HV12, Is12, Is13] for other structural/rigidity results for type III factors involving the continuous core decomposition. Following [HR14, Oz04], we say that a von Neumann algebra M is ω-semisolid if for any von Neumann subalgebra N ⊂ M with expectation such that the relative commutant N ′ ∩ M ω has no type I direct summand, we have that N is amenable. The next corollary strengthens the indecomposability properties of crossed product von Neumann algebras B ⋊ Γ arising from arbitrary actions Γ y B of bi-exact discrete groups on abelian von Neumann algebras (see [Oz04, HV12, Is12] for previous results). Corollary B. Let Γ be any bi-exact discrete group, B any abelian σ-finite von Neumann algebra and Γ y B any action. Let ω ∈ β(N) \ N be any nonprincipal ultrafilter. Then the crossed product von Neumann algebra B ⋊ Γ is ω-semisolid. In particular, if B ⋊ Γ is a nonamenable factor, then B ⋊ Γ is prime, that is, B ⋊ Γ cannot be written as a tensor product Q1 ⊗ Q2 of diffuse factors. Our second main result, Theorem C below, is an answer to the general problem mentioned earlier in the case when the acting group is bi-exact. Indeed, using Theorem A in the case when the action Γ y B arises from a strongly ergodic free nonsingular action Γ y (X, µ) of a bi-exact countable discrete group on a standard probability space, we show that the group measure space factor L∞(X) ⋊ Γ is full. Theorem C. Let Γ be any bi-exact countable discrete group and Γ y (X, µ) any strongly ergodic free nonsingular action on a standard probability space. Then the group measure space factor L∞(X) ⋊ Γ is full. The proof of Theorem C uses a combination of Theorem A and the useful Lemma 5.1 below which proves the existence of a nontrivial centralizing sequence (un)n in every nonfull factor M = L(R) arising from a strongly ergodic nonsingular equivalence relation R defined on a standard probability space such that (un)n "does not embed" into the Cartan subalgebra L∞(X). Our Lemma 5.1 is a nonsingular generalization of a recent result of Hoff (see the first part of the proof of [Ho15, Proposition C]). In view of Choda's result [Ch81], we do not know whether Theorem C holds true more generally for (arbitrary strongly ergodic free nonsingular actions of) arbitrary non-inner amenable groups instead of bi-exact groups. We point out that Ozawa recently showed in [Oz16] that Theorem C holds true for arbitrary strongly ergodic free nonsingular actions of SL3(Z), which is not bi-exact by [Sa09]. We finally exploit recent results of Boutonnet -- Ioana -- Salehi Golsefidy [BISG15] to construct, for every 0 < λ ≤ 1, examples of type IIIλ strongly ergodic free nonsingular actions of a free group on a standard probability space. It is shown in [BISG15, Theorem A] that for any (not necessarily compact) connected simple Lie group and any countable dense subgroup Λ < G with "algebraic entries" (e.g. (Λ < G) = (SLn(Q) < SLn(R)) for n ≥ 2), the left translation action Λ y G is strongly ergodic. By taking a suitable non-unimodular closed subgroup P < G, the quotient action Λ y G/P is still strongly ergodic and of type III. The quotient action Λ y G/P need not be essentially free in general. However, using a "direct product" construction similar to the one used in [HV12, Corollary B], we can then construct strongly ergodic essentially free nonsingular actions of free groups and we obtain the following corollary. 4 CYRIL HOUDAYER AND YUSUKE ISONO Corollary D. For every 0 < λ ≤ 1, there exists a strongly ergodic free nonsingular action F∞ y (Xλ, µλ) of type IIIλ so that the group measure space factor L∞(Xλ, µλ) ⋊ F∞ is of type IIIλ and is full. Moreover, there exists a strongly ergodic free nonsingular action F∞ y (X∞, µ∞) of type II∞ so that the group measure space factor L∞(X∞, µ∞) ⋊ F∞ is of type II∞ and is full. The first examples of full factors of type III were discovered by Connes in [Co74]. He showed that the factors Mn,k,ϕ =(cid:18)OFn (Mk(C), ϕ)(cid:19) ⋊ Fn arising from Connes -- Størmer Bernoulli shifts of free groups Fn y NFn(Mk(C), ϕ) are full if n, k ≥ 2 and of type III if ϕ is not tracial. Observe that the factors Mn,k,ϕ possessNFn Ck as a Cartan subalgebra. This implies that the underlying ergodic nonsingular equivalence relation is strongly ergodic [FM75]. However, Connes -- Størmer Bernoulli crossed products need not be ∗-isomorphic to group measure space factors. In this respect, Corollary D above provides the first class of group measure space type III factors with no nontrivial central sequence. We finally point out that the group measure space type III factors in Corollary D possess a unique Cartan subalgebra, up to unitary conjugacy, by [HV12, Theorem A] (see [PV11] for the trace preserving case). Moreover, Corollary D provides new examples of group measure type III factors with a unique Cartan subalgebra, up to unitary conjugacy. Indeed, the examples of ergodic free nonsingular actions considered in [HV12, Corollary B] are not strongly ergodic since they have an amenable action as a quotient. In particular, the group measure space type III factors in [HV12, Corollary B] are not full while the ones in Corollary D are full. Acknowledgments. It is our pleasure to thank Adrian Ioana, Dimitri Shlyakhtenko, Yoshimichi Ueda and Stefaan Vaes for their valuable comments. Contents 1. Introduction and statement of the main results 2. Preliminaries 3. Bi-exactness and Ozawa's condition (AO) 4. Proofs of Theorem A and Corollary B 5. Proof of Theorem C 6. Group measure space type III factors with no central sequence 7. Further remarks References 1 4 8 13 15 17 19 21 2. Preliminaries For any von Neumann algebra M , we will denote by Z(M ) the centre of M , by U (M ) the group of unitaries in M and by (M, L2(M ), J M , PM ) a standard form for M . We will say that an inclusion of von Neumann algebras P ⊂ 1P M 1P is with expectation if there exists a faithful normal conditional expectation EP : 1P M 1P → P . BI-EXACT GROUPS, STRONGLY ERGODIC ACTIONS AND TYPE III FACTORS 5 Crossed product von Neumann algebras. We will use the following terminology and notation regarding crossed product von Neumann algebras. Let Γ be any discrete group, B any σ-finite von Neumann algebra and Γ y B any action. Denote by M := B ⋊ Γ the corresponding crossed product von Neumann algebra and by EB : M → B the canonical faithful normal conditional expectation given by EB(bλg) = δg,eb for all g ∈ Γ and all b ∈ B. Fix a standard form (B, L2(B), J B, PB) for B. Denote by u : Γ → U (L2(B)) the canonical unitary representation implementing the action Γ y B. A standard form (M, L2(M ), J M , PM ) for M is given by L2(M ) = L2(B) ⊗ ℓ2(Γ) and J M (ξ ⊗ δg) = u∗ gJ Bξ ⊗ δg−1 for all ξ ∈ L2(B) and all g ∈ Γ. The Jones projection eB : L2(M ) → L2(B) is then simply given by eB = 1 ⊗ PCδe where PCδe : ℓ2(Γ) → Cδe is the orthogonal projection onto Cδe. For crossed product von Neumann algebras M = B ⋊ Γ, we will always use such a standard form (M, L2(M ), J M , PM ) as defined above. Ultraproduct von Neumann algebras. Let M be any σ-finite von Neumann algebra and ω ∈ β(N) \ N any nonprincipal ultrafilter. Define Iω(M ) = {(xn)n ∈ ℓ∞(M ) xn → 0 ∗ -strongly as n → ω} Mω(M ) = {(xn)n ∈ ℓ∞(M ) (xn)n Iω(M ) ⊂ Iω(M ) and Iω(M ) (xn)n ⊂ Iω(M )} . The multiplier algebra Mω(M ) is a C∗-algebra and Iω(M ) ⊂ Mω(M ) is a norm closed two- sided ideal. Following [Oc85, §5.1], we define the ultraproduct von Neumann algebra M ω by M ω := Mω(M )/Iω(M ), which is indeed known to be a von Neumann algebra. We denote the image of (xn)n ∈ Mω(M ) by (xn)ω ∈ M ω. For every x ∈ M , the constant sequence (x)n lies in the multiplier algebra Mω(M ). We will then identify M with (M +Iω(M ))/Iω(M ) and regard M ⊂ M ω as a von Neumann subalgebra. The map Eω : M ω → M : (xn)ω 7→ σ-weak limn→ω xn is a faithful normal conditional expectation. For every faithful state ϕ ∈ M∗, the formula ϕω := ϕ ◦ Eω defines a faithful normal state on M ω. Observe that ϕω((xn)ω) = limn→ω ϕ(xn) for all (xn)ω ∈ M ω. Following [Co74, §2], we define Mω(M ) :=n(xn)n ∈ ℓ∞(M ) lim n→ω kxnϕ − ϕxnk = 0, ∀ϕ ∈ M∗o . for every faithful state ϕ ∈ M∗. We have Iω(M ) ⊂ Mω(M ) ⊂ Mω(M ). The asymptotic centralizer is defined by Mω := Mω(M )/Iω(M ). We have Mω ⊂ M ω. Moreover, by [Co74, Proposition 2.8] (see also [AH12, Proposition 4.35]), we have Mω = M ′ ∩ (M ω)ϕω Let Q ⊂ M be any von Neumann subalgebra with faithful normal conditional expectation EQ : M → Q. Choose a faithful state ϕ ∈ M∗ in such a way that ϕ = ϕ◦EQ. We have ℓ∞(Q) ⊂ ℓ∞(M ), Iω(Q) ⊂ Iω(M ) and Mω(Q) ⊂ Mω(M ). We will then identify Qω = Mω(Q)/Iω(Q) with (Mω(Q)+Iω(M ))/Iω(M ) and be able to regard Qω ⊂ M ω as a von Neumann subalgebra. Observe that the norm k · k(ϕQ)ω on Qω is the restriction of the norm k · kϕω to Qω. Observe moreover that (EQ(xn))n ∈ Iω(Q) for all (xn)n ∈ Iω(M ) and (EQ(xn))n ∈ Mω(Q) for all (xn)n ∈ Mω(M ). Therefore, the mapping EQω : M ω → Qω : (xn)ω 7→ (EQ(xn))ω is a well- defined conditional expectation satisfying ϕω ◦ EQω = ϕω. Hence, EQω : M ω → Qω is a faithful normal conditional expectation. For more on ultraproduct von Neumann algebras, we refer the reader to [AH12, Oc85]. We record the following observation that will be used throughout. Let Γ be any discrete group, B any σ-finite von Neumann algebra and Γ y B any action. Put M := B ⋊ Γ and denote by EB : M → B the canonical faithful normal conditional expectation. Choose any faithful state ϕ ∈ M∗ such that ϕ ◦ EB = ϕ. Then the von Neumann subalgebra Bω ∨ M ⊂ M ω is globally invariant under the modular automorphism group σϕω and hence is with expectation. 6 CYRIL HOUDAYER AND YUSUKE ISONO Observe that we have Bω ∨ M = Bω ⋊ Γ canonically. Therefore the von Neumann subalgebra Bω ⋊ Γ ⊂ M ω is with expectation. Denote by EBω : M ω → Bω and by EBω ⋊Γ : M ω → Bω ⋊ Γ the unique ϕω-preserving conditional expectations. By uniqueness of the ϕω-preserving conditional expectation EBω : M ω → Bω, we have EBω ◦ EBω ⋊Γ = EBω . We thank Hiroshi Ando for pointing out to us the following well-known result. Lemma 2.1. Let M be any σ-finite von Neumann algebra and ω ∈ β(N) \ N any nonprincipal ultrafilter. For any u ∈ U (M ω), there exists a sequence (un)n ∈ Mω(M ) such that u = (un)ω and un ∈ U (M ) for every n ∈ N. Proof. Denote by f : T → (−π, π] the unique Borel function such that exp(if (z)) = z for all z ∈ T. Let u ∈ U (M ω) and put h = f (u) ∈ M ω. Then h∗ = h and exp(ih) = u. Write h = (hn)ω for some (hn)n ∈ Mω(M ). Since h∗ = h, up to replacing each hn by 1 n), we may assume that h∗ n = hn for every n ∈ N. Put un = exp(ihn) ∈ U (M ) for every n ∈ N. Since [−π, π] → T : t 7→ exp(it) is a continuous function, it is a uniform limit of polynomial functions by Stone -- Weierstrass theorem. It follows that (un)n = (exp(ihn))n = exp(i(hn)n) ∈ Mω(M ) and u = exp(ih) = exp(i(hn)ω) = (exp(ihn))ω = (un)ω. (cid:3) 2 (hn + h∗ We next recall the construction of the Groh -- Raynaud ultraproduct. For any Hilbert space H, define the ultraproduct Hilbert space Hω as the completion/separation of ℓ∞(H) with respect to the semi-inner product given by h(ξn)n, (ηn)ni := limn→ωhξn, ηniH for all (ξn)n, (ηn)n ∈ ℓ∞(H). We denote the image of (ξn)n ∈ ℓ∞(H) by (ξn)ω ∈ Hω. Let M ⊂ B(H) be any von Neumann algebra. We define the unital ∗-representation πω : ℓ∞(M ) → B(Hω) by πω((xn)n)(ξn)ω = (xnξn)ω for all (xn)n ∈ ℓ∞(B(H)) and all (ξn)n ∈ ℓ∞(H). Let (M, L2(M ), J M , PM ) be a standard form for M . The Groh -- Raynaud ultraproduct N := Qω M is the von Neumann algebra generated by πω(ℓ∞(M )). It is known that the inclusion N ⊂ B(L2(M )ω) is in standard form with modular conjugation given by J N (ξn)ω := (J M ξn)ω for all (ξn)ω ∈ L2(M )ω (see [Ra99, Corollary 3.9] and [AH12, Theorem 3.18]). By [AH12, Theorem 3.7], the Ocneanu ultraproduct is ∗-isomorphic to a corner of the Groh -- Raynaud ultraproduct. More precisely, for any faithful state ϕ ∈ M∗, denote by ξϕ ∈ PM the canonical representing vector. Then the isometry given by wϕ : L2(M ω) → L2(M )ω : (xn)ωξϕω 7→ (xnξϕ)ω satisfies w∗ ϕN wϕ = M ω. Define the ultraproduct state ϕω = h · (ξϕ)ω, (ξϕ)ωi ∈ N∗ and de- note by p ∈ N the support projection of ϕω ∈ N∗. We have wϕw∗ ϕ = pJ N pJ N . Then the ϕN wϕ = M ω implies that pJ N pJ N N pJ N pJ N ∼= pN p ∼= M ω so that the stan- condition w∗ dard representation of M ω is given by L2(M ω) = pJ N pJ N L2(M )ω with modular conjugation J M ω = pJ N p. Lemma 2.2. Let B ⊂ M be any inclusion of σ-finite von Neumann algebras with faithful nor- mal conditional expectation EB : M → B. Denote by eB : L2(M ) → L2(B) the corresponding Jones projection. Denote by N :=Qω M the Groh -- Raynaud ultraproduct. Let ϕ ∈ M∗ be any faithful state such that ϕ ◦ EB = ϕ and denote by p ∈ N the support projection of ϕω ∈ N∗. Then πω((eB)n) commutes with p and J N . Proof. Since eB commutes with J M , πω((eB)n) commutes with J N . Denote by ξϕ ∈ PM the canonical vector representing ϕ ∈ M∗. Since eBM eB = BeB and eBξϕ = ξϕ, we have πω((eB)n)J N πω(ℓ∞(M ))(ξϕ)ω = J N πω((eB)n)πω(ℓ∞(M ))πω((eB)n)(ξϕ)ω = J N πω(ℓ∞(B))πω((eB)n)(ξϕ)ω ⊂ J N πω(ℓ∞(M ))(ξϕ)ω. BI-EXACT GROUPS, STRONGLY ERGODIC ACTIONS AND TYPE III FACTORS 7 Since p is the projection onto the closure of J N πω(ℓ∞(M ))(ξϕ)ω, we obtain that πω((eB)n) commutes with p. (cid:3) Equivalence relations and von Neumann algebras. Definition 2.3 ([FM75]). Let (X, µ) be any standard probability space. A nonsingular equiv- alence relation R defined on (X, µ) is an equivalence relation R ⊂ X × X which satisfies the following three conditions: (i) R ⊂ X × X is a Borel subset, (ii) R has countable classes and (iii) for every ϕ ∈ [R], we have [ϕ∗µ] = [µ] where [R] denotes the full group of R consisting in all the Borel automorphisms ϕ : X → X such that gr(ϕ) ⊂ R. Following [FM75], to any nonsingular equivalence relation R defined on a standard probability space (X, µ), one can associate a von Neumann algebra M = L(R) which contains A = L∞(X) as a Cartan subalgebra, that is, A ⊂ M is maximal abelian with expectation and the group of normalizing unitaries NM (A) = {u ∈ U (M ) : uAu∗} generates M as a von Neumann algebra. When Γ y (X, µ) is a nonsingular Borel action of a countable discrete group on a standard probability space, we will denote by R(Γ y X) the nonsingular orbit equivalence relation defined by R(Γ y X) = {(x, γx) γ ∈ Γ, x ∈ X}. When the action Γ y (X, µ) is moreover essentially free, there is a canonical isomorphism of pairs of von Neumann algebras (L∞(X) ⊂ L(R(Γ y X))) ∼= (L∞(X) ⊂ L∞(X) ⋊ Γ) . For more information on nonsingular equivalence relations and their von Neumann algebras, we refer the reader to [FM75]. Strongly ergodic actions and full factors. We first recall the concept of strong ergodicity for group actions and equivalence relations. Definition 2.4. Let (X, µ) be any standard probability space. (i) Let Γ be any countable discrete group and Γ y (X, µ) any ergodic nonsingular ac- tion. The action Γ y (X, µ) is said to be strongly ergodic if for any sequence (Cn)n of measurable subsets of X such that limn µ(γCn△Cn) = 0 for all γ ∈ Γ, we have limn µ(Cn)(1 − µ(Cn)) = 0. (ii) Let R be any ergodic nonsingular equivalence relation defined on (X, µ). The equiva- lence relation R is said to be strongly ergodic if for any sequence (Cn)n of measurable subsets of X such that limn µ(gCn△Cn) = 0 for all g ∈ [R], we have limn µ(Cn)(1 − µ(Cn)) = 0. Put A = L∞(X) and fix any nonprincipal ultrafilter ω ∈ β(N)\N. Then the ergodic nonsingular action Γ y (X, µ) is strongly ergodic if and only if the ultraproduct action Γ y Aω defined by γ · (an)ω = (γ · an)ω is ergodic, that is, (Aω)Γ = C1. Likewise, the ergodic nonsingular equivalence relation R defined on (X, µ) is strongly ergodic if and only if L(R)′ ∩ Aω = C1. We also have that the nonsingular action Γ y (X, µ) is strongly ergodic if and only if the nonsingular orbit equivalence relation R(Γ y X) is strongly ergodic. Following [Co74], we say that a factor M with separable predual is full if Mω = C1 for some (or any) nonprincipal ultrafilter ω ∈ β(N) \ N. By [AH12, Theorem 5.2], M is full if and only if M ′ ∩ M ω = C1 for some (or any) nonprincipal ultrafilter ω ∈ β(N) \ N. Observe that for any ergodic nonsingular equivalence relation R defined on a standard probability space (X, µ), if L(R) is full then R is strongly ergodic. 8 CYRIL HOUDAYER AND YUSUKE ISONO Connes proved in [Co74, Theorem 2.12] that factors of type III0 are never full. Ueda showed in [Ue00, Corollary 11] that ergodic nonsingular equivalence relations of type III0 are never strongly ergodic. We give a short proof of Ueda's result. We refer to [Co72] for the type classification of factors. Proposition 2.5 ([Ue00, Corollary 11]). Let R be any ergodic nonsingular equivalence relation defined on a standard probability space (X, µ). If R is of type III0, then R is not strongly ergodic. Proof. Put A = L∞(X) and M = L(R). Assume that R is of type III0. Then M is of type III0. Fix a nonprincipal ultrafilter ω on N. Then Z(M ω) 6= C1 by [AH12, Theorem 6.18]. Since Aω is maximal abelian in M ω by [Po95, Theorem A.1.2], we have C1 6= Z(M ω) = (M ω)′ ∩ M ω = (M ω)′ ∩ Aω ⊂ M ′ ∩ Aω. Since M ′ ∩ Aω 6= C1, R is not strongly ergodic. (cid:3) Popa's intertwining-by-bimodules. In this subsection, we briefly recall Popa's intertwining- by-bimodules [Po01, Po03]. In the present paper, we will need a generalization of Popa's intertwining-by-bimodules to the framework of type III von Neumann algebras developed by the authors in [HI15]. We will use the following terminology (see [HI15, Definition 4.1]). Definition 2.6. Let M be any σ-finite von Neumann algebra, 1A and 1B any nonzero projec- tions in M , A ⊂ 1AM 1A and B ⊂ 1BM 1B any von Neumann subalgebras with faithful normal conditional expectations EA : 1AM 1A → A and EB : 1BM 1B → B respectively. We say that A embeds with expectation into B inside M and write A (cid:22)M B if there exist projections e ∈ A and f ∈ B, a nonzero partial isometry v ∈ eM f and a unital normal ∗- homomorphism θ : eAe → f Bf such that the inclusion θ(eAe) ⊂ f Bf is with expectation and av = vθ(a) for all a ∈ eAe. The main characterization of intertwining subalgebras we will use in this paper is the following result proven in [HI15, Theorem 4.3]. Theorem 2.7. Keep the same notation as in Definition 2.6 and moreover assume that A is finite. Then the following conditions are equivalent. (1) A (cid:22)M B. (2) There exists no net (wi)i∈I of unitaries in U (A) such that EB(b∗wia) → 0 in the σ- strong topology for all a, b ∈ 1AM 1B. 3. Bi-exactness and Ozawa's condition (AO) Bi-exactness for discrete groups. Recall from [Oz03] that a von Neumann algebra M ⊂ B(H) satisfies condition (AO) if there exist unital σ-weakly dense C∗-subalgebras A ⊂ M and B ⊂ M ′ such that A is locally reflexive and the map ν : A ⊗alg B −→ B(H)/K(H) : a ⊗ b 7→ ab + K(H) is continuous with respect to the minimal tensor norm. Recall that A is locally reflexive or equivalently has property C ′′ (see e.g. [BO08, Section 9]) if for any C∗-algebra C, the inclusion map A∗∗ ⊗alg C ֒→ (A ⊗min C)∗∗ is continuous with respect to the minimal tensor norm. In this case, any ∗-homomorphism π : A ⊗min C → B(K) has an extension eπ : A∗∗ ⊗min C → B(K) which is normal on A∗∗ ⊗ C1 (since π always has a canonical extension on (A ⊗min C)∗∗). We next recall the notion of bi-exactness for discrete groups which was introduced by Ozawa in [Oz04] (using the terminology class S) and intensively studied in [BO08, Chapter 15]. Our definition is different from the original one, but it is equivalent to it and it is moreover adapted to the framework of discrete quantum groups [Is13, Definition 3.2]. BI-EXACT GROUPS, STRONGLY ERGODIC ACTIONS AND TYPE III FACTORS 9 Definition 3.1 ([BO08, Proposition 15.2.3(2)]). Let Γ be any discrete group. We say that Γ is bi-exact if there exists a (Γ × Γ)-globally invariant unital C∗-subalgebra B ⊂ ℓ∞(Γ) such that the following two conditions are satisfied: (i) The algebra B contains c0(Γ) so that the quotient B∞ := B/c0(Γ) is well-defined. (ii) The left translation action Γ y ℓ∞(Γ) induces an amenable action Γ y B∞ and the right translation action ℓ∞(Γ) x Γ induces the trivial action on B∞. The class of bi-exact discrete groups includes amenable groups, free groups [AO74], discrete subgroups of simple connected Lie groups of real rank one [Sk88] and Gromov word-hyperbolic groups [Oz03]. Observe that for any bi-exact discrete group Γ, the group von Neumann algebra L(Γ) satisfies condition (AO). We refer the reader to [BO08, Chapter 15] for more information on bi-exact discrete groups. Ozawa's condition (AO) in crossed product von Neumann algebras. In this subsec- tion, we prove a relative version of Ozawa's condition (AO) in the framework of crossed product von Neumann algebras. This result will be used in the proof of Theorem A. Let Γ be any discrete group, B ⊂ B any inclusion of σ-finite von Neumann algebras and Γ y B any action that leaves globally invariant the subalgebra B. Denote by M := B ⋊ Γ and M = B ⋊ Γ the corresponding crossed product von Neumann algebras, by EB : M → B the canonical faithful normal conditional expectation and by eB : L2(M) → L2(B) the corresponding Jones projection. We use the notation and terminology of Section 2 for the standard forms (B, L2(B), J B, PB) of B and (M, L2(M), J M, PM) of M = B ⋊ Γ. We define a nondegenerate (and possibly nonunital) C∗-algebra and its multiplier C∗-algebra inside B(L2(M)) by KB := C∗(cid:8)aJ MxJ MeBbJ MyJ M a, b, x, y ∈ B ⋊red Γ(cid:9) ⊂ B(L2(M)) M(KB) :=(cid:8)T ∈ B(L2(M)) T KB ⊂ KB and KBT ⊂ KB(cid:9) . where C∗ {Y} ⊂ B(L2(M)) denotes the C∗-subalgebra of B(L2(M)) generated by the subset Y ⊂ B(L2(M)). We record the following elementary lemma. Lemma 3.2. We have KB ⊂ B(L2(B)) ⊗min K(ℓ2(Γ)). Proof. Denote by σ : Γ y B the action and by u : Γ → U (L2(B)) the canonical unitary representation implementing the action σ. Recall that L2(M) = L2(B) ⊗ ℓ2(Γ). Regard M = B ⋊ Γ as generated by πσ(b) = Ph∈Γ σh−1(b) ⊗ PCδh for b ∈ B and 1 ⊗ λg for g ∈ Γ where PCδh : ℓ2(Γ) → Cδh is the orthogonal projection onto Cδh. We have J M(1 ⊗ λg)J M = ug ⊗ ρg for all g ∈ Γ. Let C ⊂ B(L2(B)) be the C∗-algebra generated by B, J BBJ B and ug for all g ∈ Γ. We will show that K = C ⊗min K(ℓ2(Γ)). Recall that eB = 1 ⊗ PCδe. For all g, h ∈ Γ, denote by eg,h : Cδh → Cδg the partial isometry sending δh onto δg. For all a, b ∈ B and all g, h, s, t ∈ Γ, we have πσ(a)(J BbJ B ⊗ 1)eB = eBπσ(a)(J BbJ B ⊗ 1) = aJ BbJ B ⊗ PCδe (1 ⊗ λg)(us ⊗ ρs)eB(1 ⊗ λh)(ut ⊗ ρt) = ust ⊗ λgρsPCδeλhρt = ust ⊗ egs−1,h−1t. We then have KB = C∗(cid:8)aJ MxJ MeBbJ MyJ M a, b, x, y ∈ B ⋊red Γ(cid:9) = C∗(cid:8)aJ BbJ Bug ⊗ es,t a, b ∈ B, g, s, t ∈ Γ(cid:9) = C ⊗min K(ℓ2(Γ)). This finishes the proof of Lemma 3.2. (cid:3) 10 CYRIL HOUDAYER AND YUSUKE ISONO Consider now the following unital ∗-homomorphism: νB : (B ⋊red Γ) ⊗alg J M(B ⋊red Γ)J M → M(KB)/KB : a ⊗ J MbJ M 7→ a J MbJ M + KB. Ozawa proved in [Oz04, Proposition 4.2] that when Γ is bi-exact and B = B is finite and amenable, the map νB is continuous with respect to the minimal tensor norm. This is nothing but a relative version of the condition (AO) in the framework of crossed product von Neumann algebras. Observe that when B = B = C1, continuity of νB with respect to the minimal tensor norm implies that L(Γ) satisfies condition (AO). Since KB is the smallest C∗-algebra containing 1 ⊗ c0(Γ) and such that its multiplier algebra contains B ⋊red Γ and J M(B ⋊red Γ)J M, we can easily generalize [Oz04, Proposition 4.2] as follows. Proposition 3.3. Keep the same setting as above and assume that Γ is bi-exact and B is amenable. Then the map νB : (B ⋊red Γ) ⊗alg J M(B ⋊red Γ)J M → M(KB)/KB : a ⊗ J MbJ M 7→ a J MbJ M + KB is well-defined and continuous with respect to the minimal tensor norm. Proof. As in the proof of Lemma 3.2, regard M = B ⋊ Γ as generated by πσ(B) and (1 ⊗ λ)(Γ). Since B is amenable (i.e. semidiscrete), the map πσ(B) ⊗alg J Mπσ(B)J M → B(L2(M)) : πσ(a) ⊗ J Mπσ(b)J M 7→ πσ(a) J Mπσ(b)J M is continuous with respect to the minimal tensor norm. Then the proof of [Oz04, Proposition 4.2] applies mutatis mutandis to show that the map νB is continuous with respect to the minimal tensor norm. (cid:3) We will apply Proposition 3.3 in Theorem 4.1 (in the case when B = B) and in Theorem 4.2 (in the general case). Ozawa's condition (AO) in ultraproduct von Neumann algebras. In this subsection, we prove a version of Ozawa's condition (AO) in the framework of ultraproduct von Neumann algebras. Although we will not use this result in this paper, we nevertheless mention it since we believe it is interesting in its own right. We keep the same notation as in the previous subsection and we moreover assume that B = B. Let ω ∈ β(N) \ N be any nonprincipal ultrafilter. Denote by (M ω, L2(M ω), J M ω , PM ω ) a standard form for M ω and by eBω : L2(M ω) → L2(Bω) the Jones projection corresponding to the inclusion Bω ⊂ M ω. We define a (possibly degenerate and nonunital) C∗-subalgebra Kω and its multiplier algebra M(Kω) inside B(L2(M ω)) by Recall from Proposition 3.3 (in the case when B = B with K := KB which is exactly [Oz04, Proposition 4.2]) that the map ν : (B ⋊red Γ) ⊗alg J M (B ⋊red Γ)J M → M(K)/K : a ⊗ J M bJ M 7→ a J M bJ M + K is continuous with respect to the minimal tensor norm. We now state a version of Ozawa's condition (AO) in the ultraproduct representation L2(M ω). Proposition 3.4. Keep the same setting as above and assume that Γ is bi-exact and B is amenable. Then the map νω : (B ⋊red Γ) ⊗alg J M ω is well-defined and continuous with respect to the minimal tensor norm. → M(Kω)/Kω : a ⊗ J M ω (B ⋊red Γ)J M ω 7→ a J M ω bJ M ω bJ M ω + Kω Kω := C∗(cid:8)aJ M ω xJ M ω eBω bJ M ω yJ M ω M(Kω) :=(cid:8)T ∈ B(L2(M ω)) T Kω ⊂ Kω and KωT ⊂ Kω(cid:9) . a, b, x, y ∈ B ⋊red Γ(cid:9) ⊂ B(L2(M ω)) BI-EXACT GROUPS, STRONGLY ERGODIC ACTIONS AND TYPE III FACTORS 11 Proof. Put C := C∗(cid:8)B ⋊red Γ, J M (B ⋊red Γ)J M(cid:9) ⊂ B(L2(M )) Cω := C∗(cid:8)B ⋊red Γ, J M ω (B ⋊red Γ)J M ω(cid:9) ⊂ B(L2(M ω)). Observe that C + K (resp. Cω + Kω) is a C∗-algebra since it is the sum of a C∗-subalgebra and an ideal in M(K) (resp. M(Kω)). Claim. There is a ∗-homomorphism θ : C + K → B(L2(M ω)) such that θ(x) = x and θ(J M yJ M ) = J M ω yJ M ω for all x, y ∈ B ⋊red Γ and θ(eB) = eBω . Proof of the Claim. Indeed, fix any faithful state ϕ ∈ M∗ such that ϕ ◦ EB = ϕ and denote by p the support projection in N = Qω M of the ultraproduct state ϕω ∈ N∗. By Lemma 2.2, πω((eB)n) commutes with p and J N and hence πω((eB)n) commutes with ep := pJ N pJ N . Since ep commutes with πω(M ) and πω(J M M J M ) = J N πω(M )J N , ep commutes with πω(C + K). Recall that epNep ∼= pN p ∼= M ω and epL2(M )ω = L2(M ω). Then the ∗-homomorphism θ : C + K → B(L2(M ω)) : T 7→ epπω(T )ep satisfies all the conditions of the Claim. (cid:3) Since θ(C) = Cω and θ(K) = Kω, θ induces a ∗-homomorphism Denote by ι : (B ⋊red Γ) ⊗min J M ω (B ⋊red Γ)J M ω → (B ⋊red Γ) ⊗min J M (B ⋊red Γ)J M the tautological ∗-isomorphism. Then the composition map eθ : (C + K)/K → (Cω + Kω)/Kω ⊂ M(Kω)/Kω. νω = eθ ◦ ν ◦ ι : (B ⋊red Γ) ⊗alg J M ω is continuous with respect to the minimal tensor norm. (B ⋊red Γ)J M ω → M(Kω)/Kω (cid:3) Weak exactness for C∗-algebras. To obtain structural results for von Neumann algebras M satisfying Ozawa's (relative) condition (AO) [Oz03, Oz04], it is usually necessary to impose local reflexivity or exactness of the given unital σ-weakly dense C∗-algebra in M . Observe that in the setting of Proposition 3.3, the reduced crossed product C∗-algebra B ⋊red Γ need not be locally reflexive since it contains the von Neumann algebra B. To avoid this difficulty, Ozawa assumed in [Oz04, Theorem 4.6] that Γ is exact and B is abelian so that B ⋊red Γ is exact and hence locally reflexive. In [Is12], the second named author introduced a notion of weak exactness for C∗-algebras and could settle this problem. Namely, he generalized [Oz04, Theorem 4.6] under the assumptions that Γ is exact and B is amenable (not necessarily abelian). The main idea behind this generalization was to use some exactness (or equivalently property C ′) of the opposite algebra (B ⋊red Γ)op, instead of local reflexivity of B ⋊red Γ. In the present paper, to study more general cases, we will make use of this notion of weak exactness for C∗-algebras. Recall from [Is12, Theorem 3.1.3(1)(ii)] that for an inclusion of a unital C∗-algebra A ⊂ M in a von Neumann algebra M , we say that A is weakly exact in M if for any unital C∗-algebra C, any ∗-homomorphism π : A ⊗min C → B(K) which is σ-weakly continuous on A ⊗ C1 has an extension eπ : A ⊗min C ∗∗ → B(K) which is normal on C1 ⊗ C ∗∗. In the case when A = M , we simply say that M is weakly exact. Here we recall the following fundamental fact. Proposition 3.5 ([Is12, Proposition 4.1.7]). Let Γ be any exact discrete group, B any σ-finite amenable (and hence weakly exact) von Neumann algebra and Γ y B any action. Then the reduced crossed product C∗-algebra B ⋊red Γ is weakly exact in B ⋊Γ. If moreover Γ is countable and B has separable predual, then B ⋊ Γ is weakly exact. 12 CYRIL HOUDAYER AND YUSUKE ISONO Using this property, we prove an important lemma, which is a variant of [Oz03, Lemma 5] (see also [BO08, Proposition 15.1.6] and [Is12, Lemma 5.1.1]). The proof is essentially the same as the one of [BO08, Proposition 15.1.6] but we nevertheless include it for the reader's convenience. Lemma 3.6. Let M ⊂ M be any inclusion of σ-finite von Neumann algebras with expectation and (M, L2(M), J M, PM) a standard form for M. Let C ⊂ M be any unital σ-weakly dense C∗-subalgebra, p ∈ M any nonzero projection and ϕ : M → pMp any normal ucp map. We will use the identification pB(L2(M))p = B(pL2(M)). Assume that the following two conditions hold: • The map Φ : C ⊗alg J MCJ M → B(pL2(M)) : nXi=1 xi ⊗ J MyiJ M 7→ nXi=1 ϕ(xi) J MyiJ Mp is continuous with respect to the minimal tensor norm. • The C∗-algebra C is locally reflexive or C is weakly exact in M . Then the ucp map ϕ : M → pMp has a ucp extension eϕ : B(L2(M)) → (J MCJ Mp)′ ∩ B(pL2(M)). Proof. To simplify the notation, we will write J = J M. Observe that Φ = ν ◦ (ϕ ⊗ idJCJ ) where ν : pMp ⊗alg JCJ → B(pL2(M)) is the multiplication map. We first prove the following result. Claim. The ucp map Φ : C ⊗min JCJ → B(pL2(M)) can be extended to a ucp map eΦ : M ⊗min JCJ → B(pL2(M)) which is normal M ⊗ C1. In particular, we have eΦ(x ⊗ 1) = ϕ(x) for all x ∈ M . Proof of the Claim. Indeed, let (π, V, K) be a minimal Stinespring dilation for Φ : C ⊗min JCJ → B(pL2(M)), that is, π : C ⊗min JCJ → B(K) is a unital ∗-representation and V : pL2(M) → K is an isometry such that the subspace π(C ⊗min JCJ)V pL2(M) is dense in K and Φ(x) = V ∗π(x)V for all x ∈ C. By minimality of (π, V, K) and since Φ is σ-weakly continuous on C ⊗ C1 (resp. C1 ⊗ JCJ), we have that π is also σ-weakly continuous on C ⊗ C1 (resp. C1 ⊗ JCJ). Indeed, it suffices to notice that for all c1, c2 ∈ C, all x, y ∈ C ⊗min JCJ and all ξ, η ∈ pL2(M), we have hπ(c1 ⊗ Jc2J) π(x)V ξ, π(y)V ηiK = hV ∗π(y∗(c1 ⊗ Jc2J)x)V ξ, ηiK = hΦ(y∗(c1 ⊗ Jc2J)x)ξ, ηiK . Since C is assumed to be locally reflexive or weakly exact in M (which is equivalent to saying that JCJ is weakly exact in JM J), the unital ∗-homomorphism π : C ⊗min JCJ → B(K) always has an extension eπ : C ∗∗ ⊗min JCJ → B(K) which is normal on C ∗∗ ⊗ C1. Observe that we do not need σ-weak continuity on C1 ⊗ JCJ when C is locally reflexive. Let z ∈ C ∗∗ be the central projection such that zC ∗∗ = M canonically and let zi ∈ C be a bounded net converging to z in the σ-weak topology in C ∗∗. Observe that zi → 1M σ-weakly We then have that in M and recall that eπ (resp. π) is σ-weakly continuous on C ∗∗ ⊗ C1 (resp. C ⊗ C1 ⊂ M ⊗ C1). and hence eπ((1 − z) ⊗ 1) = 0. Since eπ is a ∗-homomorphism, it satisfies eπ((z ⊗ 1)x) = eπ(x) for all x ∈ C ∗∗ ⊗min JCJ. In particular, we have eπ((z ⊗ 1) · )C⊗minJCJ = π. Using moreover eπ((z ⊗ 1) · ) : M ⊗min JCJ → B(K) is an extension of π : C ⊗min JCJ → B(K) which is normal the identification M ⊗min JCJ = zC ∗∗ ⊗min JCJ, we obtain that the unital ∗-homomorphism eπ(z ⊗ 1) = lim π(zi ⊗ 1) = π(1M ⊗ 1) = 1 i BI-EXACT GROUPS, STRONGLY ERGODIC ACTIONS AND TYPE III FACTORS 13 is an extension of Φ : C ⊗min JCJ → B(pL2(M)) which is normal on M ⊗ C1. In particular, (cid:3) on M ⊗ C1. Therefore the ucp map eΦ = Ad(V ∗) ◦eπ((z ⊗ 1) · ) : M ⊗min JCJ → B(pL2(M)) we have eΦ(x ⊗ 1) = ϕ(x) for all x ∈ M . We next apply Arveson's extension theorem to the ucp map eΦ : M ⊗min JCJ → B(pL2(M)) and we obtain a ucp extension map that we still denote by eΦ : B(L2(M)) ⊗min JCJ → B(pL2(M)). Since eΦC1⊗JCJ : C1 ⊗ JCJ → B(pL2(M)) : 1 ⊗ JxJ 7→ JxJp is a unital ∗-homomorphism, C1 ⊗ JCJ is contained in the multiplicative domain of eΦ (see e.g. [BO08, and hence Φ(B(L2(M)) ⊗ 1) ⊂ (JCJp)′ ∩ B(pL2(M)). Thus, eϕ := Φ( · ⊗ 1) : B(L2(M)) → Φ(x ⊗ 1) JuJp = Φ(x ⊗ 1)Φ(1 ⊗ JuJ) = Φ(x ⊗ JuJ) = Φ(1 ⊗ JuJ)Φ(x ⊗ 1) = JuJp Φ(x ⊗ 1) Section 1.5]). Therefore, for all u ∈ U (C) and all x ∈ B(L2(M)), we have (JCJp)′ ∩ B(pL2(M)) is the desired ucp extension map. (cid:3) 4. Proofs of Theorem A and Corollary B We first prove two intermediate results, namely Theorems 4.1 and 4.2, from which we will deduce Theorem A. While these two results are independent from each other, their proofs are in fact very similar and use Ozawa's condition (AO) for crossed product von Neumann algebras from Section 3. Theorem 4.1 below is a spectral gap rigidity result for subalgebras with expectation N ⊂ M of crossed product von Neumann algebras M = B ⋊ Γ arising from arbitrary actions of bi-exact discrete groups on amenable von Neumann algebras. Theorem 4.1. Let Γ be any bi-exact discrete group, B any amenable σ-finite von Neumann algebra and Γ y B any action. Put M := B ⋊ Γ. Let p ∈ M be any nonzero projection and N ⊂ pM p any von Neumann subalgebra with expectation. Let ω ∈ β(N)\N be any nonprincipal ultrafilter. Then at least one of the following conditions holds true. • The von Neumann algebra N has a nonzero amenable direct summand. • We have N ′ ∩ pM ωp ⊂ p(Bω ⋊ Γ)p. Proof. Assume that N ′∩pM ωp 6⊂ p(Bω ⋊Γ)p. Let Y ∈ N ′∩pM ωp be such that Y /∈ p(Bω ⋊Γ)p. Up to replacing Y by Y −EBω ⋊Γ(Y ) 6= 0 which still lies in N ′ ∩pM ωp, we may assume that Y ∈ N ′ ∩ pM ωp, Y 6= 0 and EBω ⋊Γ(Y ) = 0. Put y = Eω(Y ∗Y ) ∈ (N ′ ∩ pM p)+. Define the nonzero 2 kyk∞,kyk∞](y) ∈ N ′ ∩ pM p and put c := (yp0)−1/2 ∈ N ′ ∩ pM p. We spectral projection p0 := 1[ 1 have Eω((Y c)∗(Y c)) = Eω(c Y ∗Y c) = c Eω(Y ∗Y ) c = c y c = p0. Up to replacing Y by Y c which still lies in N ′ ∩ pM ωp, we may assume that Y ∈ N ′ ∩ pM ωp, Y 6= 0, EBω ⋊Γ(Y ) = 0 and Eω(Y ∗Y ) = p0. Write Y = (yn)ω for some (yn)n ∈ Mω(M ). Observe that σ-weak limn→ω y∗ Denote by (M, L2(M ), J M , PM ) a standard form for M = B ⋊ Γ as in Section 2. To further simplify the notation, we will write J = J M and P = PM . Define the cp map nyn = Eω(Y ∗Y ) = p0. Ψ : B(L2(M )) → B(L2(M )) : T 7→ σ-weak lim n→ω y∗ nT yn. Observe that Ψ(1) = p0. Since Ψ is a cp map and Ψ(1) = p0 is a projection, we have Ψ(T ) = Ψ(1)Ψ(T )Ψ(1) = p0Ψ(T )p0 for every T ∈ B(L2(M )) and hence Ψ(B(L2(M ))) ⊂ B(p0L2(M )) using the identification p0B(L2(M ))p0 = B(p0L2(M )). We will then regard Ψ : B(L2(M )) → B(p0L2(M )) as a ucp map. Observe that Ψ(x) = Eω(Y ∗xY ) for all x ∈ M and hence Ψ(M ) ⊂ p0M p0 and ΨM is normal. Moreover, observe that ΨN : N → B(p0L2(M )) : x 7→ xp0 and 14 CYRIL HOUDAYER AND YUSUKE ISONO ΨJM J : JM J → B(p0L2(M )) : JxJ 7→ JxJp0 are unital ∗-homomorphisms. We will denote by ψ := ΨM : M → p0M p0 : x 7→ Ψ(x). Let KB as in Proposition 3.3 (for B = B). For all a, b ∈ M , we have EBω (b∗Y a) = EBω (EBω ⋊Γ(b∗Y a)) = EBω (b∗EBω ⋊Γ(Y )a) = 0. Since 0 = EBω (b∗Y a) = (EB(b∗yna))ω, we obtain that EB(b∗yna) → 0 σ-strongly as n → ω. Choose any cyclic unit vector ξ ∈ P such that eBξ = ξ. For all a, b, c, d ∈ M , we have = lim (cid:12)(cid:12)(cid:12)hΨ(aeBb) cξ, dξiL2(M )(cid:12)(cid:12)(cid:12) = lim n→ω(cid:12)(cid:12)(cid:12)hy∗ naeBbyn cξ, dξiL2(M )(cid:12)(cid:12)(cid:12) n→ω(cid:12)(cid:12)(cid:12)heBbyncξ, eBa∗yndξiL2(M )(cid:12)(cid:12)(cid:12) n→ω(cid:12)(cid:12)(cid:12)heB bync eBξ, eB a∗ynd eBξiL2(M )(cid:12)(cid:12)(cid:12) n→ω(cid:12)(cid:12)(cid:12)hEB(bync)eBξ, EB(a∗ynd)eBξiL2(M )(cid:12)(cid:12)(cid:12) n→ω(cid:12)(cid:12)(cid:12)hEB(bync)ξ, EB(a∗ynd)ξiL2(M )(cid:12)(cid:12)(cid:12) kEB(bync)ξkL2(M )kEB(a∗ynd)ξkL2(M ) = lim = lim = lim ≤ lim n→ω = 0. This implies that Ψ(aeBb) = 0. By construction, we have M = B ⋊ Γ and hence eB corre- sponds to the projection 1 ⊗ PCδe. Taking a = λg and b = λh for g, h ∈ Γ, we then obtain Ψ(cid:0)C1 ⊗ K(ℓ2(Γ))(cid:1) = 0. Since Ψ is a ucp map, we obtain Ψ(cid:0)B(L2(B)) ⊗min K(ℓ2(Γ))(cid:1) = 0 and hence Ψ(KB) = 0 using Lemma 3.2. Define the ucp map Ψ : M(KB)/KB → B(p0L2(M )) : a + KB 7→ Ψ(a). Using Proposition 3.3 in the case when B = B, we may then define the ucp composition map Φ = Ψ ◦ ν : (B ⋊red Γ) ⊗min J(B ⋊red Γ)J → B(p0L2(M )) : a ⊗ JbJ 7→ Ψ(a JbJ). Since ΨJM J is a unital ∗-homomorphism and since ψ = ΨM by definition, we have Φ(a ⊗ JbJ) = Ψ(a JbJ) = Ψ(a) Ψ(JbJ) = ψ(a) JbJp0 for all a, b ∈ B ⋊red Γ. Since (J(B ⋊red Γ)Jp0)′ ∩ B(p0L2(M )) = p0(JM J)′p0 ∩ B(p0L2(M )) = p0M p0, Proposition 3.5 and Lemma 3.6 imply that the normal ucp map ψ : M → p0M p0 has a ucp extension eψ : B(L2(M )) → p0M p0. Observe that N p0 ⊂ p0M p0 is still with expectation by [HU15, Proposition 2.2]. Denote by EN p0 : p0M p0 → N p0 a faithful normal conditional expectation. Define the unital ∗-homomorphism ι : N → N p0 : x 7→ xp0 and denote by z ∈ Z(N ) the unique central projection such that ker(ι) = N z⊥. Then N z ∼= N p0 and N zp0 = N p0. Define the ucp map Θ = ι−1 ◦ EN p0 ◦ eψ(z · z) : B(zL2(M )) → N z. Since ΘN z = idN z, Θ is a norm one projection and hence N z is amenable. We have therefore proved that if N ′ ∩ pM ωp 6⊂ p(Bω ⋊ Γ)p, then N has a nonzero amenable direct summand. (cid:3) Theorem 4.2. Let Γ be any bi-exact discrete group, B ⊂ B any inclusion of σ-finite von Neumann algebras with expectation and Γ y B any action that leaves the subalgebra B globally invariant. Assume moreover that B is amenable. Put M := B ⋊ Γ ⊂ B ⋊ Γ =: M. Let p ∈ M be any nonzero projection and N ⊂ pM p any von Neumann subalgebra with expectation. Then at least one of the following conditions holds true. • The von Neumann algebra N is amenable. • We have A (cid:22)M B for any finite von Neumann subalgebra A ⊂ N ′ ∩ pMp with expecta- tion. BI-EXACT GROUPS, STRONGLY ERGODIC ACTIONS AND TYPE III FACTORS 15 Proof. Since the proof is very similar to the one of Theorem 4.1, we will simply sketch it and point out the necessary changes compared to Theorem 4.1. Denote by (M, L2(M), J M, PM) a standard form for M = B ⋊ Γ as in Section 2. Suppose that there exists a finite von Neumann subalgebra A ⊂ N ′ ∩ pMp with expectation such that A 6(cid:22)M B. Observe that since N ′ ∩ pMp ⊂ pMp is with expectation, so is A ⊂ pMp. We will show N is amenable. We will use the identifications pB(L2(M ))p = B(pL2(M )) and pB(L2(M))p = B(pL2(M)). Take a net of unitaries (ui)i∈I in U (A) as in Theorem 2.7(ii) such that EB(b∗uia) → 0 σ-strongly for any a, b ∈ M. Fix a cofinal ultrafilter U on the directed set I and define the ucp map Ψ : B(L2(M)) → B(pL2(M)) : T 7→ σ-weak lim i→U u∗ i T ui. Observe that ΨM : M → pMp is normal. Indeed, since A ⊂ pMp is finite and with expecta- tion and since M is σ-finite, there exists a faithful state ϕ ∈ (pMp)∗ such that A ⊂ (pMp)ϕ. Since ui ∈ U (A) for all i ∈ I, this implies that (ϕ ◦ Ψ)(pxp) = ϕ(x) for all x ∈ pMp. Since ϕ is faithful and normal and since Ψ = Ψ(p · p), it follows that ΨM : M → pMp is indeed normal. Moreover, we have Ψ(x) = x for all x ∈ N . Let KB be as in Proposition 3.3. By a reasoning entirely similar to the one of the proof of Theorem 4.1, we have Ψ(KB) = 0. Define the ucp map Ψ : M(KB)/KB → B(pL2(M)) : a + KB 7→ Ψ(a). Using Proposition 3.3, we may then define the ucp composition map Ψ ◦ νB : (B ⋊red Γ) ⊗min J M(B ⋊red Γ)J M → B(pL2(M)) : a ⊗ J MbJ M 7→ Ψ(a J MbJ M). Proposition 3.5 and Lemma 3.6 imply that the normal ucp map ψ := ΨM : M → pMp L2(M) → L2(M ) the Jones projection corresponding to the inclusion M ⊂ M. We then have the identifications eM B(pL2(M))eM = B(pL2(M )) and eM J MeM = J M . Then we have has a ucp extension eψ : B(L2(M)) → (J M(B ⋊red Γ)J Mp)′ ∩ B(pL2(M)). Denote by eM : eM(cid:0)(J M(B ⋊red Γ)J Mp)′ ∩ B(pL2(M))(cid:1) eM = (J M (B ⋊red Γ)J M p)′ ∩ B(pL2(M )) = pM p and hence the ucp map eΨ := eM eψ( · ) eM : B(L2(M)) → pM p takes indeed values in pM p. Moreover, we have eΨ(x) = x for all x ∈ N . If we denote by EN : pM p → N a faithful normal conditional expectation, the ucp map Θ = EN ◦ eΨ(p · p) : B(pL2(M)) → N is a norm one projection. Therefore, N is amenable. (cid:3) Proof of Theorem A. Suppose that N has no amenable direct summand. Then by Theorem 4.1, we have N ′ ∩ pM ωp ⊂ p(Bω ⋊ Γ)p. We then apply Theorem 4.2 to N in the case when B := Bω and we obtain A (cid:22)Bω ⋊Γ Bω for any finite von Neumann subalgebra A ⊂ N ′ ∩ pM ωp with expectation. (cid:3) Proof of Corollary B. Let N ⊂ M be any von Neumann subalgebra with expectation such that N ′ ∩M ω has no type I direct summand. Denote by p ∈ Z(N ) the unique central projection such that N p has no amenable direct summand and N (1 − p) is amenable. Assume by contradiction that p 6= 0. Then (N p)′ ∩ pM ωp ⊂ p(Bω ⋊ Γ)p by Theorem A and (N p)′ ∩ pM ωp = p(N ′ ∩ M ω)p has no type I direct summand. By [CS78, Corollary 8] (see also [HS90, Theorem 11.1]), (N p)′ ∩ pM ωp contains a copy of the hyperfinite II1 factor R with expectation. We then have R (cid:22)Bω ⋊Γ Bω by Theorem A. Since R is of type II1 and Bω is abelian and hence of type I, we obtain a contradiction. Therefore, p = 0 and N is amenable. (cid:3) 5. Proof of Theorem C We start by proving a useful lemma which can be regarded as a generalization of the first part of the proof of [Ho15, Proposition C]. 16 CYRIL HOUDAYER AND YUSUKE ISONO Lemma 5.1. Let R be any strongly ergodic nonsingular equivalence relation defined on a standard probability space (X, µ). Put A = L∞(X) and M = L(R). Denote by EA : M → A the unique faithful normal conditional expectation. Fix any faithful state τ ∈ A∗ and put ϕ = τ ◦ EA ∈ M∗. If M is not full, then there exists a sequence of unitaries un ∈ U (M ) such that the following conditions hold: (i) limn kunϕ − ϕunk = 0, (ii) limn kxun − unxkϕ = 0 for all x ∈ M and (iii) limn kEA(xuny)kϕ = 0 for all x, y ∈ M . Proof. Assume that M is not full. Then M ′ ∩ (M ω)ϕω is diffuse by [HR14, Corollary 2.6] for any nonprincipal ultrafilter ω ∈ β(N) \ N. Then a combination of the first part of the proof of [HR14, Theorem A] and Lemma 2.1 shows that there exists a sequence of unitaries un ∈ U (M ) such that the following conditions hold: (i) limn kunϕ − ϕunk = 0, (ii) limn kxun − unxkϕ = 0 for all x ∈ M and (iii) un → 0 σ-weakly as n → ∞. It remains to prove that Conditions (i), (ii), (iii) imply that limn kEA(xuny)kϕ = 0 for all x, y ∈ M . The rest of the proof is entirely analogous to the first part of the proof of [Ho15, Proposition C] and we only give the details for the sake of completeness. Observe that for every nonprincipal ultrafilter ω ∈ β(N) \ N, Condition (i) implies that (un)n ∈ Mω(M ) and Conditions (ii) and (iii) imply that (un)ω ∈ M ′ ∩ (M ω)ϕω and ϕω((un)ω) = 0. We start by proving the following claim. Claim. We have limn kEA(un)kϕ = 0 Proof of the Claim. Let g ∈ [R] be any element and denote by ug ∈ U (L(R)) the corresponding unitary element. Since ugEA(un)u∗ g = EA(ugunu∗ g), we have kEA(un)u∗ g − u∗ gEA(un)kϕ = kugEA(un)u∗ = kEA(ugunu∗ ≤ kugunu∗ = kunu∗ g − u∗ g − unkϕ g − EA(un)kϕ g − un)kϕ gunkϕ → 0 as n → ∞. Define E = span {aug a ∈ A, g ∈ [R]} and observe that E is a unital σ-strongly dense ∗- subalgebra of M . The above calculation implies that limn kxEA(un) − EA(un)xkϕ = 0 for every x ∈ E. Let ω ∈ β(N) \ N be any nonprincipal ultrafilter. For every x ∈ E, we have kxEAω ((un)ω) − EAω ((un)ω)xkϕω = limn→ω kxEA(un) − EA(un)xkϕ = 0 and hence we have xEAω ((un)ω) = EAω ((un)ω)x. Since E is σ-strongly dense in M , this further implies that xEAω ((un)ω) = EAω ((un)ω)x for every x ∈ M . Since R is strongly ergodic, this implies that EAω ((un)ω) = ϕω((un)ω)1 = 0 and hence limn→ω kEA(un)kϕ = kEAω ((un)ω)kϕω = 0. Since this is true for every ω ∈ β(N) \ N, we finally obtain that limn kEA(un)kϕ = 0. (cid:3) We can now finish the proof of Lemma 5.1. Let g ∈ [R] be any element such that g2 = 1. Put Xg = {s ∈ X g · s = s} and observe that ug = u∗ g = zg. Since A is abelian and hence tracial, a combination of the fact that u∗ gzg ∈ A and the Claim implies that g, zg := EA(ug) = 1Xg and u∗ gzg = zgu∗ kEA(unu∗ g)zgkϕ = kEA(un u∗ gzg)kϕ = kEA(un) u∗ gzgkϕ ≤ kEA(un)kϕ → 0 as n → ∞. Denote by J the nonempty directed set (for the inclusion) of all the families of projections (zi)i∈I in A such that zi ≤ 1 − zg, zi ⊥ zj for all i 6= j ∈ I and zi ⊥ ugzju∗ g for all i, j ∈ I. BI-EXACT GROUPS, STRONGLY ERGODIC ACTIONS AND TYPE III FACTORS 17 By Zorn's lemma, let (zi)i∈I be a maximal element in J . Put z = Pi∈I zi and assume by g 6= 0. Since g2 = 1, we have contradiction that z + ugzu∗ ugeu∗ g. Then the family ((zi)i∈I , z′) is in J and this contradicts the maximality of the family (zi)i∈I in J . Therefore, we have z + ugzu∗ g = 1 − zg. A calculation entirely analogous to [Ho15, Proposition C, Equation (6.6)] shows that g = e. Since e ≤ 1 − zg = 1{s∈Xg·s6=s}, we can find 0 6= z′ ≤ e such that z′ ⊥ ugz′u∗ g 6= 1 − zg. Put e = 1 − zg − z − ugzu∗ kEA(unu∗ g)(1 − zg)k2 g)k2 ϕ ϕ = kEA(unu∗ = kEA(unu∗ = kEA(unu∗ = kEA((zun − unz)u∗ g)(z + ugzu∗ g)zk2 g)(z − ugzu∗ g)k2 ϕ g)k2 ϕ. ϕ + kEA(unu∗ g)ugzu∗ gk2 ϕ Since zun −unz → 0 σ-strongly as n → ∞, we also have that EA((zun −unz)u∗ as n → ∞. The above calculation implies that limn kEA(unu∗ implies that g) → 0 σ-strongly g)(1 − zg)kϕ = 0. This further ϕ(cid:1) = 0. lim sup n kEA(unu∗ g)k2 ϕ = lim sup n g)zgk2 ϕ + kEA(unu∗ g)(1 − zg)k2 (cid:0)kEA(unu∗ Define F = span{aug a ∈ A, g ∈ [R], g2 = 1}. By the proof of [FM75, Theorem 1], it follows that F is a σ-strongly dense linear ∗-subspace of M . The previous reasoning shows that limn kEA(unx)kϕ = 0 for every x ∈ F. Let ω ∈ β(N) \ N be any nonprincipal ultrafilter. For every x ∈ F, we have kEAω ((un)ωx)kϕω = limn→ω kEA(unx)kϕ = 0 and hence EAω ((un)ωx) = 0. Since F is σ-strongly dense in M , this further implies that EAω ((un)ωx) = 0 for every x ∈ M . Using Condition (ii), we also have EAω (x(un)ωy) = EAω ((un)ωxy) = 0 for every x, y ∈ M . This implies that limn→ω kEA(xuny)kϕ = kEAω (x(un)ωy)kϕω = 0. Since this is true for every ω ∈ β(N) \ N, we finally obtain that limn kEA(xuny)kϕ = 0 for all x, y ∈ M . (cid:3) Proof of Theorem C. Simply write B = L∞(X) and M = B ⋊ Γ. Assume by contradiction that M is not full. Fix any nonprincipal ultrafilter ω ∈ β(N) \ N. Since Γ y (X, µ) is strongly ergodic, Lemma 5.1 shows that there exists u ∈ U (Mω) such that EBω (uλs) = 0 for every s ∈ Γ. Then for every s ∈ Γ, we have EBω (EBω ⋊Γ(u)λs) = EBω (EBω ⋊Γ(uλs)) = EBω (uλs) = 0. This implies that EBω ⋊Γ(u) = 0. Since M is a nonamenable factor, Theorem A shows that Mω ⊂ M ′ ∩ M ω ⊂ Bω ⋊ Γ and hence u ∈ U (Bω ⋊ Γ). We then have u = EBω ⋊Γ(u) = 0. This is a contradiction. (cid:3) 6. Group measure space type III factors with no central sequence Definition 6.1. Let G be any locally compact second countable group. Let G y (X, µ) and G y (Y, ν) be any nonsingular Borel actions on standard probability spaces. We say that • G y (Y, ν) is a measurable quotient of G y (X, µ) if, after discarding null G-invariant Borel subsets, there exists a G-equivariant Borel quotient map q : X → Y such that [q∗µ] = [ν]. • G y (Y, ν) is measurably conjugate to G y (X, µ) if, after discarding null G-invariant Borel subsets, there exists a G-equivariant Borel isomorphism θ : X → Y such that [θ∗µ] = [ν]. Let G be any locally compact second countable group and H < G any closed subgroup. En- dowed with the quotient topology, G/H is a continuous G-space, that is, the action G y G/H defined by (g, hH) 7→ ghH is continuous. The quotient space G/H carries, up to equiva- lence, a unique G-quasi-invariant regular Borel probability measure ν ∈ Prob(G/H). Any such 18 CYRIL HOUDAYER AND YUSUKE ISONO G-quasi-invariant regular Borel probability measure is associated with a rho-function for the pair (G, H) (see e.g. [BdlHV08, Appendix B]). The action G y G/H is indeed a measurable quotient of the translation action G y G (see [BdlHV08, Theorem B.1.4]). Let G be any noncompact connected simple Lie group and P < G any minimal parabolic subgroup (e.g. G = SLn(R) and P = subgroup of upper triangular matrices, for n ≥ 2). Fix a G-quasi-invariant Borel regular probability measure ν ∈ Prob(G/P ). Denote by ∆P : P → R∗ + the modular homomorphism and observe that ∆P (P ) = R∗ + (this follows from [BdlHV08, Proposition B.1.6 (ii)] and [Zi84, Proposition 4.3.2]). Put L = ker(∆P ). The Radon-Nikodym cocycle associated with the action G y G/P is the map defined by Ω : G × G/P → R : (g, hP ) 7→ log(cid:18) dg∗ν dν (hP )(cid:19) . Observe that Ω : G × G/P → R is a continuous map by [BdlHV08, Theorem B.1.4]. The Maharam extension G y G/P × R is the continuous action defined by g · (hP, t) = (ghP, t + Ω(g, hP )). By [BdlHV08, Lemma B.1.3], we have Ω(g, P ) = (log ◦∆P )(g) for every g ∈ P . Since moreover (log ◦∆P )(P ) = R, the Maharam extension G y G/P × R is transitive and the stabilizer of the point (P, 0) is equal to L. The mapping θ : G/L → G/P × R : gL 7→ (gP, Ω(g, P )) is a well-defined G-equivariant homeomorphism that yields a measurable conjugacy between the action G y G/L and the Maharam extension G y G/P × R. Therefore, we have proved the following useful fact. Proposition 6.2 (see [BN11, Proposition 4.7]). The Maharam extension of G y G/P is measurably conjugate to G y G/L. From now on, fix n ≥ 2, G = SLn(R) and Λ = SLn(Q) and denote by P < G the minimal subgroup of upper triangular matrices. By [BISG15, Theorem A and Proposition 7.4], the translation action Λ y G is strongly ergodic and so is the nonsingular action Λ y G/P (recall that strong ergodicity is stable under taking measurable quotients). Fix a surjective group homomorphism π : F∞ → Λ such that ker(π) < F∞ is a nonamenable subgroup. For simplicity, write Γ = F∞. Put X = [0, 1]Γ and µ = Leb⊗Γ and consider the Bernoulli shift action Γ y X defined by γ · (xγ ′)γ ′∈Γ = (xγ−1γ ′)γ ′∈Γ. Observe that Γ y X preserves the Borel probability measure µ and is essentially free and strongly ergodic. Since ker(π) is nonamenable, the restricted action ker(π) y X is also strongly ergodic. Since ker(π) < F∞ is a nonamenable free subgroup and hence not inner amenable, the crossed product II1 factor L∞(X) ⋊ ker(π) is full by [Ch81]. Define the action Γ y X × G/P by γ · (x, hP ) = (γx, π(γ)hP ). Observe that Γ y X × G/P quasi-preserves the product measure µ ⊗ ν and is essentially free. Theorem 6.3. Keep the same notation as above. The following assertions hold true: (i) The nonsingular action Γ y X × G/P is essentially free and strongly ergodic and its Maharam extension Γ y X × G/P × R is also essentially free and strongly ergodic. (ii) The group measure space factor M = L∞(X × G/P ) ⋊ Γ is a full type III1 factor and its continuous core c(M ) is a full type II∞ factor. Proof. (i) As we already pointed out, the nonsingular action Γ y X × G/P is essentially free and so is its Maharam extension Γ y X × G/P × R. BI-EXACT GROUPS, STRONGLY ERGODIC ACTIONS AND TYPE III FACTORS 19 We next prove that the nonsingular action Γ y X × G defined by γ · (x, h) = (γx, π(γ)h) is strongly ergodic. Put A = L∞(X) and B = L∞(G) so that L∞(X × G) = A ⊗ B. Write N = (A ⊗ B) ⋊ Γ. Fix a nonprincipal ultrafilter ω ∈ β(N) \ N. We need to show that N ′ ∩ (A ⊗ B)ω = C1. Observe that (A ⊗ B) ⋊ ker(π) = (A ⋊ ker(π)) ⊗ B and N ′ ∩ (A ⊗ B)ω ⊂ (A⋊ker(π))′ ∩((A⋊ker(π))⊗B)ω. Since A⋊ker(π) is a full type II1 factor, [Co75, Theorem 2.1] implies that (A ⋊ ker(π))′ ∩ ((A ⋊ ker(π)) ⊗ B)ω = Bω and hence N ′ ∩ (A ⊗ B)ω = N ′ ∩ Bω = (Bω)Λ. By [BISG15, Theorem A], the nonsingular action Λ y G is strongly ergodic, that is, (Bω)Λ = C1. This implies that N ′ ∩ (A ⊗ B)ω = C1 and hence the nonsingular action Γ y X × G is strongly ergodic. Since the nonsingular action Γ y X × G/P is a quotient of the strongly ergodic nonsingular action Γ y X × G, it follows that Γ y X × G/P is also strongly ergodic. Consider the Maharam extension Λ y G/P × R of the nonsingular action Λ y G/P . By Proposition 6.2, the Maharam extension Λ y G/P × R is measurably conjugate to the nonsingular action Λ y G/L where L = ker(∆P ) and ∆P : P → R∗ + is the modular homomorphism. Since Γ y X is pmp, the action Γ y X × G/L defined by γ · (x, hL) = (γx, π(γ)hL) can be identified with the Maharam extension of the nonsingular action Γ y X × G/P . Since the nonsingular action Γ y X ×G/L is a quotient of the strongly ergodic nonsingular action Γ y X ×G, it follows that Γ y X × G/L is also strongly ergodic. Therefore, the Maharam extension Γ y X × G/P × R of the nonsingular action Γ y X × G/P is strongly ergodic. (ii) This is a consequence of Theorem C. (cid:3) Keep the same notation as above. Fix 0 < λ < 1, put T = 2π Define the nonsingular action Γ y X × G/P × T by log λ and identify T = R/(T Z). γ · (x, hP, t + T Z) = (γx, π(γ)hP, t + Ω(π(γ), hP ) + T Z). Observe that the nonsingular action Γ y X × G/P × T is a measurable quotient of the nonsingular action Γ y X × G/P × R and hence is strongly ergodic by Theorem 6.3(i). Moreover, we have a canonical identification L∞(X × G/P × T) ⋊ Γ = M ⋊σϕ T Z. It follows that L∞(X × G/P × T) ⋊ Γ is a type IIIλ factor by [Co85, Lemma 1]. Observe that L∞(X × G/P × T) ⋊ Γ is full by Theorem C. Alternatively, since c(M ) is full by Theorem 6.3(ii), L∞(X × G/P × T) ⋊ Γ = M ⋊σϕ Z is full by [TU14, Lemma 6]. T Proof of Corollary D. This is a consequence of Theorem 6.3 and the above construction. (cid:3) 7. Further remarks In [HR14], the first named author and Raum investigated the asymptotic structure of Shlyakht- enko's free Araki -- Woods factors [Sh96]. Among other things, they proved in [HR14, Theorem A] that any diffuse von Neumann algebra M with separable predual satisfying Ozawa's con- dition (AO) is ω-solid, that is, for any von Neumann subalgebra with expectation N ⊂ M such that the relative commutant N ′ ∩ M ω is diffuse, we have that N is amenable. The proof was based on a combination of Ozawa's C∗-algebraic techniques and an analysis of the relative commutant N ′ ∩ M ω and its centralizer [HR14, Theorem 2.3] (see [Io12a, Lemma 2.7] for the tracial case). In this subsection, we observe that ω-solidity can be easily obtained using the same proof as the one of Theorem 4.1 without relying on the analysis of the relative commutant N ′ ∩ M ω from [HR14, Theorem 2.3]. We moreover remove the separability assumption of the predual. 20 CYRIL HOUDAYER AND YUSUKE ISONO Theorem 7.1 ([HR14, Theorem A]). Let M be any diffuse σ-finite von Neumann algebra satisfying Ozawa's condition (AO). Let p ∈ M be any nonzero projection and N ⊂ pM p any von Neumann subalgebra with expectation. Then at least one of the following conditions holds true: • The von Neumann algebra N has a nonzero amenable direct summand. • We have N ′ ∩ pM ωp ⊂ pM p. In that case, N ′ ∩ pM ωp = N ′ ∩ pM p is moreover discrete. Proof. Suppose that N has no amenable direct summand. Then the exact same argument as in the proof of Theorem 4.1 using Ozawa's condition (AO) in lieu of Proposition 3.3 shows that N ′ ∩ pM ωp ⊂ pM p and hence N ′ ∩ pM ωp = N ′ ∩ pM p. Since M is diffuse and solid [Oz03, Theorem 6] (see also [VV05, Theorem 2.5]), it follows that pM p is also diffuse and solid. Since N ⊂ pM p has no amenable direct summand, it follows that N ′ ∩ pM p is necessarily discrete. (cid:3) In view of Proposition 3.4, we finally observe the following condition (AO) in the ultraproduct representation. Proposition 7.2. Let M be any σ-finite von Neumann algebra and ω ∈ β(N)\N any nonprin- cipal ultrafilter. Denote by (M, L2(M ), J M , PM ) (resp. (M ω, L2(M ω), J M ω , PM ω )) a standard form for M (resp. M ω). Assume there are unital C∗-subalgebras A, B ⊂ M such that the map ν : A ⊗alg J M BJ M → B(L2(M ))/K(L2(M )) : a ⊗ J M bJ M 7→ a J M bJ M + K(L2(M )) is continuous with respect to the minimal tensor norm. Then the map νω : A⊗alg J M ω BJ M ω → B(L2(M ω))/K(L2(M ω)) : a⊗J M ω bJ M ω 7→ a J M ω bJ M ω +K(L2(M ω)) is continuous with respect to the minimal tensor norm. Proof. The proof is a variation of the one of Proposition 3.4. Put C := C∗(cid:8)M, J M M J M(cid:9) ⊂ B(L2(M )) Cω := C∗(cid:8)M, J M ω M J M ω(cid:9) ⊂ B(L2(M ω)). Observe that C + K(L2(M )) (resp. Cω + K(L2(M ω))) is a C∗-subalgebra of B(L2(M )) (resp. B(L2(M ω))). Fix a faithful state ϕ ∈ M∗. Denote by e : L2(M ) → Cξϕ and f : L2(M ω) → Cξϕω the corresponding orthogonal projections. Observe that K(L2(M )) is the norm closure in B(L2(M )) of M eM . Denote by N :=Qω M the Groh -- Raynaud ultraproduct and by p ∈ N the support projection of the ultraproduct state ϕω ∈ N∗. Claim. There is a ∗-homomorphism θ : C + K(L2(M )) → B(L2(M ω)) such that θ(x) = x and θ(J M yJ M ) = J M ω for all x, y ∈ M and θ(e) = f . yJ M ω Proof of the Claim. Keep the same notation as in the proof of the Claim of Proposition 3.4. By Lemma 2.2, πω((e)n) commutes with p and J N and hence πω((e)n) commutes with ep := pJ N pJ N . Since ep commutes with πω(M ) and πω(J M M J M ) = J N πω(M )J N , ep commutes with πω(C + K(L2(M ))). Recall that epNep ∼= pN p ∼= M ω and epL2(M )ω = L2(M ω). Then the ∗-homomorphism satisfies all the conditions of the Claim. θ : C + K(L2(M )) → B(L2(M ω)) : T 7→ epπω(T )ep (cid:3) BI-EXACT GROUPS, STRONGLY ERGODIC ACTIONS AND TYPE III FACTORS 21 Since θ(C) = Cω and θ(K(L2(M ))) ⊂ K(L2(M ω)), θ induces a ∗-homomorphism eθ :(cid:0)C + K(L2(M ))(cid:1) /K(L2(M )) →(cid:0)Cω + K(L2(M ω))(cid:1) /K(L2(M ω)). BJ M ω → A ⊗min J M BJ M the tautological ∗-isomorphism. Then the Denote by ι : A ⊗min J M ω composition map is continuous with respect to the minimal tensor norm. (cid:3) νω = eθ ◦ ν ◦ ι : A ⊗alg J M ω BJ M ω → B(L2(M ω))/K(L2(M ω)) References [AO74] [AH12] C.A. Akemann, P.A. Ostrand, On a tensor product C∗-algebra associated with the free group on two generators. J. Math. Soc. Japan 27 (1975), 589 -- 599. H. Ando, U. Haagerup, Ultraproducts of von Neumann algebras. J. Funct. Anal. 266 (2014), 6842 -- 6913. [BdlHV08] M. Bekka, P. de la Harpe, A. Valette, Kazhdan's property (T). New Mathematical Monographs, 11. Cambridge University Press, Cambridge, 2008. xiv+472 pp. [BISG15] R. Boutonnet, A. Ioana, A. Salehi Golsefidy, Local spectral gap in simple Lie groups and [BN11] [BO08] [Ch81] [Co72] [Co74] [Co75] [Co85] [CJ81] [CS78] [FM75] [Ga10] [HS90] [Ho15] [HI15] [HR14] [HU15] [HV12] [Io12a] [Io12b] [Is12] [Is13] λ and closure of inner automorphisms. J. Operator applications. arXiv:1503.06473 L. Bowen, A. Nevo, Pointwise ergodic theorems beyond amenable groups. Ergodic Theory Dynam. Systems 33 (2013), 777 -- 820. N.P. Brown, N. Ozawa, C∗-algebras and finite-dimensional approximations. Graduate Studies in Mathematics, 88. American Mathematical Society, Providence, RI, 2008. M. Choda, Inner amenability and fullness. Proc. Amer. Math. Soc. 86 (1982), 663 -- 666. A. Connes, Une classification des facteurs de type III. Ann. Sci. ´Ecole Norm. Sup. 6 (1973), 133 -- 252. A. Connes, Almost periodic states and factors of type III1. J. Funct. Anal. 16 (1974), 415 -- 445. A. Connes, Classification of injective factors. Cases II1, II∞, IIIλ, λ 6= 1. Ann. of Math. 74 (1976), 73 -- 115. A. Connes, Factors of type III1, property L′ Theory 14 (1985), 189 -- 211. A. Connes, V.F.R. Jones, A II1 factor with two non-conjugate Cartan subalgebras. Bull. Amer. Math. Soc. 6 (1982), 211 -- 212. A. Connes, E. Størmer, Homogeneity of the state space of factors of type III1. J. Funct. Anal. 28 (1978), 187 -- 196. J. Feldman, C.C. Moore, Ergodic equivalence relations, cohomology, and von Neumann algebras. I, II. Trans. Amer. Math. Soc. 234 (1977), 289 -- 324, 325 -- 359. D. Gaboriau, Orbit equivalence and measured group theory. Proceedings of the International Con- gress of Mathematicians (Hyderabad, 2010), Vol. III, Hindustan Book Agency (2010), 1501 -- 1527. U. Haagerup, E. Størmer, Equivalence of normal states on von Neumann algebras and the flow of weights. Adv. Math. 83 (1990), 180 -- 262. D.J. Hoff, Von Neumann algebras of equivalence relations with nontrivial one-cohomology. J. Funct. Anal. 270 (2016), 1501 -- 1536. C. Houdayer, Y. Isono, Unique prime factorization and bicentralizer problem for a class of type III factors. arXiv:1503.01388 C. Houdayer, S. Raum, Asymptotic structure of free Araki-Woods factors. Math. Ann. 363 (2015), 237 -- 267. C. Houdayer, Y. Ueda, Rigidity of free product von Neumann algebras. arXiv:1507.02157 C. Houdayer, S. Vaes, Type III factors with unique Cartan decomposition. J. Math. Pures Appl. 100 (2013), 564 -- 590. A. Ioana, Cartan subalgebras of amalgamated free product II1 factors. Ann. Sci. ´Ecole Norm. Sup. 48 (2015), 71 -- 130. A. Ioana, Classification and rigidity for von Neumann algebras. Proceedings of the 6th European Congress of Mathematics (Krakow, 2012), European Mathematical Society Publishing House Y. Isono, Weak exactness for C∗-algebras and application to condition (AO). J. Funct. Anal. 264 (2013), 964 -- 998. Y. Isono, On bi-exactness of discrete quantum groups. Int. Math. Res. Not. Volume 2014, Article ID rnu043. [MvN43] F. Murray, J. von Neumann, Rings of operators. IV. Ann. of Math. 44 (1943), 716 -- 808. 22 [Oc85] [Oz03] [Oz04] [Oz16] [Pe06] [Po95] [Po01] [Po03] [Po06a] [Po06b] [Po06c] [PV11] [Ra99] [Sa09] [Sh96] [Sk88] [TU14] [Ue00] [Va10] [VV05] [Zi84] CYRIL HOUDAYER AND YUSUKE ISONO A. Ocneanu, Actions of discrete amenable groups on von Neumann algebras. Lecture Notes in Mathematics, 1138. Springer-Verlag, Berlin, 1985. iv+115 pp. N. Ozawa, Solid von Neumann algebras. Acta Math. 192 (2004), 111 -- 117. N. Ozawa, A Kurosh type theorem for type II1 factors. Int. Math. Res. Not. (2006), Art. ID 97560, 21 pp. N. Ozawa, A remark on fullness of some group measure space von Neumann algebras. arXiv:1602.02654 J. Peterson, L2-rigidity in von Neumann algebras. Invent. Math. 175 (2009), 417 -- 433. S. Popa, Classification of subfactors and their endomorphisms. CBMS Regional Conference Series in Mathematics, 86. Published for the Conference Board of the Mathematical Sciences, Washington, DC; by the American Mathematical Society, Providence, RI, 1995. x+110 pp. S. Popa, On a class of type II1 factors with Betti numbers invariants. Ann. of Math. 163 (2006), 809 -- 899. S. Popa, Strong rigidity of II1 factors arising from malleable actions of w-rigid groups I. Invent. Math. 165 (2006), 369 -- 408. S. Popa, Deformation and rigidity for group actions and von Neumann algebras. Proceedings of the International Congress of Mathematicians (Madrid, 2006), Vol. I, European Mathematical Society Publishing House, 2007, p. 445 -- 477. S. Popa, On the superrigidity of malleable actions with spectral gap. J. Amer. Math. Soc. 21 (2008), 981 -- 1000. S. Popa, On Ozawa's property for free group factors. Int. Math. Res. Not. IMRN 2007, no. 11, Art. ID rnm036, 10 pp. S. Popa, S. Vaes, Unique Cartan decomposition for II1 factors arising from arbitrary actions of free groups. Acta Math. 212 (2014), 141 -- 198. Y. Raynaud, On ultrapowers of non commutative Lp-spaces. J. Operator Theory 48 (2002), 41 -- 68. H. Sako, Measure equivalence rigidity and bi-exactness of groups. J. Funct. Anal. 257 (2009), 3167 -- 3202. D. Shlyakhtenko, Free quasi-free states. Pacific J. Math. 177 (1997), 329 -- 368. G. Skandalis, Une notion de nucl´earit´e en K-th´eorie (d'apr`es J. Cuntz). K-Theory 1 (1988), 549 -- 573. Encyclopaedia of Mathematical Sciences, 125. Operator Algebras and Non-commutative Ge- ometry, 6. Springer-Verlag, Berlin, 2003. xxii+518 pp. R. Tomatsu, Y. Ueda, A characterization of fullness of continuous cores of type III1 free product factors. To appear in Kyoto J. Math. arXiv:1412.2418 Y. Ueda, Fullness, Connes' χ-groups, and ultra-products of amalgamated free products over Cartan subalgebras. Trans. Amer. Math. Soc. 355 (2003), 349 -- 371. S. Vaes, Rigidity for von Neumann algebras and their invariants. Proceedings of the International Congress of Mathematicians (Hyderabad, 2010), Vol. III, Hindustan Book Agency, 2010, 1624 -- 1650. S. Vaes, R. Vergnioux, The boundary of universal discrete quantum groups, exactness, and facto- riality. Duke Math. J. 140 (2007), 35 -- 84. R. Zimmer, Ergodic theory and semisimple groups. Monographs in Mathematics, 81. Birkhauser Verlag, Basel, 1984. x+209 pp. Laboratoire de Math´ematiques d'Orsay, Universit´e Paris-Sud, CNRS, Universit´e Paris-Saclay, 91405 Orsay, France E-mail address: [email protected] RIMS, Kyoto University, 606-8502 Kyoto, Japan E-mail address: [email protected]
1910.06605
1
1910
2019-10-15T09:05:31
$C^*$-Operator systems and crossed products
[ "math.OA" ]
The purpose of this paper is to introduce a consistent notion of universal and reduced crossed products by actions and coactions of groups on operator systems and operator spaces. In particular we shall put emphasis to reveal the full power of the universal properties of the the universal crossed products. It turns out that to make things consistent, it seems useful to perform our constructions on some bigger categories which allow the right framework for studying the universal properties and which are stable under the construction of crossed products even for non-discrete groups. In the case of operator systems, this larger category is what we call a $C^*$-operator system, i.e., a selfadjoint subspace $X$ of some $\mathcal B(H)$ which contains a $C^*$-algebra $A$ such that $AX=X=XA$. In the case of operator spaces, the larger category is given by what we call $C^*$-operator bimodules. After we introduced the respective crossed products we show that the classical Imai-Takai and Katayama duality theorems for crossed products by group (co-)actions on $C^*$-algebras extend one-to-one to our notion of crossed products by $C^*$-operator systems and $C^*$-operator bimodules.
math.OA
math
C ∗-OPERATOR SYSTEMS AND CROSSED PRODUCTS MASSOUD AMINI, SIEGFRIED ECHTERHOFF, AND HAMED NIKPEY Abstract. The purpose of this paper is to introduce a consistent no- tion of universal and reduced crossed products by actions and coactions of groups on operator systems and operator spaces. In particular we shall put emphasis to reveal the full power of the universal properties of the the universal crossed products. It turns out that to make things consistent, it seems useful to perform our constructions on some bigger categories which allow the right framework for studying the universal properties and which are stable under the construction of crossed prod- ucts even for non-discrete groups. In the case of operator systems, this larger category is what we call a C ∗-operator system, i.e., a selfadjoint subspace X of some B(H) which contains a C ∗-algebra A such that AX = X = XA. In the case of operator spaces, the larger category is given by what we call C ∗-operator bimodules. After we introduced the respective crossed products we show that the classical Imai-Takai and Katayama duality theorems for crossed products by group (co-)actions on C ∗-algebras extend one-to-one to our notion of crossed products by C ∗-operator systems and C ∗-operator bimodules. 1. Introduction In the world of C ∗-algebras, the construction of crossed-products A ⋊α G for an action of a locally compact group G an a C ∗-algebra A is one of the most fundamental tools in the theory -- not only to construct interesting examples of C ∗-algebras, but also in the application of C ∗-algebra theory in Harmonic Analysis, Non-commutative Geometry, Topology, and other areas of mathematics. Having this in mind it is very surprising that a serious study of a similar construction did not appear in the world of non- selfadoint operator algebras, operator systems, or operator spaces until the recent works of Katsoulis and Ramsay [20] in the setting of operator algebras and the even more recent work [16] of Harris and Kim in which they give a 2000 Mathematics Subject Classification. Primary 46L07; Secondary 47L65. Key words and phrases. Operator space, operator system, C ∗-algebra, dynamical sys- tem, crossed product. Funded by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) under Germany's Excellence Strategy EXC 2044 390685587, Mathematics Mnster: Dy- namicsGeometryStructure, and SFB 878, Groups, Geometry & Actions. Also funded by the German Academic Exchange Service DAAD. 1 2 MASSOUD AMINI, SIEGFRIED ECHTERHOFF, AND HAMED NIKPEY construction of crossed products of operator systems by actions of discrete groups, following some of the ideas developed in [20] for the construction. But we should also mention the earlier preprint [25] by Chi-Keung Ng, where he introduces reduced (i.e., spatially defined) crossed products for coactions of quantum groups on operator spaces. Just a few days before the first version of [20] appeared on the arXiv, the authors of this paper posted a preprint describing a crossed product construction for group actions on operator spaces (see [1]). Although this paper contained many ideas which were quite similar to ideas used in [20], the proof of a central theorem ([1, Theorem 4.3]) turned out to be wrong, and since we didn't see a way for a quick repair, we decided to withdraw the paper from the arXiv. We are very grateful to Elias Katsoulis for having pointed out this error to us! The problem was that in [1] we defined the full and reduced crossed products V ⋊u α G as the completions of Cc(G, V ) inside the full and reduced crossed products by a canonical action of G on the enveloping C ∗-algebra C ∗ e (X(V )), where X(V ) denotes the Paulsen system of V . Indeed, there is no problem with this in case of the reduced crossed products, but the universal crossed product V ⋊u α G should enjoy a universal property for suitable covariant representations of the system (V, G, α) (which was the content of the unfortunate [1, Theorem 4.3]). However, [16, Theorem 5.6] indicates that this cannot be true in general. The way out is to define the universal crossed product V ⋊u α G as the closure of Cc(G, V ) inside the universal crossed product C ∗ u(X(V ))⋊α,uG where C ∗ u(X(V )) denotes the universal C ∗-hull of X(V ) as introduced by Kirchberg and Wassermann in [22]. This was the approach of [20] in case of operator algebras and of [16] in case of operator systems. α G and V ⋊r If we want to exploit the full power of universal properties for the univer- sal crossed products, however, we would want to have a one-to-one corre- spondence between completely bounded covariant maps (ϕ, u) of the system (V, G, α) and the completely bounded maps Φ of V ⋊u α G via a canonically defined integrated form ϕ ⋊ u. But it turns out that in order to obtain such a correspondence we need to remember more information of the ambient crossed product C ∗ α G. Indeed, taking the completion of u(X(V )) ⋊u Cc(G, X(V )) =(cid:18) Cc(G) Cc(G, V ∗) Cc(G, V ) Cc(G) (cid:19) ⊆ C ∗ u(X(V )) ⋊α,u G and considering the convolution products between the upper diagonal entries gives V ⋊u u(G)-bimodule, and it is this structure which one needs to take into account for a good description of the universal properties. α G the structure of an operator C ∗ C ∗-OPERATOR SYSTEMS AND CROSSED PRODUCTS 3 Other problems appear if we consider crossed products by operator sys- tems instead of operator spaces. Since in the above procedure we defined crossed products by V via a crossed-product construction with the Paulsen system X(V ), it appears to be useful to consider at first the case of crossed products by general operator systems X. As mentioned above, such crossed products have been introduced by Harris and Kim in [16] for discrete groups. The reason for the restriction to the discrete case was the simple fact that the analogous construction for non-discrete groups G would result in a non- unital (but selfadjoint) operator space, hence it does not land in the right category. The other draw back is similar as the one described above for op- erator spaces crossed-products: we need to keep more structure than simply the completion of Cc(G, X) inside C ∗ u(X) ⋊u G in order to get the full power of the universal properties. Our way out is to extend the category of operator systems to what we call C ∗-operator systems: a concrete C ∗-operator system (A, X) is a pair of subsets A ⊆ X ⊆ B(H) for some Hilbert space H, such that X = X ∗, A is a non-degenerate C ∗-subalgebra of B(H), and AX = X = XA. A morphism from (A, X) to the C ∗-operator system (B, Y ) is then a ccp map ϕX : X → Y such that the restriction ϕA := ϕX A is a ∗-homomorphisms from A to B and such that ϕX (ax) = ϕA(a)ϕX (x) and ϕX (xa) = ϕX (x)ϕA(a) for all a ∈ A and x ∈ X. Of course, if X ⊆ B(H) is a classical operator system, then (C1, X) is C ∗-operator system in this sense, and every ucp map between operator systems X and Y extends to a morphism in the above sense from (C1, X) to (C1, Y ). Thus we get an inclusion of the category of operator systems into the category of C ∗-operator systems. After some preliminaries given in Section 2 we introduce C ∗-operator systems in Section 3, where we also introduce a corresponding notion of multiplier C ∗-operator systems which play an analogous role as the multiplier algebra for a C ∗- algebra. In particular, for a C ∗-operator system (A, X) the multiplier system (M (A), M (X)) can be considered as the largest unitization of (A, X) and it always contains the unitization ( A, X) if (A, X) has not unit (which means that A has no unit). An important feature of the multiplier system is that every non-degennerate morphism from (A, X) to (B, Y ) extends uniquely to a morphism from (M A), M (X)) to (M (B), M (Y )). In Section 4 we study C ∗-hulls of C ∗-operator systems, i.e., C ∗-algebras C together with completely isometric representations (jA, jX ) : (A, X) → C such that C is generated by the image jX (X) of X. We show that there always exists a largest (the universal) C ∗-hull C ∗ u(A, X) and a smallest (the enveloping) C ∗-hull C ∗ e (A, X), using well-known ideas of Kirchberg, Wasser- mann and Hamana and Ruan. As a first hint that the category of C ∗- operator systems is useful, we give in Section 5 a brief discusion of some 4 MASSOUD AMINI, SIEGFRIED ECHTERHOFF, AND HAMED NIKPEY tensor product constructions with C ∗-operator systems. In particular, the spatial tensor products X ⊗ A of an operator X with a C ∗-algebra A natu- rally carries the structure of a C ∗-operator system (1X ⊗ A, X ⊗ A) and if A is not unital, this system is not unital as well! In Section 6 we define the universal crossed products (A, X) ⋊u α G by a continuous action α of a locally compact group G on a C ∗-operator system α G, X ⋊u (A, X) as the pair of completion (A ⋊u α G) of (Cc(G, A), Cc(G, X)) inside the universal crossed product C ∗ u(A, X)⋊α,u G for the canonical action of G on the universal C ∗-hull C ∗ α G is not always isomorphic to the universal C ∗-algebra crossed product A ⋊α,u G (e.g., see part (b) of Remark 6.2). We show that in this setting we get a very satisfying picture of the universal property: every covariant morphism (ϕX , u) of the system (A, X, G, α) "integrates" to a morphism ϕX ⋊ u of (A, X)⋊u α G appears as such integrated form. In Section 7 we add a brief discussion of the spatially defined reduced crossed product (A, X) ⋊r α G and every (non-degenerate) morphism of (A, X)⋊u u(A, X). Note that the C ∗-part A ⋊u α G. In Section 8 we study coactions of groups on C ∗-operator systems and their crossed products. Since locally compact groups are always co-amenable, it is not surprising that the full and reduced crossed products coincide in the sense that the spatially defined crossed product already enjoys the universal properties for covariant representations. In Section 9 we prove versions of the Imai-Takai and Katayama duality theorems for actions and coactions of groups on C ∗-operator systems: starting with an action α : G → Aut(A, X) there are canonical dual coactions bαu and bαr on the universal and reduced crossed products, respectively, such that we get canonical bbα = α ⊗ Ad ρ equivariant isomorphisms (where ρ denotes the right regular representation of G) and (cid:0)A ⋊u (cid:0)A ⋊r α G ⋊ bα bG, X ⋊u α G ⋊cαr bG, X ⋊r α G ⋊ cαu bG(cid:1) ∼=(cid:0)A ⊗ K(L2(G)), X ⊗ K(L2(G))(cid:1) α G ⋊cαr bG(cid:1) ∼=(cid:0)A ⊗ K(L2(G)), X ⊗ K(L2(G))(cid:1). The converse direction, when starting with a coaction δ, known as Katayama's theorem in case of C ∗-algebra crossed products, is a bit more involved, and the full analogue of the Imai-Takai theorem only works under some addi- tional assumptions, like when G is amenable or if everything in sight was defined spatially (i.e., we would consider reduced group algebras and reduced crossed products only). Note that in [25], Chi-Keung Ng proves duality the- orems for spatially defined crossed products in the more general case of (co-)actions by more general quantum groups. C ∗-OPERATOR SYSTEMS AND CROSSED PRODUCTS 5 In Section 10 we come back to group actions on operator spaces V . As indicated above, also in this case it is useful to find a suitable extension of the category of operator spaces since the natural candidate for the crossed product has a canonical structure of a C ∗-algebra bimodule via a left and right action of C ∗ α G. So we found that the right category would be the category of (concrete) C ∗-operator bimodules (A, V, B) which consist of a concrete operator space V ⊆ B(K, H) for some Hilbert spaces H and K together with C ∗-subalgebras A ⊆ B(H) and B ⊆ B(K) such that u(G) on V ⋊u AV = V = V B. Again, we can identify operator spaces V ⊆ B(K, H) with the C ∗-operator bimodule (C1H , V, C1K ). In this way the category of C ∗-operator bimod- If a C ∗-operator bimod- ules extends the category of operator spaces. ule (A, V, B) is given, we get a corresponding Paulsen C ∗-operator system (cid:0)A ⊕ B, X(A, V, B)(cid:1) with X(A, V, B) =(cid:18) A V V ∗ B(cid:19) and a one-to-one correspondence between morphism of (A, V, B) and mor- the above described crossed-product constructions for C ∗-operator systems phisms of (A ⊕ B, X(A, V, B)(cid:1). Thus it is fairly straightforward to apply to the Paulsen systems (cid:0)A ⊕ B, X(A, V, B)(cid:1) to obtain complete analogues of the above described results in this setting. In particular we get complete analogues of the Imai-Takai and Katayama duality theorems. The authors are grateful to Elias Katsoulis and David Blecher for valuable discussions and comments concerning this project and in particular to the content of the preprint [1]. 2. Preliminaries If H is a Hilbert space we denote by B(H) the algebra of bounded oper- ators on H equipped with the operator norm and the canonical involution. A concrete operator space is a closed linear subspace X ⊆ B(H) for some Hilbert space H. If X ⊆ B(H) and Y ⊆ B(K) are two operator spaces, then for each n ∈ N we have the matrix operator spaces Mn(X) ⊆ B(H n) and Mn(Y ) ⊆ B(K n). If ϕ : X → Y is a linear map define ϕn : Mn(X) → Mn(Y ) by ϕn(cid:0)(xij)1≤i,j≤n(cid:1) = (ϕ(xij ))1≤i,j≤n. Then ϕ : X → Y is called completely bounded (or a cb map), if there exist a constant C ≥ 0 such that kϕn(x)kop ≤ Ckxkop 6 MASSOUD AMINI, SIEGFRIED ECHTERHOFF, AND HAMED NIKPEY for all n ∈ N and all x ∈ Mn(X). If C can be chosen to be less or equal to one, we say that ϕ : X → Y is completely contractive and if kϕn(x)kop = kxkop for all n ∈ N and x ∈ Mn(X), we say that ϕ is completely isometric. Suppose now that X = X ∗ and Y = Y ∗ are symmetric closed subspaces of B(H) and B(K), respectively, not necessarily containing the units. We then call a linear map ϕ : X → Y a ccp map (completely contractive and positive) if it satisfies the following conditions: (1) ϕ : X → Y is completely contractive; (2) ϕ(x∗) = ϕ(x)∗ for all x ∈ X; (3) ϕn(x) ≥ 0 for every positive x ∈ Mn(X); where positivity of an element x ∈ Mn(X) (resp. y ∈ Mn(Y )) means that x (resp. y) is a positive element in B(H n) (resp. B(K n)). If, in addition, ϕ is completely isometric, we call it an icp map. Of course it is well known that if X and Y are operator systems (i.e., they contain the units 1H and 1K , respectively), then every unital linear map which satisfies (3) automatically satisfies (1) and (2), hence is a ccp map. As usual, we then say that ϕ : X → Y is a ucp map. 3. C ∗-operator systems and multiplier systems In this section we introduce a category of possibly non-unital operator systems which include C ∗-algebras and classical operator systems as sub- categories. This category will play an important role in our construction of crossed products. Definition 3.1. A (concrete) C ∗-operator system (A, X) on the Hilbert space H is a pair of norm-closed self-adjoint subspaces A ⊆ X ⊆ B(H) such that (1) A is a non-degenerate C ∗-subalgebra of B(H), i.e., AH = H. (2) span{a · x : a ∈ A, x ∈ X} = X (which by an application of Cohen's factorisation theorem is equivalent to X = AX = {ax : a ∈ A, x ∈ X}). A morphism between two C ∗-operator systems (A, X) and (B, Y ) on Hilbert spaces H and K, respectively, consists of a ccp map ψ : X → Y such that (1) ψ(A) ⊆ B, and (2) for all a ∈ A and x ∈ X we have ψ(ax) = ψ(a)ψ(x). A morphism ψ : X → Y is called non-degenerate if ψ(A)Y = Y . We say that the C ∗-operator system (A, X) is unital, if A is unital. Example 3.2. (1) Clearly every classical operator system X ⊆ B(H) can be regarded as a unital C ∗-operator system with respect to the C ∗-OPERATOR SYSTEMS AND CROSSED PRODUCTS 7 C ∗-subalgebra A = C1 ⊆ X. In that case a nonegenerate morphism from (C1, X) to (C1, Y ) is just a ucp map. (2) If (A, X) is a unital C ∗-operator system on the Hilbert space H, then the unit of A coincides with the identity operator on H since A acts non-degenerately on H. Hence X is also a classical operator system. (3) Every non-degenerate C ∗-subalgebra A ⊆ B(H) gives rise to the C ∗-operator system (A, A) on H. (4) Suppose that (A, X) and (B, Y ) are C ∗-operator systems on the Hilbert spaces H and K, respectively. Then the norm-closed tensor product X ⊗ Y ⊆ B(H ⊗ K) contains the minimal tensor product A ⊗ B as a sub-C ∗-algebra such that (A ⊗ B, X ⊗ Y ) becomes a C ∗- operator system on H ⊗ K. In particular, if X ⊆ B(H) is a classical unital operator system and B ⊆ B(K) is a C ∗-algebra, then the minimal tensor product X ⊗ B has the structure of a C ∗-operator system with C ∗-subalgebra B ∼= C1 ⊗ B ⊆ X ⊗ B. This example shows that C ∗-operator systems do appear quite naturally! Definition 3.3. Let (A, X) be a C ∗-operator system on the Hilbert space H. A representation of (A, X) on a Hilbert space K is a ccp map π : X → B(K) such that π(ax) = π(a)π(x) for all a ∈ A and x ∈ X (in particular, πA is a ∗-representation of A). The representation π is called non-degenerate, if πA : A → B(K) is non-degenerate. Definition 3.4. Suppose that X ⊆ B(H) is a self-adjoint norm-closed sub- set of B(H). A norm bounded net (ui)i∈I of self-adjoint elements in X is called an approximate unit for X if for all x ∈ X, i ∈ I we have uix, xui ∈ X and uix, xui → x in the norm of B(H). Lemma 3.5. Suppose that (A, X) is a C ∗-operator system on H. Then every bounded self-adjoint approximate unit (ui)i∈I of A is an approximate unit for X in the sense of the above definition. Moreover, ui → 1H ∗-strongly in B(H). Proof. The first assertion follows immediately from the requirement AX = X for a C ∗-operator system. The second assertion follows from H = AH. (cid:3) Definition 3.6. Suppose that (A, X) is a C ∗-operator system on some Hilbert space H. By a unitization of (A, X) we understand a unital C ∗- operator system ( A, X) on H which contains (A, X) such that the following are satisfied (1) A is an ideal in A and A X ⊆ X. 8 MASSOUD AMINI, SIEGFRIED ECHTERHOFF, AND HAMED NIKPEY (2) If x ∈ X such that ax = 0 for all a ∈ A, then x = 0. It is well known that for a C ∗-algebra A, the multiplier algebra M (A) is the largest unitization of A. We shall now introduce an analogous construc- tion for C ∗-operator systems: Lemma 3.7. Suppose that (A, X) is a C ∗-operator system on the Hilbert space H. Let M (A) = {m ∈ B(H) : mA ∪ Am ⊆ A} be the realisation of the multiplier algebra of A in B(H) and let M (X) = {k ∈ B(H) : kA ∪ Ak ⊆ X} ⊆ B(H). Then (M (A), M (X)) is a unitization of (A, X) in B(H). Moreover, if π : X → B(K) is any non-degenerate ccp representation of (A, X) on some Hilbert space K, then there exists a unique unital extension ¯π : M (X) → B(K) of π as a ccp representation of (M (A), M (X)) on K. Moreover, ¯π is completely isometric iff π is completely isometric. Notation 3.8. We call (M (A), M (X)) the multiplier C ∗-operator system of (A, X). Notice that the space M (X) very much depends on the C ∗-subalgebra A ⊆ X, so that a better notation would probably be to write MA(X) instead of M (X). However, it will always be clear from the context with respect to which C ∗-subalgebra A ⊆ X the set M (X) is defined, so in order to keep notation simple we stick to M (X). Proof of Lemma 3.7. It is trivial to check that (M (A), M (X)) fulfils all properties of a unital C ∗-operator system. Note that M (A)X ⊆ X since X = AX and hence M (A)X = M (A)(AX) = (M (A)A)X = AX = X and similarly XM (A) = X. This easily implies that M (A)M (X) ⊆ M (X). Moreover, if k ∈ M (X) such that ak = 0 for all a ∈ A, then we also have k∗a = 0 for all a ∈ A, hence k∗(AH) = k∗H = {0} which then implies that k∗ = 0. But then k = 0. Thus it follows that (M (A), M (X)) is a unitization of (A, X). Suppose now that π : X → B(K) is a non-degenerate ccp representation of (A, X). Then π(A)K = K and we define the extension ¯π : M (X) → B(K) by ¯π(k)(π(a)ξ) := π(ka)ξ. To see that this is well defined, let let (ui)i∈I be an approximate unit of A consisting of positive elements of norm ≤ 1. Then, if π(a1)ξ1 = π(a2)ξ2 for some elements a1, a2 ∈ A and ξ1, ξ2 ∈ K, we get π(ka1)ξ1 = lim i π(kuia1)ξ1 = lim i π(kui)π(a1)ξ1 = lim i π(kui)π(a2)ξ2 = lim i π(kuia2)ξ2 = π(ka2)ξ2. C ∗-OPERATOR SYSTEMS AND CROSSED PRODUCTS 9 This shows that ¯π is well defined. We now need to check that ¯π(mk) = ¯π(m)¯π(k) for all m ∈ M (A) and k ∈ M (X). We first show that π(ka) = ¯π(k)π(a) for all k ∈ M (X), a ∈ A. To see this let η ∈ H. Then η = π(b)ξ for some b ∈ A, ξ ∈ K. Then π(ka)η = π(kab)ξ = ¯ππ(ab)ξ = ¯ππ(a)π(b)ξ = ¯π(k)π(a)η. Suppose now that m ∈ M (A) and k ∈ M (X). Then, for ξ, η ∈ H and a, b ∈ A we get h¯π(mk)π(a)ξ, π(b)ηi = hπ(b∗)¯π(mk)π(a)ξ, ηi = hπ(b∗)π(mka)ξ, ηi = hπ(b∗mka)ξ, ηi = hπ(b∗m)π(ka)ξ, ηi = hπ(b∗)¯π(m)¯π(k)π(a)ξ, ηi = h¯π(m)¯π(k)π(a)ξ, π(b)ηi which then implies that ¯π(mk) = ¯π(m)¯π(k). We need to show that ¯π : M (X) → B(K) is completely positive. Since it is unital, this will also imply that it is completely contractive. If (A, X) is unital, then M (X) = X and nothing has to be done. If (A, X) is not unital, let (ui)i∈I be an approximate unit of A consisting of positive elements of norm ≤ 1. Since π(A)K = K it follows that π(ui) → 1K ∗-strongly in B(K). Then, if m ∈ M (X), it follows that π(uimui) = π(ui)¯π(m)π(ui) weakly to ¯π(m) in B(K). Now let m ∈ Mn(M (X)) ⊆ B(K n) be any positive element. Let vi := ui ⊗ In ∈ Mn(X). Then vimvi is a positive element of Mn(X) such that πn(vimvi) converges weakly to ¯πn(m) in B(K n). Since weak limits of positive elements are positive, it follows that ¯πn(m) is positive. Finally assume that π : X → B(K) is completely isometric and let (eA, eX) := (π(A), π(X)) denote the image of (A, X) in B(K). Then by the first part of this proof applied to the system (eA, eX) the inverse π−1 : eX → B(H) extends uniquely to a ccp representation ¯π−1 : M (eX) → B(H). Since is the identity on M (X). Similarly, ¯π ◦ ¯π−1 is the identity on M (eX). In ¯π−1 ◦ ¯πX : X → B(H) coincides with the identity on X, it follows from the uniqueness assertion for the extension to M (X) that ¯π−1◦¯π : M (X) → B(H) particular, ¯π : M (X) → B(K) is completely isometric. (cid:3) The following lemma shows that (M (A), M (X)) is the largest unitization of (A, X). 10 MASSOUD AMINI, SIEGFRIED ECHTERHOFF, AND HAMED NIKPEY Lemma 3.9. Let (A, X) be a C ∗-operator system in B(H) and suppose that ( A, X) is a unitization of (A, X) in B(H). Then A ⊆ M (A) and X ⊆ M (X). As a consequence, if π : X → B(K) is a non-degenerate ccp representation of (A, X) on a Hilbert space K, there exists a unique ccp representation π : X → B(K) of ( A, X) which extends π. Proof. Clearly, if ( A, X) is a unitization of (A, X) in B(H), then every m ∈ X multiplies A into X. Hence X ⊆ M (X). Since A is an ideal in A we also have A ⊆ M (A). By Lemma 3.7 we know that π extends uniquely to a representation ¯π of M (X). We then put π := ¯π X. (cid:3) Definition 3.10. Suppose that (A, X) and (B, Y ) are C ∗-operator systems. We say that ϕ : X → M (Y ) is a (non-degenerate) generalized morphism from (A, X) to (B, Y ) if the following holds: (1) ϕ : X → M (Y ) is a morphism from (A, X) to (M (B), M (Y )), and (2) ϕ(A)B = B. Note that since BY = Y , the condition (2) also implies that ϕ(A)Y = Y . Lemma 3.11. Suppose that (A, X) and (B, Y ) are C ∗-operator systems. If ϕ : X → M (Y ) is a non-degenerate generalized morphism from (A, X) to (B, Y ), then there exists a unique extension ¯ϕ : M (X) → M (Y ) of ϕ as a morphism from (M (A), M (X)) to (M (B), M (Y )). In particular, if ϕ : (A, X) → (B, Y ) is a completely positive and completely isomet- ric isomorphism of C ∗-operator systems, the same holds for the extension ¯ϕ : M (X) → M (Y ). Proof. Assume that (B, Y ) and hence (M (B), M (Y )) are C ∗-operator sys- tems on the Hilbert space K. It follows then from the condition that ϕ(A)B = B that ϕ(A)K = ϕ(A)(BK) = BK = K, so ϕ : A → B(K) is a non-degenerate representation of (A, X) on B(K). By Lemma 3.7 we know that there is a unique ccp extension ¯ϕ : M (X) → B(K). We then get ¯ϕ(M (X))B = ¯ϕ(M (X))ϕ(A)B = ϕ(M (X)A)B ⊆ ϕ(X)B ⊆ M (Y )B ⊆ Y, hence ¯ϕ(M (X)) ⊆ M (Y ). A similar argument shows that ¯π(M (A)) ⊆ M (B). For the final statement assume that ϕ : X → Y is a completely isomet- ric isomorphism of the C ∗-operator systems (A, X) and (B, Y ). Let ¯ϕ : M (X) → M (Y ) and ¯ϕ−1 : M (Y ) → M (X) denote the unique extensions of ϕ and ϕ−1 to M (X) and M (Y ), respectively. Then ¯ϕ−1◦ ¯ϕ : M (X) → M (X) extends the identity on X, and hence, by the uniqueness of the extension, must be equal to the identity mal on M (X). Similarly, ¯ϕ◦ ¯ϕ−1 is the identity on M (Y ). (cid:3) C ∗-OPERATOR SYSTEMS AND CROSSED PRODUCTS 11 Corollary 3.12. Let (A, X) be a C ∗-operator system on H and suppose that π : X → B(K) is a completely isometric c.p representation of (A, X) on K. Then the unique extension ¯π : M (X) → B(K) is completely isomeric as well. The same holds for the extension π : X → B(K) for any unitization ( A, X) of (A, X). Proof. If π : X → B(K) is completely isometric, then (π(A), π(X)) is a C ∗-operator system in B(K) and π : X → π(X) is a completely iso- metric isomorphism of C ∗-operator systems. Thus the unique extension ¯π : M (X) → M (π(X)) ⊆ B(K) is completely isometric by Lemma 3.11. (cid:3) Of course there is also a smallest unitization of (A, X): Definition 3.13. Suppose that (A, X) is a C ∗-operator system on a Hilbert space H. Let X 1 = X +C1H and A1 = A+C1H ⊆ X 1. Then (A1, X 1) is the smallest unitization of (A, X) in B(H). We call it the minimal unitization of (A, X). Remark 3.14. Of course, if π : X → B(K) is any ccp representation of (A, X) on a Hilbert space K, then the unique extension π1 : X 1 → B(K) is given by π1(x + λ1H ) = π(x) + λ1K . By Corollary 3.12, if π is completely isomertric, then π1 is completely isometric as well. 4. C ∗-hulls of C ∗-operator systems If (A, X) is a C ∗-operator system, C is a C ∗-algebra, and j : X → C is a completely positive complete isometry such that j(ax) = j(a)j(x) for all a ∈ A, x ∈ X and such X generates C as a C ∗-algebra, then the pair (C, j) is called a C ∗-hull of (A, X). Two C ∗-hulls (C, j) and (C ′, j′) of (A, X) are called equivalent, if there exists a ∗-isomorphism ϕ : C → C ′ such that ϕ ◦ j = j′. In what follows below we want to show that for any C ∗-operator system (A, X) there exist C ∗-hulls (C ∗ env(A, X), jenv) such that for any given C ∗-hull (C, j) of (A, X) there exist unique surjective ∗-homomorphisms u(A, X), ju) and (C ∗ C ∗ u(A, X) ϕu ։ C ϕenv ։ C ∗ env(A, X) u(A, X), ju) and (C ∗ such that ϕu ◦ ju = j and ϕenv ◦ j = jenv. It follows directly from these universal properties of (C ∗ env(A, X), jenv) that they are unique up to equivalence (if they exist). We call (C ∗ u(A, X), ju) the universal C ∗-hull of (A, X) and we call (C ∗ env(A, X), jenv) the enveloping C ∗-algebra of (A, X). Of course, the above notion of the universal C ∗-hull of a C ∗-operator system extends the notion of the universal C ∗-hull of a classical operator system X as introduced by Kirchberg and Wassermann in [22] and the notion of the C ∗-envelope extends the well-known notion of a C ∗-envelope of an 12 MASSOUD AMINI, SIEGFRIED ECHTERHOFF, AND HAMED NIKPEY operator system due to Hamana [14]. Using the ideas of Kirchberg and Wassermann, we now construct the universal C ∗-hull (C ∗ u(A, X), ju). We need: Definition 4.1. Suppose that (A, X) is a C ∗-operator system. A represen- tation π : X → B(K) is called finitely A-generated, if there exists a finite subset {ξ1, . . . , ξl} of K such that K = span{π(A)ξ1, . . . , π(A)ξl}. If κ is the cardinality of a dense subset of A, then every finitely A- generated representation of (A, X) can be regarded, up to unitary equiv- alence, as a representation on a closed subspace of ℓ2(Iκ), where Iκ is a fixed set with cardinality κ. Theorem 4.2. For every C ∗-operator system (A, X) there exists a universal hull (C ∗ u(A, X), ju) for (A, X). Proof. Let κ denote the cardinality of a dense subset of A and let S denote the set of all non-degenerate finitely A-generated ccp representations π : X → B(Hπ) where Hπ is a closed subspace of ℓ2(Iκ). Write HS =Lπ∈S Hπ and πS =Lπ∈S π. We claim that πS : X → B(HS) is a completely isometric representation of (A, X). For this let us assume that (A, X) is represented as a concrete C ∗-operator system on the Hilbert space H. Then for each fixed n ∈ N, x ∈ Mn(X), and ε > 0 we choose a finite rank projection p ∈ B(H) such that k(p ⊗ 1n)x(p ⊗ 1n)k ≥ kxk − ε. Let Hx,ε := span{apH : a ∈ A} and let q : H → Hx,ε denote the orthogonal projection. Define πx,ε : X → B(Hx,ε); πx,ε(y) := qyq for all y ∈ X. Since Hx,ε is an A-invariant subspace of H, we see that q commutes with the elements of A, hence πx,ε(ay) = qayq = qaqyq = πx,ε(a)πx,ε(y) for all a ∈ A, y ∈ X, so πx,ε is a ccp representation of (A, X) on Hx,ε. By construction, πx,ε is finitely A-generated and kπx,ε,n(x)k ≥ kxk − ε. By choosing an isometric embedding of Hx,ε into ℓ2(Iκ) we may assume that πx,ε ∈ S. Since ε is arbitrary, it follows now that πS is completely isometric. u(A, X) as the C ∗-subalgebra of B(HS) generated by We now define C ∗ πS(X) and ju = πS : X → C ∗ u(A, X) ⊆ B(HS). We then have C ∗ u(A, X) ⊆ Yπ∈S B(Hπ) ⊆ B(HS). C ∗-OPERATOR SYSTEMS AND CROSSED PRODUCTS 13 Suppose now that (C, j) is an arbitrary C ∗-hull of (A, X). We may as- sume that C is realised as a non-degenerate subalgebra C ⊆ B(K) for some Hilbert space K. Let C 1 = C + C1K ⊆ B(K) be the unitization of C. Ap- plying the above construction to each element c ∈ Mn(C 1) yields a family of completely positive maps ρc,ε : C 1 → B(Hρc,ε) such that ρc,ε ◦ j : X → B(Hρc,ε) is a finitely A-generated representation of (A, X) and such that kρc,ε,n(c)k ≥ kck − ε. Let S′ denote the set of all such maps ρc,ε. Then ρS ′ = Lρ∈S ′ ρ : C 1 → B(Lρ∈S ′ Hρ) is a unital completely isometric map from C 1 intoQρ∈S ′ B(Hρ). By [9, Theorem 4.1] there exists a unique unital ∗-homomorphism ϕ : C ∗(ρS ′(C 1)) → C 1 such that ϕ ◦ ρS ′ = idC1. Since for each ρ ∈ S′, ρ ◦ j : X → B(Hρ) is a finitely A-generated representation of (A, X), we may identify ρ ◦ j with an element of S via a suitable em- bedding of Hρ ֒→ ℓ2(Iκ). We then obtain a map S′ → S; ρ 7→ ρ ◦ j and a ∗-homomorphism Φ : Qπ∈S B(Hπ) → Qρ∈S ′ B(Hρ) by sending a tupel u(A, X) ⊆ Qπ∈S B(Hπ) (Tπ)π∈S to (Tρ◦j )ρ∈S ′. The restriction of Φ to C ∗ sends C ∗ tion of ∗-homomorphism u(A, X) to C ∗(ρS ′(j(X))) ⊆ C ∗(ρS ′(C 1)). Thus we get a composi- C ∗ u(A, X) Φ −−−−→ C ∗(ρS ′(j(X))) ϕ −−−−→ C such that for ϕu := ϕ ◦ Φ we get ϕu ◦ ju = j. (cid:3) Lemma 4.3. Suppose that π : X → B(K) is a ccp representation of the C ∗-operator system (A, X). Then there exists a unique ∗-homomorphism u(A, X) ։ C ∗(π(X)) ⊆ B(K) such that π ◦ ju = π, where C ∗(π(X)) π : C ∗ denotes the closed C ∗-subalgebra of B(K) generated by π(X). Proof. Suppose that (A, X) is a concrete C ∗-operator system on the Hilbert space H and assume that ι : X ֒→ B(H) is the inclusion map. Then ιL π : X → B(H ⊕ K) is a completely isometric representation and therefore there exists a unique ∗-homomorphism ]ι ⊕ π : C ∗ u(A, X) → C ∗(ι ⊕ π(X)) ⊆ B(H ⊗ K). As ι ⊕ π(X) ⊆ X ⊕ π(X) ⊆ B(H) ⊕ B(K), we obtain a well defined ∗-homomorphism C ∗(ι⊕ π(X)) → C ∗(π(X)) given by T 7→ PK T PK, where PK : H ⊕ K → K denotes the orthogonal projection. Thus π = ]ι ⊕ π(·)PK will do the job. The uniqueness follows from the fact that PK C ∗ (cid:3) u(A, X) is generated by ju(X). At this point it is convenient to consider representations of C ∗-operator systems on multiplier algebras: Definition 4.4. Suppose that (A, X) is a C ∗-operator system and let D be a C ∗-algebra. A representation of (A, X) into the multiplier algebra M (D) is a ccp map Φ : X → M (D) such that Φ(ax) = Φ(a)Φ(x) for all 14 MASSOUD AMINI, SIEGFRIED ECHTERHOFF, AND HAMED NIKPEY a ∈ A, x ∈ X. We then say that Φ is non-degenerate if the restriction of Φ to A is a non-degenerate ∗-homomorphism, i.e., if Φ(A)D = D. 1 Remark 4.5. Note that every (non-degenerate) representation of a C ∗-operator system (A, X) on a Hilbert space H an be regarded as a (non-degenerate) representation into M (K(H)) = B(H). Conversely, if Φ : X → M (D) is a representation of (A, X) in M (D) and if D (and hence M (D)) is rep- resented faithfully on the Hilbert space H, then Φ can also be regarded as a representation of (A, X) on H which is non-degenerate if and only if Φ : X → M (D) is non-degenerate. But it is often more convenient to work with representations into M (D). We now get Proposition 4.6. Let (A, X) be a C ∗-operator system and let D be a C ∗- algebra. Then there is a one-to-one correspondence between (1) The non-degenerate representations of (A, X) into M (D). (2) The non-degenerate ∗-homomorphisms of C ∗ u(A, X) into M (D). If Φ : C ∗ X → M (D) gives the corresponding representation of (A, X) as in (1). u(A, X) → M (D) is as in (2), then the restriction ΦX = Φ ◦ ju : Proof. It clearly suffices to show that every non-degenerate representation of (A, X) on M (D) extends to a representation of C ∗ u(A, X). But representing C ∗ u(A, X) faithfully on a Hilbert space Hu, say, this follows easily from Lemma 4.3. (cid:3) We now proceed with a discussion of the enveloping C ∗-hull for (A, X). For this recall that an operator space V is injective if, given operator spaces W1 ⊆ W2, any completely bounded linear map ϕ1 : W1 → V can be extended to a completely bounded linear map ϕ2 : W2 → V with kϕ2kcb = kϕ1kcb. The algebra B(H) is known to be an injective operator space [37]. Hamana in [14, 15] and Ruan in [34] independently showed that for any operator space V in B(H), there is a unique minimal injective operator subspace I(V ) of B(H) containing V . It is called the injective envelope of V and enjoys the following fundamental property, which we shall use heavily throughout this paper (e.g., see [34, §5]): Proposition 4.7. Let V ⊆ B(H) be an operator space. Then every com- pletely contractive map ψ : I(V ) → I(V ) which restricts to the identity on V is the identity on I(V ). We need the following result of Choi-Effros [8] (see §6 in [13] and partic- ularly [13, Theorem 6.1.3]). 1Note that by Cohen's factorization theorem to have Φ(A)D = D it suffices to have that span{Φ(A)D} = D. C ∗-OPERATOR SYSTEMS AND CROSSED PRODUCTS 15 Theorem 4.8. If I ⊆ B(H) is an injective operator system, then there is a unique multiplication ◦ : I × I → I making I a unital C ∗-algebra with its given ∗-operation and norm and identity 1H. The multiplication is given by x ·ϕ y = ϕ(xy), where ϕ : B(H) → I is a fixed ccp onto projection. Using these results we now show Proposition 4.9. Suppose that (A, X) is a C ∗-operator system. Then there exists an enveloping C ∗-hull (C ∗ env(A, X), jenv) of (A, X). Proof. Suppose that (A, X) is a C ∗-operator system on H. Let (A1, X 1) be the unitization of (A, X) as in Definition 3.13. By Theorem 4.8 the injective envelope I(X 1) of the unital operator system X 1 is a unital C ∗-algebra with multiplication x ·ϕ y = ϕ(xy) for some fixed ccp onto projection ϕ : B(H) → I(X 1). Now, for each a ∈ A and x ∈ X we have ax ∈ X and therefore a ·ϕ x = ϕ(ax) = ax. Therefore the inclusion map X ֒→ B(H) induces a completely isometric embedding j : X → I(X 1) such that j(ax) = j(a)j(x) env(A, X) to be the C ∗-subalgebra of I(X 1) for all a ∈ A, x ∈ X. Define C ∗ generated by j(X) and we let jenv = j : X ֒→ C ∗ env(A, X) denote the env(A, X)1 is inclusion map. Note that by construction, the unitization C ∗ env(X 1) of the unital operator system X 1 just the enveloping C ∗-algebra C ∗ in the sense of Hamana [14]. To see that (C ∗ env(A, X), jenv) satisfies the universal property let (C, j) be any given C ∗-hull of (A, X). Choose a non-degenerate embedding C ֒→ B(K) for some Hilbert space K and let C 1 = C + C1K ⊆ B(K). Then j1 : X 1 → C 1 is a completely isometric embedding of the operator system X 1. It follows therefore from the universal property of the enveloping C ∗- env(X 1) (see [14]) that there exists a ∗-homomorphism ϕ : C 1 → algebra C ∗ env(A, X)1 which intertwines the inclusions of X 1 into these C ∗ algebras. Restricting ϕ to C ⊆ C 1 then gives the desired ∗-homomorphism ϕenv : C → C ∗ (cid:3) env(X 1) = C ∗ env(A, X). We close this section with the following useful result: Lemma 4.10. Let (C, j) be any C ∗-hull of the C ∗-operator system (A, X). Then the inclusion map j : (A, X) → C extends to a completely isometric inclusion Moreover, we have ¯j(M (A)) ∩ C = j(A). ¯j :(cid:0)M (A), M (X)(cid:1) → M (C). Proof. Since A contains an approximate identity of X, and since C is gener- ated by j(X) as a C ∗-algebra, it follows that j(A) contains an approximate 16 MASSOUD AMINI, SIEGFRIED ECHTERHOFF, AND HAMED NIKPEY It follows that j : (A, X) → C ⊆ M (C) is a completely identity of C. isometric non-degenerate (generalized) morphism, where we identify C with the C ∗-operator system (C, C). The first assertion then follows from Lemma 3.11. To see the second assertion let c ∈ C such that j(A)c ⊆ j(A) ⊆ M (C). Let (ai)i∈I be an approximate unit in A. Then (ai)i∈I is also an approximate unit in X, and (j(ai))i∈I is an approximate unit in C. But then it follows that c = limi j(ai)c ∈ j(A). (cid:3) 5. Tensor products In this section we want to give a brief discussion on certain tensor product constructions of C ∗-operator systems. In particular we want to discuss ana- logues of the commutative maximal tensor product S ⊗c T of two operator systems S and T as introduced in [21] and of the minimal (or spacial) tensor product. Definition 5.1. Suppose that (A, X) and (B, Y ) are C ∗-operator systems. Let A ⊗c B and X ⊗c Y denote the closures of the algebraic tensor products A ⊙ B and X ⊙ Y inside the maximal C ∗-tensor product C ∗ u(A, X) ⊗max C ∗ u(B, Y ), respectively. Then (A ⊗c B, X ⊗c Y ) is a C ∗-operator system which we call the commuting universal tensor product of (A, X) with (B, Y ). Lemma 5.2. There are unique completely isometric generalized morphisms iX : X → M (X ⊗c Y ) and iY : Y → M (X ⊗c Y ) such that iX (x)iX (y) = x ⊗ y ∈ X ⊗c Y for all x ∈ X, y ∈ Y . u(A, X) ⊗max C ∗ u(A,X) : C ∗ u(A,X)(c)iC ∗ u(B,Y ) : C ∗ u(B, Y ), respectively. Let iX and iY denote the restrictions of iC ∗ u(A, X) → M (D) and iC ∗ u(B,Y )(d) = c ⊗ d for all c ∈ C ∗ Proof. Write D := C ∗ u(B, Y ) and assume that D is repre- sented faithfully and non-degenerately on a Hilbert space K, say. By the properties of the maximal tensor product of C ∗-algebras, there are isometric u(B, Y ) → ∗-homomorphisms iC ∗ u(A, X), d ∈ M (D) such that iC ∗ C ∗ u(A,X) and iC ∗ u(B,Y ) to X and Y , respectively. Then iX and iY are completely iso- metric representation of (A, X) and (B, Y ) into M (D) ⊆ B(K) such that iX(x)iY (y) = x ⊗ y for all x ∈ X, y ∈ Y , if we regard the algebraic ten- sor product X ⊙ Y as a subspace of X ⊗c Y . So all we need to check is that iX and iY have image in M (X ⊗c Y ), which follows easily from iX(x)(a ⊗ b) = iX(x)iX (a)iY (b) = iX (xa)iY (b) = xa ⊗ b ∈ X ⊗c Y hence iX(X)(A ⊗c B) ⊆ X ⊗c B and, similarly, iY (Y )(A ⊗c B) ⊆ A ⊗c Y , where A ⊗c B, A ⊗c Y, X ⊗c B are defined as the closures of the respective algebraic tensor products in X ⊗c Y . (cid:3) C ∗-OPERATOR SYSTEMS AND CROSSED PRODUCTS 17 Lemma 5.3. The tensor product (A⊗cB, X⊗cY ) has the following universal property: whenever (ϕX , ϕY ) is a pair of non-degenerate ccp representations ϕX : X → M (D), ϕY : Y → M (D) of (A, X) and (B, Y ) into the multiplier algebra M (D) for some C ∗-algebra D such that ϕX (x)ϕY (y) = ϕY (y)ϕX (x) for all x ∈ X and y ∈ Y , then there exists a unique ccp representation ϕ = ϕX ⋊ ϕY : X ⊗c Y → M (D) of (A ⊗c B, X ⊗c Y )such that ϕ(x ⊗ y) = ϕX (x)ϕY (y) for all x ∈ X, y ∈ Y . Remark 5.4. If H is a Hilbert space and D = K(H), we obtain a version of the above lemma for non-degenerate ccp representations on Hilbert space. u(A, X) → B(H) and ϕY : C ∗ u(A, X) and C ∗ Proof of Lemma 5.3. It follows from Proposition 4.6 that there exist unique ∗-homomorphisms ϕX : C ∗ u(B, Y ) → B(H) such that ϕX ◦ ju = ϕX and ϕY ◦ ju = ϕY . Since ϕX (x) commutes with ϕY (y) for all x ∈ X, y ∈ Y , and since C ∗ u(B, Y ) are generated by ju(X) and ju(Y ), respectively, it follows that the ranges of ϕX and ϕY commute as well. Therefore, by the universal properties of the maximal tensor product, there exists a (unique) ∗-homomorphism ϕ : C ∗ u(A, X) ⊗max C ∗ u(A, X) and d ∈ C ∗ It is then easily checked on elementary tensors that ϕ(cz) = ϕ(c)ϕ(z) for all c ∈ A ⊗c B and z ∈ X ⊗c Y . (cid:3) u(B, Y ) → B(H) such that ϕ(c ⊗ d) = ϕX(c) ϕY (d) for all c ∈ C ∗ u(B, Y ). Remark 5.5. It is an interesting question, whether there exists a converse of the above lemma, i.e., whether every non-degenerate ccp representation π : X ⊗c Y → B(K) can be realised as π = πX ⋊ πY for a pair of representa- tions (πX, πY ) as in the lemma. Indeed, this is only true if the representation π preserves some more of the multiplicativity structure of X ⊗c Y , which is not directly part of the structure of (A⊗cB, X⊗cY ) as a C ∗-operator system. Realised as a subspace of D := C ∗ u(B, Y ), we see that an ele- mentary tensor x ⊗ y can be written as a product iX (x)iY (y) = iY (y)iX (x), where iX , iY are the canonical inclusions of X and Y into M (X ⊗c Y ) as in Lemma 5.2. The representations constructed in Lemma 5.3 are precisely those whose extension ¯π to M (X ⊗c Y ) preserves these relations: If it does, then πX = ¯π ◦ iX and πY = ¯π ◦ iY , satisfy the conditions of the lemma such that π = πX ⋊ πY . u(A, X) ⊗max C ∗ But we believe that a general ccp representation π : X ⊗c Y → B(K) does not need to satisfy these relations. But, as we see below, it does if Y = B is a C ∗-algebra. Lemma 5.6. Suppose that (A, X) is a C ∗-operator system and B is a C ∗- algebra (viewed as the C ∗-operator system (B, B)). Let ϕ : X ⊗c B → M (D) 18 MASSOUD AMINI, SIEGFRIED ECHTERHOFF, AND HAMED NIKPEY be any non-degenerate ccp representation of (A ⊗c B, X ⊗c B). Then there is a unique ccp representation ϕX : X → M (B) and a ∗-representation ϕB : B → M (B) such that ϕ = ϕX ⋊ ϕB. A similar statement holds for the tensor product (B ⊗c A, B ⊗c X). Proof. Let ϕX = ¯ϕ ◦ iX and ϕB = ¯ϕ ◦ iB as in the above remark. Note that iX maps X into M (X ⊗c Y ) but iB maps B into M (A ⊗c B) ⊆ M (X ⊗c B), since (a⊗ b)iB(c) = a⊗ bc ∈ A⊗c B for all a⊗ b ∈ A⊙ B. Since the extension ¯ϕ : M (X ⊗c B) → B(K) is a unital c.p representation of the C ∗-operator system (M (A ⊗c B), M (X ⊗c B)), we get ϕX (x)ϕB(b) = ¯ϕ(iX (x)) ¯ϕ(iB(b)) = ¯ϕ(iX (x)iB(b)) = ¯ϕ(x ⊗ b) and similarly ϕB(b)ϕX (x) = ¯ϕ(x ⊗ b). It is then clear that ϕ = ϕX ⋊ ϕB as in Lemma 5.3. (cid:3) Definition 5.7. Suppose that (A, X) and (B, Y ) are concrete C ∗-operator systems on the Hilbert spaces H and K, respectively. Then we define the spacial tensor product (A ⊗B, X ⊗Y ) via the closures of the algebraic tensor products A ⊙ B and X ⊙ Y in B(H ⊗ K). It is well known that the spatial tensor product does not depend, up to isomorphism, on the particular embeddings of X in B(H) and Y in B(K). Let is Y (y) = 1H ⊗ y denote the canonical embeddings of X and Y into B(H ⊗ K). It then follows from Lemma 5.3 that there exists a canonical surjective morphism X : X → B(H ⊗ K), is Y : Y → B(H ⊗ K); is X (x) = x ⊗ 1K and is Φ := iX × iY : X ⊗c Y → X ⊗Y from (A ⊗c B, X ⊗c Y ) onto (A ⊗B, X ⊗Y ) . The following proposition is now an easy consequence of our constructions: Proposition 5.8. Suppose that (A, X) is a C ∗-operator system. Then for any nuclear C ∗-algebra B, the canonical morphism (A ⊗c B, X ⊗c B) onto (A ⊗B, X ⊗B) is an isomorphism (and similarly for (B ⊗c A, B ⊗c X)). Proof. If B is nuclear, then C ∗ sult then follows from representing C ∗ Hilbert space H. u(A, X) ⊗max B = C ∗ u(A, X) ⊗B. The re- u(A, X) (and hence X) faithfully on a (cid:3) For later use we also need to consider morphisms into the multiplier C ∗-operator systems of tensor products of C ∗-operator systems with C ∗- algebras. This is the special case of the above constructions if one of the factors is a pair (C, C) for a C ∗-algebra C. Note that in this case we also have C ∗ u(C, C) = C. C ∗-OPERATOR SYSTEMS AND CROSSED PRODUCTS 19 Lemma 5.9. Suppose that (A, X) and (B, Y ) are C ∗-operator systems and let C and D be C ∗-algebras. Let ϕX : X → M (Y ) be a non-degenerate generalized homomorphism from (A, X) into (M (B), M (Y )) and let ϕC : C → M (D) be a non-degenerate generalized homomorphism from C to D. Then there exists a unique non-degenerate generalized homomorphism ϕ ⊗c ψ : X ⊗c C → M (Y ⊗c D) (resp. ϕ ⊗ψ : X ⊗C → M (Y ⊗D)) such that ϕ ⊗c ψ(x ⊗ c) = ϕ(x) ⊗ ψ(c) (resp. ϕ ⊗ψ(x ⊗ c) = ϕ(x) ⊗ ψ(c)) for all elementary tensors x ⊗ c ∈ X ⊙ C. Proof. Let π : Y ⊗c D → B(H) be a non-degenerate completely isometric representation of (B ⊗c D, Y ⊗c D) on the Hilbert space H, By Lemma 5.6 there are non-degenerate representations πY : Y → B(H) and πD : D → B(H) such that π = πY ⊗c πD. Let ψX := ¯πY ◦ϕX and ψC := ¯πD ◦ϕA, where ¯πY and ¯πD denote the unique extensions of πY , πD to M (Y ) and M (D) as in Lemma 3.11. Note that for all m ∈ M (Y ), n ∈ M (D) we have ¯πY (m)¯πD(n) = ¯πD(n)¯πY (m). Indeed, this follows from the fact that π(B ⊙ D)H is dense in H (since π is non-degenerate) and for all b ∈ B and d ∈ D, we have ¯πY (m)¯πD(n)π(b ⊗ d) = ¯πY (m)¯πD(n)πY (b)πD(d) = ¯πY (m)¯πD(n)πD(d)πY (b) = ¯πY (m)πD(nd)πY (b) = ¯πY (m)πY (b)πD(nd) = πY (mb)πD(nd) = πD(nd)πY (mb) = ¯πD(n)πD(d)πY (mb) = ¯πD(n)πY (mb)πD(d) = ¯πD(n)¯πY (m)πY (b)πD(d) = ¯πD(n)¯πY (m)π(b ⊗ d). It follows from this that ψX (x)ψC (c) = ψC(c)ψX (x) for all x ∈ X, c ∈ C. Thus, it follows from Lemma 5.3 that there exists a unique non-degenerate ccp representation ψ := ψX ⋊ ψC : X ⊗c C → B(H) given on elementary tensors by ψ(x ⊗ c) = ψX (x)ψC (c). Now, by the construction of the multiplier system as in Lemma 6.9 we may identify M (Y ⊗c D) with its image ¯π(M (Y ⊗c Y )) ⊆ B(H). Using this identification, we want to check that ψ takes values in M (Y ⊗c Y ) ∼= {m ∈ B(H) : mπ(B ⊗c D), π(B ⊗c D)m ⊆ π(Y ⊗c D). For this let x ⊗ c ∈ X ⊙ C be any elementary tensor and let b ⊗ d ∈ B ⊙ D. Then ψ(x ⊗ c)π(b ⊗ d) = ¯πY (ϕX (x))¯πD(ϕC (c))πY (b)πD(d) = ¯πY (ϕX (x))πD(ϕC (c)d)πY (b) = πY (ϕX (x)b)πD(ϕC (c)d) = π(ϕX (x)b ⊗ ϕC(c)d) ∈ π(Y ⊗c D). 20 MASSOUD AMINI, SIEGFRIED ECHTERHOFF, AND HAMED NIKPEY Hence ψ(X ⊗c C)π(B ⊗c D) ⊆ π(Y ⊗c D) and the inclusion π(B ⊗c D)ψ(X ⊗c C) ⊆ π(Y ⊗c D) follows similarly. (cid:3) 6. Universal crossed products by group actionns For a C ∗-operator system (A, X) let Aut(A, X) denote the group of all invertible morphisms α : (A, X) → (A, X). A strongly continuous action of the locally compact group G on the C ∗-operator system (A, X) is a homo- morphism α : G → Aut(A, X); g 7→ αg such that g 7→ αg(x) is continuous for all x ∈ X. It follows directly from the universal property of (C ∗ u(A, X), ju) that every automorphism of (A, X) extends to a unique automorphisms αu of C ∗ u(A, X). Since C ∗ u(A, X) is generated by a copy of X, any strongly continuous action α : G → Aut(A, X) extends to unique strongly continuous action αu : G → Aut(C ∗ u(A, X)). This leads to the following definition of the universal crossed product by an action α of G on (A, X). Definition 6.1. Let α : G → Aut(A, X) be an action as above. We define the universal (or full) crossed product (A, X)⋊u αG for the action α as the pair (A⋊u α G are the closures of Cc(G, A) and Cc(G, X) inside the universal C ∗-algebra crossed product C ∗ u(A, X) ⋊α,u G. α G), where A⋊u α G, X ⋊u α G and X ⋊u u(A, A) = A, and therefore the crossed product (A, A) ⋊u Remark 6.2. (a) If α : G → Aut(A) is an action of G on the C ∗-algebra A, and if we consider the corresponding C ∗-operator system (A, A), then C ∗ α G is given by the pair (A ⋊α,u G, A ⋊α,u G). Thus, the universal crossed product construction for C ∗-operator systems extends the well-known universal crossed product constructions for C ∗-algebras. α G, X ⋊u α G) the C ∗-algebra A ⋊u (b) In general it is not true that that in the crossed product (A, X) ⋊u α G = (A ⋊u α G coincides with the universal C ∗-algebra crossed product A ⋊α,u G. To see this let G be any (second countable) non-amenable exact group. Then it follows from [6] that there exists an amenable compact G-space Ω, which implies that the full and reduced crossed products of G by C(Ω) coincide. Now choose a faithful and non-degenerate representation of C(Ω) into B(H) for some Hilbert space H and consider X := C(Ω) ⊆ B(H) as an operator system (forgetting the multiplicative structure). As in Example 3.2 we regard this as the C ∗- u(X) denote the enveloping C ∗-algebra of X. operator system (C, X). Let C ∗ We then get completely isometric embeddings C ֒→ X ֒→ C ∗ u(X) which give rise to ccp maps between the full crossed products (in the C ∗-algebra sense) (6.1) C ∗(G) = C ⋊u G → C(Ω) ⋊u G → C ∗ u(X) ⋊u G. C ∗-OPERATOR SYSTEMS AND CROSSED PRODUCTS 21 By definition of the crossed product (C, C(X)) ⋊αu G = (C ⋊u α G), X ⋊u α G is identical to the image of C(Ω) ⋊u G under the second map and C ⋊u α G coincides with the image of C ∗(G) under the composition in (6.1). But since C(Ω) ⋊u G = C(Ω) ⋊r G by amenability of the action of G on Ω the first map in (6.1) factors through the reduced group algebra C ∗ r (G). We therefore also have C ⋊u r (G) 6= C ∗(G) = C ⋊u G. α G ∼= C ∗ α G, X ⋊u In what follows we want to show that non-degenerate representations of the full crossed product are in one-to-one relation to the non-degenerate covariant representations of the system (A, X, G, α) as in Definition 6.3. Suppose that α : G → Aut(A, X) is an action of G on the C ∗-operator system (A, X). A covariant representation of (A, X, G, α) is a pair (π, u), where π : X → B(Hπ) is a ccp representation of (A, X) on B(Hπ) and u : G → U (Hπ) is a strongly continuous unitary representation of G such that π(αg(x)) = Ugπ(x)U ∗ g ∀x ∈ X, g ∈ G. Remark 6.4. Suppose that ρ : X → B(K) is any ccp representation of (A, X) on B(K). Then we can construct a covariant representation Ind ρ = (ρ, 1K ⊗ λ) on B(K ⊗ L2(G)) by the usual formula (cid:0)ρ(x)ξ(cid:1)(g) = ρ(αg−1(x))ξ(g) and (1 ⊗ λ)hξ(g) = ξ(h−1g) for ξ ∈ L2(G, K) ∼= K ⊗ L2(G), x ∈ X, and g, h ∈ G. Observe that if ρ is completely isometric, then so is ρ. Hence there exist covariant repre- sentations (π, u) of (A, X, G, α) in which the representation π is completely isometric. It is actually useful to extend the notion of a covariant representation to allow representations into multiplier systems as in Definition 6.5. Suppose that α : G → Aut(A, X) is an action of G on the C ∗-operator system (A, X) and suppose that (B, Y ) is a C ∗-operator sys- tem. By a non-degenerate covariant homomorphism of (A, X, G, α) into the multiplier system (M (B), M (Y )) of (B, Y ) we understand a pair of maps (ϕ, u), where ϕ : X → M (Y ) is a non-degenerate generalized morphism from (A, X) to (B, Y ) and u : G → U M (B) is a strictly continuous homo- morphism such that for all x ∈ X and g ∈ G. ϕ(αg(x)) = ugϕ(x)u∗ g Remark 6.6. Note that if (B, Y ) is represented completely isometrically and non-degenerately on a Hilbert space K, then a non-degenerate covariant ho- momorphism of (A, X, G, α) into (M (B), M (Y )) turns into a non-degenerate 22 MASSOUD AMINI, SIEGFRIED ECHTERHOFF, AND HAMED NIKPEY covariant representation of (A, X, G, α) on K. But being a representation into (M (B), M (Y )) requires some additional structure of how the image interacts with (B, Y ). Example 6.7. If α : G → Aut(A, X) is an action of G on the C ∗-operator system (A, X) we can define a canonical non-degenerate covariant homo- morphism (ΛX , ΛG) of (A, X, G, α) into M (X ⊗K(L2(G))) as follows: We first define a non-degenerate representation ΛX : X → M (X ⊗ K(L2(G))) of (A, X) by the composition of maps X α−→ M (X ⊗C0(G)) idX ⊗M −→ M (X ⊗K(L2(G))), where α : X → Cb(G, X) ⊆ M (X ⊗C0(G)) sends the element x ∈ X to the bounded continuous function g 7→ αg−1(x) and where M : C0(G) → B(L2(G)) is the representation of C0(G) by multiplication operators. More- over, we define ΛG : G → M (X ⊗K(L2(G))) by ΛG(g) = 1 ⊗ λg, where λ : G → U (L2(G)) is the regular representation of G. We leave it to the reader to check that (ΛX , ΛG) satisfies the covariance condition. We call (ΛX, ΛG) the regular representation of (A, X, G, α). Note that this construction directly extends the construction of the regu- lar representation (ΛB, ΛG) of a C ∗-dynamical system (B, G, β) into M (B ⊗ K(L2(G))). In particular, the restriction (ΛX A.ΛG) of (ΛX , ΛG) to (A, G, α) coincides with the regular representation of (A, G, α). Note also that if ρ : X → B(K) is any ccp representation of (A, X), we recover the representation Ind ρ = (ρ, 1⊗λ) of Remark 6.4 as the composition (ρ ⊗ idK(L2(G)) ◦ (ΛX , ΛG). Proposition 6.8. For each non-degenerate covariant representation (ϕ, u) of (A, X, G, α) into (M (B), M (Y )) there exists a unique generalized homo- morphism ϕ⋊u : X ⋊u α G) to (M (B), M (Y )) given on f ∈ Cc(G, X) by α G → M (Y ) from (A⋊u α G, X ⋊u ϕ ⋊ u(f ) =ZG ϕ(f (g))ug dg. Proof. Let (ϕ, u) be given and let (B, Y ) be represented completely iso- metrically on a Hilbert space K. By Lemma 4.3 there exists a unique ∗- u(A, X) → C ∗(ϕ(X)) ⊆ B(K) which extends ϕ. Apply- representation ϕ : C ∗ ing this fact to the representations ϕ ◦ αg = Ad ug ◦ ϕ shows that ( ϕ, u) is a u(A, X), G, αu) into covariant representation of the C ∗-dynamical system (C ∗ B(K). It therefore integrates to a ∗-representation ϕ ⋊ u : C ∗ u(A, X) ⋊αu,u G → B(K) given on f ∈ Cc(G, C ∗ u(A, X)) by the integral formula in the lemma. The restriction of ϕ ⋊ u to X ⋊u α G is then the desired represen- tation ϕ ⋊ u. To see that it maps into M (Y ), we only need to check that (6.2) (cid:0)ϕ ⋊ u(f )(cid:1)b =ZG f (g)ugb dg C ∗-OPERATOR SYSTEMS AND CROSSED PRODUCTS 23 (cid:0)ϕ ⋊ u(f )(cid:1)b, b(cid:0)ϕ ⋊ u(f )(cid:1) ∈ Y for all b ∈ B and f ∈ Cc(G, X). But the integration formula gives Since ug ∈ M (B), we have ugb ∈ B and hence f (g)ugb ∈ Y for all g ∈ G. Thus the integral (6.2) gives an element in Y . A similar argument shows (cid:3) that to b(cid:0)ϕ ⋊ u(f )(cid:1) ∈ Y . Lemma 6.9. Suppose that α : G → Aut(A, X) is an action. Then there is a canonical covariant morphism (iX , iG) : (A, X, G, α) →(cid:0)M (A ⋊u α G), M (X ⋊u α G)(cid:1) such that for each x ∈ X, f1 ∈ Cc(G, A), f2 ∈ Cc(G, X) and g, h ∈ G, we have (iX (x)f1)(g) = xf1(g) (f1iX(x))(g) = f1(g)αg(x) and (iG(h)f2)(g) = αh(f2(h−1g)) (f2iG(h))(g) = f2(gh−1)∆(h−1). Moreover, the integrated form iX ⋊iG : X ⋊u map on X ⋊u α G. α G → M (X ⋊u α G) is the identity Proof. Suppose that C ∗ generately on a Hilbert-space Hu. Then the restriction to X ⋊u completely isometric representation of (A, X) ⋊u universal properties of the maximal crossed product C ∗ exists a unique covariant homomorphism (iC ∗ into M (C ∗ G and f ∈ Cc(G, C ∗ u(A, X) ⋊αu,u G is represented faithfully and nonde- α G gives a α G on Hu as well. By the u(A, X) ⋊αu ,u G there u(A, X), G, αu) u(A, X), g, h ∈ u(A,X), iG) of (C ∗ u(A, X) ⋊αu,u G) ⊆ B(Hu) which is given for b ∈ C ∗ u(A, X)) by the formulas (iC ∗ u(A,X)(b)f )(g) = bf (g) (f iC ∗ u(A,X)(b))(g) = f (g)αg(b) and (iG(h)f )(g) = αh(f (h−1g)) (f ig(h))(g) = f (gh−1)∆(h−1), and such that the integrated form iC ∗ representation of the crossed product on Hu. Let iX = iC ∗ ju : X → C ∗ u(A,X) ⋊ iG coincides with the original u(A,X) ◦ ju, where u(A, X) denotes the embedding. Then for each f ∈ Cc(G, X) the integral iX ⋊ iG(f ) = RG iX(f (g))iG(g) dg coincides with the inclusion u(A, X) ⋊αu,u G, and therefore extends to α G ⊆ C ∗ of f ∈ Cc(G, X) ֒→ X ⋊u the identity on X ⋊u α G. Thus we only need to check that (iX , iG) is a non-degenerate covariant morphism into (cid:0)M (A ⋊u h ∈ G, then g 7→ (iG(h)f )(g) = αh(f (h−1g)) lies in Cc(G, A) as well, hence α G), M (X ⋊u α G)(cid:1). First, if f ∈ Cc(G, A) and 24 MASSOUD AMINI, SIEGFRIED ECHTERHOFF, AND HAMED NIKPEY iG(h)(cid:0)iX ⋊ iG(f )(cid:1) = iX ⋊ iG([g 7→ αh(f (h−1g))]) ∈ A ⋊u α G, which shows that iG takes its values in M (A⋊u α G). Moreover, if f ∈ Cc(G, A) and x ∈ X, then [g 7→ (iX (x)f )(g) = xf (g)] ∈ Cc(G, X), and hence iX (x)(iX ⋊ iG(f )) = iX ⋊ iG([g 7→ xf (g)]) ∈ X ⋊u α G). This completes the proof. (cid:3) α G, which implies that iX (x) ∈ M (X ⋊u We are now ready for the converse of Proposition 6.8. Proposition 6.10. Suppose that α : G → Aut(A, X) is an action. Then for each non-degenerate generalized morphism Φ : X ⋊u α G → M (Y ) from (A ⋊u α G) to (B, Y ) there exists a unique non-degenerate covariant morphism (ϕ, u) of (A, X, G, α) to (M (B), M (Y )) such that Φ = ϕ ⋊ u. α G, X ⋊u More precisely, if (iX , iG) is the covariant morphism of (A, X, G, α) into α G) → M (Y ) α G)) as in Lemma 6.9 and if ¯Φ : M (X ⋊u (M (A ⋊u denotes the unique extension of Φ as in Lemma 3.11, then α G), M (X ⋊u ϕ = ¯Φ ◦ iX and u = ¯Φ ◦ iG. Proof. It is straightforward to check on functions in Cc(G, X) that (ϕ, u) is a covariant morphism of (A, X, G, α) into (M (B), M (Y )) such that ϕ ⋊ u = ¯Φ ◦ (iX ⋊ iG) = ¯Φ ◦ idX⋊uG = Φ. (cid:3) Note that we may regard covariant morphism of (A, X, G, α) into M (B) for any C ∗-algebra B as the special case of covariant morphisms into the multiplier system of the C ∗-operator system (B, B). Similarly, we may re- gard covariant representations on a Hilbert space H as the special case in which B = K(H). Thus Proposition 6.8 will give us Corollary 6.11. Suppose that (ϕ, u) is a covariant morphism of (A, X, G, α) into M (D) (resp. covariant represention into B(K) for a Hilbert space K). Then there exist an integrated form ϕ ⋊ u : X ⋊u α G → M (D) (resp. B(K)) given for f ∈ Cc(G, X) by ϕ ⋊ u(f ) =RG ϕ(f (g))ug dg. Conversely, if Φ : X ⋊u α G → M (D) (resp. B(K)) is any non-degenerate homomorphism (resp. representation) of (A ⋊u α G), then there exists a unique non-degenerate covariant homomorphism (resp. representation) of (A, X, G, α) into M (D) (resp. B(K)) such that Φ = ϕ ⋊ u. Indeed, if (iX , iG) are as in Lemma 6.9 and ¯Φ denotes the unique extension of Φ to M (X ⋊u G), then α G, X ⋊u ϕ = ¯Φ ◦ iX and u = ¯Φ ◦ iG. C ∗-OPERATOR SYSTEMS AND CROSSED PRODUCTS 25 Corollary 6.12. Let (A, X, G, α) be a G-C ∗-operator system. Then C ∗ u(A, X) ⋊α,u G = C ∗ u(A ⋊u α G, X ⋊u α G). α G ⊆ C ∗ α G, X ⋊u u(A, X) ⋊αu,u G is generated by X ⋊u α G) extends to a ∗-homomorphism Ψ : C ∗ Proof. Since C ∗ u(A, X) is generated by X, it is fairly straightforward to check that C ∗ u(A, X) ⋊αu,u G. So all we need to show is that any isometric representation ψ : X ⋊u α G ֒→ B(H) of (A ⋊u u(A, X) ⋊αu,u G → B(H). Passing to the closed subspace ϕ(X)H = ϕ(A)H if necessary, we may assume without loss of generality that ϕ is non-degenerate. It follows then from Corollary 6.11 that there exists a covariant representation (ϕ, u) of (A, X, G, α) such that ψ = ϕ ⋊ u. Then, as in the proof of Proposition 6.8, we see that (ϕ, u) extends uniquely to a covariant representation ( ¯ϕ, u) of (C ∗ α G → B(H) coincides with the restriction to X ⋊u u(A, X) ⋊αu ,u G → B(H). This finishes the proof. (cid:3) u(A, X), G, αu) such that ψ = ϕ ⋊ u : X ⋊u α G of the integrated form ¯ϕ ⋊ u : C ∗ 7. Reduced crossed products α G, X ⋊r We now turn our attention to the construction of the reduced crossed product (A ⋊r α G) by an action α : G → Aut(A, X) of a group on a C ∗-operator system (A, X). Indeed, we define the reduced crossed product via the regular representation (ΛA, ΛX ) of (A, X, G, α) into M (A ⊗ K(L2(G), X ⊗ K(L2(G)) as constructed in Example 6.7: Definition 7.1. Let α : G → Aut(A, X) be an action of G on the C ∗- operator system (A, X). Then we define the reduced crossed product as the image (A, X) ⋊r α G = (A ⋊r α G, X ⋊r α G) :=(cid:0)ΛA(A ⋊u α G), ΛX (X ⋊u α G)(cid:1) of the universal crossed product by the regular representation (ΛA, ΛX ) in- side (cid:0)(M (A ⊗ K(L2(G)), M (X ⊗ K(L2(G))(cid:1). For the following proposition recall from Remark 6.4 the construction of the representation Ind ρ := (ρ, 1K ⊗λ) on B(K ⊗L2(G)) given by the formula (cid:0)ρ(x)ξ(cid:1)(g) = ρ(αg−1(x))ξ(g) and (1 ⊗ λ)hξ(g) = ξ(h−1g), where ρ : X → B(K) is any given representation of (A, X) on some Hilbert space K. We call Ind ρ the covariant representation of (A, X, G, α) induced by ρ. As an easy consequence of our definition of reduced crossed product and the discussion at the end of Example 6.7 we obtain Proposition 7.2. Let α : G → Aut(A, X) be an action of G on the C ∗- operator system (A, X). Then for every non-degenerate representation ρ : X → B(K) of (A, X) the integrated form ρ ⋊ (1 ⊗ λ) : X ⋊u α G → B(K ⊗ 26 MASSOUD AMINI, SIEGFRIED ECHTERHOFF, AND HAMED NIKPEY L2(G)) of the induced representation Ind ρ = (ρ, 1 ⊗ λ) (which we shall also simply denote by Ind ρ) factors through the reduced crossed product X ⋊r α G α G into B(K ⊗ L2(G)). Moreover, if ρ to give a representation of (A, X) ⋊r is completely isometric, then so is Ind ρ. α G), ΛX (X ⋊u Proof. By the discussion at the end of Example 6.7 we have the identity. Ind ρ = (ρ ⊗ idK(L2(G))) ◦ (ΛA, ΛX) as a representation from (A, X) ⋊u α G to B(K ⊗ L2(G)). Therefore Ind ρ clearly factors through (A, X) ⋊r α G = same holds for ρ ⊗ idK(L2(G)) : (A ⊗ K(L2(G)), X ⊗ L2(G)) → B(K ⊗ L2(G)) α G)(cid:1). If ρ : X → B(K) is completely isometric, the (cid:0)ΛA(A ⋊u and its extension to the multiplier system(cid:0)M (A⊗K(L2(G)), M (X ⊗L2(G))(cid:1) (see Lemma 3.7). Therefore Ind ρ factors through a completely isometric representation of (A, X) ⋊r (cid:3) α G into B(K ⊗ L2(G)) as claimed. Remark 7.3. (a) It follows in particular from the above proposition that, up to completely isomeric isomorphism, the reduced crossed product (A, X) ⋊r α G does not depend on the particular representation of (A, X) on a Hilbert space H. (b) Suppose that (C, j) is any C ∗-hull of (A, X) such that a given action α : G → Aut(A, X) extends to an action α : G → Aut(C). Let ρ : C → B(K) be any faithful and non-degenerate representation on a Hilbert space K. Then ρX := ρ ◦ j : X → B(K) is a non-degenerate completely isometric representation of (A, X) into B(K). Then the regular representation Ind ρ : C ⋊α,u G → B(K ⊗ L2(G)) factors through a faithful representation of the C ∗-reduced crossed product C ⋊α,r G whose composition with the canonical inclusion of (Cc(G, A), Cc(G, X)) into C ⋊α,r G coincides the completely α G → B(K ⊗ L2(G)) on the isometric representation Ind ρX : (A, X) ⋊r dense subsystem (Cc(G, A), Cc(G, X)). It follows that, up to a completely isometric isomorphism, the reduced crossed product (A, X) ⋊r α G can be identified with the closure of the pair (Cc(G, A), Cc(G, X)) inside C ⋊α,r G. This observation applies in particular to the universal C ∗-hull (C ∗ u(A, X), ju) and the enveloping C ∗-algebra (C ∗ env(A, X), jenv), where it easily follows from the respective universal properties that every action on (A, X) extends to actions on C ∗ env(A, X), respectively. u(A, X) and C ∗ 8. Crossed products by coactions Recall (e.g. from [12, Appendix A]) that if G is a locally compact group, there is a canonical comultiplication δG : C ∗(G) → M (C ∗(G) ⊗C ∗(G)) on C ∗(G) which is given as the integrated form of the unitary representa- tion g 7→ ug ⊗ ug ∈ U M (C ∗(G) ⊗C ∗(G)), where u : G → U M (C ∗(G)) is the canonical representation of G into U M (C ∗(G)). A coaction of G on C ∗-OPERATOR SYSTEMS AND CROSSED PRODUCTS 27 a C ∗-algebra A is an injective non-degenerate ∗-homomorphism δ : A → M (A ⊗C ∗(G)) such that the following conditions hold: (1) δ(A)(1 ⊗C ∗(G)) ⊆ A ⊗C ∗(G). (2) The following diagram of maps commutes A δy δ−−−−→ M (A ⊗C ∗(G)) yidA ⊗δG M (A ⊗C ∗(G)) −−−−→ δ⊗idG M (A ⊗C ∗(G) ⊗C ∗(G)) where idG denotes the identity on C ∗(G). If, in addition, we have the identity (1′) δ(A)(1 ⊗C ∗(G)) = A ⊗C ∗(G) then the coaction δ is called non-degenerate. Note that condition (1′) is automatic if G is amenable or discrete (see [19, Proposition 6], [24, Lemma 3.8] and [2]). We are now going to extend the definition of a coaction of C ∗(G) to the category of C ∗-operator systems. Definition 8.1. Let G be a locally compact group. A coaction of G on the C ∗-operator system (A, X) is an injective non-degenerate generalized morphism δX : (A, X) →(cid:0)M (A ⊗C ∗(G)), M (X ⊗C ∗(G))(cid:1) such that the following holds: (1) The map δA := δXA : A → M (A ⊗C ∗(G)) is a coaction of C ∗(G) on A. (2) The following diagram of maps commutes δX−−−−→ M (X ⊗C ∗(G)) yidX ⊗δG M (X ⊗C ∗(G)) −−−−−→ δX ⊗idG M (X ⊗C ∗(G) ⊗C ∗(G)) X δXy Remark 8.2. (a) Notice that condition (1) always implies that δX (X)(1 ⊗C ∗(G)) = δX (XA)(1 ⊗C ∗(G)) = δX (X)(cid:0)δX (A)(1 ⊗C ∗(G))(cid:1) ⊆ δX (X)(A ⊗C ∗(G)) = X ⊗C ∗(G), where the last equation follows from the nondegeneracy of δX → M (X ⊗C ∗(G)). (b) Let 1G : C ∗(G) → C denote the integrated form of the trivial representation of G. Then it follows from the definition of a coaction δX : (A, X) → M (X ⊗C ∗(G)) that (idX ⊗1G) ◦ δX is the identity on X. To see this observe that it follows from (a), that for all z ∈ C ∗(G) and x ∈ X 28 MASSOUD AMINI, SIEGFRIED ECHTERHOFF, AND HAMED NIKPEY we have (idX ⊗1G)(δX (x)(1 ⊗ z)) ∈ X. Choosing z such that 1G(z) = 1 then implies that (idX ⊗1G)(δX (x)) = (idX ⊗1G)(δX (x)(1 ⊗ z)) ∈ X as well. Now, using condition (2) and the relation (idG ⊗1G) ◦ δG = idG we get δX (x) = (idX ⊗ idG ⊗1G) ◦ (idX ⊗δG) ◦ δX (x) = (idX ⊗ idG ⊗1G) ◦ (δX ⊗ idG) ◦ δX (x) = (δX ⊗ 1G) ◦ δX (x) = δX(cid:0)(idX ⊗1G)(δX (x))(cid:1), which implies δX ◦ (idA ⊗1G) ◦ δX = δX . Since δX is injective, this implies that (idX ⊗1G) ◦ δX = idX. In particular, it follows that δX is completely isometric. Example 8.3. Suppose that α : G → Aut(A, X) is an action of the locally compact group G on the C ∗-operator system (A, X). Then there is a dual coaction α G, X ⋊u α G) →(cid:0)M (A ⋊u bα : (A ⋊u α G ⊗C ∗(G)), M (X ⋊u α G ⊗C ∗(G))(cid:1) given by the integrated form of the generalized covariant homomorphism α G ⊗C ∗(G)), where (iX , iG) : (iX ⊗ 1, iG ⊗ u) of (A, X, G, α) into M (X ⋊u (X, G) → M (X ⋊u α G) denotes the canonical covariant homomorphism and u : G → U M (C ∗(G)) the universal representation. To see that this satisfies the conditions of Definition 8.1 we choose a faithful and non-degenerate representation of C ∗ u(A, X) ⋊α,u G into some B(H). This restricts to a faithful representation of (A ⋊u α G), into B(H). Moreover, by choosing a faithful representation of C ∗(G) onto some Hilbert space K, say, we obtain a faithful and non-degenerate representation u(A, X) ⋊αu,u G ⊗C ∗(G) on H ⊗ K, which restricts to a completely of C ∗ isometric representation of X ⊗C ∗(G)), and hence of M (X ⋊u α G ⊗C ∗(G)), respectively (use Corollary 3.12). Now, there is a dual coaction α G, X ⋊u u(A, X) ⋊α,u G → M (C ∗ u(A, X) ⋊α,u G ⊗C ∗(G)) u(A, X), G, α) into M (C ∗ u(A,X) ⊗ given as the integrated form of the covariant homomorphism (iC ∗ u(A, X) ⋊α,u G ⊗C ∗(G)). This repre- 1, iG ⊗ u) of (C ∗ sentation clearly restricts to the representation (iX ⊗ 1, iG ⊗ u) : (X, G) → α G ⊗C ∗(G)) and the conditions in Definition 8.1 can then easily be M (X ⋊u u(A, X)⋊α,u G. deduced from the properties of the coaction bαu of C ∗(G) on C ∗ Similarly, the dual coaction u(A, X) ⋊α,r G → M (C ∗ u(A, X) ⋊α,r G ⊗C ∗(G)) restricts to a dual coaction bαu : C ∗ cαr : C ∗ bαr := (ir X ⊗ 1) ⋊ (ir G ⊗ u) : X ⋊r α G → M (X ⋊r α G ⊗C ∗(G)) C ∗-OPERATOR SYSTEMS AND CROSSED PRODUCTS 29 of C ∗(G) on the reduced crossed product (A, X) ⋊r where (ir X , ir into M (X ⋊r α G), G) denotes the canonical covariant homomorphism of (A, X, G, α) α G) (i.e., the regular representation). α G = (A ⋊r α G, X ⋊r of C ∗(G) on C ∗ We now want to relate coactions of C ∗(G) on (A, X) with coactions In order to formulate the result, observe that (cid:0)C ∗ u(A, X) ⊗C ∗(G), ju ⊗ idC ∗(G)(cid:1) is a C ∗-hull of (A ⊗C ∗(G), X ⊗C ∗(G)) and therefore we get a canonical completely isometric embedding u(A, X). ju ⊗ idG : M (A ⊗C ∗(G), X ⊗C ∗(G)) ֒→ M (C ∗ u(A, X) ⊗C ∗(G)). Proposition 8.4. Let (A, X) be a C ∗-operator system. Then there is a one- to-one correspondence between coactions δX of G on (A, X) and coactions δu : C ∗ u(A, X) → M (C ∗ u(A, X) ⊗C ∗(G)) of G on C ∗ (8.1) u(A, X) which satisfy the conditions δu(A) ⊆ M (A ⊗C ∗(G)) and δu(X) ⊆ M (X ⊗C ∗(G)), where we regard A and X as subspaces of C ∗ ju. Given such coaction δu of G on C ∗ G on (A, X) is given by the restriction δX := δuX. u(A, X) via the inclusion map u(A, X), the corresponding coaction of Proof. Suppose first that δX : X → M (X ⊗C ∗(G)) is a coaction of C ∗(G) on (A, X). It follows then from part (b) of Remark 8.2 and Lemma 4.10 that (ju ⊗ idG) ◦ δX : X → M (C ∗ u(A, X) ⊗C ∗(G)) is a completely isometric representation of (A, X) into M (C ∗ By Proposition 4.6, it extends to a non-degenerate ∗-homomorphism u(A, X) ⊗C ∗(G)). δu : C ∗ u(A, X) → M (C ∗ u(A, X) ⊗C ∗(G)). Since C ∗ u(A, X) is generated by X and since δu(x)(1 ⊗ z) = δX (x)(1 ⊗ z) ∈ X ⊗C ∗(G) ⊆ C ∗ u(A, X) ⊗C ∗(G) for all x ∈ X, it follows that δu(C ∗ u(A, X))(1 ⊗C ∗(G)) ⊆ C ∗ u(A, X) ⊗C ∗(G). u(A,X) ⊗1G)◦δu maps C ∗ Using this, it follows that (idC ∗ Moreover, since (idC ∗ X, it follows that (idC ∗ fore must be equal to the identity on C ∗ injective. u(A, X). u(A,X) ⊗1G) ◦ δu restricts to (idX ⊗1G) ◦ δX = idX on u(A,X) ⊗1G) ◦ δu extends the identity on X and there- u(A, X). Thus it follows that δu is u(A, X) into C ∗ In order to check that (8.2) (δu ⊗ idG) ◦ δu = (idC ∗ u(A,X) ⊗δG) ◦ δu 30 MASSOUD AMINI, SIEGFRIED ECHTERHOFF, AND HAMED NIKPEY u(A, X) into M (C ∗ u(A, X) ⊗C ∗(G) ⊗C ∗(G)) we simply ob- as maps from C ∗ serve that, via the canonical embedding of M (X ⊗C ∗(G) ⊗C ∗(G)) into u(A, X) ⊗C ∗(G) ⊗C ∗(G)), the left hand side restricts to (δX ⊗ idG) ◦ δX M (C ∗ and the right hand side restricts to (idX ⊗δG) ◦ δX . But by condition (ii) of Definition 8.1, these restrictions to X coincide and then (8.2) follows from the uniqueness assertion of Proposition 4.6. u(A, X) → M (C ∗ Conversely, suppose that δ : C ∗ u(A, X) ⊗C ∗(G)) is a coaction of G on C ∗ u(A, X) such that the equations (8.1) hold. We need to check that δX := δX is a coaction of G on (A, X), where we realize M (X ⊗C ∗(G)) as a subspace of M (C ∗ u(A, X) ⊗C ∗(G)) as explained above. Since δ is injective, the same holds for δX and condition (ii) of Definition 8.1 clearly follows from the similar condition for δ. Thus, all we need to show is that δ(A)(1 ⊗C ∗(G)) ⊆ A ⊗C ∗(G). But since δ(A)(1 ⊗C ∗(G)) ⊆ C ∗ u(A, X) ⊗C ∗(G) we observe that δ(A)(1 ⊗C ∗(G)) ⊆ M (A ⊗C ∗(G)) ∩ (C ∗ u(A, X) ⊗C ∗(G)), u(A, X) ⊗C ∗(G)). But it follows where the intersection is taken inside M (C ∗ from Lemma 4.10 that this intersection equals A ⊗C ∗(G) and the result follows. (cid:3) Definition 8.5. A coaction δX : X → M (X ⊗C ∗(G)) of G on (A, X) is called non-degenerate if the corresponding coaction δu of G on C ∗ u(A, X) is non-degenerate, i.e., δu(C ∗ u(A, X))(1 ⊗C ∗(G)) = C ∗ u(A, X) ⊗C ∗(G). Remark 8.6. Of course it would be more satisfactory to define nondegeneracy of a coaction of G on (A, X) via a condition like δX (X)(1 ⊗C ∗(G)) = X ⊗C ∗(G). However, we were not able to prove that this condition is equivalent to nondegeneracy of δu, and it is the latter condition we shall need later when dealing with Imai-Takai duality. Note that nondegeneracy of a coaction on (A, X) is automatic for amenable or discrete G, since, as we remarked before, this holds true for coactions on C ∗-algebras. The same holds for all dual coactions: Lemma 8.7. Suppose that α : G → Aut(A, X) is an action of G on the α G → M (X ⋊u α C ∗-operator system (A, X). Then the dual coactions bα : X ⋊u G ⊗C ∗(G)) and cαr : X ⋊r Proof. The first assertion follows from Corollary 6.12 together with the fact that dual coactions on C ∗-algebra crossed products are non-degenerate (e.g., see the discussion at the end of [12, Example A.26]). For the dual coaction on α G ⊗C ∗(G)) are non-degenerate. α G → M (X ⋊r C ∗-OPERATOR SYSTEMS AND CROSSED PRODUCTS 31 the reduced crossed (A, X) ⋊r induces surjective morphisms α G observe that the identity map on Cc(G, X) C ∗ u(A, X) ⋊α,u G ։ C ∗ u(A ⋊r α G, X ⋊r α G) ։ C ∗ u(A, X) ⋊α,r G, u(A ⋊r u(A ⋊r α G, X ⋊r u(A⋊u u(A, X) ⋊α,u G together with the obvious morphism of (A ⋊u where the first map exists by the universal property of C ∗ C ∗ into C ∗ dual coactions, where the one on C ∗ dual coaction of C ∗(G) on (A ⋊r α G) ∼= α G) α G). These maps are equivariant for the respective α G) is induced from the α G) via Proposition 8.4. But then it u(A, X) ⋊α,u G α G). (cid:3) αG, X ⋊u α G, X ⋊u is easy to see that nondegeneracy of the dual coaction bα on C ∗ implies nondegeneracy of the (dual) coaction on C ∗ u(A ⋊r α G, X ⋊r α G, X ⋊r α G, X ⋊r We are now going to study covariant representations for coactions on C ∗-operator systems (A, X), extending the well known theory for coactions on C ∗-algebras. In what follows let wG ∈ U M (C0(G) ⊗C ∗(G)) denote the unitary given by the map [g 7→ ug] ∈ Cb(G, C ∗(G)) ⊆ M (C0(G) ⊗C ∗(G)). Recall from [12, Definition A.32] the following definition Definition 8.8. Let δ : D → M (D ⊗C ∗(G)) be a coaction of G on the C ∗-algebra D and let B be a C ∗-algebra. Then a covariant representation of (D, G, δ) into M (B) is a pair (π, µ), where π : D → M (B), µ : C0(G) → M (D) are non-degenerate ∗-homomorphism satisfying the covariance con- dition (π ⊗ idG) ◦ δ(d) = (µ ⊗ idG)(wG)(π(d) ⊗ 1)(µ ⊗ idG)(wG)∗. If B = K(H) for some Hilbert space H, then we say that (π, µ) is a covariant representation on H. If (π, µ) is a covariant representation of (D, G, δ) as above, then π(D)µ(C0(G)) := span{π(d)µ(f ) : d ∈ D, f ∈ C0(G)} is a C ∗-subalgebra of M (B) (see [12, Proposition A.36]). Moreover, it is shown in [12, Proposition A.37] that the pair (ΛD, Λ bG) :=(cid:0)(idD ⊗λ) ◦ δ, 1 ⊗ M ) where M : C0(G) → B(L2(G)) = M (K(L2(G))) denotes the representa- tion by multiplication operators, defines a covariant representation, called regular representation, of (D, G, δ) into M (D ⊗ K(L2(G))). The crossed D ⋊δ bG := ΛD(D)Λ bG(C0(G)) ⊆ M (D ⊗ K(L2(G))). product D ⋊δ bG of the co-system (D, G, δ) is then defined as the C ∗-algebra We can then view (ΛD, Λ bG) as a covariant representation into M (D ⋊δ bG) (D ⋊δ bG, ΛD, Λ bG) satisfies the following universal property: If (π, µ) is any in a canonical way. It is then shown in [12, Theorem A.41] that the triple 32 MASSOUD AMINI, SIEGFRIED ECHTERHOFF, AND HAMED NIKPEY covariant representation of (D, G, δ) into M (B), then there exists a unique ∗-homomorphism π ⋊ µ : D ⋊δ bG → M (B) such that (π ⋊ µ) ◦ ΛD = π and (π ⋊ µ) ◦ Λ bG = µ. (8.3) Moreover, we get π ⋊ µ(D ⋊δ bG) = π(D)µ(C0(G)). We are now going derive analogues of the above constructions and facts for coactions of G on C ∗-operator systems (A, X). We start with Definition 8.9. Suppose that δX : X → M (X ⊗C ∗(G)) is a coaction of G on (A, X) and let (B, Y ) be a C ∗-operator system. Then a covariant morphism of (A, X, G, δX ) into M (B, Y ) = (M (B), M (Y )) consists of a non-degenerate generalized morphism π : X → M (Y ) of (A, X) together with a non-degenerate ∗-homomorphism µ : C0(G) → M (B) such that the pair (π, µ) satisfies the covariance condition (π ⊗ idG) ◦ δX (x) = (µ ⊗ idG)(wG)(π(x) ⊗ 1)(µ ⊗ idG)(wG)∗ If (B, Y ) = (B, B) is a C ∗-algebra, we say that (π, µ) for all x ∈ X. is a covariant representation of (A, X, G, δX ) into M (B). If, in addition, B = K(H) for some Hilbert space H, we say that (π, µ) is a covariant representation of (A, X, G, δX ) on H. Remark 8.10. Observe that the restricted pair (πA, µ) with πA := πA of a covariant representation (π, µ) of (A, X, G, δX ) into M (B, Y ) is a non- degenerate covariant homomorphism of (A, G, δA) into M (B). Proposition 8.11. Suppose that (π, µ) is a covariant morphism of (A, X, G, δX ) into M (B, Y ) for some C ∗-operator system (B, Y ). Then (cid:0)π(A)µ(C0(G)), π(X)µ(C0(G))(cid:1) (closed spans!) is a C ∗-operator subsystem of M (B, Y ). Proof. We first observe that it follows directly from the above discussion that π(A)µ(C0(G)) is a non-degenerate C ∗-subalgebra of M (B). Note that this implies in particular µ(C0(G)) acts as multipliers on this C ∗-algebra. On the other hand, precisely the same arguments as used in the proof of [12, Proposition A.36] show that π(X)µ(C0(G)) = µ(C0(G))π(X), from which it follows that π(X)µ(C0(G)) is a selfadjoint subspace of M (Y ). So in order to complete the proof, we only need to show that (cid:0)π(A)µ(C0(G))(cid:1)(cid:0)π(X)µ(C0(G))(cid:1) = π(X)µ(C0(G)) =(cid:0)π(X)µ(C0(G))(cid:1)(cid:0)π(A)µ(C0(G))(cid:1), but this follows AX = X = XA and π(X)µ(C0(G)) = µ(C0(G))π(X). (cid:3) C ∗-OPERATOR SYSTEMS AND CROSSED PRODUCTS 33 Proposition 8.12. Let δX : X → M (X ⊗C ∗(G)) be a coaction of G on u(A, X) ⊗C ∗(G)) (A, X) and let B be a C ∗-algebra. Let δu : C ∗ denote the corresponding coaction of G on the universal C ∗-hull C ∗ u(A, X) of (A, X) as in Proposition 8.4. u(A, X) → M (C ∗ Then there is a one-to-one correspondence between the non-degenerate covariant representations (π, µ) of (A, X, G, δX ) into M (B) and the non- degenerate covariant representations of (C ∗ u(A, X), G, δu) into M (B), given by sending a covariant pair (π, µ) of (A, X, G, δX ) to the covariant pair (¯π, µ), where ¯π : C ∗ u(A, X) → M (B) denotes the unique ∗-homomorphism which extends π. Proof. Since (π ⊗ idG) ◦ δX = (µ ⊗ idG)(wG)(π(·) ⊗ 1)(µ ⊗ idG)(wG)∗ as maps from X into M (D ⊗C ∗(G)), it follows that both of the ∗-homomorphisms (¯π ⊗ idG) ◦ δu and (µ ⊗ idG)(wG)(π(·) ⊗ 1)(µ ⊗ idG)(wG)∗ from C ∗ u(A, X) to M (B ⊗C ∗(G)) extend the same non-degenerate generalized morphism of (A, X), and therefore they coincide by Proposition 4.6. This shows that any covariant representation of (A, X, G, δX ) has a unique extension to (C ∗ (cid:3) u(A, X), G, δu). The converse direction follows by restriction. Lemma 8.13. Suppose that (A, X, G, δ) is a coaction of G on (A, X). Then the pair (ΛX, Λ bG) := ((idX ⊗λ) ◦ δX, 1 ⊗ M ) defines a covariant represen- tation of (A, X, G, δ) into M(cid:0)A ⊗ K(L2(G)), X ⊗ K(L2(G))(cid:1) which we call the regular representation of (A, X, G, δ) into M (X ⊗C ∗(G)). Moreover, via the completely isometric embedding of (A, X) into C ∗ u(A, X) and the corresponding completely isometric embedding of (A⊗K(L2(G)), X ⊗ K(L2(G))) into C ∗ u(A, X)⊗K(L2(G)), we may view (ΛX , Λ bG) as a covariant u(A, X)⊗ K(L2(G))), which uniquely extends to the representation into M (C ∗ u(A, X), G, δu) in the sense of regular representation (ΛC ∗ Proposition 8.12. u(A,X), Λ bG) of (C ∗ Proof. For the first assertion we follow the proof of [12, Proposition A.37]. Using the identity (λ ⊗ idG) ◦ δG = Ad(M ⊗ idG)(wG) ◦ (λ ⊗ 1), which has been established in the proof of [12, Proposition A.37], we compute (cid:0)(idX ⊗λ) ◦ δX ⊗ idG(cid:1) ◦ δX (x) = (idX ⊗λ ⊗ idG) ◦ (δX ⊗ idG) ◦ δX (x) = (idX ⊗λ ⊗ idG) ◦ (idX ⊗δG) ◦ δX (x) = (idX ⊗(λ ⊗ idG) ◦ δG) ◦ δX (x) = (1 ⊗ M ⊗ idG)(wG)(cid:0)(idX ⊗λ)(δX (x) ⊗ 1)(1 ⊗ M ⊗ idG)(wG)∗. This proves the covariance condition for ((idX ⊗λ) ◦ δX , 1 ⊗ M ). The second assertion is now obvious. (cid:3) 34 MASSOUD AMINI, SIEGFRIED ECHTERHOFF, AND HAMED NIKPEY Definition 8.14. Suppose that δX : X → M (X ⊗C ∗(G)) is a coaction of G on the C ∗-operator system (A, X). Then we define the crossed product (A, X) ⋊δX G as the C ∗-operator system (A ⋊δA bG, X ⋊δX bG) :=(cid:0)ΛX(A)Λ bG(C0(G)), ΛX (X)Λ bG(C0(G))(cid:1) generated by the regular representation (ΛX , Λ bG) of (A, X, G, δX ) as in Proposition 8.11. Remark 8.15. Note that it follows directly from Lemma 8.13 and the above u(A, X))Λ bG(C0(G)) is a C ∗- definition that C ∗ u(A,X)(C ∗ hull of (A ⋊δA bG, X ⋊δX bG). Indeed, we shall see below, that it the universal C ∗-hull of (A ⋊δA bG, X ⋊δX bG). We now show that the above defined crossed product does enjoy a uni- u(A, X) ⋊δu bG = ΛC ∗ versal property for covariant representations: Proposition 8.16. Suppose that (π, µ) is a covariant morphism of (A, X, G, δX ) into M (B, Y ) for some C ∗-operator system (B, Y ). Then there is a unique generalized morphism such that (8.4) π ⋊ µ : (A ⋊δA bG, X ⋊δX bG) → M (B, Y ) (π ⋊ µ) ◦ ΛX = π and (π ⋊ µ) ◦ Λ bG = µ. Conversely, if Φ : (A ⋊δA bG, X ⋊δX bG) → M (B, Y ) is any non-degenerate generalized morphism, then there is a unique covariant morphism (π, µ) of (A, X, G, δX ) such that Φ = π ⋊ µ. Proof. As for the case of actions, we are going to use the correspondence between covariant representations of (A, X, G, δX ) and covariant represen- tations of (C ∗ u(A, X), G, δu) as established in Proposition 8.12. For this we choose a non-degenerate completely isometric embedding of (B, Y ) into B(H) for some Hilbert space H. Then, if we compose a covariant represen- tation of (π, µ) of (A, X, G, δX ) into M (B, Y ) with this inclusion, we may view (π, µ) as a representation into B(H) = M (K(H)). By Proposition 8.12, this extends to a covariant representation, say (¯π, µ) of (C ∗ u(A, X), G, δu) into B(H). By the universal property of the C ∗-algebra crossed prod- uct (C ∗ u(A,X), Λ bG) there exists a unique ∗-homomorphism ¯π ⋊ µ : C ∗ u(A, X) ⋊δu bG, ΛC ∗ u(A, X) ⋊δu bG → B(H) such that u(G,A) = ¯π and (¯π ⋊ µ) ◦ Λ bG = µ. (¯π ⋊ µ) ◦ ΛC ∗ Define π ⋊ µ as the restriction of ¯π ⋊ µ to X ⋊δX bG ⊆ C ∗ the restriction restriction of (¯π ⋊ µ) ◦ ΛC ∗ u(G,A) to X equals (π ⋊ µ) ◦ ΛX. u(A, X) ⋊δu bG. Then C ∗-OPERATOR SYSTEMS AND CROSSED PRODUCTS 35 representation which satisfies (8.4). We still need to check that π ⋊µ can be viewed as a generalized morphism Since ¯π extends π, we see that π ⋊ µ : (A ⋊δA bG, X ⋊δX bG) → B(H) is a into M (B, Y ). For this it suffices to check that π ⋊ µ(X ⋊δX bG)B ⊆ Y and π ⋊µ(A⋊δA bG)B = B. But if we apply π ⋊µ on a typical element of the form ΛX(x)Λ bG(f ) of X ⋊δX bG with x ∈ X, f ∈ C0(G), it follows from equation (8.4) that π ⋊ µ(cid:0)(ΛX (x)Λ bG(f )(cid:1)b = π(x)µ(f )b ∈ Y since, by definition of a covariant representation into M (B, Y ), we have µ(f ) ∈ M (B) and π(x) ∈ M (Y ). Moreover, since πA : A → M (B) and µ : A → M (B) are supposed to be non-degenerate, we get π ⋊ µ(A ⋊δA bG)B = (π(A)µ(C0(G)))B = π(A)(µ(C0(G))B) = π(A)B = B. If, conversely, Φ : (A ⋊δA bG, X ⋊δX bG) → M (B, Y ) is any non-degenerate generalized morphism, then we leave it as a straightforward exercise to check that the pair (π, µ) with π := Φ ◦ ΛX, µ := Φ ◦ Λ bG is a non-degenerate covariant morphism such that Φ = π ⋊ µ. (cid:3) Corollary 8.17. Suppose that (A, X, G, δX ) is a coaction of G on (A, X). Then C ∗ u(A ⋊δA bG, X ⋊δX bG) = C ∗ u(A, X) ⋊δu bG. Proof. We already observed in Remark 8.15 that C ∗ u(A, X)⋊δu bG is a C ∗-hull of (A ⋊δA bG, X ⋊δX bG). So we only need to show that every representation Φ : (A ⋊δA bG, X ⋊δX bG) → B extends to a homomorphism ¯Φ : C ∗ bG → B. For this let us assume without loss of generality that B is generated by the image Φ(X ⋊δX bG). Then Φ is non-degenerate and there exists a unique non-degenerate covariant representation (π, µ) of (A, X, G, δX ) such that Φ = π ⋊ µ. By Proposition 8.12 (π, µ) extends uniquely to a covariant homomorphism of (¯π, µ) of (C ∗ u(A, X), G, δu) into B. The arguments given in the proof of Proposition 8.16 then show that ¯π ⋊ µ is a ∗-homomorphism from C ∗ (cid:3) u(A, X) ⋊δu bG into B which restricts to π ⋊ µ on X. Remark 8.18. We should note that, different from the situation for crossed u(A, X) ⋊δu products by actions, the definition of the crossed product (A⋊δA bG, X ⋊δX bG) does not depend on the crossed product by the universal C ∗-hull C ∗ u(A, X). This algebra is only used to reduce the proof of the universal properties to the well known case of coaction crossed products by C ∗-algebras. One should observe that the definition of the crossed product for coactions is more like the definition of the reduced crossed product in case of group actions. The fact that this constructions already enjoys the universal property for covariant morphisms comes from the fact that the locally compact quantum 36 MASSOUD AMINI, SIEGFRIED ECHTERHOFF, AND HAMED NIKPEY group C ∗(G) is amenable for all G (or, in other words, every group G is coamenable (e.g., see [3, 23] for a discussion of these notions). Hence we only have one reasonable candidate for a coaction crossed product! For this reason, it also follows that the algebra part A ⋊δA G of (A ⋊δA bG, X ⋊δX bG) is the (universal and reduced) crossed product of A with respect to the coaction δA. 9. Duality We now want to deduce versions of the Imai-Takai and Katayama duality for crossed products by actions and coactions. By Example 8.3 we know that α G) and (A ⋊r the universal and reduced crossed products (A ⋊u α G, X ⋊r α G) for an action α : G → Aut(A, X) carry canonical dual coactions α G, X ⋊u X ⊗ 1) ⋊ (ir G ⊗ u) into M (X ⋊u bα andcαr which are given as the integrated forms of the covariant morphisms bα = (iX ⊗ 1) ⋊ (iG ⊗ u) andcαr = (ir α G ⊗C ∗(G)) α G ⊗C ∗(G))), where (iX , iG) and (ir (resp. M (X ⊗r G) are the canonical covariant morphisms from (A, X, G, α) into M (X ⋊u α G), respectively. Note that it follows directly from the constructions in Example 8.3, that these coactions extend to the dual coaction on C ∗ u(A, X) ⋊α,u G = C ∗ X , ir α G) and M (X ⋊r u(A, X) ⋊αu,r G, respectively. α G, X ⋊u u(A ⋊u α G) and C ∗ In a similar way, we have Proposition 9.1. Suppose that (A, X, G, δX ) is a coaction of G on (A, X). Then there is a canonical dual action wich is given on a typical element ΛX (x)Λ bG(f ) by bδ : G → Aut(A ⋊δA bG, X ⋊δX bG) bδg(cid:0)ΛX (x)Λ bG(f )(cid:1) = ΛX(x)Λ bG(σg(f )), where σ : G → Aut(C0(G)) denotes the right translation action, i.e., σg(f )(s) = f (sg) for all g, s ∈ G, f ∈ C0(G). Proof. For any covariant representation (π, µ) of (A, X, G, δX ) the pair (π, µ ◦ σg) is a covariant representation as well. Indeed, for all x ∈ X we have (µ◦σg ⊗ idG)(wG)(x ⊗ 1)(µ ◦ σg ⊗ idG)(wG)∗ = (µ ⊗ idG)(wG)(1 ⊗ ug)(x ⊗ 1)(1 ⊗ u∗ = (µ ⊗ idG)(wG)(x ⊗ 1)(µ ⊗ idG)(wG)∗ = π(x). g)(µ ⊗ idG)(wG)∗ Applying this to the regular representation (ΛX , Λ bG) we get a covariant rep- resentation (ΛX , Λ bG ◦σg) of (A, X, G, δX ) into M (X ⋊δX bG) whose integrated form bδg maps ΛX(x)Λ bG(f ) to ΛX(x)Λ bG(σg(f )). It is then clear that bδg−1 C ∗-OPERATOR SYSTEMS AND CROSSED PRODUCTS 37 the action σ : G → Aut(C0(G)) is strongly continuous, the same holds for (cid:3) inverts bδg and that g 7→ bδg is a homomorphism into Aut(X ⋊δX bG). Since bδ. We now formulate the analogue of the Imai-Takai duality theorem for crossed products of C ∗-operator systems by actions. Theorem 9.2. Suppose that α : G → Aut(A, X) is an action and let K := K(L2(G)). Then there are canonical isomorphisms and (cid:0)A ⋊u (cid:0)A ⋊r α G ⋊ bα bG, X ⋊u α G ⋊cαr bG, X ⋊r α G ⋊ cαu bG(cid:1) ∼=(cid:0)A ⊗ K, X ⊗ K(cid:1) α G ⋊cαr bG(cid:1) ∼=(cid:0)A ⊗ K, X ⊗ K(cid:1) which transfer the double dual actions bbα and (resp. ccαr) to the diagonal action α ⊗ Ad ρ, respectively, where ρ : G → U (L2(G)) denotes the right regular representation of G. Proof. Both assertions can be deduced easily from the well known Takai- Takesaki duality theorem for the corresponding action on the universal C ∗- hull C ∗ u(A, X). Indeed, it is shown in [31, Theorem 5.1] (in the even more general situation of a dual coaction of a twisted action) that for any action β : G → Aut(B) the Imai-Takai isomorphism Ψ : B ⋊β G ⋊ bβ bG ∼=→ B ⊗ K is given by the integrated form (ΛB ⋊ ΛG) ⋊ Λ bG where ΛB ⋊ ΛG : B ⋊β G → M (B ⊗ K(L2(G))) is the regular representation of B ⋊β G and Λ bG = 1 ⊗ M . It is then clear that this factors through a homomorphism of B ⋊β,r G ⋊ bβ bG, which explains that both crossed products are the same. Now, if we apply this to the system (C ∗ u(A, X), G, α) we obtain the iso- morphism (ΛC ∗ u(A,X) ⋊ ΛG) ⋊ Λ bG : C ∗ u(A, X) ⋊αu G ⋊ cαu bG ∼=→ C ∗ u(A, X) ⊗ K which then clearly restricts to an isomorphism (ΛX ⋊ ΛG) ⋊ Λ bG : X ⋊u ∼=→ X ⊗ K α G ⋊ bα bG and similarly for X ⋊r α G ⋊cαr bG. The statement on the double dual action bbα follows from the analoguous statement for the double dual action bbα on the double crossed product of C ∗ u(A, X). (cid:3) We now proceed to a discussion of Katayama duality, where we want to study double crossed products (A, X) ⋊δX bG ⋊bδX G 38 MASSOUD AMINI, SIEGFRIED ECHTERHOFF, AND HAMED NIKPEY by dual actions of coactions. Note that here it will usually matter whether we take the universal or the reduced crossed product (or any exotic crossed product in between) on the outside, so we need to clarify this point. So let us first recall the situation if we start with a coaction δ : B → M (B ⊗C ∗(G)) of G on a C ∗-algebra B. It is shown by Nilsen in [27] that there exists a surjective ∗-homomorphisms (9.1) given by the integrated form ΦB : B ⋊δ bG ⋊bδ,u G ։ B ⊗ K(L2(G)) ΦB =(cid:0)ΛB ⋊ Λ bG(cid:1) ⋊ (1 ⊗ ρ) M (B ⊗ K(L2(G))). A coaction δ is called maximal, if Φ is an isomorphism, of the covariant homomorphism (ΛB ⋊ Λ bG, 1 ⊗ ρ) of (B ⋊δ bG, G,bδ) where (ΛB, Λ bG) =(cid:0)(idB ⊗λ)◦δ, 1⊗M(cid:1) is the regular representation of (B, G, δ) into and it is called normal, if Φ factors through an isomorphism B ⋊δ bG ⋊bδ,r G ∼= B ⊗ K(L2(G)). In general, the isomorphism Φ will factor through an isomorphism B ⋊δ bG ⋊bδ,µ G ∼= B ⊗ K(L2(G)) and the reduced crossed product in the sense that it is a C ∗-completion of of some exotic crossed product B ⋊δ bG⋊bδ,µ G which lies between the maximal Cc(G, B⋊δ bG) such that the identity map on Cc(G, B⋊δ bG) induces surjective ∗-homomorphisms (9.2) B ⋊δ bG ⋊bδ,u G ։ B ⋊δ bG ⋊bδ,µ G ։ B ⋊δ bG ⋊bδ,r G. It has been shown by Quigg [29] that (B, G, δ) is normal if and only if ΛB = (idB ⊗λ) ◦ δ : B → M (B ⊗ K(L2(G))) is faithful. In general the coaction δ determines a normal coaction δn (called the normalization of δ) on the quotient Bn := B/(ker ΛB) such that the δ − δn equivariant quotient map Ψn : B ։ Bn descents to an isomorphism of the dual systems (B ⋊δ bG, G,bδ) ∼= (Bn ⋊δn bG, G, bδn). an action α of G on a C ∗-algebra A, then (B, δ) is maximal (see [11]) and If (B, δ) = (A ⋊α G,bα) is the dual coaction on the full crossed product by the normalization of (B, δ) is given by the pair (Bn, δn) = (A ⋊r G,bαr), the dual coaction of the reduced crossed product. We now want to extend this picture to crossed products by C ∗-operator systems. We start with the following obvious consequence to the above: Theorem 9.3. Suppose that (A, X, G, δX ) is a coaction of the locally com- pact group G on the C ∗-operator system (A, X). Then there is a canonical C ∗-OPERATOR SYSTEMS AND CROSSED PRODUCTS 39 surjective morphism ΦX :(cid:0)A ⋊δA bG ⋊u cδA given by the integrated form G, X ⋊δX bG ⋊u cδX G(cid:1) ։(cid:0)A ⊗ K(L2(G)), X ⊗ K(L2(G)(cid:1) Φ =(cid:0)ΛX ⋊ Λ bG(cid:1) ⋊ (1 ⊗ ρ) cδA cδX into M (X ⊗ K(L2(G))). Moreover, there is an exotic completion (A ⋊δA of the covariant homomorphism (ΛX ⋊ Λ bG, 1 ⊗ ρ) of (X ⋊δX bG, G,cδX ) where (ΛX, Λ bG) =(cid:0)(idX ⊗λ)◦δX , 1⊗M(cid:1) is the regular representation of (A, X, G, δX ) bG ⋊µ G) of the pair (Cc(G, A ⋊δA bG), (Cc(G, X ⋊δX bG), G, X ⋊δX bG ⋊µ (cid:0)A ⋊δA bG ⋊µ G(cid:1) ∼=(cid:0)A ⊗ K(L2(G)), X ⊗ K(L2(G)(cid:1). lying between the maximal and the reduced crossed products, such that ΦX factors through a completely isometric isomorphism Proof. The result follows from the above cited results for coactions on C ∗- algebras applied to the coaction (B, G, δ) = (C ∗ u(A, X), G, δu) and the re- striction of the corresponding ∗-homomorphism ΦB to A ⊆ X ⊆ X. (cid:3) G, X ⋊δX bG ⋊µ cδX cδA Corollary 9.4. Suppose that G is an amenable locally compact group. Then the morphism ΦX :(cid:0)A ⋊δA bG ⋊u cδA G, X ⋊δX bG ⋊u cδX of Theorem 9.3 is a completely isometric isomorphism. G(cid:1) ։(cid:0)A ⊗ K(L2(G)), X ⊗ K(L2(G)(cid:1) 10. C ∗-operator bimodules In this section we want to study C ∗-operator systems which are related to C ∗-operator bimodules. This will later lead to an easy way to define crossed products by group actions on C ∗-operator bimodules. Since every operator space can be regarded as a C ∗-operator bimodule in a canonical way, this will also give a construction of crossed products by group actions on operator spaces. Definition 10.1. Let H and K be Hilbert spaces. A concrete C ∗-operator bimodule (A, V, B) inside B(K, H) consists of a norm closed subset V ⊆ B(K, H) together with a C ∗-subalgebra A ⊆ B(H) and a C ∗-subalgebra B ⊆ B(K) satisfying AV = V = V B, AH = H, and BK = K. A representation of the C ∗-operator bimodule (A, V, B) on a pair of Hilbert spaces is (K ′, H ′) is a triple of maps ρ = (ρA, ρV , ρB) such that ρA : A → B(H ′) and ρB : B → B(K ′) are ∗-homomorphisms and ρV : V → B(K ′, H ′) is a completely bounded map such that ρV (avb) = ρA(a)ρV (v)ρB(b) ∀a ∈ A, v ∈ V, b ∈ B. 40 MASSOUD AMINI, SIEGFRIED ECHTERHOFF, AND HAMED NIKPEY We say that ρ is non-degenerate if ρA and ρB are non-degenerate. We say that ρ is completely contractive if ρV is completely contractive and ρ is called completely isometric if ρA and ρB are faithful and ρV is completely isometric. A morphism from the C ∗-operator bimodule (A, V, B) to the C ∗-operator bimodule (C, W, D) inside B(K ′, H ′) is a representation ϕ = (ϕA, ϕV , ϕB) of (A, V, B) to B(K ′, H ′) such that ϕA(A) ⊆ C, ϕV (V ) ⊆ W , and ϕB(B) ⊆ D. The invertible morphisms (or isomorphisms) are then precisely the surjective completely isometric morphisms. We shall often identify isomorphic C ∗- operator systems. Remark 10.2. (1) Every concrete operator spaceV ⊆ B(K, H) determines the concrete C ∗-operator system (C1H, V, C1K ). If V ′ ⊆ B(K ′, H ′) is an- other operator space and ϕV : V → V ′ is a completely bounded map, then ϕ = (ϕC1H , ϕV , ϕC1K ) with ϕ1H (λ1H ) = λ1H ′ and ϕ1K (λ1K ) = λ1K ′ is an morphism from (C1H , V, C1K ) → (C1H ′, V ′, C1K ′). Hence V and V ′ are iso- morphic as operator spaces if and only if (C1H , V, C1K ) and (C1H ′, V ′, C1K ′) are isomorphic as C ∗-operator bimodules. In this way we may regard the category of (concrete) C ∗-operator bimodules as an extension of the cate- gory of (concrete) operator spaces. (2) If (A, V, B) is a triple of subsets of(cid:0)B(H), B(K, H), B(K)(cid:1) which satisfies all requirements of a C ∗-operator bimodule as in Definition 10.1 except the non-degeneracy requirements AH = H and BK = K, let H ′ := AH ⊆ H and K ′ = BK ⊆ K. Then, via restriction, we obtain a completely isomeric and non-degenerate representation of (A, V, B) on B(K ′, H ′). (3) If ρ = (ρA, ρV , ρB) is a completely bounded representation (or mor- phism) of (A, V, B) with 0 < C := kρV kcb, then 1 C ρV , ρB) is a completely contractive representation (resp. morphism). This easy observa- tion shows that in most situations one may assume without loss of generality that ρ is completely contractive. C ρ := (ρA, 1 The following proposition extends the well-known construction which as- signs to each operator space V ⊆ B(K, H) the Paulsen-operator system X(V ) :=(cid:18)C1H V ∗ C1K(cid:19) ⊆ B(H ⊕K). For this let (A, V, B) be a C ∗-operator V bimodule in B(H, K). Let X(A, V, B) :=(cid:26)(cid:18) a w∗ v b(cid:19) : a ∈ A, v, w ∈ V, b ∈ B(cid:27) ⊆ B(H ⊕ K), and let A⊕B be viewed as the set of diagonal operators(cid:18)a 0 0 b(cid:19) ∈ B(H ⊕K) with a ∈ A, b ∈ B. Then it is easily checked that (A ⊕ B, X(A, V, B)) is a C ∗-operator system in B(H ⊕ K) as defined in Definition 3.1. C ∗-OPERATOR SYSTEMS AND CROSSED PRODUCTS 41 On the other hand, one easily checks that the set of operators Op(A, V, B) :=(cid:26)(cid:18)a v 0 b(cid:19) : a ∈ A, v ∈ V, b ∈ B(cid:27) ⊆ B(H ⊕ K), is a concrete operator algebra in B(H ⊕ K) such that each approximate unit of A ⊕ B serves as an approximate unit of Op(A, V, B). Definition 10.3. We call (A ⊕ B, X(A, V, B)) the Paulsen C ∗-operator sys- tem of (A, V, B) and we call Op(A, V, B) the Paulsen operator algebra of (A, V, B). Proposition 10.4. Let (A, V, B) be a C ∗-operator bimodule. Then there is a one-to-one correspondence between (1) non-degenerate completely contractive representations of (A, V, B); (2) non-degenerate completely positive contractive representations of the C ∗-operator system (cid:0)A ⊕ B, X(A, V, B)(cid:1); and (3) non-degenerate completely contractive operator algebra representa- tions of Op(A, V, B). Given a representation ρ = (ρA, ρV , ρB) : (A, V, B) → B(K, H) the corre- sponding representation π of X(A, V, B) into B(H ⊕ K) is given by π(cid:16)(cid:18) a w∗ v b(cid:19)(cid:17) =(cid:18) ρA(a) ρV (w)∗ ρB(b)(cid:19) a ∈ A, v, w ∈ V, b ∈ B. ρV (v) and given a representation π : X(A, V, B) → B(L), the corresponding rep- resentation of Op(A, V, B) is given by the restriction of π to Op(A, V, B) ⊆ X(A, V, B). Proof. Let ρ = (ρA, ρV , ρB) be a completely contractive representation of (A, V, B) into B(K ′, H ′). We may further assume without loss of generality that A and B are unital -- otherwise we replace A and B by their unitisations A = A + C1H and B = B + C1K and ρA and ρB by their canonical unital extensions to A and B, respectively. Suppose now that T := (cid:18) a w∗ v b(cid:19) ∈ X(A, V, B) is positive. Then w = v and a and b are positive elements in A and B, respectively. Let π be as in the proposition. In order to see that π(T ) is positive, it suffices to show that π(T + ε1) = π(T ) + ε1 is positive for all ε > 0. Writing aε := a + ε1 and 42 MASSOUD AMINI, SIEGFRIED ECHTERHOFF, AND HAMED NIKPEY bε := b + ε1, we get − 1 2 ε 0 0 ≤ a =  b  (cid:18)aε v∗ 0 − 1 2 ε b 1 a − 1 2 ε vb − 1 ε v∗a 2 − 1 2 ε 1   0 − 1 2 b ε − 1 2 ε 0 v bε(cid:19) a   , − 1 2 ε from which it follows that ka − 1 2 ε k ≤ 1. Since ρV is completely contrac- tive, we also have kρV (a )k ≤ 1. It then follows that − 1 2 ε vb − 1 2 ε − 1 2 ε vb π(T + ε1) = ρA(a 0 1 2 ε ) 0 ρB(b 1 2 ε )     1 − 1 ε v∗a 2 − 1 2 ε ) ρV (b ρV (a − 1 2 ε vb − 1 2 ε 1 1 2 ε ) )    ρA(a 0 0 ρB(b 1 2 ε )   is positive as well. A similar computation performed an matrix algebras over X(A, B, V ) then shows that π : X(A, V, B) → B(H ⊕ K) is completely positive. Since π is unital, it is also completely contractive. It is clear that every non-degenerate completely contractive representation of X(A, B, V ) restricts to a nondenerate completely contractive operator algebra representation of Op(A, V, B). So let us finally assume that we have a non-degenerate completely con- tractive operator algebra representation π : Op(A, V, B) → B(L) for some Hilbert space L. Let us regard A, V and B as subspaces of Op(A, V, B) in the canonical way. Then the restrictions πA, πV , πB of π to A, V and B are completely contractive as well. Since πA : A → B(L) and πB : B → B(L) are contractive algebra homomorphisms, it follows from [4, Proposition A.5.8] that they are ∗-homomorphisms. Writing H := πA(A)L and K = πB(B)L we get L = H ⊕ K and (πA, πV , πB) is a non-degenerate representation of (A, V, B) into B(H, K) as in Definition 10.1. (cid:3) Remark 10.5. If we allow possibly degenerate representations of (A, V, B), X(A, V, B) or Op(A, V, B) in the statement of Proposition 10.4 then we can always pass to appropriate subspaces of the representation spaces to make these representations non-degenerate. Then the one-to-one correspondence will still hold modulo the possible addition of direct sums on which all op- erators act trivially. As a direct consequence of Proposition 10.4 we now get Corollary 10.6. Suppose that (A, V, B) and (C, W, D) are C ∗-operator bi- modules. Then there is a one-to-one correspondence between C ∗-OPERATOR SYSTEMS AND CROSSED PRODUCTS 43 (1) completely contractive morphism ϕ : (A, V, B) → (C, W, D), (2) completely positiv contractive morphism φ : X(A, V, B) → X(C, W, D) preserving the corners, and (3) complete contractive homomorphisms ψ : Op(A, V, B) → Op(C, W, D) preserving the corners. If ϕ = (ϕA, ϕV , ϕB) is as morphism from (A, V, B) to (C, W, D) then the corresponding morphism φ : X(A, V, B) → X(C, W, D) is given by φ(cid:16)(cid:18) a w∗ v b(cid:19)(cid:17) =(cid:18) ϕA(a) ϕV (v) ϕV (w)∗ ϕB(b)(cid:19) a ∈ A, v, w ∈ V, b ∈ B, and if φ : X(A, V, B) → X(C, W, D) is a morphism as in (2), then its re- striction ψ to Op(A, V, B) is the corresponding morphism from Op(A, V, B) to Op(C, W, D). The correspondences are compatible with taking compositions of mor- phisms. Proof. It is clear that the constructions given above preserve all required algebraic properties. The combination of Proposition 10.4 with Remark 10.5 shows that they also preserve the property of being completely contractive. (cid:3) Remark 10.7. In what follows it is useful to consider representations of C ∗- operator bimodules into general C ∗-algebras. By such a representation we understand a triple ρ = (ρA, ρV , ρB) of (A, V, B) into a C ∗-algebra C satis- fying (1) ρA : A → C and ρB : B → C are ∗-homomorphisms such that ρA(A)ρB(B) = {0}, (2) ρV : V → C is completely contractive and ρV (bva) = ρB(b)ρV (v)ρA(a) for all a ∈ A, v ∈ V and b ∈ B. Then we have a well-defined ∗-homomorphism ρA⊕ρB : A⊕B → C mapping a ⊕ b to ρA(a) + ρB(b) and we say that ρ is non-degenerate if ρA ⊕ ρB maps approximate units of A ⊕ B to approximate units of C (this is equivalent to the fact that ρA ⊕ ρB(A ⊕ B)C = C). In general we may always pass to the subalgebra C ′ of C generated by ρA(A) ∪ ρV (V ) ∪ ρB(B) to obtain a non-degenerate representation into this subalgebra. Then, representing C (resp. C ′) faithfully and non-degenerately on a Hilbert space L, we may regard ρ as a non-degenerate representation of (A, V, B) in B(K, H) with H = ρA(A)L, K = ρB(B)L. This allows us to use the results of Proposition 10.4 also for representations into C ∗-algebras. Note that conversely any triple of subsets (A, V, B) of a C ∗-algebra C such that A and B are C ∗-subalgebras of C, V is a closed subspace of C, 44 MASSOUD AMINI, SIEGFRIED ECHTERHOFF, AND HAMED NIKPEY AB = {0} and AV = V = V B determines the structure of a C ∗-operator bimodule on (A, V, B) via a faithful representation of C on Hilbert space. Definition 10.8. Let (A, V, B) be a C ∗-operator bimodule. If j = (jA, jV , jB) is a completely isometric representation of (A, V, B) into a C ∗-algebra C such that C is generated by jA(A) ∪ jV (V ) ∪ jB(B) as a C ∗-algebra, then (cid:0)C, (jA, jV , jB)(cid:1) is called a C ∗-hull of (A, V, B). A C ∗-hull (cid:0)C ∗ u(A, V, B), (iA, iV , iB)(cid:1) is called universal if for any com- pletely contractive representation ρ = (ρA, ρV , ρB) of (A, V, B) into some C ∗-algebra D there exists a ∗-homomorphism ρC : C ∗ u(A, V, B) → D such that ρA = ρC◦iA, ρV = ρC◦iV , and ρB = ρC◦iB. e (A, V, B), (kA, kV , kB)(cid:1) is en- On the other hand, we say that a C ∗-hull (cid:0)C ∗ veloping if for any other C ∗-hull (cid:0)C, (jA, jV , jB)(cid:1) there exists a ∗-homomor- phism kC : C → C ∗ e (A, V, B) such that kA = kC◦jA, kV = kC◦jV , and kB = kC◦jB. (The ∗-homomorphisms ρC and kC are then uniquely determined by these properties.) As a consequence, if the universal and the enveloping C ∗-hulls exist, then for any C ∗-hull (cid:0)C, (jA, jV , jB)(cid:1) of (A, V, B) we obtain unique surjective ∗-homomorphisms C ∗ u(A, V, B) ։ C ։ C ∗ e (A, V, B) which commute with the embeddings of (A, V, B) into these C ∗-algebras. Moreover, it follows easily from the universal properties that the universal and enveloping C ∗-hulls are unique up to isomorphism which are compatible with the embeddings of (A, V, B). Proposition 10.9. For each C ∗-operator bimodule (A, V, B) the universal and enveloping C ∗-hulls exist. To be more precise: let (cid:0)C ∗ u(X(A, V, B)), iX(A,V,B)(cid:1) and (cid:0)C ∗ e (X(A, V, B)), kX(A,V,B)(cid:1) denote the universal and enveloping C ∗-hulls of the C ∗-operator system X(A, V, B) and let (iA, iV , iB) and (kA, kV , kB) be the compositions of iX(A,V.B) and kX(A,V,B) with the canonical inclusions of (A, V, B) into X(A, V, B). Then (cid:0)C ∗ u(X(A, V, B)), (iA, iV , iB)(cid:1) and (cid:0)C ∗ e (X(A, V, B)), (kA, kV , kB)(cid:1) are the universal and enveloping C ∗-hulls of (A, V, B). Alternatively, let (cid:0)C ∗ u(Op(A, V, B)), iOp(A,V,B)(cid:1) and (cid:0)C ∗ e (Op(A, V, B)), kOp(A,V,B)(cid:1) C ∗-OPERATOR SYSTEMS AND CROSSED PRODUCTS 45 denote the universal and enveloping C ∗-hulls of the operator algebra Op(A, V, B) as in [4, Propositions 2.4.2 and 4.3.5] and let (iA, iV , iB) and (kA, kV , kB) be the compositions of iOp(A,V.B) and kOp(A,V,B) with the canon- ical inclusions of (A, V, B) into Op(A, V, B). Then (cid:0)C ∗ u(Op(A, V, B)), (iA, iV , iB)(cid:1) and (cid:0)C ∗ e (Op(A, V, B)), (kA, kV , kB)(cid:1) are the universal and enveloping C ∗-hulls of (A, V, B). Proof. The proof is an easy consequence of the definitions of the universal and enveloping C ∗-hulls together Proposition 10.4 and Remark 10.7. So we omit further details. (cid:3) We are now turning our attention to multipliers: Definition 10.10. Let (A, V, B) be a C ∗-operator bimodule which is non- degenerately and completely isometrically represented on B(K, H). Then the multiplier bimodule of (A, V, B) is the triple (cid:0)M (A), M (V ), M (B)(cid:1) in which M (A) and M (B) are the multiplier algebras of the C ∗-algebras A and B, respectively, and where M (V ) = {T ∈ B(K, H) : AT ∪ T B ⊆ V }. Remark 10.11. One easily checks that(cid:0)M (A), M (V ), M (B)(cid:1) is again a C ∗- operator bimodule represented on B(K, H). If one of A or B is unital, we clearly have M (V ) = V . Notice that, similarly to the construction of the multiplier C ∗-operator system (M (A), M (X)) for a given C ∗-operator system (A, X), the space M (V ) heavily depends on the algebras A and B. We retain from using a notation like AMB(V ) to keep things simple. Recall (e.g., from [4]) that for any completely isometrically and faithfully represented operator algebra A ⊆ B(L) the multiplier algebra M (A) can be defined (up to completely isometric isomorphism) as M (A) = {T ∈ B(L) : T A ∪ AT ⊆ A}. Recall also the definition of the multiplier system of a C ∗-operator system as given in Lemma 3.7. of the C ∗-operator bimodule (A, V, B) in B(K, H). Then Proposition 10.12. Let (cid:0)M (A), M (V ), M (B)(cid:1) be the muliplier bimodule (cid:0)M (A ⊕ B), M (X(A, V, B))(cid:1) =(cid:0)M (A) ⊕ M (B), X(cid:0)M (A), M (V ), M (B)(cid:1)(cid:1) is the multiplier system of the C ∗-operator system (cid:0)A ⊕ B, X(A, V, B)(cid:1) and Op(cid:0)M (A), M (V ), M (B)(cid:1) = M(cid:0)Op(A, V, B)(cid:1). 46 MASSOUD AMINI, SIEGFRIED ECHTERHOFF, AND HAMED NIKPEY Proof. Recall from Lemma 3.7 that M (X(A, V, B)) is defined as the set of all elements T ∈ B(H ⊕ K) such that (A ⊕ B)T ∪ T (A ⊕ B) ⊆ X(A, V, B). Writing T =(cid:18)T11 T12 and computing diag(a, 0)T, T diag(a, 0), diag(0, b)T, T diag(0, b) ∈ X(A, V, B) T21 T22(cid:19) ∈(cid:18) B(H) easily shows that T ∈(cid:18) M (A) M (V ) M (V )∗ M (B)(cid:19) = X(cid:0)M (A), M (V ), M (B)(cid:1). Con- versely one easily checks that X(cid:0)M (A), M (V ), M (B)(cid:1) ⊆ M (X(A, V, B)). A similar argument also shows Op(cid:0)M (A), M (V ), M (B)(cid:1) = M(cid:0)Op(A, V, B)(cid:1). (cid:3) B(K, H) B(K) (cid:19) B(H, K) Definition 10.13. Let (A, V, B) and (C, W, D) be two C ∗-operator bimod- ules. A generalized morphism from (A, V, B) to (C, W, D) is a completely contractive morphism ϕ : (A, V, B) → (M (C), M (W ), M (D)) such that ϕA(A)C = C and ϕB(B)D = D. Example 10.14. Every non-degenerate representation ρ of (A, V, B) to some B(K, H) can be regarded as a generalized morphism from (A, V, B) to the C ∗-operator bimodule (cid:16)M (K(H)), M (K(K, H)), M (K(K))(cid:17). The following proposition is now a direct combination of Proposition 10.4, Proposition 10.12 and Lemma 3.11, so we leave the details to the reader: Proposition 10.15. Every generalized morphism ϕ : (A, V, B) → (M (C), M (W ), M (D)) from (A, V, B) to (C, W, D) extends uniquely to a morphism ϕ : (M (A), M (V ), M (B)) → (M (C), M (W ), M (D)). If ϕ is completely isometric, then so is ϕ. In particular, every non-degenerate (completely isometric) representation ρ of (A, V, B) into B(K, H) uniquely extends to a (completely isometric) representation of (cid:0)M (A), M (V ), M (B)(cid:1) to B(K, H). We close this section with the following analogue of Lemma 4.10: Proposition 10.16. Let (cid:0)C, (jA, jV , jB)(cid:1) be a C ∗-hull of (A, V, B). Then the inclusions j = (jA, jV , jB) : (A, V, B) → C extend to a completely iso- metric morphism ¯j := (¯jM (A), ¯jM (V ), ¯jM (B)) : (M (A), M (V ), M (B)) → M (C) such that ¯jM (A)(M (A)) ∩ C = jA(A) and ¯jM (B)(M (B)) ∩ C = jB(B). C ∗-OPERATOR SYSTEMS AND CROSSED PRODUCTS 47 Proof. Let jX(A,V,B) : X(A, V, B) → C denote the corresponding completely isometric representation of X(A, V, B) into C. Then (cid:0)C, jX(A,V,B)(cid:1) is a C ∗- hull of X(A, V, B) and by Lemma 4.10 there exists a unique extension (cid:0)¯jM (A⊕B), ¯jM (X(A,V,B))(cid:1) :(cid:0)M (A ⊕ B), M (X(A, V, B)(cid:1) → M (C) such that (¯jM (A⊕B)(M (A) ⊕ B)) ∩ C = A ⊕ B. The result now easily follows from an application of Proposition 10.12. (cid:3) 11. Crossed products by C ∗-operator bimodules g , αV g , αB C ∗-operator bimodule dynamical system. g ) such that all components g 7→ αA Let (A, V, B) be a C ∗-operator bimodule and let Aut(A, V, B) denote the group of completely isometric isomorphisms of (A, V, B) to (A, V, B). A con- tinuous action of G on (A, V, B) is a homomorphism α : G → Aut(A, V, B) with αg = (αA g (b) are continuos for all a ∈ A, v ∈ V, b ∈ B. We then call (cid:0)(A, V, B), G, α(cid:1) a A covariant morphism of (cid:0)(A, V, B), G, α(cid:1) to the C ∗-operator bimodule (C, W, D) is a quintuple (ρA, ρV , ρB, u, v) such that (ρA, ρV , ρB) is a mor- phism from (A, V, B) into (C, W, D), u : G → U M (C) and v : G → U M (D) are strictly continuous unitary representations of G such that (ρA, u) and (ρB, v) satisfy the usual cavariance conditions for the actions αA and αB, respectively, and such that for all v ∈ V and g in G we have g (v), αB g (a), αV ρV (αg(v)) = ugρV (v)vg−1 . in the sense of Definition 10.13. A covariant representation of (A, V, B) is a covariant morphism into A generalized covariant morphism of (cid:0)(A, V, B), G, α(cid:1) to (C, W, B) is a co- variant morphism (ρA, ρV , ρB, u, v) into (cid:0)M (C), M (W ), M (D)(cid:1) such that (ρA, ρV , ρB) : (A, V, B) →(cid:0)M (C), M (W ), M (D)(cid:1) is a generalized morphism (cid:0)B(H), B(K, H), B(K)(cid:1) for some pair of Hilbert spaces (H, K). (ρA ⋊ u, u ⋉ ρV ⋊ v, ρB ⋊ v) :(cid:0)Cc(G, A), Cc(G, V ), Cc(G, B)(cid:1) → (C, W, D) If (ρA, ρV , ρB, u, v) is covariant morphism of (A, V, B) into (C, W, D), we in which ρA ⋊u and ρB ⋊v are the usual integrated forms of the covariant ho- momorphisms (ρA, u) and (ρB, v) of the systems (A, G, αA) and (B, G, αB ), respectively, and where u ⋉ ρV ⋊ v : Cc(G, V ) → W is given by have integrated forms u ⋉ ρV ⋊ v(f ) =ZG ρV (f (s))vs ds =ZG usρV(cid:0)αV s−1(f (s))(cid:1) ds, where the right equation follows from the covariance condition (this should explain the notation v ⋉ ρV ⋊ u). We have the usual convolution products 48 MASSOUD AMINI, SIEGFRIED ECHTERHOFF, AND HAMED NIKPEY on Cc(G, A) and Cc(G, B) and obvious convolution formulas for pairings Cc(G, A) × Cc(G, V ) → Cc(G, V ) and Cc(G, V ) × Cc(G, B) → Cc(G, V ) which are preserved by the integrated form (ρA ⋊ u, v ⋉ ρV ⋊ u, ρB ⋊ v) of (ρA, ρV , ρB, u, v). The following is a direct consequence of Corollary 10.6 and the universal properties of the universal and the enveloping C ∗-hulls of (A, V, B). Proposition 11.1. Let (A, V, B) be a C ∗-operator bimodule and let u(A, V, B), (iA, iV , iB)(cid:1) and (cid:0)C ∗ (cid:0)C ∗ e (A, V, B), (kA, kV , kB)(cid:1) denote the uni- versal and enveloping C ∗-hulls of (A, V, B), respectively. Then there is a canonical one-to-one correspondence between (1) continuous actions α : G → Aut(A, V, B), (2) continuous actions αX : G → Aut(cid:16)(cid:0)A ⊕ B, X(A, V, B)(cid:1)(cid:17) which (3) continuous operator algebra actions αOp : G → Aut(cid:0)Op(A, V, B)(cid:1) which preserves the corners, preserve the corners, (4) continuous actions αu : G → Aut(C ∗ u(A, V, B)) by automorphisms which preserve the subspaces iA(A), iV (V ) and iB(B). (5) continuous actions αe : G → Aut(C ∗ e (A, V, B)) by automorphisms which preserve the subspaces kA(A), kV (V ) and kB(B). Moreover, there are one-to-one correspondences between the covariant repre- sentations (morphisms) (ρA, ρV , ρB, u, v) of (cid:0)(A, V, B), G, α(cid:1) and covariant representations (morphisms) of the actions in (2), (3), and (4) above via the known correspondence for representations (morphisms) of (A, V, B) and X(A, V, B), Op(A, V, B) and C ∗ u(A, V, B) with unitary parts given by the di- rect sum u ⊕ v. We now give the definition of the universal crossed product by an action of a locally compact group G on a C ∗-operator bimodule: Definition 11.2. Let α : G → Aut(A, V, B) be a strongly continuous action of the locally compact group G. We define the universal crossed product (A, V, B) ⋊u α G :=(cid:0)A ⋊u α G, V ⋊u α G, B ⋊u α G(cid:1) of (A, V, B) by G as the respective closures of(cid:0)Cc(G, A), Cc(G, V ), Cc(G, B)(cid:1) u(A, V, B) ⋊α,u G (here we identify A, V and B with iA(A), iV (V ) inside C ∗ and iB(B) inside C ∗ u(A, V, B), respectively). To see that (A, V, B) ⋊u module, let (iC ∗ product C ∗ all GNS representations associated to the states of C ∗ α G has a canonical structure of a C ∗-operator bi- u(A,V,B), iG) denote the universal representation of the crossed u(A, V, B) ⋊α,u G an the Hilbert space Lu, i.e., the direct some of u(A, V, B) ⋊α,u G. Then C ∗-OPERATOR SYSTEMS AND CROSSED PRODUCTS 49 u(A,V,B) restricts to there exists a decomposition Lu = Hu ⊕ Ku such that iC ∗ a representation of (A, V, B) on B(Ku, Hu) . It is then easy to check that the covariant representation (iC ∗ u(A,V,B), iG) restricts to the covariant repre- sentation (iA, iV , iB, iA restrictions of iC ∗ the respective corners of B(Hu ⊕ Ku)), and iA G) of (cid:0)(A, V, B), G, α(cid:1) where iA, iV , iB denote the u(A,V,B) to A, V and B, respectively (viewed as operators in G := iGHu , iB G := iGKu. G, iB G, iB α G, B ⋊u α G, V ⋊u ⋉ iV ⋊ iA α G), M (B ⋊u α G), M (V ⋊u The integrated form iC ∗ G : Cc(G, A) → B(Hu), iB G u(A,V,B) ⋊ iG restricts to the integrated forms iA ⋊ iA G : Cc(G, V ) → B(Ku, Hu) and iB ⋊ iB G : Cc(G, B) → B(Ku) and similarly for the respective com- pletions. They therefore extend to a completely isometric representation of Moreover, it is easy to check that (iA, iV , iB, iA G), M (B⋊u We call (iA, iV , iG, iB, iA α G(cid:1) as a concrete C ∗-operator bimodule in B(Ku, Hu). (cid:0)A ⋊u (cid:0)M (A⋊u α G)(cid:1). Hence (iA, iV , iG, iB, iA eralized covariant morphism of (cid:0)(A, V, B), G, α(cid:1) into (cid:0)M (A ⋊u G) take their values in G, iB G) is a gen- α G), M (V ⋊u α αG, B⋊u αG)(cid:1) which integrates to the identity on(cid:0)A⋊u αG(cid:1). G) the universal morphism of (cid:0)(A, V, B), G, α(cid:1). properties for covariant morphisms (representations) of (cid:0)(A, V, B), G, α(cid:1). every generalized covariant morphism (ρA, ρV , ρB, u, v) of (cid:0)(A, V, B), G, α(cid:1) into (cid:0)M (C), M (W ), M (D)(cid:1) the integrated form (ρA ⋊ u, u ⋉ ρV ⋊ v, ρB ⋊ Proposition 11.3. Let α : G → Aut(A, V, B) be a continuous action. For v) from (Cc(G, A), Cc(G, V ), Cc(G, D)) into (M (C), M (W ), M (D)) extends uniquely to a morphism The following proposition shows that (A, V, B)⋊u αG has the right universal αG, V ⋊u G, iB (ρA⋊u, u⋉ρV ⋊v, ρB ⋊v) : (A⋊u αG, V ⋊u αG, B⋊u αG) → (M (C), M (W ), M (D)). (which takes values in (C, W, D) if (ρA, ρV , ρB) does). If (ρA, ρV , ρB, u, v) is non-degenerate, then so is (ρA ⋊ u, u ⋉ ρV ⋊ v, ρB ⋊ v). Conversely, for every generalized morphism (πA⋊u (A⋊u alized covariant morphism (ρA, ρV ρB, u, v) of αG) of αG) into (M (C), M (W ), M (D)) there is a unique gener- into αG, V ⋊u αG, B⋊u αG, πV ⋊u αG, πB⋊u (cid:0)(A, V, B), G, α(cid:1) (cid:0)M (C), M (W ), M (D)(cid:1) such that (πA⋊u αG, πV ⋊u αG, πB⋊u αG) = (ρA ⋊ u, u ⋉ ρV ⋊ v, ρB ⋊ v) given by the composition of (πA⋊u as in Proposition 10.15) with the universal representation (iA, iV , iG, iB, iA αG) (extended to multipliers G, iB αG, πV ⋊u αG, πB⋊u G). Proof. Starting with (ρA, ρV , ρB, u, v) we obtain a corresponding covariant homomorphism (ρC ∗ (use Propositions 10.16 and 11.1). By the universal property of the maximal crossed product C ∗ u(A, V, B) ⋊α,u G we obtain the integrated form u(A, V, B), G, αu) into M (C ∗ u(A,V,B), u⊕v) of (C ∗ u(C, W, D)) ρC ∗ u(A,V,B) ⋊ u ⊕ v : C ∗ u(A, V, B) ⋊α,u G → M (C ∗ u(C, W, D)) 50 MASSOUD AMINI, SIEGFRIED ECHTERHOFF, AND HAMED NIKPEY whose restriction to (Cc(G, A), Cc(G, V ), Cc(G, B)) coincides with the inte- grated form (ρA⋊u, u⋉ρV ⋊v, ρB ⋊v) with values in (M (C), M (W ), M (D)) ⊆ M (C ∗ α G) as desired. u(C, W, D)). They therefore extend to (A ⋊u α G, V ⋊u α G, B ⋊u For the converse we need to show that the integrated form of the covariant αG) on morphism (ρA, ρV ρB, u, v) obtained by composing (πA⋊u with (cid:0)Cc(G, A), Cc(G, V ), Cc(G, B)(cid:1). But this follows from a straightforward computation which we omit. (iA, iV , iG, iB, iA agrees with αG, πV ⋊u αG, πB⋊u αG, πV ⋊u αG, πB⋊u G, iB G) (πA⋊u αG) (cid:3) Recall from Proposition 10.9 that for a C ∗-operator bimodule (A, V, B) we have the identities u(A, V, B) ∼= C ∗ C ∗ u(Op(A, V, B)) ∼= C ∗ u(X(A, V, B)) where the first isomorphism is given by the universal property of C ∗ u(A, V, B) applied to the canonical (corner) inclusions of (A, V, B) into Op(A, V, B) ⊆ C ∗ u(Op(A, V, B)) and the second isomorphism is given by the universal prop- erty of C ∗ u(Op(A, V, B)) applied to the canonical inclusion of Op(A, V, B) into X(A, V, B). If α : G → Aut(A, V, B) is an action, then these isomor- phism are G-equivariant, where C ∗ u(Op(A, V, B)) is equipped with the action extending αOp and C ∗ u(X(A, V, B)) is equipped with the action extending αX, where αOp and αX are as in Proposition 11.1. u(A)⋊α,uG and in Definition 6.1 we defined the crossed product X ⋊u In [20] Katsoulis and Ramsay defined the universal crossed product of the operator algebra system (A, G, α) as the closure of Cc(G, A) inside C ∗ αG for an action α on a C ∗-operator system X as the closure X ⋊u α G of Cc(G, X) inside C ∗ u(X) ⋊α,u G (surpressing the C ∗-part of the C ∗-operator system in our notation). Thus, identifying (Cc(G, A), Cc(G, V ), Cc(G, B)) with the three non-zero corners of Cc(G, Op(A, V, B)) and the latter as a subspace of Cc(G, X(A, V, B)) we see that these inclusions extend to completely iso- metric inclusions Op(A ⋊u α G, V ⋊u ⊆ X(A, V, B) ⋊u α G, B ⋊u α G = X(A ⋊u α G) = Op(A, V, B) ⋊u α G, B ⋊u α G, V ⋊u α G α G) Together with Corollary 6.12 and Proposition 10.9 we obtain isomorphisms u(A, V, B) ⋊α,u G ∼= C ∗ C ∗ Corollary 6.12 u(Op(A, V, B)) ⋊α,u G ∼= C ∗ u(X(A, V, B)) ⋊α,u G ∼= ∼= C ∗ ∼= C ∗ u(cid:0)A ⋊u u(cid:0)Op(A ⋊u C ∗ u(cid:0)X(A, V, B) ⋊u α G, V ⋊u α G, B ⋊u α G(cid:1) ∼= C ∗ u(cid:0)X(A ⋊u α G(cid:1) α G)(cid:1) α G, B ⋊u α G, V ⋊u α G, V ⋊u α G, B ⋊u α G)(cid:1) C ∗-OPERATOR SYSTEMS AND CROSSED PRODUCTS 51 In particular we see that the theories for universal crossed product by cor- responding actions of G on (A, V, B), Op(A, V, B) and X(A, V, B) are com- pletely equivalent! We close this section with a brief discussion of the reduced crossed product for an action α : G → Aut(A, V, B). The easiest way to do this at this point is to form the reduced crossed product X(A, V, B) ⋊r α G ⊆ M(cid:0)X(A, V, B) ⊗ K(L2(G))(cid:1) α G, V ⋊r as the image of the regular representation ΛX(A,V,B) : X(A, V, B) ⋊u M(cid:0)X(A, V, B) ⊗ K(L2(G))(cid:1) as in Definition 7.1 and define (A, V, B) ⋊r (cid:0)A ⋊r α G → α G = α G, V ⋊u α G, B ⋊u α G (which is the same as taking closures of (Cc(G, A), Cc(G, V ), Cc(G, B)) inside X(A, V, B) ⋊r α G). We leave it as an exercise to the reader to formulate this in terms of a regular covariant α G(cid:1) via the images of the corners (A ⋊u α G) inside X(A, V, B) ⋊r α G, B ⋊r G, V ⋊r (Cc(G, A), Cc(G, V ), Cc(G, B)) K = K(L2(G)) and where, as usual, "⊗" denotes the spacial tensor product. representation of(cid:0)(A, V, B), G, α(cid:1) into(cid:0)M (A⊗K), M (V ⊗K), M (B⊗K)(cid:1) for It follows from our construction and part (b) of Remark 7.3 that (cid:0)A ⋊r α G(cid:1) is completely isometrically isomorphic to the closures of (cid:0)C, (jA, jV , jB)(cid:1) of (A, V, B) (where we identify (A, V, B) with the triple (jA(A), jV (V ), jB(B)) inside C) which carries an action αC which is com- patible with the given action on (A, V, B). In particular, we may take the closures inside C ∗ e (A, V, B) ⋊αe,r G. From this we get u(A, V, B) ⋊αu,r G or C ∗ any C ∗-hull αC ,r G for inside C ⋊ α G, B ⋊r α Proposition 11.4. Let α : G → Aut(A, V, B) be an action by an amenable group G. Then (A ⋊u α G, V ⋊u α G, B ⋊u α G, V ⋊r α G, B ⋊r α G) =(cid:0)A ⋊r α G(cid:1) via the regular representation. Proof. This follows from the above discussion and the fact that C ∗ u(A, V, B) ⋊αu,u G ∼= C ∗ u(A, V, B) ⋊αu,r G if G is amenable. (cid:3) 12. Coactions and duality In this section we want to discuss the duality theorems for crossed prod- ucts for C ∗-operator bimodules. The theory is more or less a direct conse- quence of the theory for C ∗-operator systems vie the functor (A, V, B) 7→ X(A, V, B), so we'll try to be brief. Note that if (A, V, B) is a C ∗-operator system represented completely isometrically on the pair of Hilbert spaces (H, K) and if C is any C ∗-algebra which is represented faithfully on a Hilbert space L, then we can define the spatial tensor product (A ⊗ C, V ⊗ C, B ⊗ C) 52 MASSOUD AMINI, SIEGFRIED ECHTERHOFF, AND HAMED NIKPEY as the closures of the canonical inclusions of the algebraic tensor products (cid:0)A ⊙ C, V ⊙ C, B ⊙ C(cid:1) inside(cid:0)B(H ⊗ L), B(K ⊗ L, H ⊗ L), B(K ⊗ L)(cid:1). One then checks that X(A, V, B) ⊗ C = X(A ⊗ C, V ⊗ C, B ⊗ C) (and, similarly Op(A, V, B) ⊗ C = Op(A ⊗ C, V ⊗ C, B ⊗ C)). In what follows we often write (A, V, B) ⊗ C for the C ∗-operator bimodule (A ⊗ C, V ⊗ C, B ⊗ C) and we write M(cid:0)(A, V, B) ⊗ C(cid:1) for the multiplier bimodule (cid:0)M (A ⊗ C), M (V ⊗ C), M (B ⊗ C))(cid:1). Definition 12.1. Let (A, V, B) be a C ∗-operator bimodule. A coaction of the locally compact group G on (A, V, B) is a generalized morphism δ(A,V,B) = (δA, δV , δB) : (A, V, B) → M(cid:0)(A, V, B) ⊗C ∗(G)(cid:1) such that the following hold: (1) the maps δA : A → M (A ⊗C ∗(G)) and δB : B → M (B ⊗C ∗(G)) are coactions of G on the C ∗-algebras A and B, respectively. (2) The following diagram of generalized morphism commutes: δ(A,V,B)⊗idG (A, V, B) δ(A,V,B) −−−−−→ δ(A,V,B)y M(cid:0)(A, V, B) ⊗C ∗(G)(cid:1) −−−−−−−−→ M(cid:0)(A, V, B) ⊗C ∗(G)(cid:1) yid(A,V,B) ⊗δG M(cid:0)(A, V, B) ⊗C ∗(G) ⊗C ∗(G)(cid:1) M(cid:0)(A, V, B) ⊗C ∗(G)(cid:1) with generalized morphism from X(A, V, B) Using the correspondence of generalized morphisms from (A, V, B) to to X(M (A, V, B)) of Corollary 10.6 and the isomorphism X(M (A, V, B)) ∼= M (X(A, V, B)) of Proposition 10.12 we see that every coaction δ(A,V,B) of G on (A, V, B) as in the definition above determines a coaction δX(A,V,B) of G on X(A, V, B) and vice versa. We then call δ(A,V,B) non-degenerate iff δX(A,V,B) is non-degenerate in the sense of Definition 8.5. G and X(A, V, B) ⋊r α G = X(cid:0)(A, V, B) ⋊u Example 12.2. Recall that for each action α : G → Aut(A, V, B) there corresponds a unique action (which here we also denote by α) of G on X(A, V, B) such that X(A, V, B) ⋊u ilarly for the reduced crossed products). Recall from Example 8.3 that α G(cid:1) (and sim- there exist canonical dual coactions bαu and bαr of G on X(A, V, B) ⋊u α G(cid:1) and using the correspondence between coactions on X(cid:0)(A, V, B) ⋊u α G(cid:1) (and similarly for the α G and coactions on X(cid:0)(A, V, B) ⋊u reduced crossed products), we obtain dual coactions αu and bαr on the full and reduced crossed products of (A, V, B) by G, respectively. We leave it to the reader to spell out direct formulas for these coactions. Identifying X(A, V, B) ⋊u α G, respectively. α α G with (A, V, B) ⋊u C ∗-OPERATOR SYSTEMS AND CROSSED PRODUCTS 53 Definition 12.3. Let δ(A,V,B) be a coaction of G on the C ∗-operator bi- module (A, V, B). Then a (generalized) covariant morphism of the co- system (cid:0)(A, V, B), G, δ(A,V,B)(cid:1) into the muliplier bimodule M (C, W, D) = (cid:0)M (C), M (W ), M (B)(cid:1) of a C ∗-operator bimodule (C, W, D) consists of a quintuple (cid:0)ρA, ρV , ρB, µ, ν(cid:1) such that (1) ρ = (ρA, ρV , ρB) : (A, V, B) → (cid:0)M (C), M (W ), M (B)(cid:1) is a general- ized morphism of (A, V, B); (2) µ : C0(G) → M (C), ν : C0(G) → M (D) are non-degenerate ∗- homomorphisms; (3) (ρA, µ) and (ρB, ν) are covariant for (A, G, δA) and (B, G, δB), re- spectively; and (4) (ρV ⊗ idG) ◦ δV (v) = (cid:0)µ ⊗ idG(wG)(cid:1)(ρV (v) ⊗ 1)(cid:0)ν ⊗ idG(wG)(cid:1)∗ all v ∈ V . for where wG ∈ C b st(G, M (C ∗(G)) ∼= M (C0(G) ⊗C ∗(G)) is the strictly continu- ous function wG(g) = iG(g). A covariant representation of(cid:0)(A, V, B), G, δ(A,V,B)(cid:1) on the pair of Hilbert spaces (H, K) is a morphism into(cid:0)B(H), B(K, H), B(K)(cid:1). It is now an easy exercise to see that there is a one-to-one correspondence between covariant morphism of (A, V, B) into M (C, V, D) and covariant mor- phisms of(cid:0)X(A, V, B), G, δX(A,V,B)(cid:1) into M (X(C, V, D)) given by assigning to (cid:0)ρA, ρV , ρB, µ, ν(cid:1) the covariant pair (cid:0)ρX(A,V,B), µ ⊕ ν(cid:1) with ρB(cid:19) ρX(A,V,B) =(cid:18)ρA ρV ρ∗ V as in Corollary 10.6. Moreover, using the identity C ∗ u(X(A, V, B)) and Proposi- tion 8.4, we deduce easily that there is a one-to-one correspondence between coactions δ(A,V,B) of G on (A, V, B) and coactions δu of G on C ∗ u(A, V, B) which satisfy the conditions u(A, V, B) ∼= C ∗ δu(A) ⊆ M (A ⊗C ∗(G)), δu(V ) ⊆ M (V ⊗C ∗(G)), and δu(B) ⊆ M (B ⊗C ∗(G)), where we understand these inclusions with respect to the canonical inclu- u(A, V, B) and of M ((A, V, B) ⊗C ∗(G)) into sions of (A, V, B) into C ∗ M (C ∗ u(A, V, B) ⊗C ∗(G)) which can be deduced from Lemma 4.10). Example 12.4. The regular representation (cid:0)(A, V, B), G, δ(A,V,B)(cid:1) is the covariant morphism from(cid:0)(A, V, B), G, δ(A,V,B)(cid:1) into M(cid:0)(A, V, B) ⊗ K(L2(G))(cid:1) defined as the quintuple (cid:0)ΛA, ΛV ,ΛB, ΛA =(cid:0)(idA ⊗λ) ◦ δA, (idA ⊗λ) ◦ δA, (idA ⊗λ) ◦ δA, 1A ⊗ M, 1B ⊗ M(cid:1) bG(cid:1) , ΛB bG of the co-system 54 MASSOUD AMINI, SIEGFRIED ECHTERHOFF, AND HAMED NIKPEY where λ = λG denotes the regular representation of G on L2(G) and M : C0(G) → B(L2(G)) is the representation by multiplication operators. One easily checks that this representation corresponds to the regular representa- tion of the co-system(cid:0)X(A, V, B), G, δX(A,V,B)(cid:1) via the above described cor- respondence. In particular, it is a covariant morphism of(cid:0)(A, V, B), G, δ(A,V,B)(cid:1). We are now ready to define the crossed products Proposition 12.5. Let δ(A,V,B) be a coaction of G on the C ∗-operator bi- module (A, V, B). We then define the crossed product δ(A,V,B) bG =(cid:0)A ⋊δA bG, V ⋊δV bG, B ⋊δB bG(cid:1) (C0(G))}, B⋊δB bG := span{ΛB(B)ΛB (C0(G))} = span{ΛB bG bG bG (C0(G))}, (C0(G)ΛV (V )} (A, V, B) ⋊ as bG A⋊δAbG := span{ΛA(A)ΛA and V ⋊δV bG = span{ΛV (V )ΛA inside M(cid:0)(A, V, B) ⊗ K(L2(G))(cid:1). ) Note that it follows directly from the definitions that (ΛA, ΛA bG ) and (ΛB, ΛB bG are the regular representations of (A, G, δA) and (B, G, δB), respectively. the crossed products of these C ∗-co-systems as described in Section 8. Us- ing the above described correspondence between covariant morphisms of We therefore see that the C ∗-algebras A ⋊δA bG and B ⋊δB bG coincide with (cid:0)(A, V, B), G, δ(A,V,B)(cid:1) and covariant morphisms of(cid:0)X(A, V, B), G, δX(A,V,B)(cid:1) Theorem 12.6. The crossed product (A, V, B) ⋊δ(A,V,B) bG is a well defined C ∗-operator bimodule such that we now get from Proposition 8.16: and Moreover, the pair C ∗ X(cid:0)(A, V, B) ⋊δ(A,V,B) bG(cid:1) = X(A, V, B) ⋊δX(A,V,B) bG. u(cid:0)(A, V, B) ⋊δ(A,V,B) bG(cid:1) = C ∗ u(A, V, B) ⋊δu bG. bG(cid:1)(cid:17) (cid:16)(A, V, B) ⋊ δ(A,V,B) bG,(cid:0)ΛA, ΛV , ΛB, ΛA , ΛB bG satisfies the following universal property for covariant morphisms: If (cid:0)ρA, ρV , ρB, µ, ν(cid:1) is any covariant morphism of (cid:0)(A, V, B), G, δ(A,V,B)(cid:1) into M (C, W, D) then there exists a unique covariant morphism (ρA ⋊ µ, µ ⋉ ρV ⋊ ν, ρB ⋊ ν) : (A, V, B) ⋊ such that δ(A,V,B) bG → M (C, W, D) ρA = (ρA ⋊ µ) ◦ ΛA, ρV = (µ ⋉ ρV ⋊ ν) ◦ ΛV , ρB = (ρB ⋊ ν) ◦ ΛB, µ = (ρA ⋊ µ) ◦ ΛA bG and ν = (ρB ⋊ ν) ◦ ΛB bG . C ∗-OPERATOR SYSTEMS AND CROSSED PRODUCTS 55 from (A, V, B) ⋊ Conversely, if Φ =(cid:0)ΦA⋊δA variant morphism (cid:0)ρA, ρV , ρB, µ, ν(cid:1) of (cid:0)(A, V, B), G, δ(A,V,B)(cid:1) such that bG(cid:1) is any generalized morphism δ(A,V,B) bG into M (C, W, D) then there exists a unique co- bG, ΦB⋊δB bG, ΦV ⋊δV Φ = (ρA ⋊ µ, µ ⋉ ρV ⋊ ν, ρB ⋊ ν) Remark 12.7. Let σ : G → Aut(C0(G)) denote action given by right trans- lation. Then there is a dual action bδ : G → Aut(cid:0)(A, V, B) ⋊δ bG(cid:1) such that for each s ∈ G the automorphism bδs is given by the integrated form of ◦ σ(s)(cid:1)(cid:17). Of course, it the covariant morphism (cid:0)ΛA, ΛV , ΛB, ΛA corresponds to the dual action on X(A, V, B) ⋊δX bG. We now come to the duality theorems. We start with the C ∗-operator bimodule version of the Imai-Takai duality theorem. Using the version of the Imai-Takai theorem for actions on C ∗-operator systems, Theorem 9.2, we now get ◦ σ(s), ΛB bG bG Theorem 12.8. Let α : G → Aut(A, V, B) be an action. Then there exist canonical dual coactions bαu (resp. bαr) of G on the universal and reduced crossed products (A, V, B) ⋊u α G, respectively, such that α G and (A, V, B) ⋊r and (A, V, B) ⋊u (A, V, B) ⋊r α G ⋊ bαu bG ∼= (A, V, B) ⊗ K(L2(G)) α G ⋊ bαr bG ∼= (A, V, B) ⊗ K(L2(G)), and the isomorphism transforms the double dual actions cbαu and cbαr to the action α ⊗ Ad ρ on (A, V, B) ⊗ K(L2(G)), where ρ : G → U (L2(G)) denotes the right regular representation of G. Dually, as an application of Theorem 9.3 we get the following version of Katayama's duality for coactions on C ∗-operator bimodules: Theorem 12.9. Let δ = δ(A,V,B) be a coaction of G on the C ∗-operator bimodule (A, V, B). Then there exist is a surjective morphism which factors through an isomorphism bδ Θ : (A, V, B) ⋊δ bG ⋊u (A, V, B) ⋊δ bG ⋊µ bδ G ։ (A, V, B) ⊗ K(L2(G)) G ∼= (A, V, B) ⊗ K(L2(G)), where (A, V, B) ⋊δ bG ⋊µ spect to a norm which lies between the universal and reduced crossed product norms. If G is amenable, then Θ is an isomorphism. G is a completion of Cc(cid:0)G, (A, V, B) ⋊δ bG(cid:1) with re- bδ 56 MASSOUD AMINI, SIEGFRIED ECHTERHOFF, AND HAMED NIKPEY References 1. M. Amini, S. Echterhoff, and H. Nikpey, Crossed products of operator spaces, arXiv:1512.07776. 2. S. Baaj and G. Skandalis, C ∗-alg`ebres de Hopf et th´eorie de Kasparov ´equivariante, K-Theory 2 (1989), 683 -- 721. 3. S. Baaj and G. Skandalis, Unitaires multiplicatifs et dualit´e pour les produits crois´es de C ∗-alg`ebres , Ann. Sci. ´Ecole Norm. Sup. (4) 26 (1993), no. 4, 425 -- 488. 4. D. P. Blecher and C. Le Merdy, Operator algebras and their modules-an operator space approach, London Mathematical Society Monographs, New Series 30, Oxford University Press, Oxford, 2004. 5. D. P. Blecher and V. I. Paulsen, Multipliers of operator spaces, and the injective envelope, Pacific J. Math. 200 (2001), 1 -- 17. 6. J. Brodzki, C. Cave, and K. Li, Exactness of locally compact groups, Adv. Math., 312 (2017), 209 -- 233. 7. F. Combes, Crossed products and Morita equivalence, Proc. London Math. Soc. 49 (1984), 289 -- 306. 8. M.D. Choi and E.G. Effros, Injectivity and operator spaces, J. Funct. Anal. 24 (1977), 156 -- 209. 9. M.D. Choi and E.G. Effros, The completely positive lifting problem for C ∗-algebras, Ann. Math. 104 (1976), 585 -- 609. 10. A. Buss, S. Echterhoff and R. Willett, Exotic crossed products and the Baum-Connes conjecture, to appear in J. reine angew. Math., arXiv: 1409.4332v2. 11. S. Echterhoff, S. Kaliszewski, and J. Quigg, Maximal coactions, International J. Math. 15 (2004), 47 -- 61. 12. S. Echterhoff, S. P. Kaliszewski, J. Quigg, and I. Raeburn, A categorical approach to imprimitivity theorems for C ∗-dynamical systems, Mem. Amer. Math. Soc. 180 (2006) no. 850, pp. viii+169. 13. E. G. Effros and Z.-J. Ruan, Operator spaces, London Math. Soc. Monographs, New Series 23, Oxford University Press, New York, 2000. 14. M. Hamana, Injective envelopes of operator systems, Publ. RIMS, Kyoto Univ. 15 (1979), 773 -- 785. 15. M. Hamana, Triple envelopes and Shilov boundaries of operator spaces, Math. J. Toyama Univ. 22 (1999), 77 -- 93. 16. Samuel J. Harris and Se-Jin Kim, Crossed products of operator systems, J. Funct. Anal. 276 (2019), 2156 -- 2193. 17. S. Kaliszewski, M. B. Landstad, and J. Quigg, Exotic group C ∗-algebras in noncom- mutative duality, New York J. Math. 19 (2013), 689 -- 711. 18. G. G. Kasparov, Equivariant KK-theory and the Novikov conjecture, Invent. Math. 91 (1988), 147 -- 201. 19. Y. Katayama, Takesaki's duality for a non-degenerate co-action, Math. Scand. 55 (1985), 141 -- 151. 20. E.G. Katsoulis and C. Ramsey, Crossed products of operator algebras, Mem. Amer. Math. Soc. 258 (2019), vii+85 pages. 21. A. Kavruk, V. I. Paulsen, I. G. Todorov, and M. Tomforde, Tensor products of operator systems, J. Funct. Anal. 261 (2011), 267 -- 299. 22. E. Kirchberg and S. Wassermann. C ∗-algebras generated by operator systems, J. Funct. Anal. 155 (1998), 324 -- 351. C ∗-OPERATOR SYSTEMS AND CROSSED PRODUCTS 57 23. J. Kustermans and S. Vaes, Locally compact quantum groups. Ann. Sci. ´Ecole Norm. Sup. (4) 33 (2000), no. 6, 837 -- 934. 24. M.B. Landstad, Duality theory for covariant systems, Trans. Amer. Math. Soc. 248 (1979), 223 -- 267. 25. Chi-Keung Ng, Coactions on operator spaces. Preprint available at: https://www.researchgate.net/publication/288604716_Coactions_on_Operator_spaces_and_Exactness. 26. A.R. Medghalchi and H. Nikpey, Characterizing injective operator space V for which I11(V ) ∼= B(H), Publ. Math. Debrecen 82 (2013), 21 -- 30 27. M. Nilsen, Duality for full crossed products of C ∗-algebras by non-amenable groups, Proc. Amer. Math. Soc. 126 (1998), 2969 -- 2978. 28. V. Paulsen, A covariant version of EXT, Michigan Math. J. 29 (1982), 131 -- 142. 29. J. Quigg, Full C ∗-crossed product duality, J. Austral. Math. Soc. (Ser. A) 50 (1991), 34 -- 52. 30. I. Raeburn, On crossed products and Takai duality, Proc. Edinburgh Math. Soc. 31 (1988), 321 -- 330. 31. I. Raeburn, On crossed products by coactions and their representation theory, Proc. London Math. Soc. (3) 64 (1992), no. 3, 625652. 32. I. Raeburn, A. M. Sinclair, and D. P. Williams, Equivariant completely bounded oper- ators, Pacific J. Math. 139 (1989), 155 -- 194. 33. Z.-J. Ruan, Subspaces of C ∗-algebras, J. Funct. Anal. 76 (1988), 217 -- 230. 34. Z.-J. Ruan, Injectivity of operator spaces, Trans. Amer. Math. Soc. 315 (1989), 89 -- 104. 35. H. Takai, On a duality for crossed products of C ∗-algebras, J. Funct. Anal. 19 (1975), 25 -- 39. 36. D. P. Williams, Crossed products of C ∗-algebras, Mathematical Surveys and Mono- graphs 134, American Mathematical Society, Providence, 2007. 37. G. Wittstock, Extension of completely bounded C ∗-module homomorphisms, Proc. Conf. Operator Algebras and Group Representations, Pitman, New York, 1983. Department of Mathematics, Faculty of Mathematical Sciences, Tarbiat Modares University, Tehran 14115-134, Iran E-mail address: [email protected] Mathematisches Institut, Westfalische Wilhelms-Universitat Munster, Ein- steinstr. 62, 48149 Munster, Germany E-mail address: [email protected] Department of Basic Sciences, Shahid Rajaee Teacher Training University, P. O. Box 16783-163, Tehran, Iran E-mail address: [email protected]
1501.05476
2
1501
2015-01-28T01:18:23
Groupoid Fell bundles for product systems over quasi-lattice ordered groups
[ "math.OA" ]
Consider a product system over the positive cone of a quasi-lattice ordered group. We construct a Fell bundle over an associated groupoid so that the cross-sectional algebra of the bundle is isomorphic to the Nica-Toeplitz algebra of the product system. Under the additional hypothesis that the left actions in the product system are implemented by injective homomorphisms, we show that the cross-sectional algebra of the restriction of the bundle to a natural boundary subgroupoid coincides with the Cuntz-Nica-Pimsner algebra of the product system. We apply these results to improve on existing sufficient conditions for nuclearity of the Nica-Toeplitz algebra and the Cuntz-Nica-Pimsner algebra, and for the Cuntz-Nica-Pimsner algebra to coincide with its co-universal quotient.
math.OA
math
GROUPOID FELL BUNDLES FOR PRODUCT SYSTEMS OVER QUASI-LATTICE ORDERED GROUPS ADAM RENNIE, DAVID ROBERTSON, AND AIDAN SIMS Abstract. Consider a product system over the positive cone of a quasi-lattice ordered group. We construct a Fell bundle over an associated groupoid so that the cross-sectional algebra of the bundle is isomorphic to the Nica -- Toeplitz algebra of the product system. Under the additional hypothesis that the left actions in the product system are implemented by injective homomorphisms, we show that the cross-sectional algebra of the restriction of the bundle to a natural boundary subgroupoid coincides with the Cuntz -- Nica -- Pimsner algebra of the product system. We apply these results to improve on existing sufficient conditions for nuclearity of the Nica -- Toeplitz algebra and the Cuntz -- Nica -- Pimsner algebra, and for the Cuntz -- Nica -- Pimsner algebra to coincide with its co-universal quotient. 1. Introduction In [20], Pimsner associated to each C ∗-correspondence over a C ∗-algebra A two C ∗-algebras TX and OX . His construction simultaneously generalised the Cuntz -- Krieger algebras and their Toeplitz extensions, graph C ∗-algebras and crossed products by Z, and has been intensively studied ever since. It is standard these days to present TX as the universal C ∗-algebra generated by a represen- tation of the module X, and then OX as the quotient of TX determined by a natural covariance condition. However, this was not Pimsner's original definition. In [20], OX is by definition the quotient of the image of the canonical representation of X as creation operators on its Fock space by the ideal of compact operators on the Fock space. Pimsner then provided two alterna- tive presentations of OX , the second of which is the one in terms of its universal property. The first, which is the one germane to this paper, is an analogue of the realisation of C(T) by dila- tion of the canonical representation of the classical Toeplitz algebra on ℓ2. Pimsner constructed a direct-limit module X∞ over the direct limit A∞ of the algebras of compact operators on the tensor powers of X. He showed that one can make sense of X ⊗n ∞ for all integers n, and so ∞ . This space carries a natural representation of X∞ by form a 2-sided Fock space Ln∈Z X ⊗n translation operators, and the image generates OX∞ which is isomorphic to OX . More recently [12], Fowler introduced compactly aligned product systems of Hilbert A -- A bimodules over the positive cones in quasi-lattice ordered groups (G, P ), and studied associated C ∗-algebras TX and OX , and an interpolating quotient N T X (Fowler denoted it by Tcov(X), but we follow the notation of [3]). When (G, P ) = (Z, N), TX = N T X agrees with Pimsner's Toeplitz algebra, and OX with Pimsner's Cuntz-Pimsner algebra. But even for (Z2, N2) the situation is more complicated. The algebra N T X is essentially universal for the relations encoded by the natural Fock representation of X, so it is a natural analogue of Pimsner's Toeplitz algebra. But the quotient by the ideal of compact operators on the Fock space is much too large to behave like an analogue of Pimsner's OX. (This is analogous to the fact that C ∗(Z) is the quotient of C ∗(N) by the compact operators on ℓ2(N), but C ∗(Z2) is much smaller than the quotient of C ∗(N2) by the compact operators on ℓ2(N2).) Fowler also lacked an analogue of X∞; the direct limit should be taken over P , but P is typically not directed. So Fowler's approach to defining OX was to mimic Pimsner's second alternative presentation of OX : identify a natural covariance relation and define OX as the universal quotient of TX determined by this relation. Subsequent papers [26, 4] have modified Fowler's definition to accommodate various levels of 2010 Mathematics Subject Classification. 46L05. Key words and phrases. Groupoid; C ∗-correspondence; Product system; Fell bundle. 1 2 ADAM RENNIE, DAVID ROBERTSON, AND AIDAN SIMS additional generality, but have taken the same fundamental approach of defining N OX as the universal C ∗-algebra determined by a representation of TX satisfying some additional essentially ad hoc relations. Nevertheless, there is strong evidence [12, 4] that the resulting C ∗-algebra N OX can profitably be regarded as a generalised crossed product of the coefficient algebra A by the group G. In particular, in the case that (G, P ) = (Zk, Nk) and X is the product system arising from an action α of Nk on A by endomorphisms, a new characterisation and analysis of N OX , closely related to Pimsner's dilation approach, is achieved in [6] using the powerful machinery of Arveson envelopes of non-self-adjoint operator algebras. The authors answer in the affirmative a question raised in [4] about whether N OX can be recovered using Arveson's approach, and use this to show, amongst other things, that N OX is Morita equivalent (in fact isomorphic in the case that the αp are all injective) to a genuine crossed-product by Zk. In this paper we provide an analogue of Pimsner's first representation of OX that is applicable to compactly aligned product systems over quasi-lattice ordered groups, under the additional hypothesis that the left A-actions are implemented by nondegenerate injective homomorphisms φp : A → L(Xp). Our approach is to use a natural groupoid G associated to (G, P ) [17], and construct a Fell bundle over G whose cross-sectional C ∗-algebra coincides with N T X . The groupoid G has a natural boundary, which is a closed subgroupoid (see [5]), and the restriction of our Fell bundle to this boundary subgroupoid has cross-sectional algebra isomorphic to the algebra N OX of [26]. This is strong evidence that the relations recorded in [26] are the right ones, at least for nondegenerate product systems with injective left actions. As practical upshots of our results, we deduce that if the groupoid G is amenable, then: (1) each of N T X and N OX is nuclear whenever the coefficient algebra A is nuclear, and (2) N OX coincides with its co-universal quotient N Or X as in [4]. This improves on previous results along these lines, which assume that the group G is amenable, a stronger hypothesis than amenability of G. We mention that the work of Kwasniewski and Szyma´nski in [15], is related to our con- struction. There the authors consider product systems over semigroups P that satisfy the Ore condition but are not necessarily part of a quasi-lattice ordered pair, and assume that the left actions in the product system are by compact operators. Here, by contrast, we insist that P is quasi-lattice ordered, but do not require compact actions. Both approaches use the machinery of Fell bundles, but Kwasniewski and Szyma´nski construct Fell bundles over the enveloping group G of P , whereas we construct a bundle over the associated groupoid G; as mentioned above, an advantage of the latter is that G can be amenable even when G is not. 2. Preliminaries 2.1. Product systems over quasi-lattice ordered groups. Let G be a discrete group and let P be a subsemigroup of G satisfying P ∩ P −1 = {e}. Define a partial order ≤ on G by g ≤ h ⇐⇒ g−1h ∈ P. We call the pair (G, P ) a quasi-lattice ordered group if, whenever two elements g, h ∈ G have a common upper bound in G, they have a least common upper bound g ∨ h in G. We write g ∨ h < ∞ if two elements g, h ∈ G have a common upper bound and g ∨ h = ∞ otherwise. A product system over a quasi-lattice ordered group (G, P ) is a semigroup X equipped with a semigroup homomorphism d : X → P such that the following hold. For each p ∈ P , let Xp = d−1(p). Then we require that A = Xe is a C ∗-algebra, thought of as a right-Hilbert module over itself in the usual way, and that each Xp is a right-Hilbert A-module together with a left action of A by adjointable operators denoted ϕp : A → L(Xp). We require that ϕe is given by left multiplication. Furthermore, for each p, q ∈ P with p 6= e, we require that multiplication in X determines a Hilbert bimodule isomorphism Xp ⊗A Xq → Xpq satisfying xp ⊗ xq 7→ xpxq. The product system is nondegenerate if multiplication Xe × Xp → Xp also determines an isomorphism Xe ⊗a Xp → Xp for each p; that is, if each Xp is nondegenerate as a left A-module. Every right Hilbert module is automatically nondegenerate as a right A-module by the Hewitt-Cohen factorisation theorem. PRODUCT SYSTEMS AND FELL BUNDLES 3 If p, q ∈ P satisfy e 6= p ≤ q, then there is a homomorphism ip−1q : L(Xp) → L(Xq) characterised by ip−1q(S)(xy) = (Sx)y for all x ∈ Xp, y ∈ Xp−1q. If we identify A with K(Xe) in the usual way then the corresponding map ip : K(Xe) → L(Xp) is ip = ϕp. We say that a product system X is compactly aligned if, whenever S ∈ K(Xp), T ∈ K(Xq) and p ∨ q < ∞ we have If g ∈ G \ P we define ig to be 0. ip−1(p∨q)(S)iq−1(p∨q)(T ) ∈ K(Xp∨q). Example 2.1. The pair (Z, N) is a quasi-lattice ordered group, where ≤ agrees with the usual ordering on Z. Let A be a C ∗-algebra and let E be an A-correspondence; i.e. E is a right Hilbert A-module with a left action A → L(E). Let X0 := A and for each n ∈ N \ {0} let Xn := E⊗n. Then is a product system over (Z, N). With multiplication given by ξη := ξ ⊗ η. X :=Sn∈N Xn Example 2.2. For each k ≥ 1, the pair (Zk, Nk) is a quasi lattice ordered group where, for m, n ∈ Zk and 1 ≤ i ≤ k (m ∨ n)i = max{mi, ni}. Suppose that (Λ, d) is a k-graph. For each n ∈ Nk, Cc(d−1(n)) is a pre-Hilbert A = C0(Λ0) module. Let Xn = Cc(d−1(n)). Then is a product system over (Zk, Nk). (See [23].) X =Sn∈Nk Xn 2.2. Representations of product systems. For details of the following, see [4, 12, 26]. Definition 2.3. Let X be a compactly aligned product system over a quasi-lattice ordered group (G, P ). A Toeplitz representation of X in a C ∗-algebra B is a map ψ : X → B satisfying (T1) ψp := ψXp : Xp → B is linear for all p ∈ P and ψe is a homomorphism, (T2) ψ(xy) = ψ(x)ψ(y) for all x, y ∈ X, and (T3) for any p ∈ P and x, y ∈ Xp, ψ(hx, yi) = ψ(x)∗ψ(y). Given a Toeplitz respresentation ψ : X → B, for each p ∈ P there is a homomorphism ψ(p) : K(Xp) → B satisfying We call a Toeplitz representation ψ : X → B Nica covariant if (N) for all S ∈ K(Xp), T ∈ K(Xq) we have ψ(p)(θx,y) = ψp(x)ψp(y)∗. ψ(p)(S)ψ(q)(T ) =(cid:26) ψp∨q(cid:0)ip−1(p∨q)(S)iq−1(p∨q)(T )(cid:1) 0 if p ∨ q < ∞ otherwise. Following [3], we will write N T X for the universal C ∗-algebra generated by a Nica-covariant (Fowler shows that such a C ∗-algebra exists in [12], but Toeplitz representation iX of X. denotes it Tcov(X).) Given a predicate P on P , we say P is true for large s if for every q ∈ P , there exists an r ≥ q such that P(s) is true whenever s ≥ r. We now present the definition of the Cuntz -- Nica -- Pimsner algebra N OX of a product system X under the assumption that the left action on each fibre is implemented by an injective homomorphism ϕp. This hypothesis is not needed for N OX to make sense (see [26]); but if the left actions are not implemented by injective homomorphisms, then the relation (CNP) as described below does not hold in N OX . In particular, this hypothesis will be necessary in all statements that involve Cuntz-Nica-Pimsner covariance and representations of N OX: Proposition 4.2, Theorem 5.2, and the results in Section 6 4 ADAM RENNIE, DAVID ROBERTSON, AND AIDAN SIMS Definition 2.4. Let X be a compactly aligned product system over a quasi-lattice ordered group (G, P ) and suppose that for each p ∈ P the left action φp : A → L(Xp) is injective. We say a Nica covariant Toeplitz representation ψ : X → B is Cuntz -- Nica -- Pimsner covariant if it satisfies the following property: (CNP) for each finite F ⊂ P and collection of elements Tp ∈ K(Xp), p ∈ F , if Pp∈F ip−1q(Tp) = 0 for large q, then We write N OX for the universal C ∗-algebra generated by a Cuntz -- Nica -- Pimsner covariant representation jX of X. Pp∈F ψ(p)(Tp) = 0. 2.3. Fell bundles over groupoids. We say that a groupoid G is a topological groupoid if G is a topological space and the multiplication and inversion are continuous functions. We call a topological groupoid G ´etale if the unit space G(0) is locally compact and Hausdorff, and the range map r : G → G(0) is a local homeomorphism. It follows that the source map s is also a local homeomorphism. A bisection of G is an open subset U ⊆ G such that rU and sU are homeomorphisms; the topology of a Hausdorff ´etale groupoid admits a basis consisting of bisections. See [9] for an overview of ´etale groupoids. Given a Hausdorff ´etale groupoid G, a Fell bundle over G is an upper-semicontinuous Banach bundle p : E → G with a multiplication E (2) = {(e, f ) ∈ E × E : (p(e), p(f )) ∈ G(2)} → E and an involution satisfying the following properties: ∗ : E → E , e 7→ e∗ (1) the multiplication is associative and bilinear, whenever it makes sense; (2) p(ef ) = p(e)p(f ) for all (e, f ) ∈ E (2); (3) multiplication is continuous in the relative topology on E (2) ⊆ E × E ; (4) kef k ≤ kekkf k for all (e, f ) ∈ E (2); (5) p(e∗) = p(e)−1 for all e ∈ E , and involution is continuous and conjugate linear; (6) (e∗)∗ = e, ke∗k = kek and (ef )∗ = f ∗e∗ for all (e, f ) ∈ E (2); (7) ke∗ek = kek2 for all e ∈ E ; (8) e∗e ≥ 0 as an element of p−1(s(p(e))) -- which is a C ∗-algebra by (1) -- (7) -- for all e ∈ E . We denote by Eγ the fibre p−1(γ) ⊂ E . Given a Fell bundle E over a locally compact Hausdorff ´etale groupoid, we write Γc(G; E ) for the vector space of continuous, compactly supported sections ξ : G → E . If H ⊆ G is a closed subset, we will write Γc(H; E ) for the compactly supported sections of the restriction of E to H; that is, Γc(H; E ) := Γc(H; E H). There are a convolution and involution on Γc(G; E ) such that for ξ, η ∈ Γc(G; E ), (ξ ∗ η)(γ) = Xαβ=γ ξ(α)η(β) and ξ∗(γ) = ξ(γ−1)∗. This gives Γc(G; E ) the structure of a ∗-algebra. The I-norm on Γc(G; E ) is given by kf kI := sup u∈G(0)(cid:16) max(cid:16) Xs(γ)=u kf (γ)k, Xr(γ)=u kf (γ)k(cid:17)(cid:17). A ∗-homomorphism L : Γc(G; E ) → B(HL) is called a bounded representation if kL(f )k ≤ kf kI for all f ∈ Γc(G; E ). It is nondegenerate if span{L(f )ξ : f ∈ Γc(G; E ), ξ ∈ HL} = HL is dense. The universal C ∗-norm on Γc(G; E ) is kf k := sup{kL(f )k : L is an bounded representation}. We define the cross-sectional algebra C ∗(G, E ) to be the completion of Γc(G; E ) with respect to the universal C ∗-norm. If H ⊆ G is a closed subgroupoid, then we write C ∗(H, E ) for the completion of Γc(H, E ) in the universal norm on Γc(H, E ). PRODUCT SYSTEMS AND FELL BUNDLES 5 3. From a product system to a Fell bundle In this section, given a product system X over a quasi-lattice ordered group (G, P ), we construct a groupoid G and a Fell bundle E over G. We will show in Section 5 that the C ∗- algebra of this Fell bundle coincides with the Nica -- Toeplitz algebra of X, and has a natural quotient that coincides with the Cuntz -- Nica -- Pimsner algebra. Standing notation: We fix, for the duration of Section 3, a quasi-lattice ordered group (G, P ), and a nondegenerate compactly aligned product system X over P . For the time being, we do not require that the left actions on the fibres of X are implemented by injective homomor- phisms; as mentioned before, this additional hypothesis will be needed only in Proposition 4.2, Theorem 5.2, and the results of Section 6. 3.1. The groupoid. We first construct a groupoid from (G, P ). This construction is by no means new -- for example, it appears in the work of Muhly and Renault [17] in the context of Weiner-Hopf algebras. Fix a quasi-lattice ordered group (G, P ). We say that ω ⊂ G is directed if and hereditary if g, h ∈ ω =⇒ ∞ 6= g ∨ h ∈ ω h ∈ ω and g ≤ h =⇒ g ∈ ω. Let Ω = {ω ⊂ G : ω is directed and hereditary}. With the relative product topology induced by identifying Ω with a subset of {0, 1}G in the usual way, Ω is a totally disconnected compact Hausdorff space: the sets Z(A0, A1) := {ω ∈ Ω : g ∈ Ai =⇒ χω(g) = i}, indexed by pairs A0, A1 of finite subsets of G constitute a basis of compact open sets. We say that ω ∈ Ω is maximal if ω ⊂ ρ ∈ Ω implies ω = ρ. Let Ωmax = {ω ∈ Ω : ω is maximal}. Define the boundary of Ω to be Given g ∈ G and ω ∈ Ω, let ∂Ω := Ωmax ⊂ Ω. gω := {gh : h ∈ ω}. For finite A0, A1 ⊆ G and g ∈ G, we have g−1Z(A0, A1) = Z(g−1A0, g−1A1). Hence g · ω := gω defines an action of G by homeomorphisms of Ω. Given p ∈ P , the set ωp := {g ∈ G : g ≤ p} belongs to Ω, so we can regard P as a subset of Ω. Proposition 3.1. The boundary ∂Ω is invariant under the action of G. Proof. By continuity of the G-action, it suffices to show that Ωmax is invariant. Fix ω ∈ Ωmax and g ∈ G and suppose that gω ⊂ ρ for some ρ ∈ Ω. Then ω ⊂ g−1ρ and hence ω = g−1ρ, since ω is maximal. So gω = gg−1ρ = ρ. (cid:3) The set becomes a groupoid when endowed with the operations G = {(g, ω) : P ∩ ω 6= ∅ and P ∩ gω 6= ∅} (g, hω)(h, ω) = (gh, ω) and (g, ω)−1 = (g−1, gω). The unit space is {e} × Ω, which we identify with Ω, and the structure maps are r(g, ω) = (e, gω) and s(g, ω) = (e, ω). One can check that G is equal to the restriction of the transformation groupoid G ⋉ Ω to the closure of the copy of P in Ω; in symbols, G = (G ⋉ Ω)P . We write G∂Ω for the subgroupoid G∂Ω := {(g, ω) ∈ G : ω ∈ ∂Ω}. 6 ADAM RENNIE, DAVID ROBERTSON, AND AIDAN SIMS 3.2. The fibres of the Fell bundle. For a fixed r ∈ P and any p, q ∈ P there is a map ir : L(Xp, Xq) → L(Xpr, Xqr) such that, for x ∈ Xp and y ∈ Xr ir(S)(xy) = S(x)y. There is no notational dependence on p and q, but this will not cause confusion -- indeed, it is helpful to think of ir as a map from Lp,q∈P L(Xp, Xq) to Lp,q∈P L(Xpr, Xqr). For ω ∈ Ω and p ∈ ω, we define [p, ω) := {q ∈ ω : p ≤ q}. Given any (g, ω) ∈ G, we have [e ∨ g−1, ω) = {p ∈ P ∩ ω : gp ∈ P }, and this set is directed (under the usual ordering on P ). So we can form the Banach-space direct limit lim−→p∈[e∨g−1,ω) L(Xp, Xgp) with respect to the maps ir : L(Xp, Xgp) → L(Xpr, Xgpr) where pr, gpr ∈ ω. By definition of the direct limit, there are bounded linear maps L(Xp, Xgp) → lim−→ L(Xp, Xgp), p ∈ [e ∨ g−1, ω), that are compatible with the linking maps ir. To lighten notation we regard all of these maps as components of a single map i(g,ω) :Lp L(Xp, Xgp) → lim−→ L(Xp, Xgp). We define E(g,ω) := spanSp∈[e∨g−1,ω) i(g,ω)(K(Xp, Xgp)) ⊂ lim−→ L(Xp, Xgp). Lemma 3.2. Each Aω := E(e,ω) is a C ∗-algebra and each E(g,ω) is an Agω -- Aω imprimitivity bimodule. Proof. By definition of the maps ir, if T ∈ L(Xp, Xp′) and S ∈ L(Xp′, Xp′′), then ir(T )ir(S) = ir(T S), and ir(T )∗ = ir(T ∗). Using this, one checks that, identifying each L(Xp ⊕ Xgp) with the algebra of block-operator matrices (cid:16) L(Xp) morphism ir : L(Xp ⊕ Xgp) → L(Xpr ⊕ Xgpr). In the same vein as above, we use the notation ıg,ω for all of the homomorphisms L(Xp ⊕ Xgp) → lim−→ L(Xp, Xgp). L(Xp,Xgp) L(Xgp) (cid:17), the maps ir determine a homo- The following is adapted from the proof of [16, Lemma 4.1]. Since ω is directed, each finite subset H ⊆ [e ∨ g−1, ω) is contained in a finite F ⊆ [e ∨ g−1, ω) which is closed under ∨, and each such F has a maximum element pF . For each such F , let L(Xgp,Xp) BF :=Xs∈F is−1pF (K(Xs ⊕ K(Xgs)) ⊆ L(XpF ⊕ XgpF ). If F ⊆ ω is finite with more than one element and ∨-closed, and if q ∈ F is minimal, then F ′ := F \ {q} is also ∨ closed, and pF ′ = pF . We have BF = iq−1pF (K(Xq ⊕ Xgq)) + BF ′. Nica covariance and minimality of q ensures that iq−1pF (K(Xq ⊕ Xgq))is−1pF (K(Xs ⊕ Xgs)) ⊆ i(q∨s)−1pF (K(X(q ∨ s) ⊕ Xg(q∨s))) ⊆ BF ′ So BF ′BF , BF BF ′ ⊆ BF ′. Assuming as an inductive hypothesis that BF ′ is a C ∗-algebra, we deduce from [7, Corollary 1.8.4] that BF is a C ∗-algebra. Since each B{p} = K(Xp ⊕ Xgp) is clearly a C ∗-algebra, we conclude by induction that each BF is a C ∗-algebra. So spanSp∈[e∨g−1,ω) ıg,ω(K(Xp ⊕ Xgp)) ⊂ lim−→ L(Xp ⊕ Xgp) ıg,ω(BF ), so is a C ∗-algebra. Put p = is canonically isometrically isomorphic to Lg,ω := lim−→F e ∨ g−1, so p ∈ ω ∩ P and gp ∈ gω ∩ P . Since X is nondegenerate, the spaces Aω and Agω appear as the complementary full corners ıg,ω(1Xp)Lg,ωıg,ω(1Xp) and ıg,ω(1Xgp)Lg,ωıg,ω(1Xgp) of Lg,ω, so they are C ∗-algebras. Furthermore, E(g,ω) = ıg,ω(1Xgp)Lg,ωıg,ω(1Xp), and so it is an Agω -- Aω-imprimitivity bimodule. (cid:3) PRODUCT SYSTEMS AND FELL BUNDLES 7 3.3. The operations on the Fell bundle. Let Then E is a bundle over G, with π : E → G defined by π(E(g,ω)) = {(g, ω)}. E :=S(g,ω)∈G E(g,ω). Lemma 3.3. Fix p, p′, q, q′ ∈ P with p ∨ q′ < ∞ and let r = p−1(p ∨ q′), and r′ = q′−1(p ∨ q′). Then for any S ∈ K(Xp, Xp′) and T ∈ K(Xq, Xq′) we have ir(S)ir′(T ) ∈ K(Xqr′, Xp′r). Proof. Since both the left and right actions are nondegenerate, it is enough to prove the result for SU and V T where S ∈ K(Xp,p′), U ∈ K(Xp) and T ∈ K(Xq, Xq′), V ∈ K(Xq′). We have ir(SU)ir′(V T ) = ir(S)ir(U)ir′(V )ir′(T ). Since X is compactly aligned, we have ir(U)ir′(V ) ∈ K(Xp∨q′), and hence ir(SU)ir′(V T ) ∈ K(Xqr′, Xp′r) as claimed. (cid:3) Fix ((g, hω), (h, ω)) ∈ G(2), hp ∈ [e ∨ g−1, hω), q ∈ [e ∨ h−1, ω) and S ∈ K(Xhp, Xghp), T ∈ K(Xq, Xhq). Let r = p−1(p ∨ q), r′ = q−1(p ∨ q), and define i(g,hω)(S)i(h,ω)(T ) := i(gh,ω) (ir(S)ir′(T )) . The right hand side makes sense by Lemma 3.3. This extends to a multiplication E (2) := {(e, f ) ∈ E × E : (π(e), π(f )) ∈ G(2)} → E . For (g, ω) ∈ G and p ∈ [e ∨ g−1, ω), the usual adjoint operation ∗ : L(Xp, Xgp) → L(Xgp, Xp) = L(Xgp, Xg−1(gp)) is isometric. So for each (g, ω) it extends to an involution lim−→ L(Xp, Xgp) → lim−→ L(Xgp, Xp), which then restricts to an involution E(g,ω) → E(g−1,gω). 3.4. The topology on the Fell bundle. Given p, q ∈ P and S ∈ L(Xp, Xq) define f S : G → S(g,ω)∈G lim−→p∈[e∨g−1,ω) L(Xp, Xgp) by f S(g, ω) =(cid:26) i(qp−1,ω)(S) 0 if g = qp−1 and p ∈ ω otherwise. Lemma 3.4. For any p, q ∈ P and any S ∈ L(Xp, Xq), the map (g, ω) 7→ kf S(g, ω)k is upper semicontinuous. Proof. Since kf S(g, ω)k = kf S ∗S(ω)k1/2 for any (g, ω) ∈ G, it is enough to check upper semi- continuity on the unit space G(0) = Ω. Fix p ∈ P , S ∈ L(Xp) and α > 0. We must show that the set {ω : kf S(ω)k < α} is open. Since p 6∈ ω implies that f S(ω) = 0, we see that {ω : kf S(ω)k < α} = Z({p}, ∅) ∪ {ω : p ∈ ω and kiω(S)k < α} and so it is enough to show that {ω : p ∈ ω and kf S(ω)k < α} is open. Fix ω in this set. Since Aω is a direct limit we have kf S(ω)k = kiω(S)k = lim q≥p kiqp−1(S)k = inf q≥p kiqp−1(S)k. Therefore, there exists a q ≥ p such that kiqp−1(S)k < α. Suppose that ω′ ∈ Z(∅, {q}). Then p ∈ ω′, and so kf S(ω′)k = kiω′(S)k ≤ kiqp−1(S)k < α. (cid:3) 8 ADAM RENNIE, DAVID ROBERTSON, AND AIDAN SIMS Now let Γ = span{f S : p, q ∈ P, S ∈ K(Xp, Xq)}. Given finitely many pairs (p1, q1), . . . , (pn, qn) and operators Si ∈ K(Xpi, Xqi), there are finitely many maximal subsets F1, . . . , Fm of {p1, . . . , pn} such that each Fj has an upper bound rj in ip−1rj (Si) for each j, we have Tj ∈ L(Xrj ) and P . Putting Tj :=Pp∈Fj i=1 f Si =Pm Pn j=1 f Tj , where the f Tj have mutually disjoint support. So Lemma 3.4 shows that the sections in Γ are upper semicontinuous. Given (g, ω) ∈ G we have {f (g, ω) : f ∈ Γ} =(cid:8)i(g,ω)(S) : p ∈ [e ∨ g−1, ω), S ∈ K(Xp, Xgp)(cid:9) =S[e∨g−1,ω) i(g,ω)(K(Xp, Xgp)) which densely spans E(g,ω). Hence [11, Section II.13.18] shows that there is a unique topology on E such that (E , π) is a Banach bundle and all the functions in Γ are continuous cross sections of E ; and E becomes a Fell-bundle over G in this topology. 4. Representing the product system 4.1. Toeplitz representation. Let (G, P ) be a quasi-lattice ordered group, and X a nonde- generate compactly aligned product system over P . For p ∈ P , identify Xp with K(Xe, Xp) as usual: x ∈ Xp is identified with the operator a 7→ x · a. We then write x∗ for the operator y 7→ hx, yiXe in K(Xp, Xe). Define ψp : Xp → C ∗(G, E ) by ψp(x) = f x. Proposition 4.1. Let (G, P ) be a quasi-lattice ordered group, and X a nondegenerate compactly aligned product system over P . Let G and E be the groupoid and Fell bundle constructed in Section 3. The map ψ : X → C ∗(G, E ) such that ψXp = ψp is a Nica covariant Toeplitz representation of X, and for S ∈ K(Xp), we have ψ(p)(S) = f S. Proof. We need to check the conditions of Definition 2.3. For x, y ∈ Xp and a ∈ Xe, ψp(x)∗ψp(y)(g, ω) = [(f x∗) ∗ f y](g, ω) = Xhω∩P 6=∅ f x((gh−1, hω)−1)∗f y(h, ω) f x(hg−1, gω)∗f y(h, ω) = δg,ef x(p, ω)∗f y(p, ω) = Xhω∩P 6=∅ = δg,ei(p,ω)(x)∗i(p,ω)(y) = δg,ei(p−1,pω)(x∗)i(p,ω)(y) = δg,eiω(hx, yiA) = f hx,yiA(g, ω) = ψe(hx, yi). Likewise, and [ψe(a)ψp(x)](g, ω) = [f a ∗ f x](g, ω) = Xhω∩P 6=∅ f a(gh−1, hω)f x(h, ω) = δg,pipω(a)i(p,ω)(x) = δg,pi(p,ω)(ax) = f ax(g, ω) = ψp(ax) [ψp(x)ψe(a)](g, ω) = [f x ∗ f a](g, ω) = Xhω∩P 6=∅ f x(gh−1, hω)f a(h, ω) = δg,pi(p,ω)(x)iω(a) = δg,pi(p,ω)(xa) = f xa(g, ω) = ψp(xa). To see that each ψ(p)(S) = f S, consider S = θx,y and calculate: PRODUCT SYSTEMS AND FELL BUNDLES 9 ψ(p)(θx,y)(g, ω) = [ψp(x)ψp(y)∗](g, ω) = [f x ∗ f y](g, ω) = Xhω∩P 6=∅ f x(gh−1, hω)f y((h, ω)−1)∗ f x(gh−1, hω)f y(h−1, hω)∗ = δg,pi(p,p−1ω)(x)i(p,p−1ω)(y)∗ = Xhω∩P 6=∅ = δg,pi(p,p−1ω)(x)i(p−1,ω)(y∗) = δg,piω(θx,y) = f θx,y(g, ω). So continuity and linearity give ψ(p)(S) = f S for all S ∈ K(Xp). Fix p, q ∈ P with p ∨ q < ∞ and S ∈ K(Xp), T ∈ K(Xq). Then [ψ(p)(S)ψ(q)(T )](g, ω) = [f S ∗ f T ](g, ω) = Xhω∩P 6=∅ f S(gh−1, hω)f T (h, ω) = δg,eiω(S)iω(T ) = δg,eiω(ip−1(p∨q)(S)iq−1(p∨q)(T )) = f ip−1(p∨q)(S)iq−1(p∨q)(T )(g, ω) = [ψ(p∨q)(ip−1(p∨q)(S)iq−1(p∨q)(T ))](g, ω). Thus all the conditions of Definition 2.3 are satisfied. (cid:3) 4.2. Restriction of the representation to the boundary groupoid. Consider πp : Xp → C ∗(G∂Ω, E ) satisfying πp(x) = f xG∂Ω Define π : X → C ∗(G∂Ω, E ) by πXp = πp. Proposition 4.2. Let (G, P ) be a quasi-lattice ordered group, and X a nondegenerate compactly aligned product system over P . Suppose that the homomorphisms φp : A → L(Xp) implementing the left actions are all injective. Let G and E be the groupoid and Fell bundle constructed in Section 3. The map π : X → C ∗(G∂Ω, E ) is a Cuntz -- Nica -- Pimsner covariant Toeplitz representation. Before we prove this, we need two lemmas. Lemma 4.3. Suppose that ω ∈ ∂Ω and q ∈ P satisfy q ∨ p < ∞ for all p ∈ ω. Then q ∈ ω. Proof. Consider the set If q ∨ p1, q ∨ p2 ∈ q ∨ ω we have q ∨ ω := {q ∨ p : p ∈ ω} (q ∨ p1) ∨ (q ∨ p2) = q ∨ (p1 ∨ p2) ∈ q ∨ ω since p1 ∨p2 ∈ ω. So q ∨ω is directed. Let Her(q ∨ω) denote the hereditary closure Her(q ∨ω) = {g ∈ G : g ≤ p for some p ∈ q ∨ ω} of q ∨ ω. Notice that q = q ∨ e ∈ Her(q ∨ ω). For any p ∈ ω, p ≤ q ∨ p ∈ q ∨ ω and hence p ∈ Her(q ∨ ω). So ω ⊂ Her(q ∨ ω) and hence ω = Her(q ∨ ω) because ω ∈ ∂Ω. So q ∈ ω. (cid:3) Lemma 4.4. Fix a sequence (ωn)∞ Then p ∈ ω, and for T ∈ K(Xp), n=1 ⊂ Ω with p ∈ ωn for all n, and suppose that ωn → ω. iωn(T ) → iω(T ) in E as n → ∞. Proof. We know that the set Z(∅, {p}) is closed and ωn ∈ Z(∅, {p}) for all n. Hence ω ∈ Z(∅, {p}) and so p ∈ ω. Now, fix T ∈ K(Xp) and U ⊂ E open with iω(T ) ∈ U. By definition of the topology on E , the function f T is continuous, so (f T )−1(U) ⊂ G is open. Since ωn → ω and G has the relative product topology, (e, ωn) → (e, ω) in G. We have f T (e, ω) = iω(T ) ∈ U, and hence (e, ω) ∈ (f T )−1(U). Thus there exists N such that (e, ωn) ∈ (f T )−1(U) for all n > N, and so f T (e, ωn) = iωn(T ) ∈ U for all n > N, 10 ADAM RENNIE, DAVID ROBERTSON, AND AIDAN SIMS giving iωn(T ) → iω(T ). (cid:3) Proof of Proposition 4.2. Replacing an ω ∈ Ω with ω ∈ ∂Ω in the proof of Proposition 4.1 shows that π is a Nica covariant Toeplitz representation. Since all the left actions are by injective homomorphisms, the representation π is Cuntz-Nica-Pimsner covariant if it satisfies relation (CNP) of Definition 2.4. Fix a finite set F ⊂ P and elements Tp ∈ K(Xp), p ∈ F such that for large q. We must show that Pp∈F π(p)(Tp) = 0. So, since each π(p)(T ) = ψ(p)(T )∂Ω, we have to check that for all (g, ω) ∈ G∂Ω. Fix (g, ω) ∈ G∂Ω with ω ∈ Ωmax, and observe that Since F ∩ ω ⊂ P is finite and ω is directed, the element Pp∈F f Tp(g, ω) = δg,ePp∈F ∩ω iω(Tp). Pp∈F iqp−1(Tp) = 0 Pp∈F f Tp(g, ω) = 0 r :=Wp∈F ∩ω p belongs to ω, and Pp∈F ∩ω iω(Tp) = iω(cid:16)Pp∈F ∩ω ip−1r(Tp)(cid:17). Since ω is directed and countable we can choose a sequence (rn)∞ n=1 ⊂ ω satisfying • r1 ≥ r, • rn+1 ≥ rn for all n • for all q ∈ ω, there exists n with rn ≥ q. For each n, choose qn ≥ rn and ωn ∈ ∂Ω with qn ∈ ωn (and hence rn ∈ ωn) such that Then in particular, (4.1) Pp∈F ip−1qn(Tp) = 0. ip−1qn(Tp) =Pp∈F ip−1qn(Tp) = 0 Pp∈F ∩ωn since p ∈ F \ ωn implies p (cid:2) qn and so ip−1qn this fix Z(A0, A1) containing ω. Since A1 ⊂ ω, A1 is directed. Let p = 0. We claim that ωn → ω as n → ∞. To see p. s =Wp∈A1 By definition of (rn)∞ n=1 there is an n1 with rn1 ≥ s. Then A1 ⊂ ωrn for any n ≥ n1. For each q ∈ A0, let Nq := max{n : q ∈ ωn}. Suppose for contradiction that q ∈ A0 satisfies Nq = ∞. For any p ∈ ω we can find rj ≥ p. Since Nq = ∞ we can find k ≥ j with q ∈ ωk. But then q ∨ rk < ∞ =⇒ q ∨ rj < ∞ =⇒ q ∨ p < ∞. Since p ∈ ω was arbitrary we deduce that q ∨ p < ∞ for all p ∈ ω and hence q ∈ ω by Lemma 4.3. This contradicts ω ∈ Z(A0, A1). Therefore Nq is finite for every q ∈ A0. Now put Then ωn ∈ Z(A0, A1) for any n > N and ωn → ω as claimed. Since F is finite, there exists NF such that n ≥ NF implies F ∩ ωn = F ∩ ω. N := max(cid:8)n1, maxq∈A0 Nq(cid:9) < ∞. Hence, using Lemma 4.4 at the third equality and (4.1) at the last one, we have Xp∈F f Tp(g, ω) = δg,e Xp∈F ∩ω = δg,e lim n→∞ iω(Tp) = δg,eiω Xp∈F ∩ω iωn Xp∈F ∩ω n→∞ ip−1r(Tp)! = δg,e lim iωn Xp∈F ∩ωn ip−1qn(Tp)! = δg,e lim n→∞ ip−1r(Tp)! iωn Xp∈F ∩ω ip−1qn(Tp)! = 0. PRODUCT SYSTEMS AND FELL BUNDLES 11 Since Ωmax is dense in ∂Ω and Pp∈F π(p)(Tp) is a continuous section of E , we deduce that Pp∈F π(p)(Tp) = 0. 5. The isomorphisms (cid:3) In this section, we prove our main results: that the C ∗-algebra of the Fell bundle E con- structed in Section 3 is isomorphic to the Nica -- Toeplitz algebra N T X and, under the hypothesis that the left actions of A on the Xp are implemented by injective homomorphisms, that the C ∗-algebra of the restriction of E to the boundary groupoid G∂Ω is isomorphic to the Cuntz -- Nica -- Pimsner algebra N OX. Theorem 5.1. Let X be a compactly aligned product system over a quasi-lattice ordered group (G, P ). Let G and E be the groupoid and Fell bundle constructed in Section 3. Then the homo- morphism Ψ : N T X → C ∗(G, E ) induced by the Toeplitz representation ψ of Proposition 4.1 is an isomorphism. Proof. We begin by showing that Ψ is surjective. By definition of the topology on E , it suffices to show that f S ∈ Im Ψ for all S ∈ K(Xp, Xq). If S, T ∈ K(Xp, Xq) then f S + f T = f S+T , so it suffices to show that f θy,x ∈ Im Ψ for all x ∈ Xp and y ∈ Xq. Given (g, ω) ∈ Ω we have [ψq(y)ψp(x)∗](g, ω) = [f y ∗ f x∗](g, ω) = Xhω∩P 6=∅ f y(gh−1, hω)f x(h−1, hω)∗ = δg,qp−1f y(q, p−1ω)f x(p, p−1ω)∗ = δg,qp−1i(q,p−1ω)(x)i(p,p−1ω)(y)∗ = δg,qp−1i(q,p−1ω)(x)i(p−1,ω)(y∗) = δg,qp−1i(qp−1,ω)(xy∗) = f θx,y(g, ω) as required. To see that Ψ is injective, we construct an inverse. We begin by showing that there is a well-defined map Φ : span{f S : S ∈ K(Xp, Xq)} → N T X satisfying (5.1) Φ(f θy,x) = iX (y)iX(x)∗ To see that such a map exists, suppose that for all x ∈ Xp and y ∈ Xq. It suffices to show that j=1 f θyj ,xj = 0 ∈ Γc(G; E ). j=1 iX (yj)iX(xj)∗ = 0 ∈ N T X . Since the Fock representation l : X → L(F (X)) is isometric [12, page 340], this is equivalent to To see this, fix z ∈ Xr and a ∈ A. For any p ∈ P we have j=1 l(yj)l(xj)∗ = 0 ∈ L(F (X)). Pn Pn Pn Hence (cid:16)Pn n Xj=1 l(yj)l(xj)∗(z · a)! (p) = Xpj≤r qjp−1 j r=p n j=1 f θyj ,xj(cid:17) ∗ f θz,a = 0, and so 0 = (cid:16) Xj=1 = Xpj≤r f θyj ,xj(cid:17) ∗ f θz,a! (p, [e]) = Xqjp−1 j r(θyj ,xj )ie(θz,a) = Xpj≤r ip−1 qjp−1 j r=p qjp−1 j r=p yj(cid:16)ip−1 j r(xj)∗(z · a)(cid:17) . i(qj p−1 j ,[r])(θyj,xj )i(r,[e])(θ(z,a)) j r=p pj∈[r] yj(cid:16)ip−1 j r(xj)∗(z · a)(cid:17) . Hence (cid:16)Pn j=1 l(yj)l(xj)∗(z · a)(cid:17) (p) = 0. 12 ADAM RENNIE, DAVID ROBERTSON, AND AIDAN SIMS f =Pn Since z · a and p were arbitrary, we see that there is a well-defined linear map satisfying (5.1). We now show that Φ in continuous in the inductive limit topology. Suppose that fi → f in Γc(G; E ). Fix a compact subset K ⊂ G such that f and each of the fi vanishes off K. Write j=1 f Sj where each Sj ∈ K(Xpj , Xqj ). Inductively define A1 = supp(f S1) and Ak+1 = supp(f Sk+1) \(cid:16)Sk j=1 Ak(cid:17) for 1 ≤ k ≤ n. Then each Ak ⊂ G is a bisection, so that k(fi − f )AkkC ∗(G,E ) = k(fi − f )Akk∞ for all i. Define the set An+1 = K \(cid:16)Sn j=1 Ak(cid:17) . Without loss of generality, we may assume that An+1 is also a bisection. Then there exists N ≥ 1 such that for all i ≥ N and 1 ≤ k ≤ n So for i ≥ N k(fi − f )Akk∞ < ε n + 1 . n Φ(fi − f Sj )(cid:13)(cid:13)(cid:13) kΦ(fi) − Φ(f )k =(cid:13)(cid:13)(cid:13) Xj=1 Xk=1 Xj=1 n+1 ≤ n n Xj=1 =(cid:13)(cid:13)(cid:13) n+1 Φ((fi − f Sj )Ak)(cid:13)(cid:13)(cid:13) Xk=1 Xk=1 Xj=1 n+1 n kΦ((fi − f Sj )Ak)k ≤ k(fi − f Sj )Akk∞ < ε. So Φ(fi) → Φ(f ). Since the inductive limit topology on Γc(G; E ) is weaker than the norm topology, we see that Φ is bounded in norm. Since Γc(G; E ) is norm dense in C ∗(G, E ), Φ extends to a ∗-homomorphism Φ : C ∗(G, E ) → N T X which is, by construction, an inverse for Ψ. So C ∗(G, E ) ∼= N T X. (cid:3) Theorem 5.2. Let X be a nondegenerate compactly aligned product system over a quasi-lattice ordered group (G, P ). Suppose that the homomorphisms φp : A → L(Xp) implementing the left actions are all injective. Let G and E be the groupoid and Fell bundle constructed in Section 3. Then the homomorphism Π : N OX → C ∗(G∂Ω, E ), induced by the Cuntz -- Nica -- Pimsner covariant representation π of Proposition 4.2, is an isomorphism. Before we prove Theorem 5.2, we need to do some background work on coactions. The first lemma that we need is a general statement about coactions of discrete groups. The following brief summary of discrete coactions is based on [8, §A.3]. Given a discrete group G, the universal property of C ∗(G) shows that there is a homomorphism δG : C ∗(G) → C ∗(G) ⊗ C ∗(G) whose extension to MC ∗(G) satisfies δg(iG(g)) = iG(g) ⊗ iG(g). A coaction of a discrete group G on a C ∗-algebra A is a nondegenerate homomorphism δ : A → A ⊗ C ∗(G) which satisfies the coaction identity (δ ⊗ 1C ∗(G)) ◦ δ = (1 ⊗ δG) ◦ δ. The coaction δ is coaction-nondegenerate if span δ(A)(1M(A) ⊗ C ∗(G)) = A ⊗ C ∗(G). It is claimed at the beginning of Section 1 of [22] that, in our setting of discrete groups G, every coaction of a discrete group is coaction-nondegenerate. This assertion was used in results of [4] that we in turn will want to use in the proof of Theorem 5.2. However, this assertion in [22] depends on [21, Proposition 2.5], and a gap has recently been identified in the proof of this result [14]. The following simple lemma is well known, but hard to find in the literature. We will use it first to show that the coactions used in [4] are indeed coaction-nondegenerate (so the results of [4] are not affected by the issue identified in [14]), and then again in the proof of Lemma 5.5 below. Recall that if δ : A → A ⊗ C ∗(G) is a coaction of a discrete group, then for each g ∈ G, we write Ag for the spectral subspace {a ∈ A : δ(a) = a ⊗ iG(g)}. PRODUCT SYSTEMS AND FELL BUNDLES 13 Lemma 5.3. Let A be a C ∗-algebra and G a discrete group. Suppose that δ : A → A ⊗ C ∗(G) is a coaction. Then δ is coaction-nondegenerate if and only if A = spanSg∈G Ag. Proof. First suppose that δ is coaction-nondegenerate. Then [8, Proposition A.31] shows that A is densely spanned by its spectral subspaces. Now suppose that A is densely spanned by its spectral subspaces. Fix a typical spanning element a ⊗ iG(G) of A ⊗ C ∗(G). Fix ε and choose finitely many gi ∈ G and ai ∈ Agi such that ka −Pi aik < ε. Then (cid:13)(cid:13)(cid:13)Pi δ(ai)(1 ⊗ iG(g−1 i g)) − a ⊗ iG(g)(cid:13)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:13)(cid:16)Pi ai − a(cid:17) ⊗ iG(g)(cid:13)(cid:13)(cid:13) < ε. (cid:3) Corollary 5.4. The coactions of G on N T X and N OX used in [4] are coaction-nondegenerate. Proof. By construction (see [12]), the algebra N T X is the closure of the span of the elements iX (x)iX (y)∗ where x, y ∈ X. Hence N OX is densely spanned by the corresponding elements jX (x)jX (y)∗. The coactions of [4] are given by δ(iX(x)) = iX (x) ⊗ iG(g) and δ(jX(x)) = jX (x) ⊗ iG(g) whenever x ∈ Xg. So each spanning element of N T X and of N OX belongs to a spectral subspace for δ. Hence N T X and N OX are spanned by their spectral subspaces. Thus Lemma 5.3 shows that the coactions δ are coaction-nondegenerate. (cid:3) The second lemma that we need establishes that the C ∗-algebra of the Fell bundle of Section 3 carries a coaction of G that is compatible with the gauge coactions on N T X and N OX. Lemma 5.5. Let c be a continuous grading of a Hausdorff ´etale groupoid G by a discrete group G, and let E be a Fell bundle over G. Let iG : G → C ∗(G) denote the universal representation of G. There is a coaction-nondegenerate coaction δ of G on C ∗(E , G) satisfying whenever g ∈ G and f ∈ Γc(G; E ) satisfies supp(f ) ⊂ c−1({g}). δ(f ) = f ⊗ iG(g) Proof. As a vector space, Γc(G; E ) is equal to the algebraic direct sum Lg∈G Γc(c−1(g); E ). So there is a linear map δ : Γc(G; E ) → Γc(G; E ) ⊗ C ∗(G) such that δ(f ) = f ⊗ iG whenever f ∈ Γc(c−1(g); E ). It is routine to check that this map is continuous in the inductive-limit topology, and therefore extends to a homomorphism δ : C ∗(G, E ) → C ∗(G, E ) ⊗ C ∗(G). An elementary calculation checks the coaction identity on f ∈ Γc(c−1(g); E ), which suffices by linearity and continuity. To check that δ is coaction-nondegenerate, observe that the spectral subspaces C ∗(G, E )g are precisely the spaces Γc(c−1(g)); E ). By definition, C ∗(G, E ) is the closure of Γc(G; E ), which is spanned by the spaces Γc(c−1(g)); E ). It follows that C ∗(G, E ) is densely spanned by its spectral subspaces, and so δ is coaction-nondegenerate by Lemma 5.3. (cid:3) Recall that the Cuntz -- Nica -- Pimsner algebra N OX has a quotient N Or X that possesses a co-universal property described in [4, Theorem 4.1]. Proof of Theorem 5.2. To show that Π is an isomorphism, it is enough to show that the homo- morphism Φ = Ψ−1 of (5.1) factors through the quotient map ρ : C ∗(G, E ) → C ∗(G∂Ω, E ) defined on ΓC(G; E ) by ρ(f ) = f G∂Ω. To see this we use the co-universal property of N Or Lemma 5.5 gives a coaction β : C ∗(G∂Ω, E ) → C ∗(G∂Ω, E ) ⊗ C ∗(G) such that X. Since G∂Ω is G-graded via (g, ω) 7→ g, β(f S) = f S ⊗ iG(qp−1) for all X ∈ K(Xp, Xq). For any x ∈ Xp, we have β(π(x)) = β(f x) = f x ⊗ iG(p) = ((π ⊗ 1) ◦ δ)(jX (x)), where jX : X → N OX is the universal representation. So π is gauge-compatible in the sense of [4]. We aim to apply [4, Theorem 4.1] to π, so we must show that πe : A → C ∗(G∂Ω, E ) 14 ADAM RENNIE, DAVID ROBERTSON, AND AIDAN SIMS is injective. Since the φp are injective, the maps ir : L(Xp) → L(Xpr) appearing in the construction of the fibres Aω, ω ∈ G(0) in Section 3.2 are all injective. Hence the canonical map iω : A = Xe → Xω is injective for each unit ω. In particular, for each a ∈ A, the element πe(A) := f a satisfies f a(ω) = iω(a) 6= 0 for all ω, and πe is injective. Now, writing λr for the canonical quotient map from N OX to N Or X , [4, Theorem 4.1] yields a homomorphism φ : C ∗(G∂Ω, E ) → N Or X that carries f S to λr(j(p) X (S) for S ∈ K(Xp). Fix f ∈ ker(ρ). Without loss of generality, assume that supp(f ) ⊂ G is a bisection. Then φ(ρ(f )) = 0 and hence φ(ρ(f ∗f )) = 0. So we have λr(ρ(Φ(f ∗f ))) = 0. But ρ(Φ(f ∗f )) ∈ (N OX)e and λr(N OX )e is isometric because the reduction map for any coaction is isometric on each spectral subspace. Hence as required. (cid:3) kq(Φ(f ))k2 = kq(Φ(f ∗f ))k = 0 6. Applications Takeishi [27] has recently characterised nuclearity for C ∗-algebras of Fell bundles over ´etale groupoids as follows. Theorem 6.1 ([27, Theorem 4.1]). Let E be a Fell bundle over an ´etale locally compact Haus- dorff groupoid G. If G is amenable, then the following conditions are equivalent (i) The C ∗-algebra C ∗ (ii) The fibre Ex is nuclear for every x ∈ G(0). (iii) The C ∗-algebra C0(E G(0), G(0)) is nuclear. r (E ) is nuclear. For our example, the following lemma shows that (ii) holds whenever the coefficient algebra Xe of the product system X is nuclear. Lemma 6.2. Let (G, P ) be a quasi-lattice ordered group, and let X be a nondegenerate finitely aligned product system over P . If the coefficient algebra Xe of the product system is nuclear, then the fibres Aω, ω ∈ Ω = G(0) are nuclear. Proof. Fix ω ∈ Ω. Arguing as in Lemma 3.2, for each finite F ⊆ ω that is closed under ∨, writing pF for the maximum element of F the set BF = Pp∈F ip−1pF (K(Xp)) is a C ∗-algebra. If F is not a singleton and q ∈ F is minimal, then BF \{q} is an ideal of BF and the quotient BF /BF \{q} is a quotient of iq−1pF (K(Xq)) and hence a quotient of K(Xq). Each K(Xp) is nuclear because it is Morita equivalent to Xe via Xp, and nuclearity is preserved by Morita equivalence [13, Theorem 15]. Fix a finite F ⊆ ω and a minimal q ∈ F , and write F ′ = F \ {q}. Assume as an inductive hypothesis that BF ′ is nuclear. Since BF /BF ′ is a quotient of the nuclear C ∗-algebra K(Xq), it is nuclear. So BF is an extension of a nuclear C ∗- algebra by a nuclear C ∗-algebra, so also nuclear [24, Proposition 2.1.2(iv)]. Now Aω = lim−→F BF is nuclear because direct limits of nuclear C ∗-algebras are nuclear. (cid:3) We therefore have the following theorem. Theorem 6.3. Let X be a nondegenerate finitely-aligned product system over a quasi-lattice ordered group (G, P ), and suppose that the coefficient algebra Xe is nuclear. If the groupoid G of Section 3 is amenable, then N T X and N OX is nuclear. If G∂Ω is amenable and the homomorphisms φp : A → L(Xp) implementing the left actions in X are all injective, then N OX is nuclear. Proof. If G is amenable, then C ∗(G, E ) is amenable by [27, Theorem 4.1] and Lemma 6.2. Since ∼= C ∗(G, E ) by Theorem 5.1, we have N T X nuclear, and then N OX (as defined in [26]) is N T X nuclear because it is a quotient of N T X . If G∂Ω is amenable then C ∗(G∂Ω, E ) is nuclear by [27, PRODUCT SYSTEMS AND FELL BUNDLES 15 Theorem 4.1] and Lemma 6.2. If the φp are injective, then Theorem 5.2 gives an isomorphism N OX (cid:3) ∼= C ∗(G∂Ω, E ), and so N OX is nuclear. We also obtain an improvement on [4, Corollary 4.2]. There it is proved that N OX and N Or X coincide whenever the group G is amenable. But our results show that in fact N OX = N Or X whenever G∂Ω is amenable. Proposition 6.4. Let X be a nondegenerate finitely aligned product system over a quasi-lattice ordered group (G, P ), and suppose that the homomorphism φp : Xe → L(Xp) implementing the left actions in X are all injective. If G∂Ω is amenable, then the quotient map λr : N OX → N Or X is an isomorphism. Proof. Theorem 5.2 gives an isomorphism Π−1 : C ∗(G∂Ω, E ) → N OX. Write c : G → G for the continuous cocycle c(g, ω) = g. Since supp π(x) ⊆ {p} × ∂Ω whenever x ∈ Xp, we see In particular Π−1 restricts to an isomorphism that Π((N OX )g) = Γc(c−1(g); E ) for each g. of the closure of Γc(c−1(e); E ) ⊆ C ∗(G, E ) with (N OX)e. Since c−1(e) = G(0), the closure of Γc(c−1(e); E ) is Γ0(G(0); E ) ⊆ C ∗(G, E ). It is standard that restriction of compactly sup- ported sections to G(0) extends to a faithful conditional expectation C ∗ r (G, E ) → Γ0(G(0); E ). Theorem 1 of [25] implies that C ∗(G∂Ω, E ) = C ∗ r (G∂Ω, E ), so we obtain a faithful conditional expectation R : C ∗(G, E ) → Γ0(G(0); E ) extending restriction of compactly supported sections. Lemma 1.3(a) of [22] shows that there is a conditional expectation P : N OX → (N OX)e that annihilates (N OX )g for g 6= e, and it is routine to check that Π ◦ P = R ◦ Π. Since Π is an iso- morphism and R is a faithful conditional expectation, it follows that P is a faithful conditional expectation as well. That is, the coaction ν on N OX such that δ(jX(x)) = jX (x) ⊗ iG(p) for x ∈ Xp is a normal coaction, and (N OX , G, ν) is a normal cosystem. Corollary 4.6 of [4] shows that N Or X is the C ∗-algebra appearing in the normalisation of the cosystem (N OX , G, ν), and λr is the normalisation homomorphism. Since this cosystem is already normal, we conclude that λr is injective. (cid:3) Remark 6.5. It is worth pointing out, in light of the results in this section, that it is not uncommon for the groupoid G∂Ω of Section 3 to be amenable, even when G is not amenable. For example, G∂Ω is amenable when G is a finitely generated free group -- or more generally a finitely-generated right-angled Artin group -- and P its natural positive cone. References [1] C. Anantharaman-Delaroche, J. Renault, Amenable groupoids, l'Ensignement Math´ematique 36 (2000). [2] B. Blackadar, Operator algebras: Theory of C ∗-algebras and von Neumann algebras, Encyclopaedia of Mathematical Sciences 122 (2006). [3] N. Brownlowe, A. an Huef, M. Laca, I. Raeburn, Boundary quotients of the Toeplitz algebra of the affine semigroup over the natural numbers, Ergodic Theory Dynam. Systems 32 (2012), 35 -- 62. [4] T. Carlsen, N.S. Larsen, A. Sims, S. Vittadello, Co-universal algebras associated to product systems, and gauge-invariant uniqueness theorems, Proc. Lond. Math. Soc., 103 (2011), 563 -- 600. [5] J. Crisp, M. Laca, Boundary quotients and ideals of Toeplitz C ∗-algebras of Artin groups, J. Funct. Anal. 242 (2007), 127 -- 156. [6] K.R. Davidson, A.H. Fuller and E.T.A. Kakariadis, Semicrossed products of operator algebras by semigroups, preprint 2014 (arXiv:1404.1906 [math.OA]). [7] J. Dixmier, C ∗-algebras, Translated from the French by Francis Jellett, North-Holland Mathematical Li- brary, Vol. 15, North-Holland Publishing Co., Amsterdam, 1977, xiii+492. [8] S. Echterhoff, S. Kaliszewski, J. Quigg, I. Raeburn, A categorical approach to imprimitivity theorems for C ∗-dynamical systems, Mem. Amer. Math. Soc. 180 (2006), viii+169. [9] R. Exel, Inverse semigroups and combinatorial C ∗-algebras, Bull. Braz. Math. Soc. (N.S.) 39 (2008), 191 -- 313. [10] R. Exel, M. Laca, J. Quigg, Partial dynamical systems and C ∗-algebras generated by partial isometries, J. Operator Theory 47 (2002), 169 -- 186. [11] J.M.G. Fell, R.S. Doran, Representations of ∗-algebras, locally compact groups, and Banach ∗-algebraic bundles, Pure and Applied Math. 178 (1988). [12] N.J. Fowler, Discrete product systems of Hilbert bimodules, Pacific J. Math. 204 (2002), 335 -- 375. 16 ADAM RENNIE, DAVID ROBERTSON, AND AIDAN SIMS [13] A. an Huef, I. Raeburn, D.P. Williams, Properties preserved under Morita equivalence of C ∗-algebras, Proc. Amer. Math. Soc 135 (2007), 1495 -- 1503. [14] S. Kaliszewski, J. Quigg, Erratum to "Full and reduced C*-coactions". Math. Proc. Camb. Phil. Soc. 116 (1994), 435 -- 450, preprint 2014 (arXiv:1410.7767 [math.OA]). [15] B. Kwasniewski, W. Szyma´nski, Topological aperiodicity for product systems over semigroups of Ore type, preprint 2013 (arXiv:1312.7472 [math.OA]). [16] M. Laca, I. Raeburn, Semigroup crossed products and the Toeplitz algebras of nonabelian groups, J. Funct. Anal. 139 (1996), 415 -- 440. [17] P.S. Muhly, J. Renault, C ∗-algebras of multivariable Wiener-Hopf operators, Trans. Amer. Math. Soc. 274 (1982), 1 -- 44. [18] P.S. Muhly, D.P. Williams, Equivalence and disintegration theorems for Fell bundles and their C ∗-algebras, Dissertationes Mathematicae, Warszawa (2008). [19] A. Nica, C ∗-algebras generated by isometries and Weiner-Hopf operators, J. Operator Theory, 27 (1992), 17 -- 52. [20] M.V. Pimsner, A class of C ∗-algebras generalizing both Cuntz-Krieger algebras and crossed products by Z, Fields Inst. Commun., 12, Free probability theory (Waterloo, ON, 1995), 189 -- 212, Amer. Math. Soc., Providence, RI, 1997. [21] J. Quigg, Full and reduced C ∗-coactions, Math. Proc. Cambridge Philos. Soc. 116 (1994), 435 -- 450. [22] J. Quigg, Discrete C ∗-coactions and C ∗-algebraic bundles, J. Austral. Math. Soc. Ser. A 60 (1996), 204 -- 221. [23] I. Raeburn, A. Sims, Product systems of graphs and the Toeplitz algebras of higher-rank graphs, J. Operator Theory 53 (2005), no. 2, 399 -- 429. [24] M. Rørdam, Classification of nuclear, simple C ∗-algebras, Classification of nuclear C ∗-algebras. Entropy in operator algebras, 1 -- 145, Encyclopaedia Math. Sci., 126, Springer, Berlin, 2002. [25] A. Sims, D.P. Williams, Amenability for Fell bundles over groupoids, Illinois J. Math. 57 (2013), 429 -- 444. [26] A. Sims, T. Yeend, C ∗-algebras associated to product systems of Hilbert bimodules, J. Operator Theory 64 (2010), 349 -- 376. [27] T. Takeishi, On nuclearity of C ∗-algebras of Fell bundles over ´etale groupoids, Publ. Res. Inst. Math. Sci. 50 (2014), 251 -- 268. E-mail address: [email protected], [email protected], [email protected] School of Mathematics and Applied Statistics, University of Wollongong, Wollongong, NSW, 2522, AUSTRALIA
1605.07540
1
1605
2016-05-24T16:54:00
The ideal structure of algebraic partial crossed products
[ "math.OA", "math.RA" ]
Given a partial action of a discrete group $G$ on a Hausdorff, locally compact, totally disconnected topological space $X$, we consider the correponding partial action of $G$ on the algebra $L_c(X)$ consisting of all locally constant, compactly supported functions on $X$, taking values in a given field $K$. We then study the ideal structure of the algebraic partial crossed product $L_c(X)\rtimes G$. After developping a theory of induced ideals, we show that every ideal in $L_c(X)\rtimes G$ may be obtained as the intersection of ideals induced from isotropy groups, thus proving an algebraic version of the Effros-Hahn conjecture.
math.OA
math
THE IDEAL STRUCTURE OF ALGEBRAIC PARTIAL CROSSED PRODUCTS M. Dokuchaev and R. Exel Given a partial action of a discrete group G on a Hausdorff, locally compact, totally disconnected topological space X, we consider the correponding partial action of G on the algebra Lc(X) consisting of all locally constant, compactly supported functions on X, taking values in a given field K. We then study the ideal structure of the algebraic partial crossed product Lc(X) ⋊ G. After developping a theory of induced ideals, we show that every ideal in Lc(X) ⋊ G may be obtained as the intersection of ideals induced from isotropy groups, thus proving an algebraic version of the Effros-Hahn conjecture. 1. Introduction. The study of ideals in crossed product C*-algebras has a long history and is best subsumed by the quest to prove and generalize the celebrated Effros-Hahn conjecture [3] formulated roughly fifty years ago. In its original form, the conjecture states that every primitive ideal in the crossed product of a commutative C*-algebra by a locally compact group should be induced from a primitive ideal in the C*-algebra of some isotropy group. For the case of discrete amenable groups, the Effros-Hahn conjecture was proven by Sauvageot [9], and since then has been extended to various other contexts, notably to locally compact groups acting on non commutative C*-algebras, as proven by Gootman and Rosenberg [6] under separability conditions. Motivated by Fack and Skandalis' study of C*-algebras associated to foliations [10], Renault [8] realized that the Effros-Hahn conjecture also applies for groupoid C*-algebras and proved a version of it in this context. This was later refined by Ionescu and Williams in [7]. Most treatments of the Effros-Hahn conjecture focus on the conjecture itself, namely describing a pre- viously given ideal in the crossed product algebra in terms of induced ideals, rather than studiyng the relationship between the input ideal in the isotropy group algebra and its corresponding output induced ideal. To be fair, there are a few works in the literature where this delicate relationship is discussed. Among them we should mention [15: Theorem 8.39], where a complete classification is given for the collection of primitive ideals on C0(X) ⋊ G, where X is locally compact and G abelian. There, it is shown that every such ideal is induced from some isotropy group and hence arises from a pair (x, χ), where x is a point in X, and χ a character on G. Most importantly, the primitive ideals for two pairs (x1, χ1) and (x2, χ2) coincide if and only if the closure of the orbit of x1 coincides with that of x2, and χ1 coincides with χ2 on the isotropy group of x1. In the above situation, when two points have the same orbit closure, it may be shown that their isotropy groups coincide, but this relies heavily on the commutativity of G. In particular it is not even clear how to phrase the above condition in case G is not commutative. The question of when two induced ideals coincide is also taken up in [11], where the object of study k by local is the C*-algebra of the Deaconu-Renault groupoid built from an action of the semigroup  homeomorphisms on a locally compact space X. The primitive ideals of this C*-algebra are shown to be k, and parametrized by pairs (x, χ), where x is a point in X, and χ a character on  again a criterion is given for when two such pairs lead to the same primitive ideal. The condition is that the two points of X must have identical orbit closures and the two characters must coincide as characters on the interior of the isotropy of the groupoid reduced to the common orbit closure. In a sense, this result also relies on commutativity. k, i.e. an element of T Date: 22 May 2014. Key words and phrases: Effros-Hahn conjecture, partial crossed product, induced representation, induced ideal, locally constant function. Partially supported by CNPq and FAPESP. 2 m. dokuchaev and r. exel In the present paper our aim is to study the Effros-Hahn conjecture in a new setting, namely the algebraic partial crossed product Lc(X) ⋊ G, where G is a not necessarily commutative, discrete group partially acting on a locally compact, totally disconnected topological space X, and Lc(X) is the algebra consisting of all locally constant, compactly supported functions on X, taking values in a given field K. The justification for studying this setting comes from the current interest to investigate purely algebraic versions os some intensely studied C*-algebras, such as the Leavitt path algebras which may be viewed as algebraic counterparts of graph C*-algebras. In many such cases the pertinent C*-algebra is a C*-algebraic partial crossed product of the form C0(X) ⋊ G, with X totally disconnected, while its algebraic sibling is the algebraic partial crossed product Lc(X) ⋊ G. Steinberg algebras [12] in fact generalize this correspondence to the case where an ample ´etale groupoid replaces the above partial action. One of the main results of the present paper, namely Theorem (6.3), is a version of the Effros-Hahn conjecture, where we prove that every ideal of Lc(X) ⋊ G is given as the intersection of ideals induced from isotropy groups. The method of proof is entirely elementary and does not rely on the measure theoretical or analytical tools on which the main proofs in [9], [6] and [8] are based, chiefly because our setting is eminently algebraic. The strategy adopted here is as follows: given an ideal J of Lc(X) ⋊ G, we first choose a representation π of Lc(X) ⋊ G whose null space coincides with J. We then build another representation, which we call the discretization of π, whose null space coincides with that of π, and hence also with J. The discretized representation is then easily seen to decompose as a direct sum of sub-representations based on the orbits for the action of G on X. Each such sub-representation is finally shown to be equivalent to an induced representation, and hence the initially given ideal J is seen to coincide with the intersection of the null spaces of the various induced representation involved, each of which is then an induced ideal. Only a tiny amount of the theory of induced ideals is necessary to prove Theorem (6.3), our version of the Effros-Hahn conjecture, but still we have chosen to start the study of induced ideals before stating and proving (6.3), mostly in order to be able to refer to the main concepts involved. Before and after the proof of Theorem (6.3), in fact throughout the paper, we develop tools designed to understand the induction process itself, attempting to describe how exactly does an induced ideal Indx0(I) depends on point x0 and on the ideal I it is induced from. From the outset, this dependency is expected to be quite tricky for the following reason: there are numerous examples where a crossed product algebra turns out to be simple (see e.g. [5: Theorem 4.1]), and hence the assortment of ideals in the crossed product is rather boring, but still there may be points with nontrivial isotropy (not too many since the action must be topologically free by [1], but this does not ruled out all points), and hence there may be many ideals presenting themselves as input for the induction process. However, as already mentioned, the output could be totally uninteresting due to simplicity. The explanation for this phenomenon is that, when inducing from the isotropy group Hx0 , where x0 is some point in X, not all ideals in KHx0 play a relevant role. Those which do, namely the ones we call admissible, are the only ones deserving attention in the sense that for every ideal I E KHx0, there exists a unique admissible ideal I ′ ⊆ I, which induces the same ideal of Lc(X) ⋊ G as I does. This is the content of Corollary (4.7). The correspondence I 7→ Indx0(I) is consequently seen to be a one-to-one mapping from the set of admissible ideals in KHx0 to the set of ideals in Lc(X) ⋊ G. This naturally raises the question of which are the admissible ideals in KHx0, a question we answer in general in (4.10), and then in a few special cases in (8.4), (9.5), (9.6) and (10.3). We then consider the question of when two induced ideals Indx0(I) and Indx0 ( I) coincide, where x0 and x0 are two points in X, I is an admissible ideal in KHx0, and I is an admissible ideal in KHx0 (based on our study of induced ideals it suffices to consider the case where I and I are admissible). As already mentioned, when x0 = x0 one has that Indx0(I) = Indx0( I) ⇐⇒ I = I, so the remaining case is when x0 6= x0. As in [15] and [11], a necessary condition for Indx0(I) and Indx0( I) to coincide is that the orbits of x0 and x0 have the same closure (as long as I and I are proper ideals, according to (11.2)) but now, in the absence of commutativity, the isotropy groups of x0 and x0 no longer need to be the same, so comparing the extra data within KHx0 and KHx0 (namely I versus I) is no longer a straightforward matter. the ideal structure of algebraic partial crossed products 3 To deal with this situation we introduce the notion of transposition of ideals in (11.1) which is a way of comparing ideals in different isotropy groups. Our main result in that direction, namely Theorem (11.3), says that Indx0 (I) = Indx0( I) if and only if I is the transposition of I and vice versa. We have already mentioned that the algebra Lc(X) ⋊ G, which is our main object of study, may also be described as the Steinberg algebra for the transformation groupoid associated to the partial action of G on X. Steinberg's results obtained in [12] and [13] therefore apply to our situation as well. On the other hand, in all likelihood our results may be shown to hold for Steinberg algebras with minor modifications in our proofs. Our algebras are all taken to be over a fixed field K, but in most places our results hold under the more general assumption that K is just a unital commutative ring. Notable exceptions are (3.16) and (8.4), where invertibility of nonzero elements in K is crucial. The second named author would like to acknowledge financial support from the Funda¸cao de Amparo `a Pesquisa do Estado de Sao Paulo (FAPESP) during a visit to the University of Sao Paulo, where a large part of this work was conducted. He would also like to acknowledge the warm hospitality of the members of the Mathematics Department during that visit. 2. Preliminaries. Throughout most of this work we will assume the following: 2.1. Standing Hypotheses. (a) K is a field, (b) G is a discrete group, ( c) X is a Hausdorff, locally compact, totally disconnected1 topological space, (d) θ = ({θg}g∈G, {Xg}g∈G) is a (topological) partial action [4: Definition 5.1] of G on X, such that Xg is clopen (closed and open) for every g in G, ( e) whenever appropriate, we will also fix a distinguished point x0 in X. Recall that a function f : X → K is said to be locally constant if, for every x in X, there exists a neighborhood V of x, such that f is constant on V . The support of f is defined to be the set supp(f ) = {x ∈ X : f (x) 6= 0}. Observe that the support of a locally constant function f is always closed, so we will not bother to define the support as the closure of the above set, as sometimes done in analysis. By virtue of being locally compact and totally disconnected, we have that the topology of X admits a basis formed by compact-open 2 subsets. Given any compact-open set E ⊆ X, it is easy to see that its characteristic function, here denoted by 1E, is locally constant and compactly supported. Moreover, one may easily prove that every locally constant, compactly supported function f : X → K is a linear combination of the form n f = ci 1Ei, Xi=1 where the Ei are pairwise disjoint compact-open subsets and the ci lie in K. We will henceforth denote by Lc(X) the set of all locally constant, compactly supported, K-valued functions on X. With pointwise multiplication, Lc(X) is a commutative K-algebra, which is unital if and only if X is compact. 1 A locally compact topological space is totally disconnected if and only if it admits a basis of open sets consisting of sets which are also closed [14: Theorem 29.7]. 2 That is, sets which are simultaneously compact and open. 4 m. dokuchaev and r. exel For each g in G, we may also consider the K-algebra Lc(Xg), which we will identify with the set formed by all f in Lc(X) vanishing on X \ Xg. Under this identification Lc(Xg) becomes an ideal in Lc(X). Regarding the homemorphism θg : Xg−1 → Xg, we may define an isomorphism by setting αg : Lc(Xg−1 ) → Lc(Xg), αg(f ) = f ◦ θg−1 , ∀ f ∈ Lc(Xg−1 ). The collection formed by all ideals Lc(Xg), together with the collection of all αg, is then easily seen to be an (algebraic) partial action [4: Definition 6.4] of G on Lc(X). This is a unital partial action (one for which the domain ideals are unital) if and only if all of the Xg are compact. However we shall prefer to consider the more general situation where the Xg are only assumed to be closed (besides being open). The main goal of this paper is to study the algebraic crossed product as defined in [4: Definition 8.3]. A general element b ∈ Lc(X) ⋊ G will be denoted by Lc(X) ⋊ G, fg∆g, b = Xg∈G where each fg lies in Lc(Xg−1 ), and fg = 0, for all but finitely many group elements g. In many texts dealing with crossed products the above place markers "∆g" are denoted "δg", but we shall reserve the latter to denote elements in KG, such as chδh, Xh∈H where the ch are scalars in K, again equal to zero except for finitely many group elements g. In fact, throughout this paper all summations will be finite, either because the set of indices is finite, or because all but finitely many summands are supposed to vanish. Since we are asuming that every Xg is clopen, its characteristic function 1Xg , which we will abbreviate to 1g := 1Xg , is a locally constant function, although not necessarily compactly supported. However, given any f in Lc(X), one has that f 1g−1 is compactly supported, so it belongs to Lc(Xg). We may therefore define ¯αg(f ) := αg(f 1g−1 ). so that ¯αg is a globally defined endomorphism of Lc(X). Recall that if g and h are elements of G, and if e ∈ Lc(Xg), and f ∈ Lc(Xh), then the product of e∆g by f ∆h is defined by (e∆g)(f ∆h) = αg(cid:0)αg−1 (e)f(cid:1)∆gh. (2.2) In our present situation this expression may be made simpler as follows: since αg−1 (e) lies in Lc(Xg−1 ), we have that so the coefficient of ∆gh in (2.2) equals αg−1 (e) = αg−1 (e)1g−1 , αg(cid:0)αg−1 (e)f(cid:1) = αg(cid:0)αg−1 (e)1g−1 f(cid:1) = αg(cid:0)αg−1 (e)(cid:1)αg(cid:0)1g−1 f(cid:1) = e ¯αg(f ). The promissed simpler formula for the product thus reads (e∆g)(f ∆h) = e ¯αg(f )∆gh. (2.3) the ideal structure of algebraic partial crossed products 5 3. Induction. As always, we assume the conditions set out in (2.1). From here on the distinguished point x0 mentioned in (2.1.e) will become important in our development and we will henceforth use the following notations Sx0 := {g ∈ G : x0 ∈ Xg−1 }, Hx0 := {g ∈ G : x0 ∈ Xg−1 , θg(x0) = x0}, Orb(x0) := {θg(x0) : g ∈ S}. (3.1) Whenever there is no question as to which point x0 we are referring to, as it will often be the case, we will omit subscripts and write S and H in place of Sx0 and Hx0, resectively. Notice that H is a subgroup of G, often called the isotropy group of x0. On the other hand observe that SH ⊆ S, so S is a union of left H-classes. The map g ∈ S 7−→ θg(x0) ∈ Orb(x0) is clearly onto, and two elements g1 and g2 in S satisfy θg1 (x0) = θg2 (x0), if and only if they lie in the same left H-class. A central ingredient in the induction process to be introduced shortly, is the subspace M of the group algebra KG given by M = span{δg : g ∈ S}. As already observed, SH ⊆ S, so it follows that M is naturally a right KH-module. Consider the unique bilinear form such that This may also be written as h· , ·i : M × M → KH hδk, δli =( δk−1l, 0, if k−1l ∈ H, otherwise. hδk, δli = [k−1l∈H] δk−1l, where the brackets indicate boolean value 3. An important property of this form is expressed by the identity (3.2) hm, nai = hm, nia, ∀ m, n ∈ M, ∀ a ∈ KH, (3.3) which the reader may easily prove. 3.4. Proposition. If R ⊆ S is a system of representatives of left H-classes, so that . then, for all m in M , one has rH, δrhδr, mi, S = Sr∈R m = Xr∈R where the sum is always finite in the sense that there are only finitely many nonzero summands. 3 In fact we shall often use boolean values in this work, sometimes in a slightly abusive fashion, such as in [x∈X g−1 ] f (θg (x)), where f is some scalar valued function on X. The principle behind this is that, when x is not in the domain Xg−1 of θg, so that θg(x) is not defined, the zero boolean value of the expression "x ∈ Xg−1 " predominates and turns the whole expression into zero. In other words, zero times something which is not defined is taken to be zero. It is true that an excessive abuse of this principle may perhaps lead to unexpected consequences, but we promisse to use it only to shorten expressions which could otherwise be writen in two clauses, such as (3.2). 6 m. dokuchaev and r. exel Proof. Assuming that m = δl, there exists a unique k in R such that lH = kH, which is to say that k−1l ∈ H. Then Xr∈R δrhδr, δli = Xr∈R δr [r−1l∈H] δr−1l = δkδk−1l = δl. The general case folows by writing m as a linear combination of the δl. Besides being a right KH-module, M is also a left module: 3.5. Proposition. There is a left (Lc(X) ⋊ G)-module structure on M such that for every f ∈ Lc(Xg), and all l ∈ S. With this, M moreover becomes an (Lc(X) ⋊ G) - KH - bimodule. (f ∆g)δl = [gl∈S] f(cid:0)θgl(x0)(cid:1)δgl, Proof. Left for the reader. Given any left KH-module V , one may therefore build the left (Lc(X) ⋊ G)-module M ⊗KH V, (cid:3) (cid:3) henceforth denoted simply by M ⊗ V . This left module structure is well known, but it might be worth spelling it out here: given b ∈ Lc(X) ⋊ G, one has b(m ⊗ v) = (bm) ⊗ v, ∀ m ∈ M, v ∈ V. 3.6. Definition. The left (Lc(X) ⋊ G)-module M ⊗ V mentioned above is said to be the module induced by V . Recall that a module V is said to be irreducible if it has no nontrivial submodules or, equivalently, if the submodule generated by any nonzero element coincides with V . 3.7. Proposition. If V is an irreducible KH-module, then M ⊗ V is irreducible as a (Lc(X) ⋊ G)-module. Proof. Given any nonzero vector w ∈ M ⊗ V , we must show that the submodule it generates, here denoted by hwi, coincides with M ⊗ V . In order to do this, write w = n Xi=1 mi ⊗ ui, and let R ⊆ S be a system of representatives of left classes for S modulo H. So by (3.4) we have n n n w = Xi=1Xr∈R δrhδr, mii ⊗ ui = Xr∈R Xi=1 δr ⊗ hδr, miiui = Xr∈R δr ⊗ Xi=1 hδr, miiui = Xr∈R δr ⊗ vr, where the vr are defined by the last equality above. Since all sums involved are finite, the set Γ = {r ∈ R : vr 6= 0} must be finite. It is moreover nonempty, since we are assuming that w 6= 0. Fixing any s in R, we claim that δs ⊗ vs lies in hwi. To see this, notice that no two elements of R are in the same left H-class, so the points θr(x0) are pairwise distinct. We may then pick some f in Lc(X) such that f(cid:0)θs(x0)(cid:1) = 1, while f(cid:0)θr(x0)(cid:1) = 0, for all r ∈ Γ \ {s}. We then have [r∈S] f (θr(x0))δr ⊗ vr = δs ⊗ vs. hwi ∋ f w =Xr∈Γ f δr ⊗ vr =Xr∈Γ the ideal structure of algebraic partial crossed products 7 We next show that hwi contains δs ⊗ V . Since V is irreducible as a KH-module, we have that V is spanned by the set {δhvs : h ∈ H}, so it is enough to prove that δs ⊗ δhvs ∈ hwi, ∀ h ∈ H. (3.7.1) We thus fix some h in H, and put g = shs−1. Observing that we see that x0 ∈ X(gs)−1 ∩ Xs−1. Consequently θs(x0) = θs(cid:0)θh(x0)(cid:1) = θsh(x0) = θgs(x0), θgs(x0) ∈ θgs(cid:0)X(gs)−1 ∩ Xs−1(cid:1) = Xgs ∩ Xg. We may then choose some f in Lc(Xg) such that f(cid:0)θgs(x0)(cid:1) = 1, and then hwi ∋ f ∆g(δs ⊗ vs) = [gs∈S] f(cid:0)θgs(x0)(cid:1)δgs ⊗ vs = δgs ⊗ vs = δsh ⊗ vs = δsδh ⊗ vs = δs ⊗ δhvs, thus proving (3.7.1), and hence that δs ⊗ V ⊆ hwi. We will conclude the proof by showing that δk ⊗ V ⊆ hwi, for every k in S. Given any such k, set g = ks−1, and notice that x0 ∈ Xk−1 ∩ Xs−1 , so θk(x0) ∈ θk(cid:0)Xk−1 ∩ Xs−1(cid:1) = Xk ∩ Xks−1 ⊆ Xg. So we may find some f in Lc(Xg) such that f(cid:0)θk(x0)(cid:1) = 1, and for every v in V , one has hwi ∋ f ∆g(δs ⊗ v) = [gs∈S] f(cid:0)θgs(x0)(cid:1)δgs ⊗ v = [k∈S] f(cid:0)θk(x0)(cid:1)δk ⊗ v = δk ⊗ v. This shows that δk ⊗ V ⊆ hwi, as desired, and hence that hwi = M ⊗ V , concluding the proof. (cid:3) Our next goal will be to compute the annihilator of the induced module in terms of the annihilator of the original module V . We begin with a useful technical result. 3.8. Lemma. Let V be a left KH-module and let I be the annihilator of V in KH. Given m ∈ M , the following are equivalent: (i) m ⊗ v = 0, for all v in V , (ii) hn, mi ∈ I, for all n ∈ M . Proof. (ii) ⇒ (i): Let R be a system of representatives of left H-classes in S. Then for every v in V we have m ⊗ v (3.4) = Xr∈R δrhδr, mi ⊗ v = Xr∈R δr ⊗ hδr, miv = 0. (i) ⇒ (ii): Fixing n ∈ M , consider the bilinear mapping (m, v) ∈ M × V 7→ hn, miv ∈ V. By (3.3) this is KH-balanced, so there is a well defined K-linear mapping Tn : M ⊗ V → V , such that Tn(m ⊗ v) = hn, miv, for all m in M , and all v in V . Assuming that m satisfies (i) we then have that hn, miv = Tn(m ⊗ v) = 0, ∀ n ∈ M, ∀ v ∈ V, so hn, mi annihilates V , whence hn, mi ∈ I, proving (ii). (cid:3) As a consequence we obtain the following description of the annihilator of an induced module. 8 m. dokuchaev and r. exel 3.9. Corollary. Let V be a left KH-module and let I be the annihilator of V in KH. Then the annihilator of M ⊗ V in Lc(X) ⋊ G is given by {b ∈ Lc(X) ⋊ G : hn, bmi ∈ I, ∀n, m ∈ M }. In particular the annihilator of M ⊗ V depends only on I. Proof. One has that b lies in the annihilator of M ⊗ V , iff bm ⊗ v = 0, for all m and v, which is equivalent to saying that hn, bmi ∈ I, for all n and m, by (3.8). (cid:3) Since the annihilator of M ⊗ V depends only on I, rather than on V , we may think of it as built out of I. To account for this we give the following: 3.10. Definition. Given any ideal4 I E KH, we shall let Indx0 (I) := {b ∈ Lc(X) ⋊ G : hn, bmi ∈ I, ∀n, m ∈ M }. This will be referred to as the ideal induced by I. When there is no risk of confusion we shall write this simply as Ind(I). Reinterpreting (3.9) with the terminology just introduced, we have: 3.11. Proposition. Let V be a left KH-module and let I be the annihilator of V in KH. Then the annihilator of M ⊗ V coincides with the ideal induced by I. So far it is clear that Ind(I) is a right ideal, but we will shortly prove that it is indeed a two-sided ideal. The behavior of the induction process under inclusion and intersection is easy to understand: 3.12. Proposition. (i) If I1 and I2 are ideals of KH with I1 ⊆ I2, then Ind(I1) ⊆ Ind(I2). (ii) Given any family {Iλ}λ∈Λ of ideals of KH, then Ind(cid:16)Tλ∈Λ Iλ(cid:17) =Tλ∈Λ Ind(Iλ). Proof. Follows easily by inspecting the definitions involved. (cid:3) When checking that hn, bmi ∈ I, for all n, m ∈ M , as required by the above definition, it suffices to consider m = δk and n = δl, for k, l ∈ S, since these generate M . It is therefore nice to have an explicit formula for use in this situation: 3.13. Proposition. Given b =Pg∈G fg∆g in Lc(X) ⋊ G, and given k and l in S, one has that hδk, bδli = Xg∈kHl−1 fg(cid:0)θk(x0)(cid:1) δk−1gl. Proof. We have hδk, bδli = Xg∈G hδk, [gl∈S] fg(cid:0)θgl(x0)(cid:1)δgli = hδk, (fg∆g)δli = Xg∈G = Xg∈kHl−1 [gl∈S] fg(cid:0)θgl(x0)(cid:1) δk−1gl = · · · If g ∈ kHl−1, then gl lies in kH, so θgl(x0) is indeed defined, meaning that gl ∈ S, and θgl(x0) coincides with θk(x0). So the above equals · · · = Xg∈kHl−1 fg(cid:0)θk(x0)(cid:1) δk−1gl, concluding the proof. (cid:3) The above computation allows for a very concrete criteria for membersip in Ind(I), namely: 4 Unless otherwise stated, all ideals in this paper are assumed to be two-sided. the ideal structure of algebraic partial crossed products 9 3.14. Proposition. Given any ideal I E KH, and given any b = Pg∈G fg∆g in Lc(X) ⋊ G, one has that b ∈ Ind(I), if and only if, for every k and l in S, one has that Xg∈kHl−1 fg(cid:0)θk(x0)(cid:1) δk−1gl ∈ I. Proof. Follows from (3.13) and the fact that the δk generate M , as a K-vector space. (cid:3) Let us now discuss two trivial examples: 3.15. Proposition. (a) If I = KH, then Ind(I) coincides with the whole algebra Lc(X) ⋊ G. (b) If I = {0}, then Ind(I) =nPg∈G fg∆g ∈ Lc(X) ⋊ G : fg Orb(x0) = 0, ∀g ∈ Go. Proof. The first statement is clear. As for (b), first notice that a locally constant function vanishing on Orb(x0), necessarily also vanishes on the closure Orb(x0). This said, let fg∆g ∈ Lc(X) ⋊ G. b = Xg∈G Assuming that b lies in Ind(I), and given any point y in the orbit of x0, we will prove that fg(y) = 0, for all g. In case y /∈ Xg, it is clear that fg(y) = 0, since the support of fg is contained in Xg. Otherwise, if y ∈ Xg, write y = θk(x0), for some k ∈ S, and observe that y ∈ Xk ∩ Xg, so from where we see that l := g−1k lies in S. Consequently x0 = θk−1 (y) ∈ θk−1 (Xk ∩ Xg) = Xk−1 ∩ Xk−1g, {0} = I ∋ hδk, bδli (3.13) = Xg′∈kHl−1 fg′(cid:0)θk(x0)(cid:1) δk−1g′l. Among the above summands, one is to find g′ = k1l−1 = kk−1g = g, so in particular This shows that fg vanishes on the orbit of x0, and hence also on its closure. 0 = fg′(cid:0)θk(x0)(cid:1) = fg(y). Conversely, assuming that each fg vanishes on the orbit of x0, it is clear from (3.13) that hδk, bδli = 0 ∈ I, (cid:3) so b ∈ Ind(I). Much has been said about the intersection of an ideal in a crossed product algebra and its intersection with the coefficient algebra. In the case of induced ideals we have: 3.16. Proposition. Let I be a proper 5 ideal in KH. Then the intersection Ind(I) ∩ Lc(X) consists of all f in Lc(X) vanishing on Orb(x0). Proof. Let f ∈ Ind(I) ∩ Lc(X). Then, choosing any k in S, we have I ∋ hδk, f δki (3.13) = f(cid:0)θk(x0)(cid:1)δ1. Should f(cid:0)θk(x0)(cid:1) not vanish, the above would be an invertible element in I, whence I = KH, contradicting the hypothesis. Thus f(cid:0)θk(x0)(cid:1) = 0, showing that f vanishes on the orbit of x0, and hence also on its closure. Conversely, if f vanishes on Orb(x0), then by (3.15.ii) f ∈ Ind({0}) ⊆ Ind(I). (cid:3) 5 We say that an ideal in an algebra is proper when it is not equal to the whole algebra. 10 m. dokuchaev and r. exel Consider the map given by E = Ex0 : Lc(X) ⋊ G → Lc(X) ⋊ H, (3.17) This is sometimes called a conditional expectation. One of its important properties is that of being a (Lc(X) ⋊ H)-bimodule map in the sense that if a ∈ Lc(X) ⋊ H, and b ∈ Lc(X) ⋊ G, then E(cid:16)Xg∈G fg∆g(cid:17) = Xh∈H fh∆h. E(ab) = aE(b) and E(ba) = E(b)a. It is also evident that E is a projection from Lc(X) ⋊ G onto Lc(X) ⋊ H. Consider also the map given by ν = νx0 : Lc(X) ⋊ H → KH, ν(cid:16) Xh∈H fh∆h(cid:17) = Xh∈H fh(x0)δh. Since x0 is fixed by H, one may easily show that ν is an algebra homomorphism. The composition of ν and E is therefore the map (3.18) (3.19) F = Fx0 = ν ◦ E : Lc(X) ⋊ G → KH, (3.20) given by F(cid:16)Xg∈G fg∆g(cid:17) = Xh∈H fh(x0)δh. There is a useful relationship between F and the above bilinear form h· , ·i, expressed as follows: 3.21. Lemma. Let k, l ∈ G, and choose p ∈ Lc(Xk−1 ), and q ∈ Lc(Xl), so that defining one has that u and v are in Lc(X) ⋊ G. Then, for every b in Lc(X) ⋊ G, one has that u = p∆k−1 , and v = q∆l, F (ubv) =( p(x0) q(cid:0)θl(x0)(cid:1) hδk, bδli, 0, if k, l ∈ S, otherwise. Proof. Write b =Pg∈G fg∆g, so that E(ubv) = E(cid:16)Xg∈G p∆k−1 fg∆g q∆l(cid:17) = E(cid:16)Xg∈G = Xg∈kHl−1 p αk−1(fg1k) αk−1 g(q1g−1k)∆k−1gl. p αk−1 (fg1k) αk−1 g(q1g−1k)∆k−1gl(cid:17) = Therefore p(x0) αk−1 (fg1k)x0 αk−1g(q1g−1k)x0 δk−1gl = F (ubv) = ν(cid:0)E(ubv)(cid:1) = Xg∈kHl−1 = Xg∈kHl−1 p(x0) [x0∈Xk−1 ] fg(cid:0)θk(x0)(cid:1) [x0∈Xk−1 g] q(cid:0)θg−1k(x0)(cid:1)δk−1gl = · · · Notice that, whenever r−1s ∈ H, one has that θr−1s(x0) = x0, so x0 ∈ Xr−1 ⇐⇒ x0 ∈ Xs−1 , the ideal structure of algebraic partial crossed products 11 and if these equivalent conditions hold then Applying this to r = g−1k, and s = l, we see that the above equals θr(x0) = θs(x0). · · · = Xg∈kHl−1 p(x0) [x0∈Xk−1 ] fg(cid:0)θk(x0)(cid:1) [x0∈Xl−1] q(cid:0)θl(x0)(cid:1)δk−1gl = = [x0∈Xk−1 ] [x0∈Xl−1 ] p(x0) q(cid:0)θl(x0)(cid:1) Xg∈kHl−1 fg(cid:0)θk(x0)(cid:1)δk−1gl (3.13) = [k,l∈S] p(x0) q(cid:0)θl(x0)(cid:1) hδk, bδli. (cid:3) Let us now use the above result with the purpose of giving an alternative definition of Ind(I), where F is employed instead of the form h· , ·i. 3.22. Proposition. Given any ideal I E KH, one has that Ind(I) = {b ∈ Lc(X) ⋊ G : F (ubv) ∈ I, ∀u, v ∈ Lc(X) ⋊ G}. Proof. In order to prove the statement we must show that, for any given b ∈ Lc(X) ⋊ G, the following are equivalent: (i) F (ubv) ∈ I, for all u, v ∈ Lc(X) ⋊ G, (ii) hn, bmi ∈ I, for all n, m ∈ M . (i) ⇒ (ii): It is clearly enough to prove (ii) for n = δk, and m = δl, where k and l are arbitrary elements of S. In order to do this, pick p ∈ Lc(Xk−1 ), and q ∈ Lc(Xl), such that p(x0) = 1, and q(cid:0)θl(x0)(cid:1) = 1. Then (3.21) hn, bmi = hδk, bδli = p(x0) q(cid:0)θl(x0)(cid:1) hδk, bδli = F (ubv) ∈ I, proving (ii). (ii) ⇒ (i): It is clearly enough to prove (i) for u = p∆k−1 , and v = q∆l, where k and l are arbitrary elements of G, while p ∈ Lc(Xk−1 ), and q ∈ Lc(Xl). If k and l lie in S, we have that On the other hand, if either k or l are not in S, then again by (3.21), we have that F (ubv) = 0 ∈ I. (cid:3) F (ubv) (3.21) = p(x0)q(cid:0)θl(x0)(cid:1)hδk, bδli ∈ I. When I is the annihilator of a left KH-module V , we have seen that Ind(I) is the annihilator of M ⊗ V , hence a two-sided ideal. Should there be any doubt that Ind(I) is always a two-sided ideal (in case I is not presented6 as the annihilator of some left KH-module), the above description of Ind(I) may be used to dispell this doubt. 4. Admissible ideals. As always we assume (2.1). So far we have not considered the question of which ideals I E KH actually lead to nontrivial induced ideals. In case θ is topologically free and minimal, a situation well known to lead to a simple crossed product [5: Theorem 4.1], at least some points x0 in X are allowed to possess a nontrivial isotropy group H, and hence there might be plenty ideals I E KH to choose from, but the simplicity of Lc(X) ⋊ G prevents induced ideals from being nontrivial. In this section we shall begin to explore the delicate relationship between ideals of the isotropy group algebra and the induced ideals they lead to. The map Fx0 , introduced in (3.20), will play a crucial role in this section. Because we will always consider the induction process relative to the point x0 fixed in (2.1.e), we will abolish the subscript writing F in place of Fx0 . 6 Notice, however, that any ideal of a unital algebra is the annihilator of some left module, namely the quotient algebra. 12 m. dokuchaev and r. exel 4.1. Proposition. Given a and b in Lc(X) ⋊ G, such that either a or b is in Lc(X) ⋊ H, then F (ab) = F (a)F (b). Proof. Suppose that b is in Lc(X) ⋊ H. Then, using (3.18), we have F (ab) = ν(cid:0)E(ab)(cid:1) = ν(cid:0)E(a)b(cid:1) = ν(cid:0)E(a)E(b)(cid:1) = ν(cid:0)E(a)(cid:1)ν(cid:0)E(b)(cid:1) = F (a)F (b). A similar reasoning applies when a is in Lc(X) ⋊ H. In particular, if ϕ is in Lc(X), then F (aϕ) = ϕ(x0)F (a) = F (ϕa), ∀ a ∈ Lc(X) ⋊ G. (cid:3) (4.2) Another instance of (4.1) is obtained when ϕ is supported on Xh, for some h in H, in which case we have F (a ϕ∆h) = F (a)ϕ(x0)δh, and F (ϕ∆h a) = ϕ(x0)δh F (a). (4.3) 4.4. Proposition. If J is any ideal in Lc(X) ⋊ G, then F (J) is an ideal in KH. Proof. Given c ∈ F (J), and d ∈ KH, we must prove that cd and dc lie in F (J), and it clearly suffices to assume that d = δh, for some h in H. Choose b in J such that F (b) = c, and let ϕ ∈ Lc(Xh) be such that ϕ(x0) = 1. Then cd = cϕ(x0)δh = F (b)F (ϕ∆h) (4.3) = F (bϕ∆h) ∈ F (J). A similar reasoning proves that dc ∈ F (J). (cid:3) Applying this to induced ideals we get the following: 4.5. Proposition. Let I be an ideal in KH, and put I ′ = F (Ind(I)). Then (i) I ′ is an ideal of KH, (ii) I ′ ⊆ I, (iii) Ind(I ′) = Ind(I). Proof. (i) Follows from (4.4). (ii) Given a in I ′, write a = F (b), with b ∈ Ind(I). Choosing u ∈ Lc(X), with u(x0) = 1, we then have (3.22) ∋ F (ubu) I (4.2) = u(x0)F (b)u(x0) = F (b) = a. This proves (ii). (iii) Since I ′ ⊆ I, it is obvious that Ind(I ′) ⊆ Ind(I). On the other hand, if b ∈ Ind(I), then for every u and v in Lc(X) ⋊ G, one has that ubv ∈ Ind(I), whence F (ubv) ∈ F (Ind(I)) = I ′. This proves that b ∈ Ind(I ′), concluding the proof. (cid:3) The grand conclusion of this result is that, should we want to catalogue all induced ideals, we do not need to consider all ideals I E KH, since I may be replaced by I ′, without affecting the outcome of the induction process. This motivates the question of how to separate the ideals that matter from those which don't, a task we now begin to undertake. 4.6. Definition. An ideal I E KH is said to be admissible if F (Ind(I)) = I. Interpreting (4.5) from the point of view of the concept just introduced we have: the ideal structure of algebraic partial crossed products 13 4.7. Corollary. For every ideal I E KH, there exists a unique admissible ideal I ′ ⊆ I, such that Ind(I) = Ind(I ′). Proof. Set I ′ = F (Ind(I)). Then Ind(I) = Ind(I ′), by (4.5.iii). Moreover F (Ind(I ′)) = F (Ind(I)) = I ′, so I ′ is admissible. If I ′ and I ′′ are two admissible ideals inducing the same ideal of Lc(X) ⋊ G, then I ′ = F (Ind(I ′)) = F (Ind(I ′′)) = I ′′. We have in fact already encountered examples of admissible ideals: 4.8. Proposition. The two trivial ideals of KH, namely {0} and KH, itself, are admissible. Proof. Setting I = {0}, we have that I = {0} ⊆ F (Ind(I)) (4.5.ii) ⊆ I, so I is admissible. On the other hand, if I = KH, we have seen in (3.15) that Ind(I) = Lc(X) ⋊ G, so so, again, I is seen to be admissible. F (Ind(I)) = F(cid:0)Lc(X) ⋊ G(cid:1) = KH = I, (cid:3) (cid:3) In order to better understand admissible ideals, we must be able to describe the image of an ideal in Lc(X) ⋊ G through F . 4.9. Proposition. Let J E Lc(X) ⋊ G be any ideal and let c =Ph∈H chδh be any element of KH. Then c is in F (J) if and only if there exists a compact-open set V , such that whenever ch 6= 0, satisfying Proof. Assuming that cV is in J, we have x0 ∈ V ⊆ Xh, cV := Xh∈H ch1V ∆h ∈ J. F (J) ∋ F (cV ) = Xh∈H ch1V (x0)δh = Xh∈H chδh = c, so c ∈ F (J). Conversely, if c ∈ F (J), pick b in J such that c = F (b). We will initially prove that b may be chosen in Lc(X) ⋊ H. Write b =Pg∈Γ fg∆g, where Γ is a finite subset of G, and set Γ1 = {g ∈ Γ : x0 /∈ Xg−1 }, Γ2 = {g ∈ Γ : x0 ∈ Xg−1 , θg(x0) 6= x0}, Γ3 = {g ∈ Γ : x0 ∈ Xg−1 , θg(x0) = x0} = Γ ∩ H. It is then clear that Γ is the disjoint union of the Γi. 14 m. dokuchaev and r. exel For each g in Γ1, using that Xg−1 is closed, choose a compact-open set Wg, such that x0 ∈ Wg ⊆ X \ Xg−1 . For each g in Γ2, choose open sets U and V , such that x0 ∈ U , θg(x0) ∈ V , and U ∩ V = ∅. By replacing V with V ∩ Xg we may assume that V ⊆ Xg. We then set Z = U ∩ θg−1 (V ), and observe that x0 ∈ Z ⊆ Xg−1 , and that Choosing a compact-open neighborhood Wg of x0 contained in Z, we then have that Z ∩ θg(Z) ⊆ U ∩ θg(cid:0)θg−1 (V )(cid:1) = U ∩ V = ∅. Ignoring Γ3 for the time being we put x0 ∈ Wg ⊆ Xg−1 , and Wg ∩ θg(Wg) = ∅. W = \g∈Γ1∪Γ2 Wg, and observe that W ∩ Xg−1 = ∅, ∀ g ∈ Γ1, and x0 ∈ W ⊆ Xg−1 , and W ∩ θg(W ) = ∅, ∀ g ∈ Γ2. We then have that 1W b1W =Xg∈Γ 1W (fg∆g)1W =Xg∈Γ =Xg∈Γ 1W fg1θg(W ∩Xg−1 )∆g =Xg∈Γ 1W fgαg(1W 1Xg−1 )∆g =Xg∈Γ 1W fgαg(1W ∩Xg−1 )∆g = fg1W ∩θg(W ∩Xg−1 )∆g. For g ∈ Γ1 ∪ Γ2 we have that W ∩ θg(W ∩ Xg−1 ) = ∅, so the summand corresponding to g in the above sum vanishes. Therefore, 1W b1W = Xg∈Γ3 fg1W ∩θg(W ∩Xg−1 )∆g = Xg∈Γ3 f ′ g∆g, where the f ′ because Γ3 = Γ ∩ H, and moreover b′ ∈ J. Recalling that c = F (b), we also have that g are defined by the last equality above. Setting b′ = 1W b1W , we then have that b′ ∈ Lc(X) ⋊ H, F (b′) = F (1W b1W ) (4.2) = 1W (x0)F (b)1W (x0) = F (b) = c. Replacing b by b′ we have therefore proven our claim that b may be chosen in Lc(X) ⋊ H, so we are allowed to write fh∆h. b = Xh∈H For each h in H, choose a compact-open set Vh ⊆ Xh such that x0 ∈ Vh, and such that fh is constant on Vh. Letting V be the intersection of the finitely many Vh for which fh is nonzero, we have that the fh are constant on V , so that 1V fh = dh1V , where dh is the constant value attained by fh on V. We may then define observing that, as above, b′′ ∈ J, and F (b′′) = c. The latter may be expressed as b′′ := 1V b = Xh∈H 1V fh∆h = Xh∈H dh1V ∆h, Xh∈H chδh = F(cid:0) Xh∈H dh1V ∆h(cid:1) = Xh∈H dh1V (x0)δh = Xh∈H dhδh. It follows that dh = ch, for all h, whence cV = Xh∈H ch1V ∆h = Xh∈H dh1V ∆h = b′′ ∈ J. (cid:3) the ideal structure of algebraic partial crossed products 15 We may now employ (4.9) to give a characterization of admissible ideals. 4.10. Proposition. An ideal I E KH is admissible if and only if, for every c = Ph∈H chδh in I, there exists a neighborhood V of x0, such that δk−1(cid:16) Xh∈H∩kHl−1 chδh(cid:17)δl ∈ I, for all k and l in S, such that θk(x0) ∈ V . Proof. Supposing that I is admissible, pick c = Ph∈H chδh ∈ I. By hypothesis c is in F(cid:0)Ind(I)(cid:1), so (4.9) provides a compact-open set V ∋ x0, such that cV := Xh∈H ch1V ∆h ∈ Ind(I). In view of the definition of Ind(I), one has that hδk, cV δli ∈ I, for every k and l in S, and if we use (3.13) under the hypothesis that θk(x0) ∈ V , we deduce that I ∋ hδk, cV δli = Xh∈H∩kHl−1 ch1V(cid:0)θk(x0)(cid:1) δk−1hl = Xh∈H∩kHl−1 proving the condition displayed in the statement. ch δk−1hl, Conversely, assuming that this condition holds, let us show that I is admissible, namely that I ⊆ F (Ind(I)), since the reverse inclusion is granted by (4.5.ii). For this, given c = Ph∈H chδh ∈ I, pick V as in the statement. By shrinking V a bit, if necessary, we may assume that V is compact-open and V ⊆ Xh, whenever ch 6= 0, so that is a legitimate element of Lc(X) ⋊ G. We then claim that cV ∈ Ind(I). To prove this it is enough to verify that hδk, cV δli ∈ I, for all k and l in S. Using (3.13) again we have cV := Xh∈H ch1V ∆h hδk, cV δli = Xh∈H∩kHl−1 ch1V(cid:0)θk(x0)(cid:1) δk−1hl = [θk(x0)∈V ] δk−1(cid:16) Xh∈H∩kHl−1 ch δh(cid:17)δl. In case θk(x0) ∈ V , the hypothesis implies that the above belongs to I, and otherwise hδk, cV δli vanishes so it also lies in I. This shows that cV is in Ind(I), so c = F (cV ) ∈ F (Ind(I)), concluding the proof. (cid:3) Recall from (4.5.ii) that, for every ideal I E KH, one has that F (Ind(I)) ⊆ I. One may similarly inquire about the relationship between J and Ind(F (J)). The answer is given in our next: 4.11. Proposition. (i) For every ideal J E Lc(X) ⋊ G, one has that J ⊆ Ind(F (J)). (ii) For every ideal I E KH, one has that Ind(I) is the largest among the ideals J E Lc(X) ⋊ G satisfying F (J) ⊆ I. Proof. (i) Given b in J, notice that for every u and v in Lc(X) ⋊ G, one has that ubv ∈ J, so F (ubv) ∈ F (J). We then deduce from (3.22) that b lies in Ind(F (J)). This proves (i). (ii) As already mentioned, (4.5.ii) gives F (Ind(I)) ⊆ I, so Ind(I) is indeed among the ideals mentioned above. Next, given any ideal J E Lc(X) ⋊ G, with F (J) ⊆ I, we have (i) ⊆ Ind(F (J)) ⊆ Ind(I). J (cid:3) 16 m. dokuchaev and r. exel Our main interest is to construct ideals in Lc(X) ⋊ G from admissible ideals in KH, but it is interesting to remark that one may also go the other way: 4.12. Proposition. Let J E Lc(X) ⋊ G be any ideal. Then F (J) is an admissible ideal of KH. Proof. By (4.11.i) we have that J ⊆ Ind(F (J)). So, if we apply F on both sides of this inclusion, we get F (J) ⊆ F(cid:0)Ind(F (J))(cid:1) (4.5.ii) ⊆ F (J), so we see that F(cid:0)Ind(F (J))(cid:1) = F (J), which is to say that F (J) is admissible. (cid:3) 5. Representations. In this section we will begin the preparations for As before we adopt our standing assumptions (2.1). proving that any ideal (always meaning two-sided ideal) of Lc(X) ⋊ G is the intersection of ideals induced from isotropy subgroups. Our methods will largely rely on representation theory, so we begin by spelling out a trivial connection between representations and ideals. 5.1. Proposition. Let B be a K-algebra possessing local units 7 and let us be given an ideal J E B. Then there exists a vector space V , and a non-degenerate8 representation π : B → L(V ), such that J = Ker(π). Proof. Let V = B/J, denote the quotient map by q : B → V , and consider the representation π : B → L(V ) given by π(b)q(v) = q(bv), ∀ b, v ∈ B. It is then obvious that J ⊆ Ker(π), but the reverse inclusion may also be verified: in fact, if b is in Ker(π), choose e in B such that b = be, so 0 = π(b)q(e) = q(be) = q(b), whence b is in J. In order to show that π is non-degenerate, pick any ξ in V , and write ξ = q(b), for some b in B. Letting e be such that b = eb, we have ξ = q(b) = q(eb) = π(e)q(b) ∈ π(B)V. (cid:3) To see that the above result applies to our situation, we give the following: 5.2. Proposition. For every b in Lc(X) ⋊ G, there is an idempotent e ∈ Lc(X), such that eb = b = be. In particular Lc(X) ⋊ G has local units. Proof. Given b in Lc(X) ⋊ G, write fg∆g, b = Xg∈GΓ where Γ is a finite subset of G, and each fg lies in Lc(Xg). For each g in Γ, let Cg = supp(fg), so that Cg is a compact-open subset of X, contained in Xg. Put C = [g∈Γ(cid:16)Cg ∪ θg−1 (Cg)(cid:17). 7 Recall that B is said to have local units if, for every b in B, there exists an idempotent e ∈ B, such that eb = b = be. 8 We say that π is non-degenerate if V = [π(B)V ], brackets meaning linear span. the ideal structure of algebraic partial crossed products 17 It follows that C is also a compact-open subset of X, whence its characteristic function 1C is an idempotent element of Lc(X), hence also of Lc(X) ⋊ G. We next claim that 1Cb = b1C = b. To see this we observe that, for obvious reasons, 1Cfg = fg, for any g in Γ, so it is clear that 1Cb = b. On the other hand b1C = Xg∈GΓ fg∆g1C (2.3) = Xg∈GΓ fg ¯αg(1C )∆g, (5.2.1) while, for every g in Γ, we have that fg ¯αg(1C ) = 1Cg fgα(1C 1g−1 ) = 1Cg fg1θg(C∩Xg−1 )) = fg1Cg ∩θg(C∩Xg−1 )) = fg1Cg = fg, where, in the penultimate step we have used that This proves that fg ¯αg(1C ) = fg, whence the computation in (5.2.1) gives that b1C = b. Cg = θg(cid:0)θg−1 (Cg)(cid:1) = θg(cid:0)θg−1 (Cg) ∩ Xg−1(cid:1) ⊆ θg(cid:0)C ∩ Xg−1(cid:1). (cid:3) From this point on, we will fix an arbitrary ideal J E Lc(X) ⋊ G, which in view of (5.1) and (5.2), we may assume is the kernel of a likewise fixed non-degenerate representation π : Lc(X) ⋊ G → L(V ). For the time being we will forget about the ideal J mentioned above, and we will mostly focus our attention on the representation π, even though our main long term goal is to study J. 5.3. Proposition. Regarding the above representation π, its restriction to Lc(X) is non-degenerate. Proof. Given any vector ξ in V , write n ξ = π(bi)ξi, Xi=1 n n n with bi in Lc(X) ⋊ G, and ξi in V . Using (5.2), for each i we choose an idempotent ei ∈ Lc(X) such that bi = eibi, so π(bi)ξi = ξ = Xi=1 Xi=1 π(eibi)ξi = Xi=1 π(ei)π(bi)ξi ∈(cid:2)π(cid:0)Lc(X)(cid:1)V(cid:3) . (cid:3) 5.4. Proposition. (Disintegration) There exists a partial representation u : G → L(V ), such that, for all g ∈ G, and f ∈ Lc(X), one has (i) ugπ(f ) = π(cid:0) ¯αg(f )(cid:1)ug, (ii) π(f ∆g) = π(f )ug, provided f ∈ Lc(Xg), (iii) ugug−1 π(f ) = π(f 1g) = π(f )ugug−1. Proof. Given any ξ in V , use (5.3) to write where each ϕi ∈ Lc(X), and ηi ∈ V . We then define ξ = n Xi=1 π(ϕi)ηi, ugξ = n Xi=1 π(cid:0) ¯αg(ϕi)∆g(cid:1)ηi. (5.4.1) 18 m. dokuchaev and r. exel To prove that this is well defined, suppose that ξ = 0, and let n So V is a compact-open set and 1V ϕi1g−1 = ϕi1g−1, for all i = 1, . . . , n. Therefore V = supp(ϕi) ∩ Xg−1. [i=1 1θg(V )∆g ϕi = αg(1V )¯αg(ϕi)∆g = ¯αg(1V ϕi)∆g = ¯αg(ϕi)∆g, so the right-hand-side of (5.4.1) coincides with n Xi=1 π(cid:0)1θg(V )∆g ϕi(cid:1)ηi = π(cid:0)1θg(V )∆g(cid:1) This shows that ug is well defined. n Xi=1 π(ϕi(cid:1)ηi = π(cid:0)1θg(V )∆g(cid:1)ξ = 0. In order to prove (i), consider a vector ξ ∈ V of the form ξ = π(ϕ)η, for some ϕ ∈ Lc(X), and ξη ∈ V , and observe that ugπ(f )ξ = ugπ(f )π(ϕ)η = ugπ(f ϕ)η = π(cid:0)¯αg(f ϕ)∆g(cid:1)η = = π(cid:0) ¯αg(f )(cid:1)π(cid:0) ¯αg(ϕ)∆g(cid:1)η = π(cid:0) ¯αg(f )(cid:1)ugξ. Since the set of vectors ξ of the above form spans V , we have proved (i). With the goal of proving (ii), let ξ = π(ϕ)η, as above, and notice that π(f ∆g)ξ = π(f ∆g)π(ϕ)η = π(f ∆g ϕ)η = In order to prove (iii) write f = f1f2, with f1, f2 ∈ Lc(X), and let ξ ∈ V . Then = π(cid:0)f ¯αg(ϕ)∆g(cid:1)η = π(f )π(cid:0) ¯αg(ϕ)∆g(cid:1)η = π(f )ugξ. ugug−1 π(f )ξ = ugπ(cid:0) ¯αg−1 (f )∆g−1(cid:1)ξ = ugπ(cid:0) ¯αg−1 (f1)(cid:1)π(cid:0) ¯αg−1 (f2)∆g−1(cid:1)ξ = = π(cid:0) ¯αg(¯αg−1 (f1))∆g(cid:1)π(cid:0)¯αg−1 (f2)∆g−1(cid:1)ξ = π(cid:0)f11g∆g ¯αg−1 (f2)∆g−1(cid:1)ξ = = π(cid:0)f11g ¯αg(¯αg−1 (f2))(cid:1)ξ = π(cid:0)f11gf21g(cid:1)ξ = π(cid:0)f 1g(cid:1)ξ. This proves the first identity in (iii). As for the second, let ξ = π(ϕ)η, so π(f )ugug−1ξ = π(f )ugug−1π(ϕ)η = π(f )π(ϕ1g)η = π(f ϕ1g)η = π(f 1g)π(ϕ)η = π(f 1g)ξ. This concludes the proof of (iii). We leave it for the reader to prove that u is a partial representation. (cid:3) We are now about to take a major step in our quest to understand general ideals in terms of induced ones. Observe that the very definition of induced ideals requires that a point of X be chosen in advance, so we must begin to see our representation π from the point of view of a chosen point in X, a process which will eventually lead to a discretization of π (see (5.4) below). This will be accomplished by means of the following device: for each x in X, let which is clearly an ideal in Lc(X). Consequently Zx := [π(Ix)V ] Ix = {f ∈ Lc(X) : f (x) = 0}, is invariant under Lc(X), so there is a well defined representation πx of Lc(X) on making the following diagram commute V qxy Vx Vx := V /Zx, π(f ) −−−−−−−→ V πx(f ) −−−−−−−→ Vx qx y The first indication that our localization process is bearing fruit is as follows: the ideal structure of algebraic partial crossed products 19 5.5. Proposition. Given any x in X, and any f in Lc(X), one has that πx(f )η = f (x)η, for every η in Vx. Proof. Using (5.3) it is enough to verify the statement for η of the form η = qx(cid:0)π(ϕ)ξ(cid:1), where ϕ ∈ Lc(X), and ξ ∈ V . Choose a compact-open set C containing supp(ϕ) ∪ {x}, and observe that 1Cϕ = ϕ. In addition f − f (x)1C lies in Ix, since it clearly vanishes at x. Therefore Notice that πx(f )η = πx(f )qx(cid:0)π(ϕ)ξ(cid:1) = qx(cid:0)π(f )π(ϕ)ξ(cid:1) = qx(cid:0)π(f ϕ)ξ(cid:1) = · · · (5.5.1) so f ϕ =(cid:0)f (x)1C + f − f (x)1C(cid:1)ϕ = f (x)1C ϕ +(cid:0)f − f (x)1C(cid:1)ϕ ≡ π(cid:0)f (x)ϕ(cid:1)ξ = f (x)π(ϕ)ξ, π(f ϕ)ξ (mod Zx) and we then conclude that (5.5.1) equals (mod Ix ) ≡ f (x)ϕ, · · · = qx(cid:0)f (x)π(ϕ)ξ(cid:1) = f (x)qx(cid:0)π(ϕ)ξ(cid:1) = f (x)η. Putting together the definition of πx with the result above, we get the following useful formulas: for all x ∈ X, f ∈ Lc(X), and ξ ∈ V . qx(cid:0)π(f )ξ(cid:1) = πx(f )qx(ξ) = f (x)qx(ξ), (cid:3) (5.6) Having focused on Lc(X), we momentarily lost track of the ug, but there is still time to bring them back into focus: 5.7. Proposition. If x is in Xg−1 , then: (i) ug(Zx) ⊆ Zθg(x), where u is as in (5.4), (ii) there exists a linear mapping such that µx g : Vx → Vθg (x), Proof. (i) Let ξ be a vector in Zx of the form ξ = π(ϕ)η, where ϕ ∈ Ix, and η ∈ V . Then µx g(cid:0)qx(ξ)(cid:1) = qθg (x)(ugξ), ∀ ξ ∈ V. Notice that ¯αg(ϕ) lies in Iθg (x), because ugξ = ugπ(ϕ)η (5.4.i) = π(cid:0)¯αg(ϕ)(cid:1)ugη. whence ugξ ∈ Zθg(x). ¯αg(ϕ)θg (x) = ϕ(cid:0)θg−1 (θg(x))(cid:1) = ϕ(x) = 0, (ii) Follows immediately from (i). The µx g obey the following functorial property: 5.8. Proposition. If x ∈ Xg−1 ∩ Xg−1h−1 , then the composition µx g −−−−→ Vθg (x) µθg(x) h−−−−−−→ Vθhg (x) Vx coincides with µx hg. (cid:3) 20 m. dokuchaev and r. exel Proof. We initially claim that for all g in G, if x ∈ Xg, then qx(ξ) = qx(ugug−1 ξ), ∀ ξ ∈ V. This will clearly follow should we prove that ξ − ugug−1 ξ ∈ Zx, ∀ ξ ∈ V, which we will now do. By (5.3), we may assume that ξ = π(ϕ)η, for some ϕ in Lc(X), and η in V . We then have ξ − ugug−1ξ = π(ϕ)η − ugug−1 π(ϕ)η = π(ϕ)η − π(ϕ1g)η = π(ϕ − ϕ1g)η. (5.4.iii) Observing that ϕ − ϕ1g is in Ix, we have that π(ϕ − ϕ1g)η lies in Zx, proving the claim. Addressing the statement, choose any element of Vx, say qx(ξ), for some ξ in V , and notice that µθg (x) h g(cid:0)qx(ξ)(cid:1) = µθg(x) (cid:0)µx h (cid:0)qθg(x)(ugξ)(cid:1) = qθh(θg(x))(uhugξ) = = qθhg (x)(uhuh−1uhugξ) = qθhg (x)(uhuh−1uhgξ) = · · · Since θhg(x) = θh(θg(x)) ∈ Xh, and thanks to our claim, the above equals · · · (5.8) = qθhg(x)(uhgξ) = µx hg(qxξ). (cid:3) Let us now consider the representation πx Π = Yx∈X of Lc(X) on the cartesian product Qx∈X Vx. Thus, if f ∈ Lc(X), and η = (ηx)x∈X ∈Qx∈X Vx, we have (cid:0)Π(f )η(cid:1)x = πx(f )ηx, ∀ x ∈ X. Incidentally, by (5.5) the term πx(f )ηx, above, could be replaced by f (x)ηx, if desired. Thus, Π(f ) is the block diagonal operator, acting on each Vx as the scalar multiplication by f (x). Also, for each g in G, consider the linear operator Ug on Qx∈X Vx, given by ∀ η = (ηx)x∈X ∈ Yx∈X Ug(η)x = [x∈Xg] µg(cid:0)ηθg−1 (x)(cid:1), Vx. (5.9) The above occurence of µg should have actually been written as µ θg−1 (x) g , but due to the awkward nature of this notation we will rely on the context to determine the missing superscript. In what amounts to be essentially a rewording of (5.8), we have: 5.10. Proposition. Identifying Vx as a subspace of Qx∈X Vx, in the natural way, we have: (i) if x /∈ Xg−1 , then Ug vanishes on Vx, (ii) if x ∈ Xg−1 , then Ug coincides with µx (iii) if x ∈ Xg−1 , then Ug maps Vx bijectively onto Vθg (x), (iv) if x ∈ Xg−1 ∩ Xg−1h−1, then the composition g , and hence maps Vx into Vθg (x), Ug −−−−→ Vθg (x) Vx Uh−−−−→ Vθhg (x) coincides with Uhg on Vx. the ideal structure of algebraic partial crossed products 21 Proof. (i) and (ii) follow easily by inspection, while (iv) follows directly from (5.8). In order to prove that Ug is bijective from Vx to Vθg(x), it is enough to observe that, by (iv), the (cid:3) restriction of Ug−1 to Vθg (x) is the inverse of Ug. 5.11. Proposition. For every g in G, and every f ∈ Lc(X), one has that Proof. Given η = (ηx)x∈X ∈ V , one has for every x in X that UgΠ(f ) = Π(cid:0)¯αg(f )(cid:1)Ug. (cid:0)UgΠ(f )η(cid:1)x = [x∈Xg] µg(cid:16)(cid:0)Π(f )η(cid:1)θg−1 (x)(cid:17) = [x∈Xg] µg(cid:16)f(cid:0)θg−1 (x)(cid:1)ηθg−1 (x)(cid:17) = = [x∈Xg] f(cid:0)θg−1 (x)(cid:1)µg(cid:0)ηθg−1 (x)(cid:1) = ¯αg(f )x(Ugη)x =(cid:16)Π(cid:0) ¯αg(f )(cid:1)Ugη(cid:17)x . This concludes the proof. (cid:3) As a consequence, there exists a representation Π × U of Lc(X) ⋊ G on Qx∈X Vx, such that (Π × U )(f ∆g) = Π(f )Ug, ∀ f ∈ Lc(Xg−1 ). 5.12. Definition. The representation Π × U above will be referred to as the discretization of the initially given representation π. 5.13. Proposition. The mapping Q : ξ ∈ V 7→(cid:0)qx(ξ)(cid:1)x∈X ∈ Yx∈X Vx, is injective and covariant relative to the corresponding representations of Lc(X) ⋊ G on V and onQx∈X Vx, respectively. Proof. Let g ∈ G, and f ∈ Lc(Xg). Then, for every ξ in V , and every x ∈ X, we have (cid:0)(Π × U )(f ∆g)Q(ξ)(cid:1)x =(cid:0)Π(f )UgQ(ξ)(cid:1)x = f (x)(cid:0)UgQ(ξ)(cid:1)x = f (x) [x∈Xg] µg(cid:0)Q(ξ)θg−1 (x)(cid:1) = = f (x)µg(cid:0)qθg−1 (x)(ξ)(cid:1) = f (x)qx(ugξ(cid:1) = qx(cid:0)π(f )ugξ(cid:1) = Q(cid:0)π(f ∆g)ξ(cid:1)x. This proves that Q is covariant. In order to prove that Q is injective, suppose that Q(ξ) = 0, for a given ξ in V . We then claim that, for every x in X, there exists a compact-open neighborhood C of x, such that To see this, fixing x in X, recall that qx(ξ) = 0, by hypothesis, so ξ lies in Zx and hence we may write π(1Cx)ξ = 0. (5.13.1) ξ = n Xi=1 π(fi)ξi, where the fi are in Ix, and hence vanish on x. From the fact that the fi are locally constant, and finitely many, it follows that there exists a compact-open neighborhood Cx of x, where all of the fi vanish. Consequently 1Cxfi = 0, so n proving the claim. π(1Cx)ξ = Xi=1 π(1Cxfi)ξi = 0, 22 m. dokuchaev and r. exel Using (5.3), or recycling any one of the above decompositions of ξ, let us again write with fi ∈ Lc(X), and ξi ∈ V . Let so D is a compact-open subset of X, and we have D = ξ = n Xi=1 π(fi)ξi, supp(fi), n [i=1 n n ξ = π(1Dfi)ξi = π(1D) π(fi)ξi = π(1D)ξ. (5.13.2) Xi=1 Xi=1 Regarding the open cover {Cx}x∈X of D, where the Cx are as in the first part of this proof, we may find a finite set {x1, . . . , xp} ⊆ X, such that D ⊆Sp [i=1 Ek = D ∩ Cxk \ k−1 i=1 Cxi . Putting Cxi , ∀ k = 1, . . . , p, it is easy to see that the Ek are pairwise disjoint compact-open sets, whose union coincides with D. Observing that Ek ⊆ Cxk , we then have (5.13.2) = π(1D)ξ = ξ This proves that Q is injective. π(1Ek )ξ = p Xk=1 p Xk=1 π(1Ek 1Cxk )ξ = p Xk=1 π(1Ek )π(1Cxk )ξ (5.13.1) = 0. (cid:3) As an immediate consequence we have 5.14. Corollary. The null space of Π × U is contained in the null space of π. Proof. By (5.13) we see that π is equivalent to a subrepresentation of Π × U , so the conclusion follows. (cid:3) From now on we will consider the subspace consisting of the vectors with finitely many nonzero coordinates. It is easy to see that this subspace is invariant under Π(f ), for all f in Lc(X), as well as under Ug, for all g in G, consequently it is also invariant under Π × U . Mx∈X Vx ⊆ Yx∈X Vx, cides with the null space of Π × U itself. 5.15. Proposition. The null space of the representation obtained by restricting Π × U to Lx∈X Vx coin- Proof. Given b ∈ Lc(X) ⋊ G, we must show that if (Π × U )(b) vanishes on Lx∈X Vx, then it vanishes everywhere. Writing fg∆g, b = Xg∈G we have for every η = (ηx)x∈X in Qx∈X Vx, and for every x ∈ X, that (cid:0)(Π × U )(b)η(cid:1)x = Xg∈G(cid:0)Π(fg)Ugη(cid:1)x = Xg∈G fg(x) [x∈Xg] µg(cid:0)ηθg−1 (x)(cid:1). From this we see that the xth coordinate of (Π × U )(b)η depends only on the coordinates ηy, for y of the form y = θg−1 (x), where g is such that fg 6= 0, and x ∈ Xg. What matters to us is that the set of y's mentioned above is finite, so if η′ is defined to have the same y-coordinates as η, for y on the above finite set, and zero elsewhere, then η′ lies in Lx∈X Vx, and Since η and x are arbitrary we deduce that (Π × U )(b) = 0, concluding the proof. (cid:3) (cid:0)(Π × U )(b)η(cid:1)x =(cid:0)(Π × U )(b)η′(cid:1)x = 0. the ideal structure of algebraic partial crossed products 23 important aspects. Initially, regarding the space where it acts, we will identify each Vx as a subspace of As we turn out attention to the restriction of Π × U to Lx∈X Vx, it is useful to analyze some of its Lx∈X Vx, in the usual way. Thus, given ξ in V , we will think of qx(ξ) as the element of Lx∈X Vx whose coordinates all vanish, except for the xth coordinate which takes on the value qx(ξ). Once this is agreed upon, one may easily show that Π(f )qx(ξ) = πx(f )qx(ξ) = qx(cid:0)π(f )ξ(cid:1), for all f ∈ Lc(X), g ∈ G, x ∈ X, and ξ ∈ V . Ug(cid:0)qx(ξ)(cid:1) = [x∈Xg−1 ] µx g(cid:0)qx(ξ)(cid:1) = [x∈Xg−1 ] qθg(x)(ugξ), (5.16) Since Lx∈X Vx is spanned by the union of the Vx, each of which is the range of the corresponding qx, the formulas above determine the action of the Π(f ) and of the Ug on the whole space Lx∈X Vx. Putting them together, we may give the following concrete description of the restriction of Π × U to Lx∈X Vx: 5.17. Proposition. Given b =Pg∈G fg∆g in Lc(X) ⋊ G, one has for all x in X and ξ in V , that (Π × U )(b)qx(ξ) = Xg∈G [x∈Xg−1 ] qθg(x)(cid:0)π(fg)ugξ(cid:1). Proof. The proof is now a simple direct computation: (Π × U )(b)qx(ξ) = Xg∈G Π(fg)Ug(cid:0)qx(ξ)(cid:1) = Xg∈G = Xg∈G [x∈Xg−1 ] qθg (x)(cid:0)π(fg)ugξ(cid:1). Π(fg) [x∈Xg−1 ] qθg(x)(ugξ) = (cid:3) Let us now use this to describe the matrix entries of the operator (Π × U )(b) acting on Lx∈X Vx. By this we mean that, for each x and y in X, we want an expression for the yth component of the vector obtained by applying (Π × U )(b) to any given vector in Vx, say of the form qx(ξ), where ξ ∈ V . The answer is of course the yth component of the expression given in (5.17), which is in turn given by the partial sum corresponding to the terms for which θg(x) = y. The desired expression for matrix entries therefore becomes (cid:0)(Π × U )(b)qx(ξ)(cid:1)y = Xg∈G θg (x)=y qθg(x)(cid:0)π(fg)ugξ(cid:1) = qy(cid:16) Xg∈G θg(x)=y π(fg)ugξ(cid:17). (5.18) Recall that in (5.14) and (5.15) we proved the following relations among the null spaces of π, Π × U , and the restriction of the latter to Lx∈X Vx: We will now show that equality in fact holds throughout. Ker(π) ⊇ Ker(Π × U ) = Ker(cid:0)Π × U ⊕x∈X Vx(cid:1). (5.19) 5.20. Theorem. The null space of the representation obtained by restricting Π × U to Lx∈X Vx coincides with the null space of π. Proof. An important aspect of (5.18), to be used shortly, is that since (Π × U )(b) is well defined on each Vx, then so is the right-hand-side in (5.18). Precisely speaking, if ξ and ξ′ are elements of V such that qx(ξ) = qx(ξ′), then qy(cid:16) Xg∈G θg(x)=y π(fg)ugξ(cid:17) = qy(cid:16) Xg∈G θg(x)=y π(fg)ugξ′(cid:17). (5.20.1) 24 m. dokuchaev and r. exel By (5.19), in order to prove the statement, it suffices to prove that if b is in the null space of π, then for all x and y in X. (Π × U )(b) vanishes on Lx∈X Vx, which is the same as saying that its matrix entries given by (5.18) vanish Again writing b = Pg∈G fg∆g, let Γ be the subset of G consisting of those g for which fg 6= 0, and notice that Γ decomposes as the disjoint union of the following subsets: Γ1 = {g ∈ Γ : y /∈ Xg}, Γ2 = {g ∈ Γ : y ∈ Xg, θg−1 (y) 6= x}, Γ3 = {g ∈ Γ : y ∈ Xg, θg−1 (y) = x}. From our hypothesis that π(b) = 0, we conclude that, for every η in V , one has 0 = π(b)η =Xg∈Γ π(fg∆g)η =Xg∈Γ π(fg)ugη. (5.20.2) This looks enticingly like the last part of (5.18), except of course that here we are summing over all of Γ, while only the terms corresponding to Γ3 are being considered there. In order to fix this discrepancy, notice that x is not a member of the finite set {θg−1(y) : g ∈ Γ2}, so we may choose some ϕ in Lc(X) such that ϕ(x) = 1, and ϕ(cid:0)θg−1 (y)(cid:1) = 0, for all g ∈ Γ2. Observing that qx(cid:0)π(ϕ)ξ(cid:1) (5.6) = ϕ(x)qx(ξ(cid:1) = qx(ξ(cid:1), we will later use (5.20.1) in order to replace ξ by ξ′ := π(ϕ)ξ in (5.18). Meanwhile we claim that In order to prove this, observe that qy(cid:0)π(fg)ugξ′(cid:1) = 0, ∀ g ∈ Γ1 ∪ Γ2. (5.20.3) qy(cid:0)π(fg)ugξ′(cid:1) = qy(cid:0)π(fg)ugπ(ϕ)ξ(cid:1) = qy(cid:0)π(fg ¯αg(ϕ))ugξ(cid:1) (5.6) = fg(y)¯αg(ϕ)y qy(ugξ). If g ∈ Γ1, then the fact that fg is supported on Xg implies that fg(y) = 0, so the above expression vanishes. On the other hand, if g ∈ Γ2, then so the above expression again vanishes, and (5.20.3) is proved. Combining this with (5.20.2) we then have This result will have important consequences for our study of ideals in Lc(X) ⋊ G. The method we shall adopt will be to start with any ideal J E Lc(X) ⋊ G, and then use (5.1) and (5.2) to find a representation This shows that (Π × U )(b) vanishes on Lx∈X Vx, and hence the proof is concluded. π, as above, such that Ker(π) = J. By (5.20) we may replace π by Π × U acting on Lx∈X Vx, without affecting null spaces, and it will turn out that the latter is easy enough to understand since it decomposes as a direct sum of very straightforward sub-representations, which we will now describe. (cid:3) 0 = qy(cid:16)Xg∈Γ π(fg)ugξ′(cid:17) = qy(cid:16) Xg∈Γ1 = qy(cid:16) Xg∈Γ3 π(fg)ugξ′(cid:17) (5.20.1) ¯αg(ϕ)y = ϕ(cid:0)θg−1 (y)(cid:1) = 0, π(fg)ugξ′(cid:17) + qy(cid:16) Xg∈Γ2 = qy(cid:16) Xg∈Γ3 π(fg)ugξ(cid:17) (5.18) π(fg)ugξ′(cid:17) = π(fg)ugξ′(cid:17) + qy(cid:16) Xg∈Γ3 = (cid:0)(Π × U )(b)qx(ξ)(cid:1)y. the ideal structure of algebraic partial crossed products 25 5.21. Proposition. Given any x0 in X, one has that Vx Mx∈Orb(x0) is invariant under Π × U . Proof. By (5.10.ii), this space is invariant under every Ug. It is also invariant under every Π(f ), since in fact each Vx has this property. Invariance under Π × U then follows. (cid:3) We shall now study the representation obtained by restricting Π × U to the invariant space mentioned above, so we better give it a name: 5.22. Definition. Given x0 in X, we shall denote the invariant subspace referred to in (5.21) by Wx0, while the representation of Lc(X) ⋊ G obtained by restricting Π × U to Wx0 will be denoted by ρx0. If R ⊆ X is a system of representatives for the orbit relation in X, namely if R contains exactly one point of each orbit relative to the action of G on X, then surely one has Mx∈X Vx = Mx0∈R Wx0, while the restriction of Π × U to Lx∈X Vx is equivalent to Lx0∈R ρx0. fixed an arbitrary ideal J E Lc(X) ⋊ G, which incidentally has been forgotten ever since. Before we state the main result of this section we should recall that right after the proof of (5.2) we 5.23. Theorem. Let J be an arbitrary ideal of Lc(X) ⋊ G, and let π be a non-degenerate representation of Lc(X) ⋊ G, such that J = Ker(π). Considering the representations ρx constructed above, we have Ker(ρx), J = \x∈R where R ⊆ X is any system of representatives for the orbit relation in X. Proof. The null space of π coincides with the null space of the restriction of Π×U toLx∈X Vx by (5.20). Since the latter representation is equivalent to the direct sum of the ρx, as seen above, the conclusion is evident. (cid:3) 6. The representations ρx0 . In this section we shall keep the setup of the previous section, such as the ingredients listed in (2.1), the ideal J E Lc(X) ⋊ G, and the representation π : Lc(X) ⋊ G → L(V ) fixed there. The usefulness of Theorem (5.23) in describing J is obviously proportional to the extent to which we may describe the ideals Ker(ρx0 ) mentioned there, and the good news is that the representations ρx0 are well known to us. In fact they are induced from representations of isotropy group algebras. The main goal of this section is to prove that this is indeed the case. Our next result refers to the behaviour of the operators Ug : Mx∈X Vx → Mx∈X Vx, when g lies in an isotropy group. 6.1. Proposition. Fixing x0 in X, let H be the isotropy group of x0. Then, for each h in H, one has that Vx0 is invariant under Uh. Moreover, the restriction of Uh to Vx0 is an invertible operator and the correspondence h ∈ H 7→ UhVx0 ∈ GL(Vx0) is a group representation. 26 m. dokuchaev and r. exel Proof. Follows immediately from (5.10). (cid:3) The representation of H on Vx0 referred to in the above Proposition may be integrated to a representation of KH, which in turn makes Vx0 into a left KH-module. Applying the machinery of Section (3), we may then form the induced module M ⊗ Vx0 , as in (3.6), which we may also view as a representation of Lc(X) ⋊ G on M ⊗ Vx0 . 6.2. Theorem. For each x0 in X, one has that ρx0 is equivalent to the representation induced from the left KH-module Vx0 , as described above. Proof. Recalling from (5.22) that the space of ρx0 is consider the bilinear map T : M × Vx0 → Wx0 given by Wx0 = Mx∈Orb(x0) Vx, T(cid:16)Xk∈S ckδk, ξ(cid:17) = Xk∈S ckUk(ξ). Recalling that M is a right KH-module, and viewing Vx0 as a left KH-module via the representation mentioned in (6.1), we claim that T is balanced. In fact, for every k ∈ S, h ∈ H, and ξ in Vx0 , one has T (δkδh, ξ) = T (δkh, ξ) = Ukh(ξ) (5.10) = Uk(cid:0)Uh(ξ)(cid:1) = T(cid:0)δk, Uh(ξ)(cid:1) = T (δk, δh · ξ). Therefore there exists a unique linear map τ : M ⊗ Vx0 → Wx0 , such that τ (δk ⊗ ξ) = Uk(ξ). We shall next prove that τ is an isomorphism by exhibiting an inverse for it. With this goal in mind, let R be a system of representatives of left classes for S modulo H. Thus, if x is in the orbit of x0, there exists a unique r in R such that θr(x0) = x, so that Ur−1 maps Vx onto Vx0 , by (5.10). We therefore let σx : Vx → M ⊗ Vx0 be given by σx(ξ) = δr ⊗ Ur−1(ξ), for every ξ in Vx. Putting all of the σx together, let σ : Wx0 = Mx∈Orb(x0) Vx −→ M ⊗ Vx0 be the only linear map coinciding with σx on Vx, for every x in Orb(x0). We claim that σ is the inverse of τ . To see this, let k be any element of S, and let ξ be picked at random in Vx0. Writing k = rh, with r ∈ R, and h ∈ H, set x = θk(x0) = θr(x0), so Uk(ξ) ∈ Vx. We then have σ(cid:0)τ (δk ⊗ ξ)(cid:1) = σ(cid:0)Uk(ξ)(cid:1) = δr ⊗ Ur−1(cid:0)Uk(ξ)(cid:1) = δr ⊗ Uhξ = δrh ⊗ ξ = δk ⊗ ξ. This proves that στ is the identity on M ⊗ Vx0. On the other hand, given any x in Orb(x0), and any ξ ∈ Vx, write x = θr(x0), with r ∈ R, and notice that so we see that τ σ is the identity on Wx0. τ(cid:0)σ(ξ)(cid:1) = τ(cid:0)δr ⊗ Ur−1(ξ)(cid:1) = Ur(cid:0)Ur−1(ξ)(cid:1) = ξ, Therefore τ is an isomorphism between the K-vector spaces M ⊗ Vx0 and Wx0. We will next prove that τ is covariant for the respective actions of Lc(X) ⋊ G, which amount to saying that it is linear as a map between left (Lc(X) ⋊ G)-modules. For this, given g ∈ G, and f ∈ Lc(Xg), we must prove that Given k and ξ as indicated above, the left-hand-side equals τ(cid:0)(f ∆g)δk ⊗ ξ(cid:1) = ρ(f ∆g)(cid:0)τ (δk ⊗ ξ)(cid:1), ∀ k ∈ S, ∀ ξ ∈ Vx0 . (6.2.1) τ(cid:0)(f ∆g)δk ⊗ ξ(cid:1) = [gk∈S] f(cid:0)θgk(x0)(cid:1)τ (δgk ⊗ ξ) = [gk∈S] f(cid:0)θgk(x0)(cid:1)Ugk(ξ), the ideal structure of algebraic partial crossed products 27 while the right-hand-side becomes (6.2.2) ρ(f ∆g)(cid:0)τ (δk ⊗ ξ)(cid:1) = Π(f )UgUk(ξ) = · · · Observe that Uk(ξ) is in Vθk(x0), and recall from (5.10) that Ug vanishes on Vθk(x0), unless θk(x0) ∈ Xg−1 , in which case UgUk coincides with Ugk on Vx0. So UgUk(ξ) = [θk(x0)∈Xg−1 ] Ugk(ξ). Also notice that θk(x0) ∈ Xg−1 ⇐⇒ θk(x0) ∈ Xg−1 ∩ Xk ⇐⇒ x0 ∈ θk−1 (Xg−1 ∩ Xk) = Xk−1g−1 ∩ Xk−1 ⇐⇒ ⇐⇒ x0 ∈ Xk−1g−1 ⇐⇒ gk ∈ S, where we are taking into account that x0 ∈ Xk−1 by default. It follows that the expression in (6.2.2) equals because, in the nonzero case, one has that Ugk(ξ) lies in Vθgk(x0), and Π(f ) acts there by scalar multiplication (cid:3) · · · = [gk∈S] Π(f )Ugk(ξ) = [gk∈S] f(cid:0)θgk(x0)(cid:1)Ugk(ξ), by f(cid:0)θgk(x0)(cid:1), according to (5.5). This proves (6.2.1), so τ is indeed covariant. Summarizing much that we have done so far, the following is the main result of this work: 6.3. Theorem. Let θ = ({θg}g∈G, {Xg}g∈G) be a partial action of a discrete group G on a Hausdorff, locally compact, totally disconnected topological space X, such that Xg is clopen for every g in G. Then, every ideal J E Lc(X) ⋊ G is the intersection of ideals induced from isotropy groups. Proof. Let R ⊆ X be a system of representatives for the orbit relation on X. Using (5.23) we may write J as the intersection of the null spaces of the ρx, for x in R, while (6.2) tells us that ρx is equivalent to the representation induced from a representation of the isotropy group at x. The null space of ρx is therefore induced from an ideal in the group algebra of said isotropy group by (3.11), whence the result. (cid:3) Should one want to explicitly write a given ideal J E Lc(X) ⋊ G as the intersection of induced ideals, the next result should come in handy: 6.4. Proposition. Under the assumptions of (6.3), choose a system R of representatives for the orbit relation on X. For each x in R, let Hx be the isotropy group at x, and let Fx : Lc(X) ⋊ G → KHx be as in (3.20). Then, given any ideal J E Lc(X) ⋊ G, one has that I ′ KHx, and x := Fx(J) is an admissible ideal of Ind(I ′ x). J = Tx∈R into I ′ then concluding the proof. Proof. That each I ′ x is an admissible ideal follows at once from (4.12). For each x in R, let Ix be the null space of the representation ρx referred to in the proof of (6.3), so that Observe that for each x ∈ R, one has Ind(Ix). J = Tx∈R Consequently Ind(I ′ x) ⊆ Ind(Ix), whence I ′ x = Fx(J) = Fx(cid:16) Ty∈R Ind(Iy)(cid:17) ⊆ Fx(Ind(Ix)) x) ⊆ Tx∈R Ind(Ix) = J. Ind(I ′ Tx∈R (4.5) ⊆ Ix. On the other hand, one has by (4.11) that Ind(I ′ x under Fx. Since Fx(J) = I ′ x, by definition, we have that J is among such ideals, so J ⊆ Ind(I ′ x) is the largest among the ideals of Lc(X) ⋊ G mapping x), and Ind(I ′ x), J ⊆ Tx∈R (cid:3) 28 m. dokuchaev and r. exel 7. Primitive, prime and meet-irreducible ideals. Recall that an ideal J in an algebra A is said to be primitive if it coincides with the annihilator of some irreducible module. It is called prime if, whenever K and L are ideals in A, then KL ⊆ J ⇒ (K ⊆ J) ∨ (L ⊆ J). Finally, J is said to be meet-irreducible if, for any ideals K and L in A, one has K ∩ L ⊆ J ⇒ (K ⊆ J) ∨ (L ⊆ J). It is well known that every primitive ideal is prime, and since the inclusion "KL ⊆ K ∩ L" holds for any ideals K and L, it is clear that every prime ideal is meet-irreducible. The main goal of this section is to show that the induction process preserves all of the properties mentioned above. As usual we continue working under (2.1). 7.1. Proposition. If I is a primitive ideal of KH, then Ind(I) is primitive. Proof. By hypothesis I is the annihilator of some irreducible KH-module V . Employing (3.11) we then have that Ind(I) is the annihilator of the induced module M ⊗ V , which is irreducible by (3.7). Thus Ind(I) is primitive. (cid:3) In order to deal with primeness and meet-irreducibility, we first need to prove a technical result: 7.2. Lemma. Let J and K be ideals in Lc(X) ⋊ G. Then (i) F (J) ∩ F (K) ⊆ F (J ∩ K). (ii) F (J)F (K) ⊆ F (JK). Proof. We begin by proving (i). For this, let chδh ∈ F (J) ∩ F (K), c = Xh∈Γ where Γ is a finite subset of H. Applying (4.9) twice, we obtain compact-open neighborhoods V and W of x0, such that V, W ⊆ Xh, for every h ∈ Γ, satisfying cV :=XΓ ch1V ∆h ∈ J and ch1W ∆h ∈ K. cW :=XΓ Setting Z = V ∩ W , we have that Z is another compact-open neighborhood of x0, and A similar reasoning shows that cZ also lies in K, so cZ ∈ J ∩ K. Therefore J ∋ 1ZcV =XΓ ch1Z1V ∆h =XΓ ch1Z∆h =: cZ. In order to prove (ii), let b ∈ F (J) and c ∈ F (K), and write c = F (cZ) ∈ F (J ∩ K). bhδh, and b = Xh∈Γ chδh, c = Xh∈Γ where Γ is a finite subset of H. By (4.9), there are compact-open sets V and W , such that x0 ∈ V, W ⊆ Xh, for every h ∈ Γ, satisfying bV :=XΓ bh1V ∆h ∈ J and ch1W ∆h ∈ K. cW :=XΓ Observing that bV and cW lie in Lc(X) ⋊ H, we then have that bc = F (bV )F (cW ) = ν(cid:0)E(bV )(cid:1)ν(cid:0)E(cW )(cid:1) = ν(bV )ν(cW ) = ν(bV cW ) = = ν(cid:0)E(bV cW )(cid:1) = F (bV cW ) ∈ F (JK). (cid:3) the ideal structure of algebraic partial crossed products 29 We may now prove the result announced earlier: 7.3. Theorem. Let I be an ideal in KH. If I is prime or meet-irreducible, then so is Ind(I). Proof. Let us first address meet-irreducibility, so suppose that K and L are ideals in Lc(X) ⋊ G, such that K ∩ L ⊆ Ind(I). Then F (K) ∩ F (L) (7.2) ⊆ F (K ∩ L) ⊆ F (Ind(I)) (4.5.ii) ⊆ I. Assuming that I is meet-irreducible, we have that either F (K) or F (L) is contained in I. Supposing without loss of generality that the first alternative is true, that is, F (K) ⊆ I, we then have (4.11.i) ⊆ Ind(F (K)) ⊆ Ind(I), K so Ind(I) is meet-irreducible. The proof of the result for prime ideals is obtained by going through the present proof, replacing all intersections with products. (cid:3) 8. Topologically free points. As we already hinted upon, topologically free minimal actions prevent the appearance if nontrivial induced ideals. In this section we wish to further explore this aspect. We keep enforcing (2.1). 8.1. Definition. (i) We say that θ is a topologically free partial action if, for every g in G \ {1}, the fixed point set Fg := {x ∈ Xg−1 : θg(x) = x} has empty interior. (ii) We shall say that a point x0 in X is topologically free if, for every g in G \ {1}, and every open set V , with x0 ∈ V ⊆ Xg−1 , there exists some y ∈ V ∩ Orb(x0), such that θg(y) 6= y. If x0 is not fixed by θg, then the point y referred to in (8.1.ii) may clearly be taken to be x0 itself, so the condition is automatically satisfied for such a g. In other words, this condition is only relevant for g in the isotropy group of x0. Another way to describe the notion of topologically free point is to say that there is no subset of Orb(x0) containing x0, open in the relative topology, and consisting of fixed points for a nontrivial group element g. Given the relative notion of the concept of "interior", one may find a topologically free partial action admiting a invariant subspace Y ⊆ X, such that the restriction of θ to Y is no longer topologically free. However it is clear that the notion of topologically free point is not affected by restricting the action to an invariant subset, as long as the point under consideration lies in such a subset. Still another equivalent description of topologically free points is given by the following: 8.2. Proposition. Given x0 in X, the following are equivalent: (i) x0 is topologically free, (ii) the restriction of θ to Orb(x0) (the closure of the orbit of x0) is a topologically free partial action. Proof. (i) ⇒ (ii). Since we are only concerned with Orb(x0), rather than the whole of X, we may replace the latter by the former, and hence assume that the orbit of x0 is dense in X. As already observed, this restriction does not affect condition (i). Assume by contradiction that g is a nontrivial group element whose fixed point set Fg has a nontrivial interior, so there exists a nonempty open set V ⊆ Fg. Since the orbit of x0 is assumed to be dense, there is some k in S such that θk(x0) ∈ V . It is then easy to prove that θk−1gk is the identity on the open set which contains x0. In particular θk−1gk is the identity on U ∩ Orb(x0), hence contradicting (i). U := θk−1 (Xk ∩ V ), 30 m. dokuchaev and r. exel (ii) ⇒ (i). Again by contradiction, assume that 1 6= g ∈ G, and that V is an open set with x0 ∈ V ⊆ Xg−1 , and V ∩ Orb(x0) ⊆ Fg. We then claim that V ∩ Orb(x0) ⊆ Fg, as well. To see this, let y ∈ V ∩ Orb(x0). Then y is the limit of a net {ui}i ⊆ Orb(x0), and since y ∈ V , we have that ui ∈ V for all sufficiently large i. For such i's we have ui ∈ V ∩ Orb(x0) ⊆ Fg, so θg(y) = lim i θg(ui) = lim i ui = y, proving that y ∈ Fg. The claim is therefore proven, contradicting (ii). (cid:3) As a consequence we see that two points in X having the same orbit closure are either both topologically free or both fail to satisfy this property. Topologically free points actually enjoy a slightly stronger property as described next: 8.3. Proposition. Let x0 be a topologically free point, let Γ be a finite subset of G \ {1}, and let V be an open set with x0 ∈ V ⊆ \g∈Γ Xg−1 . Then there exists some y ∈ V ∩ Orb(x0), such that θg(y) 6= y, for all g in Γ. Proof. By restricting θ to the closure of the orbit of x0 we may assume that Orb(x0) is dense in X. For each g in Γ, let Φg = Fg ∩ V = {x ∈ V : θg(x) = x}. Then clearly Φg is a closed subset (relative to V ) and by (8.2) we have that Φg has no interior (relative to X, and hence also relative to V ). Consequently Sg∈G Φg is a closed set with empty interior9, whence Φg V \ [g∈G is a nonempty open set (relative to V and hence also relative to X). Since we are assuming that the orbit of x0 is dense, we conclude that there is some y in said orbit which also lies in the above open set. This concludes the proof. (cid:3) The conflict between topological freeness and induced ideals is clearly expressed by the following: 8.4. Proposition. When x0 is topologically free, the only admissible ideals of KH are the trivial ones, namely {0} and KH, itself. Consequently the only induced ideals arising from ideals in KH are the trivial ones described in (3.15). Proof. Let I E KH be a nonzero admissible ideal. We first claim that there exists some c =Ph∈H chδh ∈ I, with c1 6= 0. To see this let d =Ph∈H dhδh be any nonzero element of I. Choose h0 in H such that dh0 6= 0, and let c = dδh−1 0 dhδhh−1 , 0 = Xh∈H so that c is also in I, and c1 = dh0 6= 0, proving the claim. Working with c, choose V as in (4.10). Letting Γ = {h ∈ H : ch 6= 0}, we may clearly assume that V ⊆ Th∈Γ Xh−1. Employing (8.3), let y be an element of the orbit of x0, belonging to V , and not fixed by any h ∈ Γ \ {1}. 9 A finite union of closed sets with empty interior always has empty interior. the ideal structure of algebraic partial crossed products 31 Writing y = θk(x0), we claim that Γ ∩ kHk−1 = {1}. To see this it is enough to observe that θk(x0) is fixed by the elements of kHk−1, while the only element of Γ having this property is the unit. By (4.10) we then conclude that I ∋ δk−1(cid:16) Xh∈Γ∩kHk−1 chδh(cid:17)δk = δk−1 (c1δ1)δk = c1δ1. This implies that I contains a nonzero multiple of the unit δ1, an invertible element, whence I = KH, concluding the proof. Regarding the last sentence in the statement, if I is any ideal in KH, then by (4.7) there exists an admissible ideal I ′ such that Ind(I) = Ind(I ′). By the first part of the proof we have that I ′ is either {0} or KH, as desired. (cid:3) 9. Regular Points. In this section, still under (2.1), we will study points possessing a property which may be considered as being in the other end of the spectrum, relative to topological freeness. 9.1. Definition. A point x0 in X is said to be strongly regular (resp. regular ) if, for every h in the isotropy group of x0, there exists an open set V with x0 ∈ V ⊆ Xh−1, and such that θh is the identity on V (resp. on V ∩ Orb(x0)). Like the notion of topological free point, the notion of regular point given above is phrased in such a way as to depend only on the action of G on the orbit of the point under consideration (seen under the relative topology). This has the advantage of being mostly an atribute of the point, rather than of the action. However, the same cannot be said of the notion of strongly regular point. In any case, it is easy to see that every strongly regular point is also regular. The following result is the partial actions version of (and it follows from) [2: Lemma 3.3.a]. 9.2. Proposition. If G is countable, then the set of strongly regular points is dense. Proof. Observe that a point x0 fails to be strongly regular precisely when it lies in the fixed point set Fg for some g in G, but it does not belong to the interior of Fg. This is obviously to say that x0 ∈ ∂Fg, meaning the boundary of Fg. So the set of points which are not strongly regular is precisely the set ∂Fg. S := [g∈G On the other hand, since Fg is closed, its boundary is a closed set with empty interior. Therefore, should G be countable, we have that S is of first category in Baire's sense, hence its complement, namely the set of strongly regular points, is dense. (cid:3) For regular points, a much simpler characterization of admissibility may be given, if compared to (4.10). This will be done based on a simpler decoding of the information that "cV ∈ Ind(I)" in the first paragraph of the proof of (4.10). In order to highlight this simplification, which will be used elsewhere later, we will isolate the technicalities involved in the next two auxiliary results. 9.3. Lemma. Let Γ be any subset of G, and let k ∈ S be such that θk(x0) is fixed by θg, for all g in Γ. Then (i) Γ ⊆ kHk−1, (ii) for every l in G such that Γ∩kHl−1 is nonempty, one has that l ∈ S, that θl(x0) = θk(x0), and moreover Γ ⊆ kHl−1. Proof. (i) Given g ∈ Γ, we have that θg(cid:0)θk(x0)(cid:1) = θk(x0), so θk−1(cid:0)θg(θk(x0))(cid:1) = x0, whence k−1gk is in H, and consequently g ∈ kHk−1. This proves (i). (ii) Pick g1 in Γ ∩ kHl−1, so that h1 := k−1g1l ∈ H. Therefore θk(x0) = θg−1 1 (θk(x0)(cid:1) = θg−1 1 (cid:0)θk(θh1(x0))(cid:1) = θg−1 1 kh1 (x0) = θl(x0). 32 m. dokuchaev and r. exel This proves that l is in S, and that θk(x0) = θl(x0). Next, picking any g ∈ Γ, notice that x0 = θk−1(cid:0)θk(x0)(cid:1) = θk−1(cid:0)θg(θk(x0))(cid:1) = θk−1(cid:0)θg(θl(x0))(cid:1) = θk−1gl(x0). Thus k−1gl ∈ H, and so g ∈ kHl−1, proving (ii). 9.4. Lemma. Let I be an ideal in KH, and let c =Pg∈Γ cgδg, be an arbitrary element of KG, where Γ is a finite subset of G. Suppose that V is a compact-open set such that V ⊆ Xg, and θg−1 coincides with the identity on V ∩ Orb(x0), for all g in Γ. Then (cid:3) cg1V ∆g cV :=Xg∈Γ lies in Ind(I) if and only if, for every k in S, such that θk(x0) ∈ V , one has that δk−1 cδk ∈ I. Proof. We begin with the "only if" part, so we assume that cV ∈ Ind(I). Given k in S, with θk(x0) ∈ V , we have I ∋ hδk, cV δki (3.13) = Xg∈Γ∩kHk−1 cg1V(cid:0)θk(x0)(cid:1) δk−1gk = δk−1(cid:16) Xg∈Γ∩kHk−1 cgδg(cid:17)δk. Observing that θk(x0) lies in V ∩ Orb(x0), we have by hypotheses that θk(x0) is fixed by θg−1, and hence also by θg, for every g in Γ. We then conclude from (9.3.i) that Γ ⊆ kHk−1, so the computation above gives δk−1 cδk ∈ I, as desired. In order to prove the "if" part, let us show that cV lies in Ind(I) by employing the criteria given in (3.14). For this we must prove that, for every k and l in S, one has that Xg∈Γ∩kHl−1 cg1V(cid:0)θk(x0)(cid:1)δk−1gl ∈ I. (9.4.1) There are two situations in which the above vanishes, in which case there is nothing to do, namely when Γ ∩ kHl−1 is the empty set, or when θk(x0) /∈ V . Ignoring these, let us assume that the opposite is true, namely that Γ ∩ kHl−1 is nonempty and that θk(x0) lies in V , which in turn implies that θk(x0) is fixed by Γ. Therefore (9.3.ii) gives θk(x0) = θl(x0), and Γ ⊆ kHl−1. We then see that the term appearing in (9.4.1) is given by Xg∈Γ cgδk−1gl = δk−1(cid:16)Xg∈Γ cgδg(cid:17)δl = δk−1 c δl = δk−1 c δkδk−1l. To see that this lies in I, notice that k−1l ∈ H, because θk(x0) = θl(x0), and moreover that δk−1 c δk is in I by hypothesis. So (9.4.1) follows from the fact that I is an ideal in KH. We then conclude that cV ∈ Ind(I), thanks to (3.14). (cid:3) The promissed simplified characterization of admissibility is given next: 9.5. Proposition. Suppose that x0 is regular. Then an ideal I E KH is admissible if and only if, for every c in I, there exists a neighborhood V of x0, such that for all k in S, such that θk(x0) ∈ V . δk−1 cδk ∈ I, the ideal structure of algebraic partial crossed products 33 Proof. We begin exactly as in the proof of (4.10): supposing that I is admissible, pick c = Ph∈Γ chδh in I, where Γ is a finite subset of H. By hypothesis c is in F(cid:0)Ind(I)(cid:1), so (4.9) provides a compact-open neighborhood V of x0, such that ch1V ∆h ∈ Ind(I). cV := Xh∈Γ Given that x0 is regular, and upon shrinking V , if necessary, we may assume that θh−1 is the identity on V ∩ Orb(x0), for every h in Γ. The conclusion then follows from (9.4). Conversely, assuming that I satisfies the condition in the statement, let us prove that I is admissible, namely that F(cid:0)Ind(I)(cid:1) ⊇ I. So, pick any c = Ph∈Γ chδh in I, where Γ is a finite subset of H. Using the hypothesis, we then choose V as in the statement, which we may clearly suppose to be compact-open. Again because x0 is regular, we may assume that θh−1 is defined and coincides with the identity on V ∩ Orb(x0), for every h in Γ. By hypothesis, and by (9.4), if follows that cV ∈ Ind(I), whence as desired. c = F (cV ) ∈ F(cid:0)Ind(I)(cid:1), (cid:3) An important, albeit trivial conclusion to be drawn from the above result is: 9.6. Corollary. If G is commutative and x0 is a regular point of X, then every ideal of KH is admissible. 10. Normal ideals. It is interesting to notice that, while the admissibility condition given in (9.5) is a combination of dynamical features (viz. "θk(x0) ∈ V ") and algebraic properties (viz. "δk−1 cδk ∈ I "), the algebraic properties alone ensure admissibility in (9.6). In this section we shall discuss other purely algebraic conditions on an ideal of KH which are enough to guarantee admissibility, regardless of any other dynamical restrictions. Given a group G and a field K, recall that the well known adjoint action of G on KG is the map given by Adg(a) = δgaδg−1 , ∀ g ∈ G, ∀ a ∈ KG. Ad : G → Aut(KG) Given any subgroup H of G, observe that KH is invariant under Ad if and only if H is a normal subgroup. Regardless of normality, we may always restrict Ad to a partial action of G on KH, as in [4: 3.2]. The main ingredients of this construction are as follows: for each g in G, we let and we let Dg = KH ∩ Adg(KH), pAdg : Dg−1 → Dg be the restriction of Adg to Dg−1 . It is well known that pAd is then a partial action (in the category of sets). 10.1. Definition. The above partial action will be called the adjoint partial action of G on KH. It is easy to see that each Dg is a subalgebra of KH, while the pAdg are algebra isomorphisms. However pAd cannot be viewed as an algebraic partial action, as defined in [4: 6.4], because the Dg are not ideals in KH, but alas, pAd is a legitimate set theoretical partial action cf. [4: 2.1]. 10.2. Definition. Let H be a subgroup of a group G, and let I be an ideal in KH. We shall say that I is normal relative to G, if I is invariant [4: 2.9] under the adjoint partial action of G on KH. 34 m. dokuchaev and r. exel Thus, to say that I is normal is to say that for every c ∈ I, and every g in G such that δgcδg−1 ∈ KH, one has that δgcδg−1 ∈ I. One should view this as the best possible effort made by the ideal I in trying to embrace all element of the above form δgcδg−1 , except of course that this is impossible in the hopeless cases when such elements are not even in KH! 10.3. Proposition. Under (2.1), let x0 be a regular point of X, and let H be its isotropy group. Then every ideal I E KH which is normal relative to G, is also admissible. Proof. We will verify the conditions of (9.5). Thus, given c in I, write c = Ph∈Γ chδh, where Γ ⊆ H is a finite set, and choose a neighborhood V of x0, such that θh is the identity map on V ∩ Orb(x0), for every h in Γ. Still focusing on (9.5), pick any k in S such that θk(x0) ∈ V . We then claim that δk−1 cδk ∈ KH. To see this, notice that, for every h in Γ, one has that θh fixes θk(x0), meaning that θh(cid:0)θk(x0)(cid:1) = θk(x0), from where we deduce that So k−1hk ∈ H, whence θk−1(cid:0)θh(θk(x0))(cid:1) = x0. chδk−1hk ∈ KH. δk−1 cδk = Xh∈Γ The invariance of I under the adjoint partial action then implies that δk−1 cδk ∈ I, concluding the verification of the conditions of (9.5), and hence proving that I is admissible. (cid:3) A source of examples of normal ideals is as follows: 10.4. Proposition. Let H be a subgroup of a group G, and let J be any ideal in KG. Then the ideal I of KH given by I = J ∩ KH is normal relative to G. Proof. If c is in I, then for every g in G, one has that δgcδg−1 ∈ J. If the latter happens to also lie in KH, then it clearly belongs to I. Therefore I is normal. (cid:3) A concrete example is the augmentation ideal IH given by IH = Ker(εH , ) where εH is the augmentation map, namely the map εH : KH → K, given by 10.5. Proposition. Let H be a subgroup of a group G. Then the augmentation ideal IH is normal relative to G. εH(cid:16) Xh∈H chδh(cid:17) = Xh∈H ch. Proof. This ideal being the intersection of KH with the augmentation ideal IG of G, the conclusion follows from (10.4). (cid:3) Incidentally, the ideal referred to in [1] is related to the ideal induced by IH . In particular we have: 10.6. Proposition. (cf. [1]) Assuming (2.1) and that Lc(X) ⋊ G is simple, one has that θ is topologically free. Proof. Suppose by contradiction that θ is not topologically free. Then there exists a nontrivial g in G whose fix point set Fg has nonempty interior. Since the regular points are dense in X, we may pick a regular point x0 in V . In particular g lies in the isotropy group H of x0, so H is a nontrivial group, whence {0} ( IH ( KH, where IH is the augmentation ideal of KH. Observe that the three ideals above are admissible by (4.8), (10.5) and (10.3), so by the uniqueness part of (4.7), we have However, since Lc(X) ⋊ G is supposed to be a simple algebra, it is impossible to find three distinct (cid:3) ideals as above. This is a contradiction, and hence the statement is proved. Ind({0}) ( Ind(IH ) ( Ind(KH) . the ideal structure of algebraic partial crossed products 35 In view of (10.3) one could ask whether conditions can be found regarding an ideal I E KH, which would ensure I to be admissible regardless of any dynamical condition, as in (10.3), but also regardless of x0 being a regular point. Except for the trivial ideals treated in (3.15), this seems to be impossible in view of (8.4), where topological freeness, an eminently dynamical condition, overrides any algebraic condition one could think of. 11. Transposition. So far we have concentrated our study on induced ideals relative to a single point x0 in X, but now we would like to conduct a comparative study. So, besides assuming (2.1), and hence having fixed a point x0, we will fix another point in X, denoted x0, and we will discuss the relationship between ideals induced relative to x0 and its peer x0. Having two points in sight, it is now crucial that we distinguish the sets H and S introduced in (3.1), depending on whether x0 or x0 is concerned. One alternative would be to employ their official notation with corresponding subscripts, such as "Hx0", "Sx0", "Hx0" and "Sx0". However we will really only consider the induction process for the two points x0 and x0 chosen above, so we will prefer to save on notation by keeping the undecorated notation when x0 is considered, and writing H and S, when we are talking about x0. The maps E, ν and F , respectively introduced in (3.17), (3.19) and (3.20), also need to be distinguished, so we will adopt the above policy of decorating everything regarding x0 with a "hat". Finally, the induction process itself needs to be distinguished, so we will write Ind( I), if inducing an ideal I E K H, relative to x0, while retaining our previous notation regarding x0. The crucial way in which the two induction processes are related may be subsumed by a correspondence between ideals in KH and ideals in K H, defined as follows: given an ideal I E KH, we may form the induced ideal Ind(I), and then we have by (4.12) that F(cid:0)Ind(I)(cid:1) is an admissible ideal in K H (relative to x0, of course). 11.1. Definition. Given an ideal I E KH, we shall let so that T is a map from the set of all ideals in KH into the set of all admissible ideals in K H. We shall refer to T (I) as the transposition of I from KH to K H. Likewise, given an ideal I E K H, its transposition from K H to KH is defined by T (I) = F(cid:0)Ind(I)(cid:1), T ( I) = F(cid:0)Ind( I)(cid:1). Since we are in the business of studying induced ideals we don't really care so much about non admissible ideals, so we will shortly restrict ourselves to transposing admissible ideals only. Nevertheless one might observe that an ideal I E KH is admissible if and only if it coincides with its own transposition from KH to itself. Even before we fully understand the transposition map, we may prove a few important facts: 11.2. Proposition. Let I E KH and I E K H be admissible ideals, then the following are equivalent (i) Ind(I) ⊆ Ind( I), (ii) T (I) ⊆ I. In addition, when the above equivalent conditions hold, and both I and I are proper ideals, then Orb(x0) ⊇ Orb(x0). Proof. (i) ⇒ (ii): We have where the last equality is a consequence of the fact that I is admissible. T (I) = F(cid:0)Ind(I)(cid:1) ⊆ F(cid:0)Ind( I)(cid:1) = I, (ii) ⇒ (i): Observing that our hypothesis reads F(cid:0)Ind(I)(cid:1) ⊆ I, recall from (4.11) that Ind( I) is the largest ideal mapping into I under F , whence (i) holds. Regarding the last sentence in the statement, we have by (3.16) that the intersection Ind(I) ∩ Lc(X) consists of all f in Lc(X) vanishing on Orb(x0). Therefore (i) implies that every such f necessarily also vanishes on Orb(x0), from where the conclusion follows. (cid:3) 36 m. dokuchaev and r. exel The fact that T (I) = I is not equivalent to I = T ( I), so our result for equality of induced ideals must mention both: 11.3. Theorem. Let I E KH and I E K H be admissible ideals, then the following are equivalent: (i) Ind(I) = Ind( I), (ii) T (I) = I, and I = T ( I), (iii) T (I) ⊆ I, and I ⊇ T ( I). Proof. (i) ⇒ (ii): We have and one similarly proves that T ( I) = I. T (I) = F(cid:0)Ind(I)(cid:1) = F(cid:0)Ind( I)(cid:1) (4.6) = I, (ii) ⇒ (iii): Obvious. (iii) ⇒ (i): Follows immediately from (11.2). In order to give a concrete description of a transposed ideal, we must bring in certain important maps between the various group algebras in sight. Initially, consider the natural projection (cid:3) and, given k and l in G, define the map P :Pg∈G cgδg ∈ KG 7→Ph∈H chδh ∈ KH, Ψk,l : c ∈ K H 7→ P (δk−1 cδl) ∈ KH. For an explicit expression, let c =Ph∈ H chδh ∈ K H, and notice that Ψk,l(c) = P(cid:16) Xh∈ H chδk−1hl(cid:17) = Xh∈ H chδk−1hl = δk−1(cid:16) Xh∈ H∩ kHl−1 chδh(cid:17)δl. k−1hl∈H 11.4. Proposition. Let I be an admissible ideal in KH. Then the transposition of I to K H is given by T (I) = [V ∋x0 \k,l∈S θk(x0)∈V Ψ−1 k,l (I), where by " V ∋ x0" we mean that V ranges in the family of all neighborhoods of x0. Proof. Let c =Ph∈ H chδh ∈ K H. Then by (4.9) one has that c lies in T (I) = F(cid:0)Ind(I)(cid:1) if and only if there exists a compact-open set V , such that x0 ∈ V ⊆ Xh, (11.4.1) whenever ch 6= 0, and cV := Xh∈ H ch1V ∆h ∈ Ind(I). Using (3.14), the above is equivalent to saying that, for every k and l in S, one has that ch1V(cid:0)θk(x0)(cid:1) δk−1hl = ch δk−1hl = I ∋ Xh∈ H∩ kHl−1 = 1V(cid:0)θk(x0)(cid:1) Xh∈ H∩ kHl−1 = 1V(cid:0)θk(x0)(cid:1) Ψk,l(c). the ideal structure of algebraic partial crossed products 37 This condition is clearly meaningless unless θk(x0) is in V , in which case it says that c ∈ Ψ−1 k,l (I). If follows that c ∈ T (I) if and only if c lies in the set whose definition is almost exactly what the statement claims T (I) to be, the only difference being the family of sets where V ranges which, in the present case, consists of all compact-open neighborhoods of x0 satisfying (11.4.1). However, if we take into account that x0 admits a fundamental system of compact-open neighborhoods, and that the correspondence is decreasing, then we see that such a difference is irrelevant. V 7→ \k,l∈S θk(x0)∈V Ψ−1 k,l (I), (11.4.2) (cid:3) An interesting consequence is that, when x0 is not in the closure of the orbit of x0, the transposition of ideals leads to a triviality: 11.5. Proposition. Suppose that x0 /∈ Orb(x0). Then, for every ideal I E KH, one has that T (I) = K H. Proof. Let V be a neighborhood of x0 such that V ∩ Orb(x0) = ∅. Then there is no k in S such that θk(x0) ∈ V , whence (11.4.2) consists of the intersection of the empty family of sets, resulting in the universe where it is considered, namely K H. (cid:3) The transposition towards strongly regular points may be described in a much simpler way: 11.6. Theorem. Assume that x0 is strongly regular, and let I be an admissible ideal in KH. Then T (I) = [V ∋x0 \k∈S θk(x0)∈V δkIδk−1 . two conditions are equivalent: Proof. Let c =Ph∈Γ chδh ∈ K H, where Γ is a finite subset of H. Then by (4.9) we have that the following (i) c ∈ T (I) = F(cid:0)Ind(I)(cid:1), (ii) there exists a compact-open set V , such that x0 ∈ V ⊆ Xh, ∀ h ∈ Γ, and Let us prove that (ii) is in turn equivalent to: ch1V ∆h ∈ Ind(I). cV := Xh∈Γ (iii) there exists a compact-open set V , satisfying all of the requirements of (ii), and morever such that θh fixes V , for every h in Γ. To see that (ii) implies (iii), use the fact that x0 is strongly regular to produce a compact-open neigh- borhood W of x0, such that θh fixes W , for every h in Γ. One then has that Ind(I) ∋ 1W cV = Xh∈Γ ch1W 1V ∆h = Xh∈Γ ch1W ∩V ∆h = cW ∩V , thus proving (iii). That (iii) implies (ii) is evident. Assuming that c ∈ T (I), and hence that (iii) holds, it follows from (9.4) that, for every k in S, with θk(x0) ∈ V , one has that δk−1 cδk ∈ I. Consequently c ∈ δkI δk−1 , which is to say that δkIδk−1 , c ∈ \k∈S θk(x0)∈V which in turn implies that c belongs to the set the statement claims T (I) to be. Conversely, if c lies in that set, there exists an open neighborhood V of x0 such that, whenever k ∈ S, and θk(x0) ∈ V , one has that c ∈ δkI δk−1 . Since x0 is strongly regular, and upon shrinking V if necessary, we may suppose that V is compact-open, and that θh fixes V , for every h in Γ. It then follows from (9.4) that cV ∈ Ind(I), namely that condition (ii) above holds, so that (i) also holds, so c ∈ T (I). This completes the proof. (cid:3) 38 m. dokuchaev and r. exel References [1] J. Brown, L. Clark, C. Farthing and A, Sims, "Simplicity of algebras associated to ´etale groupoids", Semigroup Forum, 88 (2014), 433 -- 452. [2] J. H. Brown, G. Nagy, S. Reznikoff and A. Sims, "Cartan subalgebras in C*-algebras of Hausdorff etale groupoids", preprint, arXiv:1503.03521v3 [math.OA], 2016. [3] E. G. Effros, F. Hahn, "Locally compact transformation groups and C*-algebras", Memoirs of the American Mathematical Society, no. 75, 1967. [4] R. Exel, "Partial Dynamical Systems, Fell Bundles and Applications", to be published in a forthcoming NYJM book series. Available from http://mtm.ufsc.br/∼exel/papers/pdynsysfellbun.pdf. [5] D. Gon¸calves, J. Oinert and D. Royer, "Simplicity of partial skew group rings with applications to Leavitt path algebras and topological dynamics", J. Algebra, 420 (2014), 201 -- 216. [6] E. C. Gootman and J. Rosenberg, "The structure of crossed product C*-algebras: a proof of the generalized Effros-Hahn conjecture", Invent. Math., 52 (1979), no. 3, 283 -- 298. [7] M. Ionescu and D. Williams, "The generalized Effros-Hahn conjecture for groupoids", Indiana Univ. Math. J., 58 (2009), no. 6, 2489 -- 2508. [8] J. Renault, "The ideal structure of groupoid crossed product C*-algebras. With an appendix by Georges Skandalis", J. Operator Theory, 25 (1991), no. 1, 3 -- 36. [9] Jean-Luc Sauvageot, "Ideaux primitifs de certains produits croises", Math. Ann., 231 (1977), 61 -- 76. [10] T. Fack, G. Skandalis, "Sur les repr´esentations et id´eaux de la C*-alg`ebre d'un feuilletage", J. Oper. Theory, 8 (1983), 95 -- 129. [11] A. Sims and D. Williams, "The primitive ideals of some ´etale groupoid C*-algebras", preprint, arXiv:1501.02302 [math.OA], 2015. [12] B. Steinberg, "A groupoid approach to discrete inverse semigroup algebras", Adv. Math., 223 (2) (2010), 689 -- 727. [13] B. Steinberg, "Simplicity, primitivity and semiprimitivity of ´etale groupoid algebras with applications to inverse semigroup algebras", J. Pure Appl. Algebra, 220 (2016), 1035 -- 1054. [14] S. Willard, "General topology", Addison-Wesley, 1970, xii+369 pp. [15] D. Williams, "Crossed products of C*-algebras", Mathematical Surveys and Monographs, vol. 134, American Mathematical Society, 2007.
1505.06773
2
1505
2015-07-08T08:33:51
Geometric classification of unital graph C*-algebras of real rank zero
[ "math.OA" ]
We generalize the classification result of Restorff on Cuntz-Krieger algebras to cover all unital graph C*-algebras with real rank zero, showing that Morita equivalence in this case is determined by ordered, filtered K-theory as conjectured by three of the authors. The classification result is geometric in the sense that it establishes that any Morita equivalence between C*(E) and C*(F) in this class can be realized by a sequence of moves leading from E to F in a way resembling the role of Reidemeister moves on knots. As a key technical step, we prove that the so-called Cuntz splice leaves unital graph C*-algebras invariant up to Morita equivalence. We note that we have recently found a way to generalize the results of the present paper to cover general unital graph C*-algebras. The improved methods needed render some parts of the present paper obsolete, and hence we do not intend to publish it. Instead, we will present a complete solution (drawing heavily on many of the methods presented here) in a forthcoming paper.
math.OA
math
GEOMETRIC CLASSIFICATION OF UNITAL GRAPH C∗-ALGEBRAS OF REAL RANK ZERO SØREN EILERS, GUNNAR RESTORFF, EFREN RUIZ, AND ADAM P. W. SØRENSEN Abstract. We generalize the classification result of Restorff ([Res06]) on Cuntz-Krieger algebras to cover all unital graph C ∗-algebras with real rank zero, showing that Morita equivalence in this case is determined by ordered, filtered K-theory as conjectured by three of the authors. The classification result is geometric in the sense that it establishes that any Morita equivalence between C ∗(E) and C ∗(F ) in this class can be realized by a sequence of moves leading from E to F in a way resembling the role of Reidemeister moves on knots. As a key technical step, we prove that the so-called Cuntz splice leaves unital graph C ∗-algebras invariant up to Morita equivalence. The results of this preprint will be generalized in a forthcoming paper. Contents Introduction 1. 2. Preliminaries for statement of main theorem 2.1. Graphs and their matrices 2.2. Graph C∗-algebras 2.3. Filtered K-theory 2.4. Moves on graphs 3. Main result 3.1. Strategy of proof and structure of the paper 4. Derived moves 4.1. Moves on graphs 4.2. Moves on matrices 5. Cuntz splice implies stable isomorphism 6. Notation needed for the proof 6.1. Block matrices and equivalences 6.2. K-web and induced isomorphisms 7. Standard form 8. Generalization of Boyle-Huang's lifting result 9. GLP -equivalence to SLP -equivalence 10. Generalization of Boyle's positive factorization method 10.1. Factorization: Positive case 10.2. Factorization: General case 11. Putting it all together/Proof of main theorem Acknowledgements References 2 3 3 4 4 5 7 7 8 8 9 10 19 19 20 22 26 31 37 38 41 45 46 46 Date: January 13, 2018. 2010 Mathematics Subject Classification. 46L35, 46L80 (46L55, 37B10). Key words and phrases. Graph C ∗-algebras, Geometric classification, K-theory, Flow equivalence. 1 2 SØREN EILERS, GUNNAR RESTORFF, EFREN RUIZ, AND ADAM P. W. SØRENSEN 1. Introduction Ever since the inception of graph C∗-algebras, it has been a key ambition to classify these objects by their K-theory, either up to isomorphism or stable iso- morphism. With the simple case resolved by appeal to the celebrated classification results of Elliott on one hand and Kirchberg and Phillips on the other, the focus has been on the nonsimple C∗-algebras, and in fact this endeavour has evolved in parallel with the gradual realization of what invariants may prove to be complete in the case when the number of ideals is finite and the C∗-algebras in question are not stably finite. In this sense, the fundamental results obtained on the classification of certain classes of graph C∗-algebras are playing a role parallel to the one played by Rørdam's classification of simple Cuntz-Krieger algebras as a catalyst for the Kirchberg-Phillips classification mentioned above. The first two results on the classification problem for nonsimple graph C∗-alge- bras were obtained by Rørdam in [Rør97] and by Restorff in [Res06] by very different methods. Rørdam showed the importance of involving the full data contained in six-term exact sequences of the C∗-algebras given and proved a very complete classification theorem while restricting the ideal lattice to be as small as possible: only one nontrivial ideal. In Restorff's work, the ideal lattice was arbitrary among the finite ideal lattices, but as his method was to reduce the problem to classification of shifts of finite type and appeal to deep results by Boyle and Huang from symbolic dynamics ([Boy02], [BH03]), only graph C∗-algebras in the Cuntz-Krieger class were covered. Subsequent progress has mainly followed the approach in [Rør97] (see [BK11], [ET10], [ERR13b]), and hence applies only to restricted kinds of ideal lattices but with few further restrictions on the nature of the underlying graphs. The case of purely infinite graph C∗-algebras with finitely many ideals has been resolved (very interestingly, by a different invariant than what was proposed in [ERR10]) in recent work by Bentmann and Meyer ([BM14]), but as summarized in [ERR13a] there is not at present sufficient technology to take this approach much farther in the mixed cases than to C∗-algebras with three or four primitive ideals. In the paper at hand we complete the stable classification of unital graph C∗-al- gebras with real rank zero, following the strategy from [Res06] as generalized by the authors in various constellations over a period of 5 years ([Sør13], [ERR10], [ERS12], [ERS15]). Our method of proof, a substantial elaboration of key ideas from the authors' earlier work along with key ideas from the papers of Boyle and Huang, leads to a geometric classification, allowing us to conclude from Morita equivalence between a pair of graph C∗-algebras C∗(E) and C∗(F ) that a sequence of basic moves on the graphs may lead from E to F in a way resembling the role of Reidemeister moves on knots. These moves are closely related to those defining flow equivalence for shift spaces, apart from the so-called Cuntz splice which has no counterpart in dynamics and also fails to preserve the canonical diagonal Abelian subalgebra of the graph C∗-alge- bras (cf. [MM14], [BCW14]). In all cases when classification has been established, invariance of the Cuntz splice follows immediately from the fact that it will not change the K-theory, and in particular it was observed in [BM14] that Cuntz splice is invariant in the class of graph C∗-algebras which are purely infinite with finitely many ideals. But since our goal is to use the Cuntz splice to establish classification results in classes outside the scope of these results, we must prove here that in the case under investigation, the Cuntz splice leaves the C∗-algebras invariant. In fact, this result covers the full case of unital graph C∗-algebras without any reference to real rank zero. GEOMETRIC CLASSIFICATION OF UNITAL GRAPH C ∗-ALGEBRAS 3 Although the real rank zero condition is often seen to bear importance in clas- sification theory ([Ell93], [Eil96], [DG97], [ARR12]) its role in our proof is of a substantially different nature than in the papers listed. Indeed, since we gave in [ERS15] an example of two finite graphs yielding Morita equivalent graph C∗-al- gebras of real rank one for which no sequence of moves suffices to lead from one to another, we require real rank zero, through its graph algebraic characterization Condition (K), to ensure that the classification result is indeed geometric in the sense of passing through moves. After posting the first version of this paper, we realized that it was possible to obtain classification by K-theory in the general unital case by exhibiting a new move which allows us to connect the two examples mentioned above, and we have recently completed the proof that this move leaves the C∗-algebra invariant up to Morita equivalence. Consequently, we will present a complete classification result in a forthcoming paper which will contain the results in the present paper as a special case. 2. Preliminaries for statement of main theorem 2.1. Graphs and their matrices. By a graph we mean a directed graph. For- mally: Definition 2.1. A graph E is a four tuple E = (E0, E1, r, s) where E0 and E1 are sets, and r and s are maps from E1 to E0. The elements of E0 are called vertices, the elements of E1 are called edges, the map r is called the range map, and the map s is called the source map. All graphs considered will be countable, i.e., there are countably many vertices and edges. We call a graph finite, if there are only finitely many vertices and edges. As usual, two graphs E1 = (E0 2 , r2, s2) are called 2 such that s2 ◦ φ1 = φ0 ◦ s1 and isomorphic if there exist bijections φi from Ei r2 ◦ φ1 = φ0 ◦ r1. We will freely identify graphs up to isomorphism. 1 , r1, s1) and E2 = (E0 1 to Ei 1 , E1 2 , E1 Definition 2.2. A loop is an edge with the same range and source. A path µ in a graph is a finite sequence µ = e1e2 · · · en of edges satisfying r(ei) = s(ei+1), for all i = 1, 2, . . . , n − 1, and we say that the length of µ is n. We extend the range and source maps to paths by letting s(µ) = s(e1) and r(µ) = r(en). Vertices in E are regarded as paths of length 0 (also called empty paths). A cycle is a nonempty path µ such that s(µ) = r(µ). A return path is a cycle µ = e1e2 · · · en such that r(ei) 6= r(µ) for i < n. For a loop, cycle or return path, we say that it is based at the source vertex of its path. We also say that a vertex supports a certain loop, cycle or return path if it is based at that vertex. Definition 2.3. A vertex v ∈ E0 in E is called regular if s−1(v) is finite and nonempty. A vertex v ∈ E0 in E is called source if r−1(v) = ∅. A vertex v ∈ E0 in E is called a sink if s−1(v) = ∅. Note that an isolated vertex is both a sink and a source. Notation 2.4. If there exists a path from vertex u to vertex v, then we write u ≥ v -- this is a preorder on the vertex set, i.e., it is reflexive and transitive, but need not be antisymmetric. It is key to our approach to graph C∗-algebras to be able to shift between a graph and its adjacency matrix. In what follows, we let N denote the set of positive integers, while N0 denotes the set of nonnegative integers. 4 SØREN EILERS, GUNNAR RESTORFF, EFREN RUIZ, AND ADAM P. W. SØRENSEN Definition 2.5. Let E = (E0, E1, r, s) be a graph. We define its adjacency matrix AE as a E0 × E0 matrix with the (u, v)'th entry being (cid:12)(cid:12)(cid:8)e ∈ E1 s(e) = u, r(e) = v(cid:9)(cid:12)(cid:12) . As we only consider countable graphs, AE will be a finite matrix or a countably infinite matrix, and it will have entries from N0 ⊔ {∞}. Let X be a set. If A is an X × X matrix with entries from N0 ⊔ {∞} we let EA be the graph with vertex set X and between two vertices x, x′ ∈ X we have A(x, x′) edges. It will be convenient for us to alter the adjacency matrix of a graph in two very specific ways, removing singular rows and subtracting the identity, so we introduce notation for this. Notation 2.6. Let E be a graph and AE its adjacency matrix. Denote by A• E the matrix obtained from AE by removing all rows corresponding to singular vertices of E. Let BE denote the matrix AE − I, and let B• E be BE with the rows corresponding to singular vertices of E removed. 2.2. Graph C∗-algebras. We follow the notation and definition for graph C∗-al- gebras in [FLR00]; this is not the convention used in Raeburn's monograph [Rae05]. Definition 2.7. Let E = (E0, E1, r, s) be a graph. The graph C∗-algebra C∗(E) is defined as the universal C∗-algebra generated by a set of mutually orthogonal projections (cid:8)pv v ∈ E0(cid:9) and a set (cid:8)se e ∈ E1(cid:9) of partial isometries satisfying the relations • s∗ • s∗ • ses∗ esf = 0 if e, f ∈ E1 and e 6= f , ese = pr(e) for all e ∈ E1, e ≤ ps(e) for all e ∈ E1, and, • pv =Pe∈s−1(v) ses∗ e for all v ∈ E0 with 0 < s−1(v) < ∞. It is clear from the definition that an isomorphism between graphs induces a canonical isomorphism between the corresponding graph C∗-algebras. Definition 2.8. Let E be a graph. We say that E satisfies Condition (K) if for all vertices v ∈ E0 in E, either there is no return path based at v or there are at least two distinct return paths based at v. Remark 2.9. The graph C∗-algebra C∗(E) is isomorphic to a Cuntz-Krieger alge- bra if and only if the graph E is finite with no sinks, see [AR12, Theorem 3.13]. If all vertices in E support two loops, then C∗(E) is purely infinite, see [HS03, Theo- rem 2.3]. In our main result, Theorem 3.1, the graphs are assumed to have finitely many vertices and to satisfy Condition (K) -- for all such graphs the associated graph C∗-algebras are separable, unital, of real rank zero [HS03, Theorem 2.5] and have finitely many ideals. 2.3. Filtered K-theory. Definition 2.10. Let A be a C∗-algebra with finitely many ideals, and let Prim A denote the primitive ideal space of A equipped with the hull-kernel topology. A subset of Prim A is called locally closed, if it is the set difference between two open subsets of Prim A. There is a canonical lattice isomorphism between the open subsets of Prim A and the (closed, two sided) ideals of A -- let us denote this correspondence with O 7→ A(O). If V ⊆ Prim A is a difference set, then V = U \ O for some open subsets O ⊆ U ⊆ Prim A. If also V = U ′ \ O′ for some other open GEOMETRIC CLASSIFICATION OF UNITAL GRAPH C ∗-ALGEBRAS 5 subsets O′ ⊆ U ′ ⊆ Prim A, then there exists a canonical isomorphism between A(U )/A(O) and A(U ′)/A(O′). Thus we can let A(V ) = A \ U /A \ O V =U\O,O⊆U⊆Prim A V =U\O,O⊆U⊆Prim A with a slight abuse of notation, since we identify A(O) with A(O)/{0} whenever O is open. Note, that all singletons of Prim A are locally closed. For each x ∈ Prim A we let Sx denote the smallest open subset that contains x, and we let Rx = Sx \ {x}, which is an open subset. Whenever we have two open subsets O ⊆ U ⊆ Prim A, we get a cyclic six term exact sequence in K-theory: K0(A(O)) / K0(A(U )) / K0(A(U \ O)) (2.1) K1(A(U \ O)) K1(A(U )) K1(A(O)). In fact, this holds even if O and U are locally closed. If A is a real rank zero algebra, then the map from K0 to K1 will be the zero map. Let I0(A) = {Rx x ∈ Prim A, Rx 6= ∅} ∪ {Sx x ∈ Prim A} ∪ {{x} x ∈ Prim A} , I1(A) = {{x} x ∈ Prim A} , and let Imm(x) denote the set {y ∈ Prim A Sy ( Sx∧ 6 ∃z ∈ Prim A : Sy ( Sz ( Sx} . The reduced filtered K-theory of A, FKR(A), consists of the families of groups (K0(A(V )))V ∈I0(A) and (K1(A(O)))O∈I1(A) together with the maps in the sequences K1(A({x})) → K0(A(Rx)) → K0(A(Sx)) → K0(A({x})) originating from the sequence (2.1), for all x ∈ Prim A with Rx 6= ∅, and the maps in the sequences K0(A(Sy)) → K0(A(Rx)) originating from the sequence (2.1), for all pairs (x, y) ∈ Prim A with y ∈ Imm(x) and Imm(x) \ {y} 6= ∅. Let also B be a C∗-algebra with finitely many ideals. An isomorphism from FKR(A) to FKR(B) consists of a homeomorphism ρ : Prim A → Prim B and fam- ilies of isomorphisms (φV : K0(A(V )) → K0(B(ρ(V ))))V ∈I0(A) (ψO : K1(A(O)) → K1(B(ρ(O))))O∈I1 (A) such that all the ladders coming from the above sequences commute. Analogously, we define the ordered reduced filtered K-theory of A, FK+ R(A), just as FKR(A) where we also consider the order on all the K0-groups -- and for an isomorphism, we demand that the isomorphisms between the K0-groups are order isomorphisms. 2.4. Moves on graphs. In this section we describe the moves on graphs used in [Sør13]. We mention that these moves have been considered by other authors, and were previously noted to preserve the Morita equivalence class of the associated graph C∗-algebra (see [BP04]). / /   O O o o o o 6 SØREN EILERS, GUNNAR RESTORFF, EFREN RUIZ, AND ADAM P. W. SØRENSEN Definition 2.11 (Move (S): Remove a regular source). Let E = (E0, E1, r, s) be a graph, and let w ∈ E0 be a source that is also a regular vertex. Let ES denote the graph (E0 S, rS, sS) defined by S, E1 S := E0 \ {w} E1 E0 S := E1 \ s−1(w) rS := rE1 S sS := sE1 S . We call ES the graph obtained by removing the source w from E, and say ES is formed by performing move (S) to E. Definition 2.12 (Move (R): Reduction at a regular vertex). Suppose that E = (E0, E1, r, s) is a graph, and let w ∈ E0 be a regular vertex with the property that s(r−1(w)) = {x}, s−1(w) = {f }, and r(f ) 6= w. Let ER denote the graph (E0 R, rR, sR) defined by R, E1 R := E0 \ {w} E0 E1 R :=(cid:0)E1 \ ({f } ∪ r−1(w))(cid:1) ∪(cid:8)ef e ∈ E1 and r(e) = w(cid:9) and rR(ef ) := r(f ) rR(e) := r(e) if e ∈ E1 \ ({f } ∪ r−1(w)) sR(e) := s(e) if e ∈ E1 \ ({f } ∪ r−1(w)) and sR(ef ) := s(e) = x. We call ER the graph obtained by reducing E at w, and say ER is a reduction of E or that ER is formed by performing move (R) to E. Definition 2.13 (Move (O): Outsplit at a non-sink). Let E = (E0, E1, r, s) be a graph, and let w ∈ E0 be vertex that is not a sink. Partition s−1(w) as a disjoint union of a finite number of nonempty sets s−1(w) = E1 ⊔ E2 ⊔ · · · ⊔ En O, E1 with the property that at most one of the Ei is infinite. Let EO denote the graph (E0 O, rO, sO) defined by E0 E1 O :=(cid:8)v1 v ∈ E0 and v 6= w(cid:9) ∪ {w1, . . . , wn} O :=(cid:8)e1 e ∈ E1 and r(e) 6= w(cid:9) ∪(cid:8)e1, . . . , en e ∈ E1 and r(e) = w(cid:9) rEO (ei) :=(r(e)1 sEO (ei) :=(s(e)1 s(e)j wi if e ∈ E1 and r(e) 6= w if e ∈ E1 and r(e) = w if e ∈ E1 and s(e) 6= w if e ∈ E1 and s(e) = w with e ∈ Ej. We call EO the graph obtained by outsplitting E at w, and say EO is formed by performing move (O) to E. Definition 2.14 (Move (I): Insplit at a regular non-source). Suppose that E = (E0, E1, r, s) is a graph, and let w ∈ E0 be a regular vertex that is not a source. Partition r−1(w) as a disjoint union of a finite number of nonempty sets Let EI denote the graph (E0 r−1(w) = E1 ⊔ E2 · · · ⊔ En. I , E1 I , rI , sI ) defined by E0 E1 I :=(cid:8)v1 v ∈ E0 and v 6= w(cid:9) ∪ {w1, . . . , wn} I :=(cid:8)e1 e ∈ E1 and s(e) 6= w(cid:9) ∪(cid:8)e1, . . . , en e ∈ E1 and s(e) = w(cid:9) if e ∈ E1 and r(e) 6= w if e ∈ E1 and r(e) = w with e ∈ Ej rEI (ei) :=(r(e)1 sEI (ei) :=(s(e)1 r(e)j wi if e ∈ E1 and s(e) 6= w if e ∈ E1 and s(e) = w. GEOMETRIC CLASSIFICATION OF UNITAL GRAPH C ∗-ALGEBRAS 7 We call EI the graph obtained by insplitting E at w, and say EI is formed by performing move (I) to E. Definition 2.15 (Move (C): Cuntz splicing at a regular vertex supporting two return paths). Let E = (E0, E1, r, s) be a graph and let v ∈ E0 be a regular vertex that supports at least two return paths. Let EC denote the graph (E0 C , rC , sC) defined by C , E1 C := E0 ⊔ {u1, u2} C := E1 ⊔ {e1, e2, f1, f2, h1, h2}, where rC and sC extend r and s, respectively, and satisfy E0 E1 sC (e1) = v, sC (e2) = u1, sC (fi) = u1, sC (hi) = u2, and rC (e1) = u1, rC (e2) = v, rC (fi) = ui, rC (hi) = ui. We call EC the graph obtained by Cuntz splicing E at v, and say EC is formed by performing move (C) to E. We also use the notation Ev,− for this graph -- even in the case where v is not regular or not supporting two return paths. We can also Cuntz splice the vertex u1 in Ev,−, and the resulting graph we denote Ev,−−. See also Notation 5.3 and Example 5.4 for illustrations of the Cuntz splice. Definition 2.16. The equivalence relation generated by the moves (O), (I), (R), (S) together with graph isomorphism is called move equivalence, and denoted ∼M . The equivalence relation generated by the moves (O), (I), (R), (S), (C) together with graph isomorphism is called move prime equivalence, and denoted ∼M ′ . The following theorem follows from [Sør13, Propositions 3.1, 3.2 and 3.3 and Theorem 3.5]. Theorem 2.17 ([Sør13]). Let E1 and E2 be graphs such that E1 ∼M E2. Then C∗(E1) ⊗ K ∼= C∗(E2) ⊗ K. We also extend the notation of move equivalence to adjacency matrices. Definition 2.18. If A, A′ are square matrices with entries in N0 ⊔ {∞} we define them to be move equivalent, and write A ∼M A′ if EA ∼M EA′. We define move prime equivalence similarly. 3. Main result Theorem 3.1. Let E1 and E2 be graphs with finitely many vertices satisfying Condition (K). Then the following are equivalent: (1) E1 ∼M ′ E2, (2) C∗(E1) ⊗ K ∼= C∗(E2) ⊗ K, and, (3) FK+ R(C∗(E2)). R(C∗(E1)) ∼= FK+ 3.1. Strategy of proof and structure of the paper. The proof of the main theorem above, Theorem 3.1, is structured as follows. Section 5 is devoted to show that the move (C) gives stable isomorphism. Thus, (1) implies (2) follows from Theorem 2.17 and Proposition 5.8 -- a variant for finite graphs is in [ERS15]. That (2) implies (3) is clear. The rest of the paper is devoted to proving that (3) implies (1). Mainly we emu- late the previous proofs that go from filtered K-theory data to stable isomorphism or flow equivalence, as in [BH03, Boy02, Res06]. A key component of those proofs is manipulation of the matrix B• E, in particular that we can perform basic row and 8 SØREN EILERS, GUNNAR RESTORFF, EFREN RUIZ, AND ADAM P. W. SØRENSEN column operations without changing stable isomorphism class or flow equivalence class, depending on context. We prove in Section 4.2 that these matrix manipu- lation are allowed. Once we understand matrix manipulations our proof that (3) implies (1) goes through 5 steps. Step 1 First we find graphs F1 and F2 in a certain standard form such that Fi ∼M ′ Ei. This standard form will ensure that the adjacency matrices B• Fi have the same size and block structure, and that they satisfy certain additional technical conditions. This will be done in Section 7. Step 2 In Section 8 we generalize a result of Boyle and Huang ([BH03]), to show R(C∗(F2)) is induced by a GLP - that the isomorphism FK+ equivalence from B• F1 to B• R(C∗(F1)) ∼= FK+ F2 . Step 3 In Section 9 we find graphs G1, G2 such that Gi ∼M ′ Fi and B• G1 and B• G2 are SLP -equivalent. Step 4 Then, in Section 10, we generalize Boyle's positive factorization result from [Boy02] to show that there exists a positive SLP -equivalence between B• G1 and B• G2 . Step 5 It now follows from the results of Section 4.2 that G1 ∼M ′ G2 and hence that E1 ∼M ′ E2. In Section 6, we introduce some notation and concepts about block matrices needed in the proof. In Section 11, we combine the results of the previous sections to prove the main theorem. 4. Derived moves 4.1. Moves on graphs. Here we introduce the derived moves from [Sør13, Section 5]. These are shown not to change the move equivalence class, but using them simplifies working with ∼M . Definition 4.1 (Collapse a regular vertex that does not support a loop). Let E = (E0, E1, r, s) be a graph and let v be a regular vertex in E that does not support a loop. Define a graph ECOL by COL = E0 \ {v}, E0 E1 COL = E1 \ (r−1(v) ∪ s−1(v)) ⊔(cid:8)[ef ] e ∈ r−1(v) and f ∈ s−1(v)(cid:9) , the range and source maps extends those of E, and satisfy rECOL ([ef ]) = r(f ) and sECOL ([ef ]) = s(e). According to [Sør13, Theorem 5.2] E ∼M ECOL -- in fact, the collapse move can be obtained using move (O) and move (R). We also introduce move (T). Definition 4.2. Let E = (E0, E1, r, s) be a graph and let α = α1α2 · · · αn be a path such that AE(s(α1), r(α1)) = ∞. Define a graph ET by E0 E1 T = E0, T = E1 ∪ {αm m ∈ N} the range and source maps extends those of E, and satisfy rET (αm) = r(α) and sET (αm) = s(α). By [Sør13, Theorem 5.4] E ∼M ET . GEOMETRIC CLASSIFICATION OF UNITAL GRAPH C ∗-ALGEBRAS 9 4.2. Moves on matrices. Let E be a graph with finitely many vertices. In this section we perform row and column additions on BE without changing move equiv- alence class of the associated graphs. Our setup is slightly different from what was considered in [Sør13, Section 7], so we redo the proofs from there in our setting. There are no substantial changes in the proof technique. Lemma 4.3. Let E = (E0, E1, r, s) be a graph with finitely many vertices. Let u, v ∈ E0 be distinct vertices. Suppose the (u, v)'th entry of BE is nonzero ( i.e., there is an edge from u to v), and that the sum of the entries in the u'th row of BE is strictly greater than 0 ( i.e., u emits at least two edges). If B′ is the matrix formed from BE by adding the u'th column into the v'th column, then AE ∼M B′ + I. Proof. Fix an edge f from u to v. Form a graph G from E by removing f but adding for each edge e ∈ r−1(u) an edge ¯e with s(¯e) = s(e) and r(¯e) = v. We claim that B′ = BG. At any entry other than the (u, v)'th entry the two matrices have the same values, since we in both cases add entries into the v'th column that are exactly equal to the number of edges in E. At the (u, v)'th entry of BG we have E (u) = BE(u, v) + BE(u, u) = B′(u, v). E (v) − 1) + s−1 E (u) ∩ r−1 (s−1 E (u) ∩ r−1 Thus to prove this lemma it suffices to show E ∼M G. Partition s−1(u) as E1 = {f } and E2 = s−1(u) \ {f }. By assumption E2 is not empty, so we can use move (O). Doing so yields a graph just as E but where u is replaced by two vertices, u1 and u2. The vertex u1 receives a copy of everything u did and it emits only one edge. That edge has range v. The vertex u2 also receives a copy of everything u did, and it emits everything u did, except f . Since u1 is regular and not the base of a loop, we can collapse it. The resulting graph is G (after we relabel u2 as u), so G ∼M E. (cid:3) We can also add columns along a path. Proposition 4.4. Let E = (E0, E1, r, s) be a graph with finitely many vertices. Suppose u, v ∈ E0 are distinct vertices with a path from u to v going through distinct vertices u = u0, u1, u2, . . . , un = v (labelled so there is an edge from ui to ui+1 for i = 0, 1, 2, . . . , n − 1). Suppose further that for each i = 0, 1, 2, . . . , n − 1 the vertex ui emits at least two edges. If B′ is the matrix formed from BE by adding the u'th column into the v'th column, then AE ∼M B′ + I. Proof. By repeated applications of Lemma 4.3, we first add the un−1'th column into the un'th column, which we can since there is an edge from un−1 to un. Then we add the un−2'th column into the un'th column, which we can since there now is an edge from un−2 to un. Continuing this way, we end up with a matrix C which is formed from BE by adding all the columns ui, for i = 0, 1, 2, . . . , n − 1, into the the un'th column. We have that AE ∼M C + I Consider the matrix D that is formed from BE by adding all the columns u0 and ui, for i = 2, 3 . . . , n − 1, into the the un'th column. Adding the u1'th column in D into the un'th column yields C. So by Lemma 4.3, which applies since in ED+I there is an edge from u1 to un, we get that D + I ∼M C + I ∼M AE. Similarly we see that D + I is move equivalent to the matrix formed from BE by adding all the columns u0 and ui, for i = 3 . . . , n − 1, into the the un'th column. Continuing to subtract columns in this fashion, we get that AE ∼M B′ + I. (cid:3) Remark 4.5. Similar to how we used Lemma 4.3 in the above proof, we can use Proposition 4.4 "backwards" to subtract columns in BE as long as the addition that undoes the subtraction would be legal. 10 SØREN EILERS, GUNNAR RESTORFF, EFREN RUIZ, AND ADAM P. W. SØRENSEN We now turn to row additions. Lemma 4.6. Let E = (E0, E1, r, s) be a graph with finitely many vertices. Let u, v ∈ E0 be distinct vertices. Suppose the (v, u)'th entry of BE is nonzero ( i.e., there is an edge from v to u), that the sum of the entries in the u'th column of BE is strictly greater than 0 ( i.e., u receives at least two edges), and that u is a regular vertex. If B′ is the matrix formed from BE by adding the u'th row into the v'th row, then AE ∼M B′ + I. Proof. Fix an edge f from v to u. Form a graph G from E by removing f but adding for each edge e ∈ s−1(u) an edge ¯e with s(¯e) = v and r(¯e) = r(e). We claim that E ∼M G. Arguing as in the proof of Lemma 4.3 we see that this is equivalent to proving AE ∼M B′ + I. Partition r−1(u) as E1 = {f } and E2 = r−1(u) \ {f }. By our assumptions on u, E2 is nonempty, and u is regular, so we can use move (I). Doing so replaces u with two new vertices, u1 and u2. The vertex u1 only receives one edge, and that edge comes from v, the vertex u2 receives the edges u received except f . Since u1 is regular and not the base of a loop of length one we can collapse it. The resulting graph is G (after we relabel u2 as u), so G ∼M E. (cid:3) Naturally we can also add rows along a path. Proposition 4.7. Let E = (E0, E1, r, s) be a graph with finitely many vertices. Suppose u, v ∈ E0 are distinct vertices with a path from v to u going through distinct vertices v = v0, v1, v2, . . . , vn = u (labelled so there is an edge from vi to vi+1 for i = 0, 1, 2, . . . , n − 1). Suppose further that for each i = 1, 2, . . . , n the vertex vi is regular and receives at least two edges. If B′ is the matrix formed from BE by adding the u'th row into the v'th row, then AE ∼M B′ + I. Proof. The proof is completely analogous to the proof of Proposition 4.4. (cid:3) Remark 4.8. We can also use Proposition 4.7 "backwards" to subtract rows in BE ( cf. Remark 4.5). 5. Cuntz splice implies stable isomorphism In this section we prove that (1) implies (2) in Theorem 3.1. We know that the moves (O), (I), (R), (S) imply stable isomorphism, cf. Theorem 2.17. What is missing is to prove that if E1 and E2 are graphs with finitely many vertices satisfying Condition (K) and E1 is the Cuntz splice of E2 on a vertex that supports at least two distinct return paths then C∗(E1) ⊗ K ∼= C∗(E2) ⊗ K, which is what we prove in Proposition 5.8. This is also an important result needed in Section 9. First we reduce to the case where we perform a Cuntz splice on a regular vertex that supports at least two loops. Proposition 5.1. Let E be a graph with finitely many vertices, and let u ∈ E0 be a vertex that supports at least two distinct return paths. Then there exists a graph F and a regular vertex v ∈ F 0 such that (1) E ∼M F , (2) Eu,− ∼M Fv,− and Eu,−− ∼M Fv,−−, (3) v supports at least two loops, and, (4) for all w ∈ F 0 with v ≥ w ≥ v we have that w supports at least one loop, there is a path from v to w through regular vertices, and there is a path from w to v through regular vertices (we say that a path e1e2 · · · en goes through regular vertices if s(ei) is regular for all i = 2, 3, . . . , n). GEOMETRIC CLASSIFICATION OF UNITAL GRAPH C ∗-ALGEBRAS 11 Proof. Let w ∈ E0 \ {u} be a regular vertex such that u ≥ w ≥ u. If w does not support a loop we can use the collapse move (Definition 4.1) to remove it. The resulting graph will be move equivalent to E and have fewer regular vertices z with u ≥ z ≥ u that do not support a loop. So by repeatedly collapsing regular vertices that do not support loops we arrive at a graph E1 such that E ∼M E1, and since the Cuntz splice has no bearing on the collapse move we also see that Eu,− ∼M (E1)u,− and Eu,−− ∼M (E1)u,−−. For each infinite emitter in w ∈ E0 1 \ {u} with u ≥ w ≥ u, we can apply move (T) to assure that there is at least one loop based at w. Call the resulting graph E2. Again we have that E ∼M E2 and again the Cuntz splice is irrelevant for our move so Eu,− ∼M (E2)u,− and Eu,−− ∼M (E2)u,−−. Thus we have now found a graph where every vertex w 6= u with w ≥ u ≥ w supports at least one loop. We will now modify E2 to get the desired paths to and from u through regular vertices. Let w ∈ E0 2 \ {u} be a vertex with u ≥ w ≥ u. Suppose every path from u to w goes through an infinite emitter and pick a path e1e2 · · · en from u to w of minimal length (in particular it does not contain any loops nor does it visit u again). Let l be the first index such that s(el) is an infinite emitter, and note that el is not a loop. Partition s−1(s(el)) into two sets, one of them {el}, and then outsplit according to this partition. After the outsplit we can collapse the vertex that emits el, since el is the only edge it emits. Notice that in the post-collapse graph, the singular vertices are the same, and all the paths that were in the graph are still present, and each vertex z 6= u with u ≥ z ≥ u still supports at least one loop. We now have an edge from s(el−1) to r(el), so we can change our path to avoid s(el). Continuing in this fashion we eventual modify E2 in such a way that there is a path from u to w using only regular vertices. Now we continue to do this for every such vertex w. Exactly the same strategy lets us assure that there is a path from w to u through regular vertices when u ≥ w ≥ u. Call the graph that emerges after all these moves E3. Since we only did outsplits and collapses on vertices in E0 2 \ {u}, we see that these moves are unaffected by the Cuntz splice. Thus we have E ∼M E3, Eu,− ∼M (E3)u,− and Eu,−− ∼M (E3)u,−−. Now we want to modify E3 such that u has at least two loops. If not, then since u supports two distinct return paths there exists some vertex w 6= u such that w ≥ u and s−1(u) ∩ r−1(w) ≥ 1. As every vertex z 6= u with u ≥ z ≥ u supports a loop, we can use Proposition 4.4 to add the w'th column of BE3 into the u'th column twice. Call the resulting matrix B′, and let E4 = EB′+I . In E4, u will support (at least) two loops and all the other properties are preserved, since w supports a loop. The column addition is also valid in (E3)u,− and (E3)u,−−, so we have E ∼M E4, Eu,− ∼M (E4)u,− and Eu,−− ∼M (E4)u,−−. We will do the proof in cases. Case 1: If u is regular, then we can end Case 1 by letting F = E4 and v = u. Case 2: u is an infinite emitter and there exists w0 ∈ E0 such that w0 ≥ u and s−1(u) ∩ r−1(w0) = ∞. Doing what we did above and using move (T) we can find a graph E5 such that (i) E ∼M E5, (ii) Eu,− ∼M (E5)u,− and Eu,−− ∼M (E5)u,−−, (iii) u supports infinitely many loops, (iv) if u ≥ w ≥ u then there are infinitely many edges from u to w, and, (v) for all w ∈ E0 5 with u ≥ w ≥ u we have that w supports at least one loop, there is a path from u to w through regular vertices, and there is a path from w to u through regular vertices. 12 SØREN EILERS, GUNNAR RESTORFF, EFREN RUIZ, AND ADAM P. W. SØRENSEN Pick two edges e1, e2 ∈ s−1(u) ∩ r−1(u), and pick for each u ≥ w ≥ u, w 6= u, one edge ew ∈ s−1(u) ∩ r−1(w). Partition s−1(u) as into two sets, one which is E1 = {e1, e2} ∪ {ew u ≥ w ≥ u, w 6= u}. Out-splitting according to this partition we get a graph F with E ∼M F . We will show that Fu1,− ∼M (E5)u,− and Fu1 ,−− ∼M (E5)u,−−. Hence putting v = u1 will complete the proof of this case. In (E5)u,− we call the two vertices in the Cuntz splice v1 and v2, and let f be the edge from u to v1. If we outsplit at u by partitioning s−1(u) into two sets, one of which is F1 = {e1, e2, f } ∪ {ew u ≥ w ≥ u, w 6= u} we get a graph F1 ∼M (E5)u,−, which is just like Fu1,−, except that in F1, there is an edge from v1 to u2, while there is no such edge in Fu1,−. But Proposition 4.4 lets us add the v2'th column in BFu1 ,− to the u2'th, to show that F1 ∼M Fu1,−. A completely analogue argument shows that Fu1,−− ∼M (E5)u,−−. Letting v = u1 finishes Case 2. we have w (cid:3) u. Case 3: u is an infinite emitter and for all w ∈ E0 4 with s−1(u) ∩ r−1(w) = ∞ We will perform an outsplit at u, by partitioning s−1(u) into two sets, one of which is E1 = {e ∈ s−1(u) r(e) ≥ u}. Similarly to Case 2, we see that the only difference between outsplitting according to this partition before or after we perform the Cuntz splice is as edge from v1 to u2 (notation as above). Hence, we see as above that if we let F be the outsplit graph coming from E4, then E4 ∼M F , (E4)u,− ∼M Fu1,− and (E4)u,−− ∼M Fu1,−−. Letting v = u1 finishes Case 3. (cid:3) We now show that performing the Cuntz splice twice is a legal move. Proposition 5.2. Let E be a graph with finitely many vertices, and let v be a vertex that supports at least two distinct return paths. Then E ∼M Ev,−−. Proof. According to Proposition 5.1, we can assume that E satisfies the conditions of that proposition -- so we assume that v is a regular vertex that supports at least two loops. Moreover, for convenience, we let n be the number of vertices in E and we label the vertices by the numbers 1, 2, . . . , n in such a way that v gets the label n. For a given matrix size N and i, j ∈ {1, 2, . . . , N }, we let E(i,j) denote the N × N matrix that is equal to the identity matrix everywhere except for the (i, j)'th entry, that is 1. If B is a N × N matrix, then E(i,j)B is the matrix obtained from B by adding j'th row into the i'th row, and BE(i,j) is the matrix obtained from B by adding i'th column into the j'th column. Using E−1 (i,j) instead will yield subtraction. In what follows we will make extensive use of Propositions 4.4 and 4.7 and Remarks 4.5 and 4.8, we feel it will only muddle the exposition if we add all the references in. Now let B2 = E(n+2,n+3)B1 and B3 = B2E−1 B3 + I. We have that (n+3,n+4). Then B1 + I ∼M B2 + I ∼M The n+4'th vertex in EB3+I does not support a loop, so it can be collapsed yielding With B4 + I ∼M B3 + I. Now we let B5 = E−1 (n+2,n+1)B4, B6 = E(n,n+3)B5, B7 = E−1 (n+2,n+1). We then have B4 + I ∼M B5 + I ∼M B6 + I ∼M B7 + I ∼M B8 + I ∼M B9 + I. We have that (n,n+1)B6, B8 = E(n+3,n+2)B7 and B9 = B8E−1 (n,n+1)E−1 GEOMETRIC CLASSIFICATION OF UNITAL GRAPH C ∗-ALGEBRAS 13 Note that BEv,−− can be written as B1 = BE · · · · · · · · · · · · 0 0 0 1 0 0 0 0 0 0 0 0  B3 = BE 0 · · · 0 · · · 0 · · · 0 · · · 0 0 0 1 0 0 0 0  B4 = BE 0 · · · 0 · · · 0 · · · 0 0 0 0 1         . 0 ... 0 0 0 0 1 0 0 ... 0 0 0 1 0 0  0 0 ... ... 0 0 1 0 1 0 0 1 1 0 0 1 0 0 ... ... 0 0 0 1 0 1 0 0     0 0 ... ... 0 0 0 0 0 1 0 0 1 1 1 0 1 0 1 −1          1 2    0 1(cid:19) (cid:18)1 1 1 2(cid:19)  1   0  0 0 ... ... 0 0 1 0 1 0 1 1  1 1 0 0 ... ... 0 0 0 0 −1 1 0 0 ... 0 1 1 0 0 1 0 0 ... ... 0 0 0 1 0 ... 0 0 0 1 0 . . In EB9+I the n + 1'th vertex does not support a loop, so it can be collapsed to yield B9 = BE 0 · · · 0 · · · 0 · · · 0 0 0 0 1 BE B10 = (cid:18)0 · · · 0 · · · 0 1 with B9 + I ∼M B10 + I. 14 SØREN EILERS, GUNNAR RESTORFF, EFREN RUIZ, AND ADAM P. W. SØRENSEN Now we look at the graph E again, and and let BE = (bij). Since the vertex v (number n) has at least two loops, we have bnn ≥ 1. Now we can insplit by partitioning r−1(v) into two sets, one with a single set consisting of a loop based at v, and the other the rest. In the resulting graph, v is split into two vertices v1 and v2, and let E′ denote the rest of the graph. The vertex v1 has the same edges in and out of E′ as v had, but it has only bnn loops. There is one edge from v1 to v2 and v2 has one loop and there are bnn edges from v2 to v1 as well as all the same edges going from v2 into E′ as originally from v. Use the inverse collapse move to add a new vertex u to the middle of the edge from v1 to v2 and call the resulting graph F . Label the vertices such that v1, u and v2 are the n'th, n + 1'st and n + 2'nd vertex, then BF is: BF =  (cid:18) 0 bn1 eB 0 bn,n−1 · · · · · · 0 ... 0 1 0 ... 0 0   bnn(cid:19) (cid:18)−1 1 0(cid:19) 0 0 ,  where eB is BE except for on the (n, n)'th entry, which is bnn − 1. Note that bnn − 1 ≥ 0, so that there is still a loop based at the n'th vertex. This is im- portant since it allows us to do the following matrix manipulations. Let C2 = BF E(n+2,n+1)E(n+2,n+1), C3 = E(n+2,n+1)C2, C4 = E−1 (n+2,n)C3, C5 = C4E(n+1,n) and C6 = C5E(n+2,n+1). We have that C1 +I ∼M C2 +I ∼M C3 +I ∼M C4 +I ∼M C5 + I ∼M C6 + I. The matrix C6 is in fact equivalent to B10 upon relabelling of the last two vertices, thus it follows, that E ∼M Ev,−−. (cid:3) We now show that Cuntz splicing once and twice yields isomorphic graph C∗- algebras. To do this, we first set up some notation. Notation 5.3. Let E∗ and E∗∗ denote the graphs: e1 •v1 f7 * •w3 e2 e3 f6 f5 E∗ = f10 •w4 f9 f8 e4 * •v2 f1 •w1 f4 •w2 f2 f3 E∗∗ = The graph E∗ is what we attach when we Cuntz splice, if we instead attach the graph E∗∗ it is like we Cuntz spliced twice. Let E = (E0, E1, rE, sE) be a graph and let u be a vertex of E. Then Eu,− can be described as follows (up to canonical isomorphism): E0 E1 u,− = E0 ⊔ E0 ∗ u,− = E1 ⊔ E1 ∗ ⊔ {d1, d2} with rEu,− E1 = rE, sEu,−E1 = sE, rEu,− E1 ∗ = rE∗ , sEu,−E1 ∗ = sE∗ , and sEu,− (d1) = u sEu,− (d2) = v1 rEu,− (d1) = v1 rEu,− (d2) = u.   *   j j   *   * * j j   * * j j   j j GEOMETRIC CLASSIFICATION OF UNITAL GRAPH C ∗-ALGEBRAS 15 Moreover, Eu,−− can be described as follows (up to canonical isomorphism): E0 E1 u,−− = E0 ⊔ E0 ∗∗ u,−− = E1 ⊔ E1 ∗∗ ⊔ {d1, d2} with rEu,−− E1 = rE, sEu,−−E1 = sE, rEu,−− E1 ∗∗ = rE∗∗ , sEu,−−E1 ∗∗ = sE∗∗, and sEu,−−(d1) = u sEu,−−(d2) = w1 rEu,−− (d1) = w1 rEu,−− (d2) = u. Example 5.4. Consider the graph Then and E = •u e1 •v1 e4 * •v2 e2 e3 Eu,− = d1 d2 •u f10 •w4 f7 * •w3 f9 f8 f1 •w1 f6 f5 f4 •w2 f2 f3 Eu,−− = d1 d2 •u By classification of simple purely infinite graph C∗-algebras, i.e., by Kirchberg- Phillips classification, the graph C∗-algebras C∗(E∗) and C∗(E∗∗) are isomorphic (this important case is actually due to Rørdam, cf. [Rør95]). To show that C∗(Eu,−) is isomorphic to C∗(Eu,−−) we would like to know that C∗(E∗) and C∗(E∗∗) are still isomorphic if we do not enforce the summation relation at v1 and w1 respectively. Proposition 5.5. The relative graph C∗-algebras (in the sense of Muhly-Tomforde [MT04]) C∗(E∗, {v2}) and C∗(E∗∗, {w2, w3, w4}) are isomorphic. Proof. Following [MT04, Definition 3.6] we define a graph (E∗){v2} = e4 * •v2 e2 e3 e′ 3 e1 •v1 e′ 1 •v′ 1 Then by [MT04, Theorem 3.7] we have that C∗(E∗, {v2}) ∼= C∗((E∗){v2}). Similarly we define a graph (E∗∗){w2,w3,w4} = f10 •w4 f7 * •w3 f9 f8 f6 f5 f ′ 6 f1 •w1 f ′ 1 •w′ 1 f4 •w2 f2 f3 f ′ 3 % % g g     *   j j H H % % g g   *   * * j j     * * j j   j j H H % % g g     *   j j w w   * ' '   * * j j     * * j j w w   j j 16 SØREN EILERS, GUNNAR RESTORFF, EFREN RUIZ, AND ADAM P. W. SØRENSEN Using [MT04, Theorem 3.7] again, we have that C∗(E∗∗, {w2, w3, w4}) is isomorphic to C∗((E∗∗){w2,w3,w4}). Both the graphs (E∗){v2} and (E∗∗){w2,w3,w4} satisfy Condition (K). Using the well developed theory of ideal structure and K-theory for graph C∗-algebras, we see that both have exactly one nontrivial ideal, that this ideal is the compact operators, and that their six-term exact sequences are Zhv′ 1i 0 / Z 0 / 0 0 Zhw′ 1i 0 / Z 0 / 0 0 Furthermore, in K0(C∗((E∗){v2})) we have [pv1 ] = −[pv′ 1 ] = [pv2], and in K0(C∗((E∗∗){w2,w3,w4})) we have [pw1 ] = −[pw′ [pw3 ] = 0 = [pw4 ]. 1 ] = [pw2], Therefore the class of the unit is −[pv′ ], respectively. It now follows from [BD96, Theorem 2] (see also [ERR13c, Corollary 4.20]) that C∗((E∗){v2}) ∼= C∗((E∗∗){w2,w3,w4}) and hence that C∗(E∗, {v2}) ∼= C∗(E∗∗, {w2, w3, w4}). ] and −[pw′ (cid:3) 1 1 We also need a technical result about the projections in E = C∗(E∗, {v2}). Lemma 5.6. Let E = C∗(E∗, {v2}) and choose an isomorphism between E and C∗(E∗∗, {w2, w3, w4}), which exists according to the previous proposition. Let pv1, pv2, se1 , se2 , se3 , se4 be the canonical generators of C∗(E∗, {v2}) = E and let pw1, pw2, pw3, pw4, sf1 , sf2, . . . , sf10 denote the image of the canonical generators of C∗(E∗∗, {w2, w3, w4}) in E under the chosen isomorphism. Then se1 s∗ f2 + sf5 s∗ f5, f2 + sf5 s∗ f1 + sf2s∗ in E, where ∼ denotes Murray-von Neumann equivalence. e2(cid:1) ∼ pw1 −(cid:0)sf1 s∗ pv1 −(cid:0)se1 s∗ e1 + se2 s∗ e1 + se2 s∗ e2 ∼ sf1 s∗ f1 + sf2s∗ f5(cid:1) , Proof. By [AMP07, Corollary 7.2], row-finite graph C∗-algebras have stable weak cancellation, so by [MT04, Theorem 3.7], E has stable weak cancellation. Hence any two projections in E are Murray-von Neumann equivalent if they generate the same ideal and have the same K-theory class. As in the proof of Proposition 5.5, we will use [MT04, Theorem 3.7] to realize our relative graph C∗-algebras as graph C∗-algebras of the graphs (E∗){v2} and (E∗∗){w2,w3,w4}. Denote the image of the vertex projections of C∗((E∗){v2}) in- side E under this isomorphism by qv1, qv2 , qv′ and denote the image of the vertex ∼= projections of (E∗∗){w2,w3,w4} inside E under the isomorphisms (E∗∗){w2,w3,w4} C∗(E∗∗, {w2, w3, w4}) ∼= E by qw1 , qw2 , qw3, qw4 , qv′ . Using the description of the isomorphism in [MT04, Theorem 3.7], we see that we need to show that qv1 ∼ qw1 and qv′ ∼ qw′ . Since (E∗)0 {v2} satisfies Condition (K) and the smallest hereditary and saturated subset containing v1 is all of (E∗)0 {v2} we have that qv1 is a full projection ([BHRS02, Theorem 4.4]). Similarly qw1 is full. In K0(E) we have, using our calculations from the proof of Proposition 5.5, that 1 1 1 1 So by weak stable cancellation qv1 ∼ qw1 . [qv1 ] = [1] = [qw1 ]. / /   O O o o o o / /   O O o o o o GEOMETRIC CLASSIFICATION OF UNITAL GRAPH C ∗-ALGEBRAS 17 1 and qw′ Both qv′ generate the only nontrivial ideal I of E ([BHRS02, Theo- rem 4.4]). Since that ideal is isomorphic to the compact operators and both [qv′ ] ] are positive generators of K0(I) ∼= K0(K) ∼= Z, they must both represent and [qw′ the same class in K0(I), and thus also in K0(E). Therefore qv′ (cid:3) 1 1 1 ∼ qw′ . 1 1 If E is a graph and we have a set of mutually orthogonal projections(cid:8)pv v ∈ E0(cid:9) and a set(cid:8)se e ∈ E1(cid:9) of partial isometries in a C∗-algebra satisfying the relations of Definition 2.7, then we call these elements a Cuntz-Krieger E-family. In a graph E, we call a cycle e1e2 · · · en a vertex-simple cycle if r(ei) 6= r(ej ) for all i 6= j. A vertex-simple cycle e1e2 · · · en is said to have an exit if there exists an edge f such that s(f ) = s(ek) for some k = 1, 2, . . . , n with ek 6= f . Note that in [Szy02], the author uses the term loop where we use cycle. Theorem 5.7. Let E be a graph with finitely many vertices and let u be a vertex of E. Then C∗(Eu,−) ∼= C∗(Eu,−−). Proof. As above, we let E denote the C∗-algebra C∗(E∗, {v2}), and we choose an isomorphism between E and C∗(E∗∗, {w2, w3, w4}), which exists according to Proposition 5.5. Since C∗(Eu,−) and E are unital, separable, nuclear C∗-algebras, it follows from Kirchberg's embedding theorem that there exists a unital embedding C∗(Eu,−) ⊕ E ֒→ O2. In O2, we denote the vertex We will suppress this embedding in our notation. projections and the partial isometries coming from C∗(Eu,−) by pv, v ∈ E0 u,− and se, e ∈ E1 u,−, respectively, and we denote the vertex projections and the partial isometries coming from E = C∗(E∗, {v2}) by p1, p2 and s1, s2, s3, s4, respectively. Since we are dealing with an embedding, it follows from Szymański's General Cuntz- Krieger Uniqueness Theorem ([Szy02, Theorem 1.2]) that for any vertex-simple cy- cle α1α2 · · · αn in Eu,− without any exit, we have that the spectrum of sα1 sα2 · · · sαn contains the entire unit circle. We will define a new Cuntz-Krieger Eu,−-family. For each vertex v ∈ E0 we let qv = pv, we let qv1 = p1 and qv2 = p2. Since any two nonzero projections in O2 are Murray-von Neumann equivalent, we can choose partial isometries x1, x2 ∈ O2 such that x1x∗ x2x∗ 1 = sd1s∗ d1 2 = p1 − (s1s∗ 1 + s2s∗ 2) x∗ 1x1 = p1 x∗ 2x2 = pu. We let td1 = x1 and td2 = x2. Finally we let te = se for e ∈ E1 and put tei = si for i = 1, 2, 3, 4. By construction {qv v ∈ E0 u,−} is a set of orthogonal projections, and {te e ∈ E1 u,−} a set of partial isometries. Furthermore, by choice of {te e 6= d1, d2} the relations are clearly satisfied at all vertices other than v1 and u. The choice of x1, x2 ensures that the relations hold at u and v1 as well. Hence {qv, te} does indeed form a Cuntz-Krieger Eu,− family. Denote this family by S. Using the universal property of graph C∗-algebras, we get a ∗-homomorphism from C∗(Eu,−) onto C∗(S) ⊆ O2. Let α1α2 · · · αn be a vertex-simple cycle in Eu,− without any exit. Since u is where the Cuntz splice is glued on, no vertex-simple cycle without any exit uses edges connected to u, v1 or v2. Hence tα1tα2 · · · tαn = sα1 sα2 · · · sαn and so its spectrum contains the entire unit circle. It now follows from Szymański's General Cuntz-Krieger Uniqueness Theorem ([Szy02, Theorem 1.2]) that C∗(Eu,−) ∼= C∗(S). 18 SØREN EILERS, GUNNAR RESTORFF, EFREN RUIZ, AND ADAM P. W. SØRENSEN Let A be the subalgebra of O2 generated by {pv v ∈ E0} and E. Note that A has a unit, and although it does not coincide with the unit of O2 it does co- incide with the unit of C∗(S). In fact A is a unital subalgebra of C∗(S). Let us denote by {rwi, yfj i = 1, 2, 3, 4, j = 1, 2, . . . , 10} the image of the canonical generators of C∗(E∗∗, {w2, w3, w4}) in O2 under the chosen isomorphism between C∗(E∗∗, {w2, w3, w4}) and E composed with the embedding into O2. By Lemma 5.6, certain projections in E are Murray-von Neumann equivalent, hence we can find a unitary z ∈ A such that zqvz∗ = qv, for all v ∈ E0, z(cid:0)te1 t∗ z(cid:0)qv1 −(cid:0)te1 t∗ e1 + te2 t∗ e1 + te2 t∗ f1 + yf2 y∗ e2(cid:1) z∗ = yf1 y∗ e2(cid:1)(cid:1) z∗ = rw1 −(cid:0)yf1y∗ Note that this implies that zqv1z∗ = rw1 . f2 + yf5y∗ f5 , f1 + yf2y∗ f2 + yf5 y∗ f5(cid:1) . We will now define a Cuntz-Krieger Eu,−−-family in O2. For v ∈ E0, we let Pv = qv, and we let Pwi = rwi , for i = 1, 2, 3, 4. For e ∈ E1 ∪ {d1, d2}, we let Se = ztez∗, and we let Sfi = yfi for i = 1, 2, . . . , 10. Denote this family by T . By construction {Pv v ∈ E0 u,−−} is a set of orthogonal projections, and {Se u,−−} a set of partial isometries. Since z is a unitary in C∗(S) and since e ∈ E1 S is a Cuntz-Krieger Eu,−-family, T will satisfy the Cuntz-Krieger relations at all vertices in E0. Similarly, we see that since {rwi , yfj i = 1, 2, 3, 4, j = 1, 2, . . . , 10} is a Cuntz-Krieger (E∗∗, {w2, w3, w4})-family, T will satisfy the relations at the vertices w2, w3, w4. It only remains to check the summation relation at w1, for that we compute XsEu,−− (e)=w1 SeS∗ e = Sf1 S∗ = yf1 y∗ f1 + Sf2S∗ f1 + yf2 y∗ e1 + te2 t∗ e1 + te2 t∗ f2 + Sf5 S∗ f2 + yf5 y∗ f5 + Sd2S∗ d2 d2z∗ f5 + ztd2t∗ d2z∗ e2(cid:1) z∗ + ztd2t∗ d2(cid:1) z∗ e2 + td2t∗ = z(cid:0)te1 t∗ = z(cid:0)te1 t∗ = zqv1z∗ = rw1 = Pw1 . Hence T is a Cuntz-Krieger Eu,−−-family. The universal property of C∗(Eu,−−) provides a surjective ∗-homomorphism from C∗(Eu,−−) to C∗(T ) ⊆ O2. Let α1α2 · · · αn be a vertex-simple cycle in Eu,−− without any exit. We see that all the edges αi must be in E1, and hence we have Sα1Sα2 · · · Sαn = ztα1z∗ztα2z∗ · · · ztαn z∗ = zsα1sα2 · · · sαn z∗ and so its spectrum contain the entire unit circle. It now follows from Szymański's General Cuntz-Krieger Uniqueness Theorem ([Szy02, Theorem 1.2]) that C∗(Eu,−−) is isomorphic to C∗(T ). Since A ⊆ C∗(S) and since {rwi, yfj i = 1, 2, 3, 4, j = 1, 2, . . . , 10} ⊆ E ⊆ C∗(S), we have that T ⊆ C∗(S). So C∗(T ) ⊆ C∗(S). But since A is also contained in C∗(T ) and E ⊆ C∗(T ), we have that S ⊆ C∗(T ), and hence C∗(S) ⊆ C∗(T ). Therefore C∗(Eu,−) ∼= C∗(S) = C∗(T ) ∼= C∗(Eu,−−). (cid:3) Thus we have the following fundamental result. Proposition 5.8. Let E be a graph with finitely many vertices, and let v be a vertex that supports at least two distinct return paths. Then C∗(E) ⊗ K ∼= C∗(Ev,−) ⊗ K. Proof. By Theorem 5.7, C∗(Ev,−) ⊗ K ∼= C∗(Ev,−−) ⊗ K). By Proposition 5.2 and Theorem 2.17, C∗(E)⊗K ∼= C∗(Ev,−−)⊗K. Thus, C∗(E)⊗K ∼= C∗(Ev,−)⊗K. (cid:3) GEOMETRIC CLASSIFICATION OF UNITAL GRAPH C ∗-ALGEBRAS 19 6. Notation needed for the proof 6.1. Block matrices and equivalences. Notation 6.1. For m, n ∈ N0, we let M(m × n, Z) denote the set of group homo- morphisms from Zn to Zm. When m, n ≥ 1, we can equivalently view this as the m × n matrices over Z, where composition of group homomorphisms corresponds to matrix multiplication -- the (zero) group homomorphisms for m = 0 or n = 0 we will also call empty matrices with zero rows or columns, respectively. For m, n ∈ N, we let M+(m × n, Z) denote the subset of M(m × n, Z), where all entries in the corresponding matrix are positive. For a m × n matrix, we will also write B > 0 whenever B ∈ M+(m × n, Z). For a m × n matrix B, where m, n ∈ N, we let B(i, j) denote the (i, j)'th entry of the corresponding matrix, i.e., the entry in the i'th row and j'th column. Definition 6.2. Let m, n ∈ N. For a m × n matrix B over Z, we let gcd B be the greatest common divisor of the entries B(i, j), for i = 1, . . . , m, j = 1, . . . , n, if B is nonzero, and zero otherwise. Assumption 6.3. Let N ∈ N. For the rest of the paper, we let P = {1, 2, . . . , N } denote a partially ordered set with order (cid:22) satisfying i (cid:22) j ⇒ i ≤ j, for all i, j ∈ P, where ≤ denotes the usual order on N. We denote the corresponding irreflexive order by ≺. Definition 6.4. Let m = (mi)N 0 be multiindices. We write m ≤ n if mi ≤ ni for all i = 1, 2, . . . , N , and in that case, we let n − m be (ni − mi)N i=1, n = (ni)N i=1 ∈ NN i=1. We let M(m × n, Z) denote the set of group homomorphisms from Zn1 ⊕ Zn2 ⊕ · · · ⊕ ZnN to Zm1 ⊕ Zm2 ⊕ · · · ⊕ ZmN , and for such a homomorphism B, we let B{i, j} denote the component of B from the j'th direct summand to the i'th direct summand. We also use the notation B{i} for B{i, i}. Using composition of homo- morphisms we get in a natural way a category MN with objects NN 0 and with the morphisms from n to m being M(m × n, Z). Moreover, (BC){i, j} = NXk=1 B{i, k}C{k, j}, whenever B ∈ M(m × n, Z) and C ∈ M(n × r, Z) for a multiindex r. A morphism B ∈ M(m × n, Z) is said to be in MP (m × n, Z), if B{i, j} 6= 0 =⇒ i (cid:22) j, for all i, j ∈ P. It is easy to verify, that this gives a subcategory MP with the same objects but MP (m × n, Z) as morphisms. Moreover, for a subset s of P, we let -- with a slight misuse of notation -- B{s} ∈ Ms((mi)i∈s × (ni)i∈s, Z) denote the component of B from Li∈s Zni to Li∈s Zmi. We let M(n, Z) denote M(n × n, Z), and MP (n, Z) denote MP (n × n, Z). For n, we let GLP (n, Z) denote the automorphisms in MP (n, Z). Then U ∈ GLP (n, Z) if and only if U ∈ MP (n, Z) and U {i} is a group automorphism (meaning that the determinant as a matrix is ±1 whenever ni 6= 0, for every i ∈ P). An automorphism U ∈ GLP (n, Z) is in SLP (n, Z) if the determinant of U {i} is 1 for all i ∈ P with ni 6= 0. 20 SØREN EILERS, GUNNAR RESTORFF, EFREN RUIZ, AND ADAM P. W. SØRENSEN Remark 6.5. Let m, n ∈ NN 0 be multiindices. Set k1 = m1 + · · · + mN and k2 = n1 + · · · + nN . If k1 6= 0 and k2 6= 0, we can equivalently view the elements B ∈ M(m × n, Z) as block matrices B = B{1, 1} . . . B{1, N } ... ... B{N, 1} . . . B{N, N }  where B{i, j} ∈ M(mi × nj, Z) with B{i, j} the empty matrix if mi = 0 or nj = 0. Note that from this point of view, the matrices in MP (m × n, Z) are upper triangular matrices with a certain zero block structure dictated by the order on P, and the matrices in GLP (n, Z) (respectively SLP (n, Z)) are matrices in MP (m × n, Z) with all nonempty diagonal blocks having determinant ±1 (respectively 1). Note that if B ∈ M(m × n, Z) and C ∈ M(n × r, Z) for a multiindex r, then the matrix product makes sense, and -- as matrices -- we have that (BC){i, j} = B{i, k}C{k, j}, NXk∈P,nk6=0 for all i, j ∈ P with mi 6= 0 and rj 6= 0. We will therefore also allow ourselves to talk about matrices with zero rows or columns (by considering it as an element of M(m × n, Z) ); and then B{s} for a subset s of P as defined above is just the principal submatrix corresponding to indices in s (remembering the block structure). Definition 6.6. Let m and n be multiindices. Two matrices B and B′ in MP (m × n, Z) are said to be GLP -equivalent (respectively SLP -equivalent ) if there exist U ∈ GLP (m, Z) and V ∈ GLP (n, Z) (respectively U ∈ SLP (m, Z) and V ∈ SLP (n, Z)) such that U BV = B′. Note that this is a generalization of the definitions in [Boy02, BH03] (in the finite matrix case) to the cases with rectangular diagonal blocks or vacuous blocks. 6.2. K-web and induced isomorphisms. We define the K-web, K(B), of a matrix B ∈ MP (m × n, Z) and describe how a GLP -equivalence (U, V ) : B → B′ induces an isomorphism κ(U,V ) : K(B) → K(B′). For an element B ∈ M(m×n, Z) (i.e., a group homomorphism B : Zn → Zm), we define as usual cok B to be the abelian group Zm/BZn and ker B to be the abelian group {x ∈ Zn Bx = 0}. Note, that if m = 0, then cok B = {0} and ker B = Zn, and if n = 0, then cok B = Zm and ker B = {0}. For m, n ∈ N0, B, B′ ∈ M(m × n, Z), U ∈ GL(m, Z) and V ∈ GL(n, Z) with U BV = B′, it is now clear that this equivalence induces isomorphisms cok B [x]7→[U x] ξ(U,V ) / cok B′ and ker B [x]7→[V −1x] δ(U,V ) / ker B′. Lemma 6.7. Let P = P2 = {1, 2} be a partially ordered set and let B ∈ MP (m × n, Z). Then the following sequence cok B{1} [v]7→[( v 0 )] / cok B [( v w )]7→[w] / cok B{2} v7→[A{1,2}v] 0 ker B{2} ker B w←[( v w ) ( v 0 )←[v ker B{1} is exact. / / / /   O O o o o o GEOMETRIC CLASSIFICATION OF UNITAL GRAPH C ∗-ALGEBRAS 21 Moreover, if B and B′ are elements of MP (m × n, Z) and (U, V ) : B → B′ is a GLP -equivalence, then (U, V ) induces an isomorphism (ξ(U{1},V {1}), ξ(U,V ), ξ(U{2},V {2}), δ(U{1},V {1}), δ(U,V ), δ(U{2},V {2})) of (cyclic six-term) exact sequences. Proof. The first part of the lemma follows directly from the Snake lemma applied to the diagram 0 0 / Zn1 Zn1 ⊕ Zn2 Zn2 B{1} B B{2} / Zm1 / Zm1 ⊕ Zm2 / Zm2 / 0 / 0 The second part of the proof is a straightforward verification. (cid:3) Completely analogous to [BH03], we make the following definitions. Definition 6.8. A subset c of P is called convex if c is nonempty and for all k ∈ P, {i, j} ⊆ c and i (cid:22) k (cid:22) j =⇒ k ∈ c. A subset d of P is called a difference set if d is convex and there are convex sets r and s in P with r ⊆ s such that d = s \ r and i ∈ r and j ∈ d =⇒ j (cid:14) i. Whenever we have such set r, s and d = s \ r, we get a canonical functor from MP to MP2, where P2 = {1, 2} with the usual order if there exist i ∈ r and j ∈ d such that i (cid:22) j, and the trivial order otherwise. Thus such sets will also give a canonical (cyclic six-term) exact sequence as above. Definition 6.9. Let B ∈ MP (m × n, Z). The (reduced) K-web of B, K(B), con- sists of a family of abelian groups together with families of group homomorphisms between these, as described below. For each i ∈ P, let ri = {j ∈ P j ≺ i} and si = {j ∈ P j (cid:22) i}. Note that if ri in the above definition is nonempty, then {i} = si \ ri is a difference set. We let Imm(i) denote the set of immediate predecessors of i (we say that j is an immediate predecessor of i if j ≺ i and there is no k such that j ≺ k ≺ i). For each i ∈ P with ri 6= ∅, we get an exact sequence from Lemma 6.7, (6.1) ker B{i} → cok B{ri} → cok B{si} → cok B{i} Moreover, for every pair (i, j) ∈ P × P satisfying j ∈ Imm(i) and Imm(i) \ {j} 6= ∅ is sj ( ri; consequently we have a homomorphism (6.2) cok B{sj} → cok B{ri} originating from the exact sequence above (cf. Lemma 6.7 used on the division into the sets ri, sj and ri \ sj). Set I P 0 = {ri i ∈ P and ri 6= ∅} ∪ {si i ∈ P} ∪ {{i} i ∈ P} , I P 1 = {i ∈ P ri 6= ∅} . The K-web of B, denoted by K(B), consists of the families (cok B{c})c∈I P and (ker B{i})i∈I P together with all the homomorphisms from the sequences (6.1) and (6.2). Let B′ be an element of MP (m′ × n′, Z). By a K-web isomorphism, κ : K(A) → K(B), we mean families 0 1 (φc : cok B{c} → cok B′{c})c∈I P 0 / / /   / /     / / / / / 22 SØREN EILERS, GUNNAR RESTORFF, EFREN RUIZ, AND ADAM P. W. SØRENSEN and (ψi : ker B{i} → ker B′{i})i∈I P 1 of isomorphisms satisfying that the ladders coming from the sequences in K(B) and K(B′) commute. By Lemma 6.7, any GLP -equivalence (U, V ) : B → B′ induces a K-web isomor- phism from B to B′. We denote this induced isomorphism by κ(U,V ). Remark 6.10. The definitions above are completely analogous to the definitions in [BH03], and are the same in the case mi = ni 6= 0 for all i ∈ P. Note that the last homomorphism in (6.1) is really not needed, because commutativity with this map is automatic. 7. Standard form In this section, we prove that every graph with finitely many vertices is move equivalent to a graph in canonical form (see Definition 7.6). This will allow us to reduce the proof of our classification result to graphs in canonical form. In fact, we will do even better. We will reduce the proof of our classification result to graphs whose adjacency matrices are in the same block form. The first result of this type is the following that allows us to remove breaking vertices (see [BHRS02] for a definition) and regular vertices that do not support a loop. Lemma 7.1. Let E be a graph with finitely many vertices. Then E ∼M E′, where E′ is a graph with finitely many vertices such that every vertex of E′ is either a regular vertex that is the base point of a loop or a singular vertex v satisfying the property that if there exists a path of positive length from v to w, then s−1(v) ∩ r−1(w) = ∞. Proof. First we show how to modify E to get a graph with the property that if v is an infinite emitter, then v emits infinitely many edges to any vertex it emits any edges to. Let v ∈ E0 be an infinite emitter. If there exists a vertex u ∈ E0 such that v emits only finitely many edges to u, we partition s−1(v) into two sets, E1 = {e ∈ s−1(v) s−1(v) ∩ r−1(r(e)) < ∞} and E2 = {e ∈ s−1(v) s−1(v) ∩ r−1(r(e)) = ∞}, i.e. E1 consists of the edges out of v that only have finitely many parallel edges. Note that since E0 is finite, E1 is a finite set. Hence we can perform move (O) according to this partition, resulting in a graph F ′ that is move equivalent to E. Call the vertices v got split into v1 and v2. In F ′, v2 is an infinite emitter with the property that it emits infinitely many edges to any vertex it emits any edges to, and any infinite emitter in E that already had that property keeps it. On the other hand v1 is a finite emitter. Since E0 is finite, we can do the above process a finite number of times, ending with a graph F that is move equivalent to E, and with the property that if v is an infinite emitter, then v emits infinitely many edges to any vertex it emits any edges to. Now we can use move (T) a finite number of times to get a graph G that is move equivalent to F and satisfies that for every infinite emitter v ∈ G0 and every w ∈ G0 for which there exists a path of positive length from v to w we have s−1(v) ∩ r−1(w) = ∞. Finally we use the collapse move (Definition 4.1) on each regular vertex of F that does not support a loop to produce a new graph, E′ say, with E′ ∼M G ∼M E and such that every regular vertex in E′ supports a loop. Because of the way the collapse move adds edges this process maintains the property that s−1(v) ∩ r−1(w) = ∞ for any infinite emitter v and any vertex w with a path of positive length from v to w. (cid:3) GEOMETRIC CLASSIFICATION OF UNITAL GRAPH C ∗-ALGEBRAS 23 Assume E satisfies Condition (K) and satisfies the conclusion of Lemma 7.1. Then every hereditary subset of E0 is saturated and E has no breaking vertices. Moreover, every ideal of C∗(E) is gauge invariant. In particular, there is a lattice isomorphism from the ideal lattice of C∗(E) to the lattice of hereditary subsets of E0 with ordering given by set containment. Therefore, B• E ∈ MP (mE × nE, Z) (in a canonical way) for a partially ordered set P = ({1, . . . , N }, (cid:22)), where N is the number of points in Prim(C∗(E)), and "(cid:22)" is chosen so that it satisfies Assumption 6.3. More formally we have: Lemma 7.2. Let E be a graph with finitely many vertices such that every vertex of E is either a regular vertex that is the base point of a loop or a singular vertex v satisfying the property that if there exists a path of positive length from v to w, then s−1(v) ∩ r−1(w) = ∞. Suppose E satisfies Condition (K). Then B• E ∈ MP (mE × nE, Z), where nE,i = H E i,1 \ H E i,0 the prime ideal corresponding to i and IHE i,0 the maximal proper ideal of with IHE , and IHE i,1 i,1 mE,i = nE,i − (cid:8)v ∈ H E i,1 \ H E i,0 v is a singular vertex in H E i,1 \ H E i,0(cid:9) . Note that the hereditary subsets of vertices -- as usually defined for graphs, when we consider graph C∗-algebras -- correspond to subsets S of P satisfying that i (cid:22) j implies that j ∈ S whenever i ∈ S. This is due to that fact that we generally do not work with the transposed matrix in this paper, since we find it more convenient to work with the non-transposed matrix (see also the proof of Theorem 11.1). We now expand on the conditions we can put on graphs. To turn K-theory isomorphisms into GLP -equivalences or SLP -equivalences, the matrices B• E and B• E ′ must have sufficiently big diagonal blocks, this requirement is captured in (3) and (5) below. The positivity condition, (4), is also critical when dealing with matrix manipulations. Condition (2) ensures that we can apply Propositions 4.4 and 4.7 to actually do matrix manipulations. Theorem 7.3. Let E be a graph with finitely many vertices that satisfies Con- dition (K). Then there exists a graph E′ with finitely many vertices such that E ∼M E′ and E′ satisfies the following properties: (1) every vertex of E′ is either a regular vertex that is the base point of a loop or a singular vertex v satisfying the property that if there exists a path of positive length from v to w, then s−1(v) ∩ r−1(w) = ∞; (2) for all regular vertices v, w of E′ with v ≥ w, there exists a path in E′ from v to w through regular vertices in E; (3) mE ′,i ≥ 3 whenever there exists a cycle in the graph i,1 \ H E ′ i,0 , r−1(H E ′ i,1 \ H E ′ i,0 ) ∩ s−1(H E ′ i,1 ), r, s(cid:17) ; (4) if i (cid:22) j and B• (5) if B• E ′ {i, j} > 0; and E ′ {i} is not the empty matrix, then the Smith normal form of B• E ′ {i, j} is not the empty matrix, then B• E ′ {i} has at least two 1's. Proof. Lemma 7.1 lets us find a graph F such that F ∼M E and F satisfies (1). Us- ing the same technique as described in the proof Proposition 5.1, we can guarantee that F also satisfies (2). Suppose now i is such that mF,i < 3 and there exist a cycle in i,1 \ H F i,0, r−1(H F i,1 \ H F i,0) ∩ s−1(H F i,1), r, s(cid:1) , (cid:16)H E ′ (cid:0)H F 24 SØREN EILERS, GUNNAR RESTORFF, EFREN RUIZ, AND ADAM P. W. SØRENSEN We want to reduce to the case where mF,i = 2. If mF,i = 0 then all the vertices in H F i,0 are infinite emitters. Since the subgraph has a cycle and F satisfies (1) each of the vertices in H F i,0 supports an infinite number of loops. By using move (O) to split two loops of an infinite emitter, we get a graph F ′ that is move equivalent to F , satisfies (1) and (2) and where mF ′,i > 0. Hence we may assume that 1 ≤ mF,i < 3. i,1 \ H F i,1 \ H F i,1 \ H F If mF,i = 1 there are two cases. Case one is that H F i,0 only consists of one vertex. In this case, that vertex must support at least two loops (since F satisfies Condition (K)) and we can use move (O) to split the vertex in to two, thus giving us a move equivalent graph, F ′, that satisfies (1) and (2) and where mF ′,i = 2. The other case is that H F i,0 also contains an infinite emitter. The regular vertex v has to emit at least one edge to one such infinite emitter w. By the construction of H F i,0 w must emit an edge to v, and therefore we can use column addition (Proposition 4.4) to add the w'th column of BF into the v'th column. The result will be a graph F ′ that satisfies (1) and (2) and where v supports at least two loops. Outsplitting, as in case one, we reduce to the case where mF,i = 2. i,1 \ H F i,1 \ H F Suppose now that mF,i = 2. Then there are two regular vertices u, v ∈ H F i,1 \H F i,0 and there is at least one edge from u to v and at least one from v to u. Hence we can add the v'th column of BF into the u'th, using Proposition 4.4, to ensure that u supports at least 2 loops. We can now use move (O) to outsplit u, by dividing the outgoing edges into two in such a way that each partition has a loop, to yield a graph move equivalent to F that satisfies (1), (2) and (3). Hence we can assume that F also satisfies (3). By (3) each nonempty diagonal block of B• F will have a nonzero entry. Hence we may use row and column additions (Propositions 4.7 and 4.4), which are legal because of (2), to make sure that all entries in the diagonal blocks are nonzero. Then we can use column addition to guarantee that all offdiagonal blocks (that are not forced to be zero by the block structure) are strictly positive. Since adding rows and columns together will keep conditions (1), (2) and (3), we may assume that F also satisfies (4). Note, that by the above reasoning, we can assume that any entry in B• F {i} nonempty, we can find a regular vertex, v say, in H F E is not only positive, but greater than or equal to any natural number we see fit. Hence, for each i with B• i,1 such that v emits at least 4 edges to each vertex v reaches. Partition the outgoing edges of v into two sets in such a way that each partition contains at least one edges to each vertex v can reach, and at least two loops. Let d1 be the number of loops in the first partition, and let d2 be the number in the second (then d1 + d2 = d). Outsplitting according to this partition will yield a graph F ′ such that F ∼M F ′ and F satisfies (1), (2), (3) and (4). B• F ′ {i} will contain the following two rows (corresponding to the vertices v was split into) (cid:18)d1 − 1 d2 d1 ∗ d2 − 1 ∗ ∗ · · · ∗ · · ·(cid:19) , where the asterisks can be any positive numbers, with mF ′,i = mF,i + 1. Repeating this process we can increase the size of the relevant block so much that the Smith normal form must contain at least two 1's. Continuing in this fashion for each diagonal block we can construct E′ such that (cid:3) E ∼M E′ and E′ satisfies (1), (2), (3), (4) and (5). Remark 7.4. Suppose that E is a graph with finitely many vertices that satisfies (1) and (3) of Theorem 7.3. Then E satisfies Condition (K). Moreover, if H E i,0 = {vi}, then either vi is an infinite emitter that does not support a cycle or is a sink. i,1 \ H E GEOMETRIC CLASSIFICATION OF UNITAL GRAPH C ∗-ALGEBRAS 25 Remark 7.5. Let E be a graph with finitely many vertices that satisfies Condi- tion (K). It follows from the proof of Theorem 7.3 that if E satisfies (1), (2) and (3) from the theorem, then there exists a graph E′ that is move equivalent to E and satisfies (1), (2), (3) and (4), furthermore B• E ′ have the same block form and are SLP -equivalent. E and B• Since the Smith normal form of a matrix is invariant under SL-equivalence E′ will satisfy condition (5) if E does. Definition 7.6. A graph E with finitely many vertices is in canonical form if E satisfies the properties (1), (2), (3), (4), and (5) of Theorem 7.3. A pair of graphs (E, F ) with finitely many vertices are in standard form if E and F are in canonical form with mE = mF , and nE = nF . The notion of a standard form is of course only useful if we can assume that our graphs have the standard form, the next proposition shows that we can indeed assume that, if the corresponding C∗-algebras have isomorphic ordered reduced filtered K-theory. Proposition 7.7. Let there be given graphs E1 and E2 with finitely many vertices. If FK+ R(C∗(E2)), then there exists a pair of graphs (F1, F2) with finitely many vertices such that the pair (F1, F2) is in standard form and Ei ∼M Fi. R(C∗(E1)) ∼= FK+ G1 and B• Proof. It follows from Theorem 7.3 that we can find graphs G1, G2 such that Gi ∼M Ei and Gi are in canonical form, i = 1, 2. The K-theory condition gives a specific isomorphism between the primitive ideal spaces of C∗(E1) and C∗(E2), hence B• G2 can be chosen to have the same same block structure according i,1 \ H E1 to this isomorphism. Furthermore, the number of singular vertices in H E1 i,0 is determined by its K-theory, since C∗(H E1 i,0 ) is simple (see [Sør13, Lemma 9.2]). The same holds for E2 so nE1,i − mE1,i = nE2,i − mE2,i for all i, and therefore nG1,i − mG1,i = nG2,i − mG2,i for all i. i,1 \ H E1 G1 are at least 4. Similarly we can assume that all nonzero entries of B• The only potential problem is now that the we may not have nG1,i = nG2,i for all i. Since all the entries in B• G1 are positive, unless forced to be zero by the block structure, we may use row and column additions to ensure that all nonzero entries in B• G2 are at least 4. So if nG1,i < nG2,i for some i, we can use an outsplit (similar to what is described at the end of the proof of Theorem 7.3) to grow nG1,i by 1 while keeping it in canonical form. Proceeding this way, we construct graphs F1, F2 in canonical form such that F1 ∼M E1, F2 ∼M E2 and nF1,i = nF2,i for all i. Since we also have nF1,i − mF1,i = nF2,i − mF2,i we must have mF1,i = mF2,i for all i. (cid:3) When E is in canonical form, the rows of BE that are removed to form B• E either have all entries equal to 0 except on which is −1, this is in case the corresponding vertex is a sink, or it only contains 0 and ∞, in case it is an infinite emitter. It therefore follows from Proposition 4.4 and Remark 4.5 that adding one column in B• E into another will preserve move equivalence, so long as it maintains the block structure and similarly for rows by Proposition 4.7 and Remark 4.8. Hence we have: Corollary 7.8. Let E be a graph with finitely many vertices and suppose that E is in canonical form. In B• E we can add column l into column k without changing the move equivalence class of the associated graph if the diagonal entry of column l is in block i, the diagonal entry of column k is in block j and i (cid:22) j. Similarly can add row l into row k without changing move equivalence class if the diagonal entry of row l is in block i, the diagonal entry of row k is in block j and j (cid:22) i. For the results in Section 8 we need a final refinement of our standard form, we E to have has greatest common divisor 1. We also need the diagonal blocks of B• 26 SØREN EILERS, GUNNAR RESTORFF, EFREN RUIZ, AND ADAM P. W. SØRENSEN achieve this in the next proposition by making sure that each diagonal block has an entry that is equal to 1. R(C∗(E1)) ∼= FK+ Proposition 7.9. Let there be given graphs E1 and E2 with finitely many vertices. If FK+ R(C∗(E2)), then there exists a pair of graphs (F1, F2) with finitely many vertices such that the pair (F1, F2) is in standard form, Ei ∼M Fi and each nonempty diagonal block B• Fi contains a 1. Proof. By Proposition 7.7 we can find F1, F2 satisfying the conclusion of the propo- sition, except for the last condition. As in the proof of Proposition 7.7 we may use row and column operations to ensure that all nonzero entries of B• G2 are at least 4. For each nonzero diagonal block, we will now do an outsplit similar to what is described at the end of the proof of Theorem 7.3, where we find a regular vertex, v say, that supports at least two loops. Partition the outgoing edges of v into two sets in such a way that each partition contains at least one edges to each vertex v can reach, but we also insist that one partition only contains two loops. Let d1 be the number of loops in the first partition, and let d2 be the number in the second (then d1 + d2 = d). Our assumption forces either d1 or d2 to be 2, for simplicity let us say that d1 = 2. As noted in the proof of Theorem 7.3 in the resulting graph, the diagonal block will contain the rows G1 and B• (cid:18)d1 − 1 d2 d1 d2 − 1 ∗ ∗ ∗ · · · ∗ · · ·(cid:19) =(cid:18) 1 d2 2 d2 − 1 ∗ ∗ ∗ ∗ · · · · · ·(cid:19) . Hence it contains a 1. Doing this for all nonzero diagonal blocks yields the desired graphs. (cid:3) 8. Generalization of Boyle-Huang's lifting result We aim to prove Theorem 8.12 which says that -- in certain cases -- every K-web isomorphism is induced by a GLP -equivalence. This is the main result of this section. To prove Theorem 8.12, we first strengthen [BH03, Theorem 4.5]. The following theorem is a classical well-known theorem, cf. [New72, Section II.15]. Theorem 8.1 (Smith normal form). Suppose B is an m × n matrix over Z. Then there exist matrices U ∈ GL(m, Z) and V ∈ GL(n, Z) such that the matrix D = U BV satisfies the following • D(i, j) = 0 for all i 6= j, • the min(m, n) × min(m, n) principal submatrix of D is a diagonal matrix diag(d1, d2, . . . , dr, 0, 0, . . . , 0), where r ∈ {0, 1, . . . , min(m, n)} is the rank of B and d1, d2, . . . , dr are pos- itive integers such that didi+1 for i = 1, . . . , r − 1. For each matrix B, the matrix D is unique and is called the Smith normal form of B. We now recall some terminology that was introduced in [BH03]. Definition 8.2. Let B be an element of M(m × n, Z). A GL self-equivalence of B is a GL-equivalence (U, V ) : B → B. We say that an automorphism φ of cok B is GL-allowable if there exists a GL self-equivalence, (U, V ), of B such that the isomorphism κ(U,V ) induces φ. Lemma 8.3. Let B be an m × n matrix over Z, and let U ∈ GL(m, Z) and V ∈ GL(n, Z) be given invertible matrices. Then gcd B = gcd(U BV ). In particular, if D is the Smith normal form of B, then gcd B = D(1, 1) = d1. GEOMETRIC CLASSIFICATION OF UNITAL GRAPH C ∗-ALGEBRAS 27 Proof. We may assume that B 6= 0. Let d be a positive integer. Then d divides all entries of B ⇔ ∀i, j : d eT i Bej ⇔ ∀x ∈ Zm, y ∈ Zn : d xT By ⇔ ∀x ∈ Zm, y ∈ Zn : d xT U BV y ⇔ ∀i, j : d eT ⇔ d divides all entries of U BV. i U BV ej Now the lemma follows. (cid:3) Remark 8.4. Let B be an m×n matrix over Z. Then it follows from the above, that m is greater than the number of generators of cok B according to the decomposition from the Smith normal form into direct sums of nonzero cyclic groups if and only if gcd B = 1. Boyle and Huang show in their paper [BH03] the following fundamental theorem. Theorem 8.5 ([BH03, Theorem 4.4]). Let B be a n × n (square) matrix over a PID R, and let δ = gcd B. Let φ be an automorphism of cok B, and let M be any n × n matrix over R defining φ, i.e., φ([x]) = [M x] for all x ∈ Zn. Then det(M ) ≡ 1 (mod δ) if and only if there exist n × n matrices U and V over R with determinants 1 such that U BV = B and U is defining φ. Then det(M ) ≡ u (mod δ) for some unit u in R if and only if there exist n × n invertible (GL) matrices U and V over R such that U BV = B and U is defining φ. Remark 8.6. As we will see, it is possible to generalize the part about GL-allowance in this theorem to rectangular matrices, the analogous statement to the part about SL-allowance in Theorem 8.5 does not hold in general (for rectangular matrices). If we consider the matrix B =(cid:18)3 0(cid:19) M =(cid:18)−1 0 −1(cid:19) 0 and the automorphism − id on cok B ∼= Z/3 ⊕ Z induced by the matrix it is easy to see that we get a counterexample. Although it can be done, we do not investigate this further, since for our purposes we do not need to know when automorphisms can be lifted to SL-equivalences. In [BH03] there is the following useful theorem. Note that all ni's are assumed to be nonzero in [BH03]. Theorem 8.7 ([BH03, Theorem 4.5]). Suppose B and B′ are matrices in MP (n, Z) with corresponding diagonal blocks equal, and κ : K(B) → K(B′) is a K-web iso- morphism. Then there exist matrices U, V ∈ GL(n, Z) such that we have a GLP - equivalence (U, V ) : B → B′ satisfying κ(U,V ) = κ if and only if each of the auto- morphisms di : cok B{i} → cok B′{i} defined by κ is GL-allowable. Together with [BH03, Theorem 4.4] (see Theorem 8.5 above), this gives us the following useful corollary. Corollary 8.8 ([BH03, Corollary 4.7]). Let B and B′ be matrices in MP (n, Z) with gcd B{i} = 1 = gcd B′{i} for all i ∈ P. Then for any K-web isomorphism κ : K(B) → K(B′) there exist matrices U, V ∈ GL(n, Z) such that we have a GLP - equivalence (U, V ) : B → B′ satisfying κ(U,V ) = κ. 28 SØREN EILERS, GUNNAR RESTORFF, EFREN RUIZ, AND ADAM P. W. SØRENSEN But even more is true. We can generalize [BH03, Theorem 4.4 (and Proposi- tion 4.1)] (cf. Theorem 8.5) as follows (we here only consider the case R = Z). Theorem 8.9. Let B be a n × n (square) matrix over Z, and let δ = gcd B. Let φ be an automorphism of cok B, let ψ be an automorphism of ker B, and let M be any n × n matrix over Z defining φ, i.e., φ([x]) = [M x] for all x ∈ Zn. Then det(M ) ≡ ±1 (mod δ) if and only if there exist n × n invertible (GL) matrices U and V over Z such that U BV = B and U is defining φ and V −1 is defining ψ. Proof. The only thing that does not follow from [BH03, Theorem 4.4] is that we can choose the GL-equivalence (U, V ) such that it also induces the right automorphism on ker B. For this, it is clear that we may assume that B is its own Smith normal form (just like in the proof of [BH03, Theorem 4.4]). We use [BH03, Theorem 4.4] to get a GL-equivalence (U, V ) : B → B that induces φ on cok B. The matrix V −1 induces an automorphism ψ′ of ker B. Now we will find a GL-equivalence (I, V ′) : B → B that induces ψ ◦ψ′−1 on ker B -- then (U, V V ′) is a GL-equivalence that induces φ on cok B and ψ on ker B. Now, the automorphism ψ ◦ ψ′−1 on ker B uniquely determines what V ′−1 should be on the lower right block matrix (where we write the matrices as 2 × 2 block matrices according to the nonzero respectively zero part of the diagonal of B). Let V ′−1 be the block diagonal matrix that has this matrix as lower right block matrix and the identity as the upper left block matrix. (cid:3) Now we let Pmin = {i ∈ P : j ≺ i ⇒ i = j}. Using the above result, we get the following stronger version of Theorem 8.7: Theorem 8.10 (Strengthening of [BH03, Theorem 4.5]). Let n = (ni)i∈P be 6= 0, for all i ∈ P. Suppose B and B′ are matrices in a multiindex with ni MP (n, Z) with corresponding diagonal blocks equal, and κ : K(B) → K(B′) is a K-web isomorphism. Suppose that for each i ∈ Pmin, we have an automorphism ψi : ker B{i} → ker B{i}. Then there exist matrices U, V ∈ GL(n, Z) such that we have a GLP -equivalence (U, V ) : B → B′ satisfying κ(U,V ) = κ if and only if each of the automorphisms di : cok B{i} → cok B′{i} defined by κ are GL-allowable -- moreover, the GLP -equivalence can always be chosen such that V −1{i} induces ψi for each i ∈ Pmin. Proof. The only thing that does not follow from [BH03, Theorem 4.4] (cf. Theo- rem 8.7), is that we can choose the GL-equivalence (U, V ) such that it also induces the right automorphisms on ker B{i}, i ∈ Pmin. We choose a GLP -equivalence (U, V ) according to Theorem 8.7, so that it induces the given K-web isomorphism. For each i ∈ Pmin, this gives an automorphism ψ′ i of ker B{i}. Now choose GL- i ) of B{i} according to (the proof of) Theorem 8.9 so that V ′−1 eqivalences (I, V ′ induces ψi ◦ ψ′−1 for each i ∈ Pmin. Let V be the block matrix that is the identity matrix everywhere except that V {i} = V ′ i for every i ∈ Pmin. It is straight forward to verify that (I, V ′) is a GLP -equivalence from B′ to B′, and that (U, V V ′) induces exactly what we want. (cid:3) i i Together with [BH03, Theorem 4.4 (and Proposition 4.1)] (see Theorem 8.5 above), this gives us the following stronger version of Corollary 8.8: Corollary 8.11 (Strengthening of [BH03, Corollary 4.7]). Let n = (ni)i∈P be a multiindex with ni 6= 0, for all i ∈ P. Suppose B and B′ are matrices in MP (n, Z) with gcd B{i} = 1 = gcd B′{i} for all i ∈ P. Then for any K-web isomorphism κ : K(B) → K(B′) together with automorphisms ψi : ker B{i} → ker B{i}, for GEOMETRIC CLASSIFICATION OF UNITAL GRAPH C ∗-ALGEBRAS 29 i ∈ Pmin, there exist matrices U, V ∈ GL(n, Z) such that we have a GLP -equivalence (U, V ) : B → B′ satisfying κ(U,V ) = κ and V −1{i} induces ψi for each i ∈ Pmin. The following theorem is the main result of this section, and allows us -- in certain cases -- to lift K-web isomorphisms to GLP -equivalences for rectangular cases. Although it is possible to prove this directly, imitating the proof in [BH03], the present proof is much shorter and reduces the rectangular case to the square case and uses the results from [BH03]. Theorem 8.12. Let m = (mi)i∈P , n = (ni)i∈P ∈ (N0)N be multiindices. Suppose B and B′ are matrices in MP (m × n, Z) with gcd B{i} = 1 = gcd B′{i} for all i ∈ P with mi 6= 0 and ni 6= 0. Then for any K-web isomorphism κ : K(B) → K(B′) there exist matrices U ∈ GL(m, Z) and V ∈ GL(n, Z) such that we have a GLP -equivalence (U, V ) : B → B′ satisfying κ(U,V ) = κ. If, moreover, we have given an isomorphism ψi : ker B{i} → ker B′{i}, for every i ∈ Pmin, then we can choose the above GLP -equivalence (U, V ) such that -- in addition to the above -- also V −1{i} induces the ψi, for all i ∈ Pmin. i ∈ GL(mi, Z) and Vi, V ′ Proof. For each i ∈ P, choose Ui, U ′ i ∈ GL(ni, Z) such that Di = UiBVi and D′ i are the Smith normal forms of B and B′, i B′V ′ respectively (cf. Theorem 8.1). Let U, U ′ ∈ GL(m, Z) and V, V ′ ∈ GL(n, Z) be the block diagonal matrices with Ui, U ′ i in the diagonals, respectively. i , Vi and V ′ i = U ′ Then U BV and U ′B′V ′ are in MP (m × n, Z) and (U, V ) : B → U BV and (U, V ) : B′ → U ′B′V ′ are GLP -equivalences inducing K-web isomorphisms κ(U,V ) from K(B) to K(U BV ) and κ(U ′,V ′) from K(B′) to K(U ′B′V ′), respectively. More- over, Lemma 8.3 ensures that we still have gcd(U BV ){i} = 1 = gcd(U ′B′V ′){i}. Thus we can without loss of generality assume that each diagonal block is equal to its Smith normal form. Also note, that because we have a K-web isomorphism from K(B) to K(B′), now the diagonal blocks are necessarily identical. Let r be such that ri = max(mi, ni) for all i ∈ P. Let, moreover, C, C′ ∈ MP (r, Z) denote the matrices B and B′ enlarged by putting zeros outside the original matrices. Define rc and rk by rc i = max(mi−ni, 0) for all i ∈ P. In the (reduced) K-web we are considering the modules Cc(B) = cok B(c) where c is {i}, {j ∈ P : j ≺ i} 6= ∅ or {j ∈ P : j (cid:22) i} for i ∈ P -- and similarly for B′. It is clear that when we consider C and C′ we just add onto these cokernels i = max(ni−mi, 0) and rk Zrc j , Mj∈c Zrk i , and that the maps between the modules are the obvious ones. Similarly for the modules Kd(B) = ker B(d) where d is {i} where {j ∈ P : j ≺ i} 6= ∅ -- and similarly for B′. It is clear that when we consider C and C′ we just add onto these kernels where d = {i}. And connecting homomorphism will be the zero maps. setting it to be the identity on the new groups. By Corollary 8.8, we see that there exist matrices U, V ∈ GL(r, Z) such that we have a GLP -equivalence (U, V ) : C → Thus we can extend the isomorphism κ to an isomorphismeκ : K(C) → K(C′) by C′ satisfying κ(U,V ) =eκ. We may (according to Theorem 8.11) actually assume that (U, V ) induces ψi plus the identity on the new summands of ker B{i} for i ∈ Pmin as well. Now let us look at the i'th diagonal block. We now want to cut U and V down to match the original structure. Naturally there are three cases to consider. The first one, mi = ni is trivial. 0 0 0 0 0 C00 0 0  0 ,  and  where C00 is an invertible matrix over Q and the last diagonal block has size (ni − mi) × (ni − mi). We write U and V as U11 U12 U13 U21 U22 U23 U31 U32 U33  V11 V12 V13 V21 V22 V23 V31 V32 V33  , according to the block structure of C{i} (and C′{i}). The condition implies that U CV = C′ U11C00V11 = C00, Ui1C00V1j = 0, for all (i, j) 6= (1, 1). Since C00 is invertible as a matrix over Q, we see that also U11 and V11 have to be invertible over Q. Thus V12 = 0, V13 = 0, U21 = 0, and U31 = 0. Moreover, since we have to get the identity homomorphism on the new direct summand, we need to have U33 = I, U23 = 0 and U32 = 0. So now let U0 be the block matrix where we erase the rows and columns corresponding to change the size of the i'th diagonal block from ri × ri to mi × mi -- call the new size r′. Moreover, we let C0 and C′ 0 be the block matrices where we erase the rows corresponding to change the size of the i'th diagonal block from ri × ri to mi × ni. Note that the i'th diagonal block now is the matrix(cid:0) U11 U12 0 U22(cid:1). This is a GL matrix that induces the right automorphism of cok B{i}. Moreover, clearly U0{i}B{i}V {i} = B{i}. But more is true. We have that U0 is a GL(r′, Z) matrix and that U0C0V = C′ 0 and the induced K-web isomorphism agrees with the original on all parts except for the direct summands we cut out. Now consider instead the case mi > ni. In this case, cok C{i} = cok B{i} and ker C{i} = ker B{i} ⊕ Zmi−ni -- and similarly for B′ and C′. We write C{i} = C′{i} as 30 SØREN EILERS, GUNNAR RESTORFF, EFREN RUIZ, AND ADAM P. W. SØRENSEN Now consider the case mi < ni. In this case, cok C{i} = cok B{i} ⊕ Zni−mi and ker C{i} = ker B{i} -- and similarly for B′ and C′. We write C{i} = C′{i} as 0 0 0 0 0 C00 0 0  0 ,  and  where C00 is an invertible matrix over Q and the last diagonal block has size (mi − ni) × (mi − ni). We write U and V as U11 U12 U13 U21 U22 U23 U31 U32 U33  V11 V12 V13 V21 V22 V23 V31 V32 V33  , according to the block structure of C{i} (and C′{i}). The condition implies that U CV = C′ U11C00V11 = C00, Ui1C00V1j = 0, for all (i, j) 6= (1, 1). Since C00 is invertible as a matrix over Q, we see that also U11 and V11 have to be invertible over Q. Thus V12 = 0, V13 = 0, U21 = 0, and U31 = 0. Moreover, since we have to get the identity homomorphism on the new direct summand, we need to have V33 = I, V23 = 0 and V32 = 0. So now let V0 be the GEOMETRIC CLASSIFICATION OF UNITAL GRAPH C ∗-ALGEBRAS 31 block matrix where we erase the rows and columns corresponding to change the size of the i'th diagonal block from ri × ri to ni × ni -- call the new size r′. Moreover, we let C0 and C′ 0 be the block matrices where we erase the rows corresponding to change the size of the i'th diagonal block from ri × ri to mi × ni. Note that the i'th diagonal block now is the matrix(cid:0) V11 0 V21 V22(cid:1). This is a GL matrix that induces the right automorphism of ker B{i}. Moreover, clearly U {i}B{i}V0{i} = B{i}. But more is true. We have that V0 is a GL(r′, Z) matrix and that U C0V0 = C′ 0 and the induced K-web isomorphism agrees with the original on all parts except for the direct summands we cut out. Induction finishes the proof. (cid:3) 9. GLP -equivalence to SLP -equivalence In this section we are concerned with Step 3 in our proof outline in Section 3 of the proof of (3) implies (1) in Theorem 3.1. It is of course not true in general that any two GLP -equivalent matrices will be SLP -equivalent, so we will need to alter our matrices. Our first step in that direction is to create a little more room. Lemma 9.1. Let E be a graph with finitely many vertices and suppose BE = A X Y 0 B Z 0 0 C where B is an n × n matrix with entries from N0 ⊔ {∞} for some n ≥ 2 and the entries of rows n − 1 and n of B are positive integers and the vertices corresponding to these two rows are regular vertices of E. Then there exists a graph E′ such that E ∼M E′, and with B′ an (n + 2) × (n + 2) matrix with entries from N0 ⊔ {∞} and there exists V ∈ M(n + 2, Z) with det(V ) = 1 such that where BE ′ =  I 0 0 V 0 0 A X ′ Y ′ 0 B′ Z ′ 0 0 0 0 C I = E ′ = I2(cid:19) , B• 0 A0 X ′ 0 B′ 0 0 Y ′ 0 0 Z ′ 0 0 C0  E = 0 =(cid:0)X0 B• X ′′ A0 X0 Y0 0 B0 Z0 0 0 C0  , 0 =(cid:18)B0 0 B′′ 0(cid:1) , A0 X ′′ Y0 0 0 Z ′′ 0 B′′ 0 C0 0 0   , 0 =(cid:18)Z0 0(cid:19) . A0 X ′ 0 B′ 0 0 Y ′ 0 0 Z ′ 0 0 C0 and Z ′′ Moreover, if E satisfies the property that for all v, w ∈ E0 reg with v ≥ w, there exists a path in E from v to w through regular vertices in E, then E′ also satisfies the same property. Proof. Let v be the vertex in E corresponding to the entry B(n − 1, n − 1) + 1 and w be the vertex in E corresponding to the entry B(n, n) + 1. Outsplitting the vertices v and w with respect to the partitions s−1(v) = {e} ⊔(cid:0)s−1(v) \ {e}(cid:1) where r(e) = v and s−1(w) = {f } ⊔(cid:0)s−1(w) \ {f }(cid:1) where r(f ) = w, we get a graph F such that E ∼M F . 32 SØREN EILERS, GUNNAR RESTORFF, EFREN RUIZ, AND ADAM P. W. SØRENSEN Let X1 and B1 be the column vector of X and B, respectively, that corresponds to the vth-column of BE and let X2 and B2 be the column vector of X and B, respectively, that corresponds to the wth-column of BE. Then BF = where X ′ =(cid:0)X X1 X2(cid:1), Z ′ =(cid:18)Z 0(cid:19), and A X ′ Y 0 0 bB Z ′ C 0  bB = B − J 0 0 · · · · · · 0 0 1 0 0 1 B1 B2 0 0 0 0  , where J is the matrix that is zero in all entries except the last two diagonal entries which are 1. These account for the loops v and w lost when we did the outsplit. We can now use row additions (Corollary 4.6), adding the last row of bB into the third last and the second last into the fourth last, to get a graph E′ such that E′ ∼M F ∼M E and where BE ′ = A X ′ Y 0 B′ Z ′ 0 0 C , where Let eX0 be the matrix obtained from X ′ and the n'th column of X ′ from B′ but keeping that last two rows intact. Consider the matrix 0 by replacing the (n − 1)'st and the n'th column of B0 by the zero column B 0 0 1 0 0 1 0 · · · 0 · · ·  . B1 B2 0 0 0 0 0 by replacing the (n − 1)'st column of X ′ 0 B′ = 0 by the zero column and let eB0 be the matrix obtained  eX0 Y0 eB0 Z ′  . 0 0 C0 A0 0 0 Let V1 ∈ M(n + 2, Z) be the matrix obtained from In+2 by switching the (n − 1)'st and the (n+1)'st columns and let V2 ∈ M(n+2, Z) be the matrix obtained from In+2 by switching the n'th and the (n + 2)'nd columns. Then det(V1) = det(V2) = −1 and A0 0 0  eX0 Y0 eB0 Z ′ 0 0 C0  0 I 0 V1V2 0 0 0 0 I = A0 X ′′ Y0 0 0 B′′ 0 Z ′′ 0 C0 0 0  . As in the proof of Proposition 5.2 we let E(i,j) denote the elementary matrix that is equal to the identity matrix everywhere except for the (i, j)'th entry, that is 1. Let V3 = E(n+1,n−1) ∈ M(n + 2, Z), and let V4 = E(n+2,n) ∈ M(n + 2, Z). Then det(V3) = det(V4) = 1 and A0 0 0  eX0 Y0 eB0 Z ′ 0 0 C0  0 I 0 V3V4 0 0 0 0 I = A0 X ′ 0 B′ 0 0 Y0 0 Z ′ 0 0 C0  . GEOMETRIC CLASSIFICATION OF UNITAL GRAPH C ∗-ALGEBRAS 33 Set V = V −1 4 V −1 3 V1V2. Then A0 X ′ 0 B′ 0 0 Y0 0 Z ′ 0 0 C0   I 0 0 V 0 0 0 0 I = A0 X ′′ Y0 0 0 Z ′′ 0 B′′ 0 0 0 C0  . Since det(V1) = det(V2) = −1 and det(V3) = det(V4) = 1, we have that det(V ) = 1. For the last part of the lemma, let v1 and v2 be the two additional vertices obtained from the outsplitting. It is clear that if v, w ∈ E0 reg such that v ≥ w in E′, then v ≥ w in E. Thus, there exists a path in E′ from v to w through regular vertices of E′. Suppose v ∈ E0 reg and v ≥ vi. Then by the definition of outsplitting and by the assumption on E, there exists an edge e in E′ such that r(e) = vi, reg, and v ≥ s(e). It is now clear that there exists a path in E′ from v to s(e) ∈ E0 vi through regular vertices of E′. Suppose vi ≥ v with v ∈ E0 reg. By the definition of the outsplitting, there exists a path α = α1 · · · αm in E′ such that s(α1) = vi, r(αm) = v, r(α1) ∈ E0 reg, and r(α1) ≥ vi. Since r(α1) ≥ vi and vi ≥ v, we have that r(α1) ≥ v. By the previous cases, we have that there exists a path in E′ from r(α1) to v in E′ through regular vertices in E′. Hence, there exists a path in E′ from vi to v through regular vertices in E′. For the pair (v1, v2) this is clear by construction. (cid:3) We now connect the space we have created to our method of changing signs, i.e. the Cuntz splice. Lemma 9.2. Let E be a graph with finitely many vertices such that BE = A X Y 0 B Z 0 0 C where B is an n × n matrix with entries from N0 ⊔ {∞} for some n ≥ 1 and the entries of row n of B are positive integers and the vertex v corresponding to this row is a regular vertex of E. Let Ev,− be the Cuntz splice of E at the vertex v. Then det(U ) = 1, det(V ) = −1, and 0 0 I 0 U 0 0 0 I z  A0 0 0  where X ′′ 0 =(cid:0)X0 0(cid:1) , B• Ev,− (X−)0 (B−)0 0 {  Y0 C0 (Z−)0 } E = 0 =(cid:18)B0 B′′ 0 B• A0 X0 Y0 0 B0 Z0 0 0 C0 0 I2(cid:19) , 0 I 0 V 0 0 0 0 I =  , and Z ′′ A0 X ′′ Y0 0 0 Z ′′ 0 B′′ 0 0 0 C0  0(cid:19) , 0 =(cid:18)Z0 V = V1V2 where V1 is the matrix obtained from In+2 by subtracting the (n + 2)'nd column from the n'th column and V2 is the matrix obtained from In+2 by switching the (n + 1)'st and (n + 2)'nd columns, and U is the matrix obtained from Im+2 by subtracting the (m + 2)'nd row from the m'th row, with m being the number of rows of B0. Proof. The proof of the lemma is just a simple matrix computation using that the row operations to get U only involve regular vertices of Ev,−, and is left for the reader. (cid:3) 34 SØREN EILERS, GUNNAR RESTORFF, EFREN RUIZ, AND ADAM P. W. SØRENSEN The next proposition will be our first step in going from a GLP -equivalence to an SLP -equivalence. The idea is to alter our graphs in such a way that their B• matrices are GLP -equivalent, say by (U, V ), but where all the diagonal blocks of U have determinant one. Thus moving all our problems to V . Proposition 9.3. Let E1 and E2 be graphs with finitely many vertices such that (E1, E2) is in standard form. Suppose (U1, V1) : B• E2 is a GLP -equivalence, where U1 ∈ GLP (m, Z), V1 ∈ GLP (n, Z), m = (m1, . . . , mN ) and n = (n1, . . . , nN ). 6= 0, det(U1{i}) = −1, then there exist graphs F1 and F2 If, for some i, mi with finitely many vertices and there exist U2 ∈ GLP (m′, Z), V2 ∈ GLP (n′, Z), where m′ = (m1, . . . , mi−1, mi + 2, mi+1, . . . , mN ) and n′ = (n1, . . . , ni−1, ni + 2, ni+1, . . . , nN ), such that E1 → B• • Ek ∼M Fk, k = 1, 2; • (F1, F2) is in standard form; • U2B• • det(U2{i}) = 1, det(V2{i}) = − det(V1{i}); and • det(U2{j}) = det(U1{j}) and det(V2{j}) = det(V1{j}) for all j 6= i. F1 V2 = B• F2 ; A1 X1 Y1 0 B1 Z1 0 0 C1 Proof. Write B• E1 as  Define eU ∈ GLP (m, Z) by Bk = B• Ek {i}. Apply Lemma 9.1 to both Ek's to yield graphs E′ k and matrices V ′ k.  and write B• E2 as  A2 X2 Y2 0 B2 Z2 0 0 C2 , where 0 1 0  if (r, s) = (i, i) 0(cid:17) if (r, s) = (r, i), r 6= i if (r, s) = (i, s), s 6= i 0 0 U1{i} 0 0 1  (cid:16)U1{r, i} 0   U1{i, s} 0 0 U1{r, s} otherwise eU{r, s} =  0 0 I where V is the matrix defined by  if (r, s) = (i, i) 0(cid:17) if (r, s) = (r, i), r 6= i if (r, s) = (i, s), s 6= i 0 1 0 0 (V ′ 2 )−1 0 0 0 0 0 I 0 0 V1{i} 0 0 1 I V   (cid:16)V1{r, i} 0   V1{i, s} 0 0  V {r, s} = GEOMETRIC CLASSIFICATION OF UNITAL GRAPH C ∗-ALGEBRAS 35 and set eV = I 0 0 V ′ 1 0 0 V1{r, s} otherwise . E ′ E ′ 2 eU B• By Lemma 9.1, we have that 1eV = B• Note that eV ∈ GLP (n, Z). Note that det(eV {i}) = − det(V1{i}), det(eU {i}) = 1, and det(eU {j}) = det(U1{j}) and det(eV {j}) = det(V1{j}) for all j 6= i. Since Ek is in canonical form, we have that for every vertex v, w ∈ (Ek)0 reg with v ≥ w, there exists a path in Ek from v to w k)0 through regular vertices in Ek. Hence, by Lemma 9.1, for every v, w ∈ (E′ reg with v ≥ w, there exists a path in E′ k from v to w through regular vertices in E′ i. Since Ek is in canonical form and by definition of the outsplitting graph, E′ k satisfies (1), (2), and (3) of Theorem 7.3. Furthermore, the fact that Ek is in canonical form implies that all the diagonal blocks of B• have Smith normal form with at least two 1's, so Ek it follows from the constructions of Lemma 9.1 that the diagonal blocks of B• also E ′ k have this property. Therefore E′ k also satisfies (5) of Theorem 7.3. By Remark 7.5 = n′, we get a graph Fi in canonical form such that mFi = mE ′ and E′ i ∼M ′ Fi. Also, we get an SLP -equivalence (Wi, Zi) : B• = m′, nFi = nE ′ Fi → B• E ′ i Set U2 = W −1 2 . Since W −1 2 eU W1 and V2 = Z1eV Z −1 , W1 ∈ SLP (m′, Z) and since Z1, Z −1 2 ∈ SLP (n′, Z), we have that det(U2{i}) = 1, det(V2{i}) = − det(V1{i}), and det(U2{j}) = det(U1{j}) and det(V2{j}) = det(V1{j}) for all j 6= i. By construction, the pair (F1, F2) is in standard form with B• Fi ∈ MP (m′ × n′, Z), U2B• (cid:3) F1 V2 = B• We now use the Cuntz splice to fix potential sign problems on V . F2 , and Fi ∼M Ei. 2 . i i Proposition 9.4. Let E1 and E2 be graphs with finitely many vertices such that (E1, E2) is in standard form. Suppose (U1, V1) : B• E2 is a GLP -equivalence, where U1 ∈ GLP (m, Z) and V1 ∈ GLP (n, Z), and m = (m1, . . . , mN ) and n = (n1, . . . , nN ). If, for some i, mi 6= 0, det(U1{i}) = 1 and det(V1{i}) = −1, there exist graphs F1 and F2 with finitely many vertices and there exist U2 ∈ GLP (m′, Z) and V2 ∈ GLP (n′, Z), where m′ = (m1, . . . , mi−1, mi + 2, mi+1, . . . , mN ) and n′ = (n1, . . . , ni−1, ni + 2, ni+1, . . . , nN ) such that E1 → B• • Ek ∼M ′ Fk, for k = 1, 2; • (F1, F2) is in standard form; • U2B• • det(U2{i}) = det(V2{i}) = 1; and F1 V2 = B• F2 ; 36 SØREN EILERS, GUNNAR RESTORFF, EFREN RUIZ, AND ADAM P. W. SØRENSEN • det(U2{j}) = det(U1{j}) and det(V2{j}) = det(V1{j}) for all j 6= i. Proof. Write B• E1 as  A1 X1 Y1 0 B1 Z1 0 0 C1  and write B• E2 as  A2 X2 Y2 0 B2 Z2 0 0 C2 , where Bk = B• {i} and the entries of the last two rows of Bk are positive integers, and Ek the corresponding vertices of Ek are regular. Apply Lemma 9.1 to E2 to yield a graph E′ 2 . Let U− and V− be the matrices that one obtain when applying Lemma 9.2 to the graph E1. 2 and a matrix V ′ 0 0 0 0 1 0 0 0 0 U1{i, s} U{r, s} = if (r, s) = (i, s), s 6= i 0 I 0 U− 0 0 U1{i} 0 1 0 Set eU = U I where U is the matrix defined by  if (r, s) = (i, i)  (cid:16)U1{r, i} 0 0(cid:17) if (r, s) = (r, i), r 6= i   Note that eU ∈ GLP (m′, Z). I V  Set eV =  (cid:16)V1{r, i} 0   V1{i} 0 1 0 I 0 0 V− 0 0 0 (V ′ 2 )−1 0 if (r, s) = (i, s), s 6= i V {r, s} = otherwise U1{r, s} V1{i, s} 0 0 I where V is the matrix defined by  if (r, s) = (i, i) 0(cid:17) if (r, s) = (r, i), r 6= i 0 0 1 I 0 0 0 0 0 0 0 0 V1{r, s} otherwise   By Lemma 9.1 and Lemma 9.2, we have that Note that eV ∈ GLP (n′, Z). Note that det(eU {i}) = 1 and det(eV {i}) = 1; moreover, det(eU {j}) = det(U1{j}) and det(eV {j}) = det(V1{j}), for all j 6= i. Since E2 is in canonical form, by 2 has the property that for every vertex v, w ∈ (E′ Lemma 9.1, E′ there exists a path in E′ (E1)−eV = B• eU B• 2 from v to w through regular vertices in E′ reg with v ≥ w, 2. It is clear 2)0 E ′ 2 . GEOMETRIC CLASSIFICATION OF UNITAL GRAPH C ∗-ALGEBRAS 37 from the construction of (E1)− that for all regular vertices v, w of (E1)− satisfying v ≥ w, we have that there exists a path in (E1)− from v to w through regular vertices of (E1)− (since E1 has this property). have a Smith normal form with at least two 1's, so E′ Since E2 is in canonical form and by the definition of the outsplitting graph, we have that E′ 2 satisfies (1), (2) and (3) of Theorem 7.3. Similarly, (E1)− will satisfy- ing the same properties since E1 is in canonical form. Furthermore, the canonical form of Ek and the construction of Lemma 9.1 implies that the diagonal blocks of B• k also satisfies (5) of E ′ k Theorem 7.3. By Remark 7.5, there exist graphs F1, F2 in canonical form such that mF1 = m(E1)− = m′, nF1 = n(E1)− = n′, F1 ∼M (E1)−, and F2 ∼M E′ 2. Moreover, there exist SLP -equivalences (W1, Z1) : B• and (W2, Z2) : B• F2 → B• E ′ 2 , W1 ∈ SLP (m′, Z) and since Z1, Z −1 2 ∈ SLP (n′, Z), we have that det(U2{i}) = det(V2{i}) = 1, and det(U2{j}) = det(U1{j}) and det(V2{j}) = det(V1{j}) for all j 6= i. By construction, the pair (F1, F2) is in standard form with B• F2 , and Fk Fk ∼M ′ Ek. (cid:3) 2 eU W1 and V2 = Z1eV Z −1 ∈ MP (m′ × n′, Z), U2B• F1 → B• (E1)− 2 . Since W −1 2 Set U2 = W −1 F1 V2 = B• . We now have all we need to modify a GLP -equivalence to an SLP -equivalence. Theorem 9.5. Let E1 and E2 be graphs with finitely many vertices such that the pair (E1, E2) is in standard form. Suppose (U, V ) is a GLP -equivalence from B• E1 to B• E2 satisfying that V {i} = 1 whenever ni = 1. Then there exist graphs F1 and F2 such that Ei ∼M ′ Fi, the pair (F1, F2) is in standard form, and B• F1 is SLP -equivalent to B• F2 . Proof. The theorem follows from an argument similar to the argument in [Res06, Theorem 6.8] with Propositions 9.3 and 9.4 in place of [Res06, Lemma 6.7]. Briefly, the idea is that we are given a GLP -equivalence, say (U, V ). We go down the diagonal blocks and for each of them use Proposition 9.3 if necessary to make sure the U has positive determinant. Then we go down the diagonal blocks again this time using Proposition 9.4 to fix the determinant of the diagonal blocks of V when necessary. (cid:3) 10. Generalization of Boyle's positive factorization method In [Boy02], Boyle proved several factorization theorems for square matrices. These theorems are the key components to go from SLP -equivalence to flow equiv- alence. In this section, we prove similar factorization theorems for rectangular matrices. This is our key technical result to go from SLP -equivalence to move equivalence. Although the assumptions might seem restrictive, every unital graph C∗-algebra is move equivalent to another unital graph C∗-algebras whose adja- cency matrix satisfy the assumptions of the factorization theorem. The proof for rectangular matrices will closely follow the proof in [Boy02] for square matrices. First we introduce a new equivalence called "positive equivalence" of two matrices P (m × n, Z) (see Definition 10.1) and show that if ni 6= 0 for all i, then two P (m × n, Z) that are SLP -equivalent are positive equivalent. in M+ matrices in M+ Definition 10.1. Define M+ satisfying the following: P (m × n, Z) to be the set of all B ∈ MP (m × n, Z) (i) If i (cid:22) j and B{i, j} is not the empty matrix, then B{i, j} > 0. (ii) If B{i} is not the empty matrix, then B{i} > 0, the Smith normal form of B{i} has at least two 1's, and ni, mi ≥ 3. Note that condition (ii) implies that the row rank of every non-empty diagonal block is at least 2. In most of what follows, this will suffice for our purposes, but the stronger condition is needed to apply Theorem 10.7 below. 38 SØREN EILERS, GUNNAR RESTORFF, EFREN RUIZ, AND ADAM P. W. SØRENSEN Let B, B′ ∈ M+ P (m × n, Z). An SLP -equivalence (U, V ) : B → B′ is said to be a positive equivalence if U has a factorization of basic elementary matrices in SLP (m, Z) and V has a factorization of basic elementary matrices in SLP (n, Z) such that when applying these basic elementary matrices at each step we get matrices in M+ P (m × n, Z) (recall from [Boy02] that a basic elementary matrix is a matrix that is equal to the identity matrix except for on one offdiagonal entry, where it is either 1 or −1). We denote a positive equivalence by B (U,V ) + / B′ . Note that every element U ∈ SLP (n, Z) has a factorization of basic elementary matrices in SLP (n, Z). Therefore, a positive equivalence (U, V ) : B → B′ is an SL-equivalence that allows one to stay in M+ P (m × n, Z) for some factorization of U and V . 10.1. Factorization: Positive case. In this section, we prove a factorization theorem similar to that of [Boy02, Theorem 5.1] for positive rectangular matrices. The proof is imitating the proof in [Boy02] for square matrices. Definition 10.2. By a signed transposition matrix, we mean a matrix which is the matrix of a transposition, but with one of the offdiagonal 1's replaced by −1. By a signed permutation matrix we mean a product of signed transposition matrices. Note that for K > 1, any K × K permutation matrix with determinant 1 is a signed permutation matrix. A K × K matrix S is a signed permutation matrix if and only if det(S) = 1 and the matrix S is a permutation matrix (where S(i, j) := S(i, j)). For B, B′ ∈ M+(m × n, Z), we say an equivalence (U, V ) : B → B′ is a positive equivalence through M+(m×n, Z) if it can be given as a chain of positive elementary equivalences B = B0 → B1 → B2 → · · · → Bk = B′ in which every Bi is in M+(m × n, Z) (recall from [Boy02] that an equivalence (U, V ) is an elementary equivalence if one of U and V is a basic elementary matrix and the other is the identity matrix). Investigating the proof of [Boy02, Lemma 5.3 and Lemma 5.4] one can see the proofs also hold for rectangular matrices. Thus, we have the following lemmas. [Boy02, Lemma 5.3]). Suppose B ∈ M+(m × n, Z), E is a Lemma 10.3 (cf. basic elementary matrix with nonzero offdiagonal entry E(i, j), and the ith row of EB is not the zero row. Then there exists Q ∈ SL(n, Z) that is a product of nonnegative basic elementary matrices and there exists a signed permutation matrix S ∈ SL(m, Z) such that (SE, Q) : B → SEBQ is a positive equivalence through M+(m × n, Z). Lemma 10.4 (cf. [Boy02, Lemma 5.4]). Let B be an element of M(K1 × K2, Z) for K1, K2 ≥ 3 such that the row rank of B is at least 2. Suppose U ∈ SL(K1, Z) such that no row of B and U B is the zero row. Then U is the product of elementary matrices U = Ek · · · E1 such that for 1 ≤ j ≤ k the matrix Ej Ej−1 · · · E1B has no zero rows. The following lemma is inspired by the reduction step in the proof of [Boy02, Lemma 5.5]. We give the entire proof for the convenience of the reader. Lemma 10.5. Let B ∈ M+(K1 × K2, Z) with K1, K2 ≥ 3. Suppose the row rank of B is at least 2 and there exists U ∈ SL(K1, Z) such that U B > 0. Then the equivalence (U, IK2) : B → U B is a positive equivalence through M+(K1 × K2, Z). / GEOMETRIC CLASSIFICATION OF UNITAL GRAPH C ∗-ALGEBRAS 39 Proof. By Lemma 10.4, we can write U as a product of basic elementary matrices U = EkEk−1 · · · E1, such that for 1 ≤ j ≤ k, the matrix Ej · · · E1B has no zero row. By Lemma 10.3, given the pair (E1, B), there is a nonnegative Q1 which is a product of nonnegative basic elementary matrices and a signed permutation S1 such that (S1E1, Q1) : B → S1E1BQ1 is a positive equivalence through M+(K1 ×K2, Z). Note that U BQ1 = S−1 1 [S1EkS−1 1 ] · · · [S1E2S−1 1 ][S1E1]BQ1. 1 j. Since E′ Now, for 2 ≤ j ≤ k, the matrix S1Ej S−1 is again a basic elementary matrix E′ 2(S1E1BQ1) = S1Ej · · · E2E1BQ1 for 2 ≤ j ≤ k and since Ej · · · E2E1BQ1 has no zero rows, and S1 is a signed permutation, we have that E′ 2(S1E1BQ1) has no zero rows for all 2 ≤ j ≤ k. j · · · E′ j · · · E′ Using Lemma 10.3, for the pair (S1E2S−1 1 , S1E1BQ1), we get a signed permu- tation matrix S2 and a nonnegative Q2 which is a product of nonnegative basic elementary matrices such that (S2[S1E2S−1 1 ], Q2) : S1E1BQ1 → S2[S1E2S1]−1S1E1BQ1Q2 is a positive equivalence through M+(K1 × K2, Z). Thus, we get a positive equiv- alence through M+(K1 × K2, Z) ([S2S1E2S−1 1 ][S1E1], Q1Q2) : B → S2S1E2E1BQ1Q2 and we observe that U BQ1Q2 = S−1 1 S−1 · · · [S2S1E3S−1 2 [S2S1EkS−1 1 S−1 1 S−1 2 ][S2S1E2S−1 2 ] · · · 1 ][S1E1]BQ1Q2. Continue this, to obtain a signed permutation matrix S = Sk · · · S1 and a nonneg- ative matrix Q = Q1Q2 · · · Qk that is a product of nonnegative basic elementary matrices such that U BQ = S−1[Sk · · · S1EkS−1 and (SU, Q) : B → SU BQ is a positive equivalence through M+(K1 × K2, Z). 1 ][S1E1]BQ = S−1(SU BQ) k−1] · · · [S2S1E2S−1 · · · S−1 1 We claim that the equivalence (S, IK2 ) : U BQ → SU BQ is a positive equivalence through M+(K1 × K2, Z). Since S is a product of signed transposition matrices, it may be described as a permutation matrix in which some rows have been multiplied by −1. Since U BQ and SU BQ are strictly positive, it must be that S is a per- mutation matrix. Also, det(S) = 1, so if S 6= IK1, then S is a permutation matrix which is a product of 3-cycles. So it is enough to realize the positive equivalence through M+(K1 × K2, Z) in the case that S is the matrix of a 3-cycle. For this we write the matrix 0 0 1 1 0 0 1 C = 0 0 1  0 1 1 0 1 0 1 0 0 1 0 1 1 0 as the following product C0C1C2C3C4C5: 0 −1 1 0 0  0 −1 1 0 1 1 0 0 −1 1 0 0 1 0 0 0 1 0 0 0 0 1 0 0 1 1 0 0 1 0 0 0 1 1 . For 0 ≤ i ≤ 5, the matrix CiCi+1 · · · C5 is nonnegative and has no zero row. Therefore, the equivalence (C, I) : D → CD is a positive equivalence through M+(K1 × K2, Z) whenever D ∈ M+(K1 × K2, Z). Therefore, (S, IK2 ) : U BQ → SU BQ is a positive equivalence through M+(K1 ×K2, Z) proving the claim. There- fore, (S−1, IK2 ) : SU BQ → U BQ is a positive equivalence through M+(K1 × K2, Z). Since Q is the product of nonnegative basic elementary matrices and 40 SØREN EILERS, GUNNAR RESTORFF, EFREN RUIZ, AND ADAM P. W. SØRENSEN U B ∈ M+(K1 × K2, Z), the equivalence (IK1 , Q) : U B → U BQ is a positive equiv- alence through M+(K1 × K2, Z). Thus, (IK1 , Q−1) : U BQ → U B is a positive equivalence through M+(K1 × K2, Z). Now the composition of positive equiva- lences through M+(K1 × K2, Z) B (SU,Q) + / SU BQ (S−1,IK2 ) + / U BQ (IK1 ,Q−1) + / U B is positive equivalence through M+(K1 ×K2, Z) but the composition of these equiv- alences is equal to the equivalence (U, IK2 ) : B → U B. Hence, the equivalence (U, IK2) : B → U B is a positive equivalence through M+(K1 × K2, Z). (cid:3) The proof of the next lemma is similar to the proof of [Boy02, Lemma 5.5]. Since there are some differences between the two proofs we provide the entire argument. Lemma 10.6 (cf. [Boy02, Lemma 5.5]). Let B and B′ be elements of M+(K1 × K2, Z) with K1, K2 ≥ 3, and the rank of B and B′ at least 2. Suppose U ∈ SL(K1, Z) and W ∈ SL(K2, Z) such that U B has at least one strictly positive entry and U B = B′W . Then the equivalence (U, W −1) : B → B′ is a positive equivalence through M+(K1 × K2, Z). Proof. We will first reduce to the case that U B > 0. By assumption (U B)(i, j) > 0 for some (i, j). We can repeatedly add column j to other columns of until row i of U B has all entries strictly positive. This corresponds to multiplying from the right by a nonnegative matrix Q in SL(K2, Z), where Q is the product of nonnegative basic elementary matrices, giving U BQ = B′W Q. Then we can repeatedly add row i of U BQ to other rows until all entries are positive. This corresponds to multiplying from the left by a nonnegative matrix P in SL(K1, Z), where P is the product of nonnegative basic elementary matrices, giving (P U )(BQ) = (P B′)(W Q) > 0. We also have positive equivalences through M+(K1 × K2, Z) given by (I, Q) : B → BQ and (P, I) : B′ → P B′. Note that the equivalence (U, W −1) : B → B′ is the composition of equiva- lences, (I, Q) : B → BQ followed by (P U, (W Q)−1) : BQ → P B′ followed by (P −1, I) : P B′ → B′. Since (I, Q) : B → BQ and (P −1, I) : P B′ → B′ are positive equivalences through M+(K1 × K2, Z), it is enough to show that the equivalence (P U, (W Q)−1) : BQ → P B′ is a positive equivalence through M+(K1 × K2, Z). Therefore, after replacing (U, B, B′, W ) with (P U, BQ, P B′, W Q), we may assume without loss of generality that U B > 0. By Lemma 10.5, the equivalence (U, IK2) : B → U B is a positive equivalence through M+(K1 × K2, Z). Therefore, by Lemma 10.5, ((W )T , IK1 ) : (B′)T → W T (B′)T is a positive equivalence through M+(K2×K1, Z) which implies the equiv- alence (IK1 , W ) : B′ → B′W is a positive equivalence through M+(K1 × K2, Z). Thus, the equivalence (IK1 , W −1) : B′W → B′ is a positive equivalence through M+(K1 × K2, Z). Since the equivalence (U, W −1) : B → B′ is the composition of positive equivalences: (U, IK2) : B → U B followed by (IK1 , W −1) : B′W → B′, the equivalence (U, W −1) : B → B′ a positive equivalence through M+(K1×K2, Z). (cid:3) Theorem 10.7 (cf. [Boy02, Theorem 5.1]). Let K1, K2 ≥ 3 and let B ∈ M+(K1 × K2, Z). Suppose U ∈ SL(K1, Z) and V ∈ SL(K2, Z) such that U BV ∈ M+(K1 × K2, Z) and suppose that X ∈ SL(K1, Z) and Y ∈ SL(K2, Z) such that XBY = 1 0 0 1 0 0 F . / / / GEOMETRIC CLASSIFICATION OF UNITAL GRAPH C ∗-ALGEBRAS 41 Then the equivalence (U, V ) : B → U BV is a positive equivalence through M+(K1 × K2, Z). Proof. Note that for any H ∈ SL(2, Z), the K1 × K1 matrix GH,1 and K2 × K2 matrix GH,2, given by GH,1 =(cid:18)H 0 0 IK1−2(cid:19) and GH,2 =(cid:18)H 0 H,2Y −1) : B → B. 0 IK2−2(cid:19) , give a self-equivalence (X −1GH,1X, Y G−1 c1 ... cK1 For a matrix Q, we let Q(12; ∗) denote the submatrix consisting of the first two rows. Since (XBY )(12; ∗) has rank 2 and Y is invertible, we have that (XB)(12; ∗) has rank two. Therefore, there exists H ′ ∈ SL(2, Z) such that the first row r = (cid:0)r1, . . . , rK2(cid:1) of H ′[(XB)(12; ∗)] has both a positive entry and a negative entry.  denote the first column of X −1, and note that it is nonzero. Since Let c = and a negative entry. For each m ∈ N, set Hm =(cid:18)m −1 0 (cid:19) H ′. Choose m large cr is the K1×K2 matrix with (i, j) entry equal to cirj , we have that cr has a positive enough such that the entries of the two matrices X −1GHm,1XB and mcr will have the same sign wherever the entries of mcr are nonzero. In particular, X −1GHm,1XB will have a positive entry. By Lemma 10.6, (X −1GHm,1X, Y G−1 Hm,2Y −1) : B → B gives a positive equivalence through M+(K1 × K2, Z). 1 Similarly for large enough m, the entries of U X −1GHm,1XB will agree in sign with the entries U cr whenever the entries of the latter matrix are nonzero. Since U is invertible, the matrix U cr is nonzero, and thus contains positive and negative entries, because r does. Therefore, U X −1GHm,1XB contains a positive entry. By Lemma 10.6, (U X −1GHm,1X, Y G−1 Hm,2Y −1V ) : B → B′ gives a positive equivalence through M+(K1 × K2, Z) with B′ = U BV . Hence, the equivalence (U, V ) : B → B′ is a positive equivalence through M+(K1 × K2, Z) since it is the composition of positive equivalences through M+(K1 × K2, Z): (X −1G−1 Hm,1X, Y GHm,2Y −1) : B → B followed by (U X −1GHm,1X, Y G−1 Hm,2Y −1V ) : B → B′ (cid:3) 10.2. Factorization: General case. We now use the results of the previous sec- tion to prove a factorization for general B, B′ ∈ M+ P (m × n, Z) with ni 6= 0 that are SLP -equivalent. Again, many of the arguments follow the arguments of Boyle in [Boy02]. Lemma 10.8 (cf. [Boy02, Lemma 4.6]). Let B, B′ ∈ M+ P (m × n, Z) with ni 6= 0 for all i. If (U, V ) : B → B′ is an SLP -equivalence such that U {i} and V {j} are the identity matrices of the appropriate size whenever they are not the empty matrix, then (U, V ) : B → B′ is a positive equivalence. Proof. We will first find Q in SLP (n, Z) which is a product of nonnegative basic elementary matrices such that (U, Q) : B → U BQ is a positive equivalence. We may assume that U is not the identity matrix. Factor U = Un · · · U1 where for each Ut there is an associated pair (it, jt) such that the following hold • Ut = I except in the block Ut{it, jt}, where it is nonzero • if s 6= t, then (is, js) 6= (it, jt). 42 SØREN EILERS, GUNNAR RESTORFF, EFREN RUIZ, AND ADAM P. W. SØRENSEN 1 U + 1 and U + 1 , where U − Factor U1 = U − U − 1 {i1, j1} is the nonpositive part of U1{i1, j1} and U + part of U1{i1, j1}. Note that U + 1 matrices in SLP (m, Z) and U − in SLP (m, Z). It is now clear that (U + 1 are equal to I outside the block {i1, j1}, 1 {i1, j1} is the nonnegative is a product of nonnegative basic elementary 1 is a product of nonpositive basic elementary matrices 1 B is a positive equivalence. 1 , I) : B → U + Now, note that U − 1 B){i1} > 0 since (U + 1 B outside the blocks {i1, k} such that i1 ≺ j1 (cid:22) k. Also note that mi1 6= 0 (since U1 6= I). Since ni 6= 0 for all i, we have that B{i1} is not the empty matrix. Therefore, B{i1} > 0 since B ∈ M+ P (m × n, Z). Hence, (U + 1 B is a positive equivalence. We can now add columns of (U + 1 B){i1, k} for all i1 ≺ j1 (cid:22) k enough times to obtain a Q1 which is a product of nonnegative basic elementary matrices in SLP (n, Z) such that (U − 1 BQ1 is a positive equivalence. Since (U1, Q1) : B → U1B is the composition of positive equivalences 1 B){i1} to columns of (U + 1 , I) : B → U + 1 , Q1) : U + 1 B → U − 1 B = U + 1 U + 1 U + B (U + 1 ,I) + / U + 1 B (U − 1 ,Q1) + / U1BQ1 we get that the equivalence (U1, Q1) : B → U1BQ1 is a positive equivalence. Repeat the process for the matrices U1BQ1 and U2U1BQ1, we get Q2 which is the product of nonnegative elementary matrices in SLP (n, Z) such that the equiv- alence (U2, Q2) : U1BQ1 → U2U1BQ1Q2 is a positive equivalence. We continue this process to get Qi that is the product of nonnegative elementary matrices in SLP (n, Z) for 1 ≤ i ≤ n such that (U, Q) : B → U BQ is a positive equivalence, where Q = Q1 · · · Qn. We now show that there exists P that is a product of nonnegative basic el- ementary matrices in SLP (m, Z) such that (P, V −1) : B′ → P B′V −1 is a posi- tive equivalence. Throughout the rest of the proof, if M ∈ M+ P (m × n, Z), then M {{1, 2, . . . , i}} will denote the block matrix whose {s, r} block is M {s, r} for all 1 ≤ s, r ≤ i. First note that there are matrices V2, . . . , VN in SLP (n, Z) such that V −1 = V2V3 · · · VN , each Vi is the identity matrix except for the blocks Vi{l, i}, and V2 · · · Vi =(cid:18)V −1{{1, . . . , i}} 0 I(cid:19) . 0 i {l, i} is the nonpositive part of Vi{l, i} and let V + i be the matrix in SLP (n, Z) that is the identity matrix except for the blocks i be the matrix in i {l, i} is is equal to the identity matrix i {l, i} + V − i {l, i} = Vi{l, i}. Let V − Vi{l, i} and V − SLP (n, Z) that is the identity matrix except for the blocks Vi{l, i} and V + the nonnegative part of Vi{l, i}. Note that V + except for the blocks Vi{l, i} and (V + Therefore, Vi = V + i ){l, i} = V + i V − i V − i i V − i We will inductively construct matrices P2, P3, . . . , PN in SLP (m, Z) such that each Pi is the product of nonnegative basic matrices such that each Pi is the iden- tity outside of the blocks {l, i} for l ≺ i and for each 2 ≤ i ≤ N , we have that (Pi, Vi) : Pi−1 · · · P2B′V2 · · · Vi−1 → Pi · · · P2B′V2 . . . Vi is a positive equivalence. Note that if we have constructed Pi, then the composition of these positive equiva- lences gives a positive equivalence (P, V −1) : B′ → P B′V −1, where P = Pk · · · P2. Thus, the lemma holds. 2 = V − We now prove the claim. We first construct P2. Note that if 1 is not a predecessor of 2, then V + 2 = I. Therefore, (I, V2) : B′ → B′V2 is a positive equivalence. Suppose 1 (cid:22) 2. Suppose m1 = 0. Then B′V + 2 = B′ which implies that (I, V − 2 → B′V2 is a positive equivalence. So, (I, V2) : B′ → B′V2 is a positive equivalence since it is the composition of the positive equivalences (I, V + 2 ) and (I, V − 2 ). Suppose m1 6= 0. In this situation, we have two cases, m2 6= 0 and m2 = 0. 2 = B′V − 2 ) : B′V + 2 V − / / GEOMETRIC CLASSIFICATION OF UNITAL GRAPH C ∗-ALGEBRAS 43 2 V − 2 Suppose m2 6= 0. Note that B′V + 2 > 0 since B′ > 0 and (I, V + is equal to B′ except for the {1, 2} block. We have that B′V + 2 ) : B′ → B′V + is a positive 2 equivalence. Hence, we may add rows of (B′V + 2 ){2} to rows of (B′V + 2 ){1, 2} to get a matrix P2 in SLP (m, Z) that is the product of nonnegative basic matrices and is the identity outside of the block {1, 2} such that (P2, V − 2 → B′V2 is a positive equivalence. Composing the positive equivalences (I, V + 2 ), we get a positive equivalence (P2, V2) : B′ → P2BV2. 2 ) and (P2, V − 2 ) : B′V + Suppose m2 = 0. Then (B′V2){1, 2} = B′{1}V2{1, 2} + B′{1, 2}V2{2} = B′{1}V −1{1, 2} + B′{1, 2} = (B′V −1){1, 2}, since V2{{1, 2}} = V −1{{1, 2}} and V2{2} = V −1{2} = I. Therefore, (B′V + 2 )V − 2 ){1, 2} = (B′V −1){1, 2} = (U B){1, 2} = U {1}B{1, 2} + U {1, 2}B{2} = B{1, 2} > 0, since U {1, 2} is the empty matrix and U {1} = I. Therefore (I, V − is a positive equivalence and by composing the positive equivalences (I, V + (I, V − 2 ), we get a positive equivalence (I, V2) : B′ → B′V2. 2 ) : B′V + 2 → B′V2 2 ) and So, in all cases, we have found a matrix P2 in SLP (m, Z) that is the product of nonnegative basic elementary matrices and is the identity outside of the block {1, 2} such that (P2, V2) : B′ → P2B′V2 is a positive equivalence. Let 2 ≤ n ≤ N − 1 and suppose we have constructed P2, P3, . . . , Pn in SLP (m, Z) such that each Pi is the product of nonnegative basic matrices and Pi is the identity outside of the blocks {l, i} with l ≺ i and for each 2 ≤ i ≤ n, we have that (Pi, Vi) : Pi−1 · · · P2B′V2 · · · Vi−1 → Pi · · · P2B′V2 . . . Vi is a positive equivalence. To simplify the notation, we set B′ i = Pi · · · P2B′V2 . . . Vi. Since B′ a positive equivalence (I, V + to B′ n except for the blocks {i, n + 1} with i (cid:22) n + 1. n+1) : B′ n → B′ nV + n+1. Note that B′ nV + n+1V − nV + Suppose mn+1 6= 0. Then(B′ nV + n+1){n + 1} to rows of (B′ n+1){n + 1} > 0. Hence, we may add rows nV + of (B′ n+1){i, n + 1} for all i ≺ n + 1, to obtain a matrix Pn+1 in SLP (m, Z) which is the product of nonnegative basic matri- ces and is the identity outside of the blocks {i, n + 1} for i ≺ n + 1 such that (Pn+1, V − nVn+1 is a positive equivalence. Composing the positive equivalences (I, V + n → Pn+1B′ n+1), we get that (Pn+1, Vn+1) : B′ nVn+1 is a positive equivalence. n+1) and (Pn+1, V − n+1 → Pn+1B′ n+1) : B′ nV + Suppose mn+1 = 0. Let I = {i0, . . . , it} be the set of elements is ∈ P that satisfy is (cid:22) n + 1, mis 6= 0, and if is ≺ l (cid:22) n + 1, then ml = 0. Note that for all distinct is, ir ∈ I, is is not a predecessor of ir. Note that if I = ∅, then B′ n+1 = B′ nVn+1 is a positive equivalence and hence (I, Vn+1) : B′ n. This would imply that (I, V − nVn+1 is a positive equivalence. n+1 → B′ n+1 = B′ n+1) : B′ n+1V − n → B′ nV − nV + nV + n > 0, we get n+1 is equal Suppose I 6= ∅. Note that for each is ∈ I, (B′ nVn+1){is, n + 1} = Xis(cid:22)l(cid:22)n+1 = Xis(cid:22)l(cid:22)n+1 (Pn · · · P2B′){is, l}(V2 . . . VnVn+1){l, n + 1} (Pn · · · P2B′){is, l}V −1{l, n + 1} = ((Pn · · · P2B′)V −1){is, n + 1} 44 SØREN EILERS, GUNNAR RESTORFF, EFREN RUIZ, AND ADAM P. W. SØRENSEN since (V2 . . . VnVn+1){{1, . . . n + 1}} = V −1{{1, . . . , n + 1}}. Since Pn · · · P2U B = Pn · · · P2B′V −1, ((Pn · · · P2B′)V −1){is, n + 1} = ((Pn · · · P2U )B){is, n + 1} = Xis(cid:22)l(cid:22)n+1 (Pn · · · P2U ){is, l}B{l, n + 1}. Using the fact that ml = 0 for all is ≺ l (cid:22) n + 1 and (Pn · · · P2U ){is} = I, we get that (B′ nVn+1){is, n + 1} = B{is, n + 1}. nV + nV + nVn+1){is, n + 1} = B{is, n + 1} > 0 because mis 6= 0 and B ∈ Moreover, (B′ M+ P (m × n, Z). For each l ≺ n+1, there exists an s such that l (cid:22) is. Recall that (B′ n+1){is, n+1} to rows of (B′ n+1){is, n+ nV + 1} > 0, so if l ≺ is, we may add rows of (B′ n+1){l, n+ 1}, to get a matrix P l n+1 in SLP (m, Z) that is the product of nonnegative ba- sic elementary matrices and is the identity outside of the block {l, n + 1} such that (P l nVn+1){l, n + 1} > 0. Doing this for all l ≺ n + 1, we get a matrix Pn+1 in SLP (m, Z) that is the product of nonnegative basic elementary matri- ces and is the identity outside of the blocks {l, n + 1} for l ≺ n + 1 such that (Pn+1, V − nVn+1 is a positive equivalence. Composing the positive equivalences (I, V + n → Pn+1B′ n+1), we get that (Pn+1, Vn+1) : B′ nVn+1 is a positive equivalence. n+1B′ n+1) : B′ nV + n+1 → Pn+1B′ n+1) and (Pn+1, V − In all cases, we get a matrix Pn+1 in SLP (m, Z) that is the product of nonnegative basic elementary matrices and is the identity outside of the blocks {l, n + 1} for l ≺ n + 1 such that (Pn+1, Vn+1) : B′ nVn+1 is a positive equivalence. The claim now follows by induction. (cid:3) n → Pn+1B′ The next lemma allows us to reduce the general case to the case that the diagonal blocks U {i} and V {j} are the identity matrices of the appropriate sizes when they are not the empty matrices. This will allow us to use Lemma 10.8 to get the desired positive equivalence. Lemma 10.9 ([Boy02, Lemma 4.9]). Let B, B′ ∈ M+ all l. Fix i with mi 6= 0. P (m × n, Z) with nl 6= 0 for (1) Suppose E is a basic elementary matrix in SLP (m, Z) such that E{j, k} = I{j, k} when (j, k) 6= (i, i) and (E{i}, I) : B{i} → B′{i} is a positive equivalence. Then there exists V ∈ SLP (n, Z) which is the product of nonnegative basic elementary matrices in SLP (n, Z) such that V {k} = I for all k and (E, V ) : B → EBV is a positive equivalence. (2) Suppose E is a basic elementary matrix in SLP (n, Z) such that E{j, k} = I{j, k} when (j, k) 6= (i, i) and (I, E{i}) : B{i} → B′{i} is a positive equivalence. Then there exists U ∈ SLP (n, Z) which is the product of nonnegative basic elementary matrices in SLP (n, Z) such that U {k} = I for all k and (U, E) : B → U B′E is a positive equivalence. GEOMETRIC CLASSIFICATION OF UNITAL GRAPH C ∗-ALGEBRAS 45 Proof. We prove (1). The proof of (2) is similar. Let E(s, t) be the nonzero offdi- agonal entry of E. If E(s, t) = 1, then set V = I. Suppose E(s, t) = −1. So, E acts from the left to subtract row t from row s. Since mi 6= 0 and (E{i}, I) : B{i} → B′{i}, is a positive equivalence, we have that (EB){i} > 0. Thus, there exists r an index for a column through the {i, i} block such that B′(s, r) > B′(t, r). Let V be the matrix in SLP (n, Z) which acts from the right to add column r to column q, M times, for every q indexing a column through an {i, j} block for which i ≺ j. Choos- ing M large enough, we have that (E, I) : B′V → EB′V is a positive equivalence. Therefore, (E, V ) : B → EBV is a positive equivalence since it is the composition of two positive equivalences: (I, V ) : B → BV followed by (E, I) : BV → EBV . (cid:3) We are now ready to prove the main result of this section. This result will be used to show that if the adjacency matrices of E and F are SLP -equivalent, then E is move equivalent to F . Consequently, C∗(E) is Morita equivalent to C∗(F ). Theorem 10.10 ([Boy02, Theorem 4.4]). Let B, B′ ∈ M+ P (m × n, Z) with ni 6= 0 for all i. Suppose there exist U ∈ SLP (m, Z) and V ∈ SLP (n, Z) such that U BV = B′. Then (U, V ) : B → B′ is a positive equivalence. Proof. By Theorem 10.7, for each i with mi 6= 0, we have that (U {i}, V {i}) : B{i} → B′{i} is a positive equivalence since by (ii) of Definition 10.1, a 2 × 2-submatrix can be extracted as stipulated. So, we may find a string of elementary equivalences say (E1, F1), . . . , (Et, Ft), with every Et{i, j} = I, Ft{i, j} = I unless i = j with mi 6= 0, which accomplishes the elementary positive equivalences decomposition inside the diagonal blocks. By Lemma 10.9, we may find (U1, V1), . . . , (Ut, Vt) such that Us ∈ SLP (m, Z), Vs ∈ SLP (n, Z), Us{k} = I, Vs{k} = I, and such that we have the following positive equivalences B (U1,F1) / · + (E1,V1) / · · · (Ut,Ft) / · + + (Et,Vt) / B′′. + Let X = EtUt · · · E2U2E1U1 and Y = F1V1F2V2 · · · FtVt. Then for all i, we have that X{i} = U {i} and Y {i} = V {i}. Therefore, (U X −1){i} = I and (Y −1V ){i} = I for all i. Then by Lemma 10.8, B′′ (U X −1,Y −1V ) + / B′ is a positive equivalence. Thus, (U, V ) : B → B′ is a positive equivalence since it is the composition of two positive equivalences B (X,Y ) + / B′′ (U X −1,Y −1V ) + / B′. (cid:3) 11. Putting it all together/Proof of main theorem Theorem 11.1. Let E1 and E2 be graphs with finitely many vertices satisfying Condition (K) and assume that FK+ R(C∗(E2)). Let F1 and F2 be chosen according to Proposition 7.9. Then there exists a GLP -equivalence (U, V ) from B• F2 that satisfies that V {i} is the identity matrix whenever ni = 1. R(C∗(E1)) ∼= FK+ F1 to B• Proof. As usual, we define P, m and n according to the matrices B• F2 so that it reflects the ideal structure of the associated C∗-algebras. Here, B• F2 ∈ MP (m × n, Z). We let P T denote the set P with the opposite order defined by i (cid:22) j in P T if and only if N + 1 − j (cid:22) N + 1 − i in P, for i = 1, 2, . . . , N . Moreover, we let mT = (mN , . . . , m2, m1) and nT = (nN , . . . , n2, n1), we let m = m1 + m2 + · · · + mN and n = n1 + n2 + · · · + nN , and we let Jm and F1 to B• F1 , B• / / / / / / / 46 SØREN EILERS, GUNNAR RESTORFF, EFREN RUIZ, AND ADAM P. W. SØRENSEN Jn be the m × m respectively n × n permutation matrix that reverses the order. Jm ∈ MP T(nT × mT, Z). So in a similar way as in the proof of [Res06, Proposition 8.3], where we use [CET12, Theorem 4.1 and Remark 4.2] in the place of [Res06, Proposition 3.4], we see that this ordered fil- F2(cid:1)T Jm, Jn(cid:0)B• F1(cid:1)T Then Jn(cid:0)B• tered K-theory isomorphism induces a K-web isomorphism from K(cid:0)Jn(B• F1 )TJm(cid:1) to K(cid:0)Jn(B• F2 )TJm(cid:1). When ni = 1, positivity implies that the isomorphism from cok(cid:0)Jn(B• F1 )TJm(cid:1) to cok(cid:0)Jn(B• that(cid:0)JmV TJm, JnU TJn(cid:1) is a GLP -equivalence from B• (cid:0)JnU TJn(cid:1) {i} is the identity matrix whenever ni = 1. F1 )TJm to Jn(B• F2 )TJm that induces exactly this K-web isomorphism. Note that U {i} is the identity matrix whenever nN +1−i = 1. As in [Res06, Remark 8.2], we see F2 that satisfies that (cid:3) F2 )TJm(cid:1) is the identity map. Now we use Theorem 8.12 to get a GLP T-equivalence (U, V ) from Jn(B• F1 to B• Proof of Theorem 3.1. (1) =⇒ (2): It follows from Theorem 2.17 that the moves (O), (I), (R), (S) preserve stable isomorphism. By Proposition 5.8 (C) also pre- serves stable isomorphism so ∼M ′ preserves stable isomorphism. (2) =⇒ (3): Holds in general. (3) =⇒ (1): Suppose we are given graphs E1, E2 that satisfy Condition (K) and have finitely many vertices. By Theorem 11.1 we can find graphs F1 and F2 with finally many vertices such that E1 ∼M F1, E2 ∼M F2 and (F1, F2) are in standard form and there exists a GLP -equivalence (U, V ) from B• F2 that satisfies that V {i} is the identity matrix whenever ni = 1. Theorem 9.5 lets us find graphs G1, G2 in standard form such that G1 ∼M ′ F1, G2 ∼M ′ F2 and B• G2 are SLP -equivalent. By Theorem 10.10 this equivalence is a positive equivalence and so by Corollary 7.8 G1 ∼M G2. Thus we have G1 and B• F1 to B• E1 ∼M F1 ∼M ′ G1 ∼M G2 ∼M ′ F2 ∼M E2. (cid:3) Acknowledgements This work was partially supported by the Danish National Research Founda- tion through the Centre for Symmetry and Deformation (DNRF92), by VILLUM FONDEN through the network for Experimental Mathematics in Number The- ory, Operator Algebras, and Topology, by a grant from the Simons Foundation (# 279369 to Efren Ruiz), and by the Danish Council for Independent Research Natural Sciences. The third and fourth named authors would also like to thank the School of Mathematics and Applied Statistics at the University of Wollongong for hospitality during their visit where part of this work was carried out. The authors would also like to thank Mike Boyle for many fruitful discussions. References [AMP07] Pere Ara, M. Angeles Moreno, and Enrique Pardo, Nonstable K-theory for graph algebras, Algebr. Represent. Theory 10 (2007), no. 2, 157 -- 178, URL: http://dx.doi.org/10.1007/s10468-006-9044-z, doi:10.1007/s10468-006-9044-z. MR 2310414 (2008b:46094) Sara E. Arklint and Efren Ruiz, Corners of Cuntz-Krieger algebras, ArXiv e-prints (2012), to appear in Trans. Amer. Math. Soc, arXiv:1209.4336v3. [AR12] [ARR12] Sara Arklint, Gunnar Restorff, and Efren Ruiz, Filtrated K-theory for real rank zero C ∗-algebras, Internat. J. Math. 23 (2012), no. 8, 1250078, 19, URL: http://dx.doi.org/10.1142/S0129167X12500784 , doi:10.1142/S0129167X12500784. MR 2949216 [BCW14] Nathan Brownlowe, Toke Meier Carlsen, and Michael F. Whittaker, Graph algebras and orbit equivalence, ArXiv e-prints (2014), to appear in Ergodic Th. Dynam. Sys., arXiv:1410.2308. GEOMETRIC CLASSIFICATION OF UNITAL GRAPH C ∗-ALGEBRAS 47 [BD96] Lawrence G. Brown and Marius Dadarlat, Extensions of C ∗-algebras and qua- sidiagonality, J. London Math. Soc. (2) 53 (1996), no. 3, 582 -- 600, URL: http://dx.doi.org/10.1112/jlms/53.3.582 , doi:10.1112/jlms/53.3.582. MR 1396721 (97d:46086) [BH03] Mike Boyle and Danrun Huang, Trans. Amer. Math. matrices, (electronic), doi:10.1090/S0002-9947-03-02947-7 . MR 1990568 (2004f:15020) integral 3861 -- 3886 http://dx.doi.org/10.1090/S0002-9947-03-02947-7, block (2003), Poset 355 equivalence URL: Soc. no. 10, of [BHRS02] Teresa Bates, Jeong Hee Hong, Iain Raeburn, and Wojciech Szymański, The ideal structure of the C ∗-algebras of infinite graphs, Illinois J. Math. 46 (2002), no. 4, 1159 -- 1176, URL: http://projecteuclid.org/euclid.ijm/1258138472. MR 1988256 (2004i:46105) Rasmus Bentmann and Manuel Köhler, Universal coefficient theorems for C ∗-algebras over finite topological spaces, ArXiv e-prints (2011), arXiv:1101.5702v3. Rasmus Bentmann and Ralf Meyer, A more general method to classify up to equivariant KK-equivalence, ArXiv e-prints (2014), arXiv:1405.6512v1. [BM14] [BK11] [Boy02] Mike Boyle, Flow equivalence J. Math. Pacific 204 of finite of shifts torizations, http://dx.doi.org/10.2140/pjm.2002.204.273, MR 1907894 (2003f:37018) Teresa Bates graph algebras, Er- godic Theory Dynam. 2, URL: http://dx.doi.org/10.1017/S0143385703000348 , doi:10.1017/S0143385703000348. MR 2054048 (2004m:37019) and David Pask, Flow equivalence of 367 -- 382, type 2, via positive fac- URL: doi:10.2140/pjm.2002.204.273. 273 -- 317, Systems (2004), (2002), no. no. 24 [BP04] [CET12] Toke Meier Carlsen, Søren Eilers, and Mark Tomforde, Index maps in the K-theory of graph algebras, J. K-Theory 9 (2012), no. 2, 385 -- 406, URL: http://dx.doi.org/10.1017/is011004017jkt156 , doi:10.1017/is011004017jkt156. MR 2922394 [Eil96] [DG97] Marius Dadarlat and Guihua Gong, A classification result for approximately homo- geneous C ∗-algebras of real rank zero, Geom. Funct. Anal. 7 (1997), no. 4, 646 -- 711, URL: http://dx.doi.org/10.1007/s000390050023 , doi:10.1007/s000390050023. MR 1465599 (98j:46062) Søren Eilers, A complete zero and bounded torsion in K1, J. Funct. Anal. 139 (1996), no. 2, 325 -- 348, URL: http://dx.doi.org/10.1006/jfan.1996.0088 , doi:10.1006/jfan.1996.0088. MR 1402768 (97e:46096) George real URL: rank http://dx.doi.org/10.1515/crll.1993.443.179 , doi:10.1515/crll.1993.443.179. MR 1241132 (94i:46074) Elliott, J. Reine Angew. Math. for AD algebras with real classification 443 of (1993), A. zero, C ∗-algebras 179 -- 219, invariant [Ell93] rank On the of [ERR10] Søren Eilers, Gunnar Restorff, and Efren Ruiz, On graph C ∗-algebras with a linear ideal lattice, Bull. Malays. Math. Sci. Soc. (2) 33 (2010), no. 2, 233 -- 241. MR 2666426 (2012h:46086) [ERR13a] [ERR13b] , Classification of graph C ∗-algebras with no more than four primitive ideals, Operator algebra and dynamics, Springer Proc. Math. Stat., vol. 58, Springer, Hei- delberg, 2013, pp. 89 -- 129, URL: http://dx.doi.org/10.1007/978-3-642-39459-1_5, doi:10.1007/978-3-642-39459-1_5. MR 3142033 , Classifying sub- quotients, URL: http://dx.doi.org/10.1016/j.jfa.2013.05.006 , doi:10.1016/j.jfa.2013.05.006. MR 3056712 Funct. Anal. finite no. C ∗-algebras 449 -- 468, (2013), infinite with both and 265 J. 3, [ERR13c] , Strong classification of extensions of classifiable C ∗-algebras, ArXiv e-prints [ERS12] (2013), arXiv:1301.7695v1. Søren Eilers, Efren Ruiz, and Adam P. W. Sørensen, Amplified graph C ∗-algebras, Münster J. Math. 5 (2012), 121 -- 150. MR 3047630 [ERS15] , Geometric classification of C ∗-algebras over finite graphs, In preparation, [ET10] 2015. Søren Eilers nonsim- 393 -- 418, URL: ple http://dx.doi.org/10.1007/s00208-009-0403-z, doi:10.1007/s00208-009-0403-z. MR 2563693 (2010k:46072) graph C ∗-algebras, Math. Ann. 346 (2010), and Mark Tomforde, On classification no. the 2, of [FLR00] Neal J. Fowler, Marcelo Laca, and of infinite graphs, Proc. Amer. Math. Iain Raeburn, 128 Soc. The C ∗-algebras 8, (2000), no. 48 SØREN EILERS, GUNNAR RESTORFF, EFREN RUIZ, AND ADAM P. W. SØRENSEN [HS03] URL: http://dx.doi.org/10.1090/S0002-9939-99-05378-2, 2319 -- 2327, doi:10.1090/S0002-9939-99-05378-2 . MR 1670363 (2000k:46079) Jeong Hee Hong and Wojciech Szymański, Purely infinite Cuntz-Krieger algebras of directed graphs, Bull. London Math. Soc. 35 (2003), no. 5, 689 -- 696, URL: http://dx.doi.org/10.1112/S0024609303002364 , doi:10.1112/S0024609303002364. MR 1989499 (2005c:46097) [MM14] Kengo Matsumoto and Hiroki Matui, Continuous orbit equivalence of topo- shifts 54 863 -- 877, URL: http://dx.doi.org/10.1215/21562261-2801849, logical Markov (2014), no. 4, doi:10.1215/21562261-2801849 . MR 3276420 Paul S. Muhly and Mark Tomforde, Adding tails to C ∗-correspondences, Doc. Math. 9 (2004), 79 -- 106. MR 2054981 (2005a:46117) and Cuntz-Krieger algebras, Kyoto J. Math. [MT04] [New72] Morris Newman, Integral matrices, Academic Press, New York, 1972, Pure and Applied [Rae05] [Res06] Mathematics, Vol. 45. MR 0340283 (49 #5038) Iain Raeburn, Graph algebras, CBMS Regional Conference Series in Mathematics, vol. 103, Published for the Conference Board of the Mathematical Sciences, Washing- ton, DC; by the American Mathematical Society, Providence, RI, 2005. MR 2135030 (2005k:46141) Gunnar Restorff, isomorphism, ble http://dx.doi.org/10.1515/CRELLE.2006.074, MR 2270572 (2007m:46090) sta- 185 -- 210, URL: doi:10.1515/CRELLE.2006.074. J. Reine Angew. Math. 598 (2006), of Cuntz-Krieger Classification algebras up to [Rør95] Mikael Rørdam, Classification of Cuntz-Krieger algebras, K-Theory 9 (1995), no. 1, 31 -- 58, URL: http://dx.doi.org/10.1007/BF00965458 , doi:10.1007/BF00965458. MR 1340839 (96k:46103) [Rør97] [Sør13] , Classification of extensions of certain C ∗-algebras by their six term exact sequences in K-theory, Math. Ann. 308 (1997), no. 1, 93 -- 117, URL: http://dx.doi.org/10.1007/s002080050067 , doi:10.1007/s002080050067. MR 1446202 (99b:46108) Adam P. W. alge- bras, Ergodic Theory Dynam. Systems 33 (2013), no. 4, 1199 -- 1220, URL: http://dx.doi.org/10.1017/S0143385712000260 , doi:10.1017/S0143385712000260. MR 3082546 Sørensen, Geometric classification simple graph of [Szy02] Wojciech Szymański, General Cuntz-Krieger uniqueness theorem, Internat. J. Math. 13 (2002), no. 5, 549 -- 555, URL: http://dx.doi.org/10.1142/S0129167X0200137X, doi:10.1142/S0129167X0200137X. MR 1914564 (2003h:46083) Department of Mathematical Sciences, University of Copenhagen, Universitets- parken 5, DK-2100 Copenhagen, Denmark E-mail address: [email protected] Department of Science and Technology, University of the Faroe Islands, Nóatún 3, FO-100 Tórshavn, the Faroe Islands E-mail address: [email protected] Department of Mathematics, University of Hawaii, Hilo, 200 W. Kawili St., Hilo, Hawaii, 96720-4091 USA E-mail address: [email protected] Department of Mathematics, University of Oslo, PO BOX 1053 Blindern, N-0316 Oslo, Norway E-mail address: [email protected]
1806.10242
4
1806
2019-03-13T13:37:52
Non-closure of quantum correlation matrices and factorizable channels that require infinite dimensional ancilla
[ "math.OA", "math-ph", "math-ph" ]
We show that there exist factorizable quantum channels in each dimension $\ge 11$ which do not admit a factorization through any finite dimensional von Neumann algebra, and do require ancillas of type II$_1$, thus witnessing new infinite-dimensional phenomena in quantum information theory. We show that the set of n by n matrices of correlations arising as second-order moments of projections in finite dimensional von Neumann algebras with a distinguished trace is non-closed, for all $n \ge 5$, and we use this to give a simplified proof of the recent result of Dykema, Paulsen and Prakash that the set of synchronous quantum correlations $C_q^s(5,2)$ is non-closed. Using a trick originating in work of Regev, Slofstra and Vidick, we further show that the set of correlation matrices arising from second-order moments of unitaries in finite dimensional von Neumann algebras with a distinguished trace is non-closed in each dimension $\ge 11$, from which we derive the first result above.
math.OA
math
Non-closure of quantum correlation matrices and factorizable channels that require infinite dimensional ancilla Magdalena Musat∗ and Mikael Rørdam∗ (With an Appendix by Narutaka Ozawa) Abstract We show that there exist factorizable quantum channels in each dimension ≥ 11 which do not admit a factorization through any finite dimensional von Neumann algebra, and do require ancillas of type II1, thus witnessing new infinite-dimensional phenomena in quantum information theory. We show that the set of n×n matrices of correlations arising as second-order moments of projections in finite dimensional von Neumann algebras with a distinguished trace is non-closed, for all n ≥ 5, and we use this to give a simplified proof of the recent result of Dykema, Paulsen and Prakash that the set of synchronous quantum q (5, 2) is non-closed. Using a trick originating in work of Regev, Slofstra correlations C s and Vidick, we further show that the set of correlation matrices arising from second-order moments of unitaries in finite dimensional von Neumann algebras with a distinguished trace is non-closed in each dimension ≥ 11, from which we derive the first result above. 1 Introduction C. Anantharaman-Delaroche introduced in [1] the class of factorizable completely positive maps between von Neumann algebras equipped with a normal faithful state, while studying non-commutative analogues of classical ergodic theory results, including, e.g., G.-C. Rota's "Alternierende Verfahren" theorem. It was shown in [6] that not all unital completely positive, trace-preserving maps on Mn(C) (also referred to as unital quantum channels in dimension n) are factorizable when n ≥ 3, which also led to a negative answer to the so-called asymptotic quantum Birkoff conjecture, [16]. The tool was the characterization established in [6] that a unital, completely positive, trace-preserving map T : Mn(C) → Mn(C) is factorizable if and only if there exist a finite von Neumann algebra N equipped with a (faithful, normal) tracial state τ and a unitary u in Mn(C) ⊗ N such that T (x) = (idn ⊗ τ )(u(x ⊗ 1N )u∗), for all x ∈ Mn(C). Following the terminology introduced in [7], we say in this case that T admits an exact factorization through Mn(C) ⊗ N , and N is called the ancilla. In all previously studied cases of factorizable maps (see [6] and [7], and the recent paper [12] for the case n = 2), the ancilla could be taken to be finite dimensional, and even a full matrix algebra. It was, however, first remarked in [13] that one cannot always take the ancilla to be a full matrix algebra. In this paper we show that for each n ≥ 11, there are factorizable maps on Mn(C) that do not admit a finite dimensional ancilla, nor an ancilla of type I, and we give concrete examples of such maps, see Example 3.5 ∗This research was supported by a travel grant from the Carlsberg Foundation, and by a research grant from the Danish Council for Independent Research, Natural Sciences. This work was carried out in Spring 2018, while the authors were visiting the Institute for Pure and Applied Mathematics (IPAM), which is supported by the National Science Foundation. 1 and Theorem 4.1. Therefore, one needs to employ ancillas of type II1 to describe a general factorizable channel in dimension n ≥ 11. Observe that all factorizable quantum channels do admit an exact factorization through a type II1 von Neumann algebra (even a type II1 factor), by the well-known fact that each finite von Neumann algebra equipped with a fixed faithful normal tracial state embeds in a trace-preserving way into a type II1 factor. The proof of our result uses very recent developments of analysis in quantum information theory concerning the non-closure of certain sets of correlation matrices. The first such result was due to Slofstra, who proved the failure of what is referred to as the strong Tsirelson conjecture. This was recently refined by Dykema, Paulsen and Prakash, [4], to show that the set of synchronous correlation matrices C s q (n, k) is non-closed, when n ≥ 5 and k ≥ 2. We consider the set D(n) of n× n matrices arising from second-order moments of n-tuples of projections in finite von Neumann algebras with a (normal, faithful) tracial state, and the subset Dfin(n) consisting of those matrices that arise likewise from n-tuples of projections in finite-dimensional von Neumann algebras (or C∗-algebras). We show that the set Dfin(n) is not closed, when n ≥ 5. Our proof uses a theorem of Kruglyak, Rabanovich and Samoilenko, [11], also employed in [4], which describes which scalar multiples of the identity operator on a (finite dimensional) Hilbert space can arise as the sum of n projections. We use this to give a shorter and more direct proof of the main result from [4] that the set of synchronous quantum correlation matrices C s q (n, 2) is non-closed, when n ≥ 5. Kirchberg, [10], reformulated the Connes Embedding Problem in terms of the set G(n) of n × n matrices of correlations arising from unitaries in finite von Neumann algebras with a (normal, faithful) tracial state. This result was further refined by Dykema and Juschenko, [3], and in their formulation, the Connes Embedding Problem is equivalent to the statement that F(n) = G(n), for all n ≥ 3, where F(n) is the closure of the set of n × n matrices of correlations arising from unitaries in full matrix algebras. A trick originating in (as of yet unpublished) work of Regev, Slofstra and Vidick (communicated to us by W. Slofstra in May 2018), which we carry out in the setting of finite von Neumann algebras in Section 3, allows us to conclude further that the set Ffin(2n + 1) of matrices of correlations arising from unitaries in finite dimensional von Neumann algebras is non-closed, whenever Dfin(n) is non-closed, i.e., for all n ≥ 5. A connection between the set G(n) and the set of factorizable Schur multipliers on Mn(C) was established in [7]. This connection gives the final link between the established non-closure of the sets Ffin(2n + 1), for n ≥ 5, and existence of factorizable Schur multipliers with no finite dimensional ancilla (or, even stronger, non type I), in each dimension ≥ 11. In the Appendix, written by N. Ozawa, it is shown that the construction by Kruglyak, Rabanovich and Samoilenko in [11] of an n-tuple of projections with sum equal to a multiple α of the identity can be realized in the hyperfinite II1 factor R, for all admissible values of α, except, possibly, for two extremal ones. This, in turn, implies that the factorizable Schur multipliers with no finite dimensional ancilla found in this article do admit R as an ancilla (except, possibly, for the cases corresponding to the above mentioned extremal values of α). Different sets of matrices of correlations arising from the generators of L. Brown's universal C∗-algebra were shown to be non-closed by Harris and Paulsen [8]. This was used by Gao, Harris and Junge [5] to obtain further non-closure results in a matrix-valued setting. 2 2 Non-closure of sets of matrices of quantum correlations For n ≥ 2, let D(n) and Dfin(n) be the set of n × n matrices (cid:2)τ (pjpi)(cid:3)n i,j=1, where p1, . . . , pn are projections in some arbitrary finite von Neumann algebra, respectively, in some finite dimensional von Neumann algebra, equipped with a (normal) faithful tracial state τ . We show in this section that Dfin(n) is non-closed, when n ≥ 5, and we use this to give a more direct proof, avoiding graph correlation functions, of the very recent result of Dykema, Paulsen and Prakash, [4], that the set of synchronous quantum correlations C s q (5, 2) is non-closed. Standard arguments involving ultralimits, as in the proof of Proposition 2.4 (v) below, respectively, direct sums of finite von Neumann algebras, show that the set D(n) is compact and convex. One can likewise show that the set Dfin(n) is convex. The subset Dmatrix(n) of Dfin(n) consisting of n × n matrices (cid:2)trk(pjpi)(cid:3)n i,j=1, where p1, . . . , pn are projections in a matrix algebra Mk(C), for k ≥ 1, is not convex (and not closed), for any n ≥ 1, since each diagonal entry of such a matrix is the (normalized) trace of a projection in a matrix algebra, which is a rational number. It is not hard to see that Dmatrix(n) is a dense subset of Dfin(n). Note that the closure of Dmatrix(n) is equal to D(n), for all n ≥ 3, if and only if the Connes Embedding Problem has an affirmative answer (see Section 3 for a sketch of a proof of this fact). Let us also observe that Dfin(2) = D(2) is closed. Proposition 2.1. We have D(2) =n(cid:18)s u u t(cid:19) : 0 ≤ s, t ≤ 1, max{0, s + t − 1} ≤ u ≤ min{s, t}o. (2.1) Moreover, each matrix in D(2) arises from a pair of projections in the commutative finite- dimensional C∗-algebra C ⊕ C ⊕ C ⊕ C (with respect to a suitable trace). In particular, it follows that Dfin(2) = D(2) and Dfin(2) is closed. Proof. Let (M, τ ) be a finite von Neumann algebra equipped with a tracial state, and let p, q ∈ M be projections. The associated matrix in D(2) is τ (q)(cid:19) =(cid:18)s u u t(cid:19) , (cid:18) τ (p) τ (pq) τ (qp) where s = τ (p), t = τ (q) and u = τ (pq) = τ (qp) = τ (pqp) = τ (qpq). It is clear that 0 ≤ s, t ≤ 1. Since 0 ≤ pqp ≤ p and 0 ≤ qpq ≤ q, we further conclude that 0 ≤ u ≤ min{s, t}. By Kaplansky's formula (see, e.g., [9]), we get τ (p ∧ q) = τ (p) + τ (q) − τ (p ∨ q) ≥ τ (p) + τ (q) − 1. As p ∧ q = q(p ∧ q)q ≤ qpq, we infer that u = τ (qpq) ≥ τ (p ∧ q) ≥ s + t − 1. To complete the proof, we show that each matrix in the right-hand side of (2.1) arises from projections p and q in C ⊕ C ⊕ C ⊕ C, with respect to the trace with weight (α1, α2, α3, α4), satisfying αj ≥ 0 and P4 j=1 αj = 1. Set p = (1, 1, 0, 0) and q = (1, 0, 1, 0). Then pq = (1, 0, 0, 0) and (cid:18) τ (p) τ (qp) α1 α1 + α3(cid:19) . τ (pq) τ (q)(cid:19) =(cid:18)α1 + α2 α1 We must therefore choose α1 = u ≥ 0, α2 = s−u ≥ 0, α3 = t−u ≥ 0, and α4 = 1−α1−α2−α3. The inequality u ≥ s + t − 1 ensures that α1 + α2 + α3 ≤ 1, whence α4 ≥ 0. We do not know if the sets Dfin(n) are closed for n = 3, 4. 3 We proceed to prove that the sets Dfin(n) are non-closed, for n ≥ 5, following the idea of Dykema, Paulsen and Prakash to use Theorem 2.2 below from [11]. As in [11], let Σn be the set of all α ≥ 0 for which there exist projections p1, . . . , pn on a Hilbert space H such that Pn j=1 pj = α · IH. The sets Σn are completely described in [11] and have the following properties: They are symmetric, i.e., if α ∈ Σn, then n − α ∈ Σn. Moreover, Σ2 = {0, 1, 2} and Σ3 = {0, 1, 3 2 , 2, 3}. The set Σ4 is a countably infinite subset of the rational numbers, Q, with one accumulation point, namely 2. For n ≥ 5, Σn is the union of the 2 (n−√n2 − 4n), 1 2 (n+√n2 − 4n)(cid:3) and a countably infinite discrete subset of rational interval(cid:2) 1 2 (n − √n2 − 4n). Set n−1 , n − n numbers, containing {0, 1, n n(n−1)(cid:17). 2 (1 −p1 − 4/n), 1 n(n−1) , 1 − 2n−1 n−1 , n − 1, n}. Observe that n 2 (1 +p1 − 4/n)(cid:3) ⊂(cid:16) 2n−1 Πn =(cid:2) 1 n−1 < 1 (2.2) j=1 pj = α · IH if and only if α ∈ Σn ∩ Q. Then Πn is an interval with non-empty interior, and n−1Σn \ Πn is contained in Q. Theorem 2.2 (Kruglyak, Rabanovich and Samoilenko, [11, Theorem 6]). Let n ≥ 2 be an integer. Then there exist projections p1, . . . , pn on some finite dimensional Hilbert space H such that Pn B(H) to both sides of the equation Pn space H such that Pn Note that the "only if" part of the theorem above is trivial: Apply the standard trace on j=1 dim(pj) = α dim(H). This argument also gives the following quantitative result for rational values of α: If α = a/b is irreducible, with a, b positive integers, and if there exist projections p1, . . . , pn on a Hilbert j=1 pj = α · IH, to obtain Pn j=1 pj = α · IH, then dim(H) ≥ b. For each n ≥ 2 and each t ∈ [1/n, 1], define the n × n matrix A(n) i,j=1 by t =(cid:2)A(n) t (i, j)(cid:3)n A(n) t (i, j) =  t, t(nt − 1) n − 1 , i = j, i 6= j. (2.3) Proposition 2.3. Let n ≥ 2 be an integer, let 1/n ≤ t ≤ 1, and let α = nt. (i, j), for all 1 ≤ i, j ≤ n. Then Pn (i) Let A be a unital C∗-algebra with a faithful tracial state τ and let p1, . . . , pn be projections in A satisfying τ (pjpi) = A(n) j=1 pj = α · 1A. Furthermore, if t is irrational, then A is necessarily infinite dimensional. Respectively, if α = a/b is rational with a, b positive integers and a/b is irreducible, then A has no representation on a Hilbert space of dimension less than b. t (ii) Conversely, let A be a unital C∗-algebra with a tracial state τ , and let p1, . . . , pn be j=1 pj = α · 1A. Then there exist projections ep1, . . . ,epn in projections in A satisfying Pn some matrix algebra Mm(A) over A satisfying nXj=1epj = α · 1Mm(A), eτ (epjepi) = A(n) t (i, j), 1 ≤ i, j ≤ n, where eτ is the normalized trace on Mm(A) induced by τ . Proof. (i). Set q = (nt)−1Pn j=1 pj. To show that q = 1A, it suffices to check that 1 = τ (q) = τ (q2), since this will entail that τ ((1− q)∗(1− q)) = τ ((1− q)2) = 0. These are straightforward 4 and for 1 ≤ i 6= j ≤ n, eτ (epi) = n! Xσ∈Sn 1 eτ (epiepj) = τ (pσ(i)pσ(j)) = 1 n(n − 1)Xk6=ℓ τ (pkpℓ) : = s0. calculations: Indeed, τ (q) = (nt)−1Pn τ (pj) +Xi6=j τ (q2) = (nt)−2(cid:16) nXj=1 j=1 τ (pj) = 1, and τ (pipj)(cid:17) = (nt)−2(cid:18)nt + n(n − 1) · t(nt − 1) n − 1 (cid:19) = 1. If t is irrational, then it follows from (the easy part of) Theorem 2.2 that A does not have any finite dimensional representation on a Hilbert space, so A must be infinite dimensional. Respectively, if α = a/b is rational with a, b positive integers and a/b irreducible, then A has no representation on a Hilbert space of dimension less than b, by the (explicit) comment after the statement of Theorem 2.2. (ii). We follow the same strategy as in the proof of [4, Theorem 4.2]. Let m = n! and let Sn denote the collection of all permutations of the set {1, 2, . . . , n}. For 1 ≤ j ≤ n, set A ⊂ Mm(A), pσ(j) ∈ Mσ∈Sn epj = Mσ∈Sn where the inclusion above is given by identifying the diagonal of Mm(A) with the direct sum Lσ∈Sn A (for some enumeration of Sn). j=1epj = α · 1Mm(A). Moreover, for 1 ≤ i ≤ n, n! Xσ∈Sn It is easy to check that Pn τ (pk) : = t0, τ (pσ(i)) = 1 1 n nXk=1 Since Pn j=1 pj = α · 1A, we deduce that t0 = α/n = t, and that n2t2 = αnt = nXk=1 τ (α · pk) = nXk=1 τ(cid:0)pk nXℓ=1 pℓ(cid:1) = nXk,ℓ=1 τ (pkpℓ) = nt + n(n − 1)s0, from which we conclude that s0 = (n − 1)−1t(nt − 1), as desired. Combining Theorem 2.2 and Proposition 2.3, we obtain the following: Proposition 2.4. For n ≥ 2 and 1/n ≤ t ≤ 1 we have: (i) A(n) (ii) A(n) t ∈ D(n) if and only if A(n) t ∈ Dfin(n) if and only if A(n) t ∈ Dfin(n) if and only if t ∈ n−1Σn, t ∈ Dmatrix(n) if and only if t ∈ n−1Σn ∩ Q. t Moreover, for n ≥ 5, (iii) A(n) (iv) If t ∈ n−1Σn \ Q and p1, . . . , pn are projections in a von Neumann algebra (N, τ ) with (i, j) = τ (pjpi), for all 1 ≤ i, j ≤ n, then N must be belongs to Dfin(n) \ Dfin(n), for t ∈ n−1Σn \ Q = Πn \ Q. faithful tracial state satisfying A(n) of type II1. t 5 (v) For each A ∈ Dfin(n), there exist projections p1, . . . , pn in the ultrapower Rω of the hyperfinite type II1 factor R such that A(i, j) = τRω (pjpi), for all 1 ≤ i, j ≤ n. t ∈ D(n). j=1 pj = α · IH, where α = nt. t ∈ D(n) implies t ∈ n−1Σn, while its second part shows that A(n) Proof. Suppose that n ≥ 2 and 1/n ≤ t ≤ 1. From the first part of Proposition 2.3 (i) we see that A(n) t ∈ Dfin(n) implies t ∈ n−1Σn ∩ Q. Proposition 2.3 (ii) gives that t ∈ n−1Σn implies A(n) Let t ∈ n−1Σn ∩ Q. Then by Theorem 2.2, there exist projections p1, . . . , pn on a finite dimensional Hilbert space H satisfying Pn Identifying the bounded operators on H with a full matrix algebra, we may assume that the projections pj belong to Mk(C), for some k ≥ 2. By Proposition 2.3, we can find projections ep1, . . . ,epn in some larger matrix algebra Mm(C) with normalized trace tr, satisfying A(n) (i, j) = tr(epiepj), for all 1 ≤ i, j ≤ n. This shows that A(n) belongs to Dmatrix(n). As Dmatrix(n) ⊂ Dfin(n), this t ∈ Dfin(n) when t ∈ n−1Σn. This follows directly from (ii) when t is rational. Suppose that t is irrational. As remarked below (2.2), n−1Σn \ Πn ⊂ Q, for all n ≥ 2. Hence t ∈ Πn \ Q. We can therefore find a sequence {tk}∞k=1 of rational numbers in the interval Πn converging to t. Then A(n) , as k → ∞, t and A(n) belongs to the closure of Dfin(n). tk belongs to Dfin(n), for each k ≥ 1, by (i). This shows that A(n) To complete the proof of (i) we must show that A(n) (iii) follows from (i) and (ii). (iv). The finite von Neumann algebra N can have no representation on a finite dimensional completes the proof of (ii). tk → A(n) t t t Hilbert space, by the second part of Proposition 2.3 (i), whence N must be of type II1. (v). Suppose that A is the limit of a sequence {Ak}∞k=1 of matrices in Dfin(n). Since each finite dimensional C∗-algebra with a distinguished trace can be embedded in a trace preserving way into the hyperfinite type II1 factor R, we can find projections p(k) n in R satisfying τR(p(k) ) = Ak(i, j), for all 1 ≤ i, j ≤ n. Let pi be the image in Rω of the sequence {p(k) 1 , . . . , p(k) j p(k) i i }∞k=1 ∈ ℓ∞(R). Then τRω (pjpi) = lim ω τR(p(k) j p(k) i ) = lim ω Ak(i, j) = A(i, j), for all 1 ≤ i, j ≤ n, as wanted. By Proposition 2.4 (iii) and the fact that Πn is an interval with non-empty interior, when n ≥ 5, cf. (2.2), we obtain the following theorem: Theorem 2.5. The set Dfin(n) is non-compact, when n ≥ 5. Remark 2.6. It is shown in Proposition 2.4 above that one can realize the n× n matrix A(n) using projections in Rω, for all n ≥ 2 and all t ∈ n−1Σn. If moreover t is rational, then A(n) Using Proposition 2.3, this also shows that for each n ≥ 2 and for each α in Σn, one can find an n-tuple of projections summing up to α· 1N in some type II1 factor N , e.g., N = Rω. One can also reach this conclusion directly from Theorem 2.2, using an ultraproduct argument as in the proof of Proposition 2.4 (v). can be realized using projections in some matrix algebra, by Proposition 2.3 (ii). t t α ∈(cid:0) 1 In Theorem A.1 in the Appendix by N. Ozawa it is shown that for each n ≥ 5 and each 2 (n+√n2 − 4n)(cid:1), one can find an n-tuple of projections in R summing 2 (n−√n2 − 4n), 1 6 t up to α· 1R. It follows that the matrix A(n) can be realized using projections in R for all t in 2 (1+p1 − 4/n)(cid:3). n−1Σn, except, possibly, for the endpoints of the interval(cid:2) 1 Recall, e.g., from [4, Section 2], that for n, k ≥ 2, the set Cqc(n, k) consists of nk×nk quantum correlation matrices (cid:2)(p(i, jv, w)(cid:3)i,j,v,w with entries given by 2 (1−p1 − 4/n), 1 p(i, jv, w) =(cid:10)Pv,iQw,jψ, ψ(cid:11), i=1 and {Qw,j}k 1 ≤ i, j ≤ k, 1 ≤ v, w ≤ n, where, for each v and w, {Pv,i}k j=1 are projection-valued measures on some Hilbert space H, satisfying Pv,iQw,j = Qw,jPv,i, for all i, j, v, w, and where ψ is a unit vector in H. Let Cq(n, k) be the same set of quantum correlation matrices, but with the additional as- sumption that H = HA⊗HB, for some finite dimensional Hilbert spaces HA and HB, and Pv,i belongs to B(HA)⊗ IHB , while Qw,j belongs to IHA ⊗ B(HB). The closure of the set Cq(n, k) is denoted by Cqa(n, k). We have the following inclusions: Cq(n, k) ⊆ Cqa(n, k) ⊆ Cqc(n, k). Furthemore, the sets of synchronous correlation matrices, denoted by C s qa(n, k), qc(n, k), respectively, consist of those quantum correlation matrices (cid:2)(p(i, jv, w)(cid:3)i,j,v,w and C s in Cq(n, k), Cqs(n, k), and Cqc(n, k), respectively, where p(i, jv, v) = 0, whenever i 6= j. We use Theorem 2.5 and Proposition 2.7 below from [14] to give a shorter proof of [4, Theorem 4.2] of Dykema, Paulsen and Prakash, which, again, was refining Slofstra's result (in the general, non-synchronous case) from [15]. For this, using the notation from [14], for n, k ≥ 2 let Ds in a finite dimensional von Neumann algebra with a (faithful) tracial state τ , satisfying Pk i=1 ev,i = 1, for all 1 ≤ v ≤ n. Note that Ds Proposition 2.7 (Paulsen, Severini, Stahlke, Todorov and Winter, [14]). For all n, k ≥ 2, one has C s q(n, k) be the set of matrices (cid:2)τ (ev,iew,j)(cid:3)i,j,v,w, where ev,i are projections q(n, k) is a subset of D(nk). q (n, k), C s q (n, k) = Ds q(n, k). q (n, 2) of synchronous quantum Theorem 2.8 (Dykema, Paulsen and Prakash, [4]). The set C s correlation matrices is non-compact, for all n ≥ 5. Proof. Use Theorem 2.5 to find a matrix A in D(n) \ Dfin(n) and a sequence {Ak}∞k=1 of matrices in Dfin(n) converging to A. Let p1, . . . , pn be projections in Rω such that τRω (pjpi) = A(i, j), for all i, j, cf. Proposition 2.4 (v), and let further p(k) n be projections in some matrix algebra Mmk (C) with normalized trace trmk such that trmk (p(k) ) = Ak(i, j), 1 ≤ i, j ≤ n. 1 , . . . , p(k) j p(k) i 0,v = p(k) v and e(k) 1,v = 1 − p(k) v in Mmk (C), Set e0,v = pv and e1,v = 1 − pv in Rω, and set e(k) for all k ≥ 1 and 1 ≤ v ≤ n. It then follows that lim k→∞(cid:2)trmk (e(k) w,j)(cid:3)i,j,v,w =(cid:2)τRω (ev,iew,j)(cid:3)i,j,v,w, v,i e(k) and each of the matrices (cid:2)trmk (e(k) w,j)(cid:3)i,j,v,w belongs to Ds v,i e(k) tion 2.7. However, the matrix (cid:2)τ (ev,iew,j)(cid:3)i,j,v,w itself does not belong to C s matrix A =(cid:2)τ (ev,0ew,0)(cid:3)v,w does not belong to Dfin(n). Note that the set Cq(n, 2) of (non-synchronous) quantum correlation matrices contains C s as a relatively closed subset, which shows that Cq(n, 2) also is non-closed, when n ≥ 5. q(n, 2) = C s We end this section with a remark on the matrices A(n) defined in (2.3). q (n, 2), cf. Proposi- q (n, 2), since the q (n, 2) t 7 t t = A(n) is then equal to (cid:2)τ (pjpi)(cid:3)n Remark 2.9. For each integer n ≥ 2 and for all s, t ∈ [0, 1], consider the n × n matrix A(n) t,s , whose diagonal entries all are equal to t and whose off-diagonal entries are all equal to s. Note that A(n) t,s , with s := t(nt − 1)/(n − 1), when t ∈ [1/n, 1]. The purpose of this remark is to describe the set of "admissible pairs" (t, s), for which A(n) t,s belongs to D(n), and to show that s = t(nt − 1)/(n − 1) is the smallest number for which (t, s) is such an admissible pair, for each fixed t ∈ n−1Σn, Note first that for each fixed t ∈ [0, 1], the set I (n) of those s ∈ [0, 1], for which (t, s) is admissible, is a closed interval. This follows by convexity and compactness of the set D(n). Fix t ∈ [0, 1], take a projection p of trace t in some finite von Neumann algebra, and let p1 = ··· = pn = p. The matrix A(n) i,j=1, and therefore belongs to D(n). Moreover, since τ (pq) ≤ τ (p), whenever p, q are projections in a von Neumann algebra with tracial state τ , we conclude that t = max I (n) factor M with τM (pj) = t, for all j. The corresponding n × n matrix (cid:2)τ (pjpi)(cid:3)n t,0 . Hence 0 belongs to I (n) t = [0, t], when 0 ≤ t ≤ 1/n. Let s, t ∈ [0, 1], and suppose that (t, s) is an admissible pair. Let p1, . . . , pn be projections in a finite von Neumann algebra with tracial state τ such that (cid:2)τ (pjpi)(cid:3)n t,s . Put qj = 1 − pj, 1 ≤ j ≤ n. Then(cid:2)τ (qjqi)(cid:3)n 1−t,1−2t+s, which shows that (1 − t, 1− 2t + s) is an admissible pair. The map given by (t, s) 7−→ (1− t, 1− 2t + s) is involutive, and therefore it maps the set of admissible pairs in [0, 1]2 onto itself. In particular, this involution maps {t} × I (n) 1−t. Combining this fact with the result of the previous paragraph, we obtain that I (n) projections in some finite von Neumann algebra with tracial state τ such that(cid:2)τ (pjpi)(cid:3)n For 0 ≤ t ≤ 1/n we can find pairwise orthogonal projections p1, . . . , pn in any type II1 i,j=1 is equal Consider finally the case where t ∈ [1/n, 1−1/n]. Suppose that s ∈ It, and let p1, . . . , pn be i,j=1 = t = [2t − 1, t], when 1 − 1/n ≤ t ≤ 1. onto {1 − t} × I (n) . This shows that I (n) i,j=1 = A(n) i,j=1 = A(n) to A(n) . t t,t t t A(n) t,s . Then nt + n(n − 1)s = τ(cid:16)( nXi=1 pi)(cid:17)2 pi)2(cid:17) ≥(cid:16)τ ( nXi=1 which implies that s ≥ t(nt − 1)/(n − 1). In other words, I (n) t ⊆ [t(nt − 1)/(n − 1), t]. It was noted in Remark 2.6 that A(n) t,t(nt−1)/(n−1) belongs to D(n) if and only if t belongs to n−1Σn. For those values of t, we therefore obtain that s = t(nt − 1)/(n − 1) is the smallest number for which (t, s) is an admissible pair, whence I (n) t = A(n) = n2t2, It remains a curious open problem to detemine the interval I (n) (non-empty) set [1/n, 1− 1/n]\ n−1 Σn. In this case, necessarily, min I (n) t , for t belonging to the t > t(nt− 1)/(n− 1). t = [t(nt − 1)/(n − 1), t]. 3 Correlation matrices of unitary elements Recall that a correlation matrix is a positive definite matrix whose diagonal entries are equal to 1. For each integer n ≥ 2, let G(n), Fmatrix(n), and Ffin(n) be the set of n × n correlation matrices(cid:2)τ (u∗j ui)(cid:3)n i,j=1, where u1, u2, . . . , un are unitaries in some finite von Neumann algebra equipped with a faithful tracial state τ , respectively, in some full matrix algebra with its canonical tracial state, respectively, in some finite dimensional C∗-algebra with a faithful tracial state. Then the convex hull conv(Fmatrix(n)) of Fmatrix(n) is equal to Ffin(n), and the 8 two sets Fmatrix(n) and Ffin(n) have the same closure, which is denoted by F(n). The sets G(n) and F(n) are compact and convex, see [3, Proposition 1.4]. As shown by Kirchberg, [10] (cf. Dykema -- Juschenko, [3]), the Connes Embedding Problem has an affirmative answer if and only if G(n) = F(n), for all n ≥ 3. We can use this to show that Dmatrix(n) is dense in D(n), for all n ≥ 3, if and only if the Connes Embedding Problem has an affirmative answer. Indeed, the proof of the "if" part follows the same strategy as the proof of the "if" part of Kirchberg's result, where a given finite subset of Rω is approximated in trace-norm with a finite subset of a matrix subalgebra of Rω. To see the "only if" part, assume that Dmatrix(n) is dense in D(n), for all n ≥ 3. We show that this implies G(n) = F(n), for all n ≥ 3. Take an n-tuple u1, . . . , un of unitaries in some tracial von Neumann algebra (N, τ ). Fix an integer m ≥ 1. Each uj can be approximated in norm within 2π/m by unitaries v1, . . . , vn in N of the form vj = Pm k=1 ωkpk,j, where ω = exp(2πi/m) and p1,j, . . . , pm,j are pairwise orthogonal projections in N summing up to 1, for each j. Approximate the second-order moments of the collections of projections {pk,j}k,j by second-order moments of projections {qk,j}k,j in some matrix algebra (Mr(C), trr). Then trr(qk,jqℓ,j) are small when k=1 qk,j(cid:1) is close to 1, for all j. A standard lifting argument allows us to k 6= ℓ, and trr(cid:0)Pm replace the projections {qk,j}k,j with new projections, close to the old ones in trace-norm, satisfying Pm k=1 ωkqk,j. The second-order moments of the unitaries w1, . . . , wn are then close to those of u1, . . . , un. Remark 3.1. In the case where n = 2, the set Fmatrix(2) is closed and convex, and Fmatrix(2) = F(2) = G(2), which further is equal to the set of 2 × 2 matrices of the form j=1 qk,j = 1, for all j. Set wj = Pm (cid:18)1 ¯z z 1(cid:19) , for z ∈ C with z ≤ 1. To see that each of these 2 × 2 matrices belongs to Fmatrix(2), take z ∈ C with z ≤ 1 and find λ1 and λ2 on the complex unit circle T, with z = (λ1 + λ2)/2. The correlation matrix arising from the unitary 2 × 2 matrices u1 = 1 and u2 = diag(λ1, λ2) is then as desired. We show in this section that the set Ffin(n) is not closed (hence, not compact), for all n ≥ 11. This result originates in a remark made by T. Vidick during his talk at one of the workshops in the Quantitative Linear Algebra program at IPAM, Spring 2018, that led to subsequent discussions with W. Slofstra, who, in particular, communicated to us a version of the following result (to appear, in an approximate case, in a forthcoming paper by O. Regev, W. Slofstra and T. Vidick): Proposition 3.2. Let M be a finite von Neumann algebra with a faithful tracial state τM , and let p1, . . . , pn be projections in M . Further, let u0, u1, . . . , un, un+1, . . . , u2n be the unitaries in M given by u0 = 1, uj = 2pj − 1, (1 ≤ j ≤ n), uj = (uj−n + i · 1)/√2, (n + 1 ≤ j ≤ 2n). Let N be another finite von Neumann algebra with a faithful tracial state τN . Then there exist 2n + 1 unitaries v0, v1, . . . , v2n in N satisfying τN (v∗j vi) = τM (u∗j ui), (3.1) 0 ≤ i, j ≤ 2n, if and only if there exist n projections q1, . . . , qn in N satisfying 1 ≤ i, j ≤ n. τN (qjqi) = τM (pjpi), (3.2) 9 Proof. Assume that q1, . . . , qn are projections in N satisfying (3.2). Equip the vector spaces span{p1, p2, . . . , pn} and span{q1, q2, . . . , qn} with the Euclidean structure arising from the traces τM and τN , respectively. Using (3.2), we see that the map pj 7−→ qj, 1 ≤ j ≤ n, extends to a well-defined linear isometry ϕ from span{p1, . . . , pm} to span{q1, . . . , qm}. Set v0 = 1, vj = 2qj − 1, for 1 ≤ j ≤ n, and vj = (vj−n + i · 1)/√2, for n + 1 ≤ j ≤ n, and use the isometric property of ϕ to check that (3.1) holds. Conversely, assume that we are given unitaries v0, v1, . . . , v2n in N satisfying (3.1). Upon replacing vj by v∗0vj, for all 0 ≤ j ≤ 2n, we may assume that v0 = 1. As above, equip the vector spaces span{u0, u1, . . . , u2n} and span{v0, v1, . . . , v2n} with the Euclidean structure arising from the traces τM and τN , respectively. Then, by (3.1), we have a well-defined linear isometry ψ : span{u0, u1, . . . , u2n} → span{v0, v1, . . . , v2n}, mapping uj to vj, for 0 ≤ j ≤ 2n. In particular, for 1 ≤ j ≤ n, (cid:13)(cid:13)vj+n − (vj + i · v0)/√2(cid:13)(cid:13)2 =(cid:13)(cid:13)uj+n − (uj + i · u0)/√2(cid:13)(cid:13)2 = 0, so vj+n = (vj + i · 1)/√2. Note that if u and (u + i · 1)/√2 are unitaries in some unital C∗-algebra, then u is necessarily a symmetry. Indeed, if λ is a complex number such that λ = (λ + i)/√2 = 1, then λ ∈ R. Hence, if u is as stated, then its spectrum is contained in R, which entails that it is a symmetry. We conclude that v1, . . . , vn are symmetries. For 1 ≤ j ≤ n, set qj = (vj + 1)/2. Then qj is a projection and vj = 2qj − 1. Use the isometric property of ψ to check (3.2). Corollary 3.3. The set Fmatrix(m) is not compact and not convex, whenever m ≥ 3. Proof. Let 0 < α < 1 be irrational. Equip M := C ⊕ C with the trace τ given by τ (x, y) = αx + (1 − α)y, for x, y ∈ C. Consider first the case where m = 3. Let n = 1 and let p = p1 = (1, 0) ∈ M . Then τ (p) = α is irrational. Let u0, u1, u2 be the unitaries in M arising from this projection as in the proposition above (with n = 1). The matrix (cid:2)τ (u∗j ui)(cid:3)2 i,j=0 belongs to conv(Fmatrix(3)), and hence to F(3). However, by Proposition 3.2, it does not belong to Fmatrix(3) itself, because no full matrix algebra contains a projection of irrational trace, and therefore contains no projection q = q1 satisfying (3.2) (with n = 1). Assume now that m > 3, let u0, u1, u2 be as above, and let unitaries u3, u4, . . . , um−1 ∈ M be arbitrary. If v0, v1, . . . , vm−1 are unitaries in some tracial von Neumann algebra (N, τN ) satisfying (3.1), then v0, v1, v2 satisfy (3.1) with respect to the set {u0, u1, u2}, so v0, v1, v2, and hence v0, v1, . . . , vm−1 cannot be found in a full matrix algebra. These arguments also yield the non-convexity of the set Fmatrix(m) in all cases. Example 3.4. For m = 3, the correlation matrix B = (cid:2)τ (u∗j ui)(cid:3)2 u2 =(cid:18) 1 + i √2 i,j=0 from the proof above, √2 (cid:19) , , −1 + i with unitaries given by u0 = (1, 1), u1 = (1,−1), has the following explicit form in terms of the parameter α ∈ (0, 1): B =  1 γ γ+i√2 γ 1 1+γi√2 10 γ−i√2 1−γi√2 1   , where γ = 2α − 1 ∈ (−1, 1). Note that the matrix B belongs to Ffin(3), for all γ ∈ (−1, 1), while it does not belong to Fmatrix(3), whenever γ is irrational. Example 3.5. For each n ≥ 2 and each t ∈ [1/n, 1], consider the self-adjoint (1+2n)×(1+2n) complex matrix 1 X∗ Y ∗ X D1 C∗ Y C D2   , B(n) t =      1 1 ... 1 , s + i √2   1 1 ... 1   where X, Y and C, D1, D2 are the n × 1, respectively, n × n complex matrices given by X = s , Y = C = 1 + is √2 In + 4(r − t) + 1 + is √2 E, D1 = In + (4(r − t) + 1) E, D2 = In + (2(r − t) + 1) E, where s = 2t − 1 and r = (n − 1)−1t(nt − 1), and 1 1 ... 0 0 1 ··· 1 0 ··· ... . . . 1 1 ···   E = ...   ∈ Mn(C). With this definition, Proposition 3.2 yields that the (2n + 1) × (2n + 1) matrix B(n) is the correlation matrix of an (2n + 1)-tuple of unitaries in some finite von Neumann algebra M if and only if the n × n matrix A(n) is the correlation matrix of an n-tuple of projections in the same von Neumann algebra M . Indeed, if p1, . . . , pn are projections in M such that (cid:2)τM (pjpi)(cid:3)n , and if u0, u1, . . . , u2n are the 2n + 1 unitaries in M constructed i,j=0. Conversely, if i,j=0, then there are projections from these projections as in Proposition 3.2, then B(n) v0, v1, . . . , v2n are unitaries in M such that B(n) q1, . . . , qn ∈ M such that A(n) i,j=1. t = (cid:2)τM (u∗j ui)(cid:3)2n t =(cid:2)τM (v∗j vi)(cid:3)2n i,j=1 = A(n) t t t belongs to G(2n + 1), respectively, to Ffin(2n + 1), if and only if A(n) t In particular, B(n) t =(cid:2)τM (qjqi)(cid:3)n belongs to D(n), respectively, to Dfin(n). Theorem 3.6. Let n ≥ 5 and let t ∈ Πn. t (i) Then B(n) t belongs to Ffin(2n + 1), if t is rational, and B(n) t Ffin(2n + 1), if t is irrational. belongs to F(2n + 1) \ N with a faithful tracial state τN such that τN (v∗j vi) = B(n) N is necessarily of type II1. (ii) If t is irrational and if v0, v1, . . . , v2n are unitaries in some finite von Neumann algebra (i, j), for 0 ≤ i, j ≤ 2n, then 2 (1 +p1 − 4/n)(cid:1), then there are unitaries v0, v1, . . . , v2n in 2 (1 −p1 − 4/n), 1 (iii) If t ∈ (cid:0) 1 the hyperfinite II1 factor R such that τN (v∗j vi) = B(n) (i, j), for 0 ≤ i, j ≤ 2n. t t 11 t t t ∈ G(2n+1), whenever A(n) (iv) The convex sets Ffin(k) are non-compact, for all k ≥ 11. Proof. (i). The map t 7→ B(n) and Example 3.5 that B(n) Πn. Moreover, if t ∈ Πn, then B(n) when t is irrational. This implies that B(n) is non-compact, for each n ≥ 5. tracial state τN , satisfying τN (v∗j vi) = B(n) there exist projections q1, . . . , qn in N such that A(n) has no finite dimensional representations, by Proposition 2.3, so N must be of type II1. , t ∈ Πn, is clearly continuous. It follows from Proposition 2.3 t ∈ D(n), and in particular whenever t ∈ t ∈ Ffin(2n + 1), when t is rational, and B(n) /∈ Ffin(2n + 1), t ∈ F(2n + 1), for all t ∈ Πn, and that Ffin(2n + 1) (ii). Let v0, v1, . . . , v2n be unitaries in some finite von Neumann algebra N with faithful (i, j), for 0 ≤ i, j ≤ 2n. Then, by Example 3.5, (i, j) = τN (qjqi). This entails that N 2 (1 +p1 − 4/n)(cid:1). It follows from Theorem A.1 in the can be realized using projections in R, cf. Remark 2.6. The (iv). We show that if Ffin(k) is non-compact, for some positive integer k, then so is Appendix by Ozawa that A(n) claim now follows from Example 3.5. (iii). Let t ∈ (cid:0) 1 2 (1 −p1 − 4/n), 1 t t t Ffin(k + 1). To this end, define a map ρ : F(k) → F(k + 1) by i,j=1(cid:17) =(cid:2)τM (u∗j ui)(cid:3)k+1 ρ(cid:16)(cid:2)τM (u∗j ui)(cid:3)k i,j=1, whenever u1, . . . , uk are unitaries in some von Neumann algebra M with a faithful tracial state τ , and where uk+1 is chosen to be equal to u1. Then the last row and the last column i,j=1 is equal to the first row and the first column of the matrix (cid:2)τM (u∗j ui)(cid:3)k of(cid:2)τM (u∗j ui)(cid:3)k+1 i,j=1, which shows that ρ is well-defined and continuous. Moreover, ρ−1(Ffin(k + 1)) = Ffin(k). Hence, if Ffin(k + 1) is compact, and thus closed, then Ffin(k) is closed, and thus compact. 4 Factorizable maps that require infinite dimensional ancilla We prove here our claimed result about existence of factorizable quantum channels in all dimensions ≥ 11, requiring infinite dimensional ancilla. We first recall necessary prerequisites. Let UCPT(n) denote the convex and compact set of all unital completely positive trace preserving linear maps T : Mn(C) → Mn(C), n ≥ 2. Maps in UCPT(n) are also called unital quantum channels in dimension n. Anantharaman-Delaroche defined in [1] a channel to be factorizable if it admits a factorization (in a suitable way) through a finite von Neumann algebra with a faithful tracial state. This notion was studied extensively in [6], and the following characterization (which we will take to be our definition of factorizable maps) was established therein (cf., [6, Theorem 2.2]): A unital quantum channel T in dimension n is factorizable if and only if there exist a finite von Neumann algebra N , equipped with a normal faithful tracial state τN , and a unitary u ∈ Mn(C) ⊗ N such that T (x) = (idn ⊗ τN )(u(x ⊗ 1N )u∗), x ∈ Mn(C). (4.1) The von Neumann algebra N above is also called the ancilla, and, following Definition 3.1 in [7], we say that T has an exact factorization through Mn(C) ⊗ N . The set of all factorizable unital channels in dimension n is denoted by FM(n). This set is convex and compact, as shown by standard arguments. 12 constructed in Example 3.5. t t Further, let FMmatrix(n) and FMfin(n) denote the set of factorizable maps in UCPT(n) that exactly factor through a full matrix algebra, respectively, through a finite dimensional C∗-algebra, equipped with a faithful tracial state. It was shown in [7, Theorem 3.7] that a positive answer to the Connes Embedding Problem is equivalent to FMmatrix(n) being dense in FM(n), for all n ≥ 3. Moreover, if a unital quantum channel T : Mn(C) → Mn(C) belongs to the closure of FMmatrix(n), then T admits an exact factorization through an ultrapower Rω of the hyperfinite type II1 factor R. It is shown in the upcoming manuscript [13] that FMfin(n) = conv(FMmatrix(n)), and that FMmatrix(n) is non-compact and non-convex, when n ≥ 3. Let S(n) be the set of all Schur multipliers TB : Mn(C) → Mn(C), where B ∈ Mn(C), i.e., TB(x) is the Schur product of B and x, for x ∈ Mn(C). Let FMS(n) = FM(n) ∩ S(n) be the set of all factorizable Schur multipliers, and write FMS fin(n) = FMfin(n) ∩ S(n). For the theorem below, recall the definition (2.2) of the set Πn, and the (2n + 1)× (2n + 1) matrix B(n) Theorem 4.1. The set FMfin(k) is not compact, for all k ≥ 11. Moreover, for each n ≥ 5 and each irrational number t ∈ Πn, the Schur multiplier TB, where B = B(n) , is a factorizable map belonging to the closure of FMfin(2n + 1), but not to FMfin(2n + 1), and it requires an ancilla of type II1. This ancilla can be taken to be the hyperfinite II1 factor R, when 2 (1 −p1 − 4/n) < t < 1 Proof. It was shown in [6, Proposition 2.8] that if B ∈ Mk(C) is a correlation matrix, then the associated Schur multiplier TB admits an exact factorization through Mk(C) ⊗ N , where N is a finite von Neumann algebra with normal faithful tracial state τN , if and only if B = (cid:2)τN (u∗j ui)(cid:3)k i,j=1, for some unitaries u1, . . . , uk ∈ N . In particular, TB belongs to FMS(k), FMS fin(k), and the closure of FMS fin(k), respectively, if and only if B belongs to G(k), Ffin(k), and F(k), respectively. Since the map from G(k) to FM(k) given by B 7→ TB is continuous, it follows from Theorem 3.6 that FMS fin(k) is non-closed in FM(k). As the set S(k) is closed, we conclude that FMfin(k) is non-compact. To prove the second part of the theorem, let n ≥ 5, let t ∈ Πn be irrational, and let B = B(n) . Then B belongs to F(2n + 1) \ Ffin(2n + 1) by Theorem 3.6. By the argument in the first paragraph, we conclude that TB belongs to the closure of FMS fin(2n + 1), but not to FMS fin(2n + 1). Moreover, if TB admits an exact factorization through Mn(C) ⊗ N , with ancilla (N, τN ) as in the first paragraph, then B is the matrix of correlations of unitaries u0, u1, . . . , u2n in N , which by Theorem 3.6 implies that N must be of type II1. 2 (1 +p1 − 4/n). 1 t The remaining part of the theorem follows from Theorem 3.6 (iii) and the argument in the first paragraph. We also obtain the following quantitative version of the theorem above, in the case where t ∈ Πn is rational and n ≥ 5: Let B = B(n) . If nt = a/b, with a, b positive integers and a/b irreducible, then the Schur channel TB admits an exact factorization through a finite dimensional von Neumann algebra M = Mn(C)⊗ N , where the von Neumann algebra N can have no representation on a Hilbert space of dimension smaller than b. This follows as in the proof of the theorem above and by appealing to Proposition 2.3 (i). t We conclude that for every fixed integer n ≥ 11, there is a sequence of factorizable unital quantum channels in dimension n, each admitting finite-dimensional ancillas, but where the size of any such ancillas must tend to infinity. 13 Appendix Realizing the Kruglyak -- Rabanovich -- Samoilenko projections in the hyperfinite II1 factor by Narutaka Ozawa 2 (n −√n2 − 4n), 1 2 (n − √n2 − 4n), 1 Kruglyak, Rabanovich, and Samoilenko proved in [11, Theorem 6], cf. Theorem 2.2, that for 2 (n +√n2 − 4n)], there are orthogonal projections any n ≥ 5 and any α ∈ [ 1 p1, . . . , pn such thatPi pi = α. In this appendix, we observe that the Kruglyak -- Rabanovich -- Samoilenko construction shows that these projections are realized in the hyperfinite II1 factor R, possibly except for the extremities. 2 (n + √n2 − 4n)), there are Theorem A.1. For any n ≥ 5 and any α ∈ ( 1 projections p1, . . . , pn ∈ R which satisfy Pi pi = α. Let (M, τ ) be a finite von Neumann algebra. By a matricial approximation (or matri- cial microstates) of a d-tuple (a1, . . . , ad) in M sa, we mean a sequence (x1(n), . . . , xd(n)) ∈ (Mk(n)(C)sa)d such that limn tr(p(x1(n), . . . , xd(n))) = τ (p(a1, . . . , ad)) for every polynomial p in d non-commuting variables. A matricial approximation of a generating d-tuple (a1, . . . , ad) of M gives rise to an embedding of (M, τ ) into the tracial ultraproduct of (Mk(n)(C), trk(n)). Recall that M satisfies the Connes Embedding Conjecture, i.e., (M, τ ) ֒→ (Rω, τ ω), if and only if every (or some) generating d-tuple (ai)d i=1 in M sa admits a matricial approximation. We will give a sufficient condition for hyperfiniteness of M in terms of a matricial approximation. For x = (xij) ∈ Mk(C), we define its propagation to be max{i − j : xij 6= 0}. Lemma A.2. Let (M, τ ) be a finite von Neumann algebra generated by a1, . . . , ad ∈ M sa. Assume that (a1, . . . , ad) admits a matricial approximation (x1(n), . . . , xd(n)) with uniformly bounded propagations. Then M is hyperfinite. m=−l fmzm k for some fm ∈ Dk with kfmk ≤ kyk. Proof. Let k be a positive integer and consider the shift unitary matrix zk ∈ Mk(C) given by (zk)i,j = δi+1,j (modulo k). It normalizes the diagonal maximal abelian subalgebra Dk ⊂ Mk(C). Observe that any y ∈ Mk(C) that has propagation at most l can be written as y =Pl Now, let a matricial approximation (x1(n), . . . , xd(n)) be given as in the statement. We denote by Mω the tracial ultraproduct of (Mk(n)(C), trk(n)) and by Dω the subalgebra arising from the diagonal maximal abelian subalgebras Dk(n). From the above discussion, one sees that the element xi ∈ Mω that corresponds to (xi(n))n belongs to the von Neumann subal- gebra generated by Dω and z, where z is the unitary element corresponding to (zk(n))n. The von Neumann subalgebra generated by x1, . . . , xd is isomorphic to M and the von Neumann subalgebra generated by Dω and z is hyperfinite (as it is isomorphic to Dω ⋊ Z, assuming k(n) → ∞). Proof of Theorem A.1. Firstly, note that every separable finite von Neumann algebra (N, τ ) with a faithful normal tracial state is embeddable in a trace-preserving way into a separable II1 factor M , which can be taken to be the hyperfinite II1 factor R if M is hyperfinite. Indeed, as observed by U. Haagerup, we may take M to be (N∞n=1 N ) ⋊ S∞, where the infinite tensor 14 product is with respect to the standard representation of N on L2(N, τ ), and where S∞ is the (locally finite) group of permutations on the natural numbers with finite support. It therefore suffices to find the projections p1, . . . , pn in any hyperfinite finite von Neumann algebra N . For each α ∈ Q ∩ [3/2, 2], the projections P1(α), . . . , P5(α) in Mk(α)(C) that satisfy Pi Pi(α) = α are constructed in [11, Theorem 6] as Ri. The proof of Theorem 6 (and Lemma 7) in [11] reveals that the projections Ri are obtained by sewing (see [11, Defini- tion 1]) the projections P (k) ∈ Mki+2(C), ki ∈ {1, 2, 3}. Since P (k) 's have propagation at most 4, the projections Ri have propagation at most 8, regardless of α. i i Let α ∈ (3/2, 2) be given and take a rational sequence (αn)n which converges to α. Then after passing to a convergent subsequence, (P1(αn), . . . , P5(αn)) is a matricial approximation of (P1, . . . , P5) in the tracial ultraproduct Mω of (Mk(n)(C), trk(n)) and (P1, . . . , P5) satisfies Pi Pi = α. By Lemma A.2, the projections P1, . . . , P5 generate a hyperfinite von Neumann subalgebra. This proves Theorem A.1 for n = 5 and α ∈ [3/2, 2]. By [11, Lemma 5], this implies Theorem A.1 for every n ≥ 5 and α ∈ [2, n − 2]. 2 (n+√n2 − 4n)) are obtained by iterating the numerical mappings Φ+ and Φ− (see [11, Section 1.2]) starting at α ∈ [2, n − 2] (see [11, Lemma 6]). Thus it suffices to show the functors S and T constructed in Section 1.2 in [11] preserve hyperfiniteness. This is clear for the linear reflection T . For the reader's convenience, we replicate here the construction of the hyperbolic reflection S, adapted to i=1 Pi = α. We will construct 2 (n−√n2 − 4n), 1 Finally, note that all values in ( 1 our setting. Let P1, . . . , Pn ∈ N be projections such that Pn projections Q1, . . . , Qn such that Pn Vi := (α2 − α)−1/2 Pi(cid:2)−P1 Then, ViV ∗i = (α2 − α)−1Pi(α2 − 2αPi +Pn Qi := V ∗i Vi ∈ Mn(N ) is a projection. A calculation shows i=1 Qi = α α−1 . Put ··· α − Pi k=1 Pk)Pi = Pi and Vi is a partial isometry. Hence ··· −Pn(cid:3) ∈ M1,n(N ). Xk Qk = (α2 − α)−1(α2 diag(P1, . . . , Pn) − α [PiPj]i.j) ∈ Mn(N ). Note that Q := diag(P1, . . . , Pn) − α−1[PiPj]i.j is a projection and one has Pk Qk = α α−1 Q. Thus, viewing Qk as projections in QMn(N )Q, we are done. When N is hyperfinite, so is the amplification QMn(N )Q. References [1] C. Anantharaman-Delaroche, On ergodic theorems for free group actions on noncommutative spaces, Probab. Theory Related Fields 135 (2006), no. 4, 520 -- 546. [2] N. P. Brown and N. Ozawa, C ∗-algebras and finite-dimensional approximations, Graduate Studies in Mathematics, vol. 88, American Mathematical Society, Providence, RI, 2008. [3] K. Dykema and K. Juschenko, Matrices of unitary moments, Math. Scand. 109 (2011), no. 2, 225 -- 239. [4] K. Dykema, V. I. Paulsen, and J. Prakash, Non-closure of the set of quantum correlations via graphs, Comm. Math. Phys. 365 (2019), no. 3, 1125 -- 1142. [5] L. Gao, S.J. Harris, and M. Junge, Quantum Teleportation and Super-dense Coding in Operator Algebras, arXiv:1709.02785, 2017. 15 [6] U. Haagerup and M. Musat, Factorization and dilation problems for completely positive maps on von Neumann algebras, Comm. Math. Phys. 303 (2011), no. 2, 555 -- 594. [7] , An asymptotic property of factorizable completely positive maps and the Connes embed- ding problem, Comm. Math. Phys. 338 (2015), no. 2, 721 -- 752. [8] S.J. Harris and V.I. Paulsen, Unitary correlation sets, Integral Equations Operator Theory 89 (2017), no. 1, 125 -- 149. [9] R. V. Kadison and J. R. Ringrose, Fundamentals of the theory of operator algebras: Advanced theory, vol. II, Academic Press, London, 1986. [10] E. Kirchberg, On nonsemisplit extensions, tensor products and exactness of group C ∗-algebras, Invent. Math. 112 (1993), no. 3, 449 -- 489. [11] S.A. Kruglyak, V.I. Rabanovich, and Y.S. Samoılenko, On sums of projections, Funktsional. Anal. i Prilozhen. 36 (2002), no. 3, 20 -- 35, 96. [12] A. Muller-Hermes and C. Perry, All unital qubit channels are 4-noisy operations, Lett. Math. Phys. (2018), 1 -- 9. [13] M. Musat, On factorizable quantum channels with finite dimensional ancillas, In preparation, 2018. [14] V.I. Paulsen, S. Severini, D. Stahlke, I.G. Todorov, and A. Winter, Estimating quantum chromatic numbers, J. Funct. Anal. 270 (2016), no. 6, 2188 -- 2222. [15] W. Slofstra, The set of quantum correlations is not closed, arXiv:1703.08618, 2017. [16] J.A. Smolin, F. Verstraete, and A. Winter, Entanglement of assistance and multipartite state distillations, Phys. Rev. A 72 (2005), no. 5, 052317. Magdalena Musat Department of Mathematical Sciences University of Copenhagen Universitetsparken 5, DK-2100, Copenhagen Ø Universitetsparken 5, DK-2100, Copenhagen Ø Denmark [email protected] Mikael Rørdam Department of Mathematical Sciences University of Copenhagen Denmark [email protected] Narutaka Ozawa RIMS, Kyoto University Sakyo-ku, Kyoto 606-8502 Japan [email protected] 16
1908.06343
1
1908
2019-08-17T22:35:57
The Cuntz semigroup and the radius of comparison of the crossed product by a finite group
[ "math.OA" ]
Let G be a finite group, let A be an infinite-dimensional stably finite simple unital C*-algebra, and let \alpha \colon G \to Aut (A) be an action of G on A which has the weak tracial Rokhlin property. Let A^{\alpha} be the fixed point algebra. Then the radius of comparison satisfies rc (A^{\alpha}) \leq rc (A) and rc ( C* (G, A, \alpha) ) \leq ( 1 / card (G) ) rc (A). The inclusion of A^{\alpha} in A induces an isomorphism from the purely positive part of the Cuntz semigroup Cu (A^{\alpha}) to the fixed points of the purely positive part of Cu (A), and the purely positive part of Cu ( C* (G, A, \alpha) ) is isomorphic to this semigroup. We construct an example in which G is the two element group, A is a simple unital AH algebra, \alpha has the Rokhlin property, rc (A) > 0, rc (A^{\alpha}) = rc (A), and rc (C* (G, A, \alpha)) = (1/2) rc (A).
math.OA
math
THE CUNTZ SEMIGROUP AND THE RADIUS OF COMPARISON OF THE CROSSED PRODUCT BY A FINITE GROUP M. ALI ASADI-VASFI, NASSER GOLESTANI, AND N. CHRISTOPHER PHILLIPS 1 Abstract. Let G be a finite group, let A be an infinite-dimensional stably finite simple unital C*-algebra, and let α : G → Aut(A) be an action of G on A which has the weak tracial Rokhlin property. Let Aα be the fixed point algebra. Then the radius of comparison satisfies rc(Aα) ≤ rc(A) and rc(cid:0)C ∗(G, A, α)(cid:1) ≤ · rc(A). The inclusion of Aα in A induces an isomorphism from the card(G) purely positive part of the Cuntz semigroup Cu(Aα) to the fixed points of the purely positive part of Cu(A), and the purely positive part of Cu(cid:0)C ∗(G, A, α)(cid:1) is isomorphic to this semigroup. We construct an example in which G = Z/2Z, A is a simple unital AH algebra, α has the Rokhlin property, rc(A) > 0, rc(Aα) = rc(A), and rc(C ∗(G, A, α)) = 1 2 rc(A). 9 1 0 2 g u A 7 1 ] . A O h t a m [ 1 v 3 4 3 6 0 . 8 0 9 1 : v i X r a Contents 1. Introduction 2. Preliminaries 2.1. Cuntz subequivalence 2.2. Quasitraces on C*-algebras 2.3. Radius of comparison 2.4. The radius of comparison of a corner 2.5. Approximation lemmas Injectivity of Cu+(Aα) → Cu+(A)α 3. 4. Radius of comparison of the fixed point algebra and crossed product 5. Surjectivity of Cu+(Aα) → Cu+(A)α 6. An example 7. Open problems 8. Acknowledgments References 1 4 4 7 8 9 10 11 21 27 35 44 46 46 1. Introduction We prove that if G is a finite group, A is an infinite-dimensional stably finite simple unital C*-algebra, and α : G → Aut(A) is an action of G on A which has the weak tracial Rokhlin property, then the radii of comparison (see below for further Date: 17 August 2019. 2010 Mathematics Subject Classification. Primary 46L55; Secondary 19K14; 46L80. 1 2 M. ALI ASADI-VASFI, NASSER GOLESTANI, AND N. CHRISTOPHER PHILLIPS discussion) of A, the crossed product, and the fixed point algebra are related by rc(Aα) ≤ rc(A) and rc(cid:0)C∗(G, A, α)(cid:1) ≤ 1 card(G) · rc(A). See Theorem 4.1 and Theorem 4.5. These inequalities fail for general pointwise outer actions; see Example 6.22. In fact, we prove a much stronger result, relating the Cuntz semigroups (see below and Section 2 for further discussion): the inclusion of Aα in A induces an isomorphism from the subsemigroup Cu+(Aα) ⊆ Cu(Aα) consisting of zero and the purely positive elements (recalled in Definition 3.8) to the fixed points of Cu+(A) under the action induced by α. By Example 4.7, the restriction to the purely positive part is necessary. If A has stable rank one, then one can use W(A) in place of Cu(A). We consider this to be a striking result, since the Cuntz semigroup is often considered to be too complicated to compute for C*-algebras without strict comparison. We further give an example of an infinite-dimensional stably finite simple uni- tal C*-algebra A and an action α : Z/2Z → Aut(A) which even has the Rokhlin property, and for which the radii of comparison of A, the crossed product, and the fixed point algebra are all strictly positive. It is initially not obvious that such an example should exist. The algebra A even has stable rank one. Along the way, we estimate (Theorem 2.18) the radius of comparison of a corner of a simple unital C*-algebra. This result is surely known, but we have not found it in the literature. We also prove (Lemma 4.4) that if A is simple and unital, G has order n, and α : G → Aut(A) has the weak tracial Rokhlin property, then every quasitrace on C∗(G, A, α) takes the value 1 n on the average of the unitaries in the crossed product which correspond to the elements of G. The importance of the Cuntz semigroup has become apparent in work related to the Elliott classification program. See [3] for a survey of many aspects of the Cuntz semigroup. It is generally large and complicated; roughly speaking, among simple nuclear C*-algebras, the classifiable ones are those whose Cuntz semigroups are easily accessible. With the near completion of the Elliott program, attention is turning to nonclassifiable C*-algebras, and the Cuntz semigroup is the main additional available invariant. Given its complexity, it is somewhat surprising that there is such a strong connection between the Cuntz semigroup of a simple C*- algebra and the Cuntz semigroup of its crossed product by a weak tracial Rokhlin action. It seems, also by comparison with [38], that the purely positive part of the Cuntz semigroup does not see differences which are "small in trace". The radius of comparison is a numerical invariant, based on the Cuntz semigroup, which was introduced in Section 6 of [44] to distinguish examples of nonisomorphic simple separable unital AH algebras with the same Elliott invariant. Its importance goes well beyond this application. For example, it is now conjectured that if h is a minimal homeomorphism of a compact metric space X, then rc(cid:0)C∗(Z, X, h)(cid:1) is equal to half the mean dimension of h; mean dimension is an invariant introduced in dynamics which at the time had no apparent connection with C*-algebras. The radius of comparison also plays a key role in a recent example of a simple separable unital AH algebra whose Elliott invariant has an automorphism not implemented by any automorphism of the algebra [22]. The weak tracial Rokhlin property (Definition 2.2 of [15]; see Definition 3.2 below) is a generalization of the tracial Rokhlin property (Definition 1.2 of [35]) RADIUS OF COMPARISON OF THE CROSSED PRODUCT 3 which uses positive elements instead of projections. It is a slight modification of the generalized tracial Rokhlin property of Definition 5.2 of [20]. It is much more common than the Rokhlin property, any of the higher dimensional Rokhlin prop- erties with commuting towers, or even the tracial Rokhlin property. See Example 3.12 of [36] for a collection of examples of actions of finite groups which have the tracial Rokhlin property (and hence the weak tracial Rokhlin property) but not It is shown in [2] that if A is a simple C*-algebra which the Rokhlin property. is tracially Z-absorbing, and if the minimal tensor product A⊗n of n copies of A is finite, then the permutation action of Sn on A⊗n has the weak tracial Rokhlin property. Using Z ⊗n ∼= Z, one gets in particular an action of Sn on Z which has the weak tracial Rokhlin property. (This was proved for the generalized tracial Rokhlin property in Example 5.10 of [20].) However, by Corollary 4.8(1) of [21], there is no action on Z which has any higher dimensional Rokhlin property with commuting towers. The Rokhlin property case of our Cuntz semigroup is already known, even for nonunital C*-algebras (Theorem 4.1 of [16]). That paper does not consider the radius of comparison, does not consider W(A), and gives no example like that in our Section 6, in which the algebra is simple and has nonzero radius of comparison. Moreover, as pointed out above, the weak tracial Rokhlin property is much more common than the Rokhlin property. Something close to the case rc(A) = 0 of our radius of comparison result is also already known. Let A be a simple separable nuclear unital C*-algebra. If rc(A) = 0, and if one assumes that the set of extreme points of T(A) is compact and finite-dimensional then A is Z-stable (Corollary 7.9 of [29]; Corollary 1.2 of [43]; Corollary 4.7 of [45]) and, in particular, tracially Z-absorbing in the sense of Def- inition 2.1 of [20]. If G is finite and α : G → Aut(A) has the generalized tracial Rokhlin property (Definition 5.2 of [20]), then C∗(G, A, α) is tracially Z-absorbing by Theorem 5.6 of [20]. Therefore A has strict comparison by Theorem 3.3 of [20]. (The group need not be finite; see Definition 6.1 of [20] and Theorem 6.7 of [20] for results for Z, and [30] for some results for actions of countable amenable groups.) Despite the relative abundance of actions with the weak tracial Rokhlin prop- erty, it is not obvious that there are actions with this property on stably finite simple unital C*-algebras with strictly positive radius of comparison. Getting the Rokhlin property seems even harder. For example, according to Theorems 3.4 and 3.5 of [24], if in addition A is nuclear, satisfies the Universal Coefficient The- orem, and has tracial rank zero (or is a unital Kirchberg algebra satisfying the Universal Coefficient Theorem), and α : G → Aut(A) is an action of G on A which has the Rokhlin property and is trivial on K-theory, then A is stable under ten- soring with the card(G)∞ UHF algebra. The same conclusion, under somewhat different hypotheses, is obtained in Theorem 5.10 of [16]. One might naively expect something like this to be true more generally. In fact, though, we exhibit an ac- tion α : G → Aut(A) with the Rokhlin property (not just the weak tracial Rokhlin property), in which A is a simple unital AH algebra, G = Z/2Z (although a similar construction will work for any finite group), and A, Aα, and C∗(G, A, α) all have finite but nonzero radius of comparison. When G has order n and α : G → Aut(A) is an action of G on A which has the Rokhlin property, the usual method of proving properties of C∗(G, A, α) is local approximation by algebras of the form Mn(eAe) for suitable projections e ∈ A. 4 M. ALI ASADI-VASFI, NASSER GOLESTANI, AND N. CHRISTOPHER PHILLIPS See Theorem 3.2 of [32]; there are a number of applications of this method in that paper. Weaker versions of this are true for versions of the weak tracial Rokhlin property, and are implicit in Sections 5 and 6 of [20] and in [30]. This method does not seem to work even for the Rokhlin property case of our radius of comparison result; the best we could get this way is rc(cid:0)C∗(G, A, α)(cid:1) ≤ rc(A). The difficulty is with rc(eAe). Instead, we first prove that rc(Aα) ≤ rc(A). Using the notation of Definition 2.14 and Definition 2.15 below, suppose we have a, b ∈ (Aα)+ with dτ (a) + rc(A) < dτ (b) for every normalized quasitrace τ on Aα. This applies in particular for normalized quasitraces τ on A, so a -A b. Thus, there is v ∈ A such that kvbv∗ − ak is small. Now use the Rokhlin property to "average" v over G, as described in Remark 10.3.9 and Exercise 10.3.10 of [17], getting w ∈ Aα such that kwbw∗ − ak is small. The generalization to the weak tracial Rokhlin property uses the same idea, but is considerably more technical, and requires generalizations of some of the Cuntz comparison results in [38]. The construction of an action with the Rokhlin property is a modification of the idea used in [22] to find an automorphism of the Elliott invariant of a simple AH algebra which does not lift to an automorphism of the algebra. The construc- tion there "merged" two direct systems whose direct limits had different radii of comparison but the same Elliott invariant. This "merging" was done by adding a very small number of maps which go from one of the original systems to the other. Here, we "merge" two copies of the same direct system, with the system having been chosen so that its direct limit has large radius of comparison. The action exchanges the two copies of this system. The Rokhlin projections are, roughly speaking, the identities of the algebras in the two original systems. This paper is organized as follows. Section 2 contains information on the Cuntz semigroup, Cuntz comparison, quasitraces, and the radius of comparison. It also contains several approximation results which are used repeatedly. Some of this material is new or at least not in the literature, and some definitions and results are stated here for the convenience of the reader and for easy reference. In Section 3 we prove injectivity on the purely positive part for the map Cu(Aα) → Cu(A)α. This is enough to prove the bound on the radius of comparison of a crossed product by an action with the weak tracial Rokhlin property, and our surjectivity result does not seem to help with the reverse inequality, so we prove the bound in Section 4. Section 5 contains our surjectivity result on the purely positive part for the map Cu(Aα) → Cu(A)α, as well as results on W(Aα) → W(A)α when A has stable rank one. Since W(A) is not considered in [16], we also prove the corresponding result for Rokhlin actions on unital but not necessarily simple C*-algebras. In Section 6, In Section 7, we state a few open we construct the example referred to above. problems. 2. Preliminaries In this section, we collect for easy reference some information on the Cuntz semigroup, quasitraces, and the radius of comparison. A fair amount is already in the literature, but there are several facts we did not find, among them, the estimate in Theorem 2.18 for the radius of comparison of a corner. Lemma 2.6 is definitely new. 2.1. Cuntz subequivalence. RADIUS OF COMPARISON OF THE CROSSED PRODUCT 5 Notation 2.1. We use the following standard notation. If A is a C*-algebra, or if A = M∞(B) for a C*-algebra B, we write A+ for the set of positive elements of A. Parts (1) and (2) of the following definition are originally from [8]. The usual notation for Cuntz subequivalence is a - b. We include A in the notation because we need to use Cuntz subequivalence with respect to subalgebras. Definition 2.2. Let A be a C*-algebra. (1) For a, b ∈ M∞(A)+, we say that a is Cuntz subequivalent to b in A, written a -A b, if there is a sequence (vn)∞ vnbv∗ n=1 in M∞(A) such that n = a. lim n→∞ (2) We say that a and b are Cuntz equivalent in A, written a ∼A b, if a -A b and b -A a. This relation is an equivalence relation, and we write haiA for the equivalence class of a. We define W(A) = M∞(A)+/ ∼A, together with the commutative semigroup operation haiA + hbiA = ha ⊕ biA and the partial order haiA ≤ hbiA if a -A b. We write 0 for h0iA. (3) We take Cu(A) = W(K⊗A). We write the classes as haiA for a ∈ (K⊗A)+. (4) Let A and B be C*-algebras, and let ϕ : A → B be a homomorphism. We use the same letter for the induced maps Mn(A) → Mn(B) for n ∈ Z>0 and M∞(A) → M∞(B). We define W(ϕ) : W(A) → W(B) and Cu(ϕ) : Cu(A) → Cu(B) by haiA 7→ hϕ(a)iB for a ∈ M∞(A)+ or a ∈ (K ⊗ A)+ as appropriate. Definition 2.3. Let A be a C*-algebra, let a ∈ A+, and let ε ≥ 0. Let f : [0,∞) → [0,∞) be the function f (t) = max(0, t−ε) = (t−ε)+. Then, by functional calculus, define (a − ε)+ = f (a). Part (1) of the following is taken from Proposition 2.4 of [40]. Parts (2) and (3a) are Lemma 2.5(i) and Lemma 2.5(ii) of [27]. Part (3b) is Lemma 2.2 of [28]. Part (3c) is Corollary 1.6 of [38]. Part (4) is taken from the discussion after Definition 2.3 of [27] and Proposition 2.3(ii) of [12]. Lemma 2.4. Let A be a C*-algebra. (1) Let a, b ∈ A+. Then the following are equivalent: (a) a -A b. (b) (a − ε)+ -A b for all ε > 0. (c) For every ε > 0 there is δ > 0 such that (a − ε)+ -A (b − δ)+. (2) Let a ∈ A+ and let ε1, ε2 > 0. Then (cid:0)(a − ε1)+ − ε2(cid:1)+ =(cid:0)a − (ε1 + ε2)(cid:1)+. (3) Let a, b ∈ A+ and let ε > 0. If ka − bk < ε, then: (a) (a − ε)+ -A b. (b) There is a contraction d in A such that dbd∗ = (a − ε)+. (c) For any λ > 0, we have (a − λ − ε)+ -A (b − λ)+. Lemma 2.5. Suppose t ∈ [0, 1] and s ∈ [0, 1). Then: (4) Let c ∈ A and let λ ≥ 0. Then (c∗c − λ)+ ∼A (cc∗ − λ)+. (1) 2t − t2 − s > 0 if and only if t − 1 + √1 − s > 0. (2) 1 − √1 − s ≥ s 2 . Proof. These statements are easy to check. (cid:3) (2.1) (2.2) 6 M. ALI ASADI-VASFI, NASSER GOLESTANI, AND N. CHRISTOPHER PHILLIPS The following lemma is a generalization of Lemma 1.8 of [38] or Lemma 12.1.5 of [17]. Lemma 2.6. Let A be a unital C*-algebra, let a, g ∈ A satisfy 0 ≤ a, g ≤ 1, and let ε1, ε2 ≥ 0. Then Proof. We may clearly assume ε2 < 1. Set h = 2g − g2. Functional calculus and Lemma 2.5(1) imply that ε2 2(cid:17)+ . (cid:0)a − (ε1 + ε2)(cid:1)+ -A(cid:0)(1 − g)a(1 − g) − ε1(cid:1)+ ⊕(cid:16)g − (h − ε2)+ ∼A(cid:0)g − [1 − (1 − ε2)1/2](cid:1)+. (cid:0)g − 1 + (1 − ε2)1/2(cid:1)+ -A(cid:16)g − 2(cid:17)+ ε2 . Since ε2 ∈ [0, 1), it follows from Lemma 2.5(2) that 1 − √1 − ε2 ≥ ε2 2 . So Set b = (cid:0)(1 − g)a(1 − g) − ε1(cid:1)+. Using Lemma 1.5 of [38] at the first step, Lemma 2.4(4) at the second step, kak ≤ 1 and Lemma 1.7 of [38] on the second summand at the third step, (2.1) at the fourth step, and (2.2) at the last step, we get (cid:0)a − (ε1 + ε2)(cid:1)+ -A(cid:0)a1/2(1 − h)a1/2 − ε1(cid:1)+ ⊕(cid:0)a1/2ha1/2 − ε2(cid:1)+ ∼A(cid:0)(1 − g)a(1 − g) − ε1(cid:1)+ ⊕(cid:0)h1/2ah1/2 − ε2(cid:1)+ ∼ b ⊕(cid:0)g − 1 + (1 − ε2)1/2(cid:1)+ -A b ⊕(cid:16)g − 2(cid:17)+ -A b ⊕ (h − ε2)+ ε2 . This completes the proof. (cid:3) Let a, b ∈ A+. If a -A b then by definition there is a sequence (vn)∞ n=1 in A such that limn→∞ vnbv∗ n = a. But there need not be a bounded sequence with this property. As a substitute, we have the following result, originally from [2]. We give a proof for the sake of completeness. (There is a similar result in Lemma 2.4(ii) of [28], but there is a gap in the proof.) Lemma 2.7. Let A be a C*-algebra, let a, b ∈ A+, and let δ > 0. If a -A (b− δ)+, then there exists a sequence (wn)n∈Z>0 in A such that ka − wnbw∗ nk → 0 and kwnk ≤ kak1/2δ−1/2 for every n ∈ Z>0. Proof. Let n ∈ Z>0. Since a -A (b − δ)+, there exists vn ∈ A such that Using Lemma 2.4(3b), we find a contraction dn ∈ A such that nk < 1 n . ka − vn(b − δ)+v∗ (cid:16)a − n(cid:17)+ 1 = dnvn(b − δ)+v∗ nd∗ n. Now, applying Lemma 2.4(i) of [28], we get wn ∈ A such that (cid:16)a − 1 = wnbw∗ n n(cid:17)+ n → a and kwnk ≤ kak1/2δ−1/2. kwnk ≤(cid:13)(cid:13)(cid:13)(cid:16)a − and 1 n(cid:17)+(cid:13)(cid:13)(cid:13) Therefore wnbw∗ 1/2 δ−1/2. (cid:3) RADIUS OF COMPARISON OF THE CROSSED PRODUCT 7 2.2. Quasitraces on C*-algebras. The following definition is from [19]. Parts (1), (2), and (3) correspond to the definition of a quasitrace in [6]. What we and [19] call a quasitrace is called a "2-quasitrace" in [6]. Definition 2.8. Let A be a C*-algebra. A function τ : A → C is a quasitrace if the following hold: (1) τ (x∗x) = τ (xx∗) ≥ 0 for all x ∈ A. (2) τ (a + ib) = τ (a) + iτ (b) for a, b ∈ Asa. (3) τB is linear for every commutative C*-subalgebra B ⊆ A. (4) There is a function τ2 : M2(A) → C satisfying (1), (2), and (3) with M2(A) j,k=1 denoting the standard system in place of A, and such that, with (ej,k)2 of matrix units in M2(C), for all x ∈ A we have τ (x) = τ2(x ⊗ e1,1). A quasitrace τ on a unital C*-algebra is normalized if τ (1) = 1. The set of normal- ized quasitraces on A is denoted by QT(A). All quasitraces on a unital exact C*-algebra are traces, by Theorem 5.11 of [19]. Proposition 2.9 ([6]). Let A be a stably finite unital C*-algebra. Then QT(A) 6= ∅. Proof. This is in the discussion after Proposition II.4.6 of [6]. (cid:3) Part (1) of the following proposition is Corollary II.2.3 of [6] and Parts (2) through (6) are taken from Corollary II.2.5 of [6]. That paper uses kτk instead of N (τ ). We want to avoid conflict with the definition of the norm of a linear functional. Proposition 2.10 ([6]). Let τ : A → C be a quasitrace on a C*-algebra A, and define Then: N (τ ) = sup(cid:0)(cid:8)τ (a) : a ∈ A+ and kak ≤ 1(cid:9)(cid:1). (1) N (τ ) < ∞. (2) If A is unital and τ ∈ QT(A), then N (τ ) = 1. (3) τ is order-preserving. (4) If a, b ∈ Asa, then τ (a) − τ (b) ≤ N (τ )ka − bk. (5) τ is norm-continuous. (6) If a, b ∈ A+, then τ (a + b) ≤ 2(cid:0)τ (a) + τ (b)(cid:1). Proposition 2.11 ([6]). Let A be a C*-algebra and let τ be a quasitrace on A. Then τ extends uniquely to a quasitrace τ∞ on M∞(A) such that, with (ej,k)∞ denoting the standard system of matrix units in M∞(C), we have τ∞(a⊗ej,j) = τ (a) for all a ∈ A and j ∈ Z>0. j,k=1 We denote the restriction of τ∞ to Mn(A) by τn. When no confusion is likely, we abbreviate τ∞ and τn to τ . Proof of Proposition 2.11. For Mn(A) in place of M∞(A), this is Proposition II.4.1 of [6]. By uniqueness there, for all n ∈ Z>0, the restriction to Mn(A) of the extension to Mn+1(A) is the extension to Mn(A). This implies existence of the (cid:3) extension to M∞(A), and uniqueness is now immediate. 8 M. ALI ASADI-VASFI, NASSER GOLESTANI, AND N. CHRISTOPHER PHILLIPS The following lemma is part of Proposition 3.2 of [19]. (There is a misprint there: it cites Theorem I.1.1 of [6], but apparently Theorem I.1.17 is intended.) Given Proposition 2.10, we can give a simple direct proof, which is the same as for traces except for an extra factor of 2 in the proof of closure under addition. Lemma 2.12. Let τ be a quasitrace on a C*-algebra A. Then the set Jτ = {x ∈ A : τ (x∗x) = 0} is a closed two-sided ideal in A. Proof. It is obvious that Jτ is closed under scalar multiplication and x 7→ x∗. Let x, y ∈ Jτ . Then so, by Proposition 2.10(3) and Proposition 2.10(6), (x + y)∗(x + y) ≤ (x + y)∗(x + y) + (x − y)∗(x − y) = 2x∗x + 2y∗y, 0 ≤ τ(cid:0)(x + y)∗(x + y)(cid:1) ≤ τ(cid:0)2x∗x + 2y∗y(cid:1) ≤ 4(cid:0)τ (x∗x) + τ (y∗y)(cid:1) = 0. Hence x + y ∈ Jτ . Let x ∈ Jτ and let a ∈ A. Then, using Proposition 2.10(3), 0 ≤ τ ((ax)∗(ax)) ≤ ka∗akτ (x∗x) = 0, so ax ∈ Jτ . Now xa = (a∗x∗)∗ ∈ Jτ . (cid:3) We will need Murray-von Neumann equivalence. We use notation which distin- guishes it from Cuntz equivalence. Definition 2.13. Let A be a C*-algebra, and let p, q ∈ K ⊗ A be projections. We say p is Murray-von Neumann subequivalent to q, denoted p / q, if there exists v ∈ K ⊗ A such that p = vv∗ and v∗v ≤ q. We say that p and q are Murray-von Neumann equivalent , denoted p ≈ q, if there exists v ∈ K ⊗ A such that p = vv∗ and v∗v = q. It is well known that p / q if and only if p -A q. However, it is in general not true that p ∼A q implies p ≈ q. For example, this fails in a purely infinite simple C*-algebra with nonzero K0-group. However, if A is stably finite then p ∼A q and p ≈ q are equivalent. 2.3. Radius of comparison. The following definition is Definition 12.1.7 of [17]. Definition 2.14. Let A be a unital C*-algebra, and let τ ∈ QT(A). Recalling the notation of and after Proposition 2.11, define dτ : M∞(A)+ → [0,∞) by dτ (a) = lim n→∞ τ (a1/n) for a ∈ M∞(A)+. We also use the same notation for the corresponding functions on Cu(A) and W(A). The following is Definition 6.1 of [44], except that we allow r = 0 in (1). This change makes no difference. Definition 2.15. Let A be a stably finite unital C*-algebra. (1) Let r ∈ [0,∞). We say that A has r-comparison if whenever a, b ∈ M∞(A)+ satisfy dτ (a) + r < dτ (b) for all τ ∈ QT(A), then a -A b. RADIUS OF COMPARISON OF THE CROSSED PRODUCT 9 (2) The radius of comparison of A, denoted rc(A), is rc(A) = inf(cid:0)(cid:8)r ∈ [0,∞) : A has r-comparison(cid:9)(cid:1) if it exists, and ∞ otherwise. If A is simple, then the infimum in Definition 2.15(2) is attained, that is, A has rc(A)-comparison; see Proposition 6.3 of [44]. For exact C*-algebras, one only needs to consider extreme tracial states; see Lemma 2.3 of [11]. By Proposition 6.12 of [38], the radius of comparison of a simple unital C*- algebra is the same whether computed using W(A) or Cu(A). 2.4. The radius of comparison of a corner. We give bounds on the radius of comparison of a full corner in a matrix algebra over a C*-algebra. The result is surely known, but we have not seen a proof in the literature. It will be needed in Section 4 below, to relate rc(Aα) to rc(C∗(G, A, α)). Lemma 2.16. Let A be a stably finite unital C*-algebra, let n ∈ Z>0, and let p be a full projection in Mn(A). Then, recalling the notation in and after Proposition 2.11: (1) inf τ ∈QT(A) τn(p) > 0. (2) The map θ : QT(A) → QT(pMn(A)p), given by τ 7→ 1 τn(p) τnpMn(A)p, is bijective. Proof. Since τ 7→ 1 that it suffices to prove the result when n = 1. n τn is a bijection from QT(A) to QT(cid:0)Mn(A)(cid:1), it is easily seen Since A is unital and p is full in A, it follows that ApA = A. Therefore, using Lemma 2.12, τ (p) > 0 for all τ ∈ QT(A). Since QT(A) is nonempty (by Proposi- tion 2.9) and compact, and since τ 7→ τ (p) is continuous, (1) follows. τ (p) τpAp ∈ QT(pAp). Bijectivity of θ now follows from (cid:3) Proposition II.4.2 of [6] and Proposition 2.11. To prove (2), clearly 1 Lemma 2.17. Let A be a stably finite unital C*-algebra, let n ∈ Z>0, and let p be a full projection in Mn(A). Recalling the notation in and after Proposition 2.11, if λ = inf(cid:0)(cid:8)τn(p) : τ ∈ QT(A)(cid:9)(cid:1), then 0 < λ ≤ n and rc(cid:0)pMn(A)p(cid:1) ≤ 1 λ · rc(A). Proof. By Lemma 2.16(1) we have λ > 0. Since p ≤ 1Mn(A) and τn(1Mn(A)) = n for all τ ∈ QT(A), and since QT(A) 6= ∅ by Proposition 2.9, it follows from Proposition 2.10(3) that λ ≤ n. Now let m ∈ Z>0, let a, b ∈ Mm(pMn(A)p)+ ⊆ Mmn(A)+, and suppose that dρ(a) + 1 λ · rc(A) < dρ(b) for all ρ ∈ QT(pMn(A)p). By Lemma 2.16(2), this is the same as dτ (a) + τn(p) λ · rc(A) < dτ (b) for all τ ∈ QT(A). Since λ ≤ τn(p), it follows that a -A b, so a -pMn(A)p b. Theorem 2.18. Let A be a stably finite unital C*-algebra, let n ∈ Z>0, and let p be a full projection in Mn(A). Recalling the notation in and after Proposition 2.11, define (cid:3) λ = inf(cid:0)(cid:8)τn(p) : τ ∈ QT(A)(cid:9)(cid:1) and η = sup(cid:0)(cid:8)τn(p) : τ ∈ QT(A)(cid:9)(cid:1). 10 M. ALI ASADI-VASFI, NASSER GOLESTANI, AND N. CHRISTOPHER PHILLIPS Then 0 < λ ≤ η ≤ n and 1 η · rc(A) ≤ rc(cid:0)pMn(A)p(cid:1) ≤ 1 λ · rc(A). 1 (2.3) Proof. The parts involving λ are Lemma 2.17. Since QT(A) 6= ∅ (by Proposi- tion 2.9), the relations λ ≤ η ≤ n are clear. Since p is full, there are m ∈ Z>0 and a projection q ∈ Mm(pMn(A)p) such that 1A ≈ q. Then A ∼= qMm(pMn(A)p)q. Apply Lemma 2.17 with pMn(A)p in place of A, with m in place of n, and with q in place of p. We get By Lemma 2.16(2), rc(A) ≤ inf(cid:0)(cid:8)σm(q) : σ ∈ QT(pMn(A)p)(cid:9)(cid:1)! rc(pMn(A)p). QT(pMn(A)p) =(cid:8)τn(p)−1τnpMn(A)p : τ ∈ QT(A)(cid:9) . (2.4) If σ ∈ QT(pMn(A)p) and τ ∈ QT(A) is the corresponding quasitrace from (2.4), then, using q ≈ 1A, we get σm(q) = τmn(q) inf(cid:0)(cid:8)σm(q) : σ ∈ QT(pMn(A)p)(cid:9)(cid:1) = sup(cid:0)(cid:8)τn(p) : τ ∈ QT(A)(cid:9)(cid:1) = So (2.3) implies that 1 τn(p) . Therefore τn(p) = 1 1 η . (cid:3) 1 η · rc(A) ≤ rc(pMn(A)p). 2.5. Approximation lemmas. This subsection contains several approximation lemmas which will be needed frequently. Lemma 2.19. Let M ∈ (0,∞), let f : [0, M ] → C be continuous, and let ε > 0. Then there is δ > 0 such that whenever A is a C*-algebra and a, x ∈ A satisfy a ∈ A+, kak ≤ M, kxk ≤ M, and kax − xak < δ, then kf (a)x − xf (a)k < ε. Proof. The case M = 1 is Lemma 2.5 of [4]. The proof of this version is the (cid:3) same. The statement can also be gotten from Lemma 2.5 of [4] by scaling. Lemma 2.20. Let f, g : [0,∞) → [0,∞) be continuous functions such that f (0) = g(0) = 0, let ε > 0, and let M ∈ (0,∞). Then there is δ > 0 such that whenever A is a C*-algebra, and a, b ∈ A+ satisfy kabk < δ and kak, kbk ≤ M , then kf (a)g(b)k < ε. This lemma can be proved by approximating a and b by positive elements whose product is zero, but a direct proof seems just as easy. Proof of Lemma 2.20. Without loss of generality M ≥ 1 and ε < 1. Set C = max(cid:0)kf[0,M]k∞, kg[0,M]k∞(cid:1). Choose m, n ∈ Z≥0 and Pm k=1 αkλk and g0(λ) =Pn l=1 βlλl for λ ∈ [0, M ], satisfy α1, α2, . . . , αm, β1, β2, . . . , βn ∈ R and ε f0(λ) − f (λ) < 3(C + 1) g0(λ) − g(λ) < ε 3(C + 1) such that the polynomial functions with no constant term, given by f0(λ) = RADIUS OF COMPARISON OF THE CROSSED PRODUCT 11 for all λ ∈ [0, M ]. Without loss of generality m, n ≥ 1. Define R = max(cid:0)α1,α2, . . . ,αm,β1,β2, . . . ,βn(cid:1) + 1 and δ = 3mnR2M m+n . Now let A, a, and b be as in the hypotheses. Using M ≥ 1 at the second step, ε we have kf0(a)g0(b)k ≤ mXk=1 αk · kakk−1!kabk nXl=1 βl · kbkl−1! < (mRM m−1)δ(nRM n−1) ≤ Therefore, since ε < 1 implies kf0(a)k ≤ kf (a)k + 1, ε 3 . kf (a)g(b)k ≤ kf (a) − f0(a)k · kg(b)k + kf0(a)k · kg(b) − g0(b)k + kf0(a)g0(b)k ≤(cid:18) ε 3(C + 1)(cid:19) C + (C + 1)(cid:18) ε 3(C + 1)(cid:19) + ε 3 < ε. (cid:3) This completes the proof. 3. Injectivity of Cu+(Aα) → Cu+(A)α In this section, we prove that if G is finite, A is unital, stably finite, and simple, and α : G → Aut(A) has the weak tracial Rokhlin property, then the inclusion Aα → A induces an isomorphism from the ordered semigroup of purely positive elements Cu+(Aα)∪{0} (see Definition 3.8 below) to a subsemigroup of Cu(A). Example 4.7 shows that this result fails if we do not discard the classes of projections. Notation 3.1. Let α : G → Aut(A) be an action of a finite group G on a unital C*-algebra A. For g ∈ G, we let ug be the element of Cc(G, A, α) which takes the value 1 at g and 0 at the other elements of G. We use the same notation for its image in C∗(G, A, α). We denote by Aα the fixed point algebra, given by Aα =(cid:8)a ∈ A : αg(a) = a for all g ∈ G(cid:9). We extend this notation to the elements of various objects associated with A under the actions induced by α, getting, for example, (K⊗A)α, K0(A)α, W(A)α, Cu(A)α, etc. The following definition, without Condition (4) but also requiring kfgk = 1, appears in Definition 5.2 of [20] under the name generalized tracial Rokhlin property. Definition 2.2 of [15] includes Condition (4) but only has approximate orthogonality of the contractions. By Proposition 3.10 of [14], Definition 2.2 of [15] is equivalent to our definition. Condition (4) is needed to ensure that the trivial action on C or a purely infinite simple unital C*-algebra does not have the weak tracial Rokhlin property. Definition 3.2. Let G be a finite group, let A be a simple unital C*-algebra, and let α : G → Aut(A) be an action of G on A. We say that α has the weak tracial Rokhlin property if for every ε > 0, every finite set F ⊆ A, and every positive element x ∈ A with kxk = 1, there exist orthogonal positive contractions fg ∈ A for g ∈ G such that, with f =Pg∈G fg, the following hold: (1) kafg − fgak < ε for all g ∈ G and all a ∈ F . (2) kαg(fh) − fghk < ε for all g, h ∈ G. (3) 1 − f -A x. 12 M. ALI ASADI-VASFI, NASSER GOLESTANI, AND N. CHRISTOPHER PHILLIPS (4) kf xfk > 1 − ε. In Definition 3.2, if G 6= {1} the algebra A can't be type I, since α must be point- wise outer. (See Proposition 3.2 of [14].) Therefore A is infinite-dimensional. (For clarity, we often explicitly include infinite-dimensionality in hypotheses anyway.) Lemma 3.3. Let G be a finite group, let A be an infinite-dimensional simple unital C*-algebra, and let α : G → Aut(A) be an action of G on A which has the weak tracial Rokhlin property. Then for every ε > 0, every finite set F ⊆ A, and every positive element x ∈ A with kxk = 1, there exist positive contractions eg, fg ∈ A for g ∈ G such that, with e =Pg∈G eg and f =Pg∈G fg, the following hold: (1) kegehk < ε and kfgfhk < ε for all g, h ∈ G. (2) kaeg − egak < ε and kafg − fgak < ε for all g ∈ G and all a ∈ F . (3) αg(eh) = egh and αg(fh) = fgh for all g, h ∈ G. (4) (1 − f − ε)+ -A x. (5) kf xfk > 1 − ε. (6) e ∈ Aα, f ∈ Aα, and kfk = 1. (7) egfg = fg for all g ∈ G. For most applications, we do not need the elements eg. Also, presumably one can arrange to have kek ≤ 1, but we don't need this. Proof of Lemma 3.3. Set n = card(G). We may assume ε < 1 2 . Let F ⊆ A be a finite set, and let x ∈ A+ satisfy kxk = 1. Define (3.1) M = max(cid:16)1, max a∈F kak(cid:17) and ρ = ε 4(1 + 3n)n + 1 . Define continuous functions s, t : [0, 1] → [0, 1] by s(λ) =(cid:26) ρ−1λ 1 0 ≤ λ ≤ ρ ρ ≤ λ ≤ 1 and t(λ) =( 0 λ−ρ 1−ρ 0 ≤ λ ≤ ρ ρ ≤ λ ≤ 1. Thus, if c ∈ A+ satisfies kck ≤ 1, then (3.2) s(c)t(c) = t(c) and kt(c) − ck ≤ ρ. Use Lemma 2.20 to choose ε0 > 0 such that whenever B is a C*-algebra and a, b ∈ B+ satisfy kak ≤ 1, kbk ≤ 1, and kabk < ε0, then ks(a)s(b)k < ε 4 and kt(a)t(b)k < ε 4 . Use Lemma 2.19 to choose ε1 > 0 such that whenever B is a C*-algebra and a ∈ B+ and z ∈ B satisfy kak ≤ M , kzk ≤ M , and kaz − zak < ε1, then kt(a)z − zt(a)k < ks(a)z − zs(a)k < and ε 2 ε 2 . Define (3.3) ε′ = min(cid:18)ρ, ε0 2 , ε1 2M + 1(cid:19) . Applying Definition 3.2 with with F and x as given and with ε′ in place of ε, we get orthogonal positive contractions dg ∈ A for g ∈ G such that, with d =Pg∈G dg, (8) kadg − dgak < ε′ for all g ∈ G and all a ∈ F . the following hold: RADIUS OF COMPARISON OF THE CROSSED PRODUCT 13 (9) kαg(dh) − dghk < ε′ for all g, h ∈ G. (10) 1 − d -A x. (11) kdxdk > 1 − ε′. (3.4) Define ν =(cid:13)(cid:13)Pg∈G αg(t(d1))(cid:13)(cid:13). We claim that 1 2 To prove the claim, first use kxk = 1 and kdk ≤ 1 to get (3.5) ν − 1 < 3nρ ν > and . Second, using the second part of (3.2) at the second step, 1 − ρ ≤ 1 − ε′ < kdxdk ≤ kdk2 ≤ kdk. kαg(t(d1) − d1)k +Xg∈G αg(t(d1)) − d(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≤Xg∈G < nρ + nε′ ≤ 2nρ. (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xg∈G This relation implies that(cid:12)(cid:12)ν − kdk(cid:12)(cid:12) < 2nρ, so (3.6) and, using (3.5) and ε < 1 2 , ν < kdk + 2nρ ≤ 1 + 2nρ ≤ 1 + 3nρ, kαg(d1) − dgk (3.7) ν > kdk − 2nρ > 1 − ρ − 2nρ ≥ 1 − 3nρ > 1 − ε > Using (3.6) and (3.7), we get ν − 1 < 3nρ. The claim is proved. Define eg = αg(s(d1)) and fg = 1 ν αg(t(d1)) for g ∈ G, and define e =Pg∈G eg and f = Pg∈G fg. Clearly kegk ≤ 1 and kfgk ≤ 1 for all g ∈ G, and kfk = 1. Part (3) of the conclusion is immediate. Clearly e, f ∈ Aα, so we have (6), and (7) follows from (3.2). 1 2 . Now we claim that: (12) kfg − dgk < 2(1 + 3n)ρ for all g ∈ G. To prove the claim, use the second part of (3.2) and (9) at the second step and (3.4) at the third step to get kfg − dgk ≤(cid:13)(cid:13)(cid:13) αg(t(d1)) − αg(t(d1))(cid:13)(cid:13)(cid:13) + kαg(t(d1) − d1)k + kαg(d1) − dgk 1 ν 1 ν ν − 1 · kt(d1)k + ρ + ε′ < 2(1 + 3n)ρ, ≤ as desired. We prove Part (1) of the conclusion. For g 6= h, using dgdh = 0 at the first step, (9) at the second step, and (3.3) at the third step, we have kαg(d1)αh(d1)k ≤ kαg(d1)k · kαh(d1) − dhk + kαg(d1) − dgk · kdhk < 2ε′ ≤ ε0. The choice of ε0 and the relations eg = s(αg(d1)) and fg = ν−1t(αg(d1)) now imply kegehk < ε 4 ≤ ε and which is (1). kfgfhk < ν−2(cid:16) ε 4(cid:17) < 4(cid:16) ε 4(cid:17) = ε, We prove (2). So let g ∈ G and let a ∈ F . Using (8) and (9) at the second step, and using (3.3) at the third step, we get kaαg(d1) − αg(d1)ak ≤ 2kak · kαg(d1) − dgk + kadg − dgak < (2M + 1)ε′ ≤ ε1. 14 M. ALI ASADI-VASFI, NASSER GOLESTANI, AND N. CHRISTOPHER PHILLIPS The choice of ε1 implies kaeg − egak = kas(αg(d1)) − s(αg(d1))ak < ε 2 < ε and as desired. kafg − fgak = 1 ν kat(αg(d1)) − t(αg(d1))ak < 2(cid:16) ε 2(cid:17) = ε, To prove (4), we estimate, using (12) at the third step and (3.1) at the last step (3.8) k(1 − f ) − (1 − d)k = kd − fk ≤Xg∈G kfg − dgk < n · 2(1 + 3n)ρ < ε. Therefore, using Lemma 2.4(3a) at the first step and (10) at the second step, For (5), we estimate, using part of (3.8) at the second step, (1 − f − ε)+ -A 1 − d -A x. kdxd − f xfk ≤ kdxk · kd − fk + kd − fk · kxfk < 4n(1 + 3n)ρ. Therefore, using (11) at the second step, kf xfk > kdxdk − 4n(1 + 3n)ρ > 1 − ε′ − 4n(1 + 3n)ρ ≥ 1 − ε. (cid:3) This completes the proof. Lemma 3.4. Let G be a finite group, let A be a C*-algebra, let α : G → Aut(A) be an action of G on A, and let a, b ∈ (Aα)+. If a ∈ bAb, then a ∈ bAαb and a -Aα b. Proof. Let ε > 0. Choose c ∈ A such that ka − bcbk < ε. Then ka − bαg(c)bk < ε for all g ∈ G. So d = 1 card(G)Pg∈G αg(c) satisfies d ∈ Aα and ka − bdbk ≤ ka − b αg(c) bk < ε. 1 card(G)Xg∈G Since ε > 0 is arbitrary, a ∈ bAαb, whence also a -Aα b. (cid:3) Lemma 3.5. Let α : G → Aut(A) be an action of a finite group G on a C*-algebra A, let δ ∈(cid:0)0, [2 card(G)]−2(cid:1), and let a, b ∈ (Aα)+ with kbk = 1 and kak ≤ 1. If x is a positive element in bAb with a -A (x2− 1 2 )+ and kxαg(x)k < δ for all g ∈ G\{1}, then there exists t ∈ Aα such that: (1) (a − δ1/2)+ -Aα tt∗. (2) t∗t ∈ bAb. (3) (a − δ1/2)+ -Aα b. Proof. To prove (1), set η = √δ − 2 card(G) δ. Since δ ∈ (cid:0)0, [2 card(G)]−2(cid:1), it follows that η > 0. Since a -A (x2 − 1 2 )+, by Lemma 2.7 there exists w ∈ A such that ka − wx2w∗k < η and kwk ≤ √2. Using Lemma 2.4(3b), we find d ∈ A with kdk ≤ 1 such that (a − η)+ = dwx2w∗d∗. Set v = dw. Then √2. (3.9) card(G)Xg6=h card(G)1/2 Xg∈G (a − η)+ = vx2v∗ αg(vx) αh(xv∗). kvk ≤ αg(vx) Define and s = and 1 t = 1 Clearly t ∈ Aα. RADIUS OF COMPARISON OF THE CROSSED PRODUCT 15 Now we claim that ksk < 2 card(G) δ. To prove the claim, for every g, h ∈ G with g 6= h we have kx αg−1h(x)k < δ. So (3.10) kαg(x) αh(x)k < δ. Thus, using (3.9) and (3.10) at the second step, ksk ≤ kvk2 card(G)Xg6=h The claim follows. kαg(x) αh(x)k <(cid:18) 2 card(G)(cid:19) card(G)2δ = 2 card(G) δ. Using (3.9) at the second step and a ∈ Aα (so that (a − η)+ ∈ Aα) at the last step, we get tt∗ = = 1 card(G) Xg∈G card(G)Xg∈G 1 αg(vx2v∗) +Xg6=h αg(cid:0)(a − η)+(cid:1) + αg(vx) αh(xv∗)! card(G)Xg6=h 1 It follows that tt∗ − (a − η)+ = s. Using the claim, we get ktt∗ − (a − η)+k < 2 card(G) δ. αg(vx) αh(xv∗) = (a − η)+ + s. Therefore, using Lemma 2.4(2) at the first step and Lemma 2.4(3a) at the second step, This is (1). (a − δ1/2) =(cid:0)(a − η)+ − 2 card(G) δ(cid:1)+ -Aα tt∗. To prove (2), we claim that αg(xv∗) αh(vx) ∈ bAb for all g, h ∈ G. Since x ∈ bAb, there exists a sequence (rn)n∈Z>0 in A such that (3.11) x = lim n→∞ brnb. So for all g, h ∈ G we get, using b ∈ Aα at the first step, (3.12) αg(xv∗) αh(vx) = lim n→∞ b αg(rn) b αg(v∗) αh(v) b αh(rn) b. The claim is proved. Use the definition of t to compute (3.13) t∗t = 1 card(G) Xg,h∈G αg(xv∗) αh(vx). By (3.12) and (3.13), we have t∗t ∈ bAb, which is (2). Finally, we prove (3). Since b ∈ Aα and t∗t ∈ bAb, it follows from Lemma 3.4 that t∗t -Aα b. Therefore, using (1) at the first step and Lemma 2.4(4) at the second step, This completes the proof of the lemma. (a − δ1/2)+ -Aα tt∗ ∼Aα t∗t -Aα b. (cid:3) Lemma 3.6. Let A be an infinite-dimensional simple unital C*-algebra, and let α : G → Aut(A) be an action of a finite group G on A which has the weak tracial Rokhlin property. Then for every ε > 0 and b ∈ (Aα)+ with kbk = 1, there is a positive element x ∈ bAb with kxk = 1 such that kxαg(x)k < ε for all g ∈ G \ {1}. 16 M. ALI ASADI-VASFI, NASSER GOLESTANI, AND N. CHRISTOPHER PHILLIPS Proof. We may assume ε < 1 F = {b2} 2 . Set and ε′ = ε 8(1 + card(G)2) . Applying Definition 3.2 using F , ε′, and b2, we get positive contractions fg ∈ A for g ∈ G such that, with f =Pg∈G fg, the following hold: (1) kzfg − fgzk < ε′ for all g ∈ G and z ∈ F . (2) kαg(fh) − fghk < ε′ for all g, h ∈ G. (3) kf b2fk > 1 − ε′. So we have, using at the first step g 6= h (so that fgfh = 0), and using (1) at the last step, kfgk · kb2fh − fhb2k < card(G)2ε′. Using this and orthogonality of the elements fg for g ∈ G at the last step, we estimate Xg6=h kf b2fk ≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xg∈G kfgb2fhk ≤Xg6=h fgb2fg(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) +(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xg6=h fgb2fh(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Therefore, by (3), < max g∈G kfgb2fgk + card(G)2ε′. g∈G kfgb2fgk > 1 − ε′(1 + card(G)2) > 1 − ε > max 1 2 . It follows that there exists s ∈ G such that kfsb2fsk > 1 2 . Set y = bf 2 1 2 kyk = kbfsfsbk = kfsb2fsk > . s b. Then Now define x = kyk−1 · y. We claim that kxαg(x)k < ε for all g ∈ G \ {1}. To prove the claim, using b ∈ Aα at the first step, g 6= 1 (so that fsfgs = 0) at the second step, kbk = 1 at the third step, and (1) and (2) at the last step, we estimate (3.14) s b2αg(f 2 kyαg(y)k = kbf 2 s ) bk ≤ kbfsk · kfsb2 − b2fsk · kαg(f 2 s ) bk + kbfsb2fsk · kαg(fs) − fgsk · kαg(fs) bk ≤ kfsb2 − b2fsk + kαg(fs) − fgsk < 2ε′. Since kyk−1 < 2, we have kxαg(x)k = This completes the proof. 1 kyk2 · kyαg(y)k < 8ε′ < ε. (cid:3) Lemma 3.7. Let A be an infinite-dimensional simple unital C*-algebra and let α : G → Aut(A) be an action of a finite group G on A with the weak tracial Rokhlin property. Let a, b ∈ (Aα)+, and suppose that 0 is a limit point of sp(b). Then a -A b if and only if a -Aα b. This result holds when α has the Rokhlin property, without the requirement that 0 be a limit point of sp(b); in this case, A can be any unital C*-algebra. See Theorem 4.1(ii) of [16]. RADIUS OF COMPARISON OF THE CROSSED PRODUCT 17 Proof of Lemma 3.7. We need only prove the forwards implication. So assume that a -A b. We may assume kak ≤ 1 and kbk = 1. Let ε > 0. We may assume ε < [2card(G)]−2. Since a -A b, there is δ > 0 such that (a − ε)+ -A (b − δ)+. We may require δ < 1. Set a′ = (a − ε)+ and b′ = (b − δ)+. Choose w ∈ A such that (cid:13)(cid:13)wb′w∗ − a′(cid:13)(cid:13) < [40 card(G)]−1ε. Since b′, a′ ∈ Aα, it follows that (3.15) ε 40 card(G) (cid:13)(cid:13)αg(w)b′αg(w∗) − a′(cid:13)(cid:13) < for all g ∈ G. Choose λ ∈ sp(b) ∩ (0, δ). Let h : [0,∞) → [0, 1] be a continuous function such that h(λ) = 1 and supp(h) ⊆ (0, δ). Then (3.16) h(b) + b′ -Aα b. and kh(b)k = 1, h(b) ⊥ b′, Applying Lemma 3.6 with h(b) in place of b, we find a positive element x ∈ h(b)Ah(b) with kxk = 1 such that for all g ∈ G \ {1}. Now set F0 = {a′, b′, w, w∗} and (cid:13)(cid:13)xαg(x)(cid:13)(cid:13) < ε2 64 F = [g∈G αg(F0). Define ε 1 −1 1 and ε′ = 2(cid:17)+ ·(cid:16)x2 − 2(cid:17)+(cid:13)(cid:13)(cid:13) δ > 0 such that the following hold: 40(kwk + 1)2card(G)4 . s =(cid:13)(cid:13)(cid:13)(cid:16)x2 − Set M = max(cid:0)1, maxz∈F kzk(cid:1), and use Lemma 2.20 and Lemma 2.19 to choose (1) δ ≤ ε′. (2) If c, d ∈ A+ satisfy kck, kdk ≤ 1 and kcdk < δ, then(cid:13)(cid:13)c1/2d1/2(cid:13)(cid:13) < ε′. (3) If c ∈ A+ satisfies kck ≤ M and z ∈ A satisfies kzk ≤ M and kcz− zck < δ, then(cid:13)(cid:13)c1/2z − zc1/2(cid:13)(cid:13) < ε′. of x, we get positive contractions fg ∈ A for g ∈ G such that, with f =Pg∈G fg, Applying Lemma 3.3 with F as given, with δ in place of ε, and with s in place the following hold: (4) kfgfhk < δ for all g, h ∈ G with g 6= h. (5) kzfg − fgzk < δ for all g ∈ G and z ∈ F . (6) αg(fh) = fgh for all g, h ∈ G. (7) f ∈ Aα and kfk = 1. (8) (1 − f − δ)+ -A s. Then also: g h (cid:13)(cid:13) < ε′ for all g, h ∈ G with g 6= h. (9) (cid:13)(cid:13)f 1/2 g f 1/2 (10) (cid:13)(cid:13)zf 1/2 z(cid:13)(cid:13) < ε′ for all g ∈ G and z ∈ F . g − f 1/2 (11) αg(cid:0)f 1/2 h (cid:1) = f 1/2 8 , (8) implies that (cid:0)1 − f − ε Since δ ≤ ε′ < ε Lemma 3.5(3) with h(b) in place of b, with (cid:0)1 − f − ε gh for all g, h ∈ G. 8(cid:1)+ -A s ∼(cid:0)x2 − 1 2(cid:1)+. Applying 8(cid:1)+ in place of a, with x as 18 M. ALI ASADI-VASFI, NASSER GOLESTANI, AND N. CHRISTOPHER PHILLIPS given, and with ε2/64 in place of δ, we get ε (3.17) (cid:16)1 − f − 64(cid:19)1/2!+ Now define v =Pg∈G αg(f1w). Clearly v ∈ Aα. We claim that = (cid:16)1 − f − 8(cid:17)+ −(cid:18) ε2 4(cid:17)+ ε kvb′v∗ − f a′fk < ε 4 . -Aα h(b). To prove the claim, define a0 =Xg∈G fga′fg, It is immediate that b0 =Xg∈G f 1/2 g b′f 1/2 g , and v0 =Xg∈G αg(w)f 1/2 g . kvk, kv0k ≤ card(G)kwk. f 1/2 g , ef =Xg∈G (3.18) Also set giving(cid:13)(cid:13)ef(cid:13)(cid:13) ≤ card(G). (cid:13)(cid:13)(cid:2)αg(w)f 1/2 (3.19) For g ∈ G, by (1) and (5) we have kαg(w)fg − fgαg(w)k < ε′. Therefore, for all g ∈ G, using (3.15) on the last term at the second step and kb′k ≤ 1 at the last step, g g g g b′f 1/2 ≤ kαg(w)fg − fgαg(w)k · kb′k · kfgαg(w∗)k (cid:3)(cid:2)αg(w)f 1/2 (cid:3)(cid:2)f 1/2 + kfgαg(w)k · kb′k · kfgαg(w∗) − αg(w∗)fgk + kfgk · kαg(w)b′αg(w∗) − a′k · kfgk − fga′fg(cid:13)(cid:13) (cid:3)∗ < 2kwkε′ + 40 card(G) . ε (3.20) Set S = {(g, g, g) : g ∈ G} ⊆ G3. Using (3.19), (9), and kb′k ≤ 1 at the second step, we get g g g g b′f 1/2 kv0b0v∗ (cid:3)(cid:2)f 1/2 (cid:3)(cid:2)αg(w)f 1/2 (cid:3)(cid:2)f 1/2 0 − a0k ≤Xg∈G(cid:13)(cid:13)(cid:2)αg(w)f 1/2 + X(g,t,h)∈G3\S(cid:13)(cid:13)(cid:2)αg(w)f 1/2 − fga′fg(cid:13)(cid:13) (cid:3)∗ h (cid:3)∗(cid:13)(cid:13) (cid:3)(cid:2)αh(w)f 1/2 Next, we estimate, using f = Pg∈G fg at the first step, and using (1), (4), kfga′fhk ≤Xg6=h(cid:0)kfgk · ka′fh − fha′k + kfgfhk · ka′k(cid:1) ka0 − f a′fk ≤Xg6=h + card(G)3kwk2ε′ ≤ < 2 card(G)kwkε′ + b′f 1/2 4ε 40 ε 40 . g t t and (5) at the third step, (3.21) < 2 card(G)2ε′ ≤ 2ε 40 . A similar calculation, this time using (9), (10), and (1), gives (3.22) (cid:13)(cid:13)b0 − ef b′ef(cid:13)(cid:13) < 2 card(G)2ε′. RADIUS OF COMPARISON OF THE CROSSED PRODUCT 19 we get The next step is to estimate (cid:13)(cid:13)v − v0ef(cid:13)(cid:13). For all g ∈ G, using (5), (6), and (1), kαg(f1w) − αg(w)fgk = kαg(f1w − wf1)k < δ ≤ ε′. Now, by (10), (3.23) It follows, using this, (3.22), and (3.18) at the second step, that kvb′v∗ − v0b0v∗ g f 1/2 < card(G)ε′ + card(G)2kwkε′ ≤ card(G)2(kwk + 1)ε′. (cid:13)(cid:13)v − v0ef(cid:13)(cid:13) ≤Xg∈G(cid:13)(cid:13)αg(f1w) − αg(w)fg(cid:13)(cid:13) +Xg6=h(cid:13)(cid:13)αg(w)f 1/2 h (cid:13)(cid:13) 0k ≤(cid:13)(cid:13)v − v0ef(cid:13)(cid:13) · kb′k · kv∗k + kv0k ·(cid:13)(cid:13)ef(cid:13)(cid:13) · kb′k ·(cid:13)(cid:13)(cid:0)v − v0ef(cid:1)∗(cid:13)(cid:13) + kv0k ·(cid:13)(cid:13)ef b′ef − b0(cid:13)(cid:13) · kv∗ < card(G)2(kwk + 1)ε′ · card(G)kwk 0k + card(G)2kwk · card(G)2ε′(kwk + 1) + card(G)kwk · 2 card(G)2ε′ · card(G)kwk 4ε 40 . ≤ Combining this with (3.20) and (3.21), we now have 0k + kv0b0v∗ kvb′v∗ − f a′fk ≤ kvb′v∗ − v0b0v∗ 2ε = 40 4ε 40 4ε 40 ε 4 + + < . 0 − a0k + ka0 − f a′fk The claim is proved. The claim implies that (3.24) -Aα vb′v∗ -Aα b′. Applying Lemma 2.6 with ε of a, and with 1 − f in place of g, we get (3.25) 3ε ε 4(cid:17)+ (cid:16)f a′f − -Aα (cid:16)f a′f − 4(cid:17)+ -Aα(cid:16)f a′f − (cid:16)a′ − 4(cid:17)+ (a − ε)+ =(cid:16)a′ − ε 4(cid:17)+ . ε 4(cid:17)+ ⊕(cid:16)1 − f − 4(cid:17)+ 4(cid:17)+ ⊕(cid:16)1 − f − ε ε Using (3.25) at the second step, (3.24) and (3.17) at the third step, and (3.16) at the last step, we have 3ε 4 in place of ε1, with ε 2 in place of ε2, with a′ in place Therefore (a − ε)+ -Aα b. Definition 3.8. Let A be a C*-algebra. Following the discussion before Corollary 2.24 of [3] and Definition 3.1 of [38], with slight changes in notation, we define A++ =(cid:8)a ∈ A+ : there is no projection p ∈ M∞(A) such that haiA = hpiA(cid:9), Cu+(A) =(cid:8)haiA : a ∈ (K ⊗ A)++}, and W+(A) = Cu+(A) ∩ W(A). The elements of A++ are called purely positive. We recall some properties of W+(A) and Cu+(A). Lemma 3.9. Let A be a stably finite simple unital C*-algebra. Then: (1) Cu(A) \ W(A) ⊆ Cu+(A). -Aα b′ ⊕ h(b) -Aα b. (cid:3) 20 M. ALI ASADI-VASFI, NASSER GOLESTANI, AND N. CHRISTOPHER PHILLIPS Cu(A). (3) W+(A) ∪ {0} and Cu+(A) ∪ {0} are unital subsemigroups of W(A) and (2) (K ⊗ A)++ =(cid:8)a ∈ (K ⊗ A)+ : 0 is a limit point of sp(a)(cid:9). (4) Let η1, η2, . . . ∈ Cu+(A) ∪ {0} satisfy η1 ≤ η2 ≤ ··· . Then sup(cid:0){ηn : n ∈ Z≥0}(cid:1), evaluated in Cu(A), is in Cu+(A) ∪ {0}. Proof. Parts (1) and (2) are Lemma 3.2 of [38]. Part (3) for W(A) is Corollary 2.9(i) of [34]. For Cu(A) it is Corollary 3.3 of [38]. Part (4) is Lemma 3.5 of [38] (originally (cid:3) Parts (i) and (iv) of Proposition 6.4 of [12]). There are further interesting properties: still assuming A is stably finite and simple, Cu+(A) ∪ {0} is absorbing (this follows from Corollary 3.3 of [38]) and, if A is not of type I, has the same functionals as Cu(A) (Lemma 3.8 of [38]). We will use the following result several times. The main work for the last sentence of the proof is in [1]. Lemma 3.10. Let A be an infinite-dimensional simple unital C*-algebra, let G be a finite group, and let α : G → Aut(A) be an action of G on A which has the weak tracial Rokhlin property. Then, for every x ∈ (Aα)+ \ {0}, there exists c ∈ (Aα)+ such that c -Aα x and sp(c) = [0, 1]. Proof. The algebra A is not type I, so Theorem 4.1 of [39] implies that Aα is not type I. Since C∗(G, A, α) is simple, Lemma 4.3(4) below (or [42]) implies that Aα is simple. Apply Lemma 2.1 of [38] to xAαx. (cid:3) Lemma 3.11. Let A be a stably finite simple unital C*-algebra which is not of type I and let α : G → Aut(A) be an action of a finite group G on A with the weak tracial Rokhlin property. Let ι : Aα → A be the inclusion map. Then: (1) The map W(ι) : W(Aα) → W(A) induces an isomorphism of ordered semi- (2) The map Cu(ι) : Cu(Aα) → Cu(A) induces an isomorphism of ordered semi- groups from W+(Aα) ∪ {0} to its image in W(A). groups from Cu+(Aα) ∪ {0} to its image in Cu(A). Proof. In both parts, we need only prove injectivity and order isomorphism. ε (3.26) By Corollary 4.6 of [14], for every n ∈ Z>0 the action g 7→ idMn ⊗ αg of G on Mn(A) has the weak tracial Rokhlin property. With W+(Aα) in place of W+(Aα)∪ {0}, Part (1) now follows from Lemma 3.9(2) and Lemma 3.7. Part (1) as stated is then immediate. We prove (2). It suffices to prove that if a, b ∈ (K ⊗ Aα)++ satisfy a -A b, then Choose δ > 0 such that a -Aα b. Let ε > 0; we prove that (a − ε)+ -Aα b. (3.27) kh(b)k = 1, Choose λ ∈ sp(b)∩(cid:0)0, δ h(λ) = 1 and supp(h) ⊆(cid:0)0, δ (cid:13)(cid:13)(cid:13)(cid:13)a0 −(cid:16)a − 3(cid:17)+(cid:13)(cid:13)(cid:13)(cid:13) < ε 3 ε , -A (b − δ)+. (cid:16)a − 3(cid:17)+ 3(cid:1). Let h : [0,∞) → [0, 1] be a continuous function such that 3(cid:1). Then h(b) ⊥(cid:16)b − 3(cid:17)+ 3(cid:17)+(cid:13)(cid:13)(cid:13)(cid:13) < (cid:13)(cid:13)(cid:13)(cid:13)b0 −(cid:16)b − h(b) +(cid:16)b − kc0 − h(b)k < 3(cid:17)+ -Aα b. and Choose n ∈ Z>0 and a0, b0, c0 ∈ Mn(Aα)+ such that δ , and δ δ δ 3 , 1 3 . RADIUS OF COMPARISON OF THE CROSSED PRODUCT 21 It follows from Lemma 2.4(3c) that (3.28) and (3.29) (a − ε)+ -Aα (cid:16)a0 − (b − δ)+ -Aα(cid:16)b0 − ε 3(cid:17)+ 3(cid:17)+ δ -Aα(cid:16)a − -Aα(cid:16)b − ε 3(cid:17)+ 3(cid:17)+ δ . Set c1 =(cid:0)c0 − 1 3 , so c1 6= 0. Since the action induced by α on Mn(A) has the weak tracial Rokhlin property, Lemma 3.10 provides c ∈ Mn(Aα)+ such that c -Aα c1 and sp(c) = [0, 1]. In particular, 3(cid:1)+. Then kc1k > 1 (3.30) c -Aα h(b). At the first step combining the second part of (3.28), (3.26), and the first part of (3.29), we get δ 3(cid:17)+ ⊕ c. δ ε Since (3.31) 3(cid:17)+ (cid:16)a0 − 3(cid:17)+ a0,(cid:16)b0 − -A(cid:16)b0 − 3(cid:17)+ -A(cid:16)b0 − ∞[k=1 because 0 is a limit point of the spectrum of(cid:0)b0− δ -Aα(cid:16)b0 − Part (1) and (3.31) imply (cid:16)a0 − , c ∈ ε 3(cid:17)+ (3.32) δ Mk(Aα), 3(cid:1)+⊕ c, and using Lemma 3.9(2), 3(cid:17)+ ⊕ c. δ Using, in order, the first part of (3.28), (3.32), (3.30) and the second part of (3.29), and (3.27), we get (a − ε)+ -Aα (cid:16)a0 − ε 3(cid:17)+ This completes the proof. -Aα (cid:16)b0 − δ 3(cid:17)+ ⊕ c -Aα (cid:16)b − δ 3(cid:17)+ ⊕ h(b) -Aα b. (cid:3) Lemma 3.11 fails if we don't restrict to the purely positive elements. See Exam- ple 4.7. We postpone this example, since it uses Lemma 4.3. 4. Radius of comparison of the fixed point algebra and crossed product In the next section, we identify the range of the map Cu+(Aα) → Cu(A) when α has the weak tracial Rokhlin property: it is Cu+(A)α. This information is not needed for our estimate on the radius of comparison, and does not seem to help with the (still open) opposite inequality to the one we prove. So we prove the radius of comparison results now. Then we discuss what happens under weaker hypotheses on the action, and give the example promised at the end of Section 3. Theorem 4.1. Let G be a finite group, let A be an infinite-dimensional stably finite simple unital C*-algebra, and let α : G → Aut(A) be an action of G on A which has the weak tracial Rokhlin property. Then rc(Aα) ≤ rc(A). 22 M. ALI ASADI-VASFI, NASSER GOLESTANI, AND N. CHRISTOPHER PHILLIPS Proof. We use Theorem 12.4.4(ii) of [17]. Thus, let m, n ∈ Z>0 satisfy m Let l ∈ Z>0, and let a, b ∈ (Aα ⊗ Ml)+ with kak = kbk = 1 satisfy n > rc(A). (n + 1)haiAα + mh1iAα ≤ nhbiAα in W(Aα). Corollary 4.6 of [14], the action α ⊗ idMl : G → Aut(A ⊗ Ml), defined by j,k=1, also has the weak tracial property. We may therefore assume l = 1. j,k=1) = αg(a) ⊗ (λj,k)n (α ⊗ idMl)g (a ⊗ (λj,k)n We must prove that a -Aα b. Moreover, by Lemma 2.4(1b), it is enough to show that for every ε > 0 we have (a − ε)+ -Aα b. So let ε > 0. Without loss of generality ε < 1 (4.1) Then in W(Aα) we have km kn + 1 > rc(A). 2 . Choose k ∈ Z>0 such that (kn + 1)haiAα + kmh1iAα ≤ k(n + 1)haiAα + kmh1iAα ≤ knhbiAα. Let u ∈ M∞(Aα)+ be the direct sum of kn + 1 copies of a, let z ∈ M∞(Aα)+ be the direct sum of kn copies of b, and let q ∈ M∞(Aα)+ be the direct sum of km copies of 1A. Then, by definition, u ⊕ q -Aα z. Therefore Lemma 2.4(1c) provides δ > 0 such that(cid:0)u ⊕ q − ε(cid:1)+ -Aα (z − δ)+. Since ε < 1 (u ⊕ q − ε)+ = (u − ε)+ ⊕ (q − ε)+ ∼Aα (u − ε)+ ⊕ q, 2 , we have so (kn + 1)h(a − ε)+iAα + kmh1iAα ≤ knh(b − δ)+iAα. (kn + 1)ha′iAα + kmh1iAα ≤ knhb′iAα. Set a′ = (a − ε)+ and b′ = (b − δ)+. Then (4.2) Lemma 2.7 of [38] provides positive elements c ∈ Aα and y ∈ Aα \ {0} such that (4.3) in W(Aα). By Lemma 3.10, there is y0 ∈ (Aα)+ such that y0 -Aα y and sp(y0) = [0, 1]. Replacing y with this element, we may assume that y is purely positive. By (4.2) and (4.3), knhb′iAα ≤ (kn + 1)hciAα hciAα + hyiAα ≤ hbiAα and (kn + 1)ha′iAα + kmh1iAα ≤ (kn + 1)hciAα . This relation also holds in W(A). For τ ∈ QT(A), apply dτ and divide by kn + 1 to get dτ (a′) + km kn + 1 ≤ dτ (c). So a′ -A c by (4.1). Therefore, using Lemma 3.7 with c ⊕ y in place of b at the second step, and using (4.3) at the third step, This completes the proof. (a − ε)+ = a′ -Aα c ⊕ y -Aα b. (cid:3) Using [16], we get the same conclusion for Rokhlin actions on stably finite unital C*-algebras. Theorem 4.2. Let G be a finite group, let A be a stably finite unital C*-algebra, and let α : G → Aut(A) be an action of G on A which has the Rokhlin property. Then rc(Aα) ≤ rc(A). RADIUS OF COMPARISON OF THE CROSSED PRODUCT 23 Proof. We may clearly assume rc(A) < ∞. Let r ∈ [0,∞) and suppose that A has r-comparison. Let a, b ∈ M∞(Aα)+ satisfy dτ (a) + r < dτ (b) for all τ ∈ QT(Aα). Since every quasitrace on A restricts to a quasitrace on Aα, we have dτ (a)+r < dτ (b) for all τ ∈ QT(A). Since A has r-comparison, we get a -A b. Now a -Aα b by Theorem 4.1(ii) of [16]. So rc(Aα) ≤ r. Taking the infimum over r ∈ [0,∞) such that A has r-comparison, we get rc(Aα) ≤ rc(A). (cid:3) We now turn to the radius of comparison of the crossed product. Parts (1) -- (4) of the following lemma are originally taken from [42]. Since some properties of the projection p are needed in our computations, we give a more detailed statement. Lemma 4.3. Let G be a finite group, let A be a unital C*-algebra, and let α : G → Aut(A) be an action of G on A. Recalling from Notation 3.1 that (ug)g∈G is the family of standard unitaries in C∗(G, A, α), define p = 1 (1) p is a projection in C∗(G, A, α). (2) pap =(cid:16) 1 card(G)Pg∈G αg(a)(cid:17)p for all a ∈ A. (3) If a ∈ Aα, then pap = ap. (4) The map a 7→ ap is an isomorphism from Aα to the corner pC∗(G, A, α)p. (5) If C∗(G, A, α) has stable rank one, then Aα has stable rank one. (6) If α has the Rokhlin property, then p is full in C∗(G, A, α). card(G)Pg∈G ug. Then: Proof. Parts (1) -- (4) are computations. (Also see [42].) Next, if C∗(G, A, α) has stable rank one, then Theorem 3.1.8 of [25] implies that pC∗(G, A, α)p has stable rank one, so (5) follows from (4). For (6), let J ⊆ C∗(G, A, α) be the closed ideal generated by p, and set I = J∩A. Let E : C∗(G, A, α) → A be the standard conditional expectation. The proof of Proposition 10.3.13 of [17] shows that E(J) ⊆ I. Since E(card(G)· p) = 1, we have (cid:3) 1 ∈ I, so 1 ∈ J. The proof of the following lemma is easier, and well known, for tracial states. For example, the inequality (4.10) is trivial for tracial states, but it seems to require some effort for quasitraces. 1 Lemma 4.4. Let G be a finite group, let A be an infinite-dimensional stably finite simple unital C*-algebra, let α : G → Aut(A) be an action of G on A which has the weak tracial Rokhlin property, and let τ ∈ QT(cid:0)C∗(G, A, α)(cid:1). Let p = card(G)Pg∈G ug, as in Lemma 4.3. Then τ (p) = Proof. Let ε > 0. We show that(cid:12)(cid:12) card(G) − τ (p)(cid:12)(cid:12) < ε. By Corollary 2.5 of [38], there is a ∈ A+ \ {0} such that for all ρ ∈ QT(A), dρ(a) < (4.4) card(G) . 1 1 ε 4 . Applying Definition 3.2 with F = ∅, with [32 card(G)]−1ε in place of ε, and with kak−1 · a in place of x, we get orthogonal positive contractions fg ∈ A for g ∈ G such that, with f =Pg∈G fg, we have (4.5) 1 − f -A a and kαg(fh) − fghk < ε 32 card(G) 24 M. ALI ASADI-VASFI, NASSER GOLESTANI, AND N. CHRISTOPHER PHILLIPS for all g, h ∈ G. This inequality, together with kfgk,kfghk ≤ 1, implies (4.6) ghk ≤ kαg(fh)k · kαg(fh) − fghk + kαg(fh) − fghk · kfghk h) − f 2 kαg(f 2 ε < 16 card(G) . Now we claim that the following hold: 0 ≤ τ (1) − τ (f 2) < ε 4 , ε 4 , 1 0 ≤ τ (p) − τ (pf 2p) < (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)Xh∈G hp(cid:1) − τ(cid:18)(cid:20) τ(cid:0)pf 2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) τ Xh∈G hp! −Xh∈G card(G)(cid:21)f 2(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) hp)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) τ (pf 2 pf 2 < < ε 4 , ε 4 . 0 ≤ τ (1 − f 2) ≤ dτ (1 − f 2) ≤ dτ (a) < ε 4 . (4.7) (4.8) (4.9) and (4.10) (4.11) We prove (4.7). Since sp(f ) ⊆ [0, 1], we have 1 − f 2 ∼A 1 − f , so 1 − f 2 -A a by (4.5). Clearly τA ∈ QT(A). Therefore, using (4.4) at the last step, The relation (4.7) follows because 1 and f 2 commute. To prove (4.8), we start with (1 − f 2)1/2p(1 − f 2)1/2 ≤ (1 − f 2). Then, by Proposition 2.10(3), τ(cid:0)(1 − f 2)1/2p(1 − f 2)1/2(cid:1) ≤ τ (1 − f 2). Therefore, using [p, pf 2p] = 0 at the second step, the trace property (Definition 2.8(1)) at the third step, (4.11) at the fourth step, and (4.7) at the last step, ε 4 . . 1 f 2 1 1 (4.12) card(G) For (4.9), first we estimate fhαg(fh)ug −Xg∈G 0 ≤ τ (p) − τ (pf 2p) = τ(cid:0)p(1 − f 2)p(cid:1) = τ(cid:0)(1 − f 2)1/2p(1 − f 2)1/2(cid:1) ≤ τ (1 − f 2) < fhfghug(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) h(cid:13)(cid:13)(cid:13)(cid:13) < card(G)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xg∈G h(cid:13)(cid:13)(cid:13)(cid:13) = card(G)Xg∈G f 2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≤Xh∈G(cid:13)(cid:13)(cid:13)(cid:13)fhpfh − fhpfh! − τ(cid:18) (cid:13)(cid:13)(cid:13)(cid:13)fhpfh − (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xh∈G Now use Proposition 2.10(4) and N (τ ) = 1 to get kfhk · kαg(fh) − fghk < Therefore, using (4.12) at the last step, fhpfh − < ε 4 . card(G) card(G) card(G) ε 32 < ε 4 . 1 f 2 ≤ 1 1 ε 32 card(G) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) τ Xh∈G f 2(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) RADIUS OF COMPARISON OF THE CROSSED PRODUCT 25 step, hp) for h ∈ G. This completes the proof of (4.9). commute with each other. The trace property (Definition 2.8(1)) gives τ (fhpfh) = τ (pf 2 We have τ(cid:0)Ph∈G fhpfh(cid:1) =Ph∈G τ (fhpfh) since the elements fhpfh, for h ∈ G, To prove (4.10), set b =Pg∈G αg(f 2 h) − b(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≤Xg∈G(cid:13)(cid:13)αg(f 2 < 2 card(G)(cid:18) gh(cid:13)(cid:13) +Xg∈G(cid:13)(cid:13)f 2 h) − f 2 16 card(G)(cid:19) = ε 1 ). Then, for h ∈ G, using (4.6) at the second g − αg(f 2 ε 8 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xg∈G 1 )(cid:13)(cid:13) αg(f 2 (4.13) . Next, using Lemma 4.3(2) at the first step and (4.13) at the last step, (4.14) hp − (cid:13)(cid:13)(cid:13)(cid:13)pf 2 card(G) bp(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) . 1 card(G) ε 8 card(G) 1 1 bp(cid:13)(cid:13)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) card(G)Xg∈G card(G)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xg∈G ≤ kpk hp − bp(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≤Xh∈G(cid:13)(cid:13)(cid:13)(cid:13)pf 2 hp − αg(f 2 αg(f 2 < h)!p − h) − b(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) bp(cid:13)(cid:13)(cid:13)(cid:13) < 1 card(G) ε 8 . Then, using (4.14) at the last step, (4.15) pf 2 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xh∈G Finally, we get, using pbp = bp (by Lemma 4.3(3), since b ∈ Aα), N (τ ) = 1, and Proposition 2.10(4) at the second step, and using (4.14) and (4.15) at the third step, (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) pf 2 pf 2 τ (pf 2 hp)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) τ Xh∈G hp! −Xh∈G ≤(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) hp! − τ (bp)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) +(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)Xh∈G τ Xh∈G hp − bp(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xh∈G +Xh∈G(cid:13)(cid:13)(cid:13) 8 card(G)(cid:19) = + card(G)(cid:18) pf 2 ε 8 ε 4 < 1 ε . 1 card(G) card(G) bp − pf 2 τ (pf 2 hp)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) τ (bp) −Xh∈G hp(cid:13)(cid:13)(cid:13) This completes the proof of the claim. 26 M. ALI ASADI-VASFI, NASSER GOLESTANI, AND N. CHRISTOPHER PHILLIPS Now we estimate, using (4.7), (4.9), (4.10), and (4.8) at the second step, (cid:12)(cid:12)(cid:12) 1 1 card(G) − τ (p)(cid:12)(cid:12)(cid:12) ≤(cid:12)(cid:12)(cid:12)(cid:16) card(G)(cid:17)τ (1) −(cid:16) +(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)Xh∈G τ (pf 2 4 card(G) ε 4 < + + ε 1 1 card(G)if 2(cid:17) −Xh∈G card(G)(cid:17)τ (f 2)(cid:12)(cid:12)(cid:12) +(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) τ(cid:16)h hp!(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) hp) − τ Xh∈G +(cid:12)(cid:12)τ (pf 2p) − τ (p)(cid:12)(cid:12) ε 4 ≤ ε. pf 2 ε 4 + This completes the proof. τ (pf 2 hp)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:3) Theorem 4.5. Let G be a finite group, let A be an infinite-dimensional stably finite simple unital C*-algebra, and let α : G → Aut(A) be an action of G on A which has the weak tracial Rokhlin property. Then 1 card(G) · rc(Aα) and rc(cid:0)C∗(G, A, α)(cid:1) ≤ rc(cid:0)C∗(G, A, α)(cid:1) = card(G) · rc(A). Proof. By Lemma 4.4 and Lemma 4.3(4), the projection p ∈ C∗(G, A, α) of Lemma 4.3 satisfies τ (p) = card(G)−1 for all τ ∈ QT(C∗(G, A, α)) and Aα ∼= pC∗(G, A, α)p. The algebra C∗(G, A, α) is simple by Corollary 3.3 of [14]. So p is full. Now C∗(G, A, α) is stably finite (being stably isomorphic to Aα ⊆ A), so Theorem 2.18 implies that rc(cid:0)C∗(G, A, α)(cid:1) = card(G)−1rc(Aα). This is the first part of the (cid:3) conclusion. The second part now follows from Theorem 4.1. 1 We get the same outcome with the Rokhlin property and for any stably finite unital C*-algebra, not necessarily simple. Theorem 4.6. Let G be a finite group, let A be a stably finite unital C*-algebra, and let α : G → Aut(A) be an action of G on A which has the Rokhlin property. Then 1 rc(cid:0)C∗(G, A, α)(cid:1) = card(G) · rc(Aα) and rc(cid:0)C∗(G, A, α)(cid:1) ≤ card(G) · rc(A). Proof. The proof is the same as that of Theorem 4.5, except that we now use Lemma 4.3(6) rather than simplicity of C∗(G, A, α) to deduce that p is full, and we (cid:3) use Theorem 4.2 instead of Theorem 4.1 at the end. 1 If G = Z/2Z and α is the trivial action, then the conclusions of Theorem 4.1 and Theorem 4.2 hold (because Aα = A) but the conclusions of Theorem 4.5 and Theorem 4.6 generally fail (because C∗(G, A, α) ∼= A ⊕ A and rc(A ⊕ A) = rc(A)). For pointwise outer actions α, in fact the conclusions of all these theorems can fail. See Example 6.22. Example 4.7. We give an example of a stably finite simple separable unital C*- algebra D which is not of type I and an action α : Z/2Z → Aut(D) such that α has the weak tracial Rokhlin property but such that the map W(ι) : W(Dα) → W(D) of Lemma 3.11 is not injective. This example also shows that Lemma 3.7 fails when 0 is not a limit point of sp(b). Our algebra D is in fact a UHF algebra, and α actually has the tracial Rokhlin property. This example is therefore a counterexample to RADIUS OF COMPARISON OF THE CROSSED PRODUCT 27 K0(cid:0)C∗(Z/2Z, D, α)(cid:1) such that β∗(η0) 6= η0. Let D and α be as in Example 2.8 of [37]. Let β ∈ Aut(cid:0)C∗(Z/2Z, D, α)(cid:1) be the Proposition 6.2 and Corollary 6.3 of [33]. (The mistake in [33] is in the use of gδ(b) in the proof of Proposition 6.2 of [33]. Since gδ(0) 6= 0, gδ(b) 6∈ bP b.) automorphism which generates the dual action. As shown there, α has the tracial Rokhlin property but not the Rokhlin property. The algebra C∗(Z/2Z, D, α) has a unique tracial state, which we call τ . It is clearly β-invariant. The algebra D also has a unique tracial state σ; necessarily σ = τD. Moreover, there is η0 ∈ Set η = η0 − β∗(η0). Then η 6= 0, but, since τ ◦ β = τ , we have τ∗(η) = 0. It follows from Lemma 4.3(4) that Dα is isomorphic to a full corner of C∗(Z/2Z, D, α). Thus, except for the K0-class of the identity element, the Elliott invariants of Dα and C∗(Z/2Z, D, α) are isomorphic. In particular, Dα has a unique tracial state ρ (necessarily equal to σDα ), and there is µ ∈ K0(Dα) \ {0} such that ρ∗(µ) = 0. Choose projections p, q ∈ K ⊗ Dα such that µ = [p] − [q] in K0(Dα). Since [p] 6= [q] and Dα is stably finite, it follows that hpi 6= hqi in Cu(Dα). In fact, they are in W (Dα). Let ι : Dα → D be the inclusion map. Then σ(ι(p)) = σ(ι(q)). Since D is a UHF algebra, this implies that ι∗([p]) = ι∗([q]) in K0(D). Therefore W(ι)(hpi) = W(ι)(hqi). Thus W(ι) is not injective. Also, p 6-Dα q but p -D q. 5. Surjectivity of Cu+(Aα) → Cu+(A)α In this section, we prove that if G is finite, A is unital, stably finite, and simple, and α : G → Aut(A) has the weak tracial Rokhlin property, then the inclusion Aα → A induces an isomorphism of the ordered semigroups of purely positive elements Cu+(Aα)∪{0} → Cu+(A)α ∪{0}. If we assume stable rank one, then the conclusion is valid for W+(Aα) and W+(A)α as well. We also give the corresponding result for W(Aα) when A is merely unital but α is assumed to have the Rokhlin property. In this case, we need not discard the classes of the projections, just like in Theorem 4.1(ii) of [16] for Cu(Aα). Injectivity was proved in Section 3; the content of this section is the proof of surjectivity. The next lemma produces the following chain of subequivalences, for any g ∈ G: (a − ε)+ -A (a′ − δ6)+ -A (a′ − δ5)+ -A (αg(a′) − δ4)+ -A (αg(a′) − δ2)+ -A (a′ − δ1)+ -A a′ -A (a − δ)+. Lemma 5.1. Let A be a C*-algebra, and let α : G → Aut(A) be an action of a finite group G on A. Let a ∈ (K⊗A)+ satisfy a ∼A αg(a) for all g ∈ G and kak ≤ 1. Then for every ε > 0 there are m ∈ Z>0, δ, δ1, δ2, . . . , δ6 > 0, and a′ ∈ Mm(A)+ with ka′k ≤ kak, such that: (1) 0 < δ < δ1 < δ2 < δ3 < δ4 < δ5 < δ6 < ε. (2) a′ -A (a − δ)+. (3) (αg(a′) − δ2)+ -A (a′ − δ1)+ for all g ∈ G. (4) (a′ − δ5)+ -A (αg(a′) − δ4)+ for all g ∈ G. (5) (a − ε)+ -A (a′ − δ6)+. (6) sp(a′) ∩ (0, δ) 6= ∅. Proof. We may clearly assume that a 6= 0. If, in addition, 0 is a limit point of sp(a), then we may also require: 28 M. ALI ASADI-VASFI, NASSER GOLESTANI, AND N. CHRISTOPHER PHILLIPS 4 ε 3β1 s∗ 3β2 t∗ ε β1 1 2 and 1 2 and and and (5.2) , 2β−1/2 such that β2 < β1 γ = min(cid:18) β1 4 and for all g ∈ G we have(cid:0)αg(a) − β1 12M 2 , Then, for g ∈ G, by Lemma 2.7 there are sg, tg ∈ K ⊗ A such that (5.1) Let ε > 0. First use αg(a) ∼A a for g ∈ G to choose β1 > 0 such that β1 < ε and for all g ∈ G we have (cid:0)a − ε 4(cid:1)+ -A (αg(a) − β1)+. Similarly, choose β2 > 0 4(cid:1)+ -A (a − β2)+. Set 12M 2(cid:19) . (cid:1) M = max(cid:0)1, 2β−1/2 (cid:13)(cid:13)(cid:13)sg(cid:16)αg(a) − 4 (cid:17)+ g −(cid:16)a − (cid:13)(cid:13)(cid:13)tg(cid:16)a − g −(cid:16)αg(a) − 4 (cid:17)+ (cid:13)(cid:13)(cid:13)(cid:16)b − 4 (cid:17)+(cid:13)(cid:13)(cid:13) < γ, 4 (cid:17)+(cid:13)(cid:13)(cid:13) < γ, (5.4) (cid:13)(cid:13)(cid:13)(cid:16)b − ν = min(cid:18) γ Use Lemma 12.4.5 of [17] to choose µ > 0 so small that whenever B is a C*-algebra and b, c ∈ B satisfy 0 ≤ b, c ≤ 1 and kb − ck < µ, then 3β2 (5.3) 4(cid:17)+(cid:13)(cid:13)(cid:13) < γ 4(cid:17)+(cid:13)(cid:13)(cid:13) < γ (cid:13)(cid:13)(cid:13)(cid:16)b − and (cid:13)(cid:13)(cid:13)(cid:16)b − 12(cid:19) . 4 (cid:17)+ −(cid:16)c − 4(cid:17)+ −(cid:16)c − 4 (cid:17)+(cid:13)(cid:13)(cid:13) < γ, 4(cid:17)+(cid:13)(cid:13)(cid:13) < γ. 4(cid:17)+ −(cid:16)c − 4 (cid:17)+ −(cid:16)c − ksgk ≤ 2β−1/2 ktgk ≤ 2β−1/2 If we do not need Part (6) of the conclusion, simply take δ = ν. Otherwise, since 0 is a limit point of sp(a), we can choose δ > 0 such that δ ≤ ν and 5δ 2 ∈ sp(a). Choose m ∈ Z>0 such that there is b ∈ Mm(A)+ with kbk = kak and kb − ak < δ 2 . Define a′ = (b − 2δ)+. Since kb − ak < δ, it follows from Lemma 2.4(3c) that a′ -A (a− δ)+, which is (2). If 0 is a limit point of sp(a), then we arranged to have 2 ∈ sp(a), so kb − ak < δ 2 implies sp(b) ∩ (2δ, 3δ) 6= ∅. Thus sp(a′) ∩ (0, δ) 6= ∅. This is (6). We have ka′ − ak < 3δ ≤ µ. For g ∈ G, it therefore follows from (5.1), using Define β2 4 µ 3 , ε , 3β1 3β1 3β2 ε ε β1 β1 , 3 5δ . ε s∗ 3β1 3β1 (5.4) and ksgk ≤ M , that 3β1 (5.5) (cid:13)(cid:13)(cid:13)sg(cid:16)αg(a′) − 4 (cid:17)+ ≤ ksgk2(cid:13)(cid:13)(cid:13)(cid:16)a′ − +(cid:13)(cid:13)(cid:13)sg(cid:16)αg(a) − (cid:13)(cid:13)(cid:13)tg(cid:16)a′ − 4(cid:17)+(cid:13)(cid:13)(cid:13) g −(cid:16)a′ − 4 (cid:17)+(cid:13)(cid:13)(cid:13) 4 (cid:17)+ −(cid:16)a − 4(cid:17)+(cid:13)(cid:13)(cid:13) +(cid:13)(cid:13)(cid:13)(cid:16)a − 4 (cid:17)+ g −(cid:16)a − s∗ < M 2γ + γ + γ ≤ 3M 2γ, 4 (cid:17)+(cid:13)(cid:13)(cid:13) ≤ 3M 2γ. 4 (cid:17)+ g −(cid:16)αg(a′) − and similarly, using (5.2), (5.3), and ktgk ≤ M , (5.6) Define 3β2 3β1 β1 t∗ ε ε 4(cid:17)+ −(cid:16)a′ − δ1 = , δ2 = , δ3 = , δ4 = , δ5 = , and δ6 = β1 2 5β1 8 3β1 4 ε 2 3β2 4 ε 4(cid:17)+(cid:13)(cid:13)(cid:13) 3ε 4 . RADIUS OF COMPARISON OF THE CROSSED PRODUCT 29 Then (1) is clear. Moreover, we have ε 2 ε 4 and = δ5 which is (4), and So (5.5) and (5.6) imply, for g ∈ G, + 3M 2γ ≤ (a′ − δ5)+ ≤(cid:16)a′ −(cid:16) ε (αg(a′) − δ2)+ ≤(cid:16)αg(a′) −(cid:16) β1 which is (3). Finally, ka′ − ak < 3δ ≤ ε (cid:0)a′ − 3ε 4 4 β1 4 + 3M 2γ ≤ β1 2 = δ2. + 3M 2γ(cid:17)(cid:17)+ -A (αg(a′) − δ4)+, + 3M 2γ(cid:17)(cid:17)+ -A (a′ − δ1)+, 4(cid:1)+ = (a′ − δ6)+, which is (5). This completes the proof. 4 , so by Lemma 2.4(3c) we have (a− ε)+ -A (cid:3) Lemma 5.2. Let A be a simple C*-algebra which is not of type I. Let α : G → Aut(A) be an action of a finite group G on A and let x ∈ A+ \ {0}. Then there exists z ∈ (Aα)+ \ {0} such that z -A x. Proof. Set n = card(G). By Lemma 2.4 of [38], there are b1, b2, . . . , bn ∈ A+ \ {0} such that, for j 6= k, bjbk = 0, b1 ∼A b2 ∼A ··· ∼A bn, and b1 + b2 + ··· + bn -A x. Let y be the direct sum of n copies of b1. Using Lemma 2.6 of [38], choose c ∈ A+ \ {0} such that c -A αg−1 (b1) for all g ∈ G. Then αg(c) -A b1 for all g ∈ G. Set z =Pg∈G αg(c). Clearly z ∈ (Aα)+ \ {0}. Then αg(c) -A y -A b1 + b2 + ··· + bn -A x, z =Xg αg(c) -AMg∈G as desired. (cid:3) Lemma 5.3. Let A be an infinite-dimensional simple unital C*-algebra and let α : G → Aut(A) be an action of a finite group G on A which has the weak tracial Rokhlin property. Let a ∈ (K ⊗ A)+ satisfy a ∼A αg(a) for all g ∈ G and assume that 0 is a limit point of sp(a). Then for every ε > 0 there exist δ > 0, m ∈ Z>0, and b ∈ Mm(Aα)+ such that (a − ε)+ -A b -A (a − δ)+ and Proof. We may assume kak = 1. Set n = card(G). sp(a), and let the notation be as in its conclusion. [0, 1] ⊆ sp(b). Let ε > 0. Apply the version of Lemma 5.1 which assumes 0 is a limit point of By Lemma 5.1(6), there is λ ∈ sp(a′) ∩ (0, δ). Choose a continuous function h : [0,∞) → [0, 1] such that supp(h) ⊆ (0, δ) and h(λ) = 1. Use Lemma 5.2 to choose d ∈ Mm(Aα)+ \{0} such that d -A h(a′). Since the action induced by α on Mm(A) has the weak tracial Rokhlin property (Corollary 4.6 of [14]), Lemma 3.10 provides s ∈ Mm(Aα)+ such that s -Aα d and sp(s) = [0, 1]. Define (5.7) (5.8) ρ = min(cid:0)1, δ1 − δ, δ3 − δ2, δ4 − δ3, δ6 − δ5(cid:1), µ = ρ2 6n2 , and ε′ = ρ3 36n4 . 30 M. ALI ASADI-VASFI, NASSER GOLESTANI, AND N. CHRISTOPHER PHILLIPS Set and and and (5.10) g , wg, w∗ kvgk ≤ ρ−1/2 For g ∈ G use Lemma 2.7, Lemma 5.1(3), Lemma 5.1(4), and (5.7) to choose vg, wg ∈ Mm(A) such that (5.9) (cid:13)(cid:13)vg(a′ − δ)+v∗ (cid:13)(cid:13)wg(αg(a′) − δ3)+w∗ F =(cid:8)vg, v∗ g − (αg(a′) − δ2)+(cid:13)(cid:13) < ε′ g − (a′ − δ5)+(cid:13)(cid:13) < ε′ g : g ∈ G(cid:9) ∪(cid:8)(a′ − δ)+, (a′ − δ5)+(cid:9). such that, with e =Pg∈G eg and f =Pg∈G fg, the following hold: Since the induced action on Mm(A) has the weak tracial Rokhlin property, we can apply Lemma 3.3 with Mm(A) in place of A, with F as above, with ε′ in place of ε, and with s in place of x. We get positive contractions eg, fg ∈ Mm(A) for g ∈ G kwgk ≤ ρ−1/2. (1) kegehk < ε′ and kfgfhk < ε′ for all g, h ∈ G. (2) kzeg − egzk < ε′ and kzfg − fgzk < ε′ for all g ∈ G and all z ∈ F . (3) (1 − f − ε′)+ -A s. (4) αg(eh) = egh and αg(fh) = fgh for all g, h ∈ G. (5) e, f ∈ Aα and kfk = 1. (6) egfg = fg for all g ∈ G. Define egvg x =Xg∈G kvgk ≤ nρ−1/2 kxk ≤Xg∈G fg(a′ − δ5)+fg, fgwg. and y =Xg∈G kyk ≤Xg∈G eg(αg(a′) − δ2)+eg, and kwgk ≤ nρ−1/2. and c = (c0 − µ)+. Further define a0 =Xg∈G Then kc0k ≤ n and kck ≤ n. Also, c0, c ∈ (Aα)+ by (4). c0 =Xg∈G We claim that Then (5.11) (5.12) and (5.13) (cid:13)(cid:13)x(a′ − δ)+x∗ − c0(cid:13)(cid:13) < µ (cid:13)(cid:13)ycy∗ − f (a′ − δ5)+f(cid:13)(cid:13) < ρ 3 . We prove (5.12). First, if g 6= h then, using (2), the second part of (5.9), and (1) at the second step, and (5.7) at the last step, kegvg(a′ − δ)+v∗ hehk ≤ kegvg − vgegk · k(a′ − δ)+k · kv∗ hk · kehk + kvgk · keg(a′ − δ)+ − (a′ − δ)+egk · kv∗ + kvgk · k(a′ − δ)+k · kegv∗ hegk · kehk + kvgk · k(a′ − δ)+k · kv∗ h − v∗ hk · kegehk hk · kehk < ε′ρ−1/2 + ε′ρ−1 + ε′ρ−1/2 + ε′ρ−1 ≤ 4ε′ρ−1. RADIUS OF COMPARISON OF THE CROSSED PRODUCT 31 Therefore, using the first part of (5.9) and this estimate at the second step, (cid:13)(cid:13)x(a′ − δ)+x∗ − c0(cid:13)(cid:13) ≤Xg∈G g − (αg(a′) − δ2)+(cid:13)(cid:13) · kegk kegk ·(cid:13)(cid:13)vg(a′ − δ)+v∗ +Xg6=h kegvg(a′ − δ)+v∗ hehk < nε′ + 4n2ε′ρ−1 ≤ 5n2ε′ρ−1 < µ. This is (5.12). Now we prove (5.13). First, by (5.11) and (5.8), (5.14) kycy∗ − yc0y∗k ≤ n2ρ−1µ = ρ 6 . Next, for g, h, k ∈ G we have, using (2) and the second part of (5.10) at the second step, (cid:13)(cid:13)fgwgek(αk(a′) − δ2)+ekw∗ h − w∗ < ε′ρ−1/2 + ε′ρ−1/2 ≤ hfh − fgekwg(αk(a′) − δ2)+w∗ hekfh(cid:13)(cid:13) hfh(cid:13)(cid:13) ≤ kfgk · kwgek − ekwgk ·(cid:13)(cid:13)(αk(a′) − δ2)+ekw∗ +(cid:13)(cid:13)fgekwg(αk(a′) − δ2)+(cid:13)(cid:13) · kekw∗ hekfh(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) fgekwg(αk(a′) − δ2)+w∗ < 2n3ε′ρ−1/2 ≤ hekk · kfhk ρ 18 . (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) yc0y∗ − Xg,h,k∈G ρ 18 . Therefore, by (5.8), (5.15) Set S = {(g, g, g) : g ∈ G} ⊆ G3. If g, k ∈ G are distinct, then kfgekk = kfgfkekk < ε′ by (6) and (1). Similarly, if h, k ∈ G are distinct, then kekfhk < ε′. In both cases, by (5.10), Meanwhile, by (6) and the first part of (5.10), < n3ε′ρ−1 ≤ ρ 36 . So, by (5.8), (5.16) (5.17) fgekwg(αk(a′) − δ2)+w∗ hekfh(cid:13)(cid:13) < ε′ρ−1. (cid:13)(cid:13)fgekwg(αk(a′) − δ2)+w∗ hekfh(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) g egfg − a0(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≤Xg∈G(cid:13)(cid:13)fgwg(αg(a′) − δ3)+w∗ (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xg,h,k∈G\S (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xg∈G fgegwg(αg(a′) − δ3)+w∗ g fg − fg(a′ − δ5)+fg(cid:13)(cid:13) < nε′ ≤ ρ 36 . 32 M. ALI ASADI-VASFI, NASSER GOLESTANI, AND N. CHRISTOPHER PHILLIPS Finally, using (2) and (1), and using (5.8) at the last step, (5.18) ka0 − f (a′ − δ5)+fk kfg(a′ − δ5)+fhk ≤Xg6=h ≤Xg6=h(cid:2)kfg(a′ − δ5)+ − (a′ − δ5)+fgk · kfhk + k(a′ − δ5)+k · kfgfhk(cid:3) < n2(ε′ + ε′) ≤ ρ 18 . Combining (5.14), (5.15), (5.16), (5.17), and (5.18), we get (cid:13)(cid:13)ycy∗ − f (a′ − δ5)+f(cid:13)(cid:13) < ρ 3 , which is (5.13). The claim is proved. Define b = c ⊕ s, which is in M2m(Aα)+. From (5.12) we get c = (c0 − µ)+ -A x(a′ − δ)+x∗ -A (a′ − δ)+. Using s -A d -A h(a′) and h(a′) ⊥ (a′ − δ)+, as well as Lemma 5.1(2), we have b -A (a′ − δ)+ ⊕ h(a′) -A a′ -A (a − δ)+. Using Lemma 5.1(5) at the first step, (5.7) at the second step, Lemma 2.6 at the third step, (5.13) and ε′ < ρ 3 (by (5.8)) at the fourth step, and (3) at the fifth step, we get (a − ε)+ -A (a′ − δ6)+ ≤ (a′ − δ5 − ρ)+ 3(cid:17)+ ⊕(cid:16)1 − f − 3(cid:17)+ -A(cid:16)f (a′ − δ5)+f − -A ycy∗ ⊕ (1 − f − ε′)+ -A c ⊕ s = b. ρ ρ The last two relations complete the proof. (cid:3) Lemma 5.4. Let A be an infinite-dimensional stably finite simple unital C*-algebra and let α : G → Aut(A) be an action of a finite group G on A which has the weak tracial Rokhlin property. Recalling the notation of Definition 3.8, let a ∈ (K⊗A)++ satisfy a ∼A αg(a) for all g ∈ G. Then there exists b ∈ (K ⊗ Aα)++ such that: (1) hbiA = haiA. (2) There are η0, η1, . . . ∈ W+(Aα) such that η0 ≤ η1 ≤ ··· and hbiAα = supn∈Z≥0 ηn. Proof. By induction on n, we construct sequences (εn)n∈Z≥0 in (0,∞), (bn)n∈Z≥0 in (K ⊗ Aα)+, and (m(n))n∈Z≥0 in Z>0, such that limn→∞ εn = 0, b0 -A (a− ε0)+, and for all n ∈ Z≥0 we have εn+1 < εn, and bn ∈(cid:0)Mm(n)(A)α(cid:1)+, (a − εn)+ -A bn+1 -A (a − εn+1)+, [0, 1] ⊆ sp(bn). To begin, set ε0 = 1. Given εn with n ∈ Z≥0, apply Lemma 5.3 with εn in place of ε, getting δ > 0, m(n + 1) ∈ Z>0, and bn+1 ∈ Mm(n+1)(Aα)+ such that (5.19) Then set εn+1 = min(cid:0)δ, εn 2(cid:1). The induction is complete. (a − εn)+ -A bn+1 -A (a − δ)+ [0, 1] ⊆ sp(bn). and RADIUS OF COMPARISON OF THE CROSSED PRODUCT 33 We now have b0 -A b1 -A b2 -A ··· . Since idMl⊗α has the weak tracial Rokhlin property for all l ∈ Z>0 (by Corollary 4.6 of [14]), it follows from Lemma 3.7 that (5.20) By Theorem 4.19 of [3], there exists b ∈ (K ⊗ Aα)+ such that hbiAα = supn hbniAα. Therefore, using Theorem 1.16 of [38] at the third step, b0 -Aα b1 -Aα b2 -Aα ··· . (5.21) hbiA = Cu(ι)(cid:0)hbiAα(cid:1) = Cu(ι)(cid:16) sup n hbniAα(cid:17) = sup n (cid:0)Cu(ι)hbniAα(cid:1) = sup n hbniA. Moreover, for all n ∈ Z≥0, we have (a− εn)+ -A bn+1 -A a. Since limn→∞ εn = 0, it follows from Lemma 1.25(1) of [38] that supn hbniA = haiA. So hbiA = haiA, which is Part (1) of the conclusion. Part (2) follows by taking ηn = hbniAα for n ∈ Z≥0. Finally, we prove that b ∈ (K ⊗ Aα)++. If not, then (see Definition 3.8) there is a projection p ∈ K ⊗ Aα such that hbiAα = hpiAα . But then haiA = hpiA, (cid:3) contradicting a ∈ (K ⊗ A)++. Recall the definition of Cu+(A) (Definition 3.8). Theorem 5.5. Let A be an infinite-dimensional stably finite simple unital C*- algebra and let α : G → Aut(A) be an action of a finite group G on A which has the weak tracial Rokhlin property. Then the inclusion map ι : Aα → A induces an isomorphism of ordered semigroups Cu+(ι) : Cu+(Aα) ∪ {0} → Cu+(A)α ∪ {0}. By Theorem 4.1(ii) of [16], if α has the Rokhlin property, this holds for arbi- trary A and without restricting to the classes of purely positive elements. Proof of Theorem 5.5. It follows from Lemma 3.11(2), Lemma 3.9(2), and simplic- ity of Aα that the map Cu+(ι) : Cu+(Aα) → Cu(A) is injective and is an order isomorphism onto its range. It is trivial that the range is contained in Cu(A)α, it follows from Lemma 3.9(2) that the range is contained in Cu+(A), and it follows from Lemma 5.4 that the range contains Cu+(A)α. So the range is Cu+(A)α. The extension to Cu+(ι) : Cu+(Aα) ∪ {0} → Cu+(A)α ∪ {0} is immediate. (cid:3) Corollary 5.6. Let A be an infinite-dimensional simple unital C*-algebra. Let α : G → Aut(A) be an action of a finite group G on A which has the weak tracial Rokhlin property. Assume that Aα has stable rank one. Then the inclusion map ι : Aα → A induces an isomorphism of ordered semigroups W+(ι) : W+(Aα)∪{0} → W+(A)α ∪ {0}. It is presumably true that if A is an infinite-dimensional stably finite simple unital C*-algebra with stable rank one, G is a finite group, and α : G → Aut(A) has the weak tracial Rokhlin property, then C∗(G, A, α) and Aα have stable rank one. However, this has not been proved, and a proof presumably requires methods like those in [4]. It is known that if α has the tracial Rokhlin property, then C∗(G, A, α) has stable rank one. This is claimed in Theorem 3.1 of [13]. We could not follow the proof there, but a proof will appear in [18]. In this case, Aα has stable rank one by Lemma 4.3(5). We need the following fact. It is part of Theorem 5.15 of [3], except that we omit the separability hypothesis there. That hypothesis isn't actually needed for 34 M. ALI ASADI-VASFI, NASSER GOLESTANI, AND N. CHRISTOPHER PHILLIPS the proof given there. (The statement in [3] omits "nondecreasing", but, as one sees from the proof, this hypothesis is intended.) Proposition 5.7. Let A be a unital C*-algebra with stable rank one. Let (ηn)n∈Z≥0 be a bounded nondecreasing sequence in W(A). Let η = supn∈Z≥0 ηn, evaluated in Cu(A). Then η ∈ W(A). Proof. If A is separable, this is contained in Theorem 5.15 of [3]. The only use of separability in the proof of that theorem is in the use of Lemma 5.13 of [3]. One needs to know that the algebra A∞ in that proof has a strictly positive element. It is enough to show that A∞ has a countable approximate identity, which follows from the fact that, using the notation there, A∞ is the countable increasing union of subalgebras Aan = anAan, each of which clearly has a countable approximate (cid:3) identity. Proof of Corollary 5.6. Since C∗(G, A, α) is simple, Theorem 2.8 of [7] and Lemma 4.3(4) imply that Aα is stably isomorphic to C∗(G, A, α). The algebra Aα is sta- bly finite since it has stable rank one, so C∗(G, A, α) is stably finite, and there- It now follows from Theorem 5.5 that fore its subalgebra A is stably finite. W+(ι) : W+(Aα)∪{0} → W+(A)∪{0} is an order isomorphism from W+(Aα)∪{0} to some subsemigroup of Cu+(A)∪{0} which is contained in(cid:0)W+(A)∩Cu+(A)α(cid:1)∪ {0} = W+(A)α ∪ {0}. Now let η ∈ (cid:0)W+(A) ∩ Cu+(A)α(cid:1) ∪ {0}; we show that η is in the range of W+(ι). This is trivial if η = 0. Otherwise, choose m ∈ Z>0 and a ∈ Mm(A)+ such that haiA = η. Apply Lemma 5.4 to a, getting b ∈ (K ⊗ Aα)++ and a nondecreasing sequence (ηn)n∈Z≥0 in W+(Aα) such that hbiAα = supn∈Z≥0 ηn. This sequence is bounded by h1Mm(Aα)iAα. So hbiAα ∈ W(Aα) by Proposition 5.7, and hbiAα ∈ Cu+(Aα) by Lemma 3.9(4). The conclusion follows. (cid:3) Corollary 5.8. Let A be an infinite-dimensional stably finite simple unital C*- algebra and let α : G → Aut(A) be an action of a finite group G on A which has the weak tracial Rokhlin property. Then as ordered semigroups. If Aα has stable rank one, then Cu+(cid:0)C∗(G, A, α)(cid:1) ∪ {0} ∼= Cu+(A)α ∪ {0} W+(cid:0)C∗(G, A, α)(cid:1) ∪ {0} ∼= W+(A)α ∪ {0} as ordered semigroups. Proof. It suffices to prove that Cu+(cid:0)C∗(G, A, α)(cid:1) ∼= Cu+(A)α and, in the stable rank one case, that W+(cid:0)C∗(G, A, α)(cid:1) ∼= W+(A)α. Lemma 4.3(4) and simplicity of C∗(G, A, α) (Corollary 3.3 of [14]) imply that Aα is isomorphic to a full corner of C∗(G, A, α). Since Aα and C∗(G, A, α) are both unital, it is easy to check that there is n ∈ Z>0 such that C∗(G, A, α) is isomorphic to a full corner of Mn(Aα). Therefore M∞(C∗(G, A, α)) ∼= M∞(Aα). In particular, K ⊗ Aα ∼= K ⊗ C∗(G, A, α). Using Theorem 5.5 at the second step, we get When Aα has stable rank one, the isomorphism W+(cid:0)C∗(G, A, α)(cid:1) ∼= W+(A)α Cu+(cid:0)C∗(G, A, α)(cid:1) ∼= Cu+(Aα) ∼= Cu+(A)α. follows similarly, using Corollary 5.6 and M∞(C∗(G, A, α)) ∼= M∞(Aα). (cid:3) RADIUS OF COMPARISON OF THE CROSSED PRODUCT 35 There is an analog of Corollary 5.6 for Rokhlin actions on unital C*-algebras, whose proof uses Theorem 4.1(ii) of [16] instead of our Theorem 5.5. Proposition 5.9. Let A be a unital C*-algebra with stable rank one. Let α : G → Aut(A) be an action of a finite group G on A which has the Rokhlin property. Then the inclusion map ι : Aα → A induces an isomorphism of ordered semigroups W(ι) : W(Aα) → W(A)α. We need Proposition 4.1(1) of [32], but without the separability hypothesis there. We give an easy proof directly from Theorem 3.2 of [32]. Proposition 5.10. Let A be a unital C*-algebra with stable rank one. Let α : G → Aut(A) be an action of a finite group G on A which has the Rokhlin property. Then C∗(G, A, α) has stable rank one. Proof. Let a ∈ C∗(G, A, α) and let ε > 0. Use Theorem 3.2 of [32] to choose a projection f ∈ A, an integer n ∈ Z>0, a unital homomorphism ϕ : Mn(f Af ) → C∗(G, A, α), and an element b ∈ Mn(f Af ) such that kϕ(b) − ak < ε 2 . Combining Theorem 3.1.8 and Theorem 3.1.9(1) of [25], we see that Mn(f Af ) has stable rank one. Choose c ∈ Mn(f Af ) such that c is invertible and kc − bk < ε 2 . Then ϕ(c) is an invertible element of C∗(G, A, α) such that kϕ(c) − ak < ε. (cid:3) Proof of Proposition 5.9. The algebra C∗(G, A, α) has stable rank one by Proposi- tion 5.10. It now follows from Lemma 4.3(5) that Aα has stable rank one. Theorem 4.1(ii) of [16] implies that W(ι) : W(Aα) → W(A) is an order isomor- phism from W(Aα) to some subsemigroup of Cu(A), which is necessarily contained in W(A) ∩ Cu(A)α = W(A)α. Let η ∈ W(A)∩ Cu(A)α; we need to show that η is in the range of W(ι). Choose m ∈ Z>0 and a ∈ Mm(A)+ such that haiA = η. Since η ∈ Cu(A)α, by Theorem 4.1(ii) of [16] there is b ∈ (K⊗Aα)+ such that hbiA = η. The case η = 0 is trivial, so without loss of generality kbk = 1. We now construct, by induction on n, sequences (εn)n∈Z≥0 in (0,∞), (bn)n∈Z≥0 in (K ⊗ Aα)+, and (m(n))n∈Z≥0 in Z>0, such that limn→∞ εn = 0, b0 -A (a − ε0)+, and for all n ∈ Z>0 we have and (b − εn)+ -Aα bn+1 -Aα (b − εn+1)+. εn+1 < εn, To begin, set ε0 = 1 and b0 = 0. Given εn with n ∈ Z≥0, set εn+1 = εn 3 . Choose m(n + 1) ∈ Z>0, and cn+1 ∈ Mm(n+1)(Aα)+ such that kcn+1 − bk < εn+1. Two applications of Lemma 2.4(3c) give bn ∈(cid:0)Mm(n)(A)α(cid:1)+, (b − εn)+ = (b − 3εn+1)+ -Aα (cn+1 − 2εn+1)+ -Aα (b − εn+1)+. Set bn+1 = (cn+1 − 2εn+1)+. The induction is complete. For n ∈ Z>0, set ηn = hbniAα , which is in W(Aα). Then (ηn)n∈Z≥0 is a nonde- creasing sequence in W(Aα) and, by Lemma 1.25(1) of [38], we have supn∈Z≥0 ηn = hbiAα = η. This sequence is bounded by h1Mm(Aα)iAα, so Proposition 5.7 now implies hbiAα ∈ W(Aα). (cid:3) 6. An example We give an example of a simple AH algebra A with rc(A) > 0 and an action α : Z/2Z → Aut(A) which has the Rokhlin property. As discussed in the introduc- tion, it is not a priori obvious that such examples should exist, even with the weak tracial Rokhlin property in place of the Rokhlin property. In our example, we get 36 M. ALI ASADI-VASFI, NASSER GOLESTANI, AND N. CHRISTOPHER PHILLIPS equality in Theorem 4.1 and Theorem 4.5. See Theorem 6.15 and Theorem 6.21. The algebras A and Aα have stable rank one (Lemma 6.5 and Corollary 6.7), and the maps W(Aα) → W(A)α and Cu(Aα) → Cu(A)α are isomorphisms (Corollary 6.6). The construction is motivated by [22], in which two AH direct systems with simple direct limits are "merged" into a single larger system whose direct limit is still simple but which is "not very far" from the direct sum of the two original direct limits. The "merger" is accomplished by writing the systems side by side, and inserting a very small number of point evaluation maps which go from one of the original systems to the other. In [22], the essential point was that the two systems were very different but that the base spaces were all contractible. Here, we use two copies of the same system. Writing the direct system sideways, our combined system looks like the following diagram, in which the solid arrows represent many partial maps and the dotted arrows represent a small number of point evaluations: C(X1) ⊗ Mr(1) / C(X2) ⊗ Mr(2) / C(X3) ⊗ Mr(3) / ··· C(X1) ⊗ Mr(1) C(X2) ⊗ Mr(2) C(X3) ⊗ Mr(3) / ··· . The order two automorphism exchanges the two rows. Since we don't care about contractibility, we can use products of copies of S2 instead of cones over such spaces as in [22]. We compute the radius of comparison exactly, instead of just giving bounds as is done in [22]. To keep the notation simple, we carry out only the case of Z/2Z and radius of comparison less than 1. Modifications of the construction will presumably work for any finite group and give any value of the radius of comparison in [0,∞]. Construction 6.1. We define the following objects: (1) For n ∈ Z≥0, define • d(n) = 2n+1 − 1. • l(n) = 2n+1. • r(0) = 1 and r(n) =Qn k=1 2k+1. • s(0) = 1 and s(n) =Qn k=1(2k+1 − 1). k=1(cid:0)1 − 1 2k+1(cid:1). • u(n) = s(n) • t(0) = 0 and t(n + 1) = d(n + 1)t(n) + [r(n) − t(n)]. r(n) =Qn ν coordinate projection. ν dimension of Xn is dim(Xn) = 2s(n). (2) Define κ = limn→∞ u(n). (3) For n ∈ Z≥0, define a compact space by Xn = (S2)s(n). Then the covering (4) For n ∈ Z≥0 and ν = 1, 2, . . . , d(n + 1), let P (n) : Xn+1 → Xn be the (5) Choose points xm ∈ Xm for m ∈ Z≥0 such that for all n ∈ Z≥0, the set n(cid:0)P (n) ν1 ◦ P (n+1) m = n, n + 1, . . . and νj = 1, 2, . . . , d(n + j) for j = 1, 2, . . . , m − no νm−n (cid:1)(xm) : is dense in Xn. (The contribution to this set when m = n is xn.) ν2 ◦ ··· ◦ P (m−1) (6) For n ∈ Z≥0, define An =(cid:2)C(Xn) ⊕ C(Xn)(cid:3) ⊗ Mr(n). ( ( / / / / / ( ( / / / / / & & / / / / / 6 6 / / / / / / 6 6 / / / / / / 8 8 / / / / / RADIUS OF COMPARISON OF THE CROSSED PRODUCT 37 by (6.1) When convenient, we identify An in the obvious ways with C(Xn, Mr(n)) ⊕ C(Xn, Mr(n)) and C(cid:0)Xn ∐ Xn, Mr(n)(cid:1). (7) For n ∈ Z≥0, define a unital homomorphism λn : C(Xn) ⊕ C(Xn) → Ml(n+1)(cid:2)C(Xn+1) ⊕ C(Xn+1)(cid:3) λn(f, g) =(cid:16)diag(cid:0)f ◦ P (n) d(n+1), g(xn)(cid:1), d(n+1), f (xn)(cid:1)(cid:17). diag(cid:0)g ◦ P (n) , . . . , f ◦ P (n) , . . . , g ◦ P (n) , f ◦ P (n) , g ◦ P (n) 1 2 1 2 (8) For n ∈ Z≥0, define Λn+1, n : An → An+1 by Λn+1, n = λn ⊗ idMr(n) . Thus, Λn+1,n : [C(Xn) ⊕ C(Xn)] ⊗ Mr(n) → [C(Xn+1) ⊕ C(Xn+1)] ⊗ Mr(n+1) is given by (f, g) ⊗ c 7→ (cid:0)f ◦ P (n) 1  (6.2) 1 (cid:1) , g ◦ P (n) . . . 0 0 (cid:0)f ◦ P (n) d(n+1)(cid:1) d(n+1), g ◦ P (n)  ⊗ c (cid:0)g(xn), f (xn)(cid:1) for f, g ∈ C(Xn) and c ∈ Mr(n). Using standard matrix unit notation, we can also write this definition as = Λn+1,n(cid:0)(f, g) ⊗ c(cid:1) d(n+1)Xj=1 (cid:0)f ◦ P (n) +(cid:0)g(xn) · 1C(Xn+1), f (xn) · 1C(Xn+1)(cid:1) ⊗ ed(n+1)+1, d(n+1)+1 ⊗ c. (cid:1) ⊗ ej,j ⊗ c For m, n ∈ Z≥0 with m ≤ n, now define , g ◦ P (n) j j Λn,m = Λn,n−1 ◦ Λn−1, n−2 ◦ ··· ◦ Λm+1,m : Am → An. (9) Define For n ∈ Z≥0, it is clear that Λn+1, n is an injective unital homomorphism. Let Λ∞,n : An → A be the standard map associated with the direct limit. (10) Write Z/2Z = {0, 1}. For n ∈ Z≥0, define α(n) : Z/2Z → Aut(An) by A = lim −→(cid:0)An, (Λm, n)m≥n(cid:1). α(n) 1 (cid:0)(f, g) ⊗ c(cid:1) = (g, f ) ⊗ c for f, g ∈ C(Xn) and c ∈ Mr(n). We also write α(n) for the generating automorphism α(n) 1 . We then have the following diagram: Λ3, 2−−−−→ A3 −−−−→ ··· Λ2, 1−−−−→ A2 Λ1, 0−−−−→ A1 (6.3) A0 α(0)y A0 α(1)y Λ1, 0−−−−→ A1 α(2)y α(3)y Λ2, 1−−−−→ A2 Λ3, 2−−−−→ A3 −−−−→ ··· . 38 M. ALI ASADI-VASFI, NASSER GOLESTANI, AND N. CHRISTOPHER PHILLIPS Lemma 6.2. Assume the notation and choices in Construction 6.1. Then 0 ≤ t(n) < r(n) for all n ∈ Z≥0. Proof. The statement is true for n = 0 by definition. Let n ∈ Z≥0 and assume 0 ≤ t(n) < r(n). Then t(n + 1) = d(n + 1)t(n) + [r(n) − t(n)] = [d(n + 1) − 1]t(n) + r(n), which implies (using d(n + 1) − 1 = 2n+2 − 2 ≥ 0) 0 ≤ t(n + 1) ≤ [d(n + 1) − 1]r(n) + r(n) < r(n + 1). So 0 ≤ t(n) < r(n) for all n ∈ Z≥0 by induction. Lemma 6.3. Assume the notation and choices in Construction 6.1. Then (u(n))n∈Z≥0 is strictly decreasing and 0 < κ < 1. (cid:3) Proof. The first statement is clear, as is κ < 1. To prove that κ > 0, we first observe that if β1, β2 ∈ [0, 1] then (1− β1)(1− β2) ≥ Induction gives an analogous statement for n factors, so that, in (cid:3) 2k+1 . Letting n → ∞, we get κ ≥ 1 2 . k=1 1 1 − β1 − β2. particular, u(n) ≥ 1 −Pn Lemma 6.4. In Construction 6.1(10), the diagram (6.3) commutes. Moreover, α(n), and this action there is a unique action α : Z/2Z → Aut(A) such that α = lim −→ has the Rokhlin property.. Proof. For the first statement, let n ∈ Z≥0. Using 6.1(7) in the second step and 6.1(10) in the third step, for all f, g ∈ C(Xn) and for all c ∈ Mr(n) we have (cid:0)α(n+1) ◦ Λn+1, n(cid:1)(cid:0)(f, g) ⊗ c(cid:1) = α(n+1)(cid:0)λn((f, g)) ⊗ c(cid:1) 2 1 =(cid:16)diag(cid:0)g ◦ P (n) , g ◦ P (n) diag(cid:0)f ◦ P (n) , f ◦ P (n) =(cid:0)Λn+1, n ◦ α(n)(cid:1)(cid:0)(f, g) ⊗ c(cid:1). 1 , . . . , g ◦ P (n) 2 , . . . , f ◦ P (n) d(n+1), f (xn)(cid:1), d(n+1), g(xn)(cid:1)(cid:17) ⊗ c Existence of α follows immediately. It is immediate that α(n) has the Rokhlin property for all n ∈ Z≥0, and it follows easily that α has the Rokhlin property. (cid:3) Lemma 6.5. Assume the notation and choices in Construction 6.1. Then the C*-algebra A is stably finite and simple, and has stable rank one. Proof. Stable finiteness is immediate. For simplicity, it is easy to check that the hypotheses of Proposition 2.1(ii) of [9] hold. For stable rank one, we observe that the direct system in Construction 6.1(9) has diagonal maps in the sense of Definition 2.1 (cid:3) of [10]. Therefore A has stable rank one by Theorem 4.1 of [10]. Corollary 6.6. Assume the notation and choices in Construction 6.1. Then the maps Cu(Aα) → Cu(A)α and W(Aα) → W(A)α are isomorphisms. Proof. This follows from Theorem 4.1(ii) of [16] and Proposition 5.9, by Lemma 6.5 (cid:3) and Lemma 6.4. Corollary 6.7. Assume the notation and choices in Construction 6.1. Then C∗(Z/2Z, A, α) and Aα have stable rank one. Proof. The result for C∗(Z/2Z, A, α) follows from Lemma 6.5, Lemma 6.4, and Proposition 4.1(1) of [32]. The result for Aα now follows from Lemma 4.3(5). (cid:3) RADIUS OF COMPARISON OF THE CROSSED PRODUCT 39 Notation 6.8. Let p ∈ C(S2, M2) denote the Bott projection, and let L be the tautological line bundle over S2 ∼= CP1. (Thus, the range of p is the section space of L.) Recall that X0 = S2. Assuming the notation and choices in Construction 6.1, for n ∈ Z≥0 set pn = (idM2 ⊗ Λn,0)(p, 0) ∈ M2(An) In particular, p0 = (p, 0) and p′ n = (idM2 ⊗ Λn,0)(p, p) ∈ M2(An). and p′ 0 = (p, p). Lemma 6.9 ([22]). The Cartesian product L×k does not embed in a trivial bundle over (S2)k of rank less than 2k. Proof. This is Lemma 1.9 of [22]. (cid:3) Lemma 6.10. Assume the notation and choices in Construction 6.1, and adopt : (S2)s(n) → S2 be the Notation 6.8. Let n ∈ Z≥0. For j = 1, 2, . . . , s(n) let R(n) j j coordinate projection. Then: (1) There are orthogonal projections c(0) n , c(1) n , gn ∈ M2r(n)(cid:0)C(Xn)(cid:1) such that and n + c(1) n , gn(cid:1) pn =(cid:0)c(0) for j = 1, 2, . . . , s(n), c(1) c(0) n is the direct sum of the projections p ◦ R(n) n is a constant projection of rank r(n) − s(n) − t(n), and gn is a constant projection of rank t(n). (cid:0)idM2 ⊗ α(n)(cid:1)(pn) =(cid:0)gn, c(0) n (cid:1), n + c(1) j (2) For every n ∈ Z≥0 and τ ∈ T(An) we have dτ (pn) ≤ 1. t(0) = 0. Proof. We prove the formula in (1) for pn by induction on n. The formula for (cid:0)idM2 ⊗ α(n)(cid:1)(pn) then follows from the definition of α(n). The formula holds for n = 0, since r(0) = s(0) = 1, t(0) = 0, and r(0) − s(0) − Now assume that it is known for n. Recall that Λn+1, n = λn ⊗ idMr(n) . (See Construction 6.1(8).) We suppress idM2 in the notation. With this convention, first take (f, g) in (6.1) to be (cid:0)c(0) the form required for c(0) s(n). In the same manner, we see that: n+1, while Λn+1,n(cid:0)c(0) n , 0(cid:1). The first coordinate Λn+1,n(cid:0)c(0) n , 0(cid:1)1 is of n , 0(cid:1)2 is a constant function of rank n , 0(cid:1)1 is a constant projection of rank d(n + 1)[r(n)− s(n)− t(n)]. n , 0(cid:1)2 is a constant projection of rank r(n) − s(n) − t(n). • Λn+1,n(0, gn)1 is a constant projection of rank t(n). • Λn+1,n(0, gn)2 is a constant projection of rank d(n + 1)t(n). • Λn+1,n(cid:0)c(1) • Λn+1,n(cid:0)c(1) Putting these together, we get in the first coordinate of Λn+1,n(pn) the direct sum of c(0) n+1 as described and a constant function of rank d(n + 1)[r(n) − s(n) − t(n)] + t(n). A computation shows that this expression is equal to r(n + 1)− s(n + 1)− t(n + 1). In the second coordinate we get a constant projection of rank s(n) +(cid:0)r(n) − s(n) − t(n)(cid:1) + d(n + 1)t(n) = t(n + 1). This completes the induction. 40 M. ALI ASADI-VASFI, NASSER GOLESTANI, AND N. CHRISTOPHER PHILLIPS For Part (2), we may assume that τ is extreme in T(An). Then there is x ∈ Xn ∐ Xn such that τ = trr(n) ⊗ evx. Therefore dτ (pn) = τ (pn) = 1 r(n) rank(pn(x)) =( s(n) t(n) r(n) r(n) + r(n)−s(n)−t(n) r(n) In each case, Lemma 6.2 implies dτ (pn) ≤ 1. This completes the proof. Lemma 6.11. Assume the notation and choices in Construction 6.1, and adopt the : (S2)s(n) → notation of Notation 6.8. Let n ∈ Z≥0. For j = 1, 2, . . . , s(n) let R(n) S2 be the j coordinate projection. Then: (1) There are orthogonal projections fn, hn in M2r(n)(cid:0)C(Xn)(cid:1) such that p′ (fn + hn, fn + hn), fn is the direct sum of the projections p ◦ R(n) j = 1, 2, . . . , s(n), and hn is a constant projection of rank r(n) − s(n). n = for j j (2) For every n ∈ Z≥0 and τ ∈ T(An), we have dτ (p′ n) = 1. Proof. We prove Part (1). Using Lemma 6.10(1), Lemma 6.4, and the definition of α(n) in the third step, we get x ∈ Xn ∐ ∅ x ∈ ∅ ∐ Xn. (cid:3) p′ n = (idM2 ⊗ Λn,0)(p, p) = (idM2 ⊗ Λn,0)(p, 0) + (idM2 ⊗ Λn,0)(0, p) n + c(1) n + gn, c(0) n + c(1) n + c(1) n + c(1) =(cid:0)c(0) n , gn(cid:1) +(cid:0)gn, c(0) Now it is enough to set fn = c(0) n and hn = c(1) n + gn. n (cid:1) =(cid:0)c(0) n + gn(cid:1). For Part (2), we may assume that τ is extreme in T(An). Then there is x ∈ Xn ∐ Xn such that τ = trr(n) ⊗ evx. Adding up the ranks given in Part (1), we see that rank(p′ (cid:3) n(x)) = r(n) for all x ∈ Xn ∐ Xn. The conclusion follows. Definition 6.12. Let A be a unital C*-algebra and let p be a projection in M∞(A). We call p trivial if there is n ∈ Z≥0 such that p is Murray-von Neumann equivalent to 1Mn(A). When n = 0, this means p = 0. Corollary 6.13. Adopt the assumptions and notation of Notation 6.8. Let n ∈ Z≥0 and let e = (e1, e2) be a projection in M∞(An) ∼= M∞(C(Xn) ⊕ C(Xn)) such that e1 is trivial. If there exists x ∈ M∞(An) such that kxex∗ − p′ 2 , then rank(e1) ≥ r(n) + s(n). Proof. Recall the line bundle L and the projection p from Notation 6.8. Also recall from Definition 2.13 that we use ≈ for Murray-von Neumann equivalence and / for Murray-von Neumann subequivalence. Let fn, hn ∈ M2r(n)(C(Xn)) be as in Lemma 6.11, and define q = fn + hn. The range of fn is isomorphic to the section space of the s(n)-dimensional vector bundle L×s(n) and q(p′ nk < 1 nXn∐∅)q = q. Now kxex∗ − p′ nk < 1 2 implies (cid:13)(cid:13)q(xex∗Xn∐∅ )q − q(cid:13)(cid:13) < 1 2 . Since e and q are projections, q / eXn∐∅ = e1. So there is projection w ∈ M∞(C(Xn)) such that q + w ≈ e1. Also, kxex∗ − p′ n is Murray-von Neumann equivalent to a subprojection of e. Therefore hn / e1, so rank(hn) ≤ rank(e1). Take e0 ∈ M∞(C(Xn)) to be a trivial projection of rank rank(e1) − rank(hn) such that e0 ⊥ hn. Since hn and e0 are trivial, e0 + hn ≈ e1. So 2 implies that p′ nk < 1 fn + hn + w ≈ e0 + hn. RADIUS OF COMPARISON OF THE CROSSED PRODUCT 41 Define k = rank(fn + w). Then k ≥ s(n). Now: of fn + w. • Let E1 be a vector bundle whose section space is isomorphic to the range • Let E2 be a trivial vector bundle whose section space is isomorphic to the • Set l = rank(hn). • Let H l be a trivial vector bundle whose section space is isomorphic to the range of e0. range of hn. Putting these together and using Theorem 9.1.5 of [23], we get fn + w ≈ e0. There- fore fn / e0. So rank(e0) ≥ 2s(n) by Lemma 6.9. Since e0 + hn ≈ e1, we have (cid:3) rank(e1) ≥ r(n) + s(n). Remark 6.14. We will use results of Niu from [26] to obtain an upper bound on the radius of comparison of our algebra. Niu introduced a notion of mean dimension for a diagonal AH-system, [26, Definition 3.6]. Suppose we are given a direct system of homogeneous algebras of the form An =(cid:0)C(K1,n) ⊗ Mj1(n)(cid:1) ⊕(cid:0)C(K2,n) ⊗ Mj2(n)(cid:1) ⊕ ··· ⊕(cid:0)C(Km(n),n) ⊗ Mjm(n)(n)(cid:1), in which each of the spaces involved is a connected finite CW complex, and the connecting maps are unital diagonal maps. Let γ denote the mean dimension of this system, in the sense of Niu. It follows trivially from [26, Definition 3.6] that γ ≤ lim n→∞ max(cid:18)(cid:26) dim(Kl,n) jl : l = 1, 2, . . . , m(n)(cid:27)(cid:19) . Theorem 6.2 of [26] then states that if A is the direct limit of a system as above, then rc(A) ≤ γ 2 . Since the system we are considering here is of this type, Niu's theorem applies. Theorem 6.15. Assume the notation and choices in Construction 6.1 and Nota- tion 6.8. Then rc(A) = κ. dim(Xn) Proof. Since limn→∞ r(n) = 2κ and the C*-algebra A was constructed with diagonal maps, we deduce from Remark 6.14 that rc(A) ≤ κ. Now it suffices to prove that rc(A) ≥ κ. Suppose ρ < κ. We show that A does not have ρ- comparison. Choose n ∈ Z>0 such that 1/r(n) < κ − ρ. Choose M ∈ Z≥0 such that ρ + 1 < M r(n) < κ + 1. Let e ∈ M∞(An) be a trivial projection of rank M . By slight abuse of notation, we use Λm,n to denote the amplified map from M∞(An) to M∞(Am) as well. For m > n, the rank of Λm,n(e) is M · r(m) r(n) . Suppose rank(cid:0)Λm,n(e)(cid:1) ≥ r(m) + s(m). Then, by the choice of M , We claim that the rank of Λm,n(e) is strictly less than r(m) + s(m) for m > n. < (κ + 1)r(m). r(m) + s(m) ≤ M · r(m) r(n) Thus s(m) follows. r(m) < κ. This contradicts Lemma 6.3 and Construction 6.1(2). So the claim Now, for any tracial state τ on Am (and thus for any tracial state on A), we have, using Lemma 6.11(2) in the last step, dτ (Λm,n(e)) = τ (Λm,n(e)) = 1 r(m) · M · r(m) r(n) > 1 + ρ = dτ (p′ m) + ρ. 42 M. ALI ASADI-VASFI, NASSER GOLESTANI, AND N. CHRISTOPHER PHILLIPS On the other hand, if Λ∞,0(cid:0)(p, p)(cid:1) / Λ∞,n(e) then, in particular, there exists some m > n and x ∈ M∞(Am) such that kxΛm,n(e)x∗ − p′ lary 6.13, we have 2 . Using Corol- mk < 1 This is a contradiction, and we have proved that A does not have ρ-comparison. (cid:3) rank(Λm,n(e)) ≥ r(m) + s(m). We now determine the radius of comparison of the crossed product in our exam- ple. The methods are very similar. Construction 6.16. Assume the notation and choices in Parts (1), (3), (4), and (5) in Construction 6.1. Mr(n). (1) For n ∈ Z≥0, we define Bn = C(Xn)⊗ M2r(n), identified with C(Xn, M2)⊗ (2) Let s ∈ M2 be the unitary matrix s = ( 0 1 n+1,n : Bn → Bn+1 1 0 ). Define Λ′ by f ◦ P (n) 1 0 0 f ◦ P (n) 2 . . . f ◦ P (n) d(n+1) sf (xn)s∗   ⊗ c Λ′ n+1,n(f ⊗ c) = for f ∈ C(Xn, M2) and c ∈ Mr(n). With abuse of notation (the expres- sion sf (xn)s∗ · 1C(Xn+1) is the constant function Xn+1 → M2 with value sf (xn)s∗), the analog of (6.2) is (6.4) Λ′ n+1,n(f ⊗ c) = d(n+1)Xj=1 j ⊗ ej,j ⊗ c f ◦ P (n) + sf (xn)s∗ · 1C(Xn+1) ⊗ ed(n+1)+1, d(n+1)+1 ⊗ c. It is clear that Λ′ n+1, n is injective for all n ∈ Z≥0. (Bn, Λ′ n+1, n). (3) Define B = lim −→ Lemma 6.17. Assume the notation and choices in Construction 6.1 and Construc- tion 6.16. Then C∗(Z/2Z, A, α) ∼= B. Proof. For t ∈ Z/2Z, as in Notation 3.1 let ut be the standard unitary in a crossed product by Z/2Z. (In this proof, no confusion will be caused by using the same letter in all crossed products.) For n ∈ Z≥0, there is a homomorphism ψn+1,n : C∗(Z/2Z, An, α(n)(cid:1) → C∗(Z/2Z, An+1, α(n+1)(cid:1) such that for f, g ∈ C(Xn), t ∈ Z/2Z, and c ∈ Mr(n). In view of Lemma 6.4, we can apply Theorem 9.4.34 of [17] to get an isomorphism ψn+1,n(cid:0)[(f, g) ⊗ c]ut(cid:1) =(cid:2)Λn+1,n((f, g) ⊗ c)(cid:3)ut =(cid:2)λn((f, g)) ⊗ c(cid:3)ut −→(cid:0)C∗(Z/2Z, An, α(n)(cid:1), (ψn+1,n)n∈Z≥0(cid:1). C∗(Z/2Z, A, α) ∼= lim RADIUS OF COMPARISON OF THE CROSSED PRODUCT 43 On the other hand, we have an isomorphism ϕn : C∗(Z/2Z, An, α(n)) → Bn which is defined for fm, gm ∈ C(Xn) and cm ∈ Mr(n) for m = 0, 1 by [(f0, g0) ⊗ c0]u0 + [(f1, g1) ⊗ c1]u1 7→(cid:18)f0 ⊗ c0 g1 ⊗ c1 f1 ⊗ c1 g0 ⊗ c0(cid:19) . Using matrix unit notation, the right hand side is f0 ⊗ e1,1 ⊗ c0 + f1 ⊗ e1,2 ⊗ c1 + g1 ⊗ e2,1 ⊗ c1 + g0 ⊗ e2,2 ⊗ c0. Using (6.2) and (6.4), one checks that the diagram ψn+1, n −−−−−→ C∗(Z/2Z, An+1, α(n+1)) C∗(Z/2Z, An, α(n)) yϕn Bn Λ′ n+1, n −−−−−→ yϕn+1 Bn+1 (cid:3) commutes for every n ∈ Z≥0. The result follows. Notation 6.18. Let p ∈ C(X0, M2) be the Bott projection, as in Notation 6.8. Assuming the notation and choices in 6.16, for n ∈ Z≥0 set qn = Λ′ n,0(p) ∈ Bn. In particular, q0 = p. Lemma 6.19. Adopt the assumptions and notation of Notation 6.18. Let n ∈ Z≥0 and for j = 1, 2, . . . , s(n) let R(n) : (S2)s(n) → S2 be the j coordinate projection. Then: (1) There are orthogonal projections yn, zn in M2r(n)(cid:0)C(Xn)(cid:1) such that qn = yn +zn, yn is the direct sum of the projections p◦R(n) and zn is a constant projection of rank r(n) − s(n). (2) For every n ∈ Z≥0 and τ ∈ T(Bn), we have dτ (qn) = 1 2 . for j = 1, 2, . . . , s(n), j j j k ◦ P (n) Proof. The proof of (1) is very similar to that of Lemma 6.11(1), but simpler be- cause there is only one summand. The basic facts for the induction step are that n+1,n(yn) is the direct sum of the projections p ◦ R(n) Λ′ for j = 1, 2, . . . , s(n) and k = 1, 2, . . . , d(n + 1), and a constant projection of rank s(n), and that Λ′ n+1,n(zn) is a constant projection of rank l(n + 1)[r(n) − s(n)]. We omit the details. The proof of (2) is essentially the same as that of Lemma 6.11(2), and is omitted. (cid:3) Corollary 6.20. Adopt the assumptions and notation of Notation 6.18. Let n ∈ Z≥0 and let e be a trivial projection in M∞(Bn) ∼= M∞(C(Xn)). If there exists x ∈ M∞(Bn) such that kxex∗ − qnk < 1 Proof. The proof is essentially the same as that of Corollary 6.13. We use Lemma 6.19 and the projections yn and zn instead of Lemma 6.11 and the projections fn (cid:3) and hn. 2 then rank(e) ≥ r(n) + s(n). The next result is the analog of Theorem 6.15. It shows that in our example, we get equality in Theorem 4.1 and Theorem 4.5. Theorem 6.21. Assume the notation and choices in Construction 6.1 and Nota- tion 6.8. Then rc(cid:0)C∗(Z/2Z, A, α)(cid:1) = κ 2 and rc(Aα) = κ. 44 M. ALI ASADI-VASFI, NASSER GOLESTANI, AND N. CHRISTOPHER PHILLIPS The proof is similar to that of Theorem 6.15. We give details to show where the factor 1 2 comes from, and for convenient reference in a paper in preparation. Proof of Theorem 6.21. We prove the first part of the conclusion. The second part then follows from Theorem 4.5. 2 . Since limn→∞ Because C∗(Z/2Z, A, α) is isomorphic to the C*-algebra B by Lemma 6.17, it dim(Xn) suffices to show that rc(B) = κ 2r(n) = κ and the C*-algebra B was constructed with diagonal maps, we deduce from Remark 6.14 that rc(B) ≤ κ 2 . Now it suffices to prove that rc(B) ≥ κ 2 . We show that B does not have ρ-comparison. Choose n ∈ Z>0 such that 1/r(n) < κ 2 − ρ. Choose M ∈ Z≥0 such that ρ + 1 2 . Let e ∈ M∞(Bn) be a trivial projection of rank M . By slight abuse of notation, we use Λ′ m,n to denote the amplified map m,n(e) is M · r(m) from M∞(Bn) to M∞(Bm) as well. For m > n, the rank of Λ′ r(n) . We claim that the rank of Λ′ m,n(e) is strictly less than r(m) + s(m) for m > n. 2 . Suppose ρ < κ 2r(n) < κ 2 < M Suppose rank(cid:0)Λ′ m,n(e)(cid:1) ≥ r(m) + s(m). Then, considering the choice of M , 2 + 1 r(m) + s(m) ≤ M · r(m) r(n) < (κ + 1)r(m). Thus s(m) r(m) < κ. This contradicts Construction 6.1(2). So the claim follows. Now, for any extreme tracial state τ on Bm (and thus for any trace on B), we have, using Lemma 6.19(2) in the last step, dτ (Λ′ m,n(e)) = τ (Λ′ m,n(e)) = 1 2r(m) · M · r(m) r(n) > 1 2 + ρ = dτ (qm) + ρ. On the other hand, if Λ′ m > n and x ∈ M∞(Bm) such that kxΛ′ we get ∞,0(p) / Λ′ ∞,n(e) then, in particular, there exists some 2 . Using Corollary 6.20, m,n(e)x∗ − qmk < 1 This is a contradiction, and we have proved that B does not have ρ-comparison. (cid:3) rank(Λ′ m,n(e)) ≥ r(m) + s(m). Example 6.22. We show that, in Theorem 4.1 and Theorem 4.5, the weak tracial Rokhlin property can't be replaced by pointwise outerness. Let A and α : Z/2Z → Aut(A) be as in Lemma 6.4, set B = C∗(Z/2Z, A, α), Theorem 6.21, and Lemma 6.3 that the inequalities in Theorem 4.1 and Theorem 4.5 fail for the action β. and let β =bα : Z/2Z → Aut(B) be the dual action. It follows from Theorem 6.15, We already know that B is simple, and B is stably finite because it is an AH al- gebra. It remains to show that β is pointwise outer. Suppose not. Then in fact β is an inner action, that is, given by conjugation by a unitary of order 2. (See Exercise 8.2.7 of [17].) So C∗(Z/2Z, B, β) ∼= B ⊕ B. But by Takai duality C∗(Z/2Z, B, β) ∼= M2(A), which is simple. Pointwise outerness is proved. 7. Open problems The most obvious problem is whether equality always holds in Theorem 4.1 and Theorem 4.5. One might even hope that the reverse inequalities (7.1) rc(Aα) ≥ rc(A) and rc(cid:0)C∗(G, A, α)(cid:1) = rc(cid:0)C∗(G, A, α)(cid:1) ≥ 1 card(G) · rc(A)? 1 card(G) · rc(A). RADIUS OF COMPARISON OF THE CROSSED PRODUCT 45 Question 7.1. Let G be a finite group, let A be an infinite-dimensional stably finite simple unital C*-algebra, and let α : G → Aut(A) be an action of G on A which has the weak tracial Rokhlin property. Does it follow that rc(Aα) = rc(A) and hold without restrictions on the action. Quite different methods seem to be needed for this question. Suppose, for example, that we were able to prove (7.1) for point- wise outer actions. Suppose G is finite abelian, α : G → Aut(A) is pointwise outer, and, with B = C∗(G, A, α), the dual action β = bα : bG → Aut(B) is pointwise outer and B has strict comparison. We would be able to deduce that C∗(cid:0)bG, B, β(cid:1) has strict comparison. This outcome is at least heuristically related to the long standing open question of whether the crossed product of a simple C*-algebra with stable rank one by a finite group again has stable rank one. Indeed, if B is classi- fiable in the sense of the Elliott program, and the tracial state space has compact finite-dimensional extreme boundary, it would follow (see Corollary 7.9 of [29], Corollary 1.2 of [43], or Corollary 4.7 of [45]) that C∗(cid:0)bG, B, β(cid:1) is Z-stable, and therefore from Theorem 6.7 of [41] that C∗(cid:0)bG, B, β(cid:1) has stable rank one. This case of the problem has been solved [31], but the proof depends on major results in the classification program. In the example in Section 6, the group action on T(A) is highly nontrivial. Question 7.2. Does there exist an action of a nontrivial finite group with the weak tracial Rokhlin property on a simple separable unital C*-algebra A with rc(A) > 0 and such that every tracial state is invariant? One can ask for even more. Question 7.3. Does there exist an action of a nontrivial finite group with the Rokhlin property on a simple separable unital C*-algebra A with rc(A) > 0 and unique tracial state? By combining methods of Villadsen [46] with those of Section 6, one should be able to at least produce an example of a simple separable unital nuclear C*- algebra A and an action α : Z/2Z → Aut(A) such that A does not have stable rank one, α has the Rokhlin property, and A has exactly two extreme tracial states, which are interchanged by the action α. Question 7.4. Let A be an infinite-dimensional simple unital C*-algebra with stable rank one, let G be a finite group, and let α : G → Aut(A) be an action with the weak tracial Rokhlin property. Does it follow that C∗(G, A, α) and Aα have stable rank one? This is wanted for improvement of Corollary 5.6. Question 7.5. Are the stable rank one hypotheses in Proposition 5.9 and Corol- lary 5.6 really necessary? 46 M. ALI ASADI-VASFI, NASSER GOLESTANI, AND N. CHRISTOPHER PHILLIPS That is, assuming the action has the Rokhlin property or weak tracial Rokhlin property as appropriate, does one get isomorphisms W(ι) : W(Aα) → W(A)α or W+(ι) : W+(Aα) ∪ {0} → W+(A)α ∪ {0}, or rather than just Cu+(ι) : Cu+(Aα) ∪ {0} → Cu+(A)α ∪ {0}? Cu(ι) : Cu(Aα) → Cu(A)α One possible generalization of the results of this paper is to the nonunital case. This will be treated in [5] (by a different set of authors). Complications include the additional complexity of the definition of the weak tracial Rokhlin property (see Definition 3.1 of [14]), and what to substitute for the conventional definition of the radius of comparison. 8. Acknowledgments This research was done while the first author was a visiting scholar at the Uni- versity of Oregon during the period March 2018 to September 2019. He is thankful to that institution for its hospitality. He was partially supported by the University of Tehran. This paper will be part of first author's PhD. dissertation. The research of the third author was partially supported by the Simons Foun- dation Collaboration Grant for Mathematicians #587103. The first author thanks Q. Wang for pointing out Corollary II.4.3 in [6]. All three authors would like to thank M. Amini, I. Hirshberg, and S. Jamali for sharing some of their unpublished work with us. The first author is grateful to M. B. Asadi for motivating him to study operator algebras in the first year of his Ph.D. program. References [1] C. A. Akemann and F. Shultz, Perfect C*-algebras, Memoirs Amer. Math. Soc., vol. 55 no. 326(1985). [2] M. Amini, N. Golestani, S. Jamali, N. C. Phillips, Simple tracially Z-absorbing C*-algebras, in preparation. [3] P. Ara, F. Perera, and A. S. Toms, K-theory for operator algebras. Classification of C*- algebras, pages 1 -- 71 in: Aspects of Operator Algebras and Applications, P. Ara, F Lled´o, and F. Perera (eds.), Contemporary Mathematics vol. 534, Amer. Math. Soc., Providence RI, 2011. [4] D. Archey and N. C. Phillips, Permanence of stable rank one for centrally large subalgebras and crossed products by minimal homeomorphisms, J. Operator Theory, to appear. [5] M. A. Asadi-Vasfi, The Cuntz semigroup of the crossed product of a nonunital C*-algebra by a finite group, in preparation. [6] B. Blackadar and D. Handelman, Dimension functions and traces on C*-algebras, J. Funct. Anal. 45(1982), 297 -- 340. [7] L. G. Brown, Stable isomorphism of hereditary subalgebras of C*-algebras, Pacific J. Math. 71(1977), 335 -- 348. [8] J. Cuntz, Dimension functions on simple C*-algebras, Math. Ann. 233(1978), 145 -- 153. [9] M. Dadarlat, G. Nagy, A. N´emethi, and C. Pasnicu, Reduction of topological stable rank in inductive limits of C*-algebras, Pacific J. Math. 153(1992), 267 -- 276. [10] G. A. Elliott, T. M. Ho, and A. S. Toms, A class of simple C*-algebras with stable rank one, J. Funct. Anal. 256(2009), 307 -- 322. [11] G. A. Elliott and Z. Niu, On the radius of comparison of a commutative C*-algebra, Canad. Math. Bull. 56(2013), 737 -- 744. [12] G. A. Elliott, L. Robert, and L. Santiago, The cone of lower semicontinuous traces on a C*-algebra, Amer. J. Math. 133(2011), 969 -- 1005. RADIUS OF COMPARISON OF THE CROSSED PRODUCT 47 [13] Q. Fan and X. Fang, Stable rank one and real rank zero for crossed products by finite group actions with the tracial Rokhlin property, Chin. Ann. Math. Ser. B 30(2009), 179 -- 186. [14] M. Forough and N. Golestani, Weak tracial Rokhlin property for finite group actions on simple C*-algebras, preprint (arXiv: 1711.10818v2 [math.OA]). [15] E. Gardella, I. Hirshberg, and L. Santiago, Rokhlin dimension: duality, tracial properties, and crossed products, preprint (arXiv: 1709.00222v1 [math.OA]). [16] E. Gardella and L. Santiago, Equivariant *-homomorphisms, Rokhlin constraints and equi- variant UHF-absorption, J. Funct. Anal. 270(2016), 2543 -- 2590. [17] T. Giordano, D. Kerr, N. C. Phillips, and A. Toms, Crossed Products of C*-Algebras, Topo- logical Dynamics, and Classification, edited by Francesc Perera, Advanced Courses in Math- ematics, CRM Barcelona, Birkhauser/Springer, Cham, 2018. [18] N. Golestani, in preparation. [19] U. Haagerup, Quasitraces on exact C*-algebras are traces, C. R. Math. Acad. Sci. Soc. Canada 36(2014), 67 -- 92. [20] I. Hirshberg and J. Orovitz, Tracially Z-absorbing C*-algebras, J. Funct. Anal. 265(2013), 765 -- 785. [21] I. Hirshberg and N. C. Phillips, Rokhlin dimension: obstructions and permanence properties, Doc. Math. 20(2015), 199 -- 236. [22] I. Hirshberg and N. C. Phillips, Internal asymmetry, in preparation. [23] D. Husemoller, Fibre Bundles (3rd ed.), Springer-Verlag, New York, Berlin, Heidelberg, Lon- don, Paris, Tokyo, Hong Kong, Barcelona, Budapest, 1994. [24] M. Izumi, Finite group actions on C*-algebras with the Rohlin property. II , Adv. Math. 184(2004), 119 -- 160. [25] H. Lin, An Introduction to the Classification of Amenable C*-Algebras, World Scientific, River Edge NJ, 2001. [26] Z. Niu, Mean dimension and AH-algebras with diagonal maps, J. Funct. Anal. 266(2014), 4938 -- 4994. [27] E. Kirchberg and M. Rørdam, Non-simple purely infinite C*-algebras, Amer. J. Math. 122(2000), 637 -- 666. [28] E. Kirchberg and M. Rørdam, Infinite non-simple C*-algebras: absorbing the Cuntz algebra O∞, Adv. Math. 167(2002), 195 -- 264. [29] E. Kirchberg and M. Rørdam, Central sequence C*-algebras and tensorial absorption of the Jiang-Su algebra, J. reine angew. Math. 695(2014), 175 -- 214; erratum, J. reine angew. Math. 695(2014), 215 -- 216. [30] J. Orovitz, N. C. Phillips, and Q. Wang, Strict comparison and crossed products, in prepa- ration. [31] H. Osaka, Stable rank for crossed products by actions of finite groups on C*-algebras, preprint (arXiv: 1708.02665v1 [math.OA]). [32] H. Osaka and N. C. Phillips, Crossed products by finite group actions with the Rokhlin prop- erty, Math. Z. 270(2012), 19 -- 42. [33] H. Osaka and T. Teruya, The Jiang-Su absorption for inclusions of unital C*-algebras, Canad. J. Math. 70(2018), 400 -- 425. [34] F. Perera and A. S. Toms, Recasting the Elliott conjecture, Math. Ann. 338(2007), 669 -- 702. [35] N. C. Phillips, The tracial Rokhlin property for actions of finite groups on C*-algebras, Amer. J. Math. 133(2011), 581 -- 636. [36] N. C. Phillips, Freeness of actions of finite groups on C*-algebras, pages 217 -- 257 in: Operator structures and dynamical systems, M. de Jeu, S. Silvestrov, C. Skau, and J. Tomiyama (eds.), Contemporary Mathematics vol. 503, Amer. Math. Soc., Providence RI, 2009. [37] N. C. Phillips, Finite cyclic group actions with the tracial Rokhlin property, Trans. Amer. Math. Soc. 367(2015), 5271 -- 5300. [38] N. C. Phillips, Large subalgebras, preprint (arXiv: 1408.5546v2 [math.OA]). [39] M. A. Rieffel, Actions of finite groups on C*-algebras, Math. Scand. 47(1980), 157 -- 176. [40] M. Rørdam, On the structure of simple C*-algebras tensored with a UHF-algebra. II , J. Funct. Anal. 107(1992), 255 -- 269. [41] M. Rørdam, The stable and the real rank of Z-absorbing C*-algebras, Internat. J. Math. 15(2004), 1065 -- 1084. [42] J. Rosenberg, Appendix to O. Bratteli's paper on "Crossed products of UHF algebras", Duke Math. J. 46(1979), 25 -- 26. 48 M. ALI ASADI-VASFI, NASSER GOLESTANI, AND N. CHRISTOPHER PHILLIPS [43] Y. Sato, Trace spaces of simple nuclear C*-algebras with finite-dimensional extreme bound- ary, preprint (arXiv: 1209.3000v1 [math.OA]). [44] A. S. Toms, Flat dimension growth for C*-algebras, J. Funct. Anal. 238(2006), 678 -- 708. [45] A. S. Toms, S. White, and W. Winter, Z-stability and finite-dimensional tracial boundaries, Int. Math. Res. Not. IMRN 292(2015), 2707 -- 2727. [46] J. Villadsen, On the stable rank of simple C*-algebras, J. Amer. Math. Soc. 12(1999), 1091 -- 1102. [47] N. E. Wegge-Olsen, K-Theory and C*-Algebras, Oxford University Press, Oxford etc., 1993. Department of Mathematics, University of Oregon, Eugene OR 97403-1222, USA. E-mail address: [email protected] Current address: School of Mathematics, Statistics and Computer Science, College of Science, University of Tehran, Tehran, Iran. E-mail address: [email protected] Department of Pure Mathematics, Faculty of Mathematical Sciences, Tarbiat Modares University, P.O. Box 14115 -- 134, Tehran, Iran E-mail address: [email protected] Department of Mathematics, University of Oregon, Eugene OR 97403-1222, USA.
1608.01782
2
1608
2016-12-14T11:56:53
The Toeplitz noncommutative solenoid and its KMS states
[ "math.OA" ]
We use Katsura's topological graphs to define Toeplitz extensions of Latr\'emoli\`ere and Packer's noncommutative-solenoid C*-algebras. We identify a natural dynamics on each Toeplitz noncommutative solenoid and study the associated KMS states. Our main result shows that the space of extreme points of the KMS simplex of the Toeplitz noncommutative torus at a strictly positive inverse temperature is homeomorphic to a solenoid; indeed, there is an action of the solenoid group on the Toeplitz noncommutative solenoid that induces a free and transitive action on the extreme boundary of the KMS simplex. With the exception of the degenerate case of trivial rotations, at inverse temperature zero there is a unique KMS state, and only this one factors through Latr\'emoli\`ere and Packer's noncommutative solenoid.
math.OA
math
THE TOEPLITZ NONCOMMUTATIVE SOLENOID AND ITS KMS STATES NATHAN BROWNLOWE, MITCHELL HAWKINS, AND AIDAN SIMS Abstract. We use Katsura's topological graphs to define Toeplitz extensions of Latr´emoli`ere and Packer's noncommutative-solenoid C ∗-algebras. We identify a natural dynamics on each Toeplitz noncommutative solenoid and study the associated KMS states. Our main result shows that the space of extreme points of the KMS simplex of the Toeplitz noncommutative torus at a strictly positive inverse temperature is homeomorphic to a solenoid; indeed, there is an action of the solenoid group on the Toeplitz noncommutative solenoid that induces a free and transitive action on the extreme boundary of the KMS simplex. With the exception of the degenerate case of trivial rotations, at inverse temperature zero there is a unique KMS state, and only this one factors through Latr´emoli`ere and Packer's noncommutative solenoid. 1. Introduction In this paper, we describe the KMS states of Toeplitz extensions of the noncommutative solenoids constructed by Latr´emoli`ere and Packer [22]. We prove that the extreme boundary of the KMS simplex is homeomorphic to a topological solenoid. In recent years, following Bost and Connes' work [2] relating KMS theory to the Riemann zeta function, there has been a great deal of interest in the KMS structure of C ∗-algebras associated to algebraic and com- binatorial objects. In particular Laca and Raeburn's results [20] about the Toeplitz algebra of the ax + b-semigroup over N precipitated a surge of activity around computations of KMS states for Toeplitz-like extensions. Various authors have studied KMS states on Toeplitz alge- bras associated to algebraic objects [4, 21, 6], directed graphs [14, 12, 5], higher-rank graphs [29, 13, 9], C ∗-correspondences [19, 15], and topological graphs [1]. The results suggest that the KMS structure of such algebras for their natural gauge actions frequently encodes key features of the generating object. The noncommutative solenoids AS θ are C ∗-algebras introduced by Latr´emoli`ere and Packer in [22]. They are among the first examples of twisted C ∗-algebras of non-compactly-generated abelian groups to be studied in detail, and have interesting representation-theoretic properties [23, 24]. In addition to the definition of noncommutative solenoids as twisted group C ∗-algebras, Latr´emoli`ere and Packer provide a number of equivalent descriptions. The one we are interested in realises them as direct limits of noncommutative tori. Specifically, given a positive integer N and a sequence θn of real numbers such that N 2θn+1 − θn is an integer for every n, there are homomorphisms Aθn → Aθn+1 that send the canonical unitary generators of Aθn to the Nth powers of the corresponding generators of Aθn+1. The noncommutative solenoid for the sequence θ = (θn) is the direct limit of the Aθn under these homomorphisms. Latr´emoli`ere and Packer's work focusses on features like simplicity, K-theory and classification of noncommuta- tive solenoids. Here we use Katsura's theory of topological graph C ∗-algebras [16] to introduce a class of Toeplitz extensions T S of noncommutative solenoids, realised as direct limits of Toeplitz θ extensions T (Eθn) of noncommutative tori, and then study their KMS states. Our main result says that at inverse temperatures above zero, the extreme boundary of the KMS simplex of T S that θ induces a free and transitive action on the extreme KMS states. This is further evidence that KMS structure for Toeplitz-like algebras recovers key features of the underlying generating is homeomorphic to the classical solenoid S , and there is an action of S on T S θ Date: April 18, 2019. 1991 Mathematics Subject Classification. 46L55 (primary); 28D15, 37A55 (secondary). This research was supported by the Australian Research Council grant DP150101595. 1 2 BROWNLOWE, HAWKINS, AND SIMS objects. Interestingly, this homeomorphism is subtler than one might expect: though the results of [1] show that the KMS simplex of each approximating subalgebra T (Eθn) ⊆ T S θ has extreme boundary homeomorphic to the circle, these homeomorphisms are not compatible with the connecting maps in the inductive system. In fact, none of the extreme points in the KMS simplex of any T (Eθn) extend to KMS states of T S . Identifying the simplex of KMS states of a θ given T (Eθn) that do extend to KMS states of T S requires a careful analysis of the interaction θ between the subinvariance relation, described in [1], that characterises KMS states on the T (Eθn) and the compatibility relation imposed by the connecting maps T (Eθn) ֒→ T (Eθn+1). We think the ideas involved in this analysis may be applicable to other investigations of KMS states on direct-limit C ∗-algebras. Our main result also shows that at inverse temperature 0 there is a unique KMS state (unless all the θn are zero, a degenerate case that we discuss separately), and that there are no KMS states at inverse temperatures below zero. Perhaps surprisingly, for nonzero θ the structure of the KMS simplex of T S θ does not depend on whether the θn are rational. We proceed as follows. After a brief preliminaries section, we begin in Section 3 by consid- ering KMS states for actions on direct limits that preserve the approximating subalgebras. We record a general -- and presumably well known -- description of the KMS simplex as a projec- tive limit of the KMS simplices of the approximating subalgebras. The connecting maps in this projective system need not be surjective, which is the cause of the subtleties that arise in computing the KMS states of Toeplitz noncommutative solenoids later in the paper. In Sec- tion 4, we consider the topological graph Eγ that encodes rotation on the circle R/Z by angle γ ∈ R. We describe the Toeplitz algebra T (Eγ) of this topological graph as universal for an isometry S and a representation π of C(R/Z), and its topological-graph C ∗-algebra O(Eγ) as the quotient by the ideal generated by 1 − SS∗. In particular, O(Eγ) is canonically isomorphic to the noncommutative torus Aγ. In Section 5 we consider a sequence θ = (θn) in R/Z such that N 2θn+1 = θn for all n. We use our description of T (Eγ) from the preceding section to describe homomorphisms ψn : T (Eθn) → T (Eθn+1) that descend through the quotient maps to the homomorphisms τn : O(Eθn) → O(Eθn+1) for which the noncommutative solenoid AS is θ isomorphic to lim −→ (O(Eθn), τn). 0 := lim −→ In Section 6, we define the Toeplitz noncommutative solenoid as T S θ for α at an inverse temperature β > 0 as a projective limit of spaces Ωrn (T (Eθn), ψn), by analogy with the description of AS θ outlined in Section 5. We describe a dynamics α on T S θ built from the gauge actions on the approximating subalgebras T (Eθn). Though the gauge actions on the T (Eθn) are all periodic R-actions, the dynamics α is not. We are interested in the KMS states for this dynamics. The case θ = 0 := (0, 0, 0, . . . ) is a degenerate case, and we outline in Remark 6.5 how to describe the KMS states in this instance by decomposing both the algebra T S and the dynamics α as tensor products. Since θn 6= 0 implies θn+1 6= 0, we can thereafter assume, without loss of generality, that every θn is nonzero. In the remainder of Section 6, we use our results about direct limits from Section 3 to realise the KMS simplex of T S sub of probability θ measures on R/Z that satisfy a suitable subinvariance condition. This involves an interesting interplay between the subinvariance condition for KMS states on the T (Eθn) obtained from [1], and the compatibility condition coming from the connecting maps ψn. We believe that this analysis and our analysis of the space Ωrn sub in Section 7 may be of independent interest from the point of view of ergodic theory. The theorems in [1] are silent on the case β = 0, so we must argue this case separately, and our results for this case in Section 6 appear less sharp than for β > 0: they show only that the KMS0-simplex embeds in the projective limit of the spaces Ω0 sub. But we shall see later that the subinvariance condition at β = 0 has a unique solution, so that the projective limit in this case is a one-point set. So our embedding result for β = 0 is sufficient to show that there is a unique KMS0 state. In Section 7 we analyse the space Ωr sub for r > 0. We first construct a measure mr satisfying the desired subinvariance relation, and then show that the measures obtained by composing this mr with rotations are all of the extreme points of Ωr sub. This yields an isomorphism of THE TOEPLITZ NONCOMMUTATIVE SOLENOID AND ITS KMS STATES 3 Ωr sub with the space of Borel probability measures on R/Z. A key step in our analysis is the characterisation in [12] of the subinvariant measures on the vertex set of a simple-cycle graph. We then turn in Section 8 to the proof of our main theorem. The key step is to establish that the connecting maps ψn : T (Eθn) → T (Eθn+1) induce surjections Ωrn+1 sub by showing that the induced maps carry extreme points to extreme points. ։ Ωrn sub 2. Preliminaries In this section we recall the background that we need on topological graphs and their C ∗- algebras, as introduced by Katsura in [16]. We then recall the notion of a KMS state for a C ∗-algebra A and dynamics α. Topological graphs and their C ∗-algebras. For details of the following, see [16]. A topo- logical graph E = (E0, E1, r, s) consists of locally compact Hausdorff spaces E0 and E1, a continuous map r : E1 → E0, and a local homeomorphism s : E1 → E0. In [16] Katsura con- structs from each topological graph E a Hilbert C0(E0)-bimodule X(E) and two C ∗-algebras: the Toeplitz algebra T (E) and the graph C ∗-algebra O(E). In this article we only encounter topological graphs of the form E = (Z, Z, id, h), where h : Z → Z is a homeomorphism of a compact Hausdorff space Z, so we only discuss the details of X(E), T (E) and O(E) in this setting. When E = (Z, Z, id, h), where Z is compact, the module X(E) is a copy of C(Z) as a Banach space. The left and right actions are given by g1 · f · g2(z) = g1(z)f (z)g2(h(z)), for g1, g2 ∈ C(Z), f ∈ X(E), and the inner product by hf1, f2i(z) = f1(h−1(z))f2(h−1(z)), for f1, f2 ∈ X(E). We denote by ϕ the homomorphism C(Z) → L(X(E)) implementing the left action. In this case ϕ is injective. A representation of X(E) in a C ∗-algebra B is a pair (ψ, π), consisting of a linear map ψ : X(E) → B and a homomorphism π : C(Z) → B satisfying ψ(f · h) = ψ(f )π(h), ψ∗(f )ψ(g) = π(hf, gi) and ψ(h · f ) = π(h)ψ(f ) for all f, g ∈ X(E) and h ∈ C(Z). The Toeplitz algebra T (E) is the Toeplitz algebra of X(E), in the sense of [10], which is the universal C ∗-algebra generated by a representation of X(E). We denote by (i1 X(E)) the representation generating T (E). X(E), i0 For f1, f2 ∈ X(E) there is an adjointable operator Θf1,f2 ∈ L(X(E)) given by Θf1,f2(g) = f1hf2, giC(Z) = f1f ∗ 2 g. The algebra of generalised compact operators on X(E) is K(X(E)) := span{Θf1,f2 : f1, f2 ∈ X(E)}. Since Θ1,1 = 1L(X(E)), we have K(X(E)) = L(X(E)). For a representation (ψ, π) of X(E) in B there is a homomorphism (ψ, π)(1) : K(X(E)) → B satisfying (ψ, π)(1)(Θf1,f2) = ψ(f1)ψ(f2)∗ (see [26, page 202]). The graph algebra O(E) is the Cuntz -- Pimsner algebra of X(E). So O(E) is the quotient of T (E) by the ideal generated by {(i1 X(E), i0 X(E))(1)(ϕ(h)) − i0 X(E)(h) : h ∈ C(Z)}, and is the universal C ∗-algebra generated by a covariant representation of X(E) -- that is, a representation (ψ, π) satisfying (ψ, π)(1)(ϕ(h)) = π(h) for all h ∈ C(Z). We denote the quotient map T (E) → O(E) by q, and we define (j1 i0 X(E)), the covariant representation generating O(E). X(E), j0 X(E)) := (q ◦ i1 X(E), q ◦ 4 BROWNLOWE, HAWKINS, AND SIMS KMS states. For details of the following, see [3]. Given a C ∗-algebra A and an action α : R → Aut(A), we say that a ∈ A is analytic for α if the function t 7→ αt(a) is the restriction of an analytic function z 7→ αz(a) from C into A. The set of analytic elements is always norm dense in A. A state φ of A is a KMS0-state if it is an α-invariant trace on A. For β ∈ R \ {0}, a state φ of A is a KMSβ-state, or a KMS-state at inverse temperature β, for the system (A, α) if it satisfies the KMS condition φ(ab) = φ(bαiβ(a)) for all analytic a, b ∈ A. It suffices to check this condition for all a, b in any α-invariant set of analytic elements that spans a dense subspace of A. The collection of KMSβ-states for a dynamics α on a unital C ∗-algebra A forms a Choquet simplex, and we will denote it by KMSβ(A, α). 3. KMS structure of direct limit C ∗-algebras The C ∗-algebras of interest to us in this paper are examples of direct-limit C ∗-algebras. In this short section we show that the simplex of KMS states of a direct-limit C ∗-algebra, for an action that preserves the approximating subalgebras, is the projective limit of the simplices of KMS states of the approximating subalgebras. Proposition 3.1. Suppose β ∈ [0, ∞), and that {(Aj, ϕj, αj) : j ∈ N} is a sequence of unital C ∗-algebras Aj, injective unital homomorphisms ϕj : Aj → Aj+1, and strongly continuous actions αj : R → Aut Aj satisfying αj+1,t ◦ ϕj = ϕj ◦ αj,t for all j ∈ N and t ∈ R. Denote by A∞ the direct limit lim (Aj, ϕj), and by ϕj,∞ the canonical maps Aj → A∞ satisfying ϕj+1,∞ ◦ ϕj = −→ ϕj,∞ for each j ∈ N. There is a strongly continuous action α : R → Aut A∞ satisfying ϕj,∞ ◦ αj,t = αt ◦ ϕj,∞ for each j ∈ N and t ∈ R. Moreover, there is an affine isomorphism from KMSβ(A∞, α) onto lim ←− (KMSβ(Aj, αj), φ 7→ φ ◦ ϕj−1) that sends φ to (φ ◦ ϕj,∞)∞ j=0. Proof. For each j ∈ N and t ∈ R we have (ϕj+1,∞ ◦ αj+1,t) ◦ ϕj = ϕj+1,∞ ◦ ϕj ◦ αj,t = ϕj,∞ ◦ αj,t. So the universal property of A∞ gives a homomorphism αt : A∞ → A∞ such that αt ◦ ϕj,∞ = ϕj,∞ ◦ αj,t for all j. It is straightforward to check that each αt is an automorphism of A∞ with inverse α−t, and that α : R → Aut A∞ is an action satisfying ϕj,∞ ◦ αj,t = αt ◦ ϕj,∞. An ε/3-argument using that continuous. each αj is strongly continuous and that Sj ϕj,∞(Aj) is dense in A∞ shows that α is strongly For j ∈ N and φ ∈ KMSβ(A∞, α) define hj(φ) := φ ◦ ϕj,∞. Since KMSβ states restrict to KMSβ states on invariant unital subalgebras, hj maps KMSβ(A∞, α) to KMSβ(Aj, αj) for each j. We have hj+1 ◦ ϕj = (φ ◦ ϕj+1,∞) ◦ ϕj = φ ◦ (ϕj+1,∞ ◦ ϕj) = φ ◦ ϕj,∞ = hj, KMSβ(Aj, αj) gives a map h from KMSβ(A∞, α) into KMSβ(Aj, αj) satisfying pj ◦ h = hj, where pj denotes the canonical projection onto and so the universal property of lim ←− lim ←− KMSβ(Aj, βj). We claim that h is the desired affine isomorphism. The map h is obviously affine. To see that h is surjective, fix (φj)∞ and take j ≤ k, a ∈ Aj and b ∈ Ak with ϕj,∞(a) = ϕk,∞(b). Then j=0 ∈ lim ←− (KMSβ(Aj, αj)), 0 = ϕk,∞(b) − ϕj,∞(a) = ϕk,∞(b) − ϕk,∞(ϕk−1 ◦ · · · ◦ ϕj(a)) = ϕk,∞(b − ϕk−1 ◦ · · · ◦ ϕj(a)). Since each ϕj is injective, each ϕj,∞ is injective, and so b = ϕk−1 ◦ · · · ◦ ϕj(a). Now φj(a) = φk(ϕk−1 ◦ · · · ◦ ϕj(a)) = φk(b), and so there is a well-defined linear map φ∞ : S∞ j=0 ϕj,∞(Aj) → C satisfying φ∞(ϕj,∞(a)) = φj(a) for all j ∈ N and a ∈ Aj. Since each ϕj,∞ is isometric and each φj is norm-decreasing, each φ∞ ◦ ϕj,∞ is norm-decreasing, so φ∞ is norm-decreasing. It therefore extends to a norm- decreasing φ∞ : A∞ → C. Since kφ∞k ≥ kφ∞ ◦ ϕjk = kφjk = 1, we see that kφ∞k = 1. Since THE TOEPLITZ NONCOMMUTATIVE SOLENOID AND ITS KMS STATES 5 and therefore a state of A∞. Sj ϕj,∞(cid:0)(Aj)+(cid:1) is dense in (A∞)+ and since each φ∞ ◦ ϕj,∞ = φj is positive, φ∞ is positive, To see that φ∞ is KMS, observe that if a ∈ Aj is αj-analytic, then ϕj,∞(a) is α-analytic. Indeed, since z 7→ ϕj,∞(αj,z(a)) is an analytic extension of t 7→ αt(ϕj,∞(a)), the analytic extension of t 7→ αt(ϕj,∞(a)) is given by αz(ϕj,∞(a)) = ϕj,∞(αj,z(a)). SoSj{ϕj,∞(a) : a ∈ Aj is analytic} is an α-invariant dense subspace of analytic elements in A∞. So it suffices to show that φ∞(cid:0)ϕj,∞(a)ϕk,∞(b)(cid:1) = φ∞(cid:0)ϕk,∞(b)αiβ(ϕj,∞(a))(cid:1) whenever a ∈ Aj and b ∈ Ak are analytic. For this, let l := max{j, k} and observe that a′ := ϕj,l and b′ := ϕk,l(b) are αl-analytic, and so φ∞(ϕj,∞(a)ϕk,∞(b)) = φ∞(ϕl,∞(a′b′) = φl(a′b′) = φl(b′αl,iβ(a′)) Since h(φ∞) = (φ∞ ◦ ϕj,∞)∞ = φ∞(cid:0)ϕl,∞(b′)αiβ(ϕl,∞(a′))(cid:1) = φ∞(cid:0)ϕj,∞(b)αiβ(ϕj,∞(a))(cid:1). j=0, we see that h is surjective. j=0 = (φj)∞ all j ∈ N, which implies that φ and ψ agree on the dense subset S∞ Checking that h is injective is straightforward: if h(φ) = h(ψ), then φ ◦ ϕj,∞ = ψ ◦ ϕj,∞ for j=0 ϕj,∞(Aj), giving φ = ψ. To see that h is continuous, let (φλ)λ∈Λ be a net in KMSβ(A∞, α) converging weak* to φ ∈ KMSβ(A∞, α). Then pj(h(φλ)) = φλ ◦ ϕj,∞ converges weak* to pj(h(φ)) = φ ◦ ϕj,∞ for each j ∈ N. Since the topology on the inverse limit is the initial topology induced by the projections pj, this says that h(φλ) converges weak* to h(φ). Hence h is continuous. (cid:3) 4. C ∗-algebras from rotations on the circle We are interested in topological graphs built from rotations on the circle. We write S := R/Z for the circle, which we frequently identify with [0, 1) under addition modulo 1. For γ ∈ R, let Rγ denote clockwise rotation of the circle S by angle γ. So Rγ(t) = t−γ (mod 1). Each Rγ is a homeomorphism of S, and we denote by Eγ := (S, S, idS, Rγ) the cor- responding topological graph. We denote the Hilbert bimodule X(Eγ) by C(S)γ, its inner prod- uct by h·, ·iγ, and the homomorphism implementing the left action by φγ : C(S) → L(C(S)γ). We can give alternative characterisations of the C ∗-algebras T (Eγ) and O(Eγ). This is certainly not new: the description of O(Eγ) goes back to Pimsner [26, page 193, Example 3]. But we could not locate the exact formulation that we want for the description of T (Eγ) in the literature. Definition 4.1. A Toeplitz pair for Eγ in a C ∗-algebra B is a pair (π, S) consisting of a homomorphism π of C(S) into B, and an isometry S ∈ B satisfying Sπ(f ) = π(f ◦ Rγ)S for all f ∈ C(S). A covariant pair for Eγ is a Toeplitz pair (π, W ) in which W is a unitary. Proposition 4.2. Let γ ∈ R and Eγ = (S, S, idS, Rγ). X(Eγ ), i1 (1) The pair (iγ, sγ) := (i0 X(Eγ )(1)) is a Toeplitz pair for Eγ that generates T (Eγ). Moreover, T (Eγ) is the universal C ∗-algebra generated by a Toeplitz pair for Eγ: if (π, S) is a Toeplitz pair in a C ∗-algebra B, then there is a homomorphism π × S : T (Eγ) → B satisfying (π × S) ◦ iγ = π and (π × S)(sγ) = S. (2) The pair (jγ, wγ) := (j0 X(Eγ )(1)) is a covariant pair for Eγ that generates O(Eγ). Moreover, O(Eγ) is the universal C ∗-algebra generated by a covariant pair for Eγ: if (π, W ) is a covariant pair in a C ∗-algebra B, then there is a homomorphism π × W : O(Eγ) → B satisfying (π × W ) ◦ jγ = π and (π × W )(wγ) = W . X(Eγ ), j1 6 BROWNLOWE, HAWKINS, AND SIMS Proof. We have s∗ f ∈ C(S) we have γsγ = i0 X(Eγ )(h1, 1iγ) = i0 X(Eγ )(1) = 1, and so sγ is an isometry. For each iγ(f ◦ Rγ)sγ = i0 = i1 X(Eγ )(f ◦ Rγ)i1 X(Eγ )(1 · f ) = i1 X(Eγ )(1) = i1 X(Eγ )(1)i0 X(Eγ )((f ◦ Rγ) · 1) X(Eγ )(f ) = sγiγ(f ), and so (iγ, sγ) is a Toeplitz pair. For f ∈ C(S)γ we have i1 (iγ, i1 η(1)) generates the ranges of both i0 X(Eγ ) and i1 X(Eγ ), and hence all of T (Eγ). X(Eγ )(f ) = iγ(f )sγ, so the pair Now suppose B is a unital C ∗-algebra and π : C(S) → B and S ∈ B form a Toeplitz pair (π, S) for Eγ in B. Define ψ : C(S)γ → B by ψ(f ) = π(f )S. We claim that (ψ, π) is a representation of C(S)γ in B. For each f ∈ C(S)γ and g ∈ C(S) we have π(g)ψ(f ) = π(g)π(f )S = π(gf )S = ψ(gf ) = ψ(g · f ) and ψ(f )π(g) = π(f )Sπ(g) = π(f (g ◦ Rγ))S = ψ(f (g ◦ Rγ)) = ψ(f · g). To check that the inner product is preserved, we let f, h ∈ C(S)γ and calculate ψ(f )∗ψ(h) = S∗π(f ∗)π(h)S = S∗π(f ∗ ◦ R−1 = π(f ∗ ◦ R−1 γ )S∗Sπ(h ◦ R−1 γ ◦ Rγ)π(h ◦ R−1 γ ). γ ) = π((f ∗h) ◦ R−1 γ ◦ Rγ)S We have hf, hiγ(z) = f (R−1 hence ψ(f )∗ψ(h) = π(hf, hiγ). This proves the claim. γ (z)) = (f ∗g) ◦ R−1 γ (z))g(R−1 γ (z). So hf, hiγ = (f ∗h) ◦ R−1 γ , and The universal property of T (Eγ) yields a homomorphism ψ × π : T (Eγ) → B satisfying (ψ × π) ◦ i1 X(Eγ ) = ψ and (ψ × π) ◦ i0 X(Eγ ) = π. Let π × S := ψ × π. Then (π × S) ◦ i0 X(Eγ ) = (ψ × π) ◦ i0 X(Eγ ) = π, and (π × S)(sγ) = (ψ × π)(i1 X(Eγ )(1)) = ψ(1) = π(1)S = S. Hence T (Eγ) is the universal C ∗-algebra generated by a Toeplitz pair for Eγ. To prove (2) it suffices to show that the ideal I generated by {(i1 X(Eγ ), i0 X(Eγ ))(1)(ϕγ(f )) − i0 X(Eγ )(f ) : f ∈ C(S)} is the ideal generated by the element sγs∗ γ − 1. We have X(Eγ ) = (i1 sγs∗ γ − 1 = (i1 X(Eγ ), i0 X(Eγ ))(1)(Θ1,1) − i0 X(Eγ ), i0 X(Eγ ))(1)(φη(1)) − i0 X(Eγ )(1) ∈ I, and hence the ideal generated by sγs∗ first note that ϕγ(f ) = Θf,1 for all f ∈ C(S). Then γ − 1 is contained in I. For the reverse containment we (i1 X(Eγ ), i0 X(Eγ ))(1)(ϕη(f )) − i0 X(Eγ ), i0 X(Eγ )(f )i1 X(Eγ )(f )i1 X(Eγ ))(1)(Θf,1) − i0 X(Eγ )(1)∗ − i0 η(f ) X(Eγ )(1)i1 η(f ) X(Eγ )(1)∗ − i0 X(Eγ )(f ) η(f ) = (i1 = i1 = i0 = i0 X(Eγ )(f )(cid:0)sγs∗ γ − 1(cid:1), and the result follows. (cid:3) Remarks 4.3. (1) We saw in the proof of Proposition 4.2 that a Toeplitz pair (π, S) for Eγ gives a representation (ψ, π) of X(Eγ) such that ψ(f ) = π(f )S. We denote the homomorphism (ψ, π)(1) of K(X(Eγ)) by (π, S)(1); so (π, S)(1)(Θf,g) = π(f )SS∗π(g)∗. (2) In [17, Theorem 6.2] Katsura proved a gauge-invariant uniqueness theorem for the Toeplitz algebra of a Hilbert bimodule. Suppose A is a C ∗-algebra, X is a Hilbert A-bimodule, and (ψ, π) is a representation of X in a C ∗-algebra B. The gauge-invariant THE TOEPLITZ NONCOMMUTATIVE SOLENOID AND ITS KMS STATES 7 uniqueness theorem says that ψ × π : T (X) → B is injective if B carries a gauge action, ψ × π intertwines the gauge actions on T (X) and B, and the ideal {a ∈ A : π(a) ∈ (ψ, π)(1)(K(X))} of A is trivial. If (π, S) is a Toeplitz pair for Eγ, then this ideal is {f ∈ C(S) : π(f ) ∈ (π, S)(1)(K(X(Eγ)))}, which we can write as {f ∈ C(S) : π(f ) ∈ span{π(g)SS∗π(h) : g, h ∈ C(S)}}. We denote this ideal by I(π,S). (3) Proposition 4.10 of [17] says that I(iγ ,sγ) = 0. We can give spanning families for T (Eγ) and O(Eγ) using Toeplitz and covariant pairs. Proposition 4.4. Let γ ∈ R and Eγ = (S, S, idS, Rγ). Then T (Eγ) = span{sm γ iγ(f )s∗ γ n : m, n ∈ N, f ∈ C(S)}, and O(Eγ) = span{wm γ iγ(f )s∗ γ γ jγ(f )w∗ γ n : m, n ∈ N, f ∈ C(S)}. n : m, n ∈ N, f ∈ C(S)} contains the generators of T (Eγ), so Proof. The set span{sm it suffices to show that it is a ∗-subalgebra. It is obviously closed under involution; that it is closed under multiplication follows from the calculation sm γ iγ(f )s∗ γ nsp γiγ(g)s∗ γ q if n ≥ p if n < p γ q =(sm =(sm n−piγ(g)s∗ γ iγ(f )s∗ γ γ q iγ(g)s∗ γ iγ(f )sp−n sm γ γ iγ(f (g ◦ R−(n−p) sm+p−n γ γ jγ(f )w∗ γ γ n−p+q ))s∗ γ iγ((f ◦ R−(p−n) )g)s∗ γ n under the quotient map T (Eγ) → O(Eγ), we (cid:3) if n ≥ p if n < p. γ q Since each sm have O(Eγ) = span{wm γ iγ(f )s∗ γ n is mapped to wm γ jγ(f )w∗ γ n : m, n ∈ N, f ∈ C(S)}. 5. An alternative description of the noncommutative solenoid Throughout the rest of this paper we fix a natural number N ≥ 2. In [22], given a sequence n=1 in S = R/Z such that N 2θn+1 = θn for all n, Latr´emoli`ere and Packer define θ as a twisted group C ∗-algebra involving the N-adic ratio- θ . We will take this θ = (θn)∞ the noncommutative solenoid AS nals. In [22, Theorem 3.7] they give an equivalent characterisation of AS characterisation as our definition. We recall it now. Let ΞN := {(θn)∞ n=0 : θn ∈ S and N 2θn+1 = θn for each n}. Recall that for γ ∈ S the rotation algebra Aγ is the universal C ∗-algebra generated by unitaries Uγ and Vγ satisfying UγVγ = e2πiγVγUγ. Definition 5.1. Let θ = (θn)∞ homomorphism satisfying n=0 ∈ ΞN , and for each n ∈ N let ϕn : Aθn → Aθn+1 be the The noncommutative solenoid AS θ ϕn(Uθn) = U N θn+1 and ϕn(Vθn) = V N θn+1. (Aθn, ϕn). is the direct limit lim −→ Remark 5.2. We have taken a slightly different point of view to [22] in describing AS θ . In [22], Latr´emoli`ere and Packer consider collections of (θn) such that Nθn+1 − θn ∈ Z, and take the direct limit lim −→ Aθ2n, with intertwining maps going from Aθ2n to Aθ2n+2. We now give an alternative characterisation of the noncommutative solenoid using topological graphs built from rotations of the circle as discussed in Section 4. Notation 5.3. We denote by ι : S → T the homeomorphism t 7→ e2πit, and by pN : S → S the function t 7→ Nt. 8 BROWNLOWE, HAWKINS, AND SIMS Proposition 5.4. Let N ≥ 2, and θ = (θn)∞ homomorphism τn : O(Eθn) → O(Eθn+1) satisfying n=0 ∈ ΞN . For each n ∈ N there is an injective τn(jθn(f )) = jθn+1(f ◦ pN ) and τn(wθn) = wN θn+1, for all f ∈ C(S). Moreover lim −→ (O(Eθn), τn) ∼= AS θ . We will prove the existence of the injective homomorphisms τn using the following result. Lemma 5.5. Let N ∈ N with N ≥ 2, and take γ, η ∈ S with N 2η − γ ∈ Z. Then there is an injective homomorphism ψ : T (Eγ) → T (Eη) satisfying ψ(iγ(f )) = iη(f ◦ pN ) and ψ(sγ) = sN η , for all f ∈ C(S). The map ψ descends to an injective homomorphism τ : O(Eγ) → O(Eη) satisfying τ (jγ(f )) = jη(f ◦ pN ) and τ (wγ) = wN η for all f ∈ C(S). Proof. Consider π : C(S) → T (Eη) given by π(f ) = iη(f ◦ pN ) and let S := sN η . Since Rγ ◦ pN = RN 2η ◦ pN = RN 2 η ◦ pN = pN ◦ RN η , we have π(f ◦ Rγ)S = iη(f ◦ Rγ ◦ pN )sN η = iη(f ◦ pN ◦ RN η )sN η = sN η iη(f ◦ pN ) = Sπ(f ). So (π, S) is a Toeplitz pair for Eγ. The universal property of T (Eγ) now gives a homomorphism ψ : T (Eγ) → T (Eη) satisfying ψn(iγ(f )) = iη(f ◦ pN ) for all f ∈ C(S), and ψ(sγ) = sN η . To see that ψ is injective, we aim to apply the gauge-invariant uniqueness theorem discussed 6= 0. To see this, suppose that in Remarks 4.3. We claim that I(π,S) 0 6= f ∈ I(π,S). Fix ǫ > 0, and choose gi, hi ∈ C(S) with 6= 0 =⇒ I(iη,sη) π(f ) − (cid:13)(cid:13)(cid:13) k Xi=1 k Xi=1 π(gi)SS∗π(hi)(cid:13)(cid:13)(cid:13) < ǫ. iη(gi ◦ pN )sN η s∗ η < ǫ. N iη(hi ◦ pN )(cid:13)(cid:13)(cid:13) iη(f ◦ pN ) − (cid:13)(cid:13)(cid:13) For every function g ∈ C(S) we have iη(g ◦ RN −1 −η ) = s∗ η N −1sN −1 η iη(g ◦ RN −1 −η ) = s∗ η N −1iη(g)sN −1 η . Hence iη(f ◦ pN ◦ RN −1 −η ) − iη(gi ◦ pN ◦ RN −1 −η )sηs∗ ηiη(hi ◦ pN ◦ RN −1 So (cid:13)(cid:13)(cid:13) k s∗ η Xi=1 =(cid:13)(cid:13)(cid:13) ≤(cid:13)(cid:13)(cid:13) N −1iη(f ◦ pN )sη N −1 − k s∗ η k Xi=1 iη(f ◦ pN ) − iη(gi ◦ pN )sN η s∗ η Xi=1 N −1iη(gi ◦ pN )sN η s∗ η N iη(hi ◦ pN )sη N −1(cid:13)(cid:13)(cid:13) −η )(cid:13)(cid:13)(cid:13) < ǫ. N iη(hi ◦ pN )(cid:13)(cid:13)(cid:13) It follows that iη(f ◦ pN ◦ RN −1 This proves the claim. −η ) ∈ (iη, sη)(1)(K(X(Eη))), and hence that f ◦ pN ◦ RN −1 −η ∈ I(iη,sη). By Remarks 4.3(3), I(iη,sη) = 0, so the claim gives I(π,S) = 0. We have ψ(T (Eγ)) ⊆ : f ∈ C(S), a, b ∈ N}. Hence the gauge action ρη of T on T (Eγ) satis- z+e2πi/N ◦ ψ for all z ∈ T. So there is an action ρη of T on ψ(T (Eγ)) such that η iη(f )s∗bN z ◦ ψ = ρη span{saN fies ρη η Xi=1 THE TOEPLITZ NONCOMMUTATIVE SOLENOID AND ITS KMS STATES 9 ρη e2πit ◦ ψ = ρeeπit/N for all t ∈ R. In particular, e2πit/N (saN = e2πit(a−b)saN ρη e2πit ◦ ψ(sa γiη(f )s∗b γ ) = ρη η iη(f )s∗bN ) η iη(f )s∗bN η η = ψ ◦ ργ e2πit(saN η iη(f )s∗bN η ). So continuity and linearity gives ρη theorem [17, Theorem 6.2] shows that ψ is injective. e2πit = ψ ◦ ργ e2πit. Hence the gauge-invariant uniqueness To see that ψ descends to the desired injective homomorphism τ : O(Eγ) → O(Eη), it suffices to show that the image under ψ of the kernel of the quotient map T (Eγ) → O(Eγ) is contained in the kernel of T (Eη) → O(Eη). For this, it suffices to show that ψ(1 − sγs∗ γ) is in the ideal generated by 1 − sηs∗ η, which it is because N ψ(1 − sγs∗ γ) = 1 − sN η s∗ η N = sN −i η (1 − sηsη)s∗ η N −i. (cid:3) Proof of Proposition 5.4. For each n ∈ N, Lemma 5.5 applied to γ = θn and η = θn+1 gives the desired injective homomorphism τn. Proposition 4.2 says that each O(Eη) is the crossed product C(S) ⋊ Z for the automorphism f 7→ f ◦ Rη of C(S), which is the rotation algebra Aη (see [7, Example VIII.1.1] for details). So for each n ∈ N there is an isomorphism from O(Eθn) to Aθn carrying jθn(ι) to Uθn and wθn to Vθn. Since each τn satisfies τn(jθn(ι)) = jθn+1(ι ◦ pN ) = jθn+1(ι)N and τn(wθn) = wN θn+1, the diagrams O(Eθn) ∼= commute. Hence lim −→ Aθn (O(Eθn), τn) ∼= AS θ . τn ϕn O(Eθn+1) ∼= Aθn+1 (cid:3) Remark 5.6. In [18, Section 2], Katsura studies factor maps between topological-graph C ∗- algebras, and the C ∗-homomorphisms that they induce. He shows that the projective limit of a sequence (En) of topological graphs under factor maps is itself a topological graph. He then proves that the C ∗-algebra O(lim En) of this topological graph is isomorphic to the direct limit ←− O(En) of the C ∗-algebras of the En under the homomorphisms induced by the factor maps. lim −→ So it is natural to ask whether the maps τn : O(Eθn) → O(Eθn+1) correspond to factor maps. This is not the case: as observed on page 88 of [11], there is no factor map from Eθn+1 → Eθn that induces the homomorphism of C ∗-algebras described in Lemma 5.5. 6. The Toeplitz noncommutative solenoid and its KMS structure In this section we introduce our Toeplitz noncommutative solenoids T S θ . We introduce a and apply Proposition 3.1 to begin to study its KMS structure. n=0 ∈ ΞN , Lemma 5.5 gives a sequence of injective homomorphisms ψn : natural dynamics on T S θ Given θ = (θn)∞ T (Eθn) → T (Eθn+1) satisfying ψn(iθn(f )) = iθn+1(f ◦ pN ) and ψn(sθn) = sN θn+1, for all f ∈ C(S). Definition 6.1. We define T S := lim θ −→ solenoid. We write ψn,∞ : T (Eθn) → T S θ for all n. (T (Eθn), ψn) and call it the Toeplitz noncommutative for the canonical inclusions, so that ψn,∞ = ψn+1,∞◦ψn The following lemma indicates why it is sensible to regard T S θ as a natural Toeplitz extension of the noncommutative solenoid. 10 BROWNLOWE, HAWKINS, AND SIMS Lemma 6.2. In the notation established in Proposition 5.4, there is a surjective homomorphism q : T S such that q(ψn,∞(iθn(f ))) = τn,∞(jθn(f )) and q(ψn,∞(sθn)) = τn,∞(wθn) for all n ∈ N and all f ∈ C(S). Moreover, ker(q) is generated as an ideal by ψ1,∞(iθ1(1) − sθ1s∗ θ → AS θ θ1). Proof. For the first statement observe that the canonical homomorphisms qn : T (Eθn) → O(Eθn) intertwine the ψn with the τn. For the second statement, let I be the ideal of T S θ generated by ψ1,∞(iθ1(1) − sθ1s∗ θ1), we have I ⊆ ker(q). For the reverse inclusion, note that for n ≥ 1, θn s∗nN θn s∗(nN −1) θn θn) ≤ ψn,∞(ψ1,n(iθ1(1) − sθ1s∗ θn ≥ iθn(1) − sθns∗ θn, θ1)) = ψ1,∞(iθ1(1) − sθ1s∗ θ1), which be- θ1) = iθn(1 ◦ ιN n) − snN = iθn(1) − sθn(snN −1 θ1). Since ker(q) clearly contains ψ1,∞(iθ1(1) − sθ1s∗ ψ1,n(iθ1(1) − sθ1s∗ )s∗ θn θn) ∈ I. Since ker(q) = Sn ker(q) ∩ ψn,∞(T (Eθn)) = Sn ψn,∞(ker(qn)), it therefore suffices to show that each ker(qn) is generated by iθn(1) − sθns∗ which follows from Proposition 4.2. θn, (cid:3) so each ψn,∞(iθn(1) − sθns∗ longs to I. Thus ψn,∞(iθn(1) − sθns∗ Proposition 6.3. There is a strongly continuous action α : R → Aut T S θ θj iθj (f )s∗n θj )) = eit(m−n)/N j θj iθj (f )s∗n αt(ψj,∞(sm ψj,∞(sm θj ), satisfying (6.1) for each j, m, n ∈ N and f ∈ C(S). This α descends to a strongly continuous action, also written α, on the noncommutative solenoid AS θ . Proof. For each j ∈ N we denote by ρ the gauge action on T (Eθj ), and by ρj the strongly continuous action t 7→ ρeit/Nj of R on T (Eθj ); so ρj,t ◦ iθj = iθj and ρj,t(sθj ) = eit/N j sθj for each t ∈ R. For each j ∈ N and t ∈ R we have ρj+1,t ◦ ψj(sm θj iθj (f )s∗n θj ) = eit(N m−N n)/N j+1 sN m θj+1iθj+1(f )s∗N n = eit(m−n)/N j θj+1iθj+1(f )s∗N n sN m θj+1 θj+1 = ψj ◦ ρj,t(sm θj iθj (f )s∗n θj ). Hence ρj+1,t ◦ ψj = ψj ◦ ρj,t, and Proposition 3.1 applied to each (Aj, αj) = (T (Eθj ), ρj) gives the desired action α : R → Aut T S θ . For the final statement, observe that the αt all fix ψ1,∞(iθ1(1) − sθ1s∗ that it generates invariant; so they descend to AS θ by Lemma 6.2. θ1), and so leave the ideal (cid:3) Remark 6.4. The actions on graph C ∗-algebras and their analogues studied in, for example, [5, 8, 1, 12] are lifts of circle actions, and so are periodic in the sense that αt = αt+2π for all t. By contrast, while the action α of the preceding proposition restricts to a periodic action on each approximating subalgebra ψj,∞(T (Eθj )), it is itself not periodic: αt = αs =⇒ t = s. We now wish to study the KMS structure of the Toeplitz noncommutative solenoid T S θ under the dynamics α of Proposition 6.3. (T , κ) → C(S ) ⊗ C(S ). (S, pN ) Remark 6.5. The case θ = 0 = (0, 0, . . . ) is relatively easy to analyse. Let S = lim ←− denote the classical solenoid, and T the Toeplitz algebra. Write s for the isometry generating ∼= C(S ) ⊗ lim T , and κ : T → T for the homomorphism given by κ(s) = sN . Then T S (T , κ). −→ This isomorphism intertwines the quotient map q : T S 0 → AS 0 with the canonical quotient map id ⊗q : C(S ) ⊗ lim It also intertwines α with 1 ⊗ α where −→ αt(κj,∞(s)) = eit/N j κj,∞(s). That is, α is equivariant over κj,∞ with an action αj on T that is a rescaling of the gauge dynamics studied in [12]. Theorems 3.1 and 4.3 of [12] imply that (T , αj) has a unique KMSβ state for every β ≥ 0 and has no KMSβ states for β < 0, and that the KMS0 state is the only one that factors through C(S). So Proposition 3.1 implies that (lim (T , κ), α) has a unique KMSβ state φβ for each β ≥ 0 and has no KMSβ states for −→ β < 0, and that the KMS0 state is the only one that factors through C(S ). Hence the map ψ 7→ ψ ⊗ φβ determines an affine isomorphism of the state space of C(S ) onto KMSβ(T S 0 , α) 0 THE TOEPLITZ NONCOMMUTATIVE SOLENOID AND ITS KMS STATES 11 for each β ≥ 0, there are no KMSβ states for β < 0, and the KMS0 states are the only ones that factor through AS 0 . In light of Remark 6.5, we will from now on consider only those θ ∈ ΞN such that θj 6= 0 for n=j) for any some j. Since θj 6= 0 implies θj+1 6= 0, and since lim −→ j, we may therefore assume henceforth that θj 6= 0 for all j. n=1) = lim −→ ((Aθn, ϕn)∞ ((Aθn, ϕn)∞ Our main result is the following. Theorem 6.6. Take N ∈ {2, 3, . . . }, take θ = (θj)∞ θj 6= 0 for all j. Then KMSβ(T S θ measures on the solenoid S := lim ←− a free and transitive action of S on the extreme boundary of KMSβ(T S θ KMS0-state on T S θ are no KMSβ states for β < 0. j=0 ∈ ΞN , and take β ∈ (0, ∞). Suppose that , α) is isomorphic to the Choquet simplex of Borel probability that induces , α). There is a unique θ . There for α, and this is the only KMS state for α that factors through AS (S, pN ), and there is an action λ of S on T S θ The first step in proving Theorem 6.6 is to combine the results of [1] on KMS states of local in homeomorphism C ∗-algebras with Proposition 3.1 to characterise the KMS states of T S θ terms of subinvariant probability measures on the circle. We start with some notation. It is helpful to recall what the results of [1] say in the context of the topological graphs Eγ. Recall that ρ denotes the gauge action on T (Eγ); we also use ρ for the lift of the gauge action to an action of R on T (Eγ). Combining Proposition 4.2 and Theorem 5.1 of [1], we see that for each Borel probability measure µ on S that is subinvariant in the sense that µ(Rγ(U)) ≤ eβµ(U) for every Borel U ⊆ S, there is a KMSβ-state φµ ∈ KMSβ(T (Eγ), ρ) satisfying φµ(sa γiγ(f )s∗b γ ) = δa,be−aβZS f dµ; (6.2) and moreover, the map µ 7→ φµ is an affine isomorphism of the simplex of subinvariant Borel probability measures on S to KMSβ(T (Eγ), ρ). Definition 6.7. Fix r, s ∈ [0, ∞), and γ ∈ S. Let M(S) denote the set of Borel probability measures on S. We define Msub(s, γ) := {m ∈ M(S) : m(Rγ(U)) ≤ esm(U) for all Borel U ⊆ S} and Ωr sub := {m ∈ M(S) : m(Rt(U)) ≤ ertm(U) for all t ∈ [0, ∞) and Borel U ⊆ S}. (6.3) Notation 6.8. For the rest of the section we fix θ = (θj)∞ and β ∈ [0, ∞). We define j=0 ∈ ΞN such that θj 6= 0 for all j, rj := β/N jθj for all j ∈ N. Theorem 6.9. Take N ∈ N with N ≥ 2, θ = (θj)∞ θj 6= 0 for all j. Then there is an affine injection j=0 ∈ ΞN , and β ∈ [0, ∞). Suppose that ω : KMSβ(T S θ , α) → lim ←− (Ωrj sub, m 7→ m ◦ p−1 N ) such that for each φ ∈ KMSβ(T S θ φ ◦ ψj,∞(sa θj iθj (f )s∗b θj ) = δa,be−aβ/N jZS , α) and j ∈ N. If β > 0, then ω is an isomorphism. f dω(φ)j (6.4) Since the dynamics α on T S θ Write ρ for the gauge action on T (Eθj ), and ρj for the action t 7→ ρeit/Nj of R on T (Eθj ). is induced by the ρj, Proposition 3.1 yields an affine isomorphism , α) ∼= lim ←− (KMSβ(T (Eθj ), ρj), φ 7→ φ ◦ ψj−1). KMSβ(T S θ For each j ∈ N and t ∈ R we have ρj,t = ρt/N j , so KMSβ(T (Eθj ), ρj) = KMSβ/N j (T (Eθj ), ρ). For β > 0, the KMSβ simplex of each (T (Eθj ), ρ) is well understood by the results of [1] (see the discussion preceding (6.2)), and we use these results to prove the following. 12 BROWNLOWE, HAWKINS, AND SIMS Proposition 6.10. With the hypotheses of Theorem 6.9, there is an affine injection τ : lim ←− (KMSβ/N j (T (Eθj ), ρ), φ 7→ φ ◦ ψj−1) → lim ←− (Msub(β/N j, θj), m 7→ m ◦ p−1 N ) such that φj = φτ ((φk)∞ isomorphism. k=0)j , as defined at (6.2), for all (φk)∞ k=0 and j ∈ N. If β > 0 then τ is an Throughout the rest of this section we suppress intertwining maps in projective limits. Proof of Proposition 6.10. We first claim that for each j ∈ N there is an affine injection τj of KMSβ/N j (T (Eθj ), ρ) onto Msub(β/N j, θj) satisfying φ(iθj (f )) =ZS f d(τj(φ)) for all φ ∈ KMSβ/N j (T (Eθj ), ρ) and f ∈ C(S), and that for β > 0, this τj is an isomorphism. The statement for β > 0 follows directly from [1, Theorem 5.1] (see (6.2). To prove the claim for β = 0, recall that the KMS0 states on T (Eθj ) for ρ are the ρ-invariant traces. Let (iθj , sθj ) be the universal Toeplitz pair for Eθj . If φ is a KMS0-state, then (with the convention that sn := s∗n for n < 0), φ(sn θj iθj (f )s∗m θj ) = φ(iθj (f )s∗m θj sn θj ) = φ(iθj (f )sn−m θj ) =ZS φ(ρt(iθj (f )sn−m θj )) dµ(t) = δm,nφ(iθj (f )). (6.5) So, by the Riesz -- Markov -- Kakutani representation theorem [27, Theorem 2.14], there exists a Borel probability measure mφ on S such that φ(sn we have θj ) =RS f (t) dmφ(t). For f ∈ C(S)+, θj iθj (f )s∗m φ(iθj (f )) ≥ φ(cid:0)iθj(cid:0)pf(cid:1)sθj s∗ θj iθj(cid:0)pf(cid:1)(cid:1) = φ(cid:0)s∗ θj iθj(cid:0)pf(cid:1)iθj(cid:0)pf(cid:1)sθj(cid:1) = φ(s∗ θj iθj (f )sθj ) = φ(iθj (f ◦ R−θj ). Hence mφ(Rθj (U)) ≤ mφ(U) for all Borel U. So φ 7→ mφ is an affine map from KMS0(T (Eθj ), ρ) and (6.5) shows that it is injective. This completes the proof of the claim. For each j ∈ N let pj be the projection from lim ←− Msub(β/N j, θj) to Msub(β/N j, θj). Fix an element (φj)∞ KMSβ/N j (T (Eθj ), ρ) to KMSβ/N j (T (Eθj ), ρ), j=0 of and πj the projection from lim ←− lim ←− KMSβ/N j (T (Eθj ), ρ). For each k ≥ 1 and f ∈ C(S) we have ZS f d(τk−1(φk−1)) = φk−1(iθk−1(f )) = φk(ψk−1(iθk−1(f ))) = φk(iθk (f ◦ pN )) =ZS (f ◦ pN ) d(τk(φk)) =ZS f d(τk(φk) ◦ p−1 N ), and hence τk−1(φk−1) = τk(φk) ◦ p−1 N . It follows that τk−1 ◦ pk−1((φj)∞ j=0) = τ (φk−1) = τk(φk) ◦ p−1 N = τk ◦ pk((φj)∞ j=0) ◦ p−1 N , for each k ≥ 1. The universal property of lim ←− Msub(β/N j, θj) yields a map τ : lim ←− KMSβ/N j (T (Eθj ), ρ) → lim ←− Msub(β/N j, θj), range(τk) = lim ←− range(τk), satisfying πk ◦ τ = τk ◦ pk for each k ∈ N. For β > 0, we whose image is lim ←− Msub(β/N j, θj), and otherwise it is a compact affine subset, so it now have lim ←− suffices to prove that τ is an affine isomorphism onto its range. Since τ is an injective map from a compact space to a Hausdorff space, it therefore suffices to show that it is affine and continuous. THE TOEPLITZ NONCOMMUTATIVE SOLENOID AND ITS KMS STATES 13 j=0 is a convex combination in lim ←− KMSβ/N j (T (Eθj ), ρ). For each k ∈ N i=1 λi(φi and f ∈ C(S) we have j)∞ Suppose Pq ZS f d(cid:16)τk(cid:16) q Xi=1 λiφi q q λiφi k(cid:17)(cid:17) =(cid:16) Xi=1 λiZS Xi=1 = k k(cid:17)(iθk(f )) = f d(τk(φi Xi=1 k)) =ZS λiφi k(iθk(f )) f d(cid:16) q Xi=1 λiτk(φi k)(cid:17). i=1 λiφi i=1 λiτk(φi k) by the Riesz -- Markov -- Kakutani representation theorem So τk(cid:16)Pq k(cid:17) = Pq [27, Theorem 2.14], and it follows that τ is affine. j )∞ j=0)λ∈Λ be a net in lim ←− j=0)λ∈Λ) = (φλ Straightforward arguments using that πk ◦ τ = τk ◦ pk for each k ∈ N, and that each τk is injective, show that τ is injective. We just need to show that τ is continuous. Let ((φλ j=0. Then pk(((φλ j=0) = φk for each k ∈ N. Since τk is continuous and πk ◦τ = τk ◦pk for each k ∈ N, we have that πk(τ (((φλ k)λ∈Λ) converges weak* to τk(φk) = πk(τ ((φj)∞ j=0)λ∈Λ) converges in the initial topology to τ ((φj)∞ j=0). So τ is continuous. (cid:3) KMSβ/N j (T (Eθj ), ρ) converging in the initial topology to (φj)∞ k)λ∈Λ converges weak* to pk((φj)∞ j=0)). Hence τ (((φλ j=0)λ∈Λ)) = τk((φλ j )∞ j )∞ j )∞ Remark 6.11. Fix β > 0. Let h be the affine isomorphism of Proposition 3.1 and let τ be the affine isomorphism of Proposition 6.10. Setting ω := τ ◦ h gives an affine isomorphism , α) → lim ←− satisfying φ ◦ ψj,∞ = φω(φ)j for each φ ∈ KMSβ(T S θ Msub(β/N j, θj) ∼= lim now suffices to show that lim ←− ←− ω : KMSβ(T S θ Msub(β/N j, θj) , α) and j ∈ N. So to prove Theorem 6.9 it Ωrj sub. Fix (mj)∞ j=0 ∈ lim ←− sub. Taking t = θj in the definition of Ωrj Ωrj that Ωrj reverse containment. We start with a lemma. sub ⊆ Msub(β/N j, θj). Hence lim ←− Ωrj sub is contained in lim ←− sub (see Definition 6.7) shows Msub(β/N j, θj). So we need the Lemma 6.12. Let m be a Borel probability measure on S, and fix γ ∈ (0, 1), s ∈ [0, ∞) and N ∈ N with N ≥ 2. Suppose that m(Rγ/N k (U)) ≤ es/N km(U) for every k ∈ N and every Borel set U ⊆ S. Then m ∈ Ωs/γ sub. Proof. We need to show that m(Rt(U)) ≤ e(s/γ)tm(U) for all t ≥ 0 and Borel U ⊆ S; or equivalently, that m(Rtγ(U)) ≤ estm(U) for all t ≥ 0 and Borel U ⊆ S. By the Riesz -- Markov -- Kakutani representation theorem [27, Theorem 2.14], it suffices to show that ZS f ◦ R−tγ dm ≤ estZS f dm (6.6) for every t ≥ 0 and every f ∈ C(S)+. Furthermore, if (6.6) holds whenever 0 ≤ t ≤ 1, then for arbitrary T ∈ [0, ∞), we can iterate (6.6) ⌈T ⌉ times for t = T ⌈T ⌉ to obtain (6.6) for T ; so it suffices to establish (6.6) for t ∈ [0, 1]. Fix t ∈ [0, 1] and f ∈ C(S). Write ∞ t = ai N i Xi=1 ai N i . So tn is a monotone increasing sequence in [0, 1] converging to t. Since the action s 7→ Rs of R on S by rotations where each ai ∈ {0, . . . , N − 1}. For each n ∈ N, let tn := Pn is uniformly continuous, we have f ◦ R−tnγ → f ◦ R−tγ in (cid:0)C(S), k · k∞(cid:1). Since m is a Borel probability measure, the functional f 7→RS f dm is a state, and so i=1 ZS f ◦ R−tnγ dm →ZS f ◦ R−tγ dm. 14 BROWNLOWE, HAWKINS, AND SIMS So it suffices to show that eachRS f ◦ R−tnγ ≤ estRT f dm. So fix n ∈ N. Let K :=Pn N n . By hypothesis, for every Borel U, we have so that t > tn = K i=1 aiN n−i, m(R Kγ Nn (U)) ≤ e s Nn m(R (K−1)γ Nn (U)) ≤ · · · ≤ e sK Nn m(U) ≤ estm(U), and it follows that RS f ◦ R−tnγ ≤ estRS f dm as required. Msub(β/N j, θj) Proof of Theorem 6.9. As described in Remark 6.11, it suffices to show that lim ←− Ωrj sub. For each γ ∈ S we have pN ◦ Rγ = RN γ ◦ pN , which implies that is contained in lim ←− p−1 N (RN γ(U)) = Rγ(p−1 N (U)) for all Borel U ⊆ S. An iterative argument shows that (cid:3) p−k N (RN kγ(U)) = Rγ(p−k N (U)) for all Borel U ⊆ S and k ∈ N. (6.7) Fix (mj)∞ sub are the same, it suffices to show that mj ∈ Ωrj Ωrj Msub(β/N j, θj). Since the connecting maps in lim ←− Msub(β/N j, θj) and sub for each j ∈ N. Fix j ∈ N. For each j=0 ∈ lim ←− lim ←− k ∈ N we have N 2kθj+k = θk and mj+k ◦ p−k N = mj. These identities and (6.7) give mj(Rθj /N k (U)) = mj(RN kθj+k(U)) = mj+k(p−k N (RN kθj+k (U))) N (U)) = eβ/N j+k for every Borel U ⊆ S. So Lemma 6.12 with γ = θj and s = β/N j gives mj ∈ Ωrj N (U))) ≤ eβ/N j+k = mj+k(Rθj+k(p−k mj+k(p−k mj(U), sub. (cid:3) 7. Subinvariant measures on S Throughout the section we fix r ∈ [0, ∞) and denote Lebesgue measure on S by µ. The sub of (6.3). Define main result of this section gives a concrete description of the simplex Ωr Wr : S → [0, ∞) by For each Borel U ⊆ S, define r 1 − e−r(cid:17)e−rt. Wr(t) =(cid:16) mr(U) :=ZU Wr(t) dt. (7.1) This defines a Borel probability measure mr on S. Theorem 7.1. The space Ωr r = 0, then mr = µ and Ωr sub = {µ}. sub is the weak∗-closed convex hull conv{mr ◦ Rs : 0 ≤ s < 1}. If We need a number of results to prove this theorem. sub and n ∈ N. For 0 ≤ j < 2n, let U n j = [j/2n, (j + 1)/2n) ⊆ S, and Lemma 7.2. Let m ∈ Ωr let vn j be the vector 1 − e−r/2n vn j := 1 − e−r (cid:0)e−(2n−j)r/2n , . . . , e−(2n−1)r/2n , 1, e−r/2n , e−2r/2n , . . . , e−(2n−(j+1))r/2n(cid:1) ∈ R2n . (7.2) 0 ), m(U n 1 ), . . . , m(U n Then (cid:0)m(U n Proof. Let x = (x0, x1, . . . , x2n−1) be the vector (cid:0)m(U n 2n−1)(cid:1) ∈ conv{vn j < 2n we have j : 0 ≤ j < 2n}. 0 ), m(U n 1 ), . . . , m(U n 2n−1)(cid:1). For each 0 ≤ xj = m(U n j ) = m(R2−n(U n j+1)) ≤ er/2n m(U n j+1) = er/2n xj+1, where addition in indices is modulo 2n. Let C 2n denote the graph with vertices Z/2nZ and edges {ej : j ∈ Z/2nZ} with s(ej) = j and r(ej) = j + 1 (mod 2n), and let AC2n denote the adjacency matrix of C 2n. Then x satisfies AC2n x ≤ er/2nx. So x is subinvariant for AC2n in the THE TOEPLITZ NONCOMMUTATIVE SOLENOID AND ITS KMS STATES 15 sense of [12, Theorem 3.1], and is a probability measure because m is. By [12, Theorem 3.1(a)], there is a vector y ∈ [1, ∞)Z/2nZ such that yj = Xµ∈(C2n )∗,s(µ)=j e−r/2nµ = ∞ Xk=0 e−kr/2n =(cid:0)1 − e−r/2n(cid:1)−1 for j ∈ Z/2nZ. For 0 ≤ j < 2n, define ǫj ∈ [0, ∞)Z/2nZ by ǫj(k) =(1 − e−r/2n 0 if k = j otherwise. We have (I − e−r/2n AC2n )vn j , . . . , e−(2n−1)r/2n , 1, e−r/2n AC2n )(cid:0)e−(2n−j)r/2n , . . . , e−(2n−1)r/2n , 1, e−r/2n , . . . , e−(2n−(j+1))r/2n(cid:1) , 1, e−r/2n , . . . , e−(2n−1)r/2n , . . . , e−(2n−(j+1))r/2n(cid:1) , . . . , e−(2n−(j+2))r/2n(cid:1)(cid:17) 1 − e−r/2n 1 − e−r (I − e−r/2n 1 − e−r/2n 1 − e−r (cid:16)(cid:0)e−(2n−j)r/2n = = = 1 − e−r/2n − e−r/2n(cid:0)e−(2n−(j+1))r/2n 1 − e−r (cid:0)0, . . . , 0, 1 − e−r, 0, . . . , 0(cid:1) = (1 − e−r/2n = ǫj. )(cid:0)0, . . . , 0, 1, 0, . . . , 0(cid:1) j = (I − e−r/2nAC2n )−1ǫj. Since the ǫj are the extreme points of the simplex {ǫ : ǫ · y = 1}, So vn it follows from [12, Theorem 3.1(c)] that the vn j are the extreme points of the simplex of subinvariant probability measures on Z/2nZ. Since x is a subinvariant probability measure, it follows that it is a convex combination of the vn j . (cid:3) We now approximate mr by convex combinations of restrictions of Lebesgue measure. Lemma 7.3. For n ∈ N and j ∈ Z/2nZ, let U n simple function j = [j/2n, (j + 1)/2n) ⊆ S, and let Wn,r be the Wn,r = 2n−1 Xj=0 2n(vn 0 )j1U n j . Proof. Fix n ∈ N and 0 ≤ j < 2−n. Then the average value of Wr over the interval U n j Let mn,r be the measure mn,r(U) = RU Wn,r(t) dµ(t) for Borel U ⊆ S. Then limn→∞(cid:13)(cid:13)mr − mn,r(cid:13)(cid:13)1 = 0. 2nZU n Wr(t) dµ(t) = 2nZ (j+1)/2n 1 − e−r(cid:17)e−rt dµ(t) = 2nh(cid:16) −1 (cid:16) j/2n is r j =(cid:16) −2n 1 − e−r(cid:17)(cid:16)e−(j+1)r/2n − e−jr/2n(cid:17) = 2n(cid:16) 1 − e−r/2n j/2n 1 − e−r(cid:17)e−rti(j+1)/2n 1 − e−r (cid:17)e−jr/2n = 2n(vn 0 )j, the constant value of Wn,r on U n that there exists cn j ∈ (j/2n, (j + 1)/2n) such that Wr(cn j . The Mean Value Theorem -- applied toR Wr(t) dµ(t) -- implies j ) = Wn,r(cn Fix ǫ > 0. The function Wr is uniformly continuous on [0, 1), and so there exists N ∈ N such that Wr(s) − Wr(t) < ǫ whenever s, t ∈ [0, 1) satisfy s − t < 2−N . In particular, for n ≥ N j ). 16 BROWNLOWE, HAWKINS, AND SIMS and 0 ≤ j < 2n, the point cn j of the preceding paragraph satisfies sup{Wr(t) − Wn,r(t) : j/2n ≤ t < (j + 1)/2n} = sup{Wr(t) − Wn,r(cn = sup{Wr(t) − Wr(cn j ) : j/2n ≤ t < (j + 1)/2n} j ) : j/2n ≤ t < (j + 1)/2n} ≤ ǫ. So for n ≥ N, Wr(t) − Wn,r(t) dµ(t) (cid:13)(cid:13)mr − mn,r(cid:13)(cid:13)1 =Z 1 0 2n−1 = Xj=0 Z (j+1)/2n and hence limn→∞(cid:13)(cid:13)mr − mn,r(cid:13)(cid:13)1 = 0. Corollary 7.4. Given a sequence (λn)∞ all n, we have j/2n Wr(t) − Wn,r(t) dµ(t) ≤ ǫ dµ(t) = ǫ, n=1 of vectors λn ∈ [0, 1]2n j=0 λn j = 1 for 2n−1 j/2n Xj=0 Z (j+1)/2n satisfying P2n−1 = 0. (cid:3) (cid:3) λn j (mr ◦ Rj/2n) − 2n−1 Xj=0 2n−1 Xj=0 lim n→∞(cid:13)(cid:13)(cid:13) Proof. The triangle inequality gives λn j (mr ◦ Rj/2n) − 2n−1 Xj=0 (cid:13)(cid:13)(cid:13) 2n−1 Xj=0 λn j (mn,r ◦ Rj/2n)(cid:13)(cid:13)(cid:13)1 and so the result follows from Lemma 7.3. λn 2n−1 j (mn,r ◦ Rj/2n)(cid:13)(cid:13)(cid:13)1 Xj=0 =(cid:13)(cid:13)mr − mn,r(cid:13)(cid:13)1, λn ≤ j(cid:13)(cid:13)mr ◦ Rj/2n − mn,r ◦ Rj/2n(cid:13)(cid:13)(cid:13)1 Proof of Theorem 7.1. We first have to show that each mr ◦ Rs ∈ Ωr sub, it suffices to prove that Wr(Rt(t0)) ≤ ertWr(t0) for all t0 ∈ S and t ∈ [0, ∞). Fix such a t0 and t, and write t0 − t = t1 + k for t1 ∈ [0, 1) and 0 ≥ k ∈ Z. Then r sub. To see that mr ∈ Ωr r Wr(Rt(t0)) = Wr(t1) =(cid:16) 1 − e−r(cid:17)e−rt1 =(cid:16) 1 − e−r(cid:17)erke−r(t1+k) 1 − e−r(cid:17)erke−r(t0−t) = erkertWr(t0) ≤ ertWr(t0), r =(cid:16) where the inequality follows because rk ≤ 0. So mr ∈ Ωr sub. For 0 ≤ s < 1 and Borel U ⊆ S, we have mr ◦ Rs(Rt(U)) = mr(Rt(Rs(U))) ≤ ertmr ◦ Rs(U) for all t ∈ [0, ∞) and Borel U ⊆ S, and hence mr ◦ Rs ∈ Ωr sub. Since Ωr sub is convex and weak∗ closed, we have conv{mr ◦ Rs : 0 ≤ s < 1} ⊆ Ωr sub. For each n ∈ N and 0 ≤ j < 2n we let U n j sub. For := the reverse containment, fix m ∈ Ωr [j/2n, (j + 1)/2n), and By Lemma 7.2 we can express xn as a convex combination xn = P2n−1 2n−1} described at (7.2). We claim that the measures 0 , . . . , vn {vn j=1 λn j vn j of the vectors converge weak∗ to m. To see this, fix f ∈ C(S)+. It suffices to prove that R f dMn →R f dm. For each n, let 2n−1 xn :=(cid:0)m(U n j )(cid:1)2n−1 j=0 ∈ [0, 1]2n . Mn := λn j (mr ◦ Rj/2n) 2n−1 Xj=0 M ′ n := Xj=0 λn j (mn,r ◦ Rj/2n). THE TOEPLITZ NONCOMMUTATIVE SOLENOID AND ITS KMS STATES 17 Corollary 7.4 shows that kMn − M ′ suffices to prove that For each n, define fn : S → R by n → 0. So it nk1 → 0 and in particular, R f dMn −R f dM ′ R f dM ′ n →R f dm. Xj=0 f (j/2n)1U n j . fn = 2n−1 Since f is uniformly continuous on S we have fn → f pointwise on S. Since f and each fn are bounded above by kf k∞, the Dominated Convergence Theorem implies thatR fn dm →R f dm. So it now suffices to prove that → 0. j )k, and hence n(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)R fn dm −R f dM ′ 0 )j−kZU n k f dµ = 2n(vn f dµ. j )kZU n k Fix j, k ∈ Z/2Z. Then (vn 0 )j−k = (vn Hence ZU n k f d(cid:0)mn,r ◦ Rj/2n(cid:1) = 2n(vn (cid:12)(cid:12)(cid:12)Z fn dm −Z f dM ′ n(cid:12)(cid:12)(cid:12) 2n−1 2n−1 2n−1 2n−1 − λn 2n−1 2n−1 2n−1 2n−1 2n−1 2n−1 2n−1 λn l λn j λn j i ) − l vn λn Xl=0 f (i/2n)m(U n Xk=0 ZU n =(cid:12)(cid:12)(cid:12) j(cid:16) Xi=0 Xj=0 =(cid:12)(cid:12)(cid:12) l(cid:17)i f (i/2n)(cid:16) Xi=0 Xj=0 =(cid:12)(cid:12)(cid:12) Xi=0 (cid:0)f (i/2n)(vn Xl=0 Xj=0 l )i(cid:1) − =(cid:12)(cid:12)(cid:12) Xk=0 (cid:16)2n(vn j(cid:16) Xj=0 Xi=0 (cid:0)f (i/2n)(vn j )i(cid:1) − j )kZU n =(cid:12)(cid:12)(cid:12) j(cid:16) Xi=0 (cid:16)f (i/2n)(vn Xj=0 j k1 = 1 and each Pj λn (cid:12)(cid:12)(cid:12)Z fn dm −Z f dM ′ n(cid:12)(cid:12)(cid:12) j = 1, the triangle inequality gives j )i − 2n(vn 2n−1 2n−1 2n−1 2n−1 2n−1 2n−1 λn λn ≤ ≤ max 0≤j<2n ≤ max i λn 2n−1 (vn Xj=0 Xi=0 (cid:16)f (i/2n) − 2nZU n j(cid:12)(cid:12)(cid:12) f (i/2n) − 2nZU n j )i(cid:12)(cid:12)(cid:12) Xi=0 f (i/2n) − 2nZU n 0≤i<2n(cid:12)(cid:12)(cid:12) 0≤j<2n(cid:16) max f (i/2n) − 2nZU n f dµ(cid:12)(cid:12)(cid:12) 0≤i<2n(cid:12)(cid:12)(cid:12) . i i i = max k k k k 2n−1 f d(cid:0)mn,r ◦ Rj/2n(cid:1)(cid:17)(cid:12)(cid:12)(cid:12) j )kZU n f dµ(cid:17)(cid:12)(cid:12)(cid:12) Xk=0 (cid:16)2n(vn j )kZU n f dµ(cid:17)(cid:12)(cid:12)(cid:12) Xk=0 (cid:16)2n(vn j )kZU n f dµ(cid:17)(cid:17)(cid:12)(cid:12)(cid:12) f dµ(cid:17)(cid:17)(cid:12)(cid:12)(cid:12) j )i(cid:17)(cid:12)(cid:12)(cid:12) f dµ(cid:17)(vn f dµ(cid:12)(cid:12)(cid:12) f dµ(cid:12)(cid:12)(cid:12)(cid:17) . k Since each kvn Fix 0 ≤ i ≤ 2n. The quantity 2nRU n i the Mean Value Theorem implies that there exists c ∈ U n f dµ is the average value of f over U n i . Since f is continuous, f dµ. Hence i such that f (c) = 2nRU n f (i/2n) − f (c). i (cid:12)(cid:12)(cid:12)Z fn dm −Z f dM ′ n(cid:12)(cid:12)(cid:12) ≤ max 0≤i<2n sup c∈U n i 18 BROWNLOWE, HAWKINS, AND SIMS Fix ǫ > 0. By uniform continuity of f there exists N such that x − y < 2−N =⇒ f (x) − f (y) < ǫ. For n ≥ N we have supc∈U n i f (i/2n) − f (c) ≤ ǫ for all i, giving (cid:12)(cid:12)(cid:12)R fn dm − → 0. So m ∈ conv{mr ◦ Rs : 0 ≤ s < 1}, giving n(cid:12)(cid:12)(cid:12) R f dM ′ ≤ ǫ. Hence (cid:12)(cid:12)(cid:12)R fn dm −R f dρn(cid:12)(cid:12)(cid:12) sub ⊆ conv{mr ◦ Rs : 0 ≤ s < 1} as required. For the final statement, observe that Ωr Ω0 sub = {m ∈ M(S) : m(Rt(U)) ≤ m(U) for all t ∈ [0, ∞) and Borel U ⊆ S}. So if m ∈ Ω0 sub, then m(U) = m(R1−t(Rt(U))) ≤ m(Rt(U)) ≤ m(U) for all U, t, forcing m(U) = m(Rt(U)) for all U, t. Uniqueness of the Haar measure µ on the compact group S therefore gives m = µ. So Ω0 (cid:3) sub ⊆ {µ}. The reverse containment is trivial. We can use Theorem 7.1 to describe the extreme points of Ωr sub. Proposition 7.5. The set {mr ◦ Rs : 0 ≤ s < 1} is the set of extreme points of Ωr sub. The first step in proving Proposition 7.5 will be to show that mr itself is an extreme point of Ωr sub. The following lemma will help. Lemma 7.6. Let m ∈ Ωr m([ i n )) = mr([ i n , i+1 n , i+1 n )) for all 0 ≤ i < n. sub and n ∈ N with n ≥ 1. If m([ n−1 n , 1)) ≤ mr([ n−1 n , 1)), then Proof. First observe that by definition of mr, we have mr(Rt(U)) = ertmr(U) whenever U ∪ U − t ⊆ [0, 1). Using this at the fourth equality, we note that if m([ n−1 n , 1)), then subinvariance forces n , 1)) ≤ mr([ n−1 1 = m(S) = n−1 Xi=0 , m(cid:16)h i n i + 1 n (cid:17)(cid:17) = ≤ ≤ = n−1 n−1 n−1 Xi=0 Xi=0 Xi=0 Xi=0 n−1 , 1(cid:17)(cid:17)(cid:17) , 1(cid:17)(cid:17) , 1(cid:17)(cid:17) n n m(cid:16)R(n−1−i)/n(cid:16)h n − 1 e(n−1−i)r/nm(cid:16)h n − 1 e(n−1−i)r/nmr(cid:16)h n − 1 mr(cid:16)h i n (cid:17)(cid:17) i + 1 n n , = 1. So we have equality throughout. From this we deduce first that n−1 n−1 , n n i + 1 n , i+1 n , i+1 m(cid:16)h i e(n−1−i)r/nm(cid:16)h n − 1 n (cid:17)(cid:17) = , 1(cid:17)(cid:17). Xi=0 Xi=0 Since the subinvariance relation forces m(cid:0)(cid:2) i n (cid:1)(cid:1) ≤ e(n−1−i)r/nm(cid:0)(cid:2) n−1 n (cid:1)(cid:1) = e(n−1−i)r/nm(cid:0)(cid:2) n−1 deduce that m(cid:0)(cid:2) i n , 1(cid:1)(cid:1) for each i. Since e(n−1−i)r/nm(cid:16)h n − 1 e(n−1−i)r/nmr(cid:16)h n − 1 , 1(cid:17)(cid:17) = Xi=0 Xi=0 n , 1(cid:1)(cid:1) = mr(cid:0)(cid:2) n−1 we also have m(cid:0)(cid:2) n−1 n (cid:17)(cid:17) = e(n−1−i)r/nm(cid:16)h n−1 n , 1(cid:1)(cid:1). Hence for each i we have n , 1(cid:17)(cid:17) = e(n−1−i)r/nmr(cid:16)h n−1 m(cid:16)h i n , i+1 n−1 n−1 n n n , 1(cid:1)(cid:1) for each i, we , 1(cid:17)(cid:17), n , 1(cid:17)(cid:17) = mr(cid:16)h i n, i+1 n (cid:17)(cid:17). (cid:3) THE TOEPLITZ NONCOMMUTATIVE SOLENOID AND ITS KMS STATES 19 Proof of Proposition 7.5. We first show that mr is an extreme point of Ωr m ∈ Ωr For each n define fn : S → R by sub. First suppose n , 1)) for all n. We claim that m = mr. Fix f ∈ C(S)+. sub satisfies m([ n−1 n , 1)) ≤ mr([ n−1 fn = n−1 Xi=0 f (i/n)1[ i n ). n , i+1 mr([ i n , i+1 n )) = n )) for all n ≥ 1 and 0 ≤ i < n. Hence the Dominated Convergence Theorem gives The Dominated Convergence Theorem gives R fn dm →R f dm. By Lemma 7.6, m([ i R fn dm =R fn dmr →R f dmr. It follows that m = mr. sub, t ∈ (0, 1) and that one of m1 and m2 is not equal to mr; Now suppose that m1, m2 ∈ Ωr say m1 6= mr. The above claim yields n such that m1([ n−1 n , i+1 n , 1)) > mr([ n−1 n , 1)). So (tm1 + (1 − t)m2)(cid:16)h n − 1 n , 1(cid:17)(cid:17) > (tmr + (1 − t)m2)(cid:16)h n − 1 n , 1(cid:17)(cid:17) ≥ mr(cid:16)h n − 1 n , 1(cid:17)(cid:17), and hence tm1 +(1−t)m2 6= mr. So mr cannot be expressed as a nontrivial convex combination of subinvariant probability measures, and hence is an extreme point of Ωr For s ∈ S, the map m 7→ m ◦ Rs is an affine homeomorphism of Ωr sub, so each m ◦ Rs is an extreme point of Ωr sub. For the reverse containment, observe that the space Ωr sub. This gives {mr ◦ Rs : s ∈ S} ⊆ ∂Ωr sub of all subinvariant probability measures on S is a weak∗-compact convex subset of the Banach space of all signed Borel measures on S. The map s 7→ mr ◦ Rs is a homeomorphism of S onto Z := {mr ◦ Rs : s ∈ S}. So Z is compact and in particular closed. Since Ωr sub is the closed convex hull of Z it follows from [25, Proposition 1.5] that the set of extreme points of Ωr sub is contained in the closure of Z and therefore in Z itself. (cid:3) sub. 8. Proof of the main theorem We are now almost ready to prove Theorem 6.6. We saw in Theorem 6.9 that the KMSβ simplex of T S sub under the maps induced θ by the covering maps pN : S → S. So we now show that these induced maps carry extreme points to extreme points. is affine isomorphic to the projective limit of the Ωrj Lemma 8.1. Let N ∈ N with N ≥ 2, θ = (θj)∞ for all j. For each j ∈ N, let rj := β N jθj by (7.1). For each s ∈ [0, 1), we have mrj+1 ◦ Rs ◦ p−1 j=0 ∈ ΞN , and β ∈ (0, ∞). Suppose that θj 6= 0 , and let mrj be the subinvariant measure on S defined N = mrj ◦ RN s. Proof. We first establish the result with s = 0. Fix 0 ≤ a < b ≤ 1. It suffices to prove that mrj+1 ◦ p−1 N (cid:0)(a, b)(cid:1) = mrj(cid:0)(a, b)(cid:1). We have mrj(cid:0)(a, b)(cid:1) =Z b Wrj (t) dt = a rj 1 − e−rj Z b a e−rjt dt = −1 1 − e−rj(cid:0)e−rj b − e−rja(cid:1). (8.1) We also have Since mrj+1 ◦ p−1 Xi=0 N (cid:0)(a, b)(cid:1) = Z Wrj+1(t) dt =Z (cid:16) N mrj+1(cid:18)(cid:16) a + i N , b + i N (cid:17)(cid:19) = N Xi=0 Z b+i N a+i N Wrj+1(t) dt. (8.2) rj+1 1 − e−rj+1(cid:17)e−rj+1t dt = −1 1 − e−rj+1 e−rj+1t, 20 BROWNLOWE, HAWKINS, AND SIMS Equation (8.2) gives mrj+1 ◦ p−1 N (cid:0)(a, b)(cid:1) = = = = Since N 2θj+1 = θj, we have −1 1 − e−rj+1 −1 1 − e−rj+1 −1 1 − e−rj+1 N b+i N N N a+i e− i Xi=0 he−rj+1ti N rj+1(cid:0)e− b Xi=0 N (cid:0)e− b 1 − e−rj+1 rj+1 1 − e− N rj+1 − e− a −1 1 − e− rj+1 N (cid:0)e− b N rj+1 − e− a N rj+1 − e− a N rj+1(cid:1) N rj+1(cid:1) N rj+1(cid:1). (8.3) rj+1 N = β/(N j+1θj+1) N = β/(N j · N 2θj+1) = β/N jθj = rj, and so (8.3) is precisely (8.1). Now for s 6= 0, observe that pN ◦ Rs = RN s ◦ pN so that Rs(p−1 N (U)) = p−1 N (RN s(U)) for all U ⊆ S. Hence (cid:3) mrj+1 ◦ Rs ◦ p−1 N = mrj+1 ◦ p−1 N ◦ RN s = mrj ◦ RN s. sub, m 7→ m ◦ p−1 (Ωrj We now describe the extreme points of the space lim ←− N ). Given a Borel map ψ : X → Y , we write ψ∗ : M(X) → M(Y ) for the induced map ψ∗(m)(U) = m(ψ−1(U)). Lemma 8.2. Take N ∈ {2, 3, . . . }, fix θ = (θj)∞ θj 6= 0 for all j. For each j ∈ N, let rj := β N j θj defined by (7.1). The map π : (sj)∞ (Ωrj the set of extreme points of lim ←− Proof. Since the Ωrj limit lim ←− hence closed. So to see that the image of π contains all of the extreme points of lim ←− suffices by [25, Proposition 1.5] to show that lim ←− sub are compact convex sets and (pN )∗ is affine and continuous, the projective sub is a compact convex set. The map π is continuous, so its range is compact and sub, it sub is contained in the closed convex hull of j=0 ∈ ΞN , and fix β ∈ (0, ∞). Suppose that , and let mrj be the subinvariant measure on S (S, pN ) onto j=1 7→ (mrj ◦ Rsj )∞ sub, (pN )∗). j=1 is a homeomorphism of lim ←− Ωrj Ωrj Ωrj the π(cid:0)(sj)∞ j=1(cid:1). For this, fix a point (mj)∞ of the projective-limit topology, there exist k ∈ N and Uk ⊆ Ωrk set Z(Uk) satisfies (mj)∞ sub. Take an open neighbourhood U of (mj). By definition sub open such that the cylinder j=1 ∈ Z(Uk) ⊆ U. By Theorem 7.1, there exist t1, . . . , tL ∈ [0, 1] with j=1 ∈ lim ←− Ωrj P tl = 1 such that Now for each j ∈ N, define m′ we have m′ j = (pN )j ′−j (m′ ∗ π(cid:0)(N j−lsl)∞ j=1(cid:1), and so That is, lim ←− lim ←− Ωrj sub, (pN )∗). (Ωrj sub ⊆ conv(cid:0)π(cid:0) lim ←− PL j :=PL j ′), and so (m′ l=1 tl(mrk ◦ Rsl) ∈ Uk. l=1 tl(mrl ◦ RN j−lsl). Lemma 8.1 shows that for j ≤ j′ ∈ N j=1 = sub. For l ≤ L, we have (mrl ◦ RN j−lsl)∞ j)∞ Ωrj j=1 ∈ lim ←− For the reverse containment, it suffices to show that each π(cid:0)(sj)∞ sub. For this, suppose that t ∈ (0, 1) and m′, m′′ ∈ lim ←− Ωrj Ωrj lim ←− (m′ j)∞ ←− (S, pN )(cid:1) ∩ U. j=1 ∈ conv π(cid:0) lim S(cid:1)(cid:1). So the range of π contains all the extreme points of j=1(cid:1) is an extreme point of sub satisfy For each j, j=1(cid:1) = tm′ + (1 − t)m′′. π(cid:0)(sj)∞ j=1(cid:1)j = (tm′ + (1 − t)m′′)j = tm′ mrj ◦ Rsj = π(cid:0)(sj)∞ j + (1 − t)m′′ j . THE TOEPLITZ NONCOMMUTATIVE SOLENOID AND ITS KMS STATES 21 Proposition 7.5 shows that each mrj ◦ Rsj is an extreme point of Ωrj sub, forcing m′ j = m′′ j = mrj ◦ Rsj . So m′ = m′′ = π(cid:0)(sj)∞ j=1(cid:1). compact space to a Hausdorff space. Finally, π is a homeomorphism onto its range because it is a continuous injection from a (cid:3) The final ingredient needed for the proof of Theorem 6.6 is a suitable action λ of S on T S θ . Lemma 8.3. There is an action λ of S = lim ←− (S, pN ) on T S θ for all j, a, b ≥ 0 and f ∈ C(S). λ(sj)∞ j=1(cid:0)ψj,∞(sa θj iθj (f )s∗b θj )(cid:1) = ψj,∞(cid:0)sa such that θj iθj (f ◦ Rsj )s∗b θj(cid:1) Proof. For each j ∈ N, and each t ∈ S, there is an automorphism of the topological graph Eθj given by s 7→ s + t for s ∈ E0 θj = S. This automorphism induces an automorphism λj,t of T (Eθj ) such that λj,t(sa θj for all j, a, b ≥ 0 and f ∈ C(S). θj = S, and s 7→ s + t for s ∈ E1 θj iθj (f ◦ Rt)s∗b θj iθj (f )s∗b θj ) = sa Since λj,t(sθj ) = sθj and λj,t(iθj (f )) = iθj (f ◦ Rt) for all f ∈ C(S), a routine calculation shows j=1 ∈ S , we have ψj ◦ λj,sj = λj+1,sj+1 ◦ ψj, and so the universal property of the (cid:3) that for (sj)∞ direct limit yields the desired action λ of S on lim −→ (T (Eθj ), ψj) = T S θ . Proof of Theorem 6.6. Theorem 6.9 yields an affine isomorphism ω : KMSβ(T S θ , α) → lim ←− (Ωrj sub, (pN )∗). Ωrj S, so the extreme boundary of KMSβ(T S θ sub is homeomorphic to the solenoid Lemma 8.2 shows that the space of extreme points of lim ←− S. As discussed on lim ←− pages 141 and 138 of [28], the set of KMS states for a given dynamics on a unital C ∗-algebra at given inverse temperature β is a Choquet simplex. So KMSβ(T S , α) is a Choquet simplex, θ and therefore affine isomorphic to the simplex of Borel probability measures on its extreme boundary. , α) is homeomorphic to lim ←− We claim that the action λ of Lemma 8.3 induces a free and transitive action of S on the extreme boundary of the KMSβ-simplex. The formula (6.4) shows that for l ∈ N, we have ω(φ ◦ λ(sj )∞ j=1 )l = ω(φ)l ◦ Rsl. That is, for (mj)∞ particular, if π : lim ←− j=1 ∈ lim ←− S → lim ←− (Ωrj sub), we have ω−1((mj)∞ j=1) ◦ λ(sj )∞ Ωrj sub is the map of Lemma 8.2, then j=1 = ω−1((mj ◦ Rsj )∞ j=1). In ω−1(π((tj)∞ j=1)) ◦ λ(sj)∞ j=1 = ω−1(π((tj − sj)∞ j=1)). That is, the homeomorphism ω−1 ◦ π of S onto the extreme boundary of KMSβ(T S θ twines λ with the action of S on itself by translation, which is free and transitive. Now suppose that β = 0. Then each Ωrj sub = {µ}, and so Theorem 6.9 gives an affine injection of KMS0(T S , α) into the 1-point space lim ({µ}, id). So there is at most one KMS0- ←− θ state. That there is one follows from a standard argument: Choose βn ∈ (0, ∞) converging to , α). Weak∗-compactness of the state space ensures that the 0. For each n, fix φn ∈ KMSβn(T S θ φn have a convergent subsequence. Its limit is a KMS0-state by [3, Proposition 5.3.23]. sub = Ω0 , α) inter- It remains to show that the KMS0 state is the only one that factors through AS there are no KMSβ states for β < 0. For any β, if φ is a KMSβ state of T S θ θ , and that , then in particular, φ(ψ1,∞(sθ1s∗ θ1)) = φ(ψ1,∞(s∗ θ1)αiβ(ψ1,∞(sθ1))) = e−βφ(ψ1,∞(s∗ θ1sθ1)) = e−βφ(1T S θ ), (8.4) and since φ is a state, we deduce that φ(1T S we have 1T S states for β < 0. − ψ1,∞(sθ1s∗ θ θ1)) = 1 − e−β. Since sθ1 is an isometry, θ1) ≥ 0 forcing 1 − e−β ≥ 0 and hence β ≥ 0. So there are no KMSβ − ψ1,∞(sθ1s∗ θ If β > 0, then (8.4) shows that φ(1T S − ψ1,∞(sθ1s∗ θ1)) > 0, whereas the image of 1T S − θ θ ψ1,∞(sθ1s∗ θ1) in AS θ is equal to zero. Hence φ does not factor through AS θ . 22 BROWNLOWE, HAWKINS, AND SIMS It remains to prove that if φ is a KMS0 state, then φ factors through AS implies that φ(1T S Lemma 6.2 implies that it generates the kernel of the quotient map q : T S Lemma 2.2] implies that φ factors through AS θ . θ1)) = 0. The projection 1T S − ψ1,∞(sθ1s∗ θ − ψ1,∞(sθ1s∗ θ θ . Equation 8.4 θ1) is fixed by α, and θ . So [12, (cid:3) θ → AS References [1] Z. Afsar, A. an Huef and I. Raeburn, KMS states on C ∗-algebras associated to local homeomorphisms, Internat. J. Math. 25 (2014), 1450066, 28pp. [2] J.-B. Bost and A. Connes, Hecke algebras, type III factors and phase transitions with spontaneous symmetry breaking in number theory, Selecta Math. (N.S.) 1 (1995), 411 -- 457. [3] O. Bratteli and D.W. Robinson, Operator algebras and quantum statistical mechanics. 2, Equilibrium states. Models in quantum statistical mechanics, Springer-Verlag, Berlin, 1997, xiv+519. [4] N. Brownlowe, A. an Huef, M. Laca and I. Raeburn, Boundary quotients of the Toeplitz algebra of the affine semigroup over the natural numbers, Ergodic Theory Dynam. Systems 32 (2012), 35 -- 62. [5] J. Christensen and K. Thomsen, Finite digraphs and KMS states, J. Math. Anal. Appl. 433 (2016), 1626 -- 1646. [6] L.O. Clarke, A. an Huef and I. Raeburn, Phase transitions on the Toeplitz algebras of Baumslag -- Solitar semigroups, Indiana University Mathematical Journal, to appear (arXiv:1503.04873 [math.OA]). [7] K.R. Davidson, C ∗-algebras by example, American Mathematical Society, Providence, RI, 1996, xiv+309. [8] M. Enomoto, M. Fujii and Y. Watatani, KMS states for gauge action on OA, Math. Japon. 29 (1984), 607 -- 619. [9] C. Farsi, E. Gillaspy, S. Kang and J.A. Packer, Separable representations, KMS states, and wavelets for higher-rank graphs, J. Math. Anal. Appl. 434 (2016), 241 -- 270. [10] N. J. Fowler and I. Raeburn, The Toeplitz algebra of a Hilbert bimodule, Indiana Univ. Math. J. 48 (1999), 155 -- 181. [11] M. Hawkins, Applications of compact topological graph C ∗-algebras to noncommutative solenoids, PhD Thesis, University of Wollongong 2015. [12] A. an Huef, M. Laca, I. Raeburn and A. Sims, KMS states on the C ∗-algebras of finite graphs, J. Math. Anal. Appl. 405 (2013), 388 -- 399. [13] A. an Huef, M. Laca, I. Raeburn and A. Sims, KMS states on the C ∗-algebra of a higher-rank graph and periodicity in the path space, J. Funct. Anal. 268 (2015), 1840 -- 1875. [14] T. Kajiwara and Y. Watatani, KMS states on finite-graph C ∗-algebras, Kyushu J. Math. 67 (2013), 83 -- 104. [15] E.T.A. Kakariadis, KMS states on Pimsner algebras associated with C ∗-dynamical systems, J. Funct. Anal. 269 (2015), 325 -- 354. [16] T. Katsura, A class of C ∗-algebras generalizing both graph algebras and homeomorphism C ∗-algebras I, Fundamental results, Trans. Amer. Math. Soc. 356 (2004), 4287 -- 4322. [17] T. Katsura, On C ∗-algebras associated with C ∗-correspondences, J. Funct. Anal. 217 (2004), 366 -- 401. [18] T. Katsura, A class of C ∗-algebras generalizing both graph algebras and homeomorphism C ∗-algebras II. Examples, Internat. J. Math. 17 (2006), 791 -- 833. [19] M. Laca and S. Neshveyev, KMS states of quasi-free dynamics on Pimsner algebras, J. Funct. Anal. 211 (2004), 457 -- 482. [20] M. Laca and I. Raeburn, Phase transition on the Toeplitz algebra of the affine semigroup over the natural numbers, Adv. Math. 225 (2010), 643 -- 688. [21] M. Laca, I. Raeburn, J. Ramagge and M.F. Whittaker, Equilibrium states on the Cuntz -- Pimsner algebras of self-similar actions, J. Funct. Anal. 266 (2014), 6619 -- 6661. [22] F. Latr´emoli`ere and J.A. Packer Noncommutative solenoids, New York J. Math., to appear (arXiv:1110.6227 [math.OA]). [23] F. Latr´emoli`ere and J.A. Packer, Noncommutative solenoids and their projective modules, Contemp. Math., 603, Commutative and noncommutative harmonic analysis and applications, 35 -- 53, Amer. Math. Soc., Providence, RI, 2013. [24] F. Latr´emoli`ere and J.A. Packer, Explicit construction of equivalence bimodules between noncommutative solenoids, Contemp. Math., 650, Trends in harmonic analysis and its applications, 111 -- 140, Amer. Math. Soc., Providence, RI, 2015. [25] R.R. Phelps, Lectures on Choquet's theorem, Springer -- Verlag, Berlin, 2001, viii+124. [26] M.V. Pimsner, A class of C ∗-algebras generalizing both Cuntz -- Krieger algebras and crossed products by Z, Fields Inst. Commun., 12, Free probability theory (Waterloo, ON, 1995), 189 -- 212, Amer. Math. Soc., Providence, RI, 1997. [27] W. Rudin, Real and complex analysis, McGraw -- Hill Book Co., New York, 1987, xiv+416. THE TOEPLITZ NONCOMMUTATIVE SOLENOID AND ITS KMS STATES 23 [28] M. Takesaki and M. Winnink, Local normality in quantum statistical mechanics, Comm. Math. Phys. 30 (1973), 129 -- 152. [29] D. Yang, Endomorphisms and modular theory of 2-graph C ∗-algebras, Indiana Univ. Math. J. 59 (2010), 495 -- 520. Nathan Brownlowe, School of Mathematics and Statistics, University of Sydney, NSW 2006, AUSTRALIA E-mail address: [email protected] Mitchell Hawkins and Aidan Sims, School of Mathematics and Applied Statistics, University of Wollongong, NSW 2522, AUSTRALIA E-mail address: [email protected], [email protected]
1901.08056
2
1901
2019-08-14T19:33:24
Measures of weak non-compactness in preduals of von Neumann algebras and JBW$^*$-triples
[ "math.OA", "math.FA" ]
We prove, among other results, that three standard measures of weak non-compactness coincide in preduals of JBW$^*$-triples. This result is new even for preduals of von Neumann algebras. We further provide a characterization of JBW$^*$-triples with strongly WCG predual and describe the order of seminorms defining the strong$^*$ topology. As a byproduct we improve a characterization of weakly compact subsets of a JBW$^*$-triple predual, providing so a proof for a conjecture, open for almost eighteen years, on weakly compact operators from a JB$^*$-triple into a complex Banach space.
math.OA
math
MEASURES OF WEAK NON-COMPACTNESS IN PREDUALS OF VON NEUMANN ALGEBRAS AND JBW∗-TRIPLES JAN HAMHALTER, OND REJ F.K. KALENDA, ANTONIO M. PERALTA, AND HERMANN PFITZNER Abstract. We prove, among other results, that three standard measures of weak non-compactness coincide in preduals of JBW∗-triples. This result is new even for preduals of von Neumann algebras. We further provide a characteri- zation of JBW∗-triples with strongly WCG predual and describe the order of seminorms defining the strong∗ topology. As a byproduct we improve a char- acterization of weakly compact subsets of a JBW∗-triple predual, providing so a proof for a conjecture, open for almost eighteen years, on weakly compact operators from a JB∗-triple into a complex Banach space. 1. Introduction Measures of weak non-compactness are an important tool for a deeper under- standing of weak compactness of sets and operators. They are used to prove more precise versions of known results and to establish new results as well. As an il- lustration we mention a fixed-point theorem [21], quantitative versions of Krein's theorem [26, 31], James' compactness theorem [18, 32], Eberlein-Smulyan theorem [3] and Gantmacher's theorem [4]. There are several procedures how one can measure weak non-compactness of sets. There are two basic non-equivalent ways -- on the one hand the De Blasi measure introduced and used in [21], and on the other hand various mutually equivalent quantities used in the other above-quoted papers. Their non-equivalence follows from [5, 4]. However, the counterexample witnessing their non-equivalence is an artificially constructed Banach space -- it is constructed as the ℓ1-sum of suitable renormings of the space ℓ1. It seems to be still an open question whether there is a 'natural' Banach space where they fail to be equivalent. This question seems to be rather interesting as in many classical spaces all these measures of weak non-compactness are equivalent. This applies, for example, to the Lebesgue spaces L1(µ) for an arbitrary non-negative σ-additive measure µ [46, The- orem 7.5], including the special case ℓ1(Γ) [46, Proposition 7.3], the space c0(Γ) [46, Proposition 10.2], the space of nuclear operators N (ℓq(Λ), ℓp(J)) for p, q ∈ (1,∞) 2010 Mathematics Subject Classification. 46B50, 46L70, 17C65. Key words and phrases. measure of weak non-compactness, von Neumann algebra, JBW∗- triple, JBW∗-algebra, strong∗ topology. The first two authors were in part supported by the Research Grant GA CR 17-00941S. The first author was partly supported further by the project OP VVV Center for Advanced Applied Science CZ.02.1.01/0.0/0.0/16 019/000077. The third author was partially supported by the Spanish Ministry of Science, Innovation and Universities (MICINN) and European Regional Development Fund project no. PGC2018-093332-B-I00 and Junta de Andaluc´ıa grant FQM375. 1 2 J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER [34, Theorem 2.1], and preduals of atomic von Neumann algebras [34, Theorem 2.2]. In the present paper we prove the equivalence of the measures of weak non- compactness in preduals of JBW∗-triples. This covers, in particular, the preduals of general von Neumann algebras. The machinery developed in this paper also have some important consequences in improving our current knowledge on relatively weakly compact subsets in the predual of a JBW∗-triple. More concretely, the result established in [62] asserts that a bounded subset, K, in the predual of a JBW∗-triple, M , is relatively weakly compact if and only if there is a couple of normal functionals ϕ1, ϕ2 in M∗ whose associated preHilbertian seminorm k · kϕ1,ϕ2 controls uniformly the values of all functionals in K on the closed unit ball of M (see sections 3, 6, and 7 for definitions). As shown, for example in [17, pages 340 and 341], the existence of control seminorms k·kϕ1,ϕ2 for relatively weakly compact subsets in M∗ can be applied to characterize weakly compact operators from a complex Banach space into the predual of a JBW∗-triple and from a JB∗-triple into a complex Banach space. It has been conjectured that, as in the case of von Neumann algebras (see [1, 45]), a single functional ϕ in the predual, M∗, of a JBW∗-triple M (and the associated seminorm k·kϕ) is enough to control relatively weakly compact subsets in M∗ and to determine those weakly compact operators from a complex Banach space into M∗. It was even so claimed in [20, Theorem 11]. However, some subtle difficulties in the Barton- Friedman's proof for the Grothendieck inequality for JB∗-triples, also affected the arguments in [20] (cf. section 11 or [17, page 341]). In this paper we also provide a complete proof for this conjecture and we show the validity of the original statement in [20, Theorem 11]. The paper is organized as follows. In Section 2 we give the definitions of the three measures of weak non-compact- ness we will deal with and provide their basic properties. Section 3 contains some background on JB∗-triples and, more specifically, on dual spaces among them, i.e., JBW∗-triples. We provide the basic notions and proper- ties, including the relationship to the classical subclasses formed by C∗-algebras (re- spectively, von Neumann algebras) and JB∗-algebras (respectively, JBW∗-algebras). In Section 4 we formulate our main result on the equivalence of the three measure of weak non-compactness in JBW∗-triple preduals (see Theorem 4.1). We further collect some consequences -- the special cases of the main result and the analogous result for real spaces. Sections 5, 6, 7 and 8 are devoted to the proof of the main result. On the way to the proof we establish several results which seem to be of an independent interest. In Section 5 we show that there is a close relationship between the equality of measures of weak non-compactness and the subsequence splitting property (see Proposition 5.1). In Section 6 we gather some properties of projections in C∗-algebras and JB∗- algebras and of tripotents in JB∗-triples. Most of the results presented there are known, but we point out that Lemmata 6.1 and 6.3 and Proposition 6.5 seem to be of independent interest. In Section 7 we investigate the relation between the strong∗ topology on a JBW∗- triple and the weakly compact subsets of its predual. The main achievements there MEASURES OF WEAK NON-COMPACTNESS 3 are Propositions 7.5 and 7.11, where we explore the connections between the natural partial order in the set of tripotents with the order among the associated seminorms and the first results leading to the existence of control functionals. Section 8 contains the culminating arguments of the proof of Theorem 4.1. In Section 9 we focus on σ-finite JBW∗-triples. We characterize those σ-finite JBW∗-triples whose predual is strongly WCG (see Theorem 9.3). This theorem reveals, in particular, a substantial difference between JBW∗-triples and JBW∗- algebras, as the predual of a σ-finite JBW∗-algebra is always strongly WCG. In Section 10 we apply the methods from Section 9 to prove Theorem 10.1 which provides a deep understanding of the structure of the strong∗ topology for general JBW∗-triples. In Section 11 we present the new advances on the characterization of relatively weakly compact subsets in the predual of a JBW∗-triple (see Theorem 11.3), and the subsequent consequences determining the weakly compact operators from a JB∗-triple into a complex Banach space (cf. Theorem 11.4), which, as we have already commented, provides a proof for a conjecture considered during the last eighteen years. The last section contains some open problems. 2. Measures of weak noncompactness Let X be a (real or complex) Banach space and A, B ⊆ X two nonempty sets. We set bd(A, B) = sup{dist(a, B); a ∈ A}. The Hausdorff measure of norm non-compactness is defined by the formula χ(A) = inf{bd(A, F ); F ⊆ X finite} = inf{bd(A, K); K ⊆ X compact} for a bounded set A ⊆ X. It is clear that χ(A) = 0 if and only if A is relatively norm compact. The De Blasi measure of weak non-compactness introduced in [21] is defined by ω(A) = inf{bd(A, K); K ⊆ X weakly compact}. (When confusion concerning the underlying space X could arise, like for example in Lemma 2.1, we will write ωX instead of ω.) It is a natural modification of the Hausdorff measure of non-compactness. Further, ω(A) = 0 if and only if A is relatively weakly compact. As remarked in [21] this was proved already by Grothendieck [33, p. 401] (compare also [22, Lemma XIII.2]). Another measure of weak non-compactness inspired by the Banach-Alaoglu the- orem is defined by w∗ , X). wkX (A) =bd(A w∗ is the closure of A in the space (X ∗∗, w∗), where X is considered to be Here A canonically embedded into its bidual. It is clear that A is relatively weakly compact if and only if A ⊆ X, i.e., if and only if wkX (A) = 0. w∗ We will use one more quantity, namely wckX (A) = sup{dist(clustw∗ (xn), X); (xn) is a sequence in A}, where clustw∗ (xn) denotes the set of all weak∗-cluster points of the sequence (xn) It follows easily from the Eberlein-Smulyan theorem that in the bidual X ∗∗. 4 J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER wckX (A) = 0 whenever A is relatively weakly compact. The converse follows from the quantitative version of the Eberlein-Smulyan theorem proven in [3]. It consists in the inequalities wckX (A) ≤ wkX (A) ≤ 2 wckX (A) which hold for any bounded subset A ⊆ X. Further, the following inequalities are easy to check: wkX (A) ≤ ω(A) ≤ χ(A). In general, the quantities wkX (·) and ω(·) (hence also wckX (·) and ω(·)) are not equivalent. As mentioned above, this was proved in [5, 4] but it follows also from [46, Theorem 2.3] using the fact that weakly compactly generated spaces are not stable under taking subspaces. If Y is a closed subspace of X and A is a bounded subset of Y , then we can consider the measures of weak non-compactness of the set A either in the space X or in the space Y . In general it is not the same but we have wkX (A) ≤ wkY (A) ≤ 2 wkX (A) , wckX (A) ≤ wckY (A) ≤ 2 wckX (A) . Indeed, in both cases the first inequality is trivial and the second one follows from [31, Lemma 11]. Further, we clearly have ωX (A) ≤ ωY (A) but the converse inequality is, in general, not true, neither up to a constant (this follows, for example, from [46, Theorem 2.3] with the same remark as above). However, if Y is 1-complemented in X, then the measures considered in X and in Y coincide. This is the content of the following easy lemma which is proved in [34, Lemma 3.8(b)]. Lemma 2.1. Let X be a Banach space, Y ⊆ X a 1-complemented subspace and A ⊆ Y a bounded set. Then wkX (A) = wkY (A) , wckX (A) = wckY (A) , ωX (A) = ωY (A). 3. JB∗-triples and their subclasses Originated as mathematical models of physical observables in quantum mechan- ics by authors like W. Heisenberg, P. Jordan and J. von Neumann, C∗-algebras constitute nowadays an area of intensive research in functional analysis. The ab- stract definition says that a C∗-algebra is a complex Banach algebra A equipped with an algebra involution ∗ satisfying the so-called Gelfand -- Naimark axiom, that is ka∗ak = kak2 for all a ∈ A. The celebrated Gelfand -- Naimark theorem represents each abstract C∗-algebra as a norm-closed selfadjoint subalgebra of the space B(H) of all continuous linear operators on a complex Hilbert space H. From the point of view of functional analysis, the class of C∗-algebras is not very stable for several properties. For example, a C∗-algebra is reflexive if and only if it is finite-dimensional, and C∗-algebras are not stable under contractive projections as we can find a contractive projection from B(H) onto the Hilbert space H. As observed by Poincar´e in 1907, for complex Banach spaces of dimension bigger than or equal to two, there exists a wide range of simply connected domains which are not biholomorphic to the unit ball. In other words, the Riemann mapping theorem cannot be easily generalized to arbitrary complex Banach spaces. JB∗- triples were introduced in 1983 by W. Kaup in a successful classification of bounded MEASURES OF WEAK NON-COMPACTNESS 5 symmetric domains in complex Banach spaces of arbitrary dimension (see [52]). A JB∗-triple is a complex Banach space E together with a (continuous) triple product {., ., .} : E3 → E, which is symmetric and bilinear in the outer variables and conjugate-linear in the middle one, satisfying the following algebraic -- analytic properties: (JB∗-1) L(x, y)L(a, b) = L(L(x, y)(a), b) −L(a, L(y, x)(b)) +L(a, b)L(x, y) for all a, b, x, y ∈ E, where given a, b ∈ E, L(a, b) stands for the (linear) operator on E given by L(a, b)(x) = {a, b, x}, for all x ∈ E (Jordan identity); (JB∗-2) The operator L(a, a) is a hermitian operator with nonnegative spectrum for each a ∈ E; (JB∗-3) k{a, a, a}k = kak3 for a ∈ E. Let us observe that the third axiom (JB∗-3) is a Jordan-geometric analogue of the Gelfand -- Naimark axiom. The mapping (x, y, z) 7→ {x, y, z} = 1 2 (xy∗z + zy∗x) can be applied to define a structure of JB∗-triple on a C∗-algebra A, or on the space B(H, K) of all bounded linear operators between complex Hilbert spaces H and K and in particular on an arbitrary Hilbert space H identified with B(H, C), in all these cases we keep the original norm of the corresponding underlying Banach spaces. Some of the lackings exhibited by the class of C∗-algebras are no longer a hand- icap for JB∗-triples. For example, JB∗-triples are stable under contractive pertur- bations (see [28, 75, 53]). One of the most interesting properties of JB∗-triples is that a linear bijection between JB∗-triples is an isometry if and only if it is a triple isomorphism, that is, it preserves triple products (see [52, Proposition 5.5]). In particular, if a complex Banach space E admits two triple products under which E is a JB∗-triple with respect to its original norm and any of these products, then both products coincide. An intermediate class between C∗-algebras and JB∗-triples is formed by JB∗- algebras. The hermitian part, Asa, of a C∗-algebra A is not, in general, closed under products. However Asa is a Jordan subalgebra of A when the latter is considered with its natural Jordan product defined by a ◦ b = 1 2 (ab + ba). P. Jordan released the notion of Jordan algebra to set a mathematical model for the algebra of observables in quantum mechanics. In abstract algebra, a Jordan algebra is a nonassociative algebra over a field whose multiplication, denoted by ◦, is commutative and satisfies the so-called Jordan identity (x◦ y)◦ x2 = x◦ (y ◦ x2). Given an element a in a Jordan algebra B, we shall write Ua for the linear mapping on B defined by Ua(b) := 2(a ◦ b) ◦ a − a2 ◦ b. A Jordan Banach algebra B is a Jordan algebra equipped with a complete norm satisfying ka ◦ bk ≤ kak · kbk for all a, b ∈ B. A complex Jordan Banach algebra B admitting an involution ∗ satisfying that kUa(a∗)k = kak3 for all a, b ∈ B is called a JB∗-algebra (see [79], In some texts the definition of JB∗-algebras includes the [16, Definition 3.3.1]). extra axiom that the involution in a JB∗-algebra is an isometry (cf. [36, §3.8]). The result in [79, Lemma 4] (see also [16, Proposition 3.3.13]) shows that this extra axiom is redundant. The self-adjoint part of a JB∗-algebra B will be denoted by Bsa. It is known that (real) JB-algebras are precisely the self-adjoint parts of JB∗- algebras (compare [78]). Any JB∗-algebra also admits a structure of a JB∗-triple when equipped with the triple product defined by {x, y, z} = (x◦ y∗)◦ z + (z ◦ y∗)◦ x − (x ◦ z) ◦ y∗ in which case Ua(b) = {a, b∗, a}. 6 J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER For a general overview of JB-algebras, JB∗-algebras and JB∗-triples the reader is referred to the monographs [36, 19, 16, 17]. A JBW∗-triple (respectively, a JBW∗-algebra) is a JB∗-triple (respectively, a JB∗-algebra) which is also a dual Banach space. JBW∗-triples can be considered as the Jordan alter-ego of von Neumann algebras. It should be noted that, as estab- lished by Sakai's theorem for von Neumann algebras, every JBW∗-triple admits a unique (isometric) predual and its product is separately weak∗-to-weak∗ continuous [8] (see also [17, Theorems 5.7.20 and 5.7.38] for new proofs). These facts rely on the fact proved in [8] (see also [70]) asserting that the bidual of any JB∗-triple E is a JB∗-triple under a triple product which extends that of E and is separately weak∗- to-weak∗ continuous. It is further known that every triple isomorphism between JBW∗-triples is automatically weak∗-continuous (cf. [40, Corollary 3.22]). A JC∗-algebra is a concrete JB∗-algebra which materializes as a norm-closed Jordan ∗-subalgebra of a C∗-algebra, and hence a norm-closed Jordan self-adjoint subalgebra of some B(H). A JW∗-algebra is a JC∗-algebra which is also a dual Ba- nach space, or equivalently, a weak∗-closed JB∗-subalgebra of some von Neumann algebra. There are examples of JB∗-algebras which are not JC∗-algebras. Macdon- ald's and Shirshov-Cohn's theorems are useful tools to describe certain important subalgebras of JB∗-algebras (see [36, Theorems 2.4.13 and 2.4.14]). For example, these structure results were originally applied by J.D.M. Wright in [78] to deduce that the JB∗-subalgebra of a JB∗-algebra generated by two hermitian elements (and the unit element) is a JC∗-algebra. This result and one of its multiple consequences are gathered in the next lemma (which is a variant of [16, Proposition 3.4.6]). Lemma 3.1. Let B be a unital JB∗-algebra and a ∈ B an arbitrary element. (i) Let N be the closed Jordan ∗-subalgebra of B generated by a and 1. Then N is Jordan (isometrically) ∗-isomorphic to a JB∗-subalgebra of some B(H), where H is a complex Hilbert space; (ii) The element a∗ ◦ a is positive in N . Moreover, a∗ ◦ a = 0 if and only if a = 0. Proof. (i) Note that N coincides with the unital Jordan ∗-subalgebra generated by two self-adjoint elements 1 2i (a − a∗). Hence the assertion follows immediately from [78, Corollary 2.2] (see also [16, Proposition 3.4.6]). Statement (ii) follows from (i) since the assertion is clearly true in B(H). (cid:3) 2 (a∗ + a) and 1 The literature offers a generous collection of structure results for JB∗-triples. The Gelfand -- Naimark theorem established by Y. Friedman and B. Russo proves that every JB∗-triple can be isometrically embedded as a JB∗-subtriple of an ℓ∞- sum of Cartan factors (see [30, Theorem 1] and details there). A consequence of this fact shows that every JB∗-triple is isometrically JB∗-triple isomorphic to a JB∗-subtriple of a JB∗-algebra. Let us fix some additional notation. Given two von Neumann algebras A ⊆ B(H) and W ⊆ B(K), the algebraic tensor product A ⊗ W is canonically embedded into B(H ⊗ K), where H ⊗ K is the hilbertian tensor product of H and K (see [77, Definition IV.1.2]). Then the von Neumann tensor product of A and W (denoted by A⊗W,) is precisely the von Neumann subalgebra generated by the algebraic tensor product A ⊗ W in B(H ⊗ K), that is, the weak∗ closure of A ⊗ W in B(H ⊗ K) (see [77, §IV.5]). MEASURES OF WEAK NON-COMPACTNESS 7 Suppose A is a commutative von Neumann algebra and M is a JBW∗-subtriple of some B(H). Following the standard notation, we shall write A⊗M for the weak∗-closure of the algebraic tensor product A ⊗ M in the usual von Neumann tensor product A⊗B(H) of A and B(H). Clearly A⊗M is a JBW∗-subtriple of It is well known that every Cartan factor C of type 1, 2, 3, or 4 is A⊗B(H). a JBW∗-subtriple of some B(H) (compare [41] and section 9 for a more detailed presentation of Cartan factors), and thus, the von Neumann tensor product A⊗C is a JBW∗-triple. The exceptional Cartan factors of type 5 and 6 can not be represented as JB∗- subtriples of some B(H). Fortunately, these factors are all finite-dimensional. Henceforth, if C denotes a finite-dimensional JB∗-triple, A⊗C will stand for the in- jective tensor product of A and C, which is clearly identified with the space C(Ω, C), of all continuous functions on Ω with values in C endowed with the pointwise op- erations and the supremum norm, where Ω denotes the spectrum of A (cf. [71, p. 49]). This convention is consistent with the definitions in the previous paragraph, because if C is a finite-dimensional Cartan factor which can be also embedded as a JBW∗-subtriple of some B(H) both definitions above give the same object (cf. [77, Theorem IV.4.14]). In the setting of JBW∗-triples, a much more precise description than that derived from the Gelfand-Naimark theorem was found by G. Horn and E. Neher in [41, (1.7)], [42, (1.20)], where they proved that every JBW∗-triple M writes in the form (1) ⊕ℓ∞ H(W, α) ⊕ℓ∞ pV, M =Mj∈J Aj⊗Cjℓ∞ where each Aj is a commutative von Neumann algebra, each Cj is a Cartan factor, W and V are continuous von Neumann algebras, p is a projection in V , α is a linear involution on W commuting with ∗, that is, a linear ∗-antiautomorphism of period 2 on W , and H(W, α) = {x ∈ W : α(x) = x}. 2] and [69, Theorem D.20]. We conclude this section by the following result originated from [20, Proposition Proposition 3.2. [17, Proposition 5.10.137] Let M be any JBW∗-triple. Then M is triple-isomorphic to a weak∗-closed subtriple of a JBW∗-algebra which is 1- complemented by a weak∗-to-weak∗ continuous projection. 4. Main results Our main result is the following theorem on coincidence of measures of weak non-compactness in preduals of JBW∗-triples. It is proved at the end of Section 8 below. Theorem 4.1. Let M be a JBW∗-triple and let A ⊆ M∗ be a bounded set. Then ω(A) = wkM∗ (A) = wckM∗ (A). Since JBW∗-algebras and von Neumann algebras are special cases of JBW∗- triples, the following two corollaries are immediate. Corollary 4.2. Let M be a JBW∗-algebra and let A ⊆ M∗ be a bounded set. Then ω(A) = wkM∗ (A) = wckM∗ (A). 8 J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER Corollary 4.3. Let M be a von Neuman algebra and let A ⊆ M∗ be a bounded set. Then ω(A) = wkM∗ (A) = wckM∗ (A). Note that Corollary 4.2 easily implies the main result Theorem 4.1 using Propo- sition 3.2 and Lemma 2.1. This is in fact the way the main result will be proved in Section 8. We further easily obtain the same results for real variants of the respective structures. Corollary 4.4. Let M be a real JBW∗-triple and let A ⊆ M∗ be a bounded set. Then ω(A) = wkM∗ (A) = wckM∗ (A). Proof. We use the following arguments explained in detail in [11, Section 6]. Let M be a real JBW∗-triple. Then there is a JBW∗-triple N and a weak∗- to-weak∗ continuous conjugate-linear isometry τ : N → N such that τ 2 = idN and For each ϕ ∈ N∗ define M = {x ∈ N ; x = τ (x)}. τ #(ϕ)(x) = ϕ(τ (x)), x ∈ N. Then τ # is a conjugate-linear isometry of N∗ onto N∗ such that (τ #)2 = idN∗ . Set N τ ∗ = {ϕ ∈ N∗; τ #(ϕ) = ϕ}. Then N τ real-linear isometry of N τ ∗ is a closed real-linear subspace of N∗ and the mapping ϕ 7→ Re ϕ is a ∗ onto M∗. Hence the formula (ϕ + τ #(ϕ)), ϕ ∈ N∗ defines a real-linear norm-one projection of N∗ onto N τ ∗ . P (ϕ) = 1 2 If X is a complex Banach space, denote by XR the underlying real space. Now, by Theorem 4.1 we know that the measures of weak non-compactness coincide for subsets of N∗. By the discussion at the end of Section 2.5 of [46] the measures of weak non-compactness with respect to N∗ coincide with those with respect for (N∗)R. Finally, N τ ∗ is 1-complemented in (N∗)R, hence Lemma 2.1 yields that the measures of weak non-compactness coincide for subsets of N τ ∗ . Since this space is isometric to M∗, the proof is complete. (cid:3) Since JBW-algebras are among the examples of a real JBW∗-triple, the following corollary follows immediately. Corollary 4.5. Let M be a JBW-algebra and let A ⊆ M∗ be a bounded set. Then ω(A) = wkM∗ (A) = wckM∗ (A). The following corollary also is a special case of our previous result, since the self-adjoint part of a von Neumann algebra M is a real JBW∗-triple and its predual is formed exactly by the self-adjoint elements of M∗. Corollary 4.6. Let M be a von Neuman algebra and let A ⊆ M∗,sa (where M∗,sa denotes the subspace of M∗ formed by self-adjoint functionals) be a bounded set. Then ω(A) = wkM∗,sa (A) = wckM∗,sa (A). MEASURES OF WEAK NON-COMPACTNESS 9 5. Subsequence splitting property In [46, Proposition 7.1] a special case of Theorem 4.1 was proved -- the equality of measures of weak non-compactness for subsets of L1(µ) for a finite measure µ. A key role in the proof is played by a variant of Rosenthal's subsequence splitting lemma. This is a motivation to define the following property of Banach spaces which turns out to be closely connected with the equality of measures of weak non-compactness. Let X be a Banach space. We say that X has • the subsequence splitting property if for any bounded sequence (xn) in X and any ε > 0 there is a subsequence (xnk ) which can be expressed as xnk = yk + zk, where (yk) is weakly convergent in X and (2) ≥ (1 − ε) nXj=1 αjkzjk for any n ∈ N and any choice of scalars α1, . . . , αn; • the isometric subsequence splitting property if for any bounded sequence (xn) in X there is a subsequence (xnk ) which can be expressed as xnk = yk + zk, where (yk) is weakly convergent in X and nXj=1 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) αjzj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) αj zj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) nXj=1 = nXj=1 αjkzjk for any n ∈ N and any choice of scalars α1, . . . , αn. That L1([0, 1]) has the isometric subsequence splitting property has been men- tioned by Bourgain and Rosenthal in [12] but a forerunner can be found in [47]. There are also splitting versions for Lp-spaces, 0 < p < ∞ but in this paper we naturally restrict to the case p = 1 because preduals of JBW∗-triples are gen- eralizations of (non-commutative) L1-spaces. The generalization of the isometric subsequence splitting property to preduals of von Neumann algebras was proved by Randrianantoanina [66] and, by different methods, by Raynaud and Xu [67]. Then Fern´andez-Polo, Ram´ırez and the third mentioned author [27] showed the isomet- ric subsequence splitting property for preduals of JBW∗-algebras and the two last named authors for JBW∗-triples [64]. The predual of a JBW∗-triple is L-embedded (see [17, Theorem 5.7.36] or [40], [8]). Recall that a Banach space X is called L-embedded if there is a projection P on X ∗∗ with image X such that kx∗∗k = kP x∗∗k + kx∗∗ − P x∗∗k for all x∗∗ ∈ X ∗∗; the standard reference for L-embedded spaces is [37, Chapter IV]. An L- embedded Banach space X is weakly sequentially complete [37, Theorem IV.2.2]. Hence if a bounded sequence (xn) of X does not contain any ℓ1-subsequence then by Rosenthal's theorem [25, Theorem 5.37] it contains a weak Cauchy subsequence (xnk ) and in this case the splitting property is trivially satisfied with yk = xnk and zk = 0. The interesting case appears when (xn) contains an ℓ1-sequence. This explains why we use Lemma 5.3 which says that the quality of an ℓ1-sequence (i.e., its James constant, see below) remains invariant, up to a subsequence, under a perturbation by a weak Cauchy sequence. The aim of this section is to show the following proposition whose main purpose for this paper is part (b). 10 J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER Proposition 5.1. Let X be a Banach space. Consider the following possible prop- erties of X. (I) For any bounded set A ⊆ X we have wckX (A) = ω(A). (II) For any bounded set A ⊆ X and any ε > 0 there is a countable subset C ⊆ A such that ω(C′) > ω(A) − ε for any infinite C′ ⊆ C. Then the following assertions are true. (a) (I)⇒(II); (b) If X has the subsequence splitting property, then (II)⇒(I); (c) If X is L-embedded and enjoys (I), then X has the subsequence splitting prop- erty. Before we pass to the proof of Proposition 5.1 we recall some known facts on ℓ1- sequences. To a bounded sequence (xn) in a Banach space X we associate its 'James con- stant ' (3) ; Xn≥m cJ (xn) = sup cm where the αnxn(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) cm = inf (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) αn = 1 Xn≥m n=m αnxnk ≥ (1 − ε)cJ (xn)P∞ for each ε > 0 there is m ∈ N such that kP∞ form an increasing sequence. If (xn) is equivalent to the canonical basis of ℓ1 then cJ (xn) > 0 and more specifically, cJ (xn) > 0 if and only if there is an integer m such that (xn)n≥m is equivalent to the canonical basis of ℓ1. The number cJ (xn) may be thought of as the approximately best l1-basis constant of (xn) in the sense that n=m αn for all (αn) ∈ ℓ1, and cJ (xn) cannot be replaced by a strictly greater constant. Further, a sequence (zl) will be called a block sequence of (xn) if there are suc- cessive finite sets Al ⊆ N (i.e., max Al < min Al+1 for l ∈ N) and a sequence of scalars (λn) such that for each l ∈ N we have Xk∈Al λk = 1 and zl = Xk∈Al λkxk. (1 − 2−m)cJ (xn) Lemma 5.2. Let (xn) be a bounded sequence in a Banach space X such that cJ (xn) > 0. Then there is a block sequence (zl) of (xn) such that kzlk → cJ (xn) as l → ∞ and, moreover, (4) ∞Xl=m αl ≤(cid:13)(cid:13)(cid:13) ∞Xl=m αlzl(cid:13)(cid:13)(cid:13) ≤ (1 + 2−m)cJ (xn) ∞Xl=m for all m ∈ N and all (αn) ∈ ℓ1. Proof. Set c = cJ (xn) and let cm have the meaning from (3). Since cm ր c, we can fix an increasing sequence (km) of natural numbers such that ckm > c(1 − 2−m). Next, having in mind that ck is defined as an infimum and ck ≤ c for each k ∈ N, we can find finite sets Am ⊆ N and constants λn, n ∈ Am, such that for each m ∈ N we have αl • km ≤ min Am ≤ max Am < min Am+1, • Pn∈Am λn = 1, • (cid:13)(cid:13)Pn∈Am λnxn(cid:13)(cid:13) < c(1 + 2−m). MEASURES OF WEAK NON-COMPACTNESS 11 c as m → ∞. Moreover, fix any sequence (αl) ∈ ℓ1 and m ∈ N. Then λnxn. Then clearly (zm) is a block sequence of (xn) with kzmk → Set zm =Pn∈Am αlzl(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≤ (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ∞Xl=m ∞Xl=m =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) αlzl(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) αlλnxn(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≥ (1 − 2−m)c ∞Xl=m Xn∈Al ∞Xl=m = (1 − 2−m)c (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ∞Xl=m αlkzlk ≤ (1 + 2−m)c αl, and αl , ∞Xl=m ∞Xl=m Xn∈Al αl · λn which completes the proof. (cid:3) It is elementary but useful that if one passes to a subsequence (xnk ) of (xn) then cJ (xnk ) ≥ cJ (xn); in particular it makes sense to define (5) cJ (xnk ). cJ (xn) = sup nk A diagonal argument shows that every bounded sequence (xn) admits a subsequence (xnk ) which is cJ -stable in the sense that cJ (xnk ) = cJ (xnk ). If one passes to a block sequence (zn) of (xn) then cJ (zn) ≥ cJ (xn). Lemma 5.3. Let (xn) be a bounded sequence in a Banach space X. Let (yn) be a weak Cauchy sequence in X. Then for each ε ∈ (0, 1) there is a subsequence (nk) such that (xnk ) and (xnk + ynk ) are cJ -stable and (6) (7) (1 − ε)cJ (xn) ≤ cJ (xnk ) ≤ cJ (xn). cJ (xnk + ynk ) = cJ (xnk ). If cJ (xn) = 0 then by Rosenthal's ℓ1- Proof. Let (xn) be a bounded sequence. theorem (xn) contains a weak Cauchy subsequence which remains weak Cauchy when a weak Cauchy sequence is added; hence the cJ -value of the sum is still 0 and we are done in this case. Assume now that cJ (xn) > 0. Let (yn) be weakly Cauchy and let 0 < ε < 1. Set zn = xn + yn. Choose a subsequence (xnk ) such that (6) holds. Passing to another subsequence, if necessary, we assume further that (xnk ) and (znk ) are cJ -stable. Write c = cJ (xnk ) for short. Note that c ≥ (1 − ε)cJ (xn) > 0. In particular we may assume that (xnk ) is an ℓ1-sequence (by omitting, if necessary, finitely many xnk ). Passing to another subsequence again, if necessary, we also get that cJ (znk ) > 0 because otherwise (znk ) would have no ℓ1-subsequence and would therefore contain a weak Cauchy subsequence (znkl ) by Rosenthal's theorem in which case xnkl would form a weak Cauchy sequence which is not possible. By Lemma 5.2 take blocks x(1) n of the sequence (xnk ) such that = znkl − ynkl (8) (1 − 2−m)c ∞Xl=m αl ≤(cid:13)(cid:13)(cid:13) ∞Xl=m αlx(1) l (cid:13)(cid:13)(cid:13) ≤ (1 + 2−m)c ∞Xl=m αl 12 J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER for all m ∈ N and (αl) ∈ ℓ1. For the corresponding blocks z(1) cJ (znk ) > 0. By Lemma 5.2 take blocks z(2) n ) such that n of (z(1) n we have cJ (z(1) n ) ≥ (9) n ) αl αlz(2) ∞Xl=m ∞Xl=m (1 − 2−m)cJ (z(1) n ) l (cid:13)(cid:13)(cid:13) ≤ (1 + 2−m)cJ (z(1) αl ≤(cid:13)(cid:13)(cid:13) ∞Xl=m Cauchy subsequence [for, if un = Pk∈An Pk∈An λk = 1) and if (unm) is a subsequence such that λ = limmPk∈Anm limn x∗(yn): x∗(unm ) − λµ = Pk∈Anm maxk∈Anm x∗(yk) − µ + (Pk∈Anm for all m ∈ N and (αl) ∈ ℓ1. Note that every block sequence of (yn) admits a weak λkyk, (with finite pairwise disjoint An, λk exists then, given x∗ ∈ X ∗, the sequence x∗(unm) converges to λµ where µ = λk) − λ)µ ≤ λk) − λµ → 0 as m → ∞]. Hence the blocks nk ) whose differ- n2k ) are weakly null and, by the Mazur theorem, there are blocks λk(x∗(yk) − µ) + ((Pk∈Anm n admit a weak Cauchy subsequence (y(2) 2 (y(2) of (y(2) n2k+1 − y(2) n which correspond to z(2) y(2) ences 1 y(3) n (which are blocks of the x(1) right hand side in n ) such that (cid:13)(cid:13)(cid:13)y(3) n (cid:13)(cid:13)(cid:13) → 0. Note that for the corresponding blocks x(3) n (cid:13)(cid:13)(cid:13) → c by (8). Hence the norm of the n ) we have(cid:13)(cid:13)(cid:13)x(3) n z(3) n = x(3) n + y(3) n converges to c while the norm of the left hand side converges to cJ (z(1) Thus c = cJ (z(1) n ) ≥ cJ (xnk + ynk ). n ). This shows that cJ (xnk ) = cJ (z(1) Apply now what has been shown so far to the bounded sequence (znk ) and the ) and (ynkl ) weak Cauchy sequence (−ynk ) in order to obtain subsequences (znkl such that, by cJ -stability of (xnk ) and (znk ), n ) by (9). cJ (xnk ) = cJ (xnkl ) = cJ (znkl + (−ynkl )) ≤ cJ (znkl ) = cJ (znk ). Now (7) follows. (cid:3) Proof of Proposition 5.1. (a) The proof of the implication (I)⇒(II) is almost im- mediate from the definition of wckX . For, given ε > 0, it is enough to set C = {xn} where (xn) is a sequence in A such that dist(x∗∗, X) > wckX (A) − ε for all weak∗- cluster points x∗∗ of (xn) which leads to ω(C′) > wckX (A) − ε = ω(A) − ε for all infinite C′ ⊆ C. (b) Suppose that X has the subsequence splitting property and (II) is satisfied. Let A be bounded in X. We need to prove that wckX (A) ≥ ω(A). If ω(A) = 0, the inequality is trivial. So, assume ω(A) > 0 and fix an arbitrary ε ∈ (0, min{1, ω(A)}). Let C = {xn} be a countable subset of A provided by (II). In particular, ω({xnk}) > 0 for every subsequence (xnk ) of (xn). Take a subsequence (xnk ) of (xn) such that cJ (xn)− ε < cJ (xnk ) ≤ cJ (xn). Take another subsequence (still denoted by (xnk )) and write xnk = yk + zk according to the subsequence splitting property where (yk) converges weakly and kP αkzkk ≥ (1 − ε/2)Pαkkzkk for all (αk) ∈ ℓ1. Note that the sequence (zk) is bounded, so without loss of generality λ := limkzkk exists. We have that λ > 0 because otherwise (xnk ) would be weakly convergent which would contradict ω({xnk}) > 0. Hence, without loss of generality λ(1 − ε/2) < kzkk < λ(1 + ε) for k ∈ N. Since the MEASURES OF WEAK NON-COMPACTNESS 13 sequence (yk) forms a relatively weakly compact set, by the assumptions we have ω(A) − ε < ω({xnk , k ∈ N}) ≤ sup Moreover, cJ (zk) ≥ (1 − ε/2)λ(1 − ε/2) > (1 − ε)λ. k kzkk ≤ λ(1 + ε). We apply Lemma 5.3 and obtain further subsequences (still denoted by (xnk ), (yk), (zk)). Then (7) yields cJ (xnk ) = cJ (zk) ≥ λ(1 − ε) ≥ 1 − ε 1 + ε (ω(A) − ε). By [51, Lemma 5(ii)] dist(clustw∗(xnk ), X) ≥ cJ (xnk ), hence 1 − ε 1 + ε wckX (A) ≥ dist(clustw∗(xnk ), X) ≥ cJ (xnk ) ≥ (ω(A) − ε). Since ε is arbitrary we are done. (c) Assume X is L-embedded and satisfies (I). Fix a bounded sequence (xn) in X. If cJ (xn) = 0 then by Rosenthal's ℓ1-theorem, (xn) contains a weak Cauchy subsequence (xnk ) which, by weak sequential completeness of L-embedded spaces [37, Theorem IV.2.2] converges weakly. The case is settled by putting yk = xnk and zk = 0. Let us now suppose that c := cJ (xn) > 0. In order to prove the subsequence splitting property in this case it is enough to produce, given ε > 0, a decomposition xnk = yk + zk where (yk) converges weakly and where (10) (1 − ε)cXαk ≤ (cid:13)(cid:13)(cid:13)X αkzk(cid:13)(cid:13)(cid:13) ≤ (1 + ε)cXαk for all (αk) ∈ ℓ1. Set A = {xn; n ∈ N}. First we claim that wckX (A) = c. The inequality "≥" follows from [51, Lemma 5(ii)]. For the other inequality note that for any η > 0, the construction (which works also for complex scalars) leading to [51, formula (8)] yields an x ∈ X and a subsequence (xnk ) such that cJ (xnk − x) > (1 − η) wckX (A). (Note that here the assumption that X is L-embedded is used). It follows cJ (xnk ) > (1 − η) wckX (A) (cf. [56, Proposition 4.2] or Lemma 5.3 for constant yn = −x). This proves "≤" and the claim. Let (xnk ) be a subsequence such that cJ (xnk ) > (1 − ε 2 )c. Since ω(A) = c by (I) and the claim, there is a weakly compact set K of X such that A ⊆ K + (1 + ε)cBX . Choose a sequence (yk) in K (which can be supposed to converge weakly) such that xnk = yk + zk with kzkk ≤ (1 + ε)c. The latter inequality implies the second one of (10). By Lemma 5.3 we pass to subsequences (still denoted by (xnk ), (yk), (zk)) in order to get cJ (zk) = cJ (xnk ). Now, by the definition of cJ (zk), if we omit at most finitely many zk then we have (cid:13)(cid:13)(cid:13)X αkzk(cid:13)(cid:13)(cid:13) ≥ (1 − ε 2 )cJ (zk)Xαk ≥ (1 − ε 2 )2cXαk (cid:3) and the first inequality of (10) follows. Proposition 5.1(b) will play a key role in the proof of Theorem 4.1 in Section 8. Let us remark that some special cases may be proved already now, as there are some spaces which are easily seen to satisfy (II). 14 J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER Recall that a Banach space X is said to be weakly compactly generated (shortly WCG) if there is a weakly compact subset K ⊆ X with spanK = X. Further, X is called strongly WCG (see [73]) if there is a weakly compact set K ⊆ X such that ∀L ⊆ X weakly compact∀ε > 0 ∃n ∈ N : L ⊆ nK + εBX . By the Krein theorem we may assume without loss of generality that K is absolutely convex. We have the following easy lemma Lemma 5.4. Let X be a strongly WCG Banach space. Then for any bounded set A ⊆ X there is a countable subset C ⊆ A such that ω(C′) = ω(A) for each C′ ⊆ C infinite. Proof. Let K ⊆ X be a weakly compact set witnessing the strong WCG property. Then clearly ω(A) = inf n∈Nbd(A, nK) for any bounded set A ⊆ X. Given any bounded set A ⊆ X, we can find a sequence (xn) in A such that dist(xn, nK) > ω(A) − 1 It is enough to take n . C = {xn; n ∈ N}. (cid:3) Since L1(µ) is strongly WCG for any finite measure µ (by [73, Example 2.3(d)]) and L1 spaces have the subsequence splitting property, then the essential part of [46, Proposition 7.1] follows immediately from the previous lemma and Proposition 5.1. Further, preduals of σ-finite von Neumann algebras and, more generally, preduals of σ-finite JBW∗-algebras are seen to be strongly WCG by combining [73, Theorem 2.1], [43, Appendice 6, Lemma 2] (cf. Lemma 7.3 below), and [69, Theorem D.21] (see Proposition 7.9 below). Hence the validity of Theorem 4.1 for σ-finite JBW∗- algebras easily follows. The case of σ-finite JBW∗-triples is more complicated, see Theorem 9.3 below. 6. Tripotents and projections A projection in a C∗-algebra A is a self-adjoint idempotent, i.e., an element p ∈ A satisfying p∗ = p = p2. If A is represented as a C∗-subalgebra of B(H), then p ∈ A is a projection if and only if it is, when viewed as an operator on H, an orthogonal projection. Similarly, a projection in a JB∗-algebra A is an element p ∈ A satisfying p∗ = p and p◦ p = p; in particular, p is positive (by Lemma 3.1). Two projections p, q ∈ A are said to be orthogonal if p ◦ q = 0 or, equivalently, if p + q is also a projection. In case A is a C∗-algebra, this is just equivalent to pq = 0, which means that the ranges of the projections are orthogonal subspaces of H. We shall consider the usual partial order on the set of projections in a JB∗-algebra defined by p ≤ q if q − p is a projection. In a C∗-algebra this order coincides with the standard one. In a JB∗-triple there is no natural notion of projections, but tripotents play a similar role. As a motivation for the latter notion, suppose A is a C∗-algebra regarded as a JB∗-triple with respect to the triple product {a, b, c} = 1 2 (ab∗c+cb∗a). It is well known that an element u in A is a partial isometry (i.e., u∗u is a projection, or equivalently, uu∗ is a projection) if and only if {u, u, u} = uu∗u = u. Given a MEASURES OF WEAK NON-COMPACTNESS 15 partial isometry u in a C∗-algebra A, the elements pi = u∗u and pf = uu∗ are called the initial and final projection of u, respectively. An element u of a JB∗-triple M is called a tripotent if {u, u, u} = u. If M is a JB∗-algebra, then u ∈ M is a tripotent if and only if which cannot be simplified. Note that the projections of a JB∗-algebra A are precisely those tripotents in A which are positive elements. u = 2(u ◦ u∗) ◦ u − (u ◦ u) ◦ u∗, 6.1. Peirce decomposition. From a purely algebraic point of view, a complex linear space E equipped with a triple product {., ., .} : E3 → E, which is symmetric and bilinear in the outer variables and conjugate-linear in the middle one satisfying the axiom (JB∗-1) in the definition of JB∗-triple is called a complex Jordan triple system. An element u in a Jordan triple system E is a tripotent if {u, u, u} = u. Each tripotent u in a complex Jordan triple system E induces a decomposition of E in terms of the eigenspaces of the mapping L(e, e) (this purely algebraic result can be seen, for example, in [58, 1.3 in page 7] or [19, page 32]). In our concrete setting, the algebraic structure assures that for each tripotent u in a JB∗-triple M , the eigenvalues of the operator L(u, u) are contained in the set {0, 1 2 , 1}. If u 6= 0, then 1 is always an eigenvalue, the witnessing eigenvector is u. For j = 0, 1, 2 we shall denote by Mj(u) the eigenspace of M with respect to the eigenvalue j 2 . Then M is the direct sum M0(u) ⊕ M1(u) ⊕ M2(u) of the three eigenspaces of L(u, u); this decomposition is called the Peirce decomposition of M with respect to u. The canonical projection, Pj(u), of M onto Mj(u) is called the (j-)Peirce projection associated with u. Peirce projections are explicitly determined by the following formulae: P2(u) = L(u, u)(2L(u, u) − idM ) = Q(u)2, P1(u) = 4L(u, u)(idM −L(u, u)) = 2(L(u, u) − Q(u)2), P0(u) = (idM −L(u, u))(idM −2L(u, u)) = idM −2L(u, u) + Q(u)2 where the quadratic operator Q(u) : M → M is defined by Q(u)(x) = {u, x, u} (compare [58, 1.3 in page 7] or [19, page 7]). In the setting of JB∗-triples, Peirce projections are all contractive (see [29, Corollary 1.2] or [19, 3.2.1]). In case M is a C∗-algebra and u ∈ M is a tripotent (i.e., a partial isometry with initial projection pi and final projection pf ), the Peirce projections are given by the following expressions: P2(u)x = pf xpi, P1(u)x = pf x(1− pi) + (1− pf )xpi, P0(u)x = (1− pf )x(1− pi), where x runs through M . If i, j, k ∈ {0, 1, 2}, then the so-called Peirce arithmetic or Peirce multiplication rules say that (11)  {Mi(u), Mj(u), Mk(u)} ⊆ Mi−j+k(u), {Mi(u), Mj(u), Mk(u)} = 0, {M2(u), M0(u), M} = {M0(u), M2(u), M} = 0, if i − j + k ∈ {0, 1, 2}, if i − j + k /∈ {0, 1, 2}, (see [58, (1.20) -- (1.22) in pages 7-8] or [19, Theorem 1.2.44]). It follows immediately from the Peirce multiplication rules that Mj(u) is a JB∗- subtriple for j = 0, 1, 2. In the case j = 2 something more can be said. In this 16 J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER case M2(u) is even a unital JB∗-algebra with respect to the product and involution given by a ◦ b = {a, u, b} , a∗ = {u, a, u} , a, b ∈ M2(u), respectively (cf.[19, §1.2 and Remark 3.2.2] or [16, Corollary 4.2.30]). It is further known that u is the unit of this JB∗-algebra. (If we wish to stress that the operations are with respect to u, we write a ◦u b and a∗u .) 6.2. Complete tripotents. A tripotent u in a JB∗-triple M is called complete if M0(u) = {0} and unitary if M = M2(u) (that is if {u, u, x} = x for all x ∈ M ). It follows from the above structure results that M admits a unitary tripotent if and only if it admits a structure of unital JB∗-algebra (cf. [16, Theorem 4.1.55]). Further, if M is a unital JB∗-algebra, then u ∈ M is a unitary tripotent if and only if it is a unitary element, i.e., an element satisfying that u is invertible in the Jordan sense (i.e., there exists a unique element b = u−1 in M such that b ◦ u = 1 and u2 ◦ b = u) and u−1 = u∗ (compare [16, §4.1.1]). In a C∗-algebra this reduces to u∗u = uu∗ = 1. A complete tripotent need not be unitary, even in the C∗-algebra case. Indeed, if u ∈ B(H) is a partial isometry whose initial projection is the identity on H but its final projection is strictly smaller than the identity on H (or vice versa), i.e., if u∗u = 1 6= uu∗ (or vice versa), then u is a complete non-unitary tripotent. This is not the unique possibility, but it is an important case as witnessed by the forthcoming lemmata. We observe that the extreme points of the closed unit ball of a C∗-algebra A are precisely the complete partial isometries (tripotents) of A (see [48, Theorem 1] or [77, Theorem I.10.2]). The same result remains true for the closed unit ball of a JB∗-triple E, that is, the complete tripotents in E are the extreme points of its closed unit ball (cf. [13, Lemma 4.1] and [55, Proposition 3.5] or [19, Theorem 3.2.3]). be the vector defined by τ (ξ) =Pk∈Λ hξ, ξkiξk. Then τ is a conjugation on H and, We recall that a conjugation on a complex Banach space X is a conjugate-linear isometry τ : X → X of period-2 (i.e., τ 2 = IdX ). Let H be a complex Hilbert space, and let us fix an orthonormal basis (ξk)k∈Λ in H. Given ξ ∈ H, let τ (ξ) ∈ H moreover, any conjugation is of that form (with a properly chosen orthonormal basis). Lemma 6.1. Let M be a unital C∗-algebra, and let u ∈ M be a complete tripotent. Then there exist a complex Hilbert space H and an isometric unital Jordan ∗- monomorphism ψ : M → B(H) such that ψ(u)∗ψ(u) = 1. Proof. By applying the Gelfand-Naimark-Segal construction, we can find a family of complex Hilbert spaces {Hi}i∈I and irreducible representations Φi : M → B(Hi), such that Φi : M → B(Hi) ⊆ B( Hi) ℓ2Mi∈I Φ =Mi∈I ℓ∞Mi∈I is an isometric ∗-monomorphism (we can consider, for example, the atomic repre- sentation of M [59, 4.3.7], where the family I is precisely the set of all pure states of M and each Φi is the irreducible representation associated with the pure state i [59, Theorem 3.13.2]). MEASURES OF WEAK NON-COMPACTNESS 17 Fix i ∈ I. Since u is tripotent in M , Φi(u) is a tripotent in Φi(M ) as well. Moreover, u is complete, i.e., hence (1 − uu∗)a(1 − u∗u) = 0 for a ∈ M, (1 − Φi(u)Φi(u)∗)x(1 − Φi(u)∗Φi(u)) = 0 for x ∈ Φi(M ). Since Φi is irreducible, its range is weak∗-dense in B(Hi), thus the above for- mula holds for all x ∈ B(Hi). In other words, Φi(u) is a complete tripotent in B(Hi). Having in mind that B(Hi) is a factor, it follows that Φi(u)∗Φi(u) = 1 or Φi(u)Φi(u)∗ = 1 (this follows e.g. from [77, Lemma V.1.7] applied to the projections 1 − Φi(u)∗Φi(u) and 1 − Φi(u)Φi(u)∗). Let I1 := {j ∈ I : Φj(u)∗Φj(u) = 1j} and I2 := I \ I1. For each j ∈ I2, we can find a ∗-anti-homomorphism Ψj : B(Hj ) → B(Hj ) (con- sider, for example a transposition on B(Hj ) defined by Ψj(a) := τ a∗τ , where τ is the conjugation on Hj described before the statement of the lemma). For j ∈ I1, j B(Hj ) → j B(Hj ). Clearly Ψ is a Jordan ∗-isomorphism. By construction we have j B(Hj ). Finally, we can embed the C∗- j Hj) via a ∗-monomorphism θ, and the Jordan j Hj) satisfies the desired property. (cid:3) Ψj will stand for the identity on B(Hj ). Let Ψ = Lj∈I Ψj : Lℓ∞ Lℓ∞ Ψ(Φ(u))∗Ψ(Φ(u)) = 1, the identity inLℓ∞ algebra Lℓ∞ ∗-monomorphism ψ = θ◦Ψ◦Φ : M → B(Lℓ2 j B(Hj ) inside B(Lℓ2 We continue with a technical result relating complete tripotents in a JC∗-algebra A with the complete tripotents in the C∗-algebra generated by A. Lemma 6.2. Let M be a unital JB∗-algebra. Let u be a complete tripotent in M , and let N denote the JB∗-subalgebra of M generated by u and the unit element. Then N is a JC∗-subalgebra of some C∗-algebra B, and u is a complete tripotent in the C∗-subalgebra of B generated by N . Proof. The first statement follows from Lemma 3.1(i), so fix a C∗-algebra B coni- taing N as a JC∗-subalgebra. Let A be the C∗-subalgebra of B generated by N . Furhter, let 1 denote the unit of M (which belongs to N ). Then for any x ∈ N we have 1x1 = {1, x, 1} = x, hence x = 1x = x1. Since A is generated by N , it follows that 1 is the unit of A. Clearly u is a tripotent (hence a partial isometry) in A, so its Peirce-0 projection is given by P0(u)(a) = (1 − uu∗)a(1 − u∗u), a ∈ A. To prove that u is complete in A it is enough to show that P0(u) vanishes on all the (associative) monomials in u and u∗. To this end, we will consider the formal degree of such monomials in the obvious way (1 is the unique monomial of degree 0, monomials of degree 1 are u and u∗, monomials of degree 2 are u2, (u∗)2, uu∗ and u∗u etc.). Since u is complete in N and 1, u, u∗ ∈ N , we deduce that P0(u) vanishes on monomials of degree 0 or 1. Assume that n ∈ N and P0(u) vanishes on all the monomials of degree at most n. Let a be a monomial of degree n+ 1. If a = un+1 or a = (u∗)n+1, then a ∈ N , hence P0(u)(a) = 0. Otherwise there are two monomials 18 J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER b, c with deg(b) + deg(c) = n + 1 such that either a = buu∗c or a = bu∗uc. If the first possibility takes place, then P0(u)(a) = (1 − uu∗)buu∗c(1 − u∗u) = (1 − uu∗)bc(1 − u∗u) − (1 − uu∗)b(1 − uu∗)c(1 − u∗u) = P0(u)(bc) − (1 − uu∗)bP0(u)(c) = 0 by the induction hypothesis. The second case is analogous. (cid:3) Lemma 6.3. Let M be a unital JB∗-algebra and u a complete tripotent in M . Let N be the closed unital Jordan ∗-subalgbebra of M generated by u. Then there is a unital Jordan ∗-monomorphism ψ : N → B(H), where H is a complex Hilbert space, such that ψ(u)∗ψ(u) = 1. Proof. By Lemma 6.2 we can assume that N is a unital JC∗-subalgebra of some C∗- algebra A such that u is a complete tripotent in A as well. The desired conclusion follows now from Lemma 6.1. (cid:3) 6.3. Orders on tripotents. There is a natural partial order on tripotents which we recall below. We start by analyzing a coarser ordering (see the subsequent Proposition 6.5) which will be useful in the next section. We start by the following easy lemma. Lemma 6.4. Let M be a JB∗-triple, e ∈ M a tripotent and x ∈ Mj(e) for some j ∈ {0, 1, 2}. Then the following assertions hold. (a) The operators L(e, e) and L(x, x) commute. (b) The operator L(x, x) commutes with the Peirce projections induced by e. (c) If x is moreover a tripotent, then the Peirce projections induced by x commute with the Peirce projections induced by e. Proof. (a) Let j ∈ {0, 1, 2} be such that x ∈ Mj(e). It means that L(e, e)x = j By the Jordan identity we deduce that given any y ∈ M we have 2 x. L(e, e)L(x, x)y = L(e, e){x, x, y} = {L(e, e)x, x, y} − {x, L(e, e)x, y} + {x, x, L(e, e)y} =(cid:26) j 2 x, x, y(cid:27) −(cid:26)x, j 2 x, y(cid:27) + {x, x, L(e, e)y} = L(x, x)L(e, e)y. This completes the proof of (a). Assertions (b) and (c) follow from (a) using the formulae for Peirce projections. (cid:3) Let us remark that statement (c) was already established by G. Horn in [40, (1.10)]. The previous result and its proof are included here for completeness reasons. A coarser ordering on the set of tripotents is considered in our next result. Proposition 6.5. Let M be a JB∗-triple and e, u be two tripotents in M . The following assertions are equivalent: (1) u ∈ M2(e); (2) P2(u)P2(e) = P2(e)P2(u) = P2(u), P1(u)P1(e) = P1(e)P1(u) and P0(u)P0(e) = P0(e)P0(u) = P0(e); (3) M2(u) ⊆ M2(e) and M0(e) ⊆ M0(u); (4) M2(u) ⊆ M2(e). MEASURES OF WEAK NON-COMPACTNESS 19 Proof. (1)⇒(2) Assume that u ∈ M2(e). By Lemma 6.4 the Peirce projections induced by u commute with the Peirce projections induced by e. Further, given x ∈ M the Peirce rules (11) yield {u, P1(e)x, u} = {u, P0(e)x, u} = 0, thus It follows that Q(u) = Q(u)P2(e), so {u, x, u} = {u, P2(e)x, u} . P2(u) = Q(u)2 = Q(u)2P2(e) = P2(u)P2(e). Let x ∈ M0(e), another application of Peirce arithmetic yields {u, u, x} ∈ {M2(e), M2(e), M0(e)} = {0}, so M0(e) ⊆ M0(u), and hence P0(u)P0(e) = P0(e). The implications (2)⇒(3)⇒(4)⇒(1) are obvious. (cid:3) Proposition 6.6. Let M be a JB∗-triple, and let e, u be two tripotents in M . The following assertions are equivalent: (1) u ∈ M2(e) and e ∈ M2(u); (2) M2(e) = M2(u); (3) The Peirce decompositions induced by e and u coincide (i.e., Mj(e) = Mj(u) for all j = 0, 1, 2). Proof. The implications (3)⇒(2)⇒(1) are obvious. (1)⇒(3) Assume u ∈ M2(e) and e ∈ M2(u). It follows from Proposition 6.5 (the implication (1)⇒(2)) that P2(e) = P2(u) and P0(e) = P0(u). Hence also P1(u) = P1(e). (cid:3) Proposition 6.7. Let M be a JB∗-triple, and let e, u be two tripotents in M . The following assertions are equivalent: (1) u ∈ M0(e); (2) e ∈ M0(u); (3) M2(u) ⊆ M0(e) and M2(e) ⊆ M0(u); (4) P2(u)P0(e) = P0(e)P2(u) = P2(u) and P0(u)P2(e) = P2(e)P0(u) = P2(e). Proof. (1)⇒(2) Assume u ∈ M0(e). Then {u, u, e} = {e, u, u} = 0 by the Peirce arithmetics (note that e ∈ M2(e) and u ∈ M0(e)). Hence e ∈ M0(u). (2)⇒(1) follows by symmetry. (1)⇒(4) Assume u ∈ M0(e). By the already proved implications we know that also e ∈ M0(u). It follows from Peirce arithmetic that M2(u) ⊆ M0(e) and M2(e) ⊆ M0(u). Therefore P2(u) = P0(e)P2(u) and P2(e) = P0(u)P2(e). Since, by Peirce arithmetics, we also have {u, M2(e), u} = {u, M1(e), u} = {0}, and P2(u) = Q(u)2, we deduce that P2(u)Pj (e) = 0, for j = 1, 2. Therefore, P2(u) = P2(u)P0(e), and similarly P2(e) = P2(e)P0(u). The implications (4)⇒(3)⇒(2) are obvious. (cid:3) Remark 6.8. Propositions 6.5 and 6.6 show that the Peirce subspace M2(e) de- termines the whole Peirce decomposition. This is not the case for M0(e) as there may exist complete tripotents with different Peirce decompositions. 20 J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER Tripotents u, e ∈ M satisfying any of the equivalent conditions in Proposition 6.7 are called orthogonal (u ⊥ e in short). In particular, e ± u are again tripotents. It is actually known that given two tripotents e, u ∈ M , then e ⊥ u if and only if e ± u is a tripotent (cf. [44, Lemma 3.6]). We are now in position to recall the natural partial order on the set of tripo- tents. If M is a JB∗-triple and e, u are two tripotents in M , we say that u ≤ e if e − u is a tripotent orthogonal to u. This order is finer than the one derived from Proposition 6.5 as can be seen from the last of the characterizations in the follow- ing proposition (originally due to Y. Friedman and B. Russo [29, Corollary 1.7], compare also [19, Proposition 1.2.43], [17, Corollary 5.10.56]). Proposition 6.9. (essentially [29, Corollary 1.7]) Let M be a JB∗-triple and e, u ∈ M two tripotents. Then the following assertions are equivalent: • u ≤ e; • u = P2(u)e; • u = {u, e, u}; • u is a projection in the JB∗-algebra M2(e); • M2(u) is a JB∗-subalgebra of M2(e). 6.4. More on the Peirce-2 subspace. Our next result gathers some properties of the Peirce-2 subspace associated with a tripotent. Most of the statements are part of the folklore in the theory of JB∗-triples, we include here the properties and basic references for completeness reasons. Lemma 6.10. Let M be a JB∗-triple and let e ∈ M be a tripotent. Consider M2(e) equipped with its structure of JB∗-algebra. (a) Assume that v ∈ M is a tripotent such that e ≤ v. Then for any a, b ∈ M we have P2(e){a, b, v} = {P2(e)a, P2(e)b, e} + {P1(e)a, P1(e)b, e} , P1(e){a, b, v} = {P1(e)a, P2(e)b, e} + {P0(e)a, P1(e)b, e} + {P2(e)a, P1(e)b, v − e} + {P1(e)a, P0(e)b, v − e} , P0(e){a, b, v} = {P0(e)a, P0(e)b, v − e} + {P1(e)a, P1(e)b, v − e} , in particular P2(e){a, b, e} = {P2(e)a, P2(e)b, e} + {P1(e)a, P1(e)b, e} , P1(e){a, b, e} = {P1(e)a, P2(e)b, e} + {P0(e)a, P1(e)b, e} , P0(e){a, b, e} = 0. (b) Assume j ∈ {1, 2} and a, b ∈ Mj(e). Then {a, b, e} ∈ M2(e) and {a, b, e}∗ = {b, a, e} . (c) Assume a, b ∈ M2(e). Then a ◦ b∗ = {a, b, e}. (d) If a ∈ M2(e) ∪ M1(e), then {a, a, e} is a positive element of the JB∗-algebra M2(e). Moreover, {a, a, e} = 0 if and only if a = 0. Proof. Fix a tripotent v ∈ M with e ≤ v. Then P2(e)v = e, P1(e)v = 0, and P0(e)v = v − e. Hence the Peirce arithmetic implies that, given x ∈ Mj(e) and (12) and (13) {x, y, e} {x, y, v − e} ∈ M2(e) ∈ M1(e) = 0 if j = k = 2 or j = k = 1, if j = 1, k = 2 or j = 0, k = 1, otherwise, ∈ M0(e) ∈ M1(e) = 0 if j = k = 0 or j = k = 1, if j = 1, k = 0 or j = 2, k = 1, otherwise. MEASURES OF WEAK NON-COMPACTNESS 21 y ∈ Mk(e) for some j, k ∈ {0, 1, 2}, we have Assertion (a) now follows from (12) and (13). Further, the first statement of asser- tion (b) follows from (12). Let us continue by proving the second statement from (b). We deduce from the Jordan identity, the definition of the involution in M2(e), and the fact that {b, a, e} ∈ M2(e), that {b, a, e} = L(b, a)e = L(b, a){e, e, e} = {L(b, a)e, e, e} − {e, L(a, b)e, e} + {e, e, L(b, a)e} = 2L(e, e){b, a, e} − (L(a, b)e)∗ = 2 {b, a, e} − {a, b, e}∗ . (c) The Peirce-2 subspace M2(e) is a JB∗-algebra, and hence it is a JB∗-triple with respect to the triple product given by {a, b, c}1 = (a ◦ b∗) ◦ c + (c ◦ b∗) ◦ It is also a JB∗-triple with the triple product inherited from a − (a ◦ c) ◦ b∗. M . Since the identity mapping from (M2(e),{., ., .}1) onto (M2(e),{., ., .}) is a surjective isometry, it follows from Kaup's Riemann mapping theorem (see [52, Proposition 5.5] or [19, Theorem 3.1.7]) that {a, b, c}1 = {a, b, c}, for all a, b, c ∈ M2(e). Consequently, {a, b, e} = {a, b, e}1 = (a◦b∗)◦e+(e◦b∗)◦a−(a◦e)◦b∗ = a◦b∗, because e is the unit of M2(e). (d) If a ∈ M2(e), then {a, a, e} = a ◦ a∗ by (c), hence the assertion follows from Lemma 3.1(ii). The case a ∈ M1(e) is covered by [29, Lemma 1.5(b)], and both cases (a ∈ M1(e) and a ∈ M2(e)) are fully studied in [63] (see also [16, Proposition 4.2.32]), where a simple proof based on the axioms of JB∗-triples can be found. (cid:3) When we combine the previous result with the properties of the functionals in the dual space of a JB∗-triple we get additional properties. We recall that a functional ϕ in the dual space of a JB∗-algebra M is called faithful if ϕ(a) = 0 for a ≥ 0 implies a = 0. Lemma 6.11. Let M be a JB∗-triple and let e ∈ M be a tripotent. Consider M2(e) equipped with its structure of unital JB∗-algebra. Let ϕ ∈ M ∗. Then the following assertions hold: (a) kϕ ◦ (P2(e) + P0(e))k = kϕ ◦ P2(e)k + kϕ ◦ P0(e)k. Moreover, if kϕk = ϕ(e), then the following assertions are valid, too: (b) ϕ = ϕ ◦ P2(e); (c) ϕM2(e) is a positive linear functional on the JB∗-algebra M2(e); (d) The mapping (x, y) 7→ ϕ({x, y, e}), x, y ∈ M, is a positive semidefinite sesquilinear form on M , and if z ∈ M is a norm-one element satisfying ϕ(z) = kϕk then ϕ({x, y, e}) = ϕ({x, y, z}) for all x, y ∈ M ; 22 J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER (e) The formula kxke,ϕ =pϕ({x, x, e}), x ∈ M defines a pre-Hilbert seminorm on M which is zero on M0(e). If moreover ϕM2(e) is faithful, then the kernel of k·ke,ϕ is exactly M0(e). Proof. Assertions (a) and (b) are proved in [29, Lemma 1.3(b) and Proposition 1(a)], compare also [17, Lemma 5.7.11, Fact 5.10.53]. (c) Since kϕk = ϕ(e) we also have kϕ ◦ P2(e)k = kϕk =(cid:13)(cid:13)ϕM2(e)(cid:13)(cid:13) = ϕM2(e)(e). Therefore ϕM2(e) is a positive functional on the JB∗-algebra M2(e) (cf. [36, Lemma 1.2.2] or [17, Lemma 5.10.2]). (d) and (e) are consequences of (c), (b) and Lemma 6.10(a) and (d). They are also explicitly proved in [6, Proposition 1.2] and [24, Lemma 4.1]. See also [17, Proposition 5.10.60] for the JBW∗-case. (cid:3) 7. Strong∗ topology and weakly compact sets In the previous section we collected many results on projections and tripotents. However, it may happen that there are no nontrivial projections or tripotents. For example, the C∗-algebra C0(R) contains no nonzero projection or tripotent and in the unital C∗-algebra C([0, 1]) the only projections are 0 and 1 and the only nonzero tripotents are the unitary ones (which coincide with the continuous functions with values in the unit circle). The situation is different in the dual case -- in a von Neumann algebra projections form a complete lattice and their linear span is norm- dense, and, as we previously commented, any JBW∗-triple provides a rich supply of tripotents (cf. [19, Theorem 3.2.3] or [16, Theorem 4.2.34]). If M is a JBW∗-triple and u ∈ M is a tripotent, then the Peirce projections are weak∗-to-weak∗ continuous and the Peirce subspaces are weak∗-closed. This follows from the separate weak∗-to-weak∗ continuity of the triple product and the explicit formulae for the Peirce projections displayed in page 15. In particular, M2(u) is a JBW∗-algebra. 7.1. Strong∗ topology on JBW∗-triples. Assume that M is a JBW∗-triple and ϕ ∈ M∗ \ {0}. By [29, Proposition 2] (see also [17, Proposition 5.10.57]) there is a unique tripotent s(ϕ) ∈ M , called the support tripotent of ϕ, such that • ϕ = ϕ ◦ P2(s(ϕ)), • ϕM2(s(ϕ)) is a faithful positive functional on the JBW∗-algebra M2(s(ϕ)). Furthermore, kϕk = ϕ(s(ϕ)). According to the notation from Lemma 6.11, we set k·kϕ = k·ks(ϕ),ϕ . Note that if M is a JBW∗-algebra (or even a von Neumann algebra) and ϕ ∈ M∗ is a positive functional, then its support tripotent s(ϕ) is even a projection because in such a case ϕ attains its norm at a positive element. Observe that in the latter case the seminorm k·kϕ writes in the form kxkϕ =pϕ{x, x, s(ϕ)} =pϕ{x, x, 1} =pϕ(x∗ ◦ x). Introduced in [7], the strong∗ topology on M is the locally convex topology gen- erated by the seminorms k·kϕ where ϕ runs in the set M∗ \ {0}. It should be noted that in the original definition (see [7, Definition 3.1]) only norm-one functionals are considered, but both definitions obviously give the same notion. Since each MEASURES OF WEAK NON-COMPACTNESS 23 k.kϕ is a preHilbertian seminorm, it follows from the Cauchy-Schwarz inequality (cf. 6.11(d)) and the properties of the support tripotent that, with x2 = P2(s(ϕ))x, ϕ(x) = ϕ(x2) = ϕ(x2 ◦s(ϕ) s(ϕ)) = ϕ({x2, s(ϕ), s(ϕ)}) 6.10 = ϕ(P2(s(ϕ)){x, s(ϕ), s(ϕ)}) = ϕ{x, s(ϕ), s(ϕ)} ≤ kxkϕ ks(ϕ)kϕ =pkϕk kxkϕ, x ∈ M, ϕ ∈ M∗ \ {0}. Consequently, the strong∗ topology is stronger than the weak∗ topology. Given ϕ1, . . . , ϕn ∈ M∗, we shall write k.kϕ1,...,ϕn for the seminorm on M defined by kxk2 ϕ1,...,ϕn := kxk2 ϕk (x ∈ M ). nXk=1 The following lemma summarizes some known properties of the strong∗ topology. Lemma 7.1. Let M be a JBW∗-triple. (a) If M is even a JBW∗-algebra, then the strong∗ topology on M coincides with the algebra strong∗ topology, i.e., with the locally convex topology generated by seminorms ϕ ∈ M∗, ϕ ≥ 0; with the restriction to N of the strong∗ topology of M ; x 7→pϕ(x∗ ◦ x), (b) If N is a weak∗ closed subtriple of M , then the strong∗ topology of N coincides (c) If M is even a von Neumann algebra (embedded to some B(H)), the strong∗ topology coincides on bounded sets with the locally convex topology generated by the seminorms x 7→ kxξk + kx∗ξk , ξ ∈ H; (d) A linear functional ϕ : M → C is strong∗ continuous if and only if it is weak∗ continuous. Furthermore a linear mapping between JBW∗-triples is strong∗-to- strong∗ continuous if and only if it is weak∗-to-weak∗ continuous. In particular, the Peirce projections associated with a tripotent are strong∗-to-strong∗ contin- uous; (e) If s(ϕ) is complete, then kxkϕ is a norm on M . Proof. Assertion (a) is proved in [68, Proposition 3], while (b) is established in [14, COROLLARY]. Let us justify assertion (c). In [77, Definition II.2.3] the name σ-strong∗ operator topology is used for the algebra strong∗ topology in B(H). By [77, Lemma II.2.5(iii)] this topology coincides on bounded sets with the topology generated by the given seminorms. Hence we can conclude by applying (a) and (b). Statement (d) is proved in [65, Corollary 9] and [68, Corollary 3] and the com- ments before [65, Theorem 9]. (e) If s(ϕ) is a complete tripotent, then M0(s(ϕ)) = {0}, thus the statement follows from Lemma 6.11(e). (cid:3) The description of the strong∗ topology is closely related to σ-finite projections and tripotents. Recall that a projection p in a JBW∗-algebra is called σ-finite if any family of pairwise orthogonal smaller nonzero projections is at most countable. If the unit of a JBW∗-algebra is σ-finite, the respective algebra is called σ-finite. The classical definitions in von Neumann algebras are exactly the same. Let us note 24 J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER that some authors also employ the term countably decomposable to refer to σ-finite projections in a von Neumann algebra (cf. [72, Definition 2.1.8] or [49, Definition 5.5.14]). Similarly, a tripotent u in a JBW∗-triple is σ-finite if any family of pairwise orthogonal nonzero smaller tripotents is at most countable. A JBW∗-triple is itself [24, §3]). The next called σ-finite if it admits a σ-finite complete tripotent (cf. result gathers some basic facts on σ-finite tripotents. Let us recall a couple of notions. A subspace I of a JB∗-triple E is an inner ideal if {I, E, I} ⊆ I. Every inner ideal of E is a subtriple. Given a norm-one element a in a JBW∗-triple M , there exists a smallest tripotent e ∈ M satisfying that a is a positive element in the JBW∗-algebra M2(e), this tripotent is called the range tripotent of a, and it will be denoted by r(a) (see, for example, [23, comments before Lemma 3.1]). For a non-zero element b ∈ M , b kbk and we set the range tripotent of b, r(b), is defined as the range tripotent of r(0) = 0. It follows from the same reference that if M is a JBW∗-algebra and x is a non-zero positive element, then r(x) is a projection and it coincides with the range projection in [36, Lemma 4.2.6]. Lemma 7.2. (a) Let u be a tripotent in a JBW∗-triple M . Then u is σ-finite if and only if u = s(ϕ) for some norm-one functional ϕ ∈ M∗; tion if and only if it is a σ-finite tripotent; (b) Let M be a JBW∗-algebra and p ∈ M a projection. Then p is a σ-finite projec- (c) Let M be a JBW∗-algebra and p ∈ M a projection. Then p is σ-finite if and only if p = s(ϕ) for some normal state (i.e., a positive norm-one functional) ϕ ∈ M∗; σ-finite projection p ∈ M such that e ∈ M2(p). (d) Let M be a JBW∗-algebra and e ∈ M a σ-finite tripotent. Then there is a Proof. Assertion (a) is proved in [24, Theorem 3.2]. Statement (b) follows from the fact that any projection is also a tripotent, and from the property that a tripotent u is smaller than or equal to a projection p if and only if u is a projection and u ≤ p (cf. Proposition 6.9). Statement (c) is a consequence of (a). (d) Let us consider the sets S = {x ∈ M : ∃p ∈ M a σ-finite projection such that x ∈ M2(p)}, Mσ = {x ∈ M : ∃u ∈ M a σ-finite tripotent such that x ∈ M2(u)}. Clearly S ⊆ Mσ. By [11, page 667 and Theorems 4.1 and 5.1] we have Mσ = {x ∈ M : r(x) is σ-finite} is (a 1-norming Σ-subspace and) a norm-closed inner ideal of M = (M∗)∗ (see the quoted paper for definitions). Since M is a JBW∗-algebra, we get Mσ ◦ Mσ = {Mσ, 1, Mσ} ⊆ Mσ, and hence Mσ is a Jordan subalgebra of M . It can be seen easily that a tripotent u ∈ M is σ-finite if and only if u∗ is, σ = Mσ is a norm-closed JB∗-subalgebra of M, therefore M ∗ which is also hereditary in the Jordan terminology, that is, if 0 ≤ a ≤ b in M with b ∈ Mσ, then a ∈ Mσ. Let a = h + ik be an element in Mσ, where h, k ∈ (Mσ)sa. The elements h2, k2 are positive and belong to Mσ, thus h2 + k2 ∈ Mσ. Therefore, the range tripotent p = r(h2+k2) is a σ-finite projection in M . Clearly, h2, k2 ≤ h2+k2 ∈ M2(p) ⊆ Mσ, σ ⊆ Mσ, and hence M ∗ MEASURES OF WEAK NON-COMPACTNESS 25 and consequently, h, k ∈ M2(p), which implies that p◦h = h and p◦k = k. Therefore {p, p, a} = p ◦ a = a as desired. (cid:3) Lemme 2 in Appendice 6 in [43] offers a sufficient condition to guarantee the metrizability of the strong∗ topology on the bounded subsets of a JBW-algebra. We shall next adapt the result to JBW∗-algebras. Lemma 7.3. Let M be a σ-finite JBW∗-algebra. (a) M admits a faithful normal state; (b) Let ϕ be a faithful normal state on M . Then the topology induced by the norm x 7→pϕ(x∗ ◦ x) = kxkϕ, x ∈ M, coincides with the strong∗ topology on bounded sets. Proof. (a) Since 1 is a σ-finite projection, by Lemma 7.2(c) there is a positive norm- one functional ϕ ∈ M∗ (i.e., a normal state) such that s(ϕ) = 1. It follows that ϕ is faithful. (b) We know that Msa is a JBW-algebra and a real JBW∗-subtriple of M , and that ϕMsa is a faithful normal state on Msa. [43, Appendice 6, Lemme 2] implies that the strong∗ topology on the closed unit ball of Msa is metrized by the seminorm kxk2 ϕ = ϕ(x◦x) = ϕ(x2), x ∈ Msa. Let (aλ)λ be a net in BM , and a ∈ BM such that kaλ − akϕ → 0. If we write aλ = hλ + ikλ and a = h + ik with hλ, kλ, h, k ∈ BMsa, then we get the inequalities ϕ((hλ − h)2), ϕ((kλ − k)2) ≤ ϕ((hλ − h)2 + (kλ − k)2) = kaλ − ak2 ϕ. Therefore, (hλ)λ → h and (kλ)λ → k in the strong∗ topology of Msa, and by [14, COROLLARY] they also converge to the same limits with respect to the strong∗ topology of M , and consequently, (aλ)λ → a in the strong∗ topology of M . (cid:3) The previous lemma says, in particular, that the strong∗ topology is metrizable on bounded sets of a σ-finite JBW∗-algebra. The analogous statement for von Neumann algebras is proved already in [77, Proposition III.5.3]. The analogy for JBW∗-triples fails, as we will explain below (see Example 9.1). We continue now with a lemma characterizing strong∗ convergence of bounded positive nets. Lemma 7.4. Let (xν ) be a bounded net of positive elements in a JBW∗-algebra M . Then the following assertions are equivalent. strong∗ −→ 0; weak∗ −→ 0. (1) xν (2) ϕ(xν ) → 0 for each positive ϕ ∈ M∗; (3) xν Proof. The implication (1) ⇒ (3) follows from the fact that the strong∗ topology is stronger than the weak∗ one, and (3) ⇒ (2) is trivial. (2)⇒(1): By Lemma 7.1(a) the net (xν ) strong∗ converges to zero if and only if for any positive ϕ ∈ M∗ we have ϕ(x2 ν ≤ kxνk xν proves the implication (2)⇒(1). ν ) → 0. Now the double inequality 0 ≤ x2 (cid:3) 26 J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER 7.2. Order on seminorms generating the strong∗ topology. To describe the strong∗ topology it is not necessary (in some cases) to use all the defining seminorms. This is witnessed by Lemma 7.1(a) and, on bounded sets, by Lemma 7.3. In this section we investigate this feature in detail. A key result is the following proposition relating the order on seminorms with the order from Proposition 6.5. Proposition 7.5. Let M be a JBW∗-triple and ϕ, ψ ∈ M∗ \ {0}. (i) Assume s(ψ) ∈ M2(s(ϕ)) (or, equivalently, M2(s(ψ)) ⊆ M2(s(ϕ))). Then the (ii) Assume M2(s(ψ)) $ M2(s(ϕ)). Then on BM , the seminorm k·kψ is strictly seminorm k·kψ is weaker than the seminorm k·kϕ on bounded sets. weaker than the seminorm k·kϕ. To prove this proposition we will need some lemmata. The first lemma char- acterizes convergence of sequences in a fixed seminorm. Note that the same char- acterization applies to nets, but since we are comparing seminorms, sequences are enough. Lemma 7.6. Let M be a JBW∗-triple and ϕ ∈ M∗ \ {0}. Let e = s(ϕ) and let (an) be a bounded sequence in M . Then the following assertions are equivalent: (1) kankϕ → 0; (2) {P2(e)(an), P2(e)(an), e} (3) {P2(e)(an), P2(e)(an), e} (4) {P2(e)(an), P2(e)(an), e} + {P1(e)(an), P1(e)(an), e} Proof. First notice, that it does not matter whether the convergence is considered in the JBW∗-triple M or in the JBW∗-algebra M2(e) (cf. [14, COROLLARY]). The strong∗ case follows from Lemma 7.1(a), (b), the weak∗ case is obvious. strong∗ −→ 0 and {P1(e)(an), P1(e)(an), e} weak∗ −→ 0 and {P1(e)(an), P1(e)(an), e} weak∗ −→ 0. strong∗ −→ 0; weak∗ −→ 0; (1) ⇒ (2) Assume that kankϕ → 0, i.e., ϕ({an, an, e}) → 0. Since ϕ = ϕ ◦ P2(e), by Lemma 6.10(a) we have ϕ({an, an, e}) = ϕ({P2(e)(an), P2(e)(an), e} + {P1(e)(an), P1(e)(an), e}). We actually know that the elements wn = {P2(e)(an), P2(e)(an), e} and zn = {P1(e)(an), P1(e)(an), e} are positive in the JBW∗-algebra M2(e) by Lemma 6.10(d). Note that 0 ≤ w2 kwnk · wn. Since the sequence (wn) is bounded, we deduce ϕ(w2 wn strong∗ −→ 0 by Lemma 7.3(b). Similarly we get zn n ≤ n) → 0, hence strong∗ −→ 0. (2) ⇒ (3) This is clear, as the weak∗ topology is weaker than the strong∗ one. (3) ⇒ (4) This follows by the linearity of the weak∗ topology. (4) ⇒ (1) This follows from the fact that kank2 ϕ = ϕ({P2(e)(an), P2(e)(an), e} + {P1(e)(an), P1(e)(an), e}). (cid:3) The following lemma, together with the preceding one, provides a proof of as- sertion (i) of Proposition 7.5. MEASURES OF WEAK NON-COMPACTNESS 27 Lemma 7.7. Let M be a JBW∗-triple and p, u ∈ M tripotents such that u ∈ M2(p). Then for any bounded sequence (an) in M we have {P2(p)(an), P2(p)(an), p} + {P1(p)(an), P1(p)(an), p} weak∗ −→ 0 =⇒ {P2(u)(an), P2(u)(an), u} + {P1(u)(an), P1(u)(an), u} weak∗ −→ 0. Proof. Let us start by noticing that we can identify M with a JB∗-subtriple of a unital JB∗-algebra B in such a way that p is a projection in B. This is proved in the first paragraph of the proof of [15, Proposition 2.4] (using [30, Corollary 1] and [15, Lemma 2.3]). (Note that B can be assumed to be a JBW∗-algebra -- just pass to B∗∗.) Let us consider the bounded linear mapping G : B → B2(u) defined by G(x) = P2(u)(x ◦ u), Further, observe that for any x ∈ B we have (x ∈ B). {x, x, u} + {x∗, x∗, u} = (x ◦ x∗) ◦ u + (x∗ ◦ u) ◦ x − (x ◦ u) ◦ x∗ + (x ◦ x∗) ◦ u + (u ◦ x) ◦ x∗ − (x∗ ◦ u) ◦ x = 2(x ◦ x∗) ◦ u, thus P2(u)({x, x, u} + {x∗, x∗, u}) = 2P2(u)((x ◦ x∗) ◦ u) = 2G(x ◦ x∗) = 2G({x, x, 1}). Further, given any x ∈ B, we have P2(p)({x, x, 1} ◦ u) = P2(p){{x, x, 1} , 1, u} = {P2(p){x, x, 1} , P2(p)(1), u} + {P1(p){x, x, 1} , P1(p)(1), u} = {P2(p){x, x, 1} , p, u} = {P2(p){x, x, 1} , 1, u} = (P2(p){x, x, 1}) ◦ u = (P2(p){x, x, p}) ◦ u. Indeed, the first equality is obvious, the second one follows from Peirce arithmetic as u ∈ M2(p) ⊆ B2(p). In the third equality we use the facts that P2(p)(1) = p and P1(p)(1) = 0. The fourth equality follows by Peirce arithmetic using the fact that 1 − p ∈ B0(p), and thus 1 − p ⊥ u. The fifth equality is obvious and the sixth one follows from Lemma 6.10(a). Thus, for each x ∈ B we have G({x, x, 1}) = P2(u)({x, x, 1} ◦ u) 6.5(2) = P2(u)P2(p)({x, x, 1} ◦ u) = P2(u)((P2(p){x, x, p}) ◦ u) = G(P2(p){x, x, p}), so P2(u)({x, x, u} + {x∗, x∗, u}) = 2G(P2(p){x, x, p}). Let us check what happens in M . If x ∈ M , then G(x) = P2(u)(x ◦ u) = P2(u)P2(p){x, 1, u} = P2(u)({P2(p)(x), P2(p)(1), u} + {P1(p)(x), P1(p)(1), u}) = P2(u){P2(p)x, p, u} . It follows that G maps M into P2(u)(M ) = M2(u) and, moreover, G restricted to M is weak∗-to-weak∗ continuous. 28 J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER So, assume (an) is a bounded sequence in M such that {P2(p)(an), P2(p)(an), p} + {P1(p)(an), P1(p)(an), p} weak∗ −→ 0, equivalently, Since this sequence lives in M , using weak∗-to-weak∗ continuity of G we get P2(p){an, an, p} weak∗ −→ 0. G(P2(p){an, an, p}) weak∗ −→ 0, and this sequence is contained in M2(u). Note that by the above calculation and Lemma 6.10(a) we have G(P2(p){an, an, p}) = = 1 2 1 n, a∗ n, u}) P2(u)({an, an, u} + {a∗ 2(cid:16){P2(u)(an), P2(u)(an), u} + {P1(u)(an), P1(u)(an), u} n), u}(cid:17). + {P2(u)(a∗ n), u} + {P1(u)(a∗ n), P1(u)(a∗ n), P2(u)(a∗ Moreover, all the four summands in the right hand side are positive elements in the JB∗-algebra B2(u) by Lemma 6.10(d), hence their sum is positive as well. Moreover, the sum belongs to M2(u) and the first two summands as well, and thus n), u} ∈ M2(u), n), u} + {P1(u)(a∗ {P2(u)(a∗ n), P2(u)(a∗ n), P1(u)(a∗ too. Since 0 ≤ {P2(u)(an), P2(u)(an), u} + {P1(u)(an), P1(u)(an), u} ≤ {P2(u)(an), P2(u)(an), u} + {P1(u)(an), P1(u)(an), u} n), P1(u)(a∗ n), P2(u)(a∗ n), u} + {P1(u)(a∗ the equivalence (2) ⇔ (3) in Lemma 7.4(2) shows that + {P2(u)(a∗ n), u} , {P2(u)(an), P2(u)(an), u} + {P1(u)(an), P1(u)(an), u} which completes the proof. weak∗ −→ 0, (cid:3) The following lemma together with Lemma 7.6 provides the proof for assertion (ii) in Proposition 7.5. Lemma 7.8. Let M be a JBW∗-triple and e, u ∈ M two tripotents such that M2(u) $ M2(e). Then there is a bounded sequence (an) in M such that {P2(u)(an), P2(u)(an), u} strong∗ −→ 0 and {P1(u)(an), P1(u)(an), u} strong∗ −→ 0 but {P2(e)(an), P2(e)(an), e} + {P1(e)(an), P1(e)(an), e} Proof. If M0(u) ∩ M2(e) contains a nonzero element a, then strong∗ 6−→ 0. {P2(u)(a), P2(u)(a), u} = {P1(u)(a), P1(u)(a), u} = 0 but {P2(e)(a), P2(e)(a), e} + {P1(e)(a), P1(e)(a), e} = {a, a, e} 6= 0 by Lemma 6.10(d). It follows that the constant sequence an = a works. MEASURES OF WEAK NON-COMPACTNESS 29 Next assume that M0(u) ∩ M2(e) is trivial, hence u is a complete tripotent in M2(e). Since M2(e) is a JBW∗-algebra, by Lemma 7.1(b) it is enough to consider the case in which M = M2(e). We shall therefore assume that M is a JBW∗-algebra, e = 1 and u ∈ M is a complete non-unitary tripotent. Let N denote the unital JB∗-subalgebra of M generated by u. By Lemma 6.3 we can assume without loss of generality that N is a JB∗-subalgebra of B(H) for a suitable complex Hilbert space H and, moreover, u∗u = 1 in B(H). Since u is not unitary, necessarily uu∗ 6= 1. Set q = uu∗. Then q is a projection in B(H). Moreover, q ∈ N , as q = uu∗ = uu∗ + u∗u − 1 = 2u ◦ u∗ − 1. Let us define a sequence in N by x0 = 1 − q and xn = xn−1 ◦ u∗ for n ∈ N. We claim that xn = 2−n(1 − q)(u∗)n, n ∈ N ∪ {0}. Indeed, for n = 0 the equality holds. Assume that n ∈ N and the equality holds for n − 1. Then xn = xn−1 ◦ u∗ = (xn−1u∗ + u∗xn−1) = 2−n((1 − q)(u∗)n + u∗(1 − q)(u∗)n−1) 1 2 = 2−n(1 − q)(u∗)n, as obviously u∗(1 − q) = 0. Set an = 2nxn = (1− q)(u∗)n. Then (an) is a bounded sequence in N , and hence in M . Further, {an, an, 1} = an ◦ a∗ n = 1 2 (ana∗ n + a∗ nan) = = 1 2 1 2 (un(1 − q)(u∗)n + (1 − q)(u∗)nun(1 − q)) (un(1 − q)(u∗)n + (1 − q)) as u∗u = 1. Therefore, and it then follows from Lemma 7.4 that {an, an, 1} does not converge to zero in the strong∗ topology. {an, an, 1} ≥ 1 − q, Further, observe that q = P2(u)(1), hence q ∈ M2(u) and 1 − q ∈ M1(u). We claim that xn ∈ M1(u) for every n ∈ N ∪ {0}. The case, n = 0 is clear. The Peirce arithmetic yields by the induction hypothesis that xn = xn−1 ◦ u∗ = {xn−1, u, 1} = {xn−1, u, q} + {xn−1, u, 1 − q} = {xn−1, u, q} ∈ M1(u), where we used that {xn−1, u, 1 − q} ∈ M0(u) = {0}. So an ∈ M1(u) for each n ∈ N ∪ {0} as well. It follows by Lemma 6.10(a) that nu + ua∗ P2(u){an, an, u} = {an, an, u} = (ana∗ nan) 1 2 = = 1 2 1 2 ((1 − q)(u∗)nun(1 − q)u + uun(1 − q)(1 − q)(u∗)n) un+1(1 − q)(u∗)n 30 J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER as (1 − q)u = 0. We shall show that the sequence 1 converges to zero. 2 un+1(1 − q)(u∗)n strong∗ To this end note that un is a partial isometry for each n ∈ N and, moreover, (un)∗un = 1. Thus its final projection qn = un(un)∗ belongs to N . Let Y = [n∈N ker((un)∗). Then Y is a closed subspace of H, and lim n Furthermore, Y ⊥ = \n∈N (ker((un)∗))⊥ = \n∈N (u∗)n(ξ) = 0 for each ξ ∈ Y . (ker qn)⊥ = \n∈N qn(H). Thus for any ξ ∈ Y ⊥ and n ∈ N we have ξ = qn(ξ) and so (1 − q)(u∗)n(ξ) = (1 − q)(u∗)nqn+1(ξ) = (1 − q)(u∗)nun+1(u∗)n+1(ξ) = (1 − q)u(u∗)n+1(ξ) = (1 − q)q(u∗)n(ξ) = 0. It follows that the sequence (1 − q)(u∗)n SOT converges to zero, hence clearly ( 1 2 un+1(1 − q)(u∗)n) strong∗ converges to zero. This completes the proof. 7.3. Weakly compact sets in the predual of a JBW∗-triple. There is a close connection of the strong∗ topology, the generating seminorms and the weakly compact subsets of the predual. It is witnessed, for example, by the following proposition which is proved in [69, Theorem D.21], see also [17, Theorem 5.10.138]. (cid:3) Proposition 7.9. Let M be a JBW∗-triple. The strong∗ topology on bounded subsets of M coincides with the Mackey topology (i.e., with the topology of uniform convergence on weakly compact subsets of M∗). We shall analyze the relationship in more detail in the following lemma. Lemma 7.10. Assume that M is a JBW∗-triple, e ∈ M is a tripotent, and ϕ ∈ M∗ satisfies kϕk = ϕ(e). Define the mapping Φ = Φe,ϕ : M → M∗ by (a) Φ is a conjugate linear mapping of M into M∗ which is moreover weak∗-to-weak Φ(a)(x) = ϕ({x, a, e}), x ∈ M, a ∈ M. continuous and kΦk ≤ kϕk; compact subset of M∗; (b) Set K = K(e, ϕ) = Φ(BM ) ⊆ kϕk BM∗ . Then K is an absolutely convex weakly (c) Let a ∈ M be arbitrary. Then sup{ψ(a) ; ψ ∈ K} = kΦ(a)k ; (d) For each x ∈ BM we have e,ϕ ≤ kΦ(x)k ≤pkϕk · kxke,ϕ . kxk2 In particular, the topologies induced by the seminorms k·ke,ϕ and kΦ(·)k coin- cide on BM . Proof. (a) The mapping (x, y) 7→ ϕ({x, y, e}), x, y ∈ M, MEASURES OF WEAK NON-COMPACTNESS 31 is a separately weak∗-continuous sesquilinear form on M (cf. Lemma 6.11). Indeed, the separate weak∗ continuity follows from the assumption ϕ ∈ M∗ together with the separate weak∗-to-weak∗ continuity of the Jordan product. It follows that for each a ∈ M its image Φ(a) is a weak∗ continuous linear functional on M , hence Φ(a) ∈ M∗. Further, Φ is clearly conjugate linear. The estimate of the norm is immediate from the inequality k{x, y, z}k ≤ kxk kykkzk (x, y, z ∈ M ) [30, Corollary 3], [16, Corollary 4.1.114]. Finally, Φ is weak∗-to-weak continuous because for each x ∈ (M∗)∗ = M the mapping is weak∗ continuous on M . a 7→ Φ(a)(x) = ϕ({x, a, e}) (b) This follows from (a) as BM is weak∗ compact and absolutely convex. (c) Let us compute: sup{ψ(a) ; ψ ∈ K} = sup{Φ(x)(a) ; x ∈ BM} = sup{ϕ({a, x, e}) ; x ∈ BM} = sup{(cid:12)(cid:12)(cid:12)ϕ({x, a, e})(cid:12)(cid:12)(cid:12) ; x ∈ BM} = sup{(cid:12)(cid:12)(cid:12)Φ(a)(x)(cid:12)(cid:12)(cid:12) ; x ∈ BM} = kΦ(a)k , where we used that the sesquilinear form from the proof of (a) is hermitian (because it is even positive semidefinite by Lemma 6.11(d)). (d) Fix any x ∈ BM . Then kxk2 e,ϕ = ϕ({x, x, e}) = Φ(x)(x) ≤ kΦ(x)k , which proves the first inequality. Further, for any y ∈ BM , by the Cauchy-Schwarz inequality, we have Φ(x)(y) = ϕ({y, x, e}) ≤ kxke,ϕ · kyke,ϕ ≤ kxke,ϕ ·pkΦ(y)k ≤pkϕk · kxke,ϕ . This proves the second inequality. The 'in particular part'then follows immediately. (cid:3) If ϕ ∈ M∗ \ {0}, we set Φϕ = Φs(ϕ),ϕ, K(ϕ) = K(s(ϕ), ϕ), k·kK ϕ = kΦϕ(·)k , where we use the notation from the previous lemma. We can next state a characterization of relatively weakly compact subsets in the predual of a JBW∗-triple. Proposition 7.11. Let M be a JBW∗-triple. Let A ⊆ M∗ \ {0} be such that the topology on M generated by the family {k·kϕ : ϕ ∈ A} coincides on bounded sets with the strong∗ topology. Then the following assertions are satisfied. (a) Let L ⊆ M∗ be a weakly compact subset and ε > 0. Then there are ϕ1, . . . , ϕk ∈ A and n ∈ N such that L ⊆ n · conv(K(ϕ1) ∪ ··· ∪ K(ϕk)) + εBM∗ ; (b) Assume moreover that the family {M2(s(ϕ)); ϕ ∈ A} is up-directed by inclu- sion. Then for any weakly compact set L ⊆ M∗ and any ε > 0 there are ϕ ∈ A and n ∈ N such that L ⊆ nK(ϕ) + εBM∗ . In particular, the assumption is satisfied if ∀ψ ∈ M∗ \ {0}, ∃ϕ ∈ A : s(ψ) ∈ M2(s(ϕ)). 32 J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER Proof. (a) We may assume that 0 < ε ≤ 1. We will use the following notation. For a bounded set D ⊆ M∗ denote by qD the seminorm on M defined by qD(x) = sup{ϕ(x) ; ϕ ∈ D}. Then qL is a Mackey continuous seminorm on M , so qLBM is strong∗ continuous by Proposition 7.9. The assumption together with Lemma 7.10(d) yields the existence of ϕ1, . . . , ϕk ∈ A and a natural number m such that for δ = 1 m > 0 we have {x ∈ BM ; kxkK ϕj ≤ δ for j = 1, . . . , k} ⊆ {x ∈ BM ; qL(x) ≤ ε}, hence Clearly {x ∈ BM ; kxkK ϕj ≤ δ for j = 1, . . . , k}◦ ⊃ {x ∈ BM ; qL(x) ≤ ε}◦. {x ∈ BM ; qL(x) ≤ ε}◦ ⊃ 1 ε L. Further, by Lemma 7.10(c) we have k·kK ϕ = qK(ϕ) for any ϕ ∈ A, hence {x ∈ BM ; kxkK ϕj ≤ δ for j = 1, . . . , k}◦ =BM ∩ \j≤k = BM∗ ∪ =BM ∩ \j≤k(cid:18) 1 K(ϕj)(cid:19)◦◦ {x ∈ M ; qK(ϕj)(x) ≤ δ}◦ ◦◦ K(ϕj ) δ [j≤k 1 δ 1 δ ⊆ It follows that conv(K(ϕ1) ∪ ··· ∪ K(ϕk)) + BM∗ . L ⊆ ε δ conv(K(ϕ1) ∪ ··· ∪ K(ϕk)) + εBM∗ ⊆ m · conv(K(ϕ1) ∪ ··· ∪ K(ϕk)) + 2εBM∗ , which completes the proof. (b) We proceed in the same way as in the proof of (a). We find ϕ1, . . . , ϕk and δ. The assumption then yields ϕ ∈ A such that M2(s(ϕ)) contains s(ϕ1), . . . , s(ϕk). By Proposition 7.5(i) and Lemma 7.10(d) we get some η > 0 such that {x ∈ BM ; kxkK ϕ ≤ η} ⊆ {x ∈ BM ; kxkK ϕj ≤ δ for j = 1, . . . , k} ⊆ {x ∈ BM ; qL(x) ≤ ε}. The arguments in the second part of the proof of (a) complete the proof here. The 'in particular' statement concerning the family {k·kϕ : ϕ ∈ A} follows from (cid:3) Proposition 7.5(i). 8. Proof of the main result In this section we provide a proof of Theorem 4.1. We begin with a technical lemma. MEASURES OF WEAK NON-COMPACTNESS 33 Lemma 8.1. Let M be a JBW∗-algebra and let (pn) be an increasing sequence of projections in M with supremum p. Then for any bounded sequence (ak) in M we have strong∗- lim k P2(p){ak, ak, p} = 0 ⇔ ∀n ∈ N, strong∗- lim k P2(pn){ak, ak, pn} = 0. Proof. (⇒) Fix n ∈ N. By Lemma 6.10(a) we have P2(pn){ak, ak, pn} = P2(pn){ak, ak, p} = P2(pn)P2(p){ak, ak, p}. Thus we can conclude by the strong∗-to-strong∗ continuity of P2(pn). (⇐) Arguing by contradiction, we assume that (ak) is a bounded sequence in M such that ∀n ∈ N : strong∗- lim strong∗ 6−→ 0. We may assume, without loss of generality that P2(pn){ak, ak, pn} = 0 k but P2(p){ak, ak, p} (ak) ⊆ BM . Since P2(p){ak, ak, p} = {P2(p)(ak), P2(p)(ak), p} + {P1(p)(ak), P1(p)(ak), p} is a positive element of M (by Lemma 6.10(a),(d)), there is, due to Lemma 7.4, a positive norm-one functional ϕ ∈ M∗ (i.e., a normal state on M ) such that ϕ(P2(p){ak, ak, p}) 6→ 0. Up to passing to a subsequence we may assume that there is some c > 0 such that ϕ(P2(p){ak, ak, p}) > c, for all k ∈ N. By [10, Lemma 3.2] there is some m ∈ N with kP2(p)∗ϕ − P2(pm)∗ϕk < c 2 . Then ϕ(P2(pm){ak, ak, pm}) = ϕ(P2(pm){ak, ak, p}) > c + ϕ(P2(pm){ak, ak, p}) − ϕ(P2(p){ak, ak, p}) = c + P2(pm)∗ϕ{ak, ak, p} − P2(p)∗ϕ{ak, ak, p} ≥ c − kP2(pm)∗ϕ − P2(p)∗ϕk > c 2 for all k ∈ N. Thus, Lemma 7.4 implies that (P2(pm){ak, ak, pm})k leading to a contradiction. strong∗ 6−→ 0, (cid:3) Lemma 8.2. Let M be a σ-finite JBW∗-algebra and let (ϕn) be a sequence of nonzero positive functionals in M∗ such that their support projections s(ϕn) form an increasing sequence with supremum 1. Then the strong∗ topology on bounded , n ∈ N. subsets of M coincides with the topology generated by the seminorms k·kϕn Proof. Since M is σ-finite, there exists a normal state ϕ ∈ M∗ with s(ϕ) = 1 and, moreover, the norm k·kϕ generates the strong∗ topology on bounded sets of M (cf. Lemma 7.3). Hence we can conclude using Lemmata 7.6 and 8.1. Lemma 8.3. Let M be a JBW∗-algebra and A ⊆ M∗ a bounded set. Then there is a countable set B ⊆ A such that ω(B′) = ω(A) for any B′ ⊆ B infinite. Proof. For any σ-finite projection p ∈ M , the Peirce-2 subspace M2(p) is a σ-finite JBW∗-algebra, hence by Lemma 7.3 we can fix a faithful normal state ωp on M2(p). Let us set ϕp = ωp ◦ P2(p). Then ϕp is a normal positive functional on M such that kϕpk = ϕp(p) = 1 and ϕpM2(p) is faithful. (cid:3) 34 J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER Let Φp = Φp,ϕp using the notation from Lemma 7.10. Let Kp = K(p, ϕp) = Φp(BM ). Then Kp is a weakly compact set in M∗ (by Lemma 7.10(b)). Let A ⊆ M∗ be a bounded set such that c = ω(A) > 0. Let us construct, by induction, two sequences (γn) ⊆ A and (pn) ⊆ M such that n+1 . (i) kγ1k > c − 1; (ii) pn is a σ-finite projection such that γn = γn ◦ P2(pn); (iii) pn ≥ pk for k < n; (iv) dist(γn+1, n conv(Kp1 ∪ ··· ∪ Kpn)) > c − 1 This construction can be done by just applying the definition of ω(A). Indeed, the existence of γ1 ∈ A satisfying (i) is obvious. Assume that n ∈ N and we have already constructed γj for 1 ≤ j ≤ n and pj for 1 ≤ j < n. By [10, Lemma 3.6] there is a σ-finite projection r ∈ M such that γn = γn◦P2(r). By [10, Lemma 3.5], there is a σ-finite projection pn ≥ r satisfying (iii). Clearly pn satisfies (ii) as well. Finally, find γn+1 ∈ A satisfying (iv) by the definition of ω(A). Set B = {γn; n ∈ N}. Then B is a countable subset of A and we claim that ω(B′) = c for each infinite subset B′ ⊆ B. Let p = supn pn. Then p is σ-finite (see, e.g., [24, Theorem 3.4] or [10, Lemma 3.5]) and B ⊆ P2(p)∗M∗. Since P2(p) is a norm-one projection, an application of Lemma 2.1 shows that ω(B) = ωM2(p)(B). We continue by working in the JBW∗- algebra M2(p). Lemma 8.2 implies that the strong∗ topology on BM2(p) is generated by the . Let L ⊆ (M2(p))∗ = P2(p)∗M∗ a weakly sequence of seminorms k·kϕpn M2(p) compact set and ε > 0. By Proposition 7.11(b) there are m, n ∈ N such that It follows that L ⊆ nΦpm (BM2(p)) + εBM2(p)∗ ⊆ n(Kpm ∩ P2(p)∗M∗) + εBM2(p)∗ . bd(B, L) ≥bd(B, nKpm ) − ε ≥bd(B, kKpm) − ε ≥ c − k + 1 − ε 1 for each k ≥ max{n, m}. Hence bd(B, L) ≥ c − ε. Since L is an arbitrary weakly compact set, we get ω(B) ≥ c − ε, and by the arbitrariness of ε > 0 we have ω(B) ≥ c. The same procedure applies to each infinite subset B′ ⊆ B, so the proof is completed. (cid:3) Proof of Theorem 4.1. If M is a JBW∗-algebra, the result follows from Lemma 8.3, [27, Corollary 4.3], and Proposition 5.1(b). The general case of a JBW∗-triple follows from the JBW∗-algebra case applying Lemma 2.1 and Proposition 3.2. (cid:3) 9. Preduals of JBW∗-triples which are strongly WCG Strongly WCG spaces (see the end of Section 5 for definitions) are a nice class of Banach spaces in which the computation of the De Blasi measure of weak non- compactness is easy. As explained in the end of Section 5 they include the spaces L1(µ) for a σ-finite measure µ or, more generally, preduals of σ-finite von Neu- mann algebras and preduals of σ-finite JBW∗-algebras. In the present section we characterize JBW∗-triples whose preduals are strongly WCG. Let us explain why it is not clear. By [73, Theorem 2.1] a Banach space X is strongly WCG if and only if the Mackey topology on X ∗ is metrizable on bounded MEASURES OF WEAK NON-COMPACTNESS 35 sets. Therefore, it follows from Proposition 7.9 that the predual M∗ of a JBW∗- triple M is strongly WCG if and only if the strong∗ topology on BM is metrizable. For σ-finite JBW∗-algebras this is the case by Lemma 7.3. However, for σ-finite JBW∗ triples it need not be the case as witnessed by the following example. Example 9.1. Let Γ be an uncountable set and C ⊆ Γ be an infinite countable set. Then M = B(ℓ2(Γ), ℓ2(C)) is a σ-finite JBW∗-triple whose strong∗ topology is not metrizable on the closed unit ball of M . Proof. A family of pairwise orthogonal partial isometries of M = B(ℓ2(Γ), ℓ2(C)) has pairwise orthogonal final projections and is therefore countable as ℓ2(C) is separable. Hence M is σ-finite. M is 1-complemented in B(ℓ2(Γ)) by a weak∗- to-weak∗ continuous projection, thus by [72, §1.15] and [14, COROLLARY], the strong∗ topology on bounded sets of M is given by the seminorms a 7→ ka(eγ)k + ka∗(eγ)k (a ∈ M ), where γ is a fixed element in Γ and {eγ : γ ∈ Γ} is the canonical orthonormal basis of ℓ2(Γ). If the strong∗ topology of BM were metrizable, it would be first countable, hence there would exist a countable set D ⊆ Γ such that the seminorms a 7→ ka(eγ)k + ka∗(eγ)k, γ ∈ D, generate this topology. We may assume without loss of generality that D ⊃ C. Let C = (cn)n∈N. Let {γn : n ∈ N} be a set of pairwise distinct elements in Γ \ D. Define the sequence of operators ak ∈ M by if γ = γn otherwise. Then 0 ak(eγ) =(ecn+k k(eγ) =(eγn−k a∗ if γ = cn and n > k otherwise. 0 In this case kak(eγn)k = 1, so ak do not converge strong∗ to zero. However a∗ k → 0 in SOT and ak(eγ) = 0 for γ ∈ D, so kak(eγ)k + ka∗ k(eγ)k → 0 for γ ∈ D, leading to a contradiction. (cid:3) On the other hand, σ-finiteness of a JBW∗-triple is a necessary condition for its predual to be strongly WCG. Indeed, any strongly WCG space is clearly WCG and the predual of a JBW∗-triple is WCG if and only if the triple is σ-finite by [11, Theorem 1.1]. In order to find a sufficient and necessary condition we get back to the structure results of JBW∗-triples due to G. Horn and E. Neher presented in (1) in page 7 (see [41, (1.7)], [42, (1.20)]). Every JBW∗-triple M decomposes (uniquely) as an (orthogonal) ℓ∞-sum of the form M = (cid:16)Lj∈J Aj⊗Cj(cid:17)ℓ∞ ⊕ℓ∞ H(W, α) ⊕ℓ∞ pV, where each Aj is a commutative von Neumann algebra, each Cj is a Cartan factor, W and V are continuous von Neumann algebras, p is a projection in V , α is a linear involution on W commuting with ∗, that is, a linear ∗-antiautomorphism of period 2 on W , and H(W, α) = {x ∈ W : α(x) = x}. Clearly, H(W, α) is a JBW∗-subalgebra of W when the latter is equipped whit its natural structure of JBW∗-algebra. A Cartan factor of type 1 is a JBW∗-triple C1 which coincides with the space B(H, K) of all bounded linear operators between two complex Hilbert spaces H and K. We can always assume that K is a closed subspace of H. Therefore, denoting 36 J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER by p the orthogonal projection of H onto K, we have C1 = B(H, K) = pB(H). Suppose A is a commutative von Neumann algebra. Now takingbp = 1⊗ p ∈ A⊗C1, we deduce that A⊗C1 =bp (A⊗B(H)) is a right ideal of the von Neumann algebra A⊗B(H). Cartan factors of types 2 and 3 are the subtriples of B(H) defined by C2 = {x ∈ B(H) : x = −jx∗j} and C3 = {x ∈ B(H) : x = jx∗j}, respectively, where j is a conjugation (i.e., a conjugate-linear isometry of period 2) on H. By a little abuse of notation, each x ∈ B(H) can be identified with a "matrix " (xγδ)γ,δ∈Γ. It is easy to check that the representing matrix of jx∗j is the transpose of the representing matrix of x. Hence, C2 consists of operators with antisymmetric representing matrix and C3 of operators with symmetric ones. The properties around Peirce decomposition show that, if a JBW∗-triple M admits a unitary element u, then M = M2(u) is a JBW∗-algebra with product ◦u and involution ∗u (cf. page 16). It is shown in the proof of [38, Proposition 2] that every Cartan factor of type 2 with dim(H) even, or infinite, and every Cartan factor of type 3 contains a unitary element. The same result actually proves that Cartan factors of type 2 with dim(H) even, or infinite, and all Cartan factors of type 3 are JBW∗-algebras. Consequently, if C is a Cartan factor of type 2 with dim(H) even, or infinite, or a Cartan factor of type 3, and A is a commutative von Neumann algebra, then A⊗C is a JBW∗-algebra. A Cartan factor of type 4 (also called a spin factor ) is a complex Hilbert space (with inner product h., .i) equipped with a conjugation x 7→ x, triple product {x, y, z} = hx, yiz + hz, yix − hx, ziy, and norm given by kxk2 = hx, xi +phx, xi2 − hx, xi2. Let u be an element in a spin factor C4 satisfying u = u and kuk = hu, ui = 1. It is not hard to check that {u, u, x} = hu, uix + hx, uiu − hu, xiu = hu, uix + hx, uiu − hx, uiu = x, for all x ∈ C4. This shows that u is a unitary in C4, and consequently A⊗C4 is a JBW∗-algebra whenever A is a commutative von Neumann algebra. mutative von Neumann algebra. Then A is isomorphic to Lℓ∞ Thus the JBW∗-triple A⊗C can be identified withLℓ∞ Finally, assume that C is a finite-dimensional Cartan factor and A is a com- j∈J L∞(µj), where (µj)j∈J is a family of finite (or, equivalently, probability) measures (cf. [72, §1.18]). j∈J L∞(µj , C), (cf. [39, 41]). Let us observe that the remaining Cartan factors, that is, the exceptional Cartan factors of types 5 and 6, are all finite-dimensional (they have dimensions 16 and 27, respectively). Combining the arguments in the previous paragraphs we get the following rep- resentation of JBW∗-triples. Proposition 9.2. Let M be any JBW∗-triple. Then M is (isometrically) JB∗triple isomorphic to a JBW∗-triple of the form (14) ℓ∞Mk∈Λ1 ℓ∞M N ℓ∞M pV, L∞(µk, Ck)! ℓ∞M ℓ∞Mj∈Λ2 L∞(µj , Dj) where • (µk)k∈Λ1 and (µj)j∈Λ2 are two (possibly empty) families of probability measures; MEASURES OF WEAK NON-COMPACTNESS 37 • Each Ck is a Cartan factor of type 5 or 6 for any k ∈ Λ1, and each Dj is a finite-dimensional Cartan factor of type 2 with dim(H) ∈ N odd for any j ∈ Λ2; • N is a JBW∗-algebra; • V is a von Neumann algebra and p ∈ V is a projection such that the triple pV has no nonzero direct summand triple-isomorphic to a JBW∗-algebra. Moreover, such a representation is unique, in the sense that if M admits two such representations, the respective four summands in one of them are triple-isomorphic to the respective four summands in the second one. Thanks to the structure result in the previous proposition, the promised charac- terization of JBW∗-triples is now stated in the following theorem. Theorem 9.3. Let M be a JBW∗-triple. Consider its representation provided by Proposition 9.2. (a) M is σ-finite if and only if the sets Λ1 and Λ2 are countable, the JBW∗-algebra N is σ-finite and the projection p is σ-finite. (b) M∗ is WCG if and only if M is σ-finite. (c) The following assertions are equivalent. (i) M∗ is strongly WCG. (ii) M is σ-finite and, moreover, the projection p is finite. (iii) There is ϕ ∈ M∗ \ {0} such that the strong∗ topology on BM is generated (iv) There is a σ-finite tripotent u ∈ M whose Peirce-2 subspace M2(u) is by k·kϕ. maximal with respect to inclusion. Assertions (a) and (b) follow from [24] and [11], respectively. More concretely, the 'only if part'in (a) is obvious; to see the 'if part' it is enough to use the known fact that exceptional Cartan factors are finite-dimensional and every Dj is finite- dimensional too, hence each of the summands L∞(µα, Cα) and L∞(µj, Dj) is σ- finite (cf. [24, Theorem 4.4]). Assertion (b) follows from [11, Theorem 1.1]. It remains to prove (c). In view of (b) we may restrict our attention to the σ-finite case. Let us observe that some implications in (c) are easy at this point. Indeed, (iii) implies that the strong∗ topology on BM is metrizable, hence we get (iii) ⇒ (i). Further, (iii) ⇒ (iv) follows from Proposition 7.5(ii). Recall that a projection p in a von Neumann algebra V is finite if there is no partial isometry in V with final projection p and initial projection stricly less than p. The argument will follow after considering the individual summands in the representation. However, we first give the following corollary on JBW∗-triples with separable predual. Note that while any separable Banach space is trivially WCG, c0 is an example of a separable space which is not strongly WCG by [73, Theorem 2.5]. A similar example cannot be a predual of a JBW∗-triple. Corollary 9.4. Let M be a JBW∗-triple with separable predual M∗. Then M∗ is strongly WCG. Proof. First observe that M is σ-finite. Indeed, being separable, M∗ is WCG, thus M is σ-finite by Theorem 9.3(b). (There is also an alternative way of proving this. Assume that M∗ is separable and fix e ∈ M a complete tripotent. Then M2(e)∗ is also separable. Since M2(e) is a JBW∗-algebra, we can choose a countable family of normal states {ϕn : n ∈ N} which is norm-dense in the set of normal states of 38 J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER M2(e). Then 1 2n ϕn is a faithful normal state of M2(e). Therefore, M2(e) is ∞Xn=1 σ-finite, so e is σ-finite and M is σ-finite as well.) So, assume M∗ is separable and fix a representation of M given by Proposi- tion 9.2. It follows that (pV )∗ is separable as well. Without loss of generality there is no nonzero central projection in V orthogonal to p (if z is such a projection, then pV = p(1 − z)V ). We claim that in this case necessarily V is σ-finite. As- sume it is not the case. Then there is an uncountable family of pairwise orthogonal nonzero projections (rγ )γ∈Γ in V . It follows from [77, Theorem V.1.8] that for each γ ∈ Γ there is a nonzero partial isometry uγ ∈ V such that its initial projection pi(uγ) ≤ rγ and its final projection pf (uγ) ≤ p. Then clearly uγ ∈ pV for γ ∈ Γ. Fix ϕγ ∈ (pV )∗ of norm one with uγ = s(ϕγ ). Since ϕγ(uγ) = 1 and for δ 6= γ ϕγ(uδ) = ϕγP2(uγ)(uδ) = 0, we see that (ϕγ)γ∈Γ is a 1-discrete set, contradicting the separability of (pV )∗. Hence the strong∗ topology on BV is metrizable, so by Lemma 7.1(b) the same holds for BpV , thus (pV )∗ is strongly WCG. Using Theorem 9.3 we now see that p is finite and hence M∗ is strongly WCG as well. (cid:3) To prove assertion (c) in Theorem 9.3 we will describe the structure of all pre- Hilbertian seminorms generating the strong∗ topology using Proposition 7.5 and some complements to that. We will do it first for σ-finite triples and then (in the next section) we shall discuss the general case. We start by analyzing the individual summands appearing in Proposition 9.2. 9.1. JBW∗-algebras. In the case of JBW∗-algebras we can conclude by applying the existing literature. The desired conclusion is covered by the following proposi- tion. Proposition 9.5. Let M be a JBW∗-algebra. (a) Let (en) be a sequence of σ-finite tripotents in M . Then there is a σ-finite (b) Assume M is not σ-finite. Then for each σ-finite projection p ∈ M there is a (c) The strong∗ topology on BM is metrizable if and only if M is σ-finite. In this projection p ∈ M such that M2(p) contains en for each n ∈ N. σ-finite projection q ∈ M such that q > p (and hence M2(p) $ M2(q)). case it is metrizable by k·kω, where ω is any faithful normal state. Proof. (a) This follows from Lemma 7.2(d) and [10, Lemma 3.5]. (b) This follows easily from the definitions. If M is not σ-finite and p is σ-finite, then 1 − p 6= 0, hence there is a σ-finite projection r ∈ M2(1 − p). It is enough to take q = p + r. (c) The 'if part' follows from Lemma 7.3. Conversely, assume that BM is metriz- able in the strong∗ topology. Then there is a countable base of strong∗ neighbor- hoods of zero in BM . It follows that the strong∗ topology on BM is generated by countably many seminorms. By (a) and Proposition 7.5(i) it is generated by one seminorm. By (b) and Proposition 7.5(ii) we deduce that M is σ-finite. (cid:3) 9.2. Finite dimensional Cartan factors. In this subsection we shall deal with summands of the form L∞(µ, C) where C is an exceptional Cartan factor (i.e., the Cartan factor of type 5 or 6) or a finite-dimensional Cartan factor of type 2 with dim(H) ∈ N odd. We start with properties of a finite-dimensional JB∗-triple. MEASURES OF WEAK NON-COMPACTNESS 39 Let E be a JB∗-triple. Following the most employed notation, the symbol U(E) will stand for the set of all tripotents in E. We shall write U(E)∗ for the set of all nonzero tripotents in E, and we shall employ the symbol Umax(E) to denote the set of all complete tripotents in E. By Kaup's Riemann mapping theorem [52, Proposition 5.5], a linear bijection between JB∗-triples E and F is a triple isomorphism if and only if it is an isometry. Henceforth, we denote by Iso(E, F ) the set of all surjective isometries (equivalently, triple isomorphisms) from E to F . We write Iso(E) =Iso(E, E) for the set of all triple automorphisms of E. Fix Φ ∈ Iso(E). Then Φ, being a JB∗-triple automorphism, preserves all the triple structure. In particular, it maps tripotents to tripotents and complete tripo- tents to complete tripotents, that is, (15) Φ(E)(U(E)∗) = U(E)∗, and Φ(E)(Umax(E)) = Umax(E). Moreover, the equality Φ(Pj (e)(x)) = Pj(Φ(e))(Φ(x)) holds for every e ∈ U(E), j ∈ {0, 1, 2} and x ∈ E. In particular, Φ(E2(e)) = E2(Φ(e)) and Φ is a (unital) JB∗-algebra isomorphism of E2(e) onto E2(Φ(e)). Let us fix e ∈ U(E) and ϕ ∈ E∗ a functional satisfying ϕ = ϕP2(e). Then ϕ◦Φ−1 = ϕ◦P2(e)◦Φ−1 = ϕ◦Φ−1◦P2(Φ(e)) and (ϕ ◦ Φ−1)E2(Φ(e)) = ϕE2(e) ◦ Φ−1. It is natural to ask about the orbit of a fixed e ∈ Umax(E) under the group Iso(E). In general, Iso(E)(e) is not easy to be determined (cf. [13] and [54]). If E is a finite-dimensional JB∗-triple, then any two complete (maximal) tripotents in E are interchanged by an element in Iso(E) (see [57, Theorem 5.3(b)]). This can be also seen by applying that E being finite-dimensional implies that E coincides with a finite ℓ∞-sum of finite-dimensional Cartan factors, and it is known that on a finite-dimensional Cartan factor C the group Iso(C) acts transitively on Umax(C). Therefore, for e ∈ Umax(E) and dim(E) < ∞ we have (16) Iso(E)(e) = Umax(E). Lemma 9.6. Let E be a finite-dimensional JB∗-triple, let e ∈ Umax(E), and let ϕ ∈ E∗ be a norm-one functional such that e = s(ϕ). Then the following assertions hold: (a) For each Φ ∈ Iso(E) we have Φ(e) ∈ Umax(E) and Φ(e) = s(ϕ ◦ Φ−1); (b) U(E), U(E)∗, and Umax(E) are compact subsets of E and Iso(E) is a compact (c) There is a constant α > 0 such that for each Φ ∈ Iso(E) we have subset of B(E); αkxk ≤ kxkϕ◦Φ−1 ≤ kxk , x ∈ E. (d) There is a Borel measurable mapping θ : Umax(E) → Iso(E) such that u = θ(u)(e) for each u ∈ Umax(E). Proof. Since E is finite-dimensional, it is a σ-finite JBW∗-triple, so ϕ can be found. (a) This was justified in (15). (b) Since the triple product is jointly norm continuous, U(E) and U(E)∗ = U(E)\{0} are closed subsets of the closed unit ball and the unit sphere of E, respectively, so they are compact. Elements of Iso(E) are precisely (surjective) isometries, so Iso(E) is a closed subset of the unit sphere of B(E), hence it is compact. 40 J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER We next consider the mapping Ψ : Iso(E) → E defined by Ψ(Φ) = Φ(e), Φ ∈ Iso(E). It is clearly a continuous mapping and by (a) it maps Iso(E) into Umax(E). We deduce from (16) that Ψ is onto, so Umax(E) is compact. (c) For any Φ ∈ Iso(E) and x ∈ E we have (due to (a)) kxk2 ϕ◦Φ−1 = (ϕ ◦ Φ−1){x, x, Φ(e)} = ϕ(cid:8)Φ−1(x), Φ−1(x), e(cid:9) =(cid:13)(cid:13)Φ−1(x)(cid:13)(cid:13)2 Since ϕ . (x, Φ) 7→(cid:13)(cid:13)Φ−1(x)(cid:13)(cid:13)ϕ , x ∈ SE, Φ ∈ Iso(E) is a strictly positive continuous mapping on the compact space SE × Iso(E), it has some strictly positive minimum and maximum. Thus, the existence of the constant α easily follows. Clearly, kxkφ ≤ kxk for all x ∈ M and every norm-one functional φ in M∗. (d) The mapping Ψ from the proof of (b) is a continuous mapping of a compact metric space Iso(E) onto a compact metric space Umax(E), hence the inverse set- valued map u 7→ Ψ−1(u) admits a Borel-measurable selection by the Kuratowski -- Ryll-Nardzewski theorem (see [2, Theorem 18.13]). (cid:3) The reader may already guess at this stage that the constant α > 0 given by Lemma 9.6(c) is directly linked to the dimension of the JB∗-triple E. If we have a family {Ck : k ∈ Λ} of finite-dimensional Cartan factors for which the dim(Ck) is uniformly bounded for all k ∈ Λ (for example, a family of exceptional Cartan factors of types 5 and 6), then the constant α can be chosen to be valid for all k ∈ Λ. Proposition 9.7. Let E be a finite-dimensional JB∗-triple, and let (Ω, Σ, µ) be a probability space. Consider the JBW∗-triple M = L∞(µ, E) (equipped with the pointwise triple product). Let e, ϕ, θ, α be as in Lemma 9.6. Then the following assertions hold: (a) An element f ∈ M is a tripotent if and only if f (ω) ∈ U(E) µ-almost every- (b) An element f ∈ M is a complete tripotent if and only if f (ω) ∈ Umax(E) (c) Assume that f ∈ M is a complete tripotent. Let µ-almost everywhere; where; v(ω) = ϕ ◦ θ(f (ω))−1, ω ∈ Ω. Then v ∈ L1(µ, E∗) = L∞(µ, E)∗ and s(v) = f ; (d) Let f and v be as in (c). Then 2 (e) The strong∗ topology on BM coincides with the topology generated by the norm ≤ kgkv ≤(cid:18)Z kg(ω)k2 dµ(ω)(cid:19) 1 α(cid:18)Z kg(ω)k2 dµ(ω)(cid:19) 1 k·kv and also with the topology generated by the norm g 7→(cid:16)R kg(ω)k2 dµ(ω)(cid:17) 1 Proof. Assertion (a) follows immediately from the fact that the triple product is defined pointwise. g ∈ M ; . 2 2 , MEASURES OF WEAK NON-COMPACTNESS 41 (b) Since the triple product is defined pointwise, we have, for a given tripotent f ∈ M , P0(f )(g)(ω) = P0(f (ω))(g(ω)) µ-a.e. Hence, if f (ω) ∈ Umax(E) µ-almost everywhere, then clearly P0(f ) = 0. Conversely, assume that it is not true that f (ω) ∈ Umax(E) µ-almost everywhere. Since Umax(E) is a closed set, there is a measurable set A ⊆ Ω of positive measure such f (ω) /∈ Umax(E), for all ω ∈ A. For any u ∈ U(E) there is u′ ∈ Umax(E) with u ≤ u′ (cf. [40, Lemma 3.12]). Moreover, the set {(u, u′) ∈ U(E) × Umax(E); u ≤ u′} is closed, hence compact, and thus the set-valued mapping U(E) ∋ u 7→ {u′ ∈ Umax(E); u ≤ u′} is upper-semicontinuous and compact-valued. By the Kuratowski -- Ryll-Nardzewski theorem we find a Borel-measurable mapping ζ : U(E) → Umax(E) such that u ≤ ζ(u) for u ∈ U(E). Then the mapping g = ζ ◦ f belongs to M and P0(f )(g)(ω) = P0(f (ω))(g(ω)) = P0(f (ω))(ζ(f (ω)) = ζ(f (ω)) − f (ω) which is nonzero on A. Thus f is not complete. (c) By (b) we know that f (ω) ∈ Umax(E) µ-almost everywhere, so the mapping ω 7→ θ(f (ω)) is a µ-almost everywhere defined measurable mapping from Ω into Iso(E). Since taking an inverse is a continuous transformation, we see that v is a µ-almost everywhere defined measurable mapping from Ω into E∗. Moreover, since kϕk = 1 and elements of Iso(E) are isometries, kv(ω)k = 1 µ-almost everywhere. Thus v ∈ L1(µ, E∗) and kvk = 1 (as µ is a probability measure). Moreover, hv, fi =Z hv(ω), f (ω)i dµ =Z ϕ ◦ θ(f (ω))−1(f (ω)) dµ =Z ϕ(e) dµ = 1. Furthermore, assume that h ∈ M2(f ) is positive with hv, hi = 0. Then h(ω) is a positive element of E2(f (ω)) for µ-almost all ω ∈ Ω, hence θ(f (ω))−1(h(ω)) is a positive element of E2(e) for µ-almost all ω. Hence 0 = hv, hi =Z hv(ω), h(ω)i dµ =Z ϕ(θ(f (ω))−1(h(ω))) dµ, so ϕ(θ(f (ω))−1(h(ω))) = 0 µ-a.e. Since ϕ is faithful on E2(e), we deduce that θ(f (ω))−1(h(ω)) = 0 µ-a.e., so h(ω) = 0 µ-a.e. (d) For any g ∈ M we have v = hv,{g, g, f}i =Z hv(ω),{g(ω), g(ω), f (ω)}i dµ kgk2 =Z (cid:10)ϕ ◦ θ(f (ω))−1,{g(ω), g(ω), f (ω)}(cid:11) dµ =Z kg(ω)k2 so we can conclude by the choice of α. ϕ◦θ(f (ω))−1 dµ, (e) For any tripotent h ∈ M there is a complete tripotent f ∈ M with f ≥ h. For any complete tripotent f let v(f ) ∈ M∗ be as in (c). By Proposition 7.5(i) the strong∗ topology on BM coincides with the topology generated by the seminorms 42 J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER k·kv(f ), f ∈ M a complete tripotent. We deduce from (d) that all these norms are equivalent to the norm g 7→(cid:16)R kg(ω)k2 dµ(ω)(cid:17) 1 2 . (cid:3) 9.3. Triples of the form pV . It turns out that the analysis of this case is more complicated than the previous two cases. We shall employ an argument which is closely related to the notion of equivalence of projections and to the theory of types of von Neumann algebras (see, for example, [50]). Given a von Neumann algebra V , two projections p, q ∈ V are said to be equiv- alent (we write p ∼ q) if there is a partial isometry in V with initital projection p and final projection q. Further, a projection p is called finite if the only projection q satisfying q ≤ p and q ∼ p is the projection p itself. A projection which is not finite is called infinite. Finally, a projection p is properly infinite if zp is infinite for any central projection z such that zp 6= 0. For any projection p ∈ V its central carrier is the smallest central projection Cp satisfying Cpp = p. It is further known that there is a unique central projection z ≤ Cp such that zp is properly infinite or zero and (1 − z)p = (Cp − z)p is finite. Indeed, if p is finite, we take z = 0, and if p is infinite we may use [50, Proposition 6.3.7]. Henceforth, assume that we have a JBW∗-triple of the form pV , where V is a von Neumann algebra, and p ∈ V is a projection. We may assume, without loss of generality, that Cp = 1 (otherwise we may replace V by CpV ). By the previous paragraph there is a central projection z ∈ V such that zp is properly infinite and (1− z)p is finite. Then pV = zpV ⊕ (1− z)pV , thus we discuss separately the cases in which p is finite or properly infinite. We begin with the following lemma on equivalence of projections. Lemma 9.8. Let V be a von Neumann algebra. Then the following assertions are true. (a) Let (pn) be a sequence of properly infinite projections in V which are all equiv- alent to one projection q ∈ V . Then the supremum of the sequence (pn) is also equivalent to q. (b) Let (pn) be an increasing sequence of projections in V with supremum p. If all (c) Assume that p1, p2 are two equivalent projections in V . Then for any projection the projections pn are equivalent to one projection q, then p ∼ q as well. q1 ≥ p1 there is a projection q2 ≥ p2 such that q1 ∼ q2. Proof. Assertion (a) is proved in [74, Lemma 3.2(1)]. (b) By [50, Proposition 6.2.8] we have Cpn = Cq for each n ∈ N, hence Cp = Cq by [49, Proposition 5.5.3]. So, denote by c the common central carrier of all the projections in question. Let z ≤ c be the central projection such that zq is finite and (c − z)q is properly infinite. Then zpn ∼ zq for each n ∈ N, so zpn ∼ zpm for m, n ∈ N. Since zpn is finite for each n, we deduce that zpn = zpm for each m, n ∈ N, thus zp = zpn for n ∈ N, hence zp ∼ zq. Further, (c − z)pn ∼ (c − z)q for n ∈ N. Since (c − z)q is properly infinite, the projections (c − z)pn are properly infinite as well. Thus by (a) we deduce that (c − z)p ∼ (c − z)q, hence by [50, Proposition 6.2.2] p ∼ q. MEASURES OF WEAK NON-COMPACTNESS 43 (c) By the comparability theorem [77, Theorem V.1.8] for the pair of projections q1 − p1 and 1 − p2, there is a central projection z such that • z(q1 − p1) is equivalent to some projection r ≤ z(1 − p2), and • (1 − z)(1 − p2) is equivalent to some projection s ≤ (1 − z)(q1 − p1). By [50, Proposition 6.2.2] we get that zq1 = zp1 + z(q1 − p1) is equivalent to r + zp2 and, moreover, 1 − z = (1 − z)p2 + (1 − z)(1 − p2) is equivalent to (1 − z)p1 + s ≤ (1 − z)q1. But this means that (1 − z)q1 is equivalent to 1 − z (by [50, Proposition 6.2.4]). Finally, one can take q2 = r + zp2 + 1 − z. (cid:3) We consider first the case in which p is finite. Lemma 9.9. Let V be a von Neumann algebra and p ∈ V be a finite and σ-finite projection such that p 6= 1. Consider the JBW∗-triple M = pV . (a) There is τ ∈ M∗ such that s(τ ) = p, τ (p) = 1 and τpV p is a trace. (b) Let u ∈ M be a complete tripotent. Then u can be extended to a unitary operator u ∈ V . Moreover, the functional τu(x) = τ (xu∗), x ∈ M, belongs to M∗, s(τu) = u and 1 √2pτ (pxx∗p) ≤ kxkτu ≤pτ (pxx∗p), x ∈ M. (c) The strong∗ topology on BM is generated by the norm k·kτ and also by the norm x 7→pτ (pxx∗p). Proof. (a) Since pV p is a finite and σ-finite von Neumann algebra with unit p, it admits a normal finite faithful trace τ with τ (p) = 1. Indeed, such a trace can be obtained by composition of the standard canonical center valued trace on pV p (see [77, Theorem V.2.6]) with any norm-one faithful positive normal functional on the center of pV p. Then τ ◦ P2(p) (i.e., the mapping x 7→ τ (xp)) is an extension of τ to pV . Clearly p = s(τ ◦ P2(p)), hence it is enough to denote the composition again by τ . (b) Let u ∈ M be a complete tripotent. Then it is a partial isometry in V with final projection pf (u) ≤ p. Since pf (u) is finite u can be extended to a unitary operator u ∈ V (by [77, Proposition V.1.38]). Moreover, observe that the final projection pf (u) must coincide with p. Indeed, pu ∈ M and, since u is complete, 0 = P0(u)(pu) = (p − pf (u))pu(1 − pi(u)) = (p − pf (u))u(1 − pi(u)). Since p is finite, we get pi(u) 6= 1. Moreover, u maps the range of 1− pi(u) onto the range of 1− pf (u), which contains the range of p− pf (u). It follows that pf (u) = p. Set q = pi(u) = u∗u and consider the operator υ : M → M defined by υ(x) = xu∗, x ∈ M. Then υ is a surjective isometry. Hence it is a triple isomorphism (this can be also easily checked directly), in particular, it is a weak∗-to-weak∗ homeomorphism. 44 J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER Since υ(u) = p, we deduce that s(τ ◦ υ) = u. Thus, kxk2 τ ◦υ = (τ ◦ υ)({x, x, u}) = τ ({υ(x), υ(x), υ(u)}) = τ ({xu∗, xu∗, p}) = = = 1 2 1 2 1 2 τ (pxu∗ ux∗p + pux∗xu∗p) = τ (pxx∗p + ux∗xu∗) 1 2 (τ (pxx∗p) + τ ((ux∗p)(pxu∗))) = 1 2 (τ (pxx∗p) + τ ((pxu∗)(ux∗p)) (τ (pxx∗p) + τ (pxqx∗p)). Since q ≤ 1, we deduce pxqx∗p ≤ pxx∗p, and thus 1 2 τ (pxx∗p) ≤ kxk2 τ ◦υ ≤ τ (pxx∗p). Since τu = τ ◦ υ, the proof is completed. (c) It follows from (b) combined with Proposition 7.5(i) that the strong∗ topology , where u ∈ M is on BM coincides with the topology generated by the norms k·kτu a complete tripotent. By a further application of (b) we see that all these norms are equivalent to the one given in (c). (cid:3) We finally consider the case in which p is properly infinite. Proposition 9.10. Let V be a von Neumann algebra and let p ∈ V be a σ-finite properly infinite projection. Consider the JBW∗-triple M = pV . Assume that M contains no nonzero direct summand triple-isomorphic to a JBW∗-algebra. Then the following assertions hold: (a) V is not σ-finite; (b) A tripotent u ∈ M is complete if and only if its final projection equals p; (c) Let (un) be a sequence of complete tripotents in M . Then there is a complete tripotent u ∈ M such that M2(un) ⊆ M2(u) for each n ∈ N; that M2(u) $ M2(v); (d) If u ∈ M is a complete tripotent, then there is a complete tripotent v ∈ M such (e) The strong∗ topology on BM is not metrizable. Proof. (a) If V is σ-finite, then p ∼ 1 (as Cp = 1 and p is purely infinite), thus M = pV would be triple-isomorphic to a JBW∗-algebra (given a partial isometry u with uu∗ = p and u∗u = 1, the mapping x 7→ xu∗ is a surjective isometry from M onto pV p). (b) The 'if part' is clear. Let us prove the 'only if part'. Assume pf (u) < p. By (a) we get pi(u) < 1. Thus p − pf (u) and 1 − pi(u) are two nonzero projections in V , thus it follows easily from the comparability theorem [77, Theorem V.1.8] that there are two nonzero projections q1 ≤ 1 − pi(u) and q2 ≤ p − pf (u) which are equivalent. Fix a partial isometry v ∈ V with initial projection q1 and final projection q2. Then v ∈ M and P0(u)(v) = (p − pf (u))v(1 − pi(u)) 6= 0 as the range of 1 − pi(u) contains the range of q1, v maps it isometrically to the range of q2 which is contained in the range of p − pf (u). (c) By (b) we know that pf (un) = p for each n. So, p ∼ pi(un) for each n ∈ N. If we set q = supn pi(un), Lemma 9.8(a) yields p ∼ q. Then u can be any partial isometry with initial projection q and final projection p. MEASURES OF WEAK NON-COMPACTNESS 45 (d) By (b) we know that pf (u) = p. Since pi(u) < 1 (by (a)), we can find a σ-finite projection q > pi(u). Then q is properly infinite, and hence p ∼ q. Then v can be any partial isometry with initial projection q and final projection p. (e) Assume that the restriction of the strong∗ topology to BM is metrizable. Then it is first countable, hence generated by countably many of the defining semi- norms. Then (c) and (d) together with Proposition 7.5(i) yield a contradiction. (cid:3) 9.4. The case of a general σ-finite JBW∗-triple. We are now ready to prove assertion (c) of Theorem 9.3. We will do it by proving the following two propositions (the final proof follows them). Proposition 9.11. Assume that M is a nontrivial JBW∗-triple of the form ℓ∞Mk∈Λ L∞(µk, Ck)! ℓ∞M N ℓ∞M pV, of type 2 in B(Hk) with dim(Hk) odd; where • Λ is a (possibly empty) countable set; • (µk)k∈Λ is a (possibly empty) family of probability measures; • Each Ck is a Cartan factor of type 5 or 6 or a finite-dimensional Cartan factor • N is a (possibly trivial) σ-finite JBW∗-algebra; • V is a (possibly trivial) von Neumann algebra and p ∈ V is a finite σ-finite pro- jection such that the triple pV has no nonzero direct summand triple-isomorphic to a JBW∗-algebra. Fix a faithful normal state φ3 ∈ N∗. Let τ ∈ (pV )∗ be as in Lemma 9.9(a). Then the following statements hold: (a) We can regard φ3 as an element in M∗ satisfying that the strong∗ topology on (b) We can regard τ as an element in M∗ satisfying that the strong∗ topology on BN is metrizable by the norm k · kφ3N ; B(pV ) is metrizable by the norm k · kτpV ; (c) Let C = L∞(µk, Ck). Fix any ϕ ∈ C∗ \ {0} such that s(ϕ) ∈ Umax(C). ℓ∞Mk∈Λ Then the norm k·kϕ is equivalent to the norm (ak)k∈Λ 7→ ∞Xn=1 4−n Z kaknk2 dµkn! 1 2 on bounded sets of C (where (kn) is an enumeration of Λ). The strong∗ topology on BC is metrized by the norm displayed above; (d) The strong∗ topology on BM is metrized by the norm k·kτ +φ3+ϕ (where the functional ϕ from (c) is considered as an element of M∗) which is equivalent to the norm k((ak)k∈Λ, x, y)k2 = ∞Xn=1 1 4nZ kaknk2 dµ! + kxk2 φ3 + τ (py∗yp). Proof. (a) and (b) are proved in Lemmata 7.3(b) and 9.9, respectively. (c) Fix any ϕ ∈ C∗ \ {0} such that s(ϕ) = (fk)k∈Λ ∈ Umax(C). 46 J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER Fix k ∈ Λ. Then fk is a maximal tripotent in L∞(µk, Ck), hence we can fix hk ∈ L1(µk, (Ck)∗) provided by Proposition 9.7(c). Let (kn) be an enumeration of Λ and set φ1((ak)k∈Λ) = 2−n hhkn , akni , ((ak)k∈Λ ∈ C). ∞Xn=1 Clearly, s(φ1) = (fk)k∈Λ = s(ϕ), so k·kφ1 sition 7.5. By Proposition 9.7 the norm k·khk and k·kϕ are equivalent on BC by Propo- is equivalent to the norm on the unit ball of L∞(µk, Ck) for each k ∈ Λ. Hence, the norm is equivalent on BC to the norm k(ak)k∈Λkφ1 4−n kaknk2 hkn!1/2 4−nZ kaknk2!1/2 . f 7→(cid:18)Z kfk2 dµk(cid:19)1/2 = ∞Xn=1 (ak)k∈Λ 7→ ∞Xn=1 (cid:13)(cid:13)(cid:13)aj k(cid:13)(cid:13)(cid:13)hk Z (cid:13)(cid:13)(cid:13)aj k(cid:13)(cid:13)(cid:13) dµk 2 j j → 0 for each k ∈ N, → 0 for each k ∈ Λ, Indeed, both norms are well defined. Moreover, a bounded sequence ((aj converges to zero in the first norm if and only if k)k∈Λ)∞ j=1 which takes place if and only if which is in turn equivalent to the convergence to zero in the second norm. Finally, it follows from Proposition 7.5 that the strong∗ topology on BC is gen- erated by the mentioned norm. Lastly, statement (d) follows from the previous statements. (cid:3) The remaining case is treated in our next result. Proposition 9.12. Assume that M is a JBW∗-triple of the form ℓ∞Mk∈Λ L∞(µk, Ck)! ℓ∞M N ℓ∞M pV ℓ∞M qW, where • Λ is a (possibly empty) countable set; • (µk)k∈Λ is a (possibly empty) family of probability measures; • Each Ck is a Cartan factor of type 5 or 6 or a finite-dimensional Cartan factor • N is a (possibly trivial) σ-finite JBW∗-algebra; • V is a (possibly trivial) von Neumann algebra and p ∈ V is a finite σ-finite pro- jection such that the triple pV has no nonzero direct summand triple-isomorphic to a JBW∗-algebra; of type 2 in B(Hj ) with dim(Hj ) odd; MEASURES OF WEAK NON-COMPACTNESS 47 • W is a nontrivial von Neumann algebra and q ∈ W is a properly infinite σ- finite projection such that the triple qW has no nonzero direct summand triple- isomorphic to a JBW∗-algebra. Then the strong∗ topology on BM is not metrizable. Proof. Proposition 9.10(e) assures that the strong∗ topology on BqW is not metriz- able, and the desired conclusion follows from Lemma 7.1(b). (cid:3) Proof of Theorem 9.3(c). (ii) ⇒ (iii) This follows from Proposition 9.11. (iii) ⇒ (i) This follows from Proposition 7.9 and [73, Theorem 2.1] as explained (i) ⇒ (ii) Assume p is not finite. By Proposition 9.12 the strong∗ topology on BM is not metrizable. Hence M∗ is not strongly WCG by Proposition 7.9 and [73, Theorem 2.1]. above. (iii) ⇒ (iv) This follows from Proposition 7.5(ii). (iv) ⇒ (ii) Assume (iv) holds but p is not finite. It follows from (iv) that M is σ-finite, hence M has the form from Proposition 9.12. Let u = ((ak), (bj), x, v, w) be any tripotent in M . Then w is a tripotent in qW . It follows from Propo- sition 9.10 that there is a tripotent w ∈ qW with (qW )2(w) $ (qW )2( w). Set u = ((ak), (bj), x, v, w). Then u is a tripotent in M and M2(u) $ M2(u). (cid:3) Remark 9.13. It follows from the analysis of the individual cases in this section that any σ-finite JBW∗-triple can be expressed as a direct sum of (countably many) summands of three different types. Type 1 -- JBW∗-algebra: If N is a σ-finite JBW∗-algebra, it admits a unit, i.e., a (σ-finite) unitary element. Then N2(1N ) = N , hence the Peirce-2 subspace is the largest possible. Type 2 -- L∞(µ, C) or pV with p finite: Assume M = L∞(µ, C), where µ is a probability measure and C is a finite-dimensional Cartan factor without a unitary element, or M = pV , where V is a von Neumann algebra and p ∈ V is a finite σ-finite projection such that M has no direct summand isomorphic to a JBW∗-algebra. Then there are tripotents whose Peirce-2 subspaces are maximal with respect to inclusion, but mutually different. But all the norms k·kϕ, where s(ϕ) is such a tripotent, are equivalent on bounded sets. Type 3 -- pV for p properly infinite: Assume that M = pV , where V is a von Neumann algebra and p ∈ V is a properly infinite σ-finite projection such that M has no direct summand isomorphic to a JBW∗-algebra. Then the family of Peirce-2 subspaces M2(u), u ∈ U(M ), is upwards σ-directed by inclusion and has no maximal element. In the next section we give some consequences of this trichotomy to the structure of general (not necessarily σ-finite) JBW∗-triples. 10. On seminorms generating the strong∗ topology In this section we provide a characterization of the natural ordering of the semi- norms generating the strong∗ topology for a JBW∗-triple, which is defined by in- clusion of the respective topologies on the unit ball. The case of σ-finite triples is covered by Propositions 9.11 and 9.12, here we deal with general triples. The promised result is contained in the following theorem. 48 J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER Theorem 10.1. Let M be a JBW∗-triple. Then it can be represented in the form M =Mα∈Λ L∞(µα, Cα) ⊕ N ⊕ pV ⊕ qW, where • Λ is an arbitrary (possibly empty) set; • µα is a probability measure and Cα is a finite-dimensional Cartan factor not • N is a (possibly trivial) JBW∗-algebra; • V is a (possibly trivial) von Neumann algebra and p ∈ V is a finite projection such that the triple pV has no nonzero direct summand triple-isomorphic to a JBW∗-algebra; containing a unitary element for any α ∈ Λ; • p =Pj∈J pj, where (pj)j∈J is an orthogonal family of (finite) σ-finite projections • W is a (possibly trivial) von Neumann algebra and q ∈ W is a properly infi- nite projection such that the triple qW has no nonzero direct summand triple- isomorphic to a JBW∗-algebra. in the center of pV p; For an element h = ((fα(h))α∈Λ, x(h), v(h), w(h)) = ((fα)α∈Λ, x, v, w) ∈ M we denote Further, set spt1 h = {α ∈ Λ; fα 6= 0}, spt2 h = {j ∈ J; pjv 6= 0}. T (M ) = {h = ((fα)α∈Λ, x, v, w) ∈ M ; spt1 h, spt2 h are countable, ∀α ∈ spt1 h : fα is a complete tripotent in L∞(µα, Cα), x is a σ-finite projection in N, vv∗ = Xj∈spt2 h ww∗ is a properly infinite σ-finite projection below q}. pj, Then the following assertions hold: (a) The elements of T (M ) are σ-finite tripotents in M . Moreover, for any σ-finite tripotent g ∈ M there is h ∈ T (M ) with M2(g) ⊆ M2(h). (b) Let ϕ, ψ ∈ M∗ \ {0} such that the support tripotents of these functionals belong to T (M ). Set h = s(ϕ) and g = s(ψ). Then k·kϕ is weaker than k·kψ on BM if and only if the following assertions hold: ◦ spt1 h ⊆ spt1 g and spt2 h ⊆ spt2 g; ◦ x(h) ≤ x(g) as projections in N ; ◦ k·kϕ is weaker than k·kψ on BqW . Before proving this theorem let us formulate some consequences. Corollary 10.2. Let M be a JBW∗-triple. Given a sequence (ϕn) in M∗ \ {0}, is weaker than k·kψ on BM for each n ∈ N, there is ψ ∈ M∗ \ {0}, such that k·kϕn i.e., the family of topologies on BM generated by the seminorms k·kϕ, ϕ ∈ M∗ \{0} is upwards σ-directed by inclusion. The proof of this corollary will use one of the lemmata below, so we postpone the the end of the section. MEASURES OF WEAK NON-COMPACTNESS 49 Corollary 10.3. Let M be a JBW∗-triple and L ⊆ M∗ any weakly compact set. Then there is ϕ ∈ M∗ such that for any ε > 0 there is n ∈ N satisfying L ⊆ nK(ϕ) + εBM∗ . Proof. We will imitate the proof of Proposition 7.11 using the same notation. The seminorm qL is Mackey continuous, so qLBM is strong∗-continuous. Hence, given m ∈ N, there are ϕm km ∈ M∗ \ {0} and δm > 0 such that 1 , . . . , ϕm {x ∈ BM ; kxkϕm j ≤ δm for j = 1, . . . , km} ⊆ {x ∈ BM ; qL(x) ≤ 1 m}. By Corollary 10.2 there is ϕ ∈ M∗\{0} such that k·kϕm for each m ∈ N and j = 1, . . . , km. Since k·kϕ is on BM equivalent to k·kK (by Lemma 7.10(c, d)) we deduce that for each m ∈ N there is ηm > 0 such that is weaker than k·kϕ on BM ϕ = qK(ϕ) j {x ∈ BM ; qK(ϕ)(x) ≤ ηm} ⊆ {x ∈ BM ; qL(x) ≤ 1 m}. The calculation of polars (see the proof of Proposition 7.11) shows that L ⊆ 1 mηm K(ϕ) + 2 m BM∗ . This completes the proof. (cid:3) Now we proceed with the proof of Theorem 10.1. This will be done in several steps. Let us start by explaining the existence of the respective representation. Let M be any JBW∗-triple. By Proposition 9.2 M can be represented as M =Mα∈Λ L∞(µα, Cα) ⊕ N ⊕ sU, where Λ is a set, µα is a probability measure and Cα is a finite-dimensional Cartan factor without unitary element for each α ∈ Λ, N is a JBW∗-algebra, U is a von Neumann algebra, and s ∈ U is a projection such that sU admits no nonzero direct summand isomorphic to a JBW∗-algebra. We may assume without loss of generality that Cs = 1U . By [50, Proposition 6.3.7] there is a unique central projection z ∈ U such that zs is poperly infinite or zero and (1− z)s is finite, hence sU = zsU ⊕ (1 − z)sU . Take W = zU , q = zs, V = (1 − z)U , p = (1 − z)s. Then we have the representation of the form from Theorem 10.1, where p is finite and q properly infinite. Finally, the existence of the relevant decomposition of p follows from [77, Corollary V.2.9]. We continue by proving assertion (a). It is clear that all the elements of T (M ) are σ-finite tripotents. To prove the second statement we will use two lemmata. Lemma 10.4. Let V be a von Neumann algebra, u ∈ V a σ-finite tripotent and (rj)j∈J an orthogonal family of projections. Then the sets are countable. {j ∈ J; rju 6= 0} and {j ∈ J; urj 6= 0} Proof. Note that u, being a tripotent, is a partial isometry with initial projection pi(u) = u∗u and final projection pf (u) = uu∗. Moreover, since u is σ-finite, both pi(u) and pf (u) are σ-finite. Further, it is clear that rj u 6= 0 if and only if rj pf (u) 6= 50 J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER 0, and that urj 6= 0 if and only if pi(u)rj 6= 0. We can therefore assume, without loss of generality, that u is a projection. So, assume u is a projection. Now consider two orthogonal families of cyclic projections in V with sum equal to 1, say (qγ)γ∈Γ and (sδ)δ∈∆, such that u is the sum of a subfamily of the first one and rj is the sum of a subfamily of the second one for each j ∈ J. The existence of these families follows easily from [49, Proposition 5.5.9]. (cid:3) Since {γ ∈ Γ; qγu 6= 0} is countable, [9, Proposition 4.1] implies that both sets {δ ∈ ∆; usδ 6= 0} and {δ ∈ ∆; sδu 6= 0} are countable. Now the assertion easily follows. Lemma 10.5. Let V be a von Neumann algebra and let p ∈ V be a properly infinite projection. Then for any σ-finite projection q ≤ p there is a properly infinite σ-finite projection r such that q ≤ r ≤ p. Proof. Without loss of generality we can assume that p = 1. According to assump- tion V is properly infinite. Therefore there is a sequence (qn) of mutually orthogonal page 97 in [76]). Therefore, there are projections r′ projections such thatPn qn = 1 and qn ∼ 1 for each n (see e.g. Proposition 4.12, each n (cf. Lemma 9.8(c)). Then r′ =Pn r′ n for n is a properly infinite σ-finite projec- tion (cf. [76, Proposition 4.12]) to which q is subequivalent (see Lemma 9.8). Now by Lemma 9.8(c) there is a projection r ≥ q with r ∼ r′. This projection is σ-finite and properly infinite. n with r′ n ≤ qn and q ∼ r′ (cid:3) Now we are ready to prove the second statement of assertion (a). Let h = ((fα)α∈Λ, x, v, w) ∈ M be a σ-finite tripotent. It is clear that spt1 h is countable. For each α ∈ spt1 h choose a complete tripotent gα ∈ L∞(µα, Cα) such that gα ≥ fα, and for each α ∈ Λ \ spt1 h set gα = 0. Further, x is a σ-finite tripotent in N , hence by Lemma 7.2(d) there is a σ-finite projection y ∈ N with x ∈ N2(y). Since v is a σ-finite tripotent in pV , by Lemma 10.4 the set spt2 h is countable. The final projection of v satisfies pf (v) ≤Pj∈spt2 h pj, so by Lemma 9.8(c) there is a projection r ≥ pi(v) such that r ∼ Pj∈spt2 h pj, so we can choose a partial isometry v ∈ V with pi(v) = r and pf (v) =Pj∈spt2 h pj. Finally, w is a σ-finite tripotent in qW , thus pf (w) is a σ-finite projection below q. It follows from Lemma 10.5 that there is a σ-finite properly infinite projection s1 with pf (w) ≤ s1 ≤ q. By Lemma 9.8(c) there is a projection s2 ≥ pi(w) equivalent to s1. Let w be any partial isometry with pi( w) = s2 and pf ( w) = s1. Now it is clear that g = ((gα)α∈Λ, y, v, w) ∈ T (M ) and M2(h) ⊆ M2(g). This completes the proof of assertion (a). We continue by proving (b). Fix ϕ, ψ ∈ M∗ \ {0} such that s(ϕ) = h = ((hα)α∈Λ, x(h), v(h), w(h)), s(ψ) = g = ((gα)α∈Λ, x(g), v(g), w(g)). It is clear that k·kϕ is weaker than k·kψ on BM if and only if MEASURES OF WEAK NON-COMPACTNESS 51 • k·kϕ is weaker than k·kψ on BL∞(µα,Cα) for each α ∈ Λ, • k·kϕ is weaker than k·kψ on BN , • k·kϕ is weaker than k·kψ on BpV , • k·kϕ is weaker than k·kψ on BqW . Observe that Proposition 9.7 yields that k·kϕ and k·kψ are equivalent on the closed unit ball of L∞(µα, Cα) whenever both hα and gα are nonzero. Further, clearly k·kϕ is weaker than k·kψ on BN if and only if x(h) ≤ x(g) (by Proposi- tion 7.5, but in fact this is an easy case). Further, set u(g) = Xj∈spt2 g pj, u(h) = Xj∈spt2 h pj. These are σ-finite projections in V which belong to pV , thus there are ϕ, ψ ∈ (pV )∗ \{0} with s( ϕ) = u(h) and s( ψ) = u(g). By Lemma 9.9 and Proposition 7.5 we see that k·kϕ and k·k ϕ are equivalent on BpV (and k·kψ and k·k ψ as well). Now we deduce, via Proposition 7.5, that k·kϕ is weaker than k·kψ on BpV if and only if spt2 h ⊆ spt2 g. Now assertion (b) follows easily. Next we provide the following postponed proof. Proof of Corollary 10.2. We use the notation from Theorem 10.1. By assertion (a) in the just quoted theorem, for each n ∈ N, we can find an element hn ∈ T (M ) such that M2(s(ϕn)) ⊆ M2(hn) for each n ∈ N. Fix the notation For any α ∈ Sn spt1 hn choose a complete tripotent fα ∈ L∞(µα, Cα) and set fα = 0 for the remaining α ∈ Λ. Further, set x = supn xn and α )α∈Λ, xn, vn, wn), n ∈ N. hn = ((f n v = Xj∈Sn spt2 hn pj . Finally, wn is a partial isometry in W such that its final projection pf (wn) is a properly infinite σ-finite projection below q for each n ∈ N. By Lemma 10.5 there is a σ-finite properly infinite projection r ∈ W with pf (wn) ≤ r ≤ q. sup n By Lemma 9.8(c) we can find, for each n ∈ N, a projection sn ≥ pi(wn) such that sn ∼ r. Then s = supn sn is equivalent to r by Lemma 9.8(a). So, we can fix a partial isometry w ∈ W with initial projection s and final projection r. Then h = ((fα)α∈Λ, x, v, w) ∈ T (M ). Choose ϕ ∈ M∗ \ {0} with s(ϕ) = h. Then k·kϕn Theorem 10.1(b) (and Proposition 7.5). is weaker than k·kϕ on BM by (cid:3) 11. Characterizations of weakly compact sets and operators As a byproduct of our investigation we improve characterizations of weakly com- pact sets in preduals of JBW∗-triples and of weakly compact operators on such spaces. We start by recalling the following known result. 52 J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER Theorem 11.1. [62, Theorem 1.1, Corollary 1.4 and Theorem 1.5] Let K be a bounded subset in the predual of a JBW∗-triple M . Then the following are equiva- lent: (a) K is relatively weakly compact; (b) There exist norm-one normal functionals ϕ1, ϕ2 ∈ M∗ satisfying the following property: Given ε > 0, there exists δ > 0 such that for every x ∈ M with kxk ≤ 1 and kxkϕ1,ϕ2 < δ, we have φ(x) < ε for every φ ∈ K; (c) The restriction, KC , of K to each maximal abelian subtriple C of M is rela- tively weakly compact in C∗; (d) For each tripotent e ∈ M the restriction of K to M2(e) is relatively weakly (e) For any monotone decreasing sequence of tripotents (en) in M with (en) → 0 compact in (M2(e))∗; in the weak∗ topology, we have lim n→+∞ φ(en) = 0 uniformly for φ ∈ K. If M is a JBW∗-algebra then statement (b) can be replaced with the following: (b′) There exists a normal state ψ ∈ M∗ satisfying the following property: Given ε > 0, there exists δ > 0 such that for every x ∈ M with kxk ≤ 1 and kxkψ < δ, we have φ(x) < ε for each φ ∈ K. The equivalence (a) ⇔ (b) is a generalization of Akemann's theorem [1] char- acterizing weakly compact sets in predual of von Neumann algebras. Recall that kxk2 , hence it gives also a more precise version of Proposi- tion 7.9 on the relationship of strong∗ and Mackey topologies. = kxk2 ϕ1,ϕ2 + kxk2 ϕ2 ϕ1 We also notice that a triple C is abelian if the operators L(a, b) and L(x, y) commute for any choice a, b, x, y ∈ C (cf. [16, p. 468]). As observed in [17, pages 340 -- 342] the previous theorem can be applied to char- acterize weakly compact operators from a complex Banach space into the predual of a JBW∗-triple and from a JB∗-triple into a complex Banach space. The concrete result in the latter case reads as follows. Theorem 11.2. [65, Theorem 10] Let E be a JB∗-triple, X a complex Banach space, and T : E → X a bounded linear operator. Then the following assertions are equivalent: (i) T is weakly compact; (ii) There exist norm-one functionals ϕ1, ϕ2 ∈ E∗ and a function N : (0, +∞) → (0, +∞) such that kT (x)k ≤ N (ε) kxkϕ1,ϕ2 + εkxk (iii) There exist a bounded linear operator G from E to a real (respectively, com- for all x ∈ E and ε > 0; plex) Hilbert space and a function N : (0, +∞) → (0, +∞) such that kT (x)k ≤ N (ε)kG(x)k + εkxk for all x ∈ E and ε > 0. This result is collected in the recent monograph [17] as Theorem 5.10.141. By quoting [17], it should be noted that "The above theorem is established in [20, Theorem 11], with k·kϕ1,ϕ2 in condition (ii) replaced with k·kϕ for a single functional ϕ in the unit sphere of E∗. Since this refinement depends on an affirmative answer to [17, Problem 5.10.131], it should remain in doubt." Problem 5.10.131 refers to MEASURES OF WEAK NON-COMPACTNESS 53 the so-called Barton-Friedman conjecture for JB∗-triples and the subtle difficulties appearing around the original statement of Grothendieck's inequality for JB∗-triples published in [6] (see [60, 65, 61], [17, Subsection 5.10.4], [35] and the final remark in page 55 for more details). Summarizing, the problem whether in Theorem 11.2(ii) (respectively, Theorem 11.1(b)) the seminorm of the form k·kϕ1,ϕ2 can be replaced with a seminorm of the form k·kϕ for a single norm-one functional ϕ ∈ E∗ remains as an open question. Our next result provides a positive solution to these problems and proves the validity of the original statement in [20, Theorem 11]. Theorem 11.3. Let K be a bounded subset in the predual of a JBW∗-triple M . Then K is relatively weakly compact if and only if there exists a norm-one normal functional ϕ ∈ M∗ satisfying the following property: Given ε > 0, there exists δ > 0 such that for every x ∈ M with kxk ≤ 1 and kxkϕ < δ, we have φ(x) < ε for every φ ∈ K. Proof. The 'if part' follows from the implication (b) ⇒ (a) in Theorem 11.1, whereas the 'only if part follows from the implication (a) ⇒ (b) in Theorem 11.1 and Corol- lary 10.2. (cid:3) We can now provide a proof of the statement in [20, Theorem 11] and close a conjecture which has remained open for over eighteen years. The proof dissipates the commented doubts expressed in [17, page 341]. Theorem 11.4. Let M be a JBW∗-triple, E a JB∗-triple, and let X be a complex Banach space. Then the following statements hold: (a) A bounded linear operator T : X → M∗ is weakly compact if and only if there exists a norm-one functional ϕ ∈ M∗ and a function N : (0, +∞) → (0, +∞) such that kT ∗(a)k ≤ N (ε) kakϕ + εkak for all a ∈ M and ε > 0; (b) A bounded linear operator T : E → X is weakly compact if and only if there exists a norm-one functional ϕ ∈ E∗ and a function N : (0, +∞) → (0, +∞) such that for all a ∈ E and ε > 0. kT (a)k ≤ N (ε) kakϕ + εkak Proof. In both statements the 'if parts' follow from Theorem 11.2. (a) Suppose T : X → M∗ is weakly compact. Since T (BX) ⊆ M∗ is relatively weakly compact, Theorem 11.3 implies the existence of a norm-one normal func- tional ϕ ∈ M∗ satisfying the following property: Given ε > 0, there exists δ > 0 such that for every a ∈ M with kak ≤ 1 and kakϕ < δ, we have T (x)(a) < ε for every x ∈ BX . a Given a ∈ M\{0}, the element b = kak+δ−1kakϕ ∈ BM and satisfies kbkϕ < δ, therefore T (x)(b) < ε for every x ∈ BX , equivalently, T ∗(a)(x) = T (x)(a) < εδ−1kakϕ + εkak, for all x ∈ BX , and thus kT ∗(a)k ≤ εδ−1kakϕ + εkak, for all a ∈ M . Statement (b) follows from (a) since, by virtue of Gantmacher's theorem, an (cid:3) operator T : E → X is weakly compact if and only if T ∗ : X ∗ → E∗ is. 54 J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER 12. Final remarks and open problems Theorem 4.1 says that the three measures of weak non-compactness considered in this paper coincide in preduals of JBW∗-triples which complements the previous results from [46, 34]. However, the mentioned results include explicit formulas for these measures. In fact, these formulas are substantially used in the proofs. In the present paper we do not get an explicit formula due to the procedure of the proof -- we use subsequence splitting property, Lemma 8.3, Proposition 3.2 and Lemma 2.1. So it is natural to ask whether there are some natural formulas for the De Blasi measure ω. The first question deals with a special case of von Neumann algebras. Question 12.1. Let M be a semifinite von Neumann algebra with a fixed normal semifinite faithful trace τ . Is there a formula for the De Blasi measure of weak noncompatness in M∗ in terms of the trace τ ? Is it so at least for finite σ-finite von Neumann algebras? We note that the special case of commutative spaces L1(µ) is settled in [46, Section 7]. Another possibility is to try to get quantitative versions of some characterizations of weakly compact sets in the predual of a JBW∗-triple given in Theorem 11.1 (or in the improvement of its assertion (b) contained in Theorem 11.3). More precisely, we have the following question. Question 12.2. Let M be a JBW∗-triple and let A ⊆ M∗ be a bounded set. Can ω(A) be expressed using quantitative versions of the characterizations from The- orem 11.1? In particular, is ω(A) equal (or at least equivalent) to the following quantities? (b) inf inf δ>0 sup{φ(x) ; x ∈ BM ,kxkϕ < δ}; ϕ∈SM∗ sup φ∈A (c) sup{ω(AC ); C ⊆ M a maximal abelian subtriple}; (d) sup{ω(AM2(e)); e ∈ M a tripotent}; (e) sup(cid:26)lim sup n→∞ sup φ∈Aφ(en) ; (en) a decreasing sequence of tripotents weak∗-converging to 0 (cid:27) . Note that these quantities naturally correspond to the respective characteriza- tions in Theorem 11.1 (or in the improvement of assertion (b) contained in Theo- rem 11.3). It is easy to check that all these quantities are bounded above by the De Blasi measure ω, but the converse inequalities seem not to be obvious. We investigated measures of weak non-compactness in preduals of JBW∗-triples. Another possibility is to look at subsets of JB∗-triples themselves. In this direction there is just one positive result -- coincidence of measures of weak non-compactness for subsets of c0(Γ) (see [34, Theorem C]) and no negative result up to now. So, the following question seems to be natural. Question 12.3. Are the above-considered measures of weak non-compactness equi- valent (or even equal) for bounded subsets of a JBW∗-triple? We have no idea how to approach this general question, so we formulate two important special cases. Question 12.4. Let K be a compact space. Are the above-considered measures of weak non-compactness equivalent (or even equal) for bounded subsets of C(K)? Is it true for K = [0, 1]? MEASURES OF WEAK NON-COMPACTNESS 55 Question 12.5. Let H be a Hilbert space. Are the above-considered measures of weak non-compactness equivalent (or even equal) for bounded subsets of K(H), the space of compact operators on H? Added during the revision process: Months after the submission of this paper we discovered a complete proof of the so-called Barton-Friedman conjecture for general JB∗-triples, and a solution to Problem 5.10.131 in [17] (treated in page 53). The result is included in the recent preprint [35]. This proof of the Barton- Friedman conjecture offers an alternative approach to derive Theorems 11.3 and 11.4 as a straightforward consequences of [65, Theorem 10]. Acknowledgements: We would like to thank the anonymous referee for the time em- ployed in writing a professional and thorough report with a wide list of constructive and enriching comments and suggestions. References [1] Akemann, C. A. The dual space of an operator algebra. Trans. Amer. Math. Soc. 126 (1967), 286 -- 302. [2] Aliprantis, C. D., and Border, K. C. Infinite-dimensional analysis, second ed. Springer- Verlag, Berlin, 1999. A hitchhiker's guide. [3] Angosto, C., and Cascales, B. The quantitative difference between countable compactness and compactness. J. Math. Anal. Appl. 343, 1 (2008), 479 -- 491. [4] Angosto, C., and Cascales, B. Measures of weak noncompactness in Banach spaces. Topol- ogy Appl. 156, 7 (2009), 1412 -- 1421. [5] Astala, K., and Tylli, H.-O. Seminorms related to weak compactness and to Tauberian operators. Math. Proc. Cambridge Philos. Soc. 107, 2 (1990), 367 -- 375. [6] Barton, T., and Friedman, Y. Grothendieck's inequality for J B∗-triples and applications. J. London Math. Soc. (2) 36, 3 (1987), 513 -- 523. [7] Barton, T., and Friedman, Y. Bounded derivations of JB∗-triples. Quart. J. Math. Oxford Ser. (2) 41, 163 (1990), 255 -- 268. [8] Barton, T., and Timoney, R. M. Weak∗-continuity of Jordan triple products and its ap- plications. Math. Scand. 59, 2 (1986), 177 -- 191. [9] Bohata, M., Hamhalter, J., and Kalenda, O. F. K. On Markushevich bases in preduals of von Neumann algebras. Israel J. Math. 214, 2 (2016), 867 -- 884. [10] Bohata, M., Hamhalter, J., and Kalenda, O. F. K. Decompositions of preduals of JBW- and JBW∗- algebras. J. Math. Anal. Appl. 446, 1 (2017), 18 -- 37. [11] Bohata, M., Hamhalter, J., Kalenda, O. F. K., Peralta, A. M., and Pfitzner, H. Preduals of JBW*-triples are 1-Plichko spaces. Q. J. Math. 69, 2 (2018), 655 -- 680. [12] Bourgain, J., and Rosenthal, H. P. Martingales valued in certain subspaces of L1. Israel J. Math. 37, 1-2 (1980), 54 -- 75. [13] Braun, R., Kaup, W., and Upmeier, H. A holomorphic characterization of Jordan C ∗- algebras. Math. Z. 161, 3 (1978), 277 -- 290. [14] Bunce, L. J. Norm preserving extensions in JBW∗-triple preduals. Q. J. Math. 52, 2 (2001), 133 -- 136. [15] Bunce, L. J., Fern´andez-Polo, F. J., Mart´ınez Moreno, J., and Peralta, A. M. A Saito-Tomita-Lusin theorem for JB*-triples and applications. Q. J. Math. 57, 1 (2006), 37 -- 48. [16] Cabrera Garc´ıa, M., and Rodr´ıguez Palacios, A. Non-associative normed algebras. Vol. 1, vol. 154 of Encyclopedia of Mathematics and its Applications. Cambridge University Press, Cambridge, 2014. The Vidav-Palmer and Gelfand-Naimark theorems. [17] Cabrera Garc´ıa, M., and Rodr´ıguez Palacios, A. Non-associative normed algebras. Vol. 2, vol. 167 of Encyclopedia of Mathematics and its Applications. Cambridge University Press, Cambridge, 2018. Representation theory and the Zel'manov approach. [18] Cascales, B., Kalenda, O. F. K., and Spurn´y, J. A quantitative version of James's com- pactness theorem. Proc. Edinb. Math. Soc. (2) 55, 2 (2012), 369 -- 386. 56 J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER [19] Chu, C.-H. Jordan structures in geometry and analysis, vol. 190 of Cambridge Tracts in Mathematics. Cambridge University Press, Cambridge, 2012. [20] Chu, C.-H., and Iochum, B. Weakly compact operators on Jordan triples. Math. Ann. 281, 3 (1988), 451 -- 458. [21] De Blasi, F. S. On a property of the unit sphere in a Banach space. Bull. Math. Soc. Sci. Math. R. S. Roumanie (N.S.) 21(69), 3-4 (1977), 259 -- 262. [22] Diestel, J. Sequences and series in Banach spaces, vol. 92 of Graduate Texts in Mathematics. Springer-Verlag, New York, 1984. [23] Edwards, C. M., and Ruttimann, G. T. On the facial structure of the unit balls in a JBW∗-triple and its predual. J. London Math. Soc. (2) 38, 2 (1988), 317 -- 332. [24] Edwards, C. M., and Ruttimann, G. T. Exposed faces of the unit ball in a JBW∗-triple. Math. Scand. 82, 2 (1998), 287 -- 304. [25] Fabian, M., Habala, P., H´ajek, P., Montesinos Santaluc´ıa, V., Pelant, J., and Zi- zler, V. Functional analysis and infinite-dimensional geometry. CMS Books in Mathemat- ics/Ouvrages de Math´ematiques de la SMC, 8. Springer-Verlag, New York, 2001. [26] Fabian, M., H´ajek, P., Montesinos, V., and Zizler, V. A quantitative version of Krein's theorem. Rev. Mat. Iberoamericana 21, 1 (2005), 237 -- 248. [27] Fern´andez-Polo, F. J., Peralta, A. M., and Ram´ırez, M. I. A Kadec-Pelczy´nski dichotomy-type theorem for preduals of JBW∗-algebras. Israel J. Math. 208, 1 (2015), 45 -- 78. [28] Friedman, Y., and Russo, B. Solution of the contractive projection problem. J. Funct. Anal. 60, 1 (1985), 56 -- 79. [29] Friedman, Y., and Russo, B. Structure of the predual of a J BW ∗-triple. J. Reine Angew. Math. 356 (1985), 67 -- 89. [30] Friedman, Y., and Russo, B. The Gel′fand-Naımark theorem for JB∗-triples. Duke Math. J. 53, 1 (1986), 139 -- 148. [31] Granero, A. S. An extension of the Krein-Smulian theorem. Rev. Mat. Iberoam. 22, 1 (2006), 93 -- 110. [32] Granero, A. S., Hern´andez, J. M., and Pfitzner, H. The distance dist(B, X) when B is a boundary of B(X ∗∗). Proc. Amer. Math. Soc. 139, 3 (2011), 1095 -- 1098. [33] Grothendieck, A. Espaces vectoriels topologiques. Instituto de Matem´atica Pura e Aplicada, Universidade de Sao Paulo, Sao Paulo, 1954. [34] Hamhalter, J., and Kalenda, O. F. K. Measures of weak non-compactness in spaces of nuclear operators. Math. Z. 292, 1-2 (2019), 453 -- 471. [35] Hamhalter, J., Kalenda, O. F. K., Peralta, A. M., and Pfitzner, H. Grothendieck's inequalities for JB-triples: Proof of the Barton-Friedman conjecture. arXiv:1903.08931. [36] Hanche-Olsen, H., and Størmer, E. Jordan operator algebras, vol. 21. Pitman Advanced Publishing Program, 1984. [37] Harmand, P., Werner, D., and Werner, W. M -ideals in Banach spaces and Banach algebras, vol. 1547 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, 1993. [38] Ho, T., Martinez-Moreno, J., Peralta, A. M., and Russo, B. Derivations on real and complex JB ∗-triples. J. London Math. Soc. (2) 65, 1 (2002), 85 -- 102. [39] Horn, G. Klassifikation der JBW∗-Tripel vomm Type 1. Dissertation. University of Tubingen, 1984. [40] Horn, G. Characterization of the predual and ideal structure of a JBW∗-triple. Math. Scand. 61, 1 (1987), 117 -- 133. [41] Horn, G. Classification of JBW∗-triples of type I. Math. Z. 196, 2 (1987), 271 -- 291. [42] Horn, G., and Neher, E. Classification of continuous J BW ∗-triples. Trans. Amer. Math. Soc. 306, 2 (1988), 553 -- 578. [43] Iochum, B. Cones autopolaires et alg`ebres de Jordan, vol. 1049 of Lecture Notes in Mathe- matics. Springer-Verlag, Berlin, 1984. [44] Isidro, J. M., Kaup, W., and Rodr´ıguez-Palacios, A. On real forms of JB∗-triples. Manuscripta Math. 86, 3 (1995), 311 -- 335. [45] Jarchow, H. On weakly compact operators on C ∗-algebras. Math. Ann. 273, 2 (1986), 341 -- 343. [46] Kacena, M., Kalenda, O. F. K., and Spurn´y, J. Quantitative Dunford-Pettis property. Adv. Math. 234 (2013), 488 -- 527. [47] Kadec, M. I., and Pe lczy´nski, A. Bases, lacunary sequences and complemented subspaces in the spaces Lp. Studia Math. 21 (1961/1962), 161 -- 176. MEASURES OF WEAK NON-COMPACTNESS 57 [48] Kadison, R. V. Isometries of operator algebras. Ann. Of Math. (2) 54 (1951), 325 -- 338. [49] Kadison, R. V., and Ringrose, J. R. Fundamentals of the theory of operator algebras. Vol. I, vol. 100 of Pure and Applied Mathematics. Academic Press, Inc. [Harcourt Brace Jovanovich, Publishers], New York, 1983. Elementary theory. [50] Kadison, R. V., and Ringrose, J. R. Fundamentals of the theory of operator algebras. Vol. II, vol. 16 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 1997. Advanced theory, Corrected reprint of the 1986 original. [51] Kalenda, O., Pfitzner, H., and Spurn´y, J. On quantification of weak sequential complete- ness. J. Funct. Anal. 260, 10 (2011), 2986 -- 2996. [52] Kaup, W. A Riemann mapping theorem for bounded symmetric domains in complex Banach spaces. Math. Z. 183, 4 (1983), 503 -- 529. [53] Kaup, W. Contractive projections on Jordan C ∗-algebras and generalizations. Math. Scand. 54, 1 (1984), 95 -- 100. [54] Kaup, W. On real Cartan factors. Manuscripta Math. 92, 2 (1997), 191 -- 222. [55] Kaup, W., and Upmeier, H. Jordan algebras and symmetric Siegel domains in Banach spaces. Math. Z. 157, 2 (1977), 179 -- 200. [56] Knaust, H., and Odell, E. On c0 sequences in Banach spaces. Israel J. Math. 67, 2 (1989), 153 -- 169. [57] Loos, O. Bounded symmetric domains and Jordan pairs. Lecture Notes, Univ. California at Irvine, 1977. [58] Neher, E. Jordan triple systems by the grid approach, vol. 1280 of Lecture Notes in Mathe- matics. Springer-Verlag, Berlin, 1987. [59] Pedersen, G. K. C ∗-algebras and their automorphism groups, vol. 14 of London Mathe- matical Society Monographs. Academic Press, Inc. [Harcourt Brace Jovanovich, Publishers], London-New York, 1979. [60] Peralta, A. M. Little Grothendieck's theorem for real JB∗-triples. Math. Z. 237, 3 (2001), 531 -- 545. [61] Peralta, A. M. New advances on the Grothendieck's inequality problem for bilinear forms on JB*-triples. Math. Inequal. Appl. 8, 1 (2005), 7 -- 21. [62] Peralta, A. M. Some remarks on weak compactness in the dual space of a JB*-triple. Tohoku Math. J. (2) 58, 2 (2006), 149 -- 159. [63] Peralta, A. M. Positive definite hermitian mappings associated with tripotent elements. Expo. Math. 33, 2 (2015), 252 -- 258. [64] Peralta, A. M., and Pfitzner, H. The Kadec-Pe lczy´nski-Rosenthal subsequence splitting lemma for JBW∗-triple preduals. Studia Math. 227, 1 (2015), 77 -- 95. [65] Peralta, A. M., and Rodr´ıguez Palacios, A. Grothendieck's inequalities for real and complex JBW∗-triples. Proc. London Math. Soc. (3) 83, 3 (2001), 605 -- 625. [66] Randrianantoanina, N. Kadec-Pe lczy´nski decomposition for Haagerup Lp-spaces. Math. Proc. Cambridge Philos. Soc. 132, 1 (2002), 137 -- 154. [67] Raynaud, Y., and Xu, Q. On subspaces of non-commutative Lp-spaces. J. Funct. Anal. 203, 1 (2003), 149 -- 196. [68] Rodr´ıguez-Palacios, A. On the strong∗ topology of a JBW∗-triple. Quart. J. Math. Oxford Ser. (2) 42, 165 (1991), 99 -- 103. [69] Rodr´ıguez-Palacios, A. Jordan structures in analysis. In Jordan algebras (Oberwolfach, 1992). de Gruyter, Berlin, 1994, pp. 97 -- 186. [70] Rodr´ıguez Palacios, A., and Cabrera Garc´ıa, M. A new proof of the Barton-Timoney theorem on the bidual of a J B∗-triple. Math. Nachr. 292, 3 (2019), 640 -- 644. [71] Ryan, R. A. Introduction to tensor products of Banach spaces. Springer Science & Business Media, 2013. [72] Sakai, S. C*-algebras and W*-algebras. Springer Science & Business Media, 2012. [73] Schluchtermann, G., and Wheeler, R. F. On strongly WCG Banach spaces. Math. Z. 199, 3 (1988), 387 -- 398. [74] Sherman, D. On the dimension theory of von Neumann algebras. Math. Scand. 101, 1 (2007), 123 -- 147. [75] Stacho, L. L. A projection principle concerning biholomorphic automorphisms. Acta Sci. Math 44 (1982), 99 -- 124. 58 J. HAMHALTER, O.F.K. KALENDA, A.M. PERALTA, AND H. PFITZNER [76] Stratila, S¸., and Zsid´o, L. Lectures on von Neumann algebras. Editura Academiei, Bucharest; Abacus Press, Tunbridge Wells, 1979. Revision of the 1975 original, Translated from the Romanian by Silviu Teleman. [77] Takesaki, M. Theory of operator algebras. I. Springer-Verlag, New York-Heidelberg, 1979. [78] Wright, J. D. M. Jordan C ∗-algebras. Michigan Math. J. 24, 3 (1977), 291 -- 302. [79] Youngson, M. A. A Vidav theorem for Banach Jordan algebras. Math. Proc. Cambridge Philos. Soc. 84, 2 (1978), 263 -- 272. Czech Technical University in Prague, Faculty of Electrical Engineering, Depart- ment of Mathematics, Technicka 2, 166 27, Prague 6, Czech Republic E-mail address: [email protected] Charles University, Faculty of Mathematics and Physics, Department of Mathemat- ical Analysis, Sokolovsk´a 86, 186 75 Praha 8, Czech Republic E-mail address: [email protected] Departamento de An´alisis Matem´atico, Facultad de Ciencias, Universidad de Grana- da, 18071 Granada, Spain. E-mail address: [email protected] Institut Denis Poisson, Universit´e d'Orl´eans, Universit´e de Tours, CNRS, Rue de Chartres, BP 6759, F-45067 Orl´eans Cedex 2, France E-mail address: [email protected]
1203.5548
1
1203
2012-03-25T22:05:43
Classification of Noncommutative Domain Algebras
[ "math.OA" ]
Noncommutative domain algebras are noncommutative analogues of the algebras of holomorphic functions on domains of $\C^n$ defined by holomorphic polynomials, and they generalize the noncommutative Hardy algebras. We present here a complete classification of these algebras based upon techniques inspired by multivariate complex analysis, and more specifically the classification of domains in hermitian spaces up to biholomorphic equivalence.
math.OA
math
CLASSIFICATION OF NONCOMMUTATIVE DOMAIN ALGEBRAS ALVARO ARIAS AND FRÉDÉRIC LATRÉMOLIÈRE Abstract. Noncommutative domain algebras are noncommutative analogues of the algebras of holomorphic functions on domains of Cn defined by holo- morphic polynomials, and they generalize the noncommutative Hardy alge- bras. We present here a complete classification of these algebras based upon techniques inspired by multivariate complex analysis, and more specifically the classification of domains in hermitian spaces up to biholomorphic equivalence. 1. Introduction Noncommutative domain algebras, introduced by Popescu in [11] are universal objects in the category of operator algebras for certain polynomial relations and are noncommutative analogues of algebras of holomorphic functions on domains in hermitian spaces, as well as generalizations of Hardy Algebras. This note presents a classification of these algebras up to completely isometric isomorphism, based upon their defining symbols, and thus complete the work initiated by the authors in [2] and [3]. Our methods are based on the deep interplay between analysis of this type of operator algebras and analysis of multivariate holomorphic functions. This note also exploits an observation made in [9] which helps us conclude Theorem 2.2. Algebras of weighted shifts on Fock spaces [10, 5, 7, 6] find their origin in the study of row contractions, dilations, and commutant lifting theorems. Their ap- plications became important in interpolation problems [8, 4] and the study of E0- semigroups [9] or noncommutative complex analysis [2, 1, 3]. In [2] and [3], the authors of this note studied the geometry of the spectra of noncommutative domains. They used Thullen's [13] and Sunada's [12] classification of Reinhardt domains and combinatorial arguments to fully classify a large class of domain algebras. However, one important case remained unsolved, and this note present the solution for this last unresolved class, thus completing the classification of all noncommutative domain algebras with polynomial symbols. Noncommutative domain algebras are naturally -- and best -- represented as norm closed algebras of operators on full Fock spaces. For any (complex) Hilbert space H , the associated full Fock space, denoted by F (H ), is the Hilbert sum ⊕n∈NH ⊗n with the convention that H 0 = C. It is useful for the definition of noncommutative domain algebras to introduce the following alternative description of F (H ). Let Fn + be the free semigroup over {1, . . . , n} whose neutral element will be denoted by ∅ by abuse of notation. For any elements t1, . . . , tn of any unital associative algebra with unit 1, and for any nonempty word α = i1 · · · ik in Fn +, Date: October 14, 2018. 2000 Mathematics Subject Classification. (Primary) 47L15, (Secondary) 47A63, 46L52, 32A07. Key words and phrases. Non-selfadjoint operator algebras, weighted shifts, biholomorphisms. 1 2 ALVARO ARIAS AND FRÉDÉRIC LATRÉMOLIÈRE with i1, . . . , ik ∈ {1, . . . , k}, we denote by tα the product Qk j=1 tij , while t∅ will In particular, F (Cn) is a unital associative algebra be set to 1 by convention. for the tensor product. We fix henceforth the canonical basis {e1, . . . , en} of Cn is a Hilbert basis of F (Cn), which allows us to identify and observe that (eα)α∈Fn F (Cn) with ℓ2(Fn + +). +, where each generator i ∈ {1, . . . , n} of Fn Fix n ∈ N, n > 0. Let Pn = C[X1, . . . , Xn] be the algebra of polynomials in n noncommuting indeterminates X1, . . . , Xn -- which is the free associative algebra generated by Fn + is identified with Xi. aαXα for an almost zero family aαXα ∈ Pn and any Hilbert (aα)α∈Fn space H whose C*-algebra of bounded linear operators will be denoted by B(H ), the noncommutative domain Df (H ) is defined by: + + + Thus any element of Pn is of the form Pα∈Fn of complex numbers. For any f =Pα∈Fn Df (H ) = f = Xα∈Fn (T1, . . . , Tn) ∈ B(H )n : Xα∈Fn aαXα(cid:12)(cid:12)a∅ = 0, ∀α ∈ Fn where we used the notation T ∗ α for (Tα)∗. + + If f is an element of the subset Sn ⊆ Pn defined by: + aα ≥ 0 ∧ (α = 1 =⇒ aα > 0) , aαTαT ∗ α ≤ 1H , Sn = fix an n-symbol f =Pα∈Fn j (eα) =s bα F (Cn) by extending: (1.1) W f bj·α + where α is the word-length of α ∈ Fn +, one can construct an explicit, universal n-tuples of operators in Df (F (Cn)). We shall refer to elements of Sn as n-symbols, while Popescu refers to them as positive regular n-free formal polynomials. Let us n on aαXα. We define the weighted shifts W f 1 , . . . , W f (ej ⊗ eα) where bα = αXk=1 Xγ1···γk=α γ1≥1,··· ,γk≥1 aγ1 · · · aγk for all α ∈ Fn +. We then have the following fundamental universal property: 1 , . . . , W f Theorem 1.1. (Popescu, [11]) Let n ∈ N, n > 0 and let f ∈ Sn. Let Af be the norm closure of the associative algebra generated by the set of operators {W f n ) ∈ Df (F (Cn)). Moreover, for any Hilbert space H and any (T1, . . . , Tn) ∈ Df (H ), there exists a unique completely contractive algebra morphism ϕ from Af onto the norm closure of the algebra generated by T1, . . . , Tn such that ϕ(W f n } defined by (1.1) on F (Cn). We have (W f j ) = Tj for j = 1, . . . , n. 1 , . . . , W f The purpose of this note is to completely classify the noncommutative domain algebras Af defined for f ∈ Sn in Theorem (1.1) in term of their symbol f . 2. Main result We define the following equivalence relation on the set Sn of symbols [3, Defi- nition 2.1]: CLASSIFICATION OF NONCOMMUTATIVE DOMAINS 3 Definition 2.1. Let n ∈ N, n > 0. Two elements f, g ∈ Sn are scale-permutation equivalent when there exists a permutation σ of {1, . . . , n} and scalars λ1, . . . , λn ∈ R such that f (X1, . . . , Xn) = g(λ1Xσ(1), . . . , λnXσ(n)). Our main theorem is the following complete classification result: Theorem 2.2. Let f ∈ Sn and g ∈ Sm be two symbols. The noncommutative domain algebras Af and Ag are completely isometrically isomorphic if and only if n = m and f and g are scale-permutation equivalent. Before we provide the proof of this result, let us recall from [2] the following fundamental duality construction upon which our work relies. Let f ∈ Sn for some n ∈ N, n > 0. Let k ∈ N, k > 0. Let T = (T1, . . . , Tn) ∈ Df (Ck). By universality of Af , there exists a completely contractive morphism hT, ·ik : Af → B(Ck) such = Tj for j = 1, . . . , n. Moreover, the function h·, aik is holomorphic on the interior of Df (Ck) and extends by continuity to Df (Ck). We shall abuse the terminology and call such a function holomorphic on the compact Df (Ck). jEk thatDT, W f Proof. In this proof, isomorphisms are always meant for completely isometric iso- morphisms of operator algebras. By [2, Lemma 4.4], whenever two symbols are scale-permutation equivalent, their associated noncommutative domains are isomorphic. It is thus sufficient to prove the converse here. Let us denote the set of all isomorphisms from Af to Ag by I (Af , Ag). By [2, Theorem 3.7], for any Ψ ∈ I (Af , Ag), there exists a necessarily unique biholomor- . Moreover, by [2, Theorem 3.18] and [3, Thus, let us assume that there exists an isomorphism Φ : Af → Ag. is also a disk. remains to consider the case where Df (C) and Dg(C) are Bn, up to replacing f and g by scale-permutation equivalent symbols. f and g are scale-permutation equivalent. Henceforth, we shall assume ω 6= 0. We In [3, Theorem 3.4], we were able to show that if either Df (C) or Dg(C) is not biholo- j=1 zjzj ≤ 1} of Cn, then phic map bΨ : Dg(C) → Df (C) such that, for all λ = (λ1, · · · , λn) ∈ Dg(C) and a ∈ Af , we have hλ, Ψ(a)i1 =DbΨ(λ), aE1 Theorem 3.2], if bΨ(0) = 0 then f and g are scale-permutation equivalent. morphic to the unit ball Bn = {(z1, . . . , zn) ∈ Cn : Pn bΨ(0) = 0, which implies that f and g are scale-permutation equivalent. Thus, it Let ω =bΦ−1(0). If ω = 0 then by [3, Theorem 3.2], we can already conclude that adapt an argument in [9]. Let Dg = Cω∩Bn and Df =bΦ(Dg). By construction, Dg is a disk in the plane Cω. Now, as a conformal self-map of the unit ball, bΦ = Υ ◦ ϕω is the identity and ϕω(0) = ω. Thus bΦ ◦ ϕω(0) = 0 and hence, by Cartan's Lemma, bΦ ◦ ϕω is a unitary Υ. Hence bΦ =bΦ ◦ ϕω ◦ ϕω = Υ ◦ ϕω. Now, ϕω maps the plane Cω to itself by construction, so Df = bΦ(Dg) = Υ(Dg), and since Υ is unitary, Df We now set G = {z ∈ Dg : ∃Ψ ∈ I (Af , Ag) bΨ(z) = 0} and F = {z ∈ Df : ∃Ψ ∈ I (Af , Af ) bΨ(0) = z}. The first observation is that G and F are circular domains. Indeed, note first that for any λ ∈ C, λ = 1, we can extend the function which, to each j ∈ {1, . . . , n}, maps W g j , to an automorphism Λ of Ag where ϕω is the conformal map ϕω = ψω ω−x2 , and Υ is unitary. Indeed, ϕω is a easily seen to be a conformal map such that ϕω ◦ϕω ψω2 where ψω : z ∈ Bn 7→ ω+(1−ω2) ω−x j to λW g 4 ALVARO ARIAS AND FRÉDÉRIC LATRÉMOLIÈRE by [2, Lemma 4.4]. Now, for any b ∈ G there exists Ψ ∈ I (Af , Ag) such that Now, by construction, ω ∈ G. Thus G contains the circle Tg = {λω : λ = 1}. λb ∈ G since λb ∈ Dg as Dg is a disk, hence circular. The same argument of course applies to F . bΨ(b) = 0. We have \Λ ◦ Ψ = bΨ ◦bΛ so \Λ ◦ Ψ(λb) = 0 and Ψ ◦ Λ ∈ I (Af , Ag), so Since bΦ restricted to the disk Dg is a Möbius map whose poles lie outside of Dg, it preserves circles. Let Tf = bΦ(Tg). On the other hand, if b ∈ G then there exists Ψ ∈ I (Af , Ag) such that bΨ(b) = 0, so bΦ(b) =bΦ ◦bΨ−1(0) = \Ψ−1 ◦ Φ(0) and Ψ ◦ Φ ∈ I (Af , Af ). Hence bΦ(b) ∈ F . Since ω ∈ Tg we conclude that 0 ∈ Tf . checks easily that F contains the interior of Tf . Since G = bΦ−1(F ) by as similar argument as above, we conclude that G contains the interior of Tg = bΦ−1(Tf ), which in turn contains 0. Hence, there exists Ψ ∈ I (Af , Ag) such that bΨ(0) = 0. Hence, F contains the circle Tf containing the origin. Since F is circular, one By [3, Theorem 3.2], we conclude that f and g are scale-permutation equivalent. (cid:3) References 1. A. Arias and F. Latrémolière, Ergodic actions of convergent fuchsian groups on noncommuta- tive hardy algebras, Proc. Amer. Math. Soc. 139 (2011), no. 7, 2485 -- 2496, ArXiv: 10105840. , Isomorphisms of non-commutative domain algebras, J. Oper. Theory 66 (2011), 2. no. 2, 425 -- 450, ArXiv: 09020195. 3. , Isomorphisms of non-commutative domains II, Accepted, Journal of Operator Theory (2011), 15 pages., ArXiv: 1010.5838. 4. A. Arias and G. Popescu, Noncommutative interpolation and Poisson transforms, Israel J. Math. 115 (2000), 205 -- 234. 5. W. Arveson, Subalgebras of C ∗-algebras. III. multivariable operator theory, Acta Math. 181 (1998), no. 2, 159 -- 228. 6. K. Davidson, E. Katsoulis, and D. Pitts, The structure of free semigroup algebras, J. Reine Angew. Math. 533 (2001), 99 -- 125. 7. K. Davidson and D. Pitts, The algebraic structure of non-commutative analytic toeplitz alge- bras, Math. Ann. 311 (1998), 275 -- 303. 8. K. Davidson and D. Pitts, Nevanlinna-Pick interpolation for non-commutative analytic toeplitz algebras, Integral Equations and Operator Theory 31 (1998), no. 3, 321 -- 337. 9. K. Davidson, C. Ramsey, and Shalit O., The isomorphism problem for some universal operator algebras, Advances in Math. 228 (2011), no. 1, 167 -- 218. 10. G. Popescu, Von Neumann inequality for (B(H)n)1, Math. Scand. 68 (1991), no. 2, 292 -- 304. , Operator theory on noncommutative domains, Memoirs of the American Mathemat- 11. ical Society 205 (2010), no. 954, vi+124. 12. T. Sunada, Holomorphic equivalence problem for bounded reinhardt domains, Math. Ann. 235 (1978), 111 -- 128. 13. P. Thullen, Zu den abbildungen durch analytische funktionen mehrerer veränderlichen, Math. Ann. 104 (1931), 244 -- 259. Department of Mathematics, University of Denver, Denver CO 80208 E-mail address: [email protected] URL: http://www.math.du.edu/~aarias Department of Mathematics, University of Denver, Denver CO 80208 E-mail address: [email protected] URL: http://www.math.du.edu/~frederic
1908.07775
1
1908
2019-08-21T10:09:47
Quantum Euclidean Spaces with Noncommutative Derivatives
[ "math.OA" ]
Quantum Euclidean spaces, as Moyal deformations of Euclidean spaces, are the model examples of noncompact noncommutative manifold. In this paper, we study the quantum Euclidean space equipped with partial derivatives satisfying canonical commutation relation (CCR). This gives an example of semi-finite spectral triple with non-flat geometric structure. We develop an abstract symbol calculus for the pseudo-differential operators with noncommuting derivatives. We also obtain a simplified local index formula (even case) that is similar to the commutative setting.
math.OA
math
QUANTUM EUCLIDEAN SPACES WITH NONCOMMUTATIVE DERIVATIVES LI GAO, MARIUS JUNGE∗, AND EDWARD MCDONALD Abstract. Quantum Euclidean spaces, as Moyal deformations of Euclidean spaces, are the model examples of noncompact noncommutative manifold. In this paper, we study the quantum Euclidean space equipped with partial derivatives satisfying canonical com- mutation relation (CCR). This gives an example of semi-finite spectral triple with non-flat geometric structure. We develop an abstract symbol calculus for the pseudo-differential operators with noncommuting derivatives. We also obtain a simplified local index formula (even case) that is similar to the commutative setting. 9 1 0 2 g u A 1 2 ] . A O h t a m [ 1 v 5 7 7 7 0 . 8 0 9 1 : v i X r a 1. Introduction The theory of pseudo-differential operators (ΨDOs) plays an influential role in the index theory of elliptic operators. This approach also prevails in noncommutative geometry. In [CM95], Connes and Moscovici established the local index formula for spectral triples, which gives an analytic expression for the index pairing between K-theory of noncommutative algebras and the K-homology class induced by a Dirac type operator. This local index formula was extended to the locally compact (i.e., non-unital) setting by Carey, Gayral, Rennie and Sukochev [CGRS14]. In both proofs of the local index formula [CM95, CGRS14], an abstract theory of ΨDOs is crucial to the analysis. On the prototypical example of a noncommutative geometry -- quantum tori, pseudo-differential operators been widely used in studying curvatures and other geometric structures (see e.g. [CT11, FK13, LM16, BM12, CM14]). Recently several works [Tao18, HLP18a, HLP18b, GJP17] give detailed accounts of the symbol calculus for ΨDOs on quantum tori. Quantum Euclidean spaces are model examples of noncommutative spaces in the locally compact setting, and can be viewed as locally compact counterparts of quantum tori. They are noncommutative deformation of Euclidean spaces which originate from the Heisenberg relation and Moyal products in quantum mechanics. Let θ = (θjk)d j,k=1 be a skew-symmetric d × d matrix. Roughly speaking, a d-dimensional quantum Euclidean space is given by the von Neumann algebra Rθ generated by the spectral projections of d self-adjoint operators x1,· · · , xd satisfying the the canonical commutation relation (CCR) [xj, xk] = −iθjk . We will review a rigorous definition of Rθ in Section 2. Despite having a relatively sim- ple algebraic structure (a type I von Neumann algebra) the connection to Euclidean spaces and quantum physics make them indispensable in various scenarios. For example, from the perspective of harmonic and functional analysis, Calderón-Zygmund theory and pseudodif- ferential operator theory on quantum Euclidean spaces was established in the recent article [GJP17] and the theory of distributions goes back to [GBV88, VGB88]. In noncommutative geometry, quantum Euclidean spaces serve as model examples for non-unital spectral triples ∗ Partially supported by NSF grant DMS-1501103 and DMS-1800872. 1 2 LI GAO, MARIUS JUNGE, AND EDWARD MCDONALD [GGBI+04]. In mathematical physics, noncommutative Euclidean spaces have been heavily studied under the name of canonical commutation relation (CCR) algebras [BR97, Section 5.2.2.2] and in the context of Weyl quantization [Hal13, Chapter 14], [Tak08, Chapter 2, Section 3]. Also, the discovery of instantons on noncommutative R4 makes an influential connection to string theory [CL01, NS98, SW99]. In this paper, we revisit the connection between ΨDOs and the local index formula for quantum Euclidean spaces. Both topics have been considered for Rθ, with its standard geometric structure. Recall that Rθ is associated with a Weyl quantization map, defined for functions in the Schwartz class S(Rd) as: λθ : f ∈ S(Rd) 7→ f (ξ)λθ(ξ)dξ ∈ Rθ . 1 2πdZRd where λθ(ξ) = eξ1x1+···+ξdxd, ξ ∈ Rd is a projective unitary representation of Rd, λθ(ξ)λθ(η) = ei θ 2 ξηλθ(ξ + η) (see Section 2 for further details). The canonical trace associated to Rθ is defined on the image of S(Rd) under λθ as τθ(λθ(f )) =Z f . Differentiation operators ∂ ∂xj extension to Rθ, defined on λθ(S(Rd)) by Djλθ(f ) = λθ(−i ∂ f ). The operators Dj have self-adjoint extensions to the Hilbert-Schmidt space L2(Rθ, τθ). Since partial differentiation operators on S(Rd) commute, it follows immediately that [Dj, Dk] = 0 for 1 ≤ j, k ≤ d. The fact that these partial derivatives mutually commute reflects a "flat" geometry of Rθ. The scope of this paper is to consider a more general but still computable differential structure on Rθ. More precisely, we shall equip Rθ with "covariant derivatives" ξ1,· · · , ξd satisfying (another) CCR relation. Unlike the standard case ∂xj admit a canonical (1.1) [xj, xk] = −iθj,k, [Dj, xk] = −iδj,k , [Dj, Dk] = 0 , we consider that xj's and ξk's together have the commutation relations jk . [xj, xk] = −iθj,k, [ξj, xk] = −iδjk , [ξj, ξk] = −iθ′ (1.2) where δ is the Kronecker Delta notation and θ′ is an arbitrary but fixed skew-symmetric matrix. In the classical case when θ = 0 and R0 = L∞(Rd), such ξj's are covariant derivatives of connections with a constant curvature form (see Section 3.1). From this perspective, (1.2) can be viewed as a natural deformation of (1.1) by adding a nonzero curvature form. From the perspective of quantum physics, noncommuting derivatives occur in the presence of a magnetic field [AHS78]. One can view the matrix θ′ as representing a constant magnetic field on Rθ. The noncommutativity of the covariant derivatives ξj adds essential difficulty in developing the theory of ΨDOs. When θ′ = 0, the commutativity of Dj's makes the phase space (or the Fourier transform side) a commutative space, and then the symbol of a ΨDO is a operator-valued function a : Rd → Rθ. In our setting for noncommuting ξj's, the symbol will become purely abstract as operators affiliated to Rθ ⊗ R′ θ. Moreover, due to the unbounded natural of symbol functions, we have to inevitably deal with unbounded but smooth elements. The idea of incorporating noncommuting derivatives into pseudodifferential calculus has also appeared in the related context of magnetic pseudodifferential calculus [MP04, MPR05]. We now briefly explain our setting and illustrate the main results. Let Rθ⊗R′ 2d-dimensional quantum Euclidean space generated by the relations θ be the [xj, xk] = −iθj,k , [ξj, ξk] = −iθ′ j,k , [xj, ξk] = 0 QUANTUM EUCLIDEAN SPACES WITH NONCOMMUTATIVE DERIVATIVES 3 (cid:20) θ −Id Id and let RΘ be the 2d-dimensional space generated by (1.2) with parameter matrix Θ = θ′ (cid:21). We will consider pseudodifferential calclulus defined with symbols as operators affiliated to Rθ⊗R′ or quantization map "Op" sending symbols to ΨDOs is simple: for a ∈ Rθ, b ∈ Rθ′ θ and the ΨDOs themselves are operators affiliated to RΘ. The operator Op(a ⊗ b) = ab ∈ RΘ , (1.3) where Rθ, R′ following abstract symbol class. θ are viewed as subalgebras of RΘ. The domain of Op can extended to the • We say an operator a affiliated to Rθ⊗Rθ′ is a symbol of order m (write as a ∈ Σm) extends to a bounded j )− m+β 2 if for any multi-indices α and β, Dα operator in Rθ⊗Rθ′. x Dβ ξ (a)(1 +Pj ξ2 Here Dx are the canonical (commuting) differentiation operators acting on the first compo- nent Rθ and Dξ are the same for Rθ′. A priori it is not clear that this definition is closed under multiplication, and adjoint, or if we have the expected properties Σm · Σn = Σm+n and (Σm)∗ = Σm, which are important components for the development of a symbol calculus. To resolve that, we introduce in Section 3 a notation of "asymptotic degree" to measure the unboundedness of operators affiliated to Rθ. This is a notion directly inspired by the abstract pseudodifferential calculus developed by Connes and Moscovici [CM95, Appendix B] and Higson [Hig03]. With this definition of symbol class, we establish in Section 4 the two core parts of ΨDOs calculus -- the L2-boundedness theorem for 0-order ΨDOs and the composition formula. Theorem 1.1 (c.f. Theorem 4.12). Let a be a symbol of order 0 (i.e., a ∈ Σ0). Then Op(a), initially defined on λΘ(S(R2d)) has unique extension to a bounded operator on the Hilbert space L2(RΘ). Theorem 1.2 (c.f. Theorem 4.14). Let a be a symbol of order m and b be a symbol of order n. Then Op(a)Op(b) = Op(c) for some symbol c of order m + n. Moreover c ∼Xα i−α α! Dα ξ (a)Dα x (b) in the sense that for any positive integer N, c−Pα≤N m + n − N − 1. i−α α! Dα ξ (a)Dα x (b) is a symbol of order The proofs of the above theorems use the idea of co-multiplication maps. The co- multiplication maps enables us to convert the operator map Op as an operator-valued classi- cal operator map on the Rd. In particular, this gives an alternative approach to some parts of symbol calculus in [GJP17] for θ′ = 0. In Section 5, we apply the ΨDO calculus prove that (W ∞,1(Rθ), L2(RΘ) ⊗ CN , D =Xj ξj ⊗ cj) , (1.4) forms a semifinite non-unital spectral triple (in the sense of [CGRS14, Definition 2.1]). Here, cj are generators of the Clifford algebra Cld and W ∞,1(Rθ) = {aDα(a) ∈ L1(Rθ) ∀ α} is the noncommutative Sobolev spaces. We denote W ∞,1(Rθ)∼ = W ∞,1(Rθ) + C for the minimal unitalization. The triple (1.4) forms a smoothly summable semifinite spectral triple with 4 LI GAO, MARIUS JUNGE, AND EDWARD MCDONALD isolated spectrum dimension (see Section 5 for further details). We are able to apply the even case of the local index formula [CGRS14, Theorem 3.33], yielding the following: Theorem 1.3 (c.f. Corollary 5.9). Let d be even and Rθ be a d-dimensional quantum Eu- clidean space. Then (A, H, D) := (W ∞,1(Rθ), L2(RΘ)⊗ MN ,Pj ξj ⊗ cj) is an even, smoothly summable, semi-finite spectral triple with isolated spectrum dimension. Moreover, for a pro- jection e ∈ Mn(W ∞,1(Rθ)∼), the index pairing is given by d 2 (τθ ⊗ tr(γ(e − 1e) d 2 ω d 2 ! ) + d 2Xm=1 1 2m! τθ ⊗ tr(γe(de)2m ω d 2 −m ( d 2 − m)! )) , h[e] − [1e], (A, H, D)i = π where ω = i 2Pj,k θj,kcjck. Note that the Dirac Laplacian has square given by D2 = (Xj ξj ⊗ cj)2 =Xj ξ2 j − ω . Where ω plays the role of a curvature form in the index pairing. The general local index formula in [CM95, CGRS14] contains residue cocycles which involve higher order residues at z = 0 for zeta functions ζk(z) = tr(γa0da(k1) 1 m (1 + D2)− m 2 −k−z) · · · da(km) {z k-times , da]]. Theorem 1.3 basically observes where aj ∈ A, da = [D, a] and da(k) := [D2, [D2,· · · [D2 } that the above zeta functions has nonzero residue only for k = 0 and the poles are simple. For a Dirac operator on compact spin Riemannian manifolds, such a simplification was observed in [CM95] and fully developed by Ponge [Pon03] using Getzler calculus. The local index formula of Connes and Moscovici [CM95] recovers the Atiyah-Singer index theorem for spin Dirac operators. Theorem 1.3 shows that a similar simplified index formula holds for the noncommutative spectral triple (W ∞,1(Rθ), L2(RΘ) ⊗ MN ,Pj ξj ⊗ cj). We also provide a concrete example of the index pairing in d = 2 (Theorem 5.11). The paper is organized as follows: We first reviews some preliminary facts about quantum Euclidean spaces in Section 2. Section 3 introduces and discuss the notation "asymptotic degree", which is a key tool in the subsequent discussions. In Section 4, we discuss the symbol calculus of ΨDOs and prove Theorem 1.1 and 1.2. Section 5 is devoted to the local index formula and Theorem 1.4. Acknowledgement-The authors are grateful to Alexander Gorokhovsky for helpful discussion on the local index formula. 2. Preliminaries on Quantum Euclidean spaces In this section we review the basic structures of Quantum Euclidean spaces. Quan- tum Euclidean spaces in the literature has been studied under several different names: Moyal plane [GGBI+04, GBV88, VGB88], canonical commutatation relation (CCR) alge- bras [BR12, Section 5.2.2.2], noncommutative Euclidean Spaces [Gao18, SMZ18] and quan- tum Euclidean spaces [GJP17]. In particular, [BR12] gives a detail account from the oper- ator theoretic perspective. The distribution theory was studied in [GBV88, VGB88]. More recently [GJP17] studies harmonic analysis on quantum Euclidean spaces. From the non- commutative geometric perspective, an early exposition is in [GGBI+04]. QUANTUM EUCLIDEAN SPACES WITH NONCOMMUTATIVE DERIVATIVES 5 2.1. Definitions and notations. Throughout the paper we use the usual letters x1, x2,· · · , and ξ1, ξ2,· · · for operators and the boldface letters x = (x1, x2,· · · , xd), ξ = (ξ1, ξ2,· · · , ξd) for vectors and scalars. Let d ≥ 2 and θ = (θjk)d j,k=1 be a real skew-symmetric d × d matrix. Let S(Rd) the space of complex Schwartz functions (smooth, rapidly decreasing) on Rd. The Moyal product ⋆θ associated to θ is defined as (see [Rie93]), f ⋆θ g(x) := (2π)−dZRdZRd f (x + θ 2 v)g(x − w)eiv·wdvdw , f, g ∈ S(Rd) The Moyal product is bilinear, associative and reversed under complex conjugation f ⋆θ g = g ⋆θ f , which makes (S(Rd), ⋆θ) a ∗-algebra. The left Moyal multiplication gives the following ∗-homomorphism λθ : (S(Rd), ⋆θ) → B(L2(Rd)), λθ(f )g = f ⋆θ g, λθ(f )λθ(g) = λθ(f ⋆θ g) . (2.1) Definition 2.1. The quantum Euclidean space associated to θ is given by the following objects in B(L2(Rd)), i) Sθ := λθ(S(Rd)) as the quantized Schwartz class ; ii) Eθ := S · iii) Rθ := (Sθ)′′ as the von Neumann algebra generated by Sθ. as the C ∗-algebra generated by Sθ; θ When θ = 0, ⋆0 is the usual point-wise multiplication, E0 = C0(Rd) is the space of continuous functions on Rd which vanish at infinity and R0 = L∞(Rd) is the space of essentially bounded functions on Rd. An equivalent approach is the θ-twisted regular representation of the group Rd. For each vector ξ ∈ Rd, we define the unitary operator λθ(ξ) on L2(Rd), They satisfies the commutation relation (λθ(ξ)g)(x) = eiξ·xg(x − θ 2 ξ) (2.2) λθ(ξ)λθ(η) = e i 2 ξ·θηλθ(ξ + η) = eiξ·θηλθ(η)λθ(ξ) . The map λθ : Rd → B(L2(Rd) is a projective unitary representation of Rd called the twisted left regular representation. The Moyal multiplication (2.1) for (S(Rd), ⋆θ) is equivalent to the corresponding Weyl quantization 1 λθ(f ) = (2π)dZRd f (ξ)λθ(ξ)dξ , f ∈ S(Rd). Here f (ξ) = RRd f (x)e−ix·ξdx is the Fourier transform of f and the integral converges in strong operator topology. Let uj(t) = λθ(0, 0,· · · , t,· · · , 0) be the one parameter unitary group associated to the j-th coordinate. The generator xj of uj(t) satisfying uj(t) = eixj t is given by. (xjg)(x) = xjg(x) + (x) . i 2Xk θjk ∂g ∂xk (x1,· · · , xd) are d self-adjoint operators on L2(Rd) affiliated to Rθ which satisfies the CCR relation [xj, xk] = −iθjk. The projective unitary representation ξ → λθ(ξ) can be recovered from (x1,· · · , xd) using Baker -- Campbell -- Hausdorff formula i.e. λθ(ξ) := ei(ξ1x1+···+ξdxd) = e− i 2 Pj<k θjkξj ξkeiξ1x1 · · · eiξdxd , ξ ∈ Rd The generator (x1,·, xd), unitary λθ(ξ) and the quantized Schwartz class λθ(f ) are equivalent formulations of quantum Euclidean spaces. We will use them interchangeably in the paper. 6 LI GAO, MARIUS JUNGE, AND EDWARD MCDONALD 2.2. The Stone-von Neumann Theorem. We say two self-adjoint operator P, Q satisfies the Heisenberg relation [P, Q] = −iI if for any s, t ∈ R, eisP eitQ = eisteitQeisP The well-known Stone-von Neumann Theorem states that any irreducible representations of [P, Q] = −iI is unitarily equivalent to the 1-dimensional Schrodinger picture that P f = −i df dx , (Qf )(x) = xf (x) , f ∈ S(R) . Here P, Q are unbounded self-adjoint operators on L2(R) and the one-parameter unitary groups are (eitP f )(x) = f (x + t) , (eisQf )(x) = eisxf (x) , (2.3) The Stone-von Neumann Theorem extends to n pairs of Heisenberg relations that mutually commute, i.e. [Pj, Qk] =(−iI, 0, if j = k if j 6= k. , [Pj, Pk] = [Qj, Qk] = 0 , ∀ j, k (2.4) The following is the Theorem 14.8 of [Hal13]. Theorem 2.2 (Stone -- von Neumann Theorem). Suppose P1,· · · , Pd and Q1,· · · , Qd are self- adjoint operators on H satisfying the CCR relations (2.4). Then H can be decomposed as an orthogonal direct sum of closed subspaces {Hj} satisfying i) each Hl is invariant under eitPj and eitQj for all j and t. ii) there exist unitary operators Ul : Hl → L2(Rd) such that UlPjU ∗ l f = −i ∂ ∂xj f , (UlQjU ∗ l f )(x) = xjf (x) . (2.5) The above theorem says that any representation of (2.4) is a finite or infinite multiple of the n-dimensional Schrodinger picture on L2(Rn). When d = 2n is even dimensional, this gives the standard noncommutative case for Rθ that θ =(cid:20) 0 −In 0 (cid:21), where In is the n-dimensional identity matrix. In this case, Eθ ∼= K(L2(Rn)) the compact operators and Rθ ∼= B(L2(Rn)). The following proposition gives change of variables between Rθ's with different θ. In Proposition 2.3. Let T = (Tjk)d θ and θ be two skew-symmetric matrices such that θ = T θT t. Then the map ΦT : j,k=1 be a real invertible matrix and T t be its transpose. Let ΦT (λθ(ξ)) = λθ(T tξ) , ΦT (λθ(f )) = λθ(f ◦ T ) extends to a ∗-isomorphism from Eθ to Eθ and a normal ∗-isomorphism from Rθ to Rθ. Proof. Define the operator UT on L2(Rd) as follows, (UT f )(x) = f (T −1x) . UT is bounded and invertible with k UT k= det(T ) 1 function f , one verifies that 2 and (UT )−1 = UT −1. For any Schwartz (U −1 T λθ(ξ)UT f )(x) = eiξ·T xf (T −1(T x + θξ)) = ei(T tξ)·xf (x + 1 2 1 2 θT tξ) = λθ(T tξ)f (x) . QUANTUM EUCLIDEAN SPACES WITH NONCOMMUTATIVE DERIVATIVES 7 Then it is clear that U −1 then ΦT (·) = U −1 from Rθ to Rθ. T SθUT = Sθ. Since UT is a bounded invertible operator on L2(Rd), T (·)UT extends to a ∗-isomorphism from Eθ to Eθ and a normal ∗-isomorphism (cid:3) In general, let θ be a skew-symmetric matrix of rank 2n ≤ d. There exists an invertible matrix T such that θ = T θT t is the following standard form 0 −In 0 In 0d−2n   , (2.6) where 0d−2n is (d − 2n) × (d − 2n) zero matrix. Let x1,· · · , xd be the generators of E (θ). Then x1,· · · , x2n by Stone-von Neumann theorem are unitary equivalent to (a multiple , x1,· · · , xn on L2(Rn), and of) the derivatives and position operators −i x2n+1,· · · , xd are d − 2n the position operators xn+1,· · · , xd−n on L2(Rd−2n). Hence if θ is of rank 2n < d, we have up to multiplicity ,· · · ,−i ∂ ∂xn ∂ ∂x1 Eθ ∼= K(L2(Rn)) ⊗ C0(Rd−2n) , Rθ ∼= B(L2(Rn))⊗L∞(Rd−2n) In particular, the C ∗-algebra Eθ is simple if and only if the matrix θ is of full rank. 2.3. Integrals and Derivatives. We start with the noncommutative integrals. Proposition 2.4. The linear functional τθ(λθ(f )) =ZRd f , f ∈ S(Rd) extends to a normal faithful semi-finite trace on Rθ. i) Let T be a real invertible matrix and θ, θ be two skew-symmetric matrix such that θ = T θT t. Then the normal ∗-isomorphism ΦT : Rθ → Rθ , ΦT (λθ(f )) = λθ(f ◦ T ), (2.7) satisfies τθ ◦ ΦT = detT−1τθ. ii) Let x ∈ Rd and αx be the translation action αx(f )(·) = f (· + x). Define the map αx(λθ(ξ)) = eiξ·xλθ(ξ) , αx(λθ(f )) = λθ(αx(f )) . Then αx is a τθ-preserving automorphism on Rθ. Proof. The fact τθ is a normal faithful trace on Rθ was proved in [GJP17] by writing Rθ as an iterated crossed product L∞(R) ⋊ R ⋊ · · · ⋊ R. Here we present a proof using change of variables, which is useful for our later discussion. A similar discussion can be found in [LSZ17]. Denote the multiplier and translation unitary groups on L2(Rn) as follows, (u(ξ)f )(x) = f (x + ξ) , (v(η)f )(x) = eiη·xf (x) . We first consider the case d = 2n and θ =(cid:20) 0 −In 0 (cid:21). By the Stone-von Neumann theorem, there exists some Hilbert space H and a unitarily W : L2(Rθ) → L2(Rn) ⊗ IH such that In W λθ(ξ, 0)W ∗ = u(ξ) ⊗ IH , W λθ(0, η)W ∗ = v(η) ⊗ IH , 8 LI GAO, MARIUS JUNGE, AND EDWARD MCDONALD where ξ ∈ Rn are the first n coordinates and η ∈ Rn are the last n coordinates. For f1, f2 ∈ S(Rn), the quantization λθ(f1 ⊗ f2) is unitarily equivalent to (a multiple of) the following operator Tf1,f2. For h ∈ L2(Rn) (Tf1,f2h)(y) = (2π)−2nZ Z f1(ξ) f2(η)e− i = (2π)−2nZ Z f1(x − y) f2(η)e− i = (2π)−nZ f1(x − y)f2( x + y 2 )h(x)dx . 2 ξ·ηeiη·(y+ξ)h(y + ξ)dξdη 2 (x−y)·ηeix·ηh(x)dxdη Bacause f1, f2 ∈ S(Rn), it follows from [Bri88, Proposition 1.1 and Theorem 3.1] that Tf1,f2 is a trace class operator on L2(Rn) and which coincides with τθ on Rθ up to a normalization constant (2π)−n. Now we consider θ tr(Tf1,f2) =(2π)−nZRn =(2π)−nZRn is a singular standard form θ = y + y 2 )dy f1(y − y)f2( f1 ·ZRn f1(0)f2(y)dy = (2π)−nZRn 0 . Let θ1 =(cid:20) 0 −In 0 (cid:21) be the nonsingular f2 , 0 −In 0 In 0 0 0 0 In part. Rθ1 ∼= B(L2(Rn)) is a Type I factor and the degenerated part gives the left regular representation λ0 : Rd−2n → B(L2(Rd−2n)). Then, Rθ ∼= Rθ1⊗R0 ∼= B(L2(Rn))⊗L∞(Rd−2n) as von Neumann algebras. The trace τθ on Rθ is the product trace τθ1 ⊗ τ0, where τ0 on L∞(Rd−2n) is the Lebesgue integral and τθ1 is up to a constant the standard trace tr on B(L2(Rn)). Then τθ is normal faithful semifinite and the case for general θ follows from i). Recall that the ∗-isomorphism ΦT is implemented by the bounded invertible operator UT : L2(Rθ) → L2(Rθ) , UT λθ(f ) = λθ(f ◦ T −1) . For f ∈ S(Rd), τθ ◦ ΦT (λθ(f )) =τθ(cid:16)ZRd f (ξ)λθ(T ξ)dξ(cid:17) = det T−1τθ(cid:16)ZRd = det T−1 f (0) = det T−1τθ(λθ(f )) . f(cid:0)T −1η(cid:1)λθ(η)dη(cid:17) For ii), αx is implemented by the shifting unitary Ux on L2(Rd)) that x , Uxf (y) = f (y + x) . αx(λθ(f )) = Uxλθ(f )U ∗ Hence αx extends to an automorphism on Rθ. (cid:3) The automorphisms αx, x ∈ Rd is called the transference action on Rθ. For 1 ≤ p ≤ ∞, we write Lp(Rθ) for the noncommutative Lp space with respect to τθ and identify L∞(Rθ) = Rθ. For all θ, L2(Rθ) ∼= L2(Rd) and λθ is exactly the left regular representation of Rθ on L2(Rθ). It is clear that Sθ is dense in Eθ and L2(Rθ). Lemma 2.5. Sθ is dense in L1(Rθ). QUANTUM EUCLIDEAN SPACES WITH NONCOMMUTATIVE DERIVATIVES Proof. If a ∈ L1(Rθ), then a = a1a2 for some a1, a2 ∈ L2(Rθ) and k a1 k2=k a2 k2=k a k Then we can find f1, f2 ∈ S(Rd) such that k λθ(fj) − aj k2≤ ǫ, j = 1, 2. Then k a − λθ(f1)λθ(f2)k ≤k a1a2 − a1λθ(f2)k1 + k a1λθ(f2) − λθ(f1)λθ(f2)k1 ≤k a1k2 ǫ+ k f2k2 ǫ ≤ (2 k ak 1 2 1 +ǫ)ǫ . 9 1 2 1 . (cid:3) The noncommutative Lorentz space Lp,∞(Rθ) is the space of measurable operators a affiliated to Rθ such that the following quasi-norm is finite tpτθ(1a>t) , k akp Lp,∞= sup t>0 where 1a>t denote the spectral projection of a. In other words, a ∈ Lp,∞(Rθ) if τθ(1a>t) is asymptotically at most O(t−p). For det(θ) 6= 0, the above (weak) Lp spaces are nothing but the (weak) Schatten p-spaces. Proposition 2.6. Denote x := (Pj x2 i) hxi−1 ∈ Ld,∞(Rθ). ii) τθ(e−tx2 2 det( ) = t− d πitθ j ) )1/2 for t > 0. 1 2 and hxi := (1 +Pj x2 j ) 1 2 . For all θ, sinh(itθ) πµ Here the function µ 7→ is the function calculus for self-adjoint matrix iθ. sinh µ is a real function continuously extended to µ = 0 and πiθ sinh(iθ) Proof. Let us first consider that θ is the standard form (2.6) of rank 2n. We have shown in Proposition 2.4 that there is (up to a factor (2π)n) a trace preserving ∗-isomorphism π : Rθ → B(L2(Rn))⊗L∞(Rd−2n) on L2(Rd−n) such that for 1 ≤ j ≤ n, 1 ≤ k ≤ d − 2n where Dyj and yj are the self-adjoint derivative and position operators on L2(Rd−n) xj 7→ Dyj , xj+n 7→ yj , x2n+k 7→ yn+k . Dyj g = −i ∂g ∂yj , (yjg)(y) = yjg(y) . Then hxi2 is unitary equivalent to (a multiple) of the following operator on L2(Rd−n), Thus τθ(1H −1/2>µ) . µ−d which implies H −1/2 ∈ Ld,∞. The case for general θ follows from the change of variable in Proposition 2.4. Moreover, if T is a real invertible matrix such H := ( D2 yj + y2 j ) ⊗ idL2(Rd−2n) + idL2(Rn) ⊗ (1 + nXj=1 y2 l ) . d−nXl=n+1 j=1 D2 yj + y2 (Pn The first part is the Hamiltonian of n-dimemsional quantum harmonic oscillator and the second part is a multiplier on L2(Rd−2n). It is known (see [Hal13, Chapter 11]) that H1 := Combined with the continuous part on L∞(Rd−2n), we have j ) has discrete spectrum µN = 2N + n and the degeneracy of µN is(cid:0)N +n−1 N (cid:1). τθ(1H≤µ) = (2π)n X2N ≤µ−n(cid:18)N + n − 1 (cid:19)ZRd−2n 1(1+y2)≤µ−2N −ndy N . µ · µn−1 · µ d−2n d 2 = µ 2 . 10 LI GAO, MARIUS JUNGE, AND EDWARD MCDONALD that T θT t is the standard form (2.6), then det(T ) = (µ1µ2 · · · µn)−1, where µ1, µ2,· · · , µn are imaginary parts of eigenvalues of θ. Thus, by the isomorphism in (2.7), we have τθ(e−tx2 e−t Pd−n j=n+1 y2 j dyn+1 · · · dyd−n j )) ·ZRd−2n d−2n 2 ) j=1 µj (D2 yj +y2 e−tµj (1+2k)(cid:1) · ( π t ) = µ1µ2 · · · µn(2π)n · tr(e−t Pn = µ1µ2 · · · µn(2π)n ·(cid:0) nYj=1Xk=0 =(cid:0) nYj=1 etµj − e−tµj(cid:1)(π) 2(cid:0) nYj=1 sinh tµj(cid:1)(π) = t− d 2πtµj πtµj d−2n d−2n 2 2 t− d 2 = t− d 2 det( πitθ sinh(itθ) )1/2 . The last equality follows from lim µ→0 πµ sinh(µ) = π. (cid:3) Let Dx1,· · · , Dxd be the partial derivatives operator Dxj f = −i f , which are un- bounded self-adjoint operators on L2(Rd) with a common core S(Rd). On Rθ, we define for λθ(f ) in Sθ ⊂ B(L2(Rd)) the partial derivatives ∂ ∂xj Djλθ(f ) := [Dxj , λθ(f )] = λθ(Dxj f ). Here ej = (0,· · · , 1,· · · , 0) is the j-th standard basis of Rd. Since Dxj is the same as Dj for θ = 0, we will often write Dxj simply as Dj. Let S ′(Rd) be the space of tempered distribution on Rd. In [GBV88, VGB88] (see also [GGBI+04]), Moyal product and the Weyl quantization are weakly extended to S ′(Rd) as follows, hT ⋆θ f, gi = hT, f ⋆θ gi ,hf ⋆θ T, gi = hT, g ⋆θ fi . where the bracket is the pairing between S(Rd) and S ′(Rd). For T ∈ S′(Rd), λθ(T ) is the quantized operator λθ(T )f = T ⋆θ f and satisfies λθ(T )λθ(f ) = λθ(T ⋆θ f ), λθ(f )λθ(T ) = λθ(f ⋆θ T ) . For all T ∈ S ′(Rd), λθ(T ) commutes with the right Moyal multiplication hence affiliates to Rθ. We will use the multiplier algebra introduced in [VGB88], Mθ = {λθ(T ) T ∈ S ′(Rd), λθ(T )Sθ ⊂ Sθ,Sθλθ(T ) ⊂ Sθ} . The pairing between S(Rd) and S ′(Rd) coincides with the τθ-trace duality for the quantiza- tion. Namely for λθ(T ) ∈ Mθ, λθ(f ) ∈ Sθ, In particular, Mθ contains the noncommutative polynomials of x1,· · · , xd as the quantized coordinate function xj, τθ(λθ(T )λθ(f )) = τθ(λθ(T ⋆θ f )) =R T ⋆θ f = hT, fi 2Pk θjkDkλθ(f ) . λθ(xj) = xj , xjλθ(f ) = λθ(xjf ) + 1 QUANTUM EUCLIDEAN SPACES WITH NONCOMMUTATIVE DERIVATIVES 11 The transference automorphism αx and the partial derivatives Dj weakly extend to Mθ hαx(a), λθ(f )i := ha, α−xλθ(f )i ,hDj(a), λθ(f )i = ha, Djλθ(f )i . Viewing a ∈ Mθ as an unbounded operator densely defined on S(Rd) ⊂ L2(Rd), the weak derivatives satisfies Dj(a) = [Dj, a]. 3. Asymptotic degrees In this section, we introduce a notation of "asymptotic degrees" to measure the "growth" of unbounded elements in Rθ, which serves as a key technical tool for later discussions. The idea is inspired from the abstract ΨDOs introduced by Connes and Moscovici in [CM90, CM95]. We briefly recall the basic setting here. Let D be a (possibly unbounded) self- adjoint operator on a Hilbert space H such that D is strictly positive. For each s ∈ R, put H s = Dom(Ds) with inner product hv1, v2iH s := hDsv1,Dsv2iH , v1, v2 ∈ Dom(Ds) Let H ∞ = ∩s∈ZH s. Because Dom(eD2) ⊂ H ∞, H ∞ is a dense subspace of H. Let F be a closed operator on H such that H ∞ ⊂ Dom(F ), F (H ∞) ⊂ H ∞. Because D−s : H 0 → H s is an isometric isomorphism, one sees that k F : H s → H s−rk=kDs−rFD−sk For a fixed r ∈ R, F extends to a bounded operator from H s to H s−r for any s if and only if Ds−rFD−s are bounded on H. Such F is considered as an abstract ΨDO of order r. We use the above idea to characterize the asymptotic degree (we use the word "degree" to distinguish with the notation "order" for ΨDOs) of elements in Mθ. We choose the strictly 1 positive operator D as hxi := (1 +Pj x2 j ) 2 . Definition 3.1. We say an operator a ∈ Mθ is of asymptotic degree r if for any s ∈ R, hxisahxi−s−r extends to a bounded operator in B(L2(Rθ)) (hence also in Rθ ⊂ B(L2(Rθ))). We denote Or the set of all elements of asymptotic degree r and write O−∞ = ∩r∈ZOr. Let Ls 2(Rθ) be the Hilbert space completion of Sθ with respect to the inner product hλθ(f ), λθ(g)is = τθ(λθ(f )∗hxi2sλθ(g)) . It is clear that a ∈ Or if and only if for any s ∈ R, the left multiplication operator λθ(f ) 7→ aλθ(f ) extends continuously from Ls (Rθ). The following theorem estimates the degrees of some common elements. We introduce the standard notation of multi-indices that for α = (α1, α2,· · · , αd), 2(Rθ) to Ls−r 2 xα := xα1 1 xα2 2 · · · xαd d , Dα := Dα1 1 Dα2 2 · · · Dαd d . Note that the product xα is ordered because xj's are noncommutative. Theorem 3.2. For all multi-indices α and r ∈ R, xα ∈ Oα , [xα,hxir] ∈ Or+α−2 , Dα(hxir) ∈ Or−α . 12 LI GAO, MARIUS JUNGE, AND EDWARD MCDONALD Proof. We divide the proof into several steps. Step 1.: [Dj,hxi−r]hxir+1, [xj,hxi−r]hxir+1 are bounded for 0 < r < 2. We use the fractional power for a positive operator A, A−s = CsZ ∞ 0 (t + A)−1t−sdt , 0 < s < 1 , where Cs is a nonzero constant depending on s. Since the constant does not affect the j . For 0 < r < 2, 0 0 2 dt 2 dt [Dj, (t + ∆)−1]t− r (t + ∆)−1xj(t + ∆)−1t− r (t + ∆)−1[(t + ∆), Dj](t + ∆)−1t− r boundedness, we suppress all constant Cs's. Denote ∆ := hxi2 = 1 +Pj x2 [Dj,hxi−r] =Z ∞ =Z ∞ = 2iZ ∞ = 2iZ ∞ = 2iZ ∞ = 2ixjZ ∞ 2ixjZ ∞ 2 dt + 2iZ ∞ 2 dt + 2iZ ∞ 2 dt + 2Xk θjkZ ∞ 2 = 2ixj∆−1− r For the first integral, xj(t + ∆)−2t− r xj(t + ∆)−2t− r (t + ∆)−2t− r (t + ∆)−2t− r 2 dt · ∆ 1+r 2 = 2ixj∆− 1 2 2 dt 2 ∆ 0 0 0 0 0 1+r 0 0 0 is bounded. For the second integral, [(t + ∆)−1, xj](t + ∆)−1t− r 2 dt (t + ∆)−1[xj, (t + ∆)](t + ∆)−2t− r 2 dt (t + ∆)−1xk(t + ∆)−2t− r 2 dt Then [xj,hxi−r]hxir+1 for 0 < r < 2 which is bounded by previous case. In particular, we also obtained hxi−rxjhxir+1 = [hxi−r, xj]hxir+1 + xjhxi−1 is bounded for 0 < r < 2. Step 2. [xj,hxi−r]hxir+1, [Dj,hxi−r]hxir+1 are bounded for all r. First for −2 < r < 0, the bounededness follows from [xj,hxi−r]hxir+1 = [xj,hxi−r−2]hxir+3 + 2iXk θjkhxi−r−2xkhxir+1 . converges absolutely. For the commutator with xj, we have 0 0 kZ ∞ (t + ∆)−1xk(t + ∆)−2t− r 2 dthxi1+rk ≤Z ∞ ≤Z ∞ [xj,hxi−r] =Z (t + ∆)−1[(t + ∆), xj](t + ∆)−1t− r θjkZ (t + ∆)−1xk(t + ∆)−1t− r = 2iXk 2 dt 0 k (t + ∆)−2+ r 2 t− r (t + 1)−2+ r 2 dt 2 k t− r 2 dt < ∞ 2 dt = 2iXk θjk[Dj,hxi−r]. QUANTUM EUCLIDEAN SPACES WITH NONCOMMUTATIVE DERIVATIVES 13 Then we have the initial case for −2 < r < 2 and use the the following induction steps r → −r + 1 for r < 0 and r → −r − 1 for r > 0, [xj,hxir]hxi−r+1 = hxi[xj,hxir−1]hxi−r+1 + [xj,hxi] = hxir[hxi−r+1, xj] + [xj,hxi] [xj,hxir]hxi−r+1 = hxi−1[xj,hxir+1]hxi−r+1 + [xj,hxi−1]hxi2 = hxir[hxi−r−1, xj]hxi2 + [xj,hxi−1]hxi2 = hxir[hxi−r−1, xj] − hxi−1[hxi2, xj] + [xj,hxi−1]hxi2 . The argument for [Dj,hxi−r]hxir+1 is similar. Step 3. xα ∈ Oα and [xα,hxir] ∈ Oα+r−2 for all α and r. First, by Step 2 we have that for all s hxisxjhxi−s−1 = [hxis, xj]hxi−s−1 + xjhxi−1 hxi−s[xj,hxir]hxi−r+s+1 = [xj,hxir−s]hxi−r+s+1 + [xj,hxi−s]hxis+1 hxi−s[Dj,hxir]hxi−r+s+1 = [Dj,hxir−s]hxi−r+s+1 + [Dj,hxi−s]hxis+1 are all bounded. This implies xj ∈ O1 , [xj,hxir] ∈ Or−1 , [Dj,hxir] ∈ Or−1 . Thus xα ∈ Oα by product. For [xα,hxir], we use the induction step that by the Leibniz's rule [xjxα,hxir] = xj[xα,hxir] + [xj,hxir]xα , and [xj, xα] is a polynomial of order less than α. Step 4. Dα(hxir) ∈ Or−α for all r ∈ R. We first do induction on α for −2 < r = −2s < 0. For 0 < s < 1, we introduce the following notation Is(a1, a2,· · · , al) :=Z ∞ 0 t−s(t + ∆)−1a1(t + ∆)−1a2(t + ∆)−1 · · · (t + ∆)−1al(t + ∆)−1dt . For α = 1, [Dj,hxi−2s] = 2iIs(xj). Note that by Leibniz rules [Dj, Iα(a1,· · · , al)] = X1≤k≤l Iα(a1,· · · , [Dj, ak] ,· · · , al) + X1≤k≤l+1 kth {z } {z } kth Iα(a1,· · · , [∆, Dj] , ak,· · · , al) . (3.1) Then all higher order derivatives of hxi−2s are sum of Is(a1, a2,· · · , al) terms with a1,· · · , al ∈ {1, x1,· · · , xn}. Moreover, their degree can be tracked inductively. Let sk be the degree of Now assume that for α ≤ N, Dα(hxir) is a sum of the terms Is(a1, a2,· · · , al) with −2l − the first part in (3.1) is lowered by 1 because [Dj, xj] = −i and [Dj, 1] = 0, and the second part has the degrees at most ak. We show in the next lemma that Is(a1,· · · , al) is at most of degree −2l − 2s +Pk sk. 2s +Pk sk ≤ r − α. Then [Dj, Dα(hxir)] is a sum of commutators as (3.1). The degree of −2(l + 1) − 2s + (1 +Xk sk) = −2l − 2s − 1 +Xk sk 14 LI GAO, MARIUS JUNGE, AND EDWARD MCDONALD because [∆, Dj] = 2ixj and the length l is increased by 1. Thus by induction on α we prove the case −2 < r < 0. For general r, one can always write r = r1 + r2 + · · · + rl as a finite sum of rk ∈ (−2, 0] ∪ 2N. Then by Leibniz rule Dα(hxir) = Xα1+···+αl=α(cid:18) α α1,· · · , αn(cid:19)Dα1(hxir1)· · · Dαl(hxirl) , α1,··· ,αn(cid:1) = α!(α1!)−1 · · · (αd!)−1 is the multi-nomial coefficient. For positive integer m, Dα(x2m) is a polynomial of degree 2m − α and the term Dα(hxirk),−2 < rk < 0 has degree at most rk − α as proved above. Therefore, Dα(hxir) is of degree at most (cid:3) α where (cid:0) Pk rk − αk = r − α. The following lemma is inspired from the abstract ΨDO calculus in [Hig03]. Lemma 3.3. Let 0 < s < 1 and let Is be the notation t−s(t + ∆)−1a1(t + ∆)−1a2(t + ∆)−1 · · · (t + ∆)−1al(t + ∆)−1dt . Is(a1, a2,· · · , al) :=Z ∞ 0 Then i) if ak ∈ Osk, Is(a1, a2,· · · , al) ∈ O−2l−2s+Pk sk+ǫ for any ǫ > 0 ii) if ak ∈ {1, x1, x2,· · · , xn}, Iα(a1, a2,· · · , al) ∈ O−2l−2s+Pk sk. Proof. Let q, r ∈ R with −q + r = −2l − 2s +Pk sk + ǫ. hxiqZ ∞ =Z ∞ 0 0 t−s(t + ∆)−1a1(t + ∆)−1a2(t + ∆)−1 · · · (t + ∆)−1al(t + ∆)−1dthxi−r t−s(t + ∆)−1+α−ǫ/2hxiq(t + ∆)−s+ǫ/2a1(t + ∆)−1 · · · (t + ∆)−1al(t + ∆)−1hxi−rdt Note that khxiq(t + ∆)−s+ǫ/2a1(t + ∆)−1a2(t + ∆)−1 · · · (t + ∆)−1an(t + ∆)−1hxi−rk ≤ khxi2q−ǫ(t + ∆)−q+ǫ/2kkhxiq−2s+ǫa1hxi−q+2s−ǫ−s1 kkhxi2(t + ∆)−1k · · · khxi2(t + ∆)−1kkhxiq+Pk≤l−1 sk−2(n−1)−2s+ǫalhxi−q−Pk≤l sk+2s+2(n−1)−ǫkkhxi2(t + ∆)−1k ≤ khxiq−2s+ǫa1hxi−q+2s−ǫ−s1 k · · · khxiq+Pk≤l−1 sk−2(l−1)−2s+ǫalhxi−q−Pk≤l sk+2s+2(l−1)−ǫk which is uniformly bounded. Thus t−s(t + ∆)−1a1(t + ∆)−1a2(t + ∆)−1 · · · (t + ∆)−1an(t + ∆)−1dthxi−rk khxiqZ ∞ .Z ∞ 0 0 For ii), note that k t−q(t + ∆)−1+s−ǫ/2k dt ≤Z ∞ ) =Z ∞ Is(1,· · · , 1 0 0 l {z } t−s(t + 1)−1+s−ǫ/2dt < ∞ . (t + ∆)−lt−sdt = Cshxi−2(l−1)−2s QUANTUM EUCLIDEAN SPACES WITH NONCOMMUTATIVE DERIVATIVES 15 Let k be the last position in Is(a1,· · · , al) such that ak is not scalar. That is, for all n ≤ k, an = xjn for some 1 ≤ jn ≤ d and am = 1 for all k < m ≤ l. =Is(a1,· · · , ak−1, 1, xj, 1,· · · , 1 ) + Is(a1,· · · , ak−1, 1, [∆, xj], 1,· · · , 1 ) Is(a1,· · · , ak−1, xj, 1,· · · , 1 ) l {z {z l l {z } )xj + Xk+1≤m≤l+1 } } l+1 {z } {z }m th =Is(a1,· · · , ak−1, 1,· · · , 1 Is(a1,· · · , ak−1, 1,· · · , [∆, xj] ,· · · , 1) Note that [∆, xj] = −2iPk θkjxk. Then by i), the second part belongs to O−2l−2+Pk sk−2s+ǫ ⊆ O−2l+Pk sk−2s. We then finish the proof by the induction on the last non-scalar position. (cid:3) Proposition 3.4. i) Let s ∈ R. If Dα(a)hxi−s is bounded for all α, then a ∈ Os. ii) Sθ = {a ∈ Rθ Dα(a) ∈ O−∞ for all α}. Moreover, the map f 7→ λθ(f ) is bi-continuous from S(Rd) equipped with the standard semi-norms to Sθ with the semi-norms k Dα(·)hxi2nk for all α and n. In particular, hxirSθ ⊂ Sθ for any r. Proof. i) Define the notation θjl(xjDl(a) + Dl(a)xj); θjlθmj(xmDl(a) + Dl(a)xm) a(1) := [∆, a] = iXl a(2) := [∆, [∆, a]] = −2Xl Xm −Xl,m θjlθkm(xjxkDlDm(a) + xjDlDm(a)xk + xkDlDm(a)xj + DlDm(a)xkxj) We first give the proof for s = 0. Assume that Dα(a) is bounded for all α. Then a(1)hxi−1 is bounded because and similarly one can verify that a(2)hxi−2 is bounded. Then for 0 < r < 2, xjDl(a)hxi−1 = Dl(a)xjhxi−1 + [xj, Dl(a)]hxi−1 θjkDkDl(a)hxi−1 . = Dl(a)xjhxi−1 −Xk (a(1))hxir (a(2), 1)hxir = a(1)hxi−1 + I r ([∆, a])hxir = I r (1)hxir + I r 2 2 2 2 [a,hxi−r]hxir = I r 2 = a(1)I r (a(2), 1)hxir . Thus we have hxi−rahxir is bounded for 0 ≤ r ≤ 2, and for −2 ≤ r ≤ 0 by taking the adjoint. Moreover, the same argument applies to Dβ(a) for all β. Consider b = hxi−rahxir. The second part is bounded because k I r 2 (a(2), 1)hxirk ≤Z ∞ .Z ∞ 0 0 t− r t− r Dα(b) = Xα1+α2+α3=α(cid:18) t− r 2 k (∆ + t)−1kk a(2)(t + ∆)−1kkhxir(t + ∆)−1k dt 2 khxir(t + ∆)−2k dt ≤Z ∞ 2 dt < ∞ α1, α2, α3(cid:19)Dα1(hxi−r)Dα2(a)Dα3(hxir) . 2 (t + 1)−2+ r α 0 16 LI GAO, MARIUS JUNGE, AND EDWARD MCDONALD is bounded for all α by Leibniz rule and Theorem 3.2. Thus we have shown that hxi−rahxir bounded for −4 ≤ r ≤ 4. By induction this can be extended for all r ∈ R which proves the case s = 0. For general s, we have Dα(ahxi−s) = Xα1+α2=α(cid:18) α α1, α2(cid:19)Dα1(a)Dα2(hxi−s) , which the assumption Dα(a)hxi−s is bounded and Dα2(hxi−s) ∈ Os−α by Theorem 3.2. Thus by the case of s = 0, we know ahxi−s ∈ O0 which implies a ∈ Os. For ii), we first show that for f ∈ S(Rd), λθ(f )hxi2m is bounded for all positive integers m. Note that hxi2m is a polynomial of x with degree 2m. And xjλθ(f ) = λθ(xjf + θjk∂jf ) , i 2Xk λθ(f )xj = (xjλθ( ¯f ))∗ = (λθ(xj ¯f + θjk∂jf ))∗ = λθ(xjf ) − θjkλθ(∂jf ) i 2Xk i 2Xk Then λθ(f )hxi2m are again in Sθ hence bounded. Therefore for any r > 0, λθ(f )hxir is bounded and similarly for the derivatives Dα(λθ(f )). Thus by i), Dα(λθ(f )) ∈ O−∞ for all α. For the other direction, a ∈ Or for r < − d 2 implies k ak2≤khxirk2khxi−rak∞< ∞ . Thus a = λθ(f ) for some f ∈ L2(Rd) and Dα(a) = λθ(Dα(f )) in the distribution sense. Then all the derivatives of f belongs to L2(Rd) and hence f is in the Sobolev space H s(Rd) = {f (1 + ∆)sf ∈ L2(Rd)} for all s. Using Sobolev embedding theorem, f ∈ C ∞ 0 (Rd) with all derivatives bounded. To see xβf are bounded functions for β, we use induction on β and (3.2) θjkλθ(Djf ) . i λθ(xjf ) = xjλθ(f ) − 2Xk Similarly we know that Dα(f )xβ are bounded for all α, β. To show the semi-norms are equivalent, let f ∈ S(Rd) and denote f as its Fourier transform. Let n be the smallest even integer greater than d 2 , k Dβ(f )hxi2mk∞≤k \Dβ(f )hxi2mk1≤khξin \Dβ(f )hxi2mk2khξi−nk2 . Let hξin \Dβ(f )hxi2m ∈ S(Rd) be the Fourier transform of g. g can be expressed as a linear combination of xβDα(f ) with α up to n, β up to 2m. Therefore, k Dβ(f )hxi2mk∞ . k λθ(g)k2. k λθ(g)hxink∞ . sup{k Dαλθ(f )xβ k∞ α ≤ n,β ≤ n + 2m} . Finally, we note that Dαλθ(f ) ∈ Sθ ⊂ O−∞ and by Theorem 3.2 Dαhxir ∈ Or−α. By product rule, Dα(hxirλθ(f )) ∈ O−∞ for all α. Then hxirSθ ⊂ Sθ. (cid:3) Lemma 3.5. Let y ∈ Rd. Denote hx + yi := (1 +Pj(xj + yj)2) i) αy(hxir) = hx + yir. ii) for any 0 < r ≤ 2n with n integer, there exists a constant cr,n such that 2 . Then 1 khx + yirhxi−rk∞≤ cr,nhyi2n , khxirhx + yi−rk∞≤ cr,nhyi2n . QUANTUM EUCLIDEAN SPACES WITH NONCOMMUTATIVE DERIVATIVES 17 Proof. It is clear that hαy(x)i2 = 1 +Pj(xj + yj)2 = αy(hxi2). Then by the fact αy is a ∗-isomorphism on Mθ, αy(hxi−2) = hαy(x)i−2. Then we apply the operator integral for 0 < s < 2, (t + hxi2)−1t− s 2 dt . 0 hxi−s = CrZ ∞ 2yjxjhxi−2 +Xj Then the general case follows from writing r = 2n − s. For ii), for r = 2, khx + yi2hxi−2k≤k 1 +Xj k (hxi−2 − hx + yi2)(t + hxi2)−1k≤kXj jhxi−2k. hyi2 y2 2yjxj(t + hxi2)−1 +Xj j (t + hxi2)−1k. t− 1 y2 2hyi2 For r = 2n, hxi2n is a 2n-degree polynomial of xj whose largest coefficient is the constant term hyi2n. By a similar argument for hxi2n, we have khx + yi2nhxi−2nk. hyi2n , k (hxi−2n − hx + yi2n)(t + hxi2n)−1k. t− 1 2nhyi2 . Using the transference, khxi2nhx + yi−2nk=k αy(hx − yi2nhxi−2n)k=khxi2nhx + yi−2nk. hyi2n This proves the inequality for r = 2n even integers. For general positive r, choose integer n such that 0 < r < 2n − 1, consider 1 − hxirhx + yi−r = hxir(hxi−r − hx + yi−r). Take s = r 2n < 1 − 1 2n , we have hxir(hxi−r − hx + yi−r) =CshxirZ ∞ =CsZ ∞ 0 (cid:16)(t + hxi2n)−1 − (t + hx + yi2n)−1(cid:17)t−sdt . 0 (cid:16)hxir(t + hxi2n)−1(cid:17)(cid:16)(hx + yi2n − hxi2n)(t + hx + yi2n)−1(cid:17)t−sdt . (3.3) Note that khxir(t + hxi2n)−1k≤ (t + 1)s−1 and k (hx + yi2n − hxi2n)(t + hxi2n)k. t− 1 2nhyi2n . Therefore, khxir(hxi−r − hx + yi−r)k.Z ∞ 0 (1 + t)s−1t− 1 2n −shyi2ndt . hyi2n This proves the inequality for hxirhx + yi−r and the other case follows from transference. (cid:3) Using the above lemma, we show that quantized partial derivatives defined in Section 2.3 are indeed the vector derivatives of transference action. Proposition 3.6. Let ej = (0,· · · , 1,· · · , 0) be the j-th basis vector. i) for λθ(f ) ∈ Sθ, Djλθ(f ) = −i lim (αhej (λθ(f )) − λθ(f )) in Sθ. ii) Let m ∈ R. If a ∈ Mθ and Dα(a)hxim ∈ Rθ for all α ≤ 2, then h→0 1 h lim h→0 1 h k(cid:16)αhej (a) − a − hDj(a)(cid:17)hximk∞= 0 . 18 LI GAO, MARIUS JUNGE, AND EDWARD MCDONALD Proof. For a Schwartz function f ∈ S(Rd), we have that yj(∂jf )(x + ty)dt . yjαty(iDjf )dt . In terms of the function f , we have 0 f (x + y) − f (x) =Xj Z 1 αy(f ) − f =Xj Z 1 0 Since {αty(iDjf ) 0 ≤ t ≤ 1} is uniformly bounded for every semi-norm of S(Rd), we have y → αy(f ) is continuous in S(Rd). Because Sθ and S(Rd) have equivalent semi-norms, we have y 7→ αy(λθ(f )) = λθ(αyf ) is also continuous. 1 h(cid:16)αhej (λθ(f )) − λθ(f ) − hλθ(iDjf )(cid:17) =Z 1 =Z 1 0 (cid:16)αthej λθ(iDjf ) − λθ(iDjf )(cid:17)dt αthj λθ(iDjf ) − λθ(iDjf )dt 0 which goes to 0 in Sθ for h → 0 because of the continuity of y → αy(λθ(Djf )). For ii), we have the integral αy(a)hxim − ahxim =Xj yjZ 1 0 αty(iDja)hximdt . (3.4) which holds weakly. Suppose ahxim and Dj(a)hxim are bounded. Then k αy(Dja)hximk≤k αy(Djahxim)kkhx + yi−mhximk≤k Djahximk hyi2n . for some 2n > m. So αy(Dja)hxim is uniformly bounded for small y, which by the integral (3.4) implies y 7→ αy(a)hxim is continuous in norm. Now if Dα(a)hxim bounded for all α ≤ 2, then 1 h(cid:16)αhej (a) − a − hDj(a)(cid:17)hximk∞≤Z 1 k 0 k(cid:0)αthej (iDja) − iDja(cid:1)hximk∞ dt This goes 0 in norm as h → 0 because y → αy(Dja)hxim is continuous. The next proposition gives an approximation of identity for Lp(Rθ). (cid:3) Proposition 3.7. There exists a sequence fn ∈ S(Rd) independent of θ such that i) for any a ∈ Eθ and p = ∞; and ii) for any a ∈ Lp(Rθ) and 1 ≤ p < ∞, n→∞ k aλθ(fn) − akp= lim lim n→∞ k λθ(fn)a − akp= 0 . Proof. Let φ ∈ S(Rd) be a smooth positive function such that φ supported on x ≤ 1 and R φ = (2π)d. Take φn = ndφ(nx) and the inverse Fourier transform φn. We first show that QUANTUM EUCLIDEAN SPACES WITH NONCOMMUTATIVE DERIVATIVES 19 2 ξθ(η−ξ)dξ. Given ǫ > 0, we can find R and n large such for any λθ(g) ∈ Sθ, k λθ(g)λθ( φn) − λθ(g)k∞→ 0. Indeed 2πdZRd λθ(g)λθ( φn) =(cid:16) 1 g(ξ)λθ(ξ)dξ(cid:17)(cid:16) 1 g(ξ)φn(η)e 2 ξθηλθ(ξ + η)dξdη i 1 2πdZRd 2π2dZRdZRd 2π2dZRd(cid:16)ZRd 1 = = φn(η)λθ(η)dη(cid:17) 2 ξθ(η−ξ)dξ(cid:17)λθ(η)dη := λθ(gn) i g(ξ)φn(η − ξ)e i i ǫ 1 1 3g1 where gn = 2 ξθη < k g − gnk1= g(ξ)φn(η − ξ)e ǫ and 1 − e 3 2πdZRd thatZξ<R g(ξ) < 2πdZRd g(η) −ZRd 2πdZRdZRd g(ξ)φn(η − ξ)(1 − e 2πdZξ>RZRd g(ξ)φn(η − ξ)(1 − e 2πdZξ<RZRd g(ξ)φn(η − ξ)(1 − e 2πdZξ>RZRd 2g(ξ)φn(η − ξ)dηdξ + g(ξ)φn(η − ξ)e ≤ ≤ ≤ + 1 1 1 1 2ǫ 3 ≤ ǫ 3 + = ǫ for all ξ < R. Then, i 2 ξθ(η−ξ)dξdη 2 ξθ(η−ξ))dξdη i i 2 ξθ(η−ξ))dηdξ i 2 ξθ(η−ξ))dηdξ 2πdZξ<RZRd 1 ǫg(ξ)φn(η − ξ)dηdξ Hence k λθ(gn) − λθ(g) k∞≤k gn − g k1→ 0. For 1 ≤ p < ∞, we apply the argument for hxidλθ(g). Note that hxid+1λθ(g) ∈ Sθ by Proposition 3.4. Thus we have k λθ(g)λθ(fn) − λθ(g)kp≤khxid+1(λθ(g)λθ(fn) − λθ(g))k∞khxi−d−1kp→ 0 . Given a ∈ L1(Rθ), we choose g ∈ Sθ so that k λθ(g) − ak1≤ ǫ/3. Note that for all n, k λθ( φn)k∞≤k φnk1= 1 . Then for n large enough, k a − aλθ( φn)k1≤ k a − λθ(g)k1 + k λθ(g) − λθ(g)λθ( φn)k1 + k λθ(g)λθ( φn) − aλθ( φn)k1 ≤ k a − λθ(g)k1 + k λθ(g) − λθ(g)λθ( φn)k1 + k λθ(g) − ak1k λθ( φn)k∞ ǫ ≤ 3 = ǫ ǫ 3 ǫ 3 + + (3.5) The argument for ∞-norm and a ∈ Eθ is similar. For 1 < p < ∞, we use interpolation inequality that k a − aλθ( φn)kp≤k a − aλθ( φn)k 1 p 1 k a − aλθ( φn)k 1− 1 p ∞ → 0 . for any a ∈ L1(Rθ)∩ L∞(Rθ). Since L1 ∩ L∞ is dense in Lp, the argument for general a ∈ Lp is similar to (3.5). (cid:3) 20 LI GAO, MARIUS JUNGE, AND EDWARD MCDONALD 4. Pseudo-differential Calculus for Non-commutative Derivatives On Rd the CCR relation for covariant derivatives corresponds to a constant curvature form. Consider connection ∇ : C∞(Rd) → Ω1(Rd) , ∇f = df + i with curvature form dω = i ∇− ∂ satisfy that ∂j 2Pj,k θjkdxj ∧ dxk . The self-adjoint covariant derivatives ∇j = (4.1) 2Pj,k θ′ j,kxjdxk ∇jf = −i ∂ ∂xj 1 2 θ′ jkxk , [∇j,∇k] = −iθ′ jk . (f ) −Pk The physical meaning behind this is a constant magnetic field perpendicular to the space Rd. In this section, we develop the symbol calculus of ΨDOs of the above structure for a noncommutative Rθ. Let Rθ be the quantum Euclidean space generated by [xj, xk] = −iθjk. We equipped Rθ with noncommuting covariant derivatives ξj satisfying [ξj, xk] = −iδjk, [ξj, ξk] = −iθ′ jk . (4.2) I where δ is the Kronecker delta notation. For θ′ = 0, [GJP17] establish the ΨDOs as operators on L2(Rθ) via ξj = Dj. For general θ and θ′, xj's and ξk's satisfying above commutation relations together generate a 2d-dimensional quantum Euclidean space RΘ with parameter because Θ can be singular. Hence we consider the ΨDOs as operators (densely) defined on θ′ (cid:21). In general xj's and ξk's do not admit a canonical representation on L2(Rθ) Θ =(cid:20) θ −I L2(RΘ) ∼= L2(Rθ) ⊗2 L2(Rθ′) affiliated to RΘ . Here ⊗2 is the Hilbert space tensor product. 4.1. Abstract symbols. In the classical case for Rd, a symbol of order m is a smooth bi-variable function a ∈ C ∞(Rd × Rd) such that the x Dβ Dα ξ (a)(x, ξ) ≤ Cα,β(1 + ξ2)(m−β)/2 . (4.3) In our setting, the symbols are operators affiliated to the von Neumann algebra tensor prod- uct Rθ⊗Rθ′. Let us denote Rθ,θ′ := Rθ⊗Rθ′, Mθ,θ′ for the multiplier algebra of Rθ,θ′ and Sθ,θ′ for the Schwartz class. Rθ,θ′ is a 2d-dimensional quantum Euclidean space with parameter 0 θ′ (cid:21), in which x and ξ variables are mutually commuting, i.e. [xj, ξk] = 0 for all j, k. We specify the canonical partial derivatives for x variables by Dx1,· · · , Dxd and for ξ variables by Dξ1,· · · , Dξd. That is, for a ∈ Mθ,θ′ matrix(cid:20) θ 0 Dxj (a) = [Dj ⊗ 1, a] , Dξj (a) = [1 ⊗ Dj, a] . We index the transference action by the position: αy ⊗ αη(a) = α1 standard multi-derivative notation that for α = (α1, α2,· · · , αd) ∈ Nd , ξ2 · · · Dαd x2 · · · Dαd ξ (a) = Dα1 x (a) = Dα1 xn , Dα ξ1 Dα2 x1 Dα2 Dα ξd (a) . Write hξi := (1 +Pj ξ2 j ) 1 2 where ξj's are the non-commuting generators for Rθ′. We start with the abstract reformulation of the definition (4.3). Definition 4.1. For a real number m, define Σm as the set of all operators a ∈ Mθ,θ′ such that for all α, β, ηα2 y(a). We use the Dα x Dβ ξ (a)hξiβ−r QUANTUM EUCLIDEAN SPACES WITH NONCOMMUTATIVE DERIVATIVES 21 extends to be a bounded operator in Rθ,θ′. We call Σm the space of symbols of order m and write Σ−∞ = ∩mΣm, Σ∞ = ∪mΣm. Apriori it is not clear that the above definition satisfy the properties that Σm·Σn = Σm+n and (Σm)∗ = Σm. To resolve it, we use the asymptotic degree discussed in Section 3. Definition 4.2. Given two real numbers s and r, we say an operator a ∈ Mθ,θ′ is of bi-degree (s, r) if for all s′, r′ ∈ R ahxi−s′−shξi−r′−r extends to a bounded element in Rθ,θ′. We denote Os,r the set of all elements of bi-degree (s, r) and write O−∞,r = ∩s∈ROs,r, O−∞,−∞ = ∩s,r∈ROs,r. hxis′ hξir′ Note that in Rθ,θ′, hxi and hξi commute so the order of the product hxishξir does not matter. The "bi-degree" gives a characterization of abstract symbols. Theorem 4.3. Let m be a real number and a ∈ Mθ,θ′. Then a ∈ Σm if and only if for all α, β, Dα x Dβ ξ (a) ∈ O0,m−β . Proof. The sufficiency is clear by the definition. Let a ∈ Σm. It follows from the Lemma 3.4 x Dβ that for all α, β, Dα ξ (a) is of degree 0 for x and degree m − β for ξ. Because hxi and hξi x Dβ commute, we have Dα (cid:3) ξ (a) ∈ O0,m−β. Proposition 4.4. Σm equipped with the seminorms k·kα,β:=k Dα spaces. In particular, for a ∈ Σm, Dxj (a) and Dξj (a) are the vector derivatives ξ (·)hξiβ−mk is a Frechet x Dβ 1 h Dxj (a) = i lim h→0 (α1 hej (a) − a) , Dξj (a) = i lim h→0 (α2 hej (a) − a) , 1 h where the limit converges in the Σm. Proof. Let an ∈ Σm be a converging sequence in Σm with respect to all the seminorms k·kα,β. Then there exists bα,β ∈ Rθ,θ′ such that x Dβ k Dα ξ (an)hξiβ−m − bα,β k∞→ 0 as n → ∞ . Denote that cα,β = bα,βhξim−β and C0,0 = b0,0hξim. Let λθ,θ′(f ) ∈ Sθ,θ′. hcα,β,hξiβ−mλθ,θ′(f )i = hbα,βhξiβ−m, λθ,θ′(f )i = hbα,β, λθ,θ′(f )i = lim ξ (an)hξiβ−m, λθ,θ′(f )i x Dβ n→∞hDα n→∞hanhξi−m,hξimDα x Dβ ξ (hξiβ−m(λθ,θ′(f ))i = lim = hb0,0,hξimDα = hDα x Dβ x Dβ ξ (hξiβ−mλθ,θ′(f ))i ξ (c0,0),hξiβ−mλθ,θ′(f )i . Note that the set hξiβ−mSθ,θ′ = Sθ,θ′ by Proposition 3.4. We have cα,β = Dα ξ (c0,0) weakly. To see that c0,0 is again in the multiplier algebra Mθ,θ′, it suffices to show that for any λθ,θ′(f ) ∈ Sθ,θ′, x Dβ x Dβ k Dα ξ (c0,0λθ,θ′(f ))(1 +Xj x2 j + ξj)γ k 22 LI GAO, MARIUS JUNGE, AND EDWARD MCDONALD is bounded for any α, β, γ. This follows from Leibniz rule and the fact λθ,θ′(f ) and all ξ (λθ,θ′(f )) are in O−∞,−∞. The vector derivatives are consequence of its derivatives Dα applying Proposition 3.6 to Rθ,θ′. (cid:3) x Dβ Corollary 4.5. For all multi-indices α and real numbers m, n, i) ξα ∈ Σα, hξim ∈ Σm; ii) if a ∈ Σm, then a∗ ∈ Σm; iii) if a ∈ Σm, b ∈ Σn, then ab ∈ Σm+n. Proof. i) is a direct consequence of Theorem 3.2. ii) follows from the fact that For iii), by the Leibniz rule Dα x Dβ . x Dβ ξ (a∗) = (−1)α+β(cid:16)Dα α1, α2(cid:19)(cid:18) β ξ (a)(cid:17)∗ β1, β2(cid:19)Dα1 Dα x Dβ ξ (ab) = Xα1+α2=α, β1+β2=β(cid:18) α x Dβ1 ξ (a)Dα2 x Dβ2 ξ (b) . (4.4) Using Theorem 4.3, Dα1 x Dβ1 ξ (a) ∈ O0,m−β1 , Dα2 x Dβ2 ξ (b) ∈ O0,n−β2 . Hence all summands in (4.4) are belongs to O0,m+n−β1−β2 = O0,m+n−β. Again by Theorem 4.3, ab ∈ Σn+m. (cid:3) 4.2. Comultiplications. One key tool that will be used in the proof of our symbol calculus is the the comultiplication maps of Rθ and Rθ,θ′. The comultiplication map of Rd as an abelian group is σ : L∞(Rd) → L∞(Rd × Rd) ∼= L∞(Rd × Rd) , σ(f )(x, y) = f (x + y) . Algebraically, σ(u(ξ)) = u(ξ) ⊗ u(ξ) where u(ξ) is the unitary function u(ξ)(x) = eiξ·x. For Rθ, we consider the a deformed comultiplication map σθ : Rθ → L∞(Rn)⊗Rθ , σθ(λθ(ξ)) = u(ξ) ⊗ λθ(ξ) , where ⊗ is the von Neumann algebra tensor product. L∞(Rn)⊗Rθ can be identified with Rθ-valued functions L∞(Rd, Rθ), and at a point x ∈ Rd, σθ(λθ(ξ))(x) = eix·ξλθ(ξ) = αx(λθ(ξ)) . A different co-multiplication map is used in [GJP17] to studied ΨDOs of Rθ with commuting derivatives. Proposition 4.6. The map σθ : Sθ → L∞(Rd, Rθ) σθ(λθ(f ))(x) = αx(λθ(f )) , i) extends to an injective normal ∗-homomorphism from Rθ to L∞(Rd, Rθ). ii) extends to an injective algebraic ∗-homomorphism from Mθ to L∞(Rd,Mθ). More- over, for all a ∈ Mθ, σθ(Dja) = Dxj (σθ(a)) = Dxj (σθ(a)) . 2(Rd) ⊗wh Rθ. Here ⊗wh iii) extends to an complete isometry Vθ right from L2(Rθ)c to Lc denotes the W ∗-Haagerup tensor product (see [BS92]) and Lc 2(Rd) is the column space. QUANTUM EUCLIDEAN SPACES WITH NONCOMMUTATIVE DERIVATIVES 23 Proof. i) follows from the fact that at each point x ∈ Rd, αx is a ∗-automorphism of Rθ. The normality was proved in [GJP17, Corollary 1.4]. ii) is similar to i). For the derivatives, let Dxj denote the jth partial derivatives for Rd and Dxj denote the partial derivatives on Rθ. For all x ∈ Rd and a ∈ Mθ, Dxj (σθ(a))(x) = lim h→0− i h(cid:0)αx+hej (a) − αx(a)(cid:1) = Dxj (αx(a)) = αx(Dxj a) . For iii), let b = Pk bkλθ(fk) with bk ∈ C and λθ(fk) being an orthonormal set in L2(Rθ). 2(Rd) ⊗wh Rθ is given by the Rθ-valued inner Then k b k2 product that for f, g ∈ L2(Rd) and a, c ∈ Rθ L2(Rθ)= Pk bk2. The norm of Lc hf ⊗ a, g ⊗ ciRθ = hf, giL2(Rd)a∗c , k BkLc 2(Rd)⊗whRθ=khB, BiRθ kRθ Note that on the Fourier transform side, Vθ(λθ(f ))(ξ) = f (ξ)λθ(ξ) . Therefore, k Vθ(Xk = k (Xk bkλθ(fk))kLc bk2)1kRθ=Xk 2(Rd)⊗whRθ =kXk,k′ bk2 . bk ¯bk′Z fk(ξ) f k′(ξ)λθ(ξ)λθ(ξ)∗dξkRθ Replacing bk ∈ C with matrices bk ∈ Mn in the above argument gives the complete isometry. (cid:3) Let us write λθ,θ′(η, y) := λθ(η) ⊗ λθ′(y) for the generators of Rθ,θ′ := Rθ⊗Rθ′. The quantization map for Rθ,θ′ is λθ,θ′(F ) = (2π)−2dZR2d F (η, y)λθ,θ′(η, y)dηdy , where F (η, y) = ZR2d F (x, ξ)e−i(xη+ξy)dxdξ is the Fourier transform. By the Proposition 4.6, we can dilate the symbols affiliated to Rθ,θ′ to operator valued symbols, σθ ⊗ σθ′ : Rθ,θ′ → L∞(Rd × Rd, Rθ⊗Rθ′) , λθ,θ′(F )(x, y) = α1 xα2 y(λθ,θ′(F )), where α1 (resp. α2) is the transference action on Rθ (resp. Rθ′). For the ΨDOs, we consider the comultiplication maps for RΘ with Θ =(cid:20) θ −In θ′ (cid:21). We use the following quantization In for RΘ, λΘ(F ) = (2π)−2dZRdZRd F (η, y)λθ(η)λθ′(y)dηdy , F ∈ S(Rd × Rd) . Note that the unitary generators in RΘ satisfy the commutation relation λθ(η)λθ′(y) = eiηyλθ′(y)λθ(η) . We have the Hilbert space isometry between two quantizations, W : L2(RΘ) → L2(Rθ,θ′) , WλΘ(F )i = λθ,θ′(F )i . Here and in the following, we will use the "ket" notation ·i to emphasis L2 vector. 24 LI GAO, MARIUS JUNGE, AND EDWARD MCDONALD Proposition 4.7. Define the unitary uθ(y) : L2(Rθ) → L2(Rθ) , vθ(y)λθ(f )i = λθ(αyf )i . The map σΘ : SΘ → B(L2(Rθ))⊗Rθ′ λΘ(F ) 7→ (2π)−2dZR2d i) satisfies that σΘ(λΘ(F )) = W λΘ(F )W ∗ by viewing F (η, y)λθ(η)vθ(y) ⊗ λθ′(y)dηdy SΘ ⊂ B(L2(RΘ)) , B(L2(Rθ))⊗Rθ′ ⊂ B(L2(Rθ) ⊗2 L2(Rθ′)) . ii) extends to an injective normal ∗-homomorphism from RΘ to B(L2(Rθ))⊗Rθ′. Proof. By linearity, it suffices to verify that W λθ(η0)λθ′(y0)W ∗ = λθ(η0)vθ(y0) ⊗ λθ′(y0). Indeed, for λθ,θ′(G) ∈ Sθ,θ′, W λθ(η0)λθ′(y0)W ∗λθ,θ′(G)i =W λθ(η0)λθ′(y0)λΘ(G)i = WλΘ(G1)i λΘ(G1) =ZR2d =ZR2d G(η, y)λθ(η0)λθ′(y0)λθ(η)λθ′(y)dydη G(η − η0, y − y0)eiηy0e 2 (ηθη0+yθ′y0)λθ(η)λθ′(y)dydη . i where Then WλΘ(G1)i = λθ,θ′(G1)i =(cid:16)λθ(η0)vθ(y0) ⊗ λθ′(y0)(cid:17)λθ,θ′(G)i . Now let us consider the GNS-construction of B(L2(Rθ)) with respect to its standard (cid:3) trace. Define for a Schwartz function F the operator TF = (2π)−2dZR2d F (η, y)λθ(η)vθ(y)dηdy . For λθ(f )i ∈ L2(Rθ), TFλθ(f )i = (2π)−2dZ F (η, y)λθ(η)vθ(y)dηdyλθ(f )i =: λθ(g)i where TF has the following kernel representation, Since F ∈ S(R2d), TF is trace class and i bg(η) = (2π)−2dZ F (η − ξ, y)eiyηe tr(TF ) = (2π)−2dZ F (0, y)eiyηdydη = (2π)−dZ F . 2 ηθξdy f (ξ)dξ . One calculates that F TF = (2π)−4dZR2d(cid:16)ZR2d T ∗ Hence tr(T ∗ F (η1, y1) F (η + η1, y + y1)e− i we have a Hilbert space isometry F TF ) = (2π)−2dZR2d V : L2(B(L2(Rθ)), tr) → L2(Rd, L2(Rθ)) , V (TF )(x) = λθ(F (x,·)) . F (η1, y1) F (η1, y1)dη1dy1 = (2π)−2d k F k2 2 ηθη1e−iη1ydη1dy1(cid:17)λθ(η)vθ(y)dηdy 2 . Up to a scalar QUANTUM EUCLIDEAN SPACES WITH NONCOMMUTATIVE DERIVATIVES 25 Write π as the GNS construction of B(L2(Rθ)) on L2(B(L2(Rθ)), tr). Then π(·) = V π(·)V ∗ gives a normal faithful ∗-homomorphism form B(L2(Rθ)) to B(L2(Rd))⊗Rθ as follow, F (η, y)v(η)u(y) ⊗ λθ(η)dηdy ∈ B(L2(Rd))⊗Rθ , π(TF ) := V π(TF )V ∗ = (2π)−2dZR2d where v(η) is translation unitary on L2(Rd). Combining π with the co-multiplication σΘ, we obtain another co-multiplication of RΘ. Proposition 4.8. The map σΘ : SΘ → B(L2(Rd))⊗Rθ,θ′ λΘ(F ) 7−→ (2π)−2dZ F (η, y)(cid:16)u(η)v(y) ⊗ λθ,θ′(η, y)(cid:17)dηdy θ′ ) for the isometry θ′ : Lc Vθ ⊗ idR i) extends to a normal injective ∗-homomorphism from RΘ to B(L2(Rd))⊗Rθ,θ′. ii) satisfies the intertwining relation (Vθ⊗idR 2(Rθ) ⊗wh Rθ′ → Lc θ′ ) ◦ σΘ. Indeed θ′ )σΘ(·) = σΘ(·)(Vθ⊗idR 2(Rd) ⊗wh (Rθ⊗Rθ′) . F (η, y)λθ(η)vθ(y) ⊗ λθ′(y)dηdy(cid:17) =(2π)−2dZ F (η, y)(cid:16)u(η)v(y) ⊗ λθ(η) ⊗ λθ′(y)(cid:17)dηdy = σΘ(λΘ(F )). θ′(cid:16)(2π)−2dZR2d Proof. i) We verify that σΘ = (π ⊗ idR θ′ )◦σΘ(λΘ(F )) = π ⊗ idR (π ⊗ idR For ii), recall that B(L2(Rθ))⊗Rθ′ is canonically isomorphic to the adjointable R′ θ-module map L(Lc 2(Rθ) ⊗wh Rθ,θ′) as Rθ,θ′- module map (see [Lan95]). The complete isometry Vθ in Proposition 4.6 give an isometry 2(Rθ) ⊗wh Rθ′) and similarly B(L2(Rd))⊗Rθ⊗Rθ′ ∼= L(Lc Vθ ⊗ idθ′ : Lc 2(Rθ) ⊗wh Rθ′ → Lc 2(Rd) ⊗wh (Rθ⊗Rθ′) . We verify that the intertwining relation (Vθ ⊗ id)σΘ(·) = σΘ(·)(Vθ⊗ id). For any λΘ(F ) ∈ SΘ and λθ,θ′(G) ∈ Sθ,θ′, we have σΘ(λΘ(F ))λθ,θ′(G)i = λθ,θ′(G1)i where G1(η, y) = (2π)−2dZ F (η − η1, y − y1) G(η1, y1)eiη1(y−y1)e On the other hand, one verifies that i 2 ηθη1e i 2 yθy1dη1dy1 σΘ ⊗ id(λΘ(F ))Vθλθ,θ′(G)i = Z G1(η, y)u(η) ⊗ λθ,θ′(η, y)dηdyi =Vθ ⊗ id(cid:16)σΘ(λΘ(F ))λθ,θ′(G)i(cid:17) We see that the representation (Vθ ⊗ id)∗σΘ(·)(Vθ ⊗ id) is a restriction of σΘ. 4.3. Pseudo-differential operator calculus. Recall that on Rd the pseudo-differential operator of a symbol a(x, ξ) is given by the singular integral form (cid:3) In [GJP17] the ΨDOs on Rθ are defined as op0(a)(f )(x) := opθ(a)(λθ(f )) = 1 (2π)dZRd (2π)dZRd 1 eix·ξa(x, ξ) f (ξ)dξ , f ∈ S(Rd) a(ξ)λθ(ξ) f (ξ)dξ , f ∈ S(Rd) . (4.5) (4.6) 26 LI GAO, MARIUS JUNGE, AND EDWARD MCDONALD where a : Rd → Rθ is the symbol as a Rθ-valued function. The ΨDOs in our setting are operators densely defined on L2(Rθ,θ′) ∼= L2(Rθ) ⊗2 L2(Rθ′). For a symbol a1 ⊗ a2 with a1 ∈ Rθ, a2 ∈ Rθ′, we define that Op(a1 ⊗ a2) = σΘ(a1a2) ∈ B(L2(Rθ,θ′)) where a1a2 is the product in RΘ by viewing Rθ, R′ representation of RΘ on L2(Rθ,θ′) defined in Proposition 4.7. Definition 4.9. For a symbol a ∈ Σm, we define the operator Op(a) : Sθ,θ′ → Sθ,θ′ as follows, θ ⊂ RΘ as subalgebras and σΘ is the Op(a)λθ,θ′(F ) = η(a) F (η, y)λθ,θ′(η, y)dηdy α2 1 (2π)2dZR2d We denote by opm the set of all ΨDOs of order m. We justifies the above definition below. Proposition 4.10. For a symbol a ∈ Σm, Op(a) is a continuous map form Sθ,θ′ to Sθ,θ′ and Op(a) is an operator affiliated to σΘ(RΘ) ⊂ B(L2(Rθ,θ′)). In particular, if a1 ∈ Rθ and a2 ∈ Rθ′, Op(a1 ⊗ a2) = σΘ(a1a2). Proof. In the calculation below, the normalization constant (2π)−d will be omitted. Recall from Proposition 4.7 that W : L2(RΘ) → L2(Rθ,θ′) , WλΘ(F )i = λθ,θ′(F )i , is the isometry such that W ∗σΘ(·)W is the left regular representation of RΘ on L2(RΘ). To verify that Op(a) is affiliated to σΘ(RΘ), it suffices to show that W Op(a)W ∗ commutes with right multiplication of RΘ. For any η0, y0 ∈ Rd, λΘ(F )λθ(η0)λθ′(y0) =(cid:16)ZR2d F (η, y)λθ(η)λθ′(y)dηdy(cid:17)λθ(η0)λ′ =ZR2d η0(cid:0)λθ,θ′(F )(cid:1)λθ,θ′(η0, y0). We verify that F (η, y)eiyη0λθ(η)λθ(η0)λθ′(y)λθ′(y0)dηdy . θ(y0) Then W (λΘ(F )λθ(η0)λθ′(y0)) = α2 η+η0(a) F (η, y)eiyη0e α2 Op(a)W(cid:16)λΘ(F )λθ(η0)λθ′(y0)(cid:17) η0(cid:0)λθ,θ′(F )(cid:1)λθ,θ′(η0, y0)(cid:17) =Op(a)(cid:16)α2 =ZR2d =(cid:16)ZR2d η0(cid:16)ZR2d η0(cid:16)Op(a)λθ,θ′(F )(cid:17)λθ,θ′(η0, y0) . η+η0(a) F (η, y)α2 α2 2 ηθη0e =α2 =α2 η0(λθ,θ′(η, y))dηdy(cid:17)λθ,θ′(η0, y0) α2 η(a) F (η, y)λθ,θ′(η, y)dηdy(cid:17)λθ,θ′(η0, y0) i i 2 yθ′y0λθ,θ′(η + η0, y + y0)dηdy Hence W ∗Op(a)W(cid:16)λΘ(F )λθ(η0)λθ′(y0)(cid:17) =(cid:16)W ∗Op(a)W λΘ(F )(cid:17)λθ(η0)λθ′(y0) , which implies Op(a) is affiliated to the representation on σ(RΘ) ⊂ B(L2(Rθ) ⊗2 L2(Rθ′)). QUANTUM EUCLIDEAN SPACES WITH NONCOMMUTATIVE DERIVATIVES 27 Now we show that Op(a) : Sθ,θ′ → Sθ,θ′ is continuous. Let us first assume that a ∈ Σ0 is η(a)k∞ for all η. Thus the a zero order symbol. Then a is bounded in Rθ,θ′ and k ak∞=k α2 singular integral kZR2d η(a) F (η, y)λθ,θ′(η, y)dηdyk∞≤k F k1k ak∞ α2 converges in Rθ,θ′. Write the set Ω := {Op(a)λΘ(F ) F ∈ S(R2d) , a ∈ Σ0} ⊂ Rθ,θ′. For ξ (a) ∈ Σ−γ. derivatives, we know Dxj (λθ(η)) = ηjλθ(η) , Dξj (λθ′(y)) = yjλθ′(y) and Dβ Using product rules in the integral, x Dγ Dξj(cid:16)Op(a)λθ,θ′(F )(cid:17) =Dξj(cid:16)ZR2d =ZR2d η(a) F (η, y)λθ(η) ⊗ λθ′(y)dηdy(cid:17) η(Dξj a) F (η, y)λθ,θ′(η, y)dηdy +ZR2d =Op(Dξj a)λθ,θ′(F ) + Op(a)λθ,θ′(Dξj F ), α2 α2 η(a) F (η, y)yjλθ,θ′(η, y)dηdy α2 x Dγ which is again in the set Ω hence bounded in Rθ,θ′. By induction, Dβ in Ω for any β, γ. On the other hand, let h ∈ R and ej = (0,· · · , 1,· · · , 0) ξ (Op(a)λθ,θ′(F )) is λθ(η)eixjh = e− i 2 Pk hθjk ηkλθ(η + hej) , λθ′(y)eiξj h = e− i 2 Pk hθ′ jkyk λθ(y + hej) . Taking derivatives at h = 0, λθ(η)xj = Dηj (λθ(η)) − holds weakly. Then 1 2Xk θjkηkλθ(η) , λθ(y)ξj = Dyj (λθ′(y)) − θ′ jkηkλθ′(y) . 1 2Xk (cid:16)Op(a)λθ,θ′(F )(cid:17)xj 2Z α2 =Z α2 η(a) F (η, y)Dηj(λθ,θ′(η, y))dηdy − η(Dξj a) F (η, y)(λθ,θ′(η, y))dηdy −Z α2 = −Z α2 2Z α2 η(a) F (η, y)(Xk θjkηk)λθ,θ′(η, y)dηdy − 1 1 = − Op(Dξj a)λθ,θ′(F ) − Op(a)λθ,θ′(ξjF ) − η(a) F (η, y)(Xk θjkηk)λθ,θ′(η, y)dηdy η(a)(Dηj F )(η, y)(λθ,θ′(η, y))dηdy θ′ jkOp(a)λθ,θ′(DξkF ) 1 2Xk which is again in the set Ω. By induction, Ω is stable under right multiplication of polyno- mials xβξγ. By Proposition 3.4, we know Ω ⊂ Sθ,θ′ because for all β1, β2, γ1, γ2 x Dγ1 k Dβ1 ξ (Op(a)λθ,θ′(F ))xβ2ξγ2 k∞< ∞ . Moreover, one can track that these norms are controlled by the semi-norms of a ∈ Σ0 and λθ,θ′(F ) ∈ Sθ,θ′. Thus we proved Op(a) : Sθ,θ′ → Sθ,θ′ is continuous for 0-order ΨDO. Now 28 LI GAO, MARIUS JUNGE, AND EDWARD MCDONALD consider b ∈ Σm with m being an even integer, we know b = bhξi−mhξim and bhξi−m is a zero order symbol, hξim is a polynomial. Note that for a ∈ Σ0, Op(aξj)λθ,θ′(F ) =ZR2d =ZR2d =ZR2d η(aξj) F (η, y)λθ,θ′(η, y)dηdy α2 (ξj + ηj)α2 η(a) F (η, y)λθ,θ′(η, y)dηdy ξjα2 η(a) F (η, y)λθ,θ′(η, y)dηdy +ZR2d =ξjOp(a)λθ,θ′(F ) + Op(a)λθ,θ′(Dxj F ) η(a) F (η, y)ηjλθ,θ′(η, y)dηdy α2 which is again in Ω. Moreover, the continuity of Op(aξj) follows from the continuity of Op(a). By induction, we obtain that Op(a) : Sθ,θ′ → Sθ,θ′ is continuous for Op(a) ∈ Σm for all m. Finally, we verify the property that Op(a1 ⊗ a2) = σ(a1a2). It suffices to consider test functions λθ,θ′(F ) = λθ(f1) ⊗ λθ′(f2) with F (x, ξ) = f1(x)f2(ξ). Then Op(a1 ⊗ a2)λθ,θ′(F ) =Z (cid:0)a1 ⊗ αη(a2)(cid:1) f1(η) f2(y)(cid:0)λθ(η) ⊗ λθ′(y)(cid:1)dηdy =Z f1(η)a1λθ(η) ⊗ (αη(a2)λθ′(f2))dη =W ∗(Z f1(η)a1λθ(η)αη(a2)λθ′(f2)dη) =W ∗(a1a2Z f1(η)λθ(η)λθ′(f2)dη) =W ∗(a1a2λθ(f1)λθ′(f2)) = W ∗(a1a2)W(cid:16)λθ(f1) ⊗ λθ′(f2)(cid:17) . Here we use the fact that for a2 ∈ Mθ′, a2λθ(η) = λθ(η)αη(a2) . This property be easily verified for a2 ∈ Sθ′ and then extends to Mθ′. (cid:3) Based on the above proposition, we can equivalently consider Op(a) are operators affili- ated to RΘ and Op(a) ∈ RΘ if it is bounded. The connection between our setting and ΨDOs on Rd and Rθ can be made explicit via the following commuting diagram. Σ0 ⊂ Rθ,θ′ Op RΘ σΘ opθ ⊗ idR θ′ B(L2(Rθ)) ¯⊗Rθ′ id ⊗ σθ′ Rθ ¯⊗L∞(Rd, Rθ′) σθ ⊗ id L∞(Rd × Rd, Rθ,θ′) Here σθ, σθ′, σΘ are the co-multiplication maps discussed in section 3.2. The composition σΘ ◦ Op gives the definition 4.9. On the second row, the co-multiplication id ⊗ σθ′(a)(η) = 2(Rd) ⊗wh Rθ,θ′) op0 ⊗ idR L(Lc 2(Rθ) ⊗wh Rθ′) Vθ(·)V ∗ θ θ,θ′ L(Lc QUANTUM EUCLIDEAN SPACES WITH NONCOMMUTATIVE DERIVATIVES 29 α2 η(a) gives Rθ′-valued symbol, and Definition 4.9 is then coincides with the Rθ′-valued oper- ator map opθ⊗id on Rθ in (4.6). Via the identification B(L2(Rθ))⊗Rθ′ ∼= L(L2(Rθ)c⊗wh Rθ′) ([Lan95]), this also gives operators on Hilbert Rθ′-module L2(Rθ)c ⊗wh Rθ′. On the bottom row, we have a Rθ,θ′-valued classical symbol σθ ⊗ σ′ ξ(a), and op0 ⊗ idθ,θ′ is the Rθ,θ′-valued operator map on Rd in (4.5). The ΨDOs are Rθ,θ′-linear operators on the Hilbert module L2(Rd)c ⊗wh Rθ,θ′. By Proposition 4.8, we have the Hilbert space isometry θ(a)(x, ξ) = α1 xα2 Vθ ⊗ idR θ′ : L2(Rθ)c ⊗wh Rθ′ → L2(Rd)c ⊗wh Rθ,θ′ . Moreover, for a symbol a ∈ Σ0, the operator Op(a) can be viewed as a restriction of the Rθ,θ′-valued ΨDO op0 ⊗ id(σθ,θ′(a)) as follows, op0 ⊗ id(cid:0)σθ ⊗ σ′ =(2π)−dZ eixξα1 =αx(cid:16)(2π)−dZ α2 ξ(a) F (ξ, y)λθ,θ′(ξ, y)dξdy . θ(a)(cid:1)(cid:16)Vθ ⊗ id(λθ,θ′(F ))(cid:17) ξ(a) F (ξ, y)λθ,θ′(ξ, y)dξdy(cid:17) = Vθ ⊗ id(Op(a)λθ,θ′(F )) xα2 This enable us to reduce the L2-boundedness to the operator-valued case. For that we recall the operator-valued Calderon-Vallicourt theorem proved by Merklen in [Mer05]. Theorem 4.11 (Theorem 2.1 of [Mer05]). Let A be a C ∗-algebra and CB∞(Rd × Rd ,A) be the set of smooth A-valued functions with bounded derivatives of all orders. Then for any a ∈ CB∞(Rd × Rd ,A), op(a)f (x) = eix·ξa(x, ξ) f (ξ)dξ , f ∈ S(Rd,A) 1 (2π)dZRd extends to a bounded operator on the Hilbert A-module L2(Rd,A). Moreover, there exists a constant C independent of a, such that k op(a)k≤ C sup{k Dα ξ (a)k∞ 0 ≤ α, β ≤ (1, 1,· · · , 1)} . xDβ Then L2-boundedness theorem in our setting follows from the commuting diagram. Theorem 4.12 (L2-boundedness). Let a ∈ Σ0 be a symbol of order 0. Then Op(a) extends to a bounded operator on L2(Rθ,θ′). Proof. By definition of Σ0, a and all its derivatives Dα L∞(Rd × Rd, Rθ,θ′) and for any α, β, xDβ ξ (σθ,θ′(a))k=k σθ,θ′(Dα k Dα x Dβ ξ (a))k x Dβ ξ (a) are in Rθ,θ′. Then σθ ⊗ σθ′(a) ∈ are bounded. Thus σθ,θ′(a) is a Rθ,θ′-valued symbol with all derivatives bounded. Then by Theorem 4.11, we know op0 ⊗ id(σθ,θ′(a)) is a bounded element in B(L2(Rd))⊗Rθ,θ′. By diagram chasing, k Op(a)k=k VθOp(a)V ∗ θ kB(L2(Rθ))⊗R and the norm estimates follows from Theorem 4.11. θ′≤k op(cid:16)σθ ⊗ σ′ θ(a)(cid:17)kL(L2(Rd,R θ,θ′ )) (cid:3) =α1 =α1 1 xα2 α1 xα2 (2π)dZR2d (2π)dZR2d ξ(cid:16) 1 (2π)dZR2d ξ(cid:16) 1 (2π)dZR2d xα2 1 α2 η−ξ(a)α1 α2 η(a)α1 y−x(b)ei(η−ξ)·(x−y)dηdy(cid:17) y(b)e−iηydηdy(cid:17) = σθ,θ′(c) 30 LI GAO, MARIUS JUNGE, AND EDWARD MCDONALD We now discuss the composition formula. Let us first identify the formula by a heuristic argument. Given two classical operator valued symbol a, b ∈ C ∞(Rd × Rd,A), the composi- tion symbol in the usual Euclidean case is c(x, ξ) = a(x, η)b(y, ξ)ei(η−ξ)·(x−y)dηdy. 1 (2π)dZR2d Given symbols a, b affiliated to Rθ,θ′, the co-multiplication σθ,θ′ gives us operator-valued symbol σθ,θ′(a)(x, ξ) = α1 xα2 ξ(a) , σθ,θ′(b)(x, ξ) = α1 xα2 ξ(b) . The operator-valued composition symbol is C(x, ξ) = η(a)α1 yα2 ξ(b)ei(η−ξ)·(x−y)dηdy where c is a Mθ,θ′-valued singular integral, α2 η(a)α1 c = y(b)e−iη·ydηdy . We first justify this singular integral and prove its formal series of the following definition. Definition 4.13. Let mj, j ≥ 0 be a decreasing sequence of real numbers and aj ∈ Σmj . We write a m0 order symbol a ∼Pj≥0 aj if for any N, a −PN ≤mj The proof adapts the argument for the classical case by Stein [Ste16] to the operator- aj ∈ ΣN . valued setting. Theorem 4.14 (Composition formula). Let a ∈ Σm and b ∈ Σn. Then there exists a symbol c ∈ Σm+n such that Op(c) = Op(a)Op(b) and iα α! ξ (a)Dα x (b) . Dα Proof. Let φ be a positive function on Rd such that φ(x) = 1 for x ≤ 1 and φ(x) = 0 for x > 2. Write c = lim ǫ→0 η(a)bǫ(y)e−iη·ydηdy , where for each ǫ, bǫ(y) = φ(ǫy)α2 y(b) is compactly supported. This is a Bochner integral, because the integrand function (η, y) 7→ α2 η(a)bǫ(y)e−iη·y is smooth in the Frechet space Σm+n by Proposition 4.4. We first prove that the above integral converges in Σm+n and admit the series expansion. For the compactly supported bǫ ∈ C(Rd, Σn), the Fourier transform with value in the Frechet space Σn is well-defined, c ∼Xα (2π)dZ α2 1 bǫ(η) =Z bǫ(y)e−iyηdy . QUANTUM EUCLIDEAN SPACES WITH NONCOMMUTATIVE DERIVATIVES 31 Note that for any compactly supported b,Z b(y)e−iηydηdy = (2π)db(0). Then for any β, Z ηβbǫ(η)dη = (−1)βZ bǫ(y)Dβ = Xβ1+β2=β(cid:18) β = (2π)d Xβ1+β2=β(cid:18) β y(e−iyη)dydη =Z Dβ β1, β2(cid:19)Z ǫβ1(Dβ1φ)(ǫy)α1 β1, β2(cid:19)ǫβ1(Dβ1φ)(0)Dβ2 ξ (Z φǫ(y)αy(b)e−iyηdy) x Dγ Dβ x Dγ ξ (bǫ(η)) =Dβ We also have y(φ(ǫy)α1 y(b))e−iyηdydη y(Dβ2 x b)e−iyηdydη x b = (2π)dDβ x b By Proposition 4.4, we use Taylor expansion with value in the Frechet space Σm, =Z φǫ(y)αy(Dβ x Dγ ξ b)e−iyηdy = \ Dβ x Dγ ξ b ǫ (η) . η(a)bǫ(y)e−iη·ydηdy = η(a)bǫ(η)dη 1 (2π)dZ α2 ηβZ 1 0 αtη(Dβ ξ a)(1 − t)N dt . (4.7) (4.8) We write c = c1 + c2 with c1 = 1 (2π)dZ α2 αη(a) = Xβ≤N iβ(Dβ ξ a)ηβ β! + (N + 1) Xβ=N +1 iβ β! Using the calculation (4.7), the first part leads to 1 (2π)dZ Xβ≤N Dβ ξ a β! ηβbǫ(η)dη = Xβ≤N iβ β! Dβ ξ aDβ x b which gives the leading terms. For the second term in (4.8), we have β = N + 1 and 0 0 α2 tη(Dα kZ 1 ξ a)(1 − t)N dthξi−m+N +1k ≤Z 1 tη(cid:0)Dβ ξ (a)hξi−m+N +1(cid:1)k · khξ + tηim−N −1hξi−m+N +1k dt (1 − t)N k α2 ≤Z 1 (1 − t)N k Dβ ξ (a)hξi−m+N +1k · khξ + tηim−N −1hξi−m+N +1k dt .Z 1 (1 − t)N (thηi)⌈−m+N +1⌉dt ≤ AN,mhηi⌈−m+N +1⌉ . 0 0 Here AN,m is some positive constant only depends on N, m, and ⌈r⌉ denote the smallest even integer greater than r. On the other hand for any β, bǫ(η)ηβ = Xβ1+β2=β β! β1!β2!Z Dβ1 y φǫ(y)α2 y(Dβ2 x (b))e−iyηdy 32 LI GAO, MARIUS JUNGE, AND EDWARD MCDONALD For each term Here we used the assumption that b, Dβ2 supported function of y, we have for any positive integer l, y(Dβ2 ≤Dβ1 y φǫ(y)α1 khξim−N −1Dβ1 y φǫ(y) · k α1 y(cid:0)hξim−N −1Dβ2 x (b))hξi−n−m+N +1k x (b)hξi−n−m+N +1(cid:1)k khξim−N −1bǫ(η)hξi−n−m+N +1k≤ Bn,m,N (1 + η−l) , x (b) ∈ Σn. Because Dβ1 y (φǫ(y)) is a compactly where Bl,n,m,N is a constant depending on (l, n, m, N) and ǫ. Thus, by choosing large enough l, kZRd(cid:16)Z 1 0 αtη(Dβ ξ a)(1 − t)N dt(cid:17)ηβbǫ(η)dηhξi−m−n+N +1k.Z hηi⌈m−N −1⌉(1 + η−l)dη < ∞ . Similar argument applies for derivatives, Therefore we obtain that Dγ1 x Dγ2 0 ξ (cid:16)ZRd(cid:0)Z 1 c1 = Xβ≤N tη(Dβ α2 ξ a)(η)(1 − t)N dt(cid:1)ηβbǫ(η)dη(cid:17) (i)−β β! Dβ ξ aDβ x b + c3 where c3 is a remainder term in Σn+m−N −1. Now take ǫ′ < ǫ and b2(y) := bǫ′(y) − bǫ(y) = (φ(ǫ′y) − φ(ǫy))αy(b) which is supported on 1/ǫ < y < 2/ǫ′. Note that in above argument, we actually show that Then for each j, we can use integration by parts the singular integralR αη(a)b(y)eiη·ydηdy converges absolutely if b is compactly supported. Z αη(a)yjy−2b2(y)eiη·ydηdy =Z αη(a)y−2b2(y)Dηj eiη·ydηdy =Z Dηj (αη)(a)y−2b2(y)eiη·ydηdy =Z αη(Dξj a)y−2b2(y)eiη·ydηdy . Here we used the property Dηj (αη)(a) = αη(Dξj a). Denote ∆η =Pj D2 ∆y =Pj D2 ξj and η(∆ξa), using the standard trick in singular integral, ηj , ∆ξ =Pj D2 yj . Because ∆η(α1 η(a)) = α1 Z αη(a)b2(y)eiη·ydηdy =Z αη(∆m1 =Z αη(∆m1 ξ a)(1 + ∆y)m2(y−2m1b2(y))hηi−2m2e−iηydηdy ξ a)y−2m1b2(y)e−iηydηdy Here y−2m1b2(y) has no singularity because b2 is supported away from y = 0. Because a ∈ Σm, b ∈ Σn, ∆m1 ξ (a) ∈ Σm−2m1 , (1 + ∆y)m2(y−2m1b2(y)) ∈ Σn . QUANTUM EUCLIDEAN SPACES WITH NONCOMMUTATIVE DERIVATIVES 33 We have ξ a)hξi−m+2m1 k≤ Am,m1hηi⌈−m+2m1⌉ , k αη(∆m1 khξim−2m1(1 + ∆y)m2(y−2m1b2(y))hξi−m+2m1−nk≤ Bm,m1,n(1 + y)−2m1χ{ 1 ǫ′ } (4.9) for some constants Am,m1 and Bm,m1,n. We can choose m1, m2 large enough such that 2m1 > N + 1 and then the integral kZ αη(a)b2(y)e−iηydηdy · hξi−m−n+N +1k≤Z η⌈−m+2m1⌉hηi−2m2(1 + y)−2m1dηdy < ∞ ǫ <y< 2 converges absolutely. The argument for the derivatives are similar. Hence Z αη(a)b2(y)e−iηydηdy ∈ Σn+m−N −1 , which is of lower order of the leading terms. Note that the above estimates is uniform for 0 < ǫ′, ǫ < 1 and when ǫ′, ǫ → 0, the norm estimates (4.9) goes to 0. So when ǫ → 0, the remainder c2 converges to 0 in Σn+m−N −1. This implies c = lim η(a)φ(ǫy)α1 y(b)eiηydηdy ǫ→0Z α2 y(b)eiηydηdy. We now show that for any λθ,θ′(F ) ∈ Sθ,θ′, converges in Σm+n. Write cǫ =Z α2 η(a)φ(ǫy)α1 Op(a)Op(b)λθ,θ′(F ) = lim ǫ→0 Op(cǫ)λθ,θ′(F ) = Op(c)λθ,θ′(F ) Indeed, since the integral in cǫ converges absolutely Op(cǫ)λθ,θ′(F ) =Z α2 η(a)α1 η+η1(a)α1 y(b)e−iηydηdy(cid:17) F (η1, y1)λθ,θ′(η1, y1)dη1dy1 η(b) F (η1, y1)λθ,θ′(η1, y1)dη1dy1dηdy yα2 Let φ be the Fourier transform of φ. Z φ(ǫy)e−iξyα1 y(λθ,θ′(G))dy =Z φ(ǫy)e−i(ξ−η1)y G(η1, y1)λθ,θ′(η1, y1)dydy1dη1 ) G(η1, y1)λθ,θ′(η1, y1)dy1dη1 =Z 1 ǫd ξ − η1 φ( ǫ η1(cid:16)Z φ(ǫy)α2 =Z φ(ǫy)e−iηyα2 =Z φ(ǫy)e−i(ξ−η1)yα2 =Z φ(ǫy)e−iξyα2 =Z φ(ǫy)e−iξyα2 ǫ→0Z φ(ǫy)e−iξyα2 ξ(a)α1 lim ξ(a)α1 ξ(a)α1 ξ(a)α1 yα2 η1(b) F (η1, y1)λθ,θ′(η1, y1)dη1dy1dξdy η1(b) F (η1, y1)λθ,θ′(η1, y1)dη1dy1(cid:17)dξdy y(cid:16)Z α2 y(cid:16)Op(b)λθ,θ′(F )(cid:17)dξdy . y(cid:16)λθ,θ′(G)(cid:17)dηdy = Op(a)λθ,θ′(G) . Then it suffices to show that for any λθ,θ′(G), 34 Here 1 ǫd φ(· ǫ LI GAO, MARIUS JUNGE, AND EDWARD MCDONALD ) approximates the delta function, Z φ(ǫy)e−iξyα2 ξ(cid:16)Op(a)α2 ξ(a)α1 y(cid:16)λθ,θ′(G)(cid:17)dηdy =Z 1 =Z 1 φ( ǫd ǫd φ( ξ ǫ )Op(α2 ξa)λθ,θ′(G)dξ ξ ǫ )α2 ξ(cid:16)Op(a)α2 −ξλθ,θ′(G)(cid:17)dξ . −ξλθ,θ′(G)(cid:17) is continuous in Sθ,θ′. When ǫ → 0, the above integral Since ξ → α2 converges to Op(a)λθ,θ′(G) in Sθ,θ′. 4.4. Integrability and trace formula. In the rest of this section we discuss the integra- bility of ΨDOs whose symbols is integrable in the first component Rθ. Definition 4.15 (Tame symbols). An element a ∈ Mθ,θ′ is a tame symbol of order m if there exists a r > d such that for any α, β and γ, (cid:3) hxirDα x Dβ ξ (a)hξiβ−m tame := ∩rΣr extends to bounded element in Rθ,θ′. We write Σm and Σ−∞ Proposition 4.16. A symbol a ∈ Σm α, β, Dα tame. x Dβ ξ (a) ∈ O−r,m−β. Moreover, if b ∈ Σn, ab, ba ∈ Σn+m tame. tame the set of all tame symbols of order m tame if and only if there exists r > d such that for all Proof. This is a direct consequence of Theorem 4.3. (cid:3) Lemma 4.17. Let a ∈ L2(Rθ) and b ∈ L2(Rθ′). Then ab ∈ L2(RΘ) and k ab kL2(RΘ)=k akL2(Rθ)k bkL2(R Proof. It can be verified from the definition of trΘ that for f ∈ Sθ, g ∈ Sθ′ θ′ ). trΘ(λθ(f )λθ′(g)) = trθ(λθ(f ))trθ′(λθ(g)) . Then we have k λθ(f )λθ′(g)k2 L2(RΘ)=trΘ(λθ′(g)∗λθ(f )∗λθ(f )λθ′(g)) = trΘ(λθ(f )∗λθ(f )λθ′(g)λθ′(g)∗) =trθ(λθ(f )∗λθ(f ))trθ′(λθ′(g)λθ′(g)∗) = k λθ(f )k2 L2(Rθ)k λθ′(g)k2 L2(R θ′ ) The assertion for general a ∈ L2(Rθ), b ∈ L2(Rθ′) follows from density. Corollary 4.18. Let a ∈ Sm tame. Then i) Op(a) ∈ L2(RΘ) if m < − d 2 ; ii) Op(a) ∈ L1(RΘ) if m < −d. (cid:3) Proof. We know from the algebraic property that Op(λθ(f1) ⊗ λθ′(f2)) = λθ(f1)λθ′(f2) for f1, f2 ∈ S(Rd). The Op is a L2-isometry and trace preserving on Sθ,θ′. Let a ∈ Σm tame. Then for some r > d, Op(a) =hxi−rhξimhξi−mhxirOp(a) = hxi−rhξimhξi−mOp(hxira) =(cid:16)hxi−rhξim(cid:17)(cid:16)hξi−mOp(hxira)(cid:17) . For m < −d, choose n = m 2 , k Op(a)k2≤khxi−rhξimk2khξi−mOp(hxira)k∞ Op(a) =(cid:16)hxinhξin(cid:17)(cid:16)hξi−nOp(hxi−na)(cid:17) . QUANTUM EUCLIDEAN SPACES WITH NONCOMMUTATIVE DERIVATIVES 35 By symbol calculus, hξi−mOp(hxira) is a ΨDO of order 0 hence in RΘ. For m < −d/2, khξimkL2(R θ′ )< ∞ and khxi−rkL2(Rθ)< ∞. Then hxi−rhξim ∈ L2(RΘ) and hξi−nOp(hxi−na) is a tame ΨDO of order less than d/2 hence in L2(RΘ) and hxi−nhξi−n is also in L2(RΘ) by the discussion in i). (cid:3) We end this section with the trace formula. Proposition 4.19. Suppose a symbol a ∈ L1(Rθ,θ′) and its operator Op(a) ∈ L1(RΘ). Then τΘ(Op(a)) = τθ,θ′(a) . Proof. Using the definition of Op(a), τΘ(Op(a)λΘ(F )) =τθ,θ′(cid:16)ZR2d F (η, y)α2 η(a)λθ,θ′(η, y)dηdy(cid:17) η(a)λθ,θ′(η, y))(cid:17)dηdy F (η, y)(cid:16)τθ,θ′(α2 −η(λθ,θ′(η, y))(cid:17)dηdy F (η, y)τθ,θ′(cid:16)aα2 F (η, y)e−iηy(cid:16)τθ,θ′(aλθ,θ′(η, y))(cid:17)dηdy =ZR2d =ZR2d =ZR2d =τθ,θ′(aλθ,θ′(F ′)) , where F ′ has the Fourier transform F ′(η, y) = F (η, y)e−iηy. Here we use the Fubini theorem because a ∈ L1(Rθ,θ′). Let Fn ∈ S(R2d) be a sequence of Schwartz function in Proposition 3.7. Then λΘ(Fn) (resp. λθ,θ′(Fn)) is an approximation of identity in L1(RΘ) (resp. L1(Rθ,θ′)). n(η, y) = Fn(η, y)e−iηy. Note that k Fn k1= 1 and Fn is Take F ′ supported in (η, y) ≤ 1 n ∈ S(R2d) such that F ′ n . When n → 1, k λθ,θ′(Fn) − λθ,θ′(F ′ n)k∞≤k F ′ Therefore, n − Fnk1=ZR2d Fn(η, y)1 − e−iηydηdy → 0 . τΘ(Op(a)) = lim n→∞ τΘ(Op(a)λΘ(Fn)) = lim n→∞ τθ,θ′(aλθ,θ′(F ′ n)) = lim n→∞ τθ,θ′(aλθ,θ′(Fn)) =τθ,θ′(a) . (cid:3) 5. Local Index formula In this section we discuss the spectral triple structure on Rθ equipped with noncom- muting partial derivatives. We first recall the definitions of semi-finite spectral triple from [CGRS14]. We shall show that the non-commuting derivatives in Section 4 gives a natural example of semi-finite spectral triple. The main results of this chapter is a simplified index formula and we calculate it for the Bott projector as an example. 36 LI GAO, MARIUS JUNGE, AND EDWARD MCDONALD 5.1. Semifinite spectral triple. Let N be a von Neumann algebra equipped with a normal faithful semi-finite trace τ . The τ -compact operators K(N , τ ) is the norm completion of L1(N , τ ) ∩ N in N . In our case K(Rθ, τθ) = Eθ. The following definitions of semi-finite spectral triple is from [CGRS14]. Definition 5.1. A semi-finite spectral triple (A, H, D), relative to a semi-finite tracial von Neumann algebra (N , τ ), is by given a Hilbert space H, a ∗-subalgebra A of N acting on H, and a densely defined unbounded self-adjoint operator D affiliated to N such that i) a · dom D ⊂ dom D for all a ∈ A, so that da := [D, a] is densely defined. Moreover, ii) a(1 + D2)−1/2 ∈ K(N , τ ). da extends to a bounded operator in N for all a ∈ A; (A, H, D) is even if there is an operator γ ∈ N such that for all a ∈ A, γ = γ∗, γ2 = 1, γa = aγ, and Dγ + γD = 0. (A, H, D) is finitely summable if there exists s > 0 such that a(1 + D2)− s all a ∈ A. Then p = inf{s > 0 for all a ∈ A, a(1 + D2)− s 2 ∈ L1(N , τ )} is called the spectral dimension of (A, H, D). 2 ∈ L1(N , τ ) for The subalgebra A plays the role of smooth functions. The main difference to the compact case is the condition ii), which simplifies to that (1 + D2)−1/2 is compact. The semi-finiteness allow locally compact space equipped with non-finite measure. We recall the following suf- ficient condition for the smooth summability of a semi-finite spectral triple and refer to [CGRS14] for the detailed definition. Proposition 5.2 (Proposition 2.21. of [CGRS14]). Let (A, H, D) be a spectral triple of spectral dimension p relative to (N , τ ). If for all a ∈ A ∪ [D,A], k ∈ N+ and s > p, (1 + D2)− s 4 Lk(a)(1 + D2)− s 4 ∈ L1(N , τ ), then (A, H, D) is smoothly summable. Here L(T ) := (1 + D2)− 1 L(Lk−1(T )). 2 [D2, T ] and Lk(T ) = Quantum Euclidean space Rθ equipped with its natural partial derivative Dj's were studied as the prototypical example of semi-finite spectral triple in [GGBI+04, CGRS14]. The rest of this subsection is to show that the non-commuting derivatives also gives a semi- finite spectral triple structure of Rθ. First, we choose the smooth subalgebra A to be the noncommutative Sobolev space W 1,∞(Rθ) = {a Dα(a) ∈ L1(Rθ) for all α} . 0 (Rd) by Sobolev embedding theorem (c.f. In the classical case W 1,∞(Rd) ⊂ C ∞ The next lemma is a weaker analog on Rθ. Lemma 5.3. If Dα(a) ∈ L1(Rθ) for all α, then Dα(a) ∈ Lp(Rθ) for all 1 ≤ p ≤ ∞ and α. In particular, the unitalization W 1,∞(Rθ)∼ := (W 1,∞(Rθ) + C) is a dense ∗-subalgebra of E∼ closed under holomorphic function calculus. xj . For λθ(f ) ∈ Sθ, [Gra09]). θ Proof. Denote ∆ =Pj D2 (1 + ∆)λθ(f ) = λθ((1 + ∆)f ) =Z hηi2 f (η)λθ(η)dη . QUANTUM EUCLIDEAN SPACES WITH NONCOMMUTATIVE DERIVATIVES 37 Choose a integer 2n > d, we have (1 + ∆)−n : L2(Rθ) → L∞(Rθ) is bounded because k (1 + ∆)−nλθ(f )k= kZ hηi−n f (η)λθ(η)dηk≤khηi−n f k1 ≤ khηi−nk2k f k2=khηi−nk2k λθ(f )k2 . By duality, we also have that (1 + ∆)−n : L1(Rθ) → L2(Rθ) is bounded. Indeed, for any λθ(f ), λθ(g) ∈ Sθ, hλθ(g), (1 + ∆)−nλθ(f )iτθ = h(1 + ∆)−nλθ(g), λθ(f )iτθ ≤k (1 + ∆)−nλθ(g)k∞k λθ(f )k1≤ C k λθ(g)k2k λθ(f )k1 Here we have used the fact (1 + ∆)−n is self-adjoint on Sθ. Thus we have that (1 + ∆)−n : L1(Rθ) → L∞(Rθ) is continuous. If Dα(a) ∈ L1(Rθ) for all α ≤ 2n, then (1 + ∆)n(a) ∈ L1(Rd) and hence a ∈ L∞(Rθ). Therefore W 1,∞(Rθ) is closed under product hence a sub- It is dense because Sθ ⊂ W 1,∞(Rθ). To show W 1,∞(Rθ) is closed under algebra of Eθ. holomorphic calculus, it suffices to consider the resolvent (λ − a)−1 for λ /∈ Spec(a). Indeed, (λ − a)−1 is bounded and λ−1 − (λ − a)−1 = λ−1(cid:0)(λ − a) − λ(cid:1)(λ − a)−1 = −λ−1a(λ − a)−1 ∈ L1(Rθ) . For the derivatives, For higher order derivatives Dα, we use induction and Leibniz rule [Dj, (λ − a)−1] = (λ − a)−1[Dj, a](λ − a)−1 ∈ L1 Dα((λ − a)−1) =Dα((λ − a)−1(λ − a)(λ − a)−1) = Xα1+α2+α3=α α! α1!α2!α3! Dα1((λ − a)−1)Dα2(λ − a)Dα3((λ − a)−1) . (cid:3) The above lemma implies that the inclusion W 1,∞(Rθ) ⊂ Eθ induces K-groups isomor- θ or θ ) can be approximated using projections (resp. unitary) in W 1,∞(Rθ)∼. To verify the In particular, every projection (resp. unitary) in E∼ phism (c.f. page 292 of [Con]). Mn(E∼ finite and smooth summability, we need the following lemma. Lemma 5.4. Let a ∈ W 1,∞(Rθ). Then hξi− r Proof. We write a as a = a1a2 with a1, a2 ∈ L2(Rθ). Then 2 a1)(a2hξi− r 2 = (hξi− r 2 ahξi− r 2 ahξi− r hξi− r 2 ) ∈ L1(RΘ) 2 , ahξi−r ∈ L1(RΘ) if r > d. because 2 a1kL2(Rθ)=khξi− r 2 [a,hξi− r khξi− r Note that hξi− r n such that 2n > r 2 [a,hξi− r hξi− r 2 kL2(R θ′ )k a2kL2(Rθ) . 2 ] ∈ L1(RΘ), choose 2 kL2(Rθ)=khξi− r θ′ )k a1kL2(Rθ) , k a2hξi− r 2 kL2(R 2 ] = hξi− r 2 ahξi− r 2Z ∞ 2Z ∞ t−s(t + hξi2n)−1(cid:16)hξi− r 2 and write s = r 2 ] =Cshξi− r =Cshξi− r = CsZ ∞ 0 0 2 − ahξi−r. To show hξi− r 2 [a,hξi− r 4n . By operator integral, t−s[a, (t + hξi2n)−1]dt t−s(t + hξi2n)−1[a, t + hξi2n](t + hξi2n)−1dt 2 [a,hξi2n]hξi−2n(cid:17)hξi2n(t + hξi2n)−1dt 0 38 LI GAO, MARIUS JUNGE, AND EDWARD MCDONALD Here Cs is some positive constant depending on s. Since [a,hξi2n] is a linear combination of a's derivatives, we know Then the integral converges in L1-norm, hξi− r 2 [a,hξi2n]hξi−2n ∈ L1(RΘ) . khξi− r .Z ∞ .Z ∞ 0 0 2 ]k1 2 [a,hξi− r t−s k (t + hξi2n)−1k∞khξi− r t−s(t + 1)−1dt < ∞ . 2 [a,hξi2n]hξi−2nk1khξi2n(t + hξi2n)−1k∞ dt (cid:3) d Recall that the Clifford algebra Cld is generated by d self-adjoint operators c1,· · · , cd satisfying the anti-commutation relation cjck + ckcj = 2δj,k. For d = 2n even, Cld is isomor- phic to the N × N matrix algebra MN with N = 2n. For d = 2n + 1 odd, Cld is isomorphic to M2n ⊕ M2n ⊂ MN with N = 2n+1. When d even, Cld is Z2 graded with the parity element γ = (−i) Theorem 5.5. (W ∞,1(Rθ)⊗ MN , L2(RΘ)⊗ CN ,Pj ξj ⊗ cj) relative to (RΘ ⊗ MN , τΘ⊗ tr) is a smooth summable semi-finite spectral triple with spectral dimension d. Moreover it is even if d = 2n is even, and γ = (−i) Proof. Note that 2 c1 · · · cd. 2 c1 · · · cd. d D2 =Xj,k ξjξk ⊗ cjck =Xj ξ2 j − i 2Xj,k θ′ j,kcjck . 2Pj,k θ′ Denote ω = i j,kcjck. Then 1 + D2 = hξi2 − ω. Since ω ∈ MN commutes with RΘ, to verify summability it is equivalent to replace 1 + D2 by hξi2. By Lemma 5.4, we know the spectral dimension is less than d. On the other hand, if ahξi−r ∈ L1(RΘ), k ahξi− r 2 k2 2≤k ahξi−da∗k1≤k a∗k∞k ahξi−dk1< ∞ which implies r > d. For smooth summability, we know [hξi2, a] ∈ L1(Rθ) and by Lemma 5.4 again, if s > d. The arguments for Lk(a) are similar. (1 + D2)− s 2 L(a)(1 + D2)− s 2 ∈ L1(RΘ) 5.2. Local Index formula. We briefly recall the local index formula for the even case and refer to [CM95, CGRS14] for detailed information. Let (A, H, D) be an even spectral triple relative to (N , τ ) and γ is the parity element. Denote H+ = γ+1 2 H For µ > 0, define Dµ =(cid:20) D µ µ D (cid:21) on H ⊕ H. Write Fµ = DµDµ−1 and 2 H and H− = 1−γ 1 + γ (Fµ)+ = ( 2 ⊗ I2)Fµ : H+ ⊕ H+ → H− ⊕ H− . Here and in the following In represents the n-dimensional identity matrix. For a projection e ∈ Mn(A∼), denote e = (cid:20) e 0 1e (cid:21) ∈ M2n(A∼) where 1e ∈ Mn(C) is the rank element of e. Following [CGRS14, Definition 2.12 and Proposition 2.13], the numerical index pairing between the K0(A) element [e] − [1e] and the even spectral triple (A, H, D) is given by 0 h[e] − [1e], (A, H, D)i = indexτ ⊗tr2n(e(Fµ,+ ⊗ In)e) (cid:3) (5.1) QUANTUM EUCLIDEAN SPACES WITH NONCOMMUTATIVE DERIVATIVES 39 Here the numerical index indexτ (F ) = τ (kerF ) − τ (cokerF ) is defined as the trace of kernel subtracting the trace of cokernel. Both quantities are topological invariants under homotopy. The local index formula express the index pairings by the following residue cocycle formulas. Definition 5.6. (A, H, D) has isolated spectral dimension if for all a0,· · · , am ∈ A, the zeta function has an analytic continuation to a deleted neighbourhood of z = 0. ζ(z) = τ (γa0da(k1) 1 · · · da(km) m (1 + D2)−k−m/2−z) Here we introduce the notation da := [D, a] and da(k) := [D2, [D2,· · · [D2 } (A, H, D) be a smoothly summable semifinite spectral triple with spectral dimension d and M be the largest integer in [0, d+1]. Suppose A has isolated spectral dimension. The residue cocycle φm : A⊗m+1 → C is the (m + 1)-linear form given by , da]]. Let {z k-times φ0(a0) =Resz=0z−1τ (γa0(1 + D2)−z) (5.2) φm(a0,· · · , am) = (−1)kα(k) M −mXk=0 k+m/2Xj=0 σk+m/2,jResz=0zj−1τ (γa0da(k1) (5.3) where α(k), σk+m/2,j are the constant defined as follows. For a multi-index k = (k1,· · · , km), (5.4) α(k) = k1!k2!· · · km!/(k1 + 1)(k1 + k2 + 2)· · · (k + m) . m (1 + D2)−k−m/2−z) . · · · da(km) 1 σn,j are the non negative constant given by the equation n−1Yj=0 (z + j) =Xj=1 σn,jzj for (5.5) In particular, α(0) = m! and σn,1 = (n − 1)!. The terms in φm is a linear combination of residue and higher order residue of the zeta function ζ(z) = τ (γa0da(k1) 1 · · · da(km) m (1 + D2)−k−m/2−z) . The isolated spectral dimension condition assumes that these residues are well-defined. Theorem 5.7 (Theorem 3.33 of [CGRS14] (even case)). Let (A, H, D) relative to (N , τ ) be an even smoothly summable semi-finite spectral triple. Suppose that (A, H, D) has isolated spectral dimension. Then the numerical index pairing can be computed by h[e] − [1e], [(A, H, D)]i = φm(Chm(e) − Chm(1e)) , MXm=0,even where for a projection e ∈ Mn(A∼), Ch0(e) = (e) and (e − ) ⊗ e ⊗ · · · ⊗ e ∈ A⊗2k+1 . 1 2 k! Ch2k(e) = (−1)k 2k! 2P θ′ CN ,Pj ξj ⊗ cj). Recall that ω = i We shall now calculate the local index formula for the spectral triple (W ∞,1(Rθ), L2(RΘ)⊗ jkcjck is the analog of curvature form. Let us denote the super trace on Cld as str(a) = tr(γa) and the corresponding super trace on RΘ ⊗ Cld (resp. Rθ ⊗ Cld) as StrΘ = τΘ ⊗ str (resp. Strθ = τθ ⊗ str). 40 LI GAO, MARIUS JUNGE, AND EDWARD MCDONALD Theorem 5.8. Let d be even. The spectral triple (W ∞,1(Rθ), L2(RΘ) ⊗ CN ,Pj ξj ⊗ cj) has isolated spectral dimension. Moreover, a0,· · · , am ∈ W ∞,1(Rθ), ), d−m d 2 π 2 ω (d−m) m! Strθ(a0da1 · · · dam 0, ! 2 if m even if m odd. . φm(a0,· · · , am) = Proof. We first consider m > 0. Let us denote Ψk = a0da(k1) linear combination of residue of the zeta functions at z = 0, ζk(z) = StrΘ(Ψk(1 + D2)−k− m 1 2 −z) . · · · da(km) m . The cocycle φm is a Because a0,· · · , am ∈ W ∞,1(Rθ)∼ and da(kj ) are derivatives of aj, Ψk ∈ W ∞,1(Rθ) ⊗ Cld. Using the same argument of Lemma 5.4, one can obtain that Ψk(1 + D2)−r ∈ L1(RΘ ⊗ MN ) if r > d 2, and hence it suffices to consider the nonzero residue of ζk at z = 0 for m + 2k ≤ d. Applying Cahen -- Mellin integral, we have (5.6) 2. Then ζk(z) is analytic for k + m e−s(1+D2)sk+ m (1 + D2)−k− m 2 + Re z > d 2 +z−1ds . 2 −z = j 1 Γ(k + m 2 + z)Z ∞ 0 For a ∈ W ∞,1(Rθ) and ν ∈ Cld, k (a ⊗ ν)e−s(1+D2)kL1(RΘ⊗MN )≤ e−s k (a ⊗ ν)(1 + D2)−rk1k (1 + D2)re−sD2 k∞ By functional calculus, k (1 + D2)re−sD2 k∞≤( rr sr , 1, if s < r if s ≥ r. Then the integral Z ∞ Re(z) > r > d 0 2. Hence by Fubini Theorem k (a ⊗ ν)e−s(1+D2) kL1(RΘ⊗MN ) sk+ m ζk(z) =Z ∞ 0 StrΘ(Ψke−s(1+D2))sk+m/2+z−1ds 2 +z−1ds converges for k + m 2 + Using the trace formula from Proposition 4.19, StrΘ(Ψke−s(1+D2)) =StrΘ(Ψk(e−s(1+ξ2) ⊗ e−sω)) = trθ′(e−s(1+ξ2))Strθ(Ψkesω) =Xn Strθ(cid:16)Ψk ωn n!(cid:17)π d 2 e−ssn− d 2 h(s) Here we used the calculation in Proposition 2.6 that tr′ θ(e−sξ2 ) = s− d 2 det( iπsθ′ sinh isθ′ ) where 1 2 = s− d 2 π d 2 h(s) , h(s) = det( isθ′ sinh isθ′ ) = Πl j=1 λjs sinh λjs , QUANTUM EUCLIDEAN SPACES WITH NONCOMMUTATIVE DERIVATIVES 41 where iλ1,−iλ1,· · · , iλl,−iλl are the nonzero eigenvalues of θ′. Using L'Hospital's Rule, we know lim s→0 s−1(h(s) − 1) = 0. Then we split the residue into two parts Resz=0ζk(z) =Resz=0StrΘ(Ψk(1 + D2)−m/2−k−z) 1 Γ(m/2 + k + z)Z ∞ 0 =Resz=0 = 1 n! d 2 π Γ(m/2 + k)Xn + Resz=0Z ∞ 0 Strθ(Ψkωn)(cid:16)Resz=0Z ∞ 2 +k+m/2+z−1ds(cid:17) 0 (h(s) − 1)e−ssn− d StrΘ(Ψke−s(1+D2))sk+m/2+z−1ds e−ssn− d 2 +k+m/2+z−1ds Note that for any j1, j2 and j3, [cj1cj2, cj3] = 0 or of order 1. Then [D2, da] = [ξ2 − ω,Pj Dj(a) ⊗ cj] =Pj[ξ2, Dj(a)] ⊗ cj +Pj Dj(a) ⊗ [ω, cj] is of Clifford order 1 and similarly for da(k0). Thus Ψk = a0da(k1) m contains Clifford term of at most order m and Ψkωn contains Clifford elements of order at most m + 2n. Hence the super trace Strθ(Ψkωn) = 0 for 2n + m < d. It suffices to consider the residue for 2n + m ≥ d. On one hand, · · · da(km) 1 Resz=0Z ∞ =Resz=0Z ∞ 0 0 (h(s) − 1)e−ssn− d h(s) − 1 e−ssn− d s 2 sk+m/2+z−1ds 2 +k+m/2+zds = 0 (5.7) because the integral converges absolutely for Re(z) > −1 ≥ −n + d other residue 2 −k− m/2− 1. For the Resz=0Z ∞ 0 e−ssn− d 2 +k+m/2+z−1ds = Resz=0Γ(n − d 2 + k + m/2 + z) is zero if n− d and it is a simple pole. Then φm vanishes for odd m and for even m ≥ 2, 2 +k + m/2 ≥ 0. Therefore, the only nonzero residue is at 2n+ m−d = k = 0 φm(a0,· · · , am) = α(0)σ m 2 ,1Resz=0ζ0(z) = Γ(m/2) d 2 π m! Γ(m/2) Resz=0Γ(z)Strθ(Ψ0 ω(d−m)/2 d−m 2 ! ) = d 2 π m! Strθ(a0da1 · · · dam ω(d−m)/2 d−m 2 ! ) . 42 LI GAO, MARIUS JUNGE, AND EDWARD MCDONALD For m = 0, we follow the same argument φ0(a0) =Resz=0z−1StrΘ(a0(1 + D2)−z) )str(esω)e−ssz−1ds str(ωn) n! h(s)e−sπ d 2 sn− d 2 +z−1ds e−ssn− d 2 +zds StrΘ(a0e−s(1+D2))sz−1ds =Resz=0 =Resz=0z−1 1 0 0 1 trθ(a0)trθ′(e−sξ2 Γ(z)Z ∞ zΓ(z)Z ∞ Γ(z + 1)Z ∞ 0 Xn=0 (cid:16)Resz=0Z ∞ (h(s) − 1)e−ssn− d str(ωn) n! 1 0 =trθ(a0)Resz=0 =π d 2 trθ(a0)Xn=0 + Resz=0Z ∞ 0 2 +z−1ds(cid:17) 2. For n ≥ d (h(s) − 1)e−ssn− d h(s) − 1 e−ssn− d s 2 +z−1ds 2 +zds = 0 Resz=0Z ∞ =Resz=0Z ∞ 0 0 The super trace str(ωn) is non-zero if n < d 2, the second residue (5.8) (cid:3) because the integral converges for integral converges absolutely for Re(z) > −1 ≥ n− d The first residue 2 − 1. Resz=0Z ∞ 0 e−ssn− d 2 +z−1ds = Resz=0Γ(n − 2 ≤ 0. Therefore, φ0(a0) = πd/2Strθ(a0 d 2 + z) ωd/2 (d/2)! ). is non-zero only if n − d For compact Spin manifolds, the isolated spectral dimension condition always holds and the only nonzero residues when j = 0 and k = 0. This simplification recovers the Atiyah- Singer index theorem for Spin Dirac operator (see [CM95], [Hig03] and [Pon03]). The above theorem gives a simplification of the cocycle formula for (W ∞,1(Rθ), L2(RΘ) ⊗ CN ,X ξj ⊗ cj) to the terms only for k = j = 0. As a consequence, the local index formula for Rθ simplifies too. We can see the term ω plays the role of the curvature form. Corollary 5.9. For any projection e ∈ Mn(W ∞,1(Rθ)) and with Fµ,+ defined as in (5.1), Index(e(Fµ,+ ⊗ idn)e) = π d 2 Strθ(cid:16)(e − 1e) ωn n! + dXm=2,even 1 m! e(de)m ωd−m (d − m)!(cid:17) . 5.3. A concrete example for d = 2. We shall now calculate a concrete example in di- In the classical case, a canonical generator for K0(C0(R2)) is the Bott mension d = 2. projector eB(x, y) = 1 1 + x2 + y2(cid:20) 1 x + iy x2 + y2 (cid:21) ∈ M2(C0(R2)∼) , 1eB =(cid:20) 0 0 0 1 (cid:21) ∈ M2(C) . x − iy QUANTUM EUCLIDEAN SPACES WITH NONCOMMUTATIVE DERIVATIVES 43 Now let θ be a real number and Rθ is the Moyal plane generated by two self-adjoint element x, y with [x, y] = −iθ. We consider an analog of Bott projection for Rθ. Write z = x+iy, R = z (cid:21). Then e := u(cid:20) R 0 0 0 (cid:21) u∗ =(cid:20) R Rz∗ zR zRz∗ (cid:21) is a projection because (1+z∗z)−1 and u =(cid:20) 1 u∗Ru = 1. The only drawback of e is that it does not belongs to M2(W ∞,1(Rθ)∼). Indeed, by Proposition 2.6 and Theorem 3.2, we know that R, zR, zRz∗ /∈ L1(Rθ). Nevertheless, dede and id ⊗ tr2(e − 1e) = R + zRz∗ − 1 do belong to L1 so that the cocycle formula in Corollary 5.9 are well defined. The next lemma shows that by approximation the cocycle formula remains valid for e. Lemma 5.10. There exists a sequence of projection en ∈ M2(W ∞,1(Rθ)∼) such that 1en = 1e and limn→∞ k en − ek∞= 0, limn→∞ k id ⊗ tr2(en − e)k1= 0. As a consequence, h[e] − [1e], (W ∞,1(Rθ), L2(RΘ) ⊗ CN ,X ξj ⊗ cj)i = πStrθ((e − 1e)ω) + πStrθ(edede) Proof. Let λθ(φn) be the approximation identity in Propsition 3.7. Define en := (λθ(φn) ⊗ 1)(e − 1e) + 1e ∈ M2(W ∞,1(Rθ)) . Because e − 1e ∈ Eθ and id ⊗ tr2(e − 1e) ∈ L1(Rθ), we have k en − ek∞=k (λθ(φn) ⊗ 1)(e − 1e) − (e − 1e)k∞→ 0 , k id ⊗ tr2(en − 1e) − id ⊗ tr2(e − 1e)k1→ 0 . Using holomorphic functional calculus, we can made projections en ∈ M2(W ∞,1(Rθ)) from en with satisfies the same limits above. It is known that if two projections e, f satisfy that k e − f k< 1 then e is homotopic to f hence [e] = [f ] (see e.g. [RLLL00]). Then by the homotopy invariance of index pairing, we know for n large enough h[e] − [1e], (A, H, D)i = h[en] − [1en], (A, H, D)i = φ0(en − 1en) + φ2(en − = πStrθ(en − 1enω) + πStrθ((en − )denden)) . 1 2 1 2 , en, en) Taking the limit n → ∞, lim n→∞ Strθ((en − 1en)ω) = Strθ((e − 1e)ω) . For the second term, we first note that Strθ(denden) = Strθ(−denden) = 0 because denγ = −γden. For the same reason, we have the cyclicity that Strθ(edende) = Strθ(d(een)de) − Strθ(d(e)ende) = Strθ(endede), Strθ(endeden) = Strθ(d(ene)den) − Strθ(d(en)eden) = Strθ(edend(en)) . Therefore, Strθ(edede) − τθ ⊗ Strθ(endenden) =Strθ(edede − endede) + Strθ(endede − endende) + Strθ(endende − endenden) =Strθ(edede − endede) + Strθ(ededen − endeen) + Strθ(edendenendenden) =Strθ(cid:0)(e − en)dede(cid:1) + Strθ(cid:0)(e − en)deden(cid:1) + Strθ(cid:0)(e − en)denden(cid:1), All the three terms above converges to 0, since k e − en k∞→ 0 and dede, deden, denden are in M2(L1(Rθ)). (cid:3) 44 LI GAO, MARIUS JUNGE, AND EDWARD MCDONALD Theorem 5.11. For any θ, θ′, h[e] − [1e], (W ∞,1(Rθ), L2(RΘ) ⊗ CN ,X ξj ⊗ cj)i = 4π2(1 − θθ′) . In particular, [e] is a generator of K0(Eθ) = Z. Proof. The super trace Strθ(edede) is of eight terms zR zRz∗ (cid:21)(cid:20) dR d(zR) d(zRz∗) (cid:21)(cid:20) dR Strθ(edede) = Strθ ⊗ tr2(cid:16)(cid:20) R Rz∗ d(zR) d(zRz∗) (cid:21)(cid:17) =Strθ(cid:16)Rd(R)d(R) + Rd(Rz∗)d(zR) + Rz∗d(zR)d(R) + Rz∗d(zRz∗)d(zR) + zRd(R)d(Rz∗) + zRd(Rz∗)d(zRz∗) + zRz∗d(zR)d(Rz∗) + zRz∗d(zRz∗)d(zRz∗)(cid:17) . We will repeatedly use Leibniz rule and cyclicity of trace (in the strong sense [BK90, Theorem 17]) that d(Rz∗) d(Rz∗) d(a1a2) = (da1)a2 + a1da2 , Strθ(da1(da2)a3) = Strθ(a3da1da2) Denote τ = Strθ in short. For the first and fifth term, τ(cid:16)Rd(R)d(R) + zRd(R)d(Rz∗)(cid:17) = τ(cid:16)d(R)d(R)R + d(R)d(Rz∗)zR(cid:17) = τ(cid:16)d(R)d(R)R + d(R)d(R)z∗zR + d(R)Rd(z∗)zR(cid:17) = τ(cid:16)d(R)d(R)R + d(R)d(R)(1 − R) + d(R)Rd(z∗)zR(cid:17) = τ(cid:16)d(R)d(R) + d(R)Rd(z∗)zR(cid:17) Similarly we have for the second and sixth term, third and seventh term , fourth and eighth term, τ(cid:16)Rd(Rz∗)d(zR) + zRd(Rz∗)d(zRz∗)(cid:17) = τ(cid:16)d(Rz∗)d(zR) + zRd(Rz∗)zRdz∗(cid:17) τ(cid:16)Rz∗d(zR)d(R) + zRz∗d(zR)d(Rz∗)(cid:17) = τ(cid:16)z∗d(zR)dR + zRz∗d(zR)Rdz∗(cid:17) τ(cid:16)Rz∗d(zRz∗)d(zR) + zRz∗d(zRz∗)d(zRz∗)(cid:17) = τ(cid:16)z∗d(zRz∗)d(zR) + zRz∗d(zRz∗)zRdz∗(cid:17) Recoupling these terms, τ(cid:16)dRdR + z∗d(zR)dR(cid:17) = τ(cid:16)R−1dRdR + z∗(dz)RdR(cid:17) τ(cid:16)zR(dR)Rdz∗ + zRz∗d(zR)Rdz∗(cid:17) = τ(cid:16)z(dR)Rdz∗ + zRz∗dzR2dz∗(cid:17) τ(cid:16)d(Rz∗)d(zR) + z∗d(zRz∗)d(zR)(cid:17) = τ(cid:16)R−1d(Rz∗)d(zR) + z∗(dz)Rz∗d(zR)(cid:17) τ(cid:16)zRd(Rz∗)zRdz∗ + zRz∗d(zRz∗)zRdz∗(cid:17) = τ(cid:16)zd(Rz∗)zRdz∗ + zRz∗(dz)Rz∗zRdz∗(cid:17) QUANTUM EUCLIDEAN SPACES WITH NONCOMMUTATIVE DERIVATIVES 45 On the right hand side, there are only three terms still contains derivatives of products. We again use Leibniz rule, τ (R−1d(Rz∗)d(zR)) =τ (R−1d(R)z∗d(zR) + dz∗d(zR)) =τ (d(R)z∗d(z) + R−1d(R)(R−1 − 1)dR) + dz∗d(z)R + dz∗zdR) τ (z∗(dz)Rz∗d(zR)) =τ (z∗(dz)(1 − R)dR + z∗(dz)Rz∗d(z)R) τ (zd(Rz∗)zRdz∗) =τ (z∗Rdz∗zRdz∗ + zdR(1 − R)dz∗) Gathering all the terms we have, ((dR)z∗dz + z∗dzdR) + (dz∗zdR + zdRdz∗)+ (zR(dz∗)zRdz∗ + R−1dRR−1dR + (dz)Rz∗(dz)Rz∗) + Rdz∗dz + zRz∗(dz)Rdz∗ . Here only the last two terms has nonzero trace. This is because for any a1, a2, a3, b1, b2b3 Strθ(cid:16)a1(da2)a3b1(db2)b3(cid:17) = −Strθ(cid:16)b1(db2)b3a1(da2)a3(cid:17), Strθ(cid:16)a1(da2)a3a1(da2)a3(cid:17) = 0. This follows from that fact a1(da2)a3 has Clifford term of order 1 hence a1(da2)a3γ = −γa1(da2)a3. It remains to calculate the trace of Rdz∗dz + zRz∗dzRdz∗. Note that zz∗ = z∗z − 2θ = R−1 − 1 − 2θ , dz = −ic1 + c2 , dz∗ = −ic1 − c2 . Then Strθ(Rdz∗dz + zRz∗(dz)Rdz∗) = 4τθ(R − zRz∗R) Finally we use the spectrum of quantum harmonic oscillator the above trace. Assume that θ > 0. By Proposition 2.4, there is a trace preserving ∗-isomorphism (up to a factor 2πθ π : Rθ → B(L2(R)) such that √θDx , y 7→ x 7→ √θx , Recall that H = D2 x + x2 is the Hamiltonian of 1-dimensional quantum harmonic oscillator which has eigenbasis ni, n ≥ 0 with Hni = (2n + 1)ni. For the creation operator a∗ = Dx + ix and the annihilation a = Dx − ix, Now take z = √θa∗, z∗ = √θa and R−1 = 1 + 2θ + zz∗ = θ(H + 1) + 1. We have a∗ni = √2n + 2n + 1i , ani = √2nn − 1i 4τθ(R − zRz∗R) = 2θπ · 4Xk=0 = 8θπXk=0 1 1 + 2θ + 2kθ − 1 1 1 2kθ 1 + 2kθ 1 + 2θ + 2kθ 1 + 2kθ 1 + 2θ + 2kθ = 4π . For φ0, we have φ0(e − 1e) = Strθ((e − 1e)ω) = τθ(R + zRz∗ − 1)tr(γω) = 2θ′τθ(R + zRz∗ − 1) Note that R−1 = 1 + z∗z = 1 + θ + x2 + y2 and [R−1, z] = [x2 + y2, x + iy] = 2θz. Then, R + zRz∗ − 1 =R(1 + z∗z) − 1 + [z, Rz∗] = [z, Rz∗] =[z, R]z∗ + R[z, z∗] = R[R−1, z]Rz∗ − 2θR = 2θ(RzRz∗ − R) 46 LI GAO, MARIUS JUNGE, AND EDWARD MCDONALD We have calculated that τθ(R − RzRz∗) = 2π. So Strθ((e − 1e)ω) = −θθ′4π. To conclude, we have the index pairing h[e] − [1e], (W ∞,1(Rθ), L2(RΘ) ⊗ MN , D)i =πStrθ((e − 1e)ω) + πStrθ(edede) = − 4π2θθ′ + 4π2 = 4π2(1 − θθ′) Recall for d = 2 that Θ = 0 θ 1 −θ 0 0 0 1 0 1 −θ′ 0 0 1 θ′ 0 . When det Θ = (1 − θθ′)2 6= 0, we have RΘ is ∗-isomorphic to B(L2(R2)) with the trace differs by a factor τΘ = (2π)21 − θθ′tr, which is exactly the normalization constant we obtained. In other words, if we replace τΘ with the matrix trace tr, Indextr(eFµ,+e) = 1 (or −1). Since for every θ, we can choose θ′ such that θθ′ θ ) is a representative of generator of the K0(Eθ) = Z. (cid:3) 6= 1, then the index pairing shows that e ∈ M2(E∼ References [AHS78] [BK90] [BM12] [BR97] [BR12] [Bri88] [BS92] J. Avron, I. Herbst, and B. Simon. Schrödinger operators with magnetic fields. I. General inter- actions. Duke Math. J., 45(4):847 -- 883, 1978. Lawrence G Brown and Hideki Kosaki. Jensen's inequality in semi-finite von neumann algebras. Journal of Operator Theory, pages 3 -- 19, 1990. Tanvir Ahamed Bhuyain and Matilde Marcolli. The ricci flow on noncommutative two-tori. Letters in Mathematical Physics, 101(2):173 -- 194, 2012. Ola Bratteli and Derek W. Robinson. Operator algebras and quantum statistical mechanics. 2. Texts and Monographs in Physics. Springer-Verlag, Berlin, second edition, 1997. Equilibrium states. Models in quantum statistical mechanics. Ola Bratteli and Derek William Robinson. Operator Algebras and Quantum Statistical Mechan- ics: Volume 1: C*-and W*-Algebras. Symmetry Groups. Decomposition of States. Springer Sci- ence & Business Media, 2012. Chris Brislawn. Kernels of trace class operators. Proceedings of the American Mathematical Society, 104(4):1181 -- 1190, 1988. David P Blecher and Roger R Smith. The dual of the haagerup tensor product. Journal of the London Mathematical Society, 2(1):126 -- 144, 1992. [CGRS14] Alan L Carey, Victor Gayral, Adam Rennie, and Fedor A Sukochev. Index theory for locally [CL01] [CM90] [CM95] [CM14] [Con] [CT11] [FK13] [Gao18] [GBV88] compact noncommutative geometries. American Mathematical Soc., 2014. Alain Connes and Giovanni Landi. Noncommutative manifolds, the instanton algebra, and isospectral deformations. Communications in mathematical physics, 221(1):141 -- 159, 2001. Alain Connes and Henri Moscovici. Cyclic cohomology, the novikov conjecture and hyperbolic groups. Topology, 29(3):345 -- 388, 1990. Alain Connes and Henri Moscovici. The local index formula in noncommutative geometry. In Geometries in Interaction, pages 174 -- 243. Springer, 1995. Alain Connes and Henri Moscovici. Modular curvature for noncommutative two-tori. Journal of the American Mathematical Society, 27(3):639 -- 684, 2014. Alain Connes. Noncommutative geometry, 1994. Alain Connes and Paula Tretkoff. The gauss-bonnet theorem for the noncommutative two torus. Noncommutative geometry, arithmetic, and related topics, pages 141 -- 158, 2011. Farzad Fathizadeh and Masoud Khalkhali. Weyl's law and connes' trace theorem for noncom- mutative two tori. Letters in Mathematical Physics, 103(1):1 -- 18, 2013. Li Gao. Continuous perturbations of noncommutative euclidean spaces and tori. Journal of Operator Theory, 79(1):173 -- 200, 2018. José M Gracia-Bondía and Joseph C Varilly. Algebras of distributions suitable for phase-space quantum mechanics. i. Journal of Mathematical Physics, 29(4):869 -- 879, 1988. QUANTUM EUCLIDEAN SPACES WITH NONCOMMUTATIVE DERIVATIVES 47 [GJP17] [GGBI+04] Victor Gayral, Jose M Gracia-Bondia, Bruno Iochum, Thomas Schücker, and Joseph C Várilly. Moyal planes are spectral triples. Communications in mathematical physics, 246(3):569 -- 623, 2004. A. M. González-Pérez, M. Junge, and J. Parcet. Singular integrals in quantum Euclidean spaces. ArXiv e-prints, May 2017. Loukas Grafakos. Modern fourier analysis, volume 250. Springer, 2009. Brian C Hall. Quantum theory for mathematicians. Springer, 2013. Nigel Higson. The local index formula in noncommutative geometry. Contemporary developments in algebraic K-theory, ICTP Lecture Notes, 15:444 -- 536, 2003. [Gra09] [Hal13] [Hig03] [HLP18a] H. Ha, G. Lee, and R. Ponge. Pseudodifferential calculus on noncommutative tori, I. Oscillating integrals. ArXiv e-prints, March 2018. [HLP18b] H. Ha, G. Lee, and R. Ponge. Pseudodifferential calculus on noncommutative tori, II. Main [Lan95] [LM16] [LSZ17] [Mer05] [MP04] properties. ArXiv e-prints, March 2018. E Christopher Lance. Hilbert C*-modules: a toolkit for operator algebraists, volume 210. Cam- bridge University Press, 1995. Matthias Lesch and Henri Moscovici. Modular curvature and morita equivalence. Geometric and Functional Analysis, 26(3):818 -- 873, 2016. Galina Levitina, Fedor Sukochev, and Dmitriy Zanin. Cwikel estimates revisited. arXiv preprint arXiv:1703.04254, 2017. Marcela I Merklen. Boundedness of pseudodifferential operators of a c∗-algebra-valued symbol. Proceedings of the Royal Society of Edinburgh Section A: Mathematics, 135(6):1279 -- 1286, 2005. Marius Măntoiu and Radu Purice. The magnetic Weyl calculus. J. Math. Phys., 45(4):1394 -- 1417, 2004. [MPR05] Marius Măntoiu, Radu Purice, and Serge Richard. Twisted crossed products and magnetic pseu- dodifferential operators. In Advances in operator algebras and mathematical physics, volume 5 of Theta Ser. Adv. Math., pages 137 -- 172. Theta, Bucharest, 2005. Nikita Nekrasov and Albert Schwarz. Instantons on noncommutative r4, and (2, 0) superconfor- mal six dimensional theory. Communications in Mathematical Physics, 198(3):689 -- 703, 1998. Raphaël Ponge. A new short proof of the local index formula and some of its applications. Communications in mathematical physics, 241(2-3):215 -- 234, 2003. Marc Aristide Rieffel. Deformation Quantization for Actions of Rd. Number 506. American Mathematical Soc., 1993. [Pon03] [NS98] [Rie93] [RLLL00] Mikael Rørdam, Flemming Larsen, Flemming Larsen, and N Laustsen. An Introduction to K- [SMZ18] [Ste16] [SW99] [Tak08] [Tao18] [VGB88] theory for C*-algebras, volume 49. Cambridge University Press, 2000. F. Sukochev, E. McDonald, and D. Zanin. A C ∗-algebraic approach to the principal symbol II. ArXiv e-prints, June 2018. Elias M Stein. Harmonic Analysis (PMS-43), Volume 43: Real-Variable Methods, Orthogonality, and Oscillatory Integrals.(PMS-43), volume 43. Princeton University Press, 2016. Nathan Seiberg and Edward Witten. String theory and noncommutative geometry. Journal of High Energy Physics, 1999(09):032, 1999. Leon A. Takhtajan. Quantum mechanics for mathematicians, volume 95 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2008. Jim Tao. The theory of pseudo-differential operators on the noncommutative n-torus. In Journal of Physics: Conference Series, volume 965, page 012042. IOP Publishing, 2018. Joseph C Várilly and José M Gracia-Bondía. Algebras of distributions suitable for phase-space quantum mechanics. ii. topologies on the moyal algebra. Journal of Mathematical Physics, 29(4):880 -- 887, 1988. 48 LI GAO, MARIUS JUNGE, AND EDWARD MCDONALD Department of Mathematics, Texas A&M University, College Station, TX 77843, USA E-mail address, Li Gao: [email protected] Department of Mathematics, University of Illinois, Urbana, IL 61801, USA E-mail address, Marius Junge: [email protected] School of Mathematics and Statistics, University of New South Wales, UNSW Sydney NSW, 2052, Australia E-mail address, Edward McDonald: [email protected]
1807.01598
2
1807
2018-07-12T21:48:38
The polar decomposition for adjointable operators on Hilbert $C^*$-modules and centered operators
[ "math.OA" ]
Let $T$ be an adjointable operator between two Hilbert $C^*$-modules and $T^*$ be the adjoint operator of $T$. The polar decomposition of $T$ is characterized as $T=U(T^*T)^\frac12$ and $\mathcal{R}(U^*)=\overline{\mathcal{R}(T^*)}$, where $U$ is a partial isometry, $\mathcal{R}(U^*)$ and $\overline{\mathcal{R}(T^*)}$ denote the range of $U^*$ and the norm closure of the range of $T^*$, respectively. Based on this new characterization of the polar decomposition, an application to the study of centered operators is carried out.
math.OA
math
THE POLAR DECOMPOSITION FOR ADJOINTABLE OPERATORS ON HILBERT C ∗-MODULES AND CENTERED OPERATORS NA LIU, WEI LUO, and QINGXIANG XU∗ Abstract. Let T be an adjointable operator between two Hilbert C ∗-modules and T ∗ be the adjoint operator of T . The polar decomposition of T is charac- 2 and R(U ∗) = R(T ∗), where U is a partial isometry, terized as T = U (T ∗T ) R(U ∗) and R(T ∗) denote the range of U ∗ and the norm closure of the range of T ∗, respectively. Based on this new characterization of the polar decompo- sition, an application to the study of centered operators is carried out. 1 8 1 0 2 l u J 2 1 ] . A O h t a m [ 2 v 8 9 5 1 0 . 7 0 8 1 : v i X r a 1. Introduction Much progress has been made in the study of the polar decomposition for Hilbert space operators [3, 5, 6, 7, 15]. Let H, K be two Hilbert spaces and B(H, K) be the set of bounded linear operators from H to K. For any T ∈ B(H, K), let T ∗, R(T ) and N (T ) denote the conjugate operator, the range and the null space of T , respectively. It is well-known [6, 7] that every operator T ∈ B(H, K) has the unique polar decomposition T = UT and N (T ) = N (U), (1.1) where T = (T ∗T ) expression of (1.1) is 1 2 and U ∈ B(H, K) is a partial isometry. An alternative T = UT and R(T ∗) = R(U ∗), (1.2) since N (T )⊥ = R(T ∗) and N (U)⊥ = R(U ∗) = R(U ∗) in the Hilbert space case. Note that if H = K, then B(H, H) abbreviated to B(H), is a von Neumann algebra. It follows from [14, Proposition 2.2.9] that the polar decomposition also works for elements in a von Neumann algebra. Nevertheless, it may be false for some elements in a general C ∗-algebra; see [14, Remark 1.4.6]. Both Hilbert spaces and C ∗-algebras can be regarded as Hilbert C ∗-modules, so one might study the polar decomposition in the general setting of Hilbert C ∗- modules. An adjointable operator between Hilbert C ∗-modules may have no polar decomposition unless some additional conditions are satisfied; see Lemma 3.5 be- low for the details. The polar decomposition for densely defined closed operators and unbounded operators are also considered in some literatures; see [2, 4, 5] for example. 2010 Mathematics Subject Classification. Primary 46L08; Secondary 47A05. Key words and phrases. Hilbert C ∗-module, polar decomposition, centered operator. 1 2 N. LIU, W. LUO, Q. XU The purpose of this paper is, in the general setting of adjointable operators on Hilbert C ∗-modules, to provide a new insight into the polar decomposition theory and its applications. We will prove in Lemma 3.6 that five equalities appearing in Lemma 3.5 (iii) can be in fact simplified to two equalities described in (3.2), which are evidently the same as that in (1.2) when the underlying spaces are Hilbert spaces. It is remarkable that (1.1) is a widely used characterization of the polar decomposition for Hilbert space operators. Nevertheless, Example 3.15 indicates that such a characterization of the polar decomposition is no longer true for adjointable operators on Hilbert C ∗-modules. This leads us to figure out a modified version of (1.1), which is stated in Theorem 3.13. Note that the verification of the equivalence of Lemma 3.5 (i) and (ii) is trivial, so it is meaningful to give another interpretation of Lemma 3.5 (ii). We have managed to do that in Theorem 3.8 (iii). One application of the polar decomposition is the study of centered operators on Hilbert spaces, which was initiated in [12] and generalized in [8, 9]. Based on the new characterization (3.2) of the polar decomposition for adjointable operators, some generalizations on centered operators are made in the framework of Hilbert C ∗-modules. The paper is organized as follows. Some elementary results on adjointable op- erators are provided in Section 2. In Section 3, we focus on the study of the polar decomposition for adjointable operators on Hilbert C ∗-modules. As an applica- tion of the polar decomposition, centered operators are studied in Section 4. 2. Some elementary results on adjointable operators Hilbert C ∗-modules are generalizations of Hilbert spaces by allowing inner products to take values in some C ∗-algebras instead of the complex field. Let A be a C ∗-algebra. An inner-product A-module is a linear space E which is a right A-module, together with a map (x, y) →(cid:10)x, y(cid:11) : E × E → A such that for any x, y, z ∈ E, α, β ∈ C and a ∈ A, the following conditions hold: (i) hx, αy + βzi = αhx, yi + βhx, zi; (ii) hx, yai = hx, yia; (iii) hy, xi = hx, yi∗; (iv) hx, xi ≥ 0, and hx, xi = 0 ⇐⇒ x = 0. An inner-product A-module E which is complete with respect to the induced norm (kxk =pkhx, xik for x ∈ E) is called a (right) Hilbert A-module. Suppose that H and K are two Hilbert A-modules, let L(H, K) be the set of operators T : H → K for which there is an operator T ∗ : K → H such that hT x, yi = hx, T ∗yi for any x ∈ H and y ∈ K. We call L(H, K) the set of adjointable operators from H to K. For any T ∈ L(H, K), the range and the null space of T are denoted by R(T ) and N (T ), respectively. In case H = K, L(H, H) which is abbreviated to L(H), is a C ∗- algebra. Let L(H)sa and L(H)+ be the set of self-adjoint elements and positive elements in L(H), respectively. THE POLAR DECOMPOSITION AND CENTERED OPERATORS 3 Definition 2.1. A closed submodule M of a Hilbert A-module E is said to be orthogonally complemented if E = M ∔ M ⊥, where M ⊥ =(cid:8)x ∈ E : hx, yi = 0 for any y ∈ M(cid:9). In this case, the projection from H onto M is denoted by PM . Throughout the rest of this paper, A is a C ∗-algebra, E, H and K are three Hilbert A-modules. Note that L(H) is a C ∗-algebra, so we begin with an elemen- tary result on C ∗-algebras. Definition 2.2. Let B be a C ∗-algebra. The set of positive elements of B is denoted by B+. For any a, b ∈ B, let [a, b] = ab − ba be the commutator of a and b. Proposition 2.3. Let B be a C ∗-algebra and let a, b ∈ B be such that a = a∗ and [a, b] = 0. Then [f (a), b] = 0 whenever f is a continuous complex-valued function on the interval [−kak, kak]. Proof. We might as well assume that B has a unit. Choose any sequence {pn}∞ n=1 of polynomials such that pn(t) → f (t) uniformly on the interval [−kak, kak]. Then kpn(a) − f (a)k → 0 as n → ∞, hence f (a)b = lim n→∞ pn(a)b = lim n→∞ b pn(a) = bf (a). (cid:3) Next, we state some elementary results on the commutativity of adjointable operators. For any α > 0, the function f (t) = tα is continuous on [0, +∞), so a direct application of Proposition 2.3 yields the following proposition: Proposition 2.4. Let S ∈ L(H) and T ∈ L(H)+ be such that [S, T ] = 0. Then [S, T α] = 0 for any α > 0. The technical result of this section is as follows: Proposition 2.5. Let T ∈ L(H)+ be such that R(T ) is orthogonally comple- mented. Then lim n→∞ kTnx − PR(T )xk = 0 for all x in H, (2.1) where Tn =(cid:0) 1 n I + T(cid:1)−1 T for each n ∈ N. Proof. For each n ∈ N, the continuous function fn associated to the operator Tn is given by Now, given any x ∈ H and any ε > 0, let x = u + v, where u ∈ R(T ) and v ∈ N (T ) ⊆ N (Tn) for any n ∈ N. Choose h ∈ H and n0 ∈ N such that ku − T hk < ε 3 and n0 > 3(khk + 1) ε . fn(t) = for t ∈ sp(T ) ⊆ [0, kT k], t 1 n + t where sp(T ) is the spectrum of T . Then for each n ∈ N, kTnk = max(cid:8)(cid:12)(cid:12)fn(t)(cid:12)(cid:12) : t ∈ sp(T )(cid:9) ≤ 1; kTnT − T k = max(cid:8)(cid:12)(cid:12)tfn(t) − t(cid:12)(cid:12) : t ∈ sp(T )(cid:9) ≤ 1 n . 4 N. LIU, W. LUO, Q. XU Then for any n ∈ N with n ≥ n0, we have kTnx − PR(T )xk = kTnu − PR(T )uk ≤ kTnu − TnT hk + kTnT h − PR(T )T hk + kPR(T )T h − PR(T )uk ≤ kTn(u − T h)k + k(TnT )h − T hk + kPR(T )(T h − u)k ≤ ku − T hk + 1 n khk + kT h − uk < ε 3 + ε 3 + ε 3 = ε. This completes the proof of (2.1). (cid:3) Based on Proposition 2.5, a result on the commutativity for adjointable oper- ators can be provided as follows: Proposition 2.6. Let S ∈ L(H) and let T ∈ L(H)+ be such that R(T ) is orthogonally complemented. If [S, T ] = 0, then hS, PR(T )i = 0. Proof. Denote PR(T ) simply by P . Since [S, T ] = 0, we have [S, Tn] = 0, where Tn (n ∈ N) are given in Proposition 2.5. It follows from (2.1) that P (Sx) = lim n→∞ Tn(Sx) = lim n→∞ S(Tnx) = S(P x) for any x ∈ H. (cid:3) We end this section by stating some range equalities for adjointable operators. Proposition 2.7. Let A ∈ L(H, K) and B, C ∈ L(E, H) be such that R(B) = R(C). Then R(AB) = R(AC). Proof. Let x ∈ E be arbitrary. Since Bx ∈ R(C), there exists a sequence {xn} in E such that Cxn → Bx. Then ACxn → ABx, which means ABx ∈ R(AC), and thus R(AB) ⊆ R(AC) and furthermore R(AB) ⊆ R(AC). Similarly, we have R(AC) ⊆ R(AB). (cid:3) Lemma 2.8. [17, Lemma 2.3] Let T ∈ L(H)+. Then R(T α) = R(T ) for any α ∈ (0, 1). Proposition 2.9. Let T ∈ L(H)+. Then R(T α) = R(T ) for any α > 0. Proof. In view of Lemma 2.8, we might as well assume that α > 1. Put S = T α. Then S ∈ L(H)+, so from Lemma 2.8 we have R(T ) = R(S 1 α ) = R(S) = R(T α). (cid:3) 3. The polar decomposition for adjointable operators In this section, we study the polar decomposition for adjointable operators on Hilbert C ∗-modules. Definition 3.1. Recall that an element U of L(H, K) is said to be a partial isometry if U ∗U is a projection in L(H). Proposition 3.2. [17, Lemma 2.1] Let U ∈ L(H, K) be a partial isometry. Then U ∗ is also a partial isometry which satisfies U U ∗U = U . THE POLAR DECOMPOSITION AND CENTERED OPERATORS 5 Lemma 3.3. [11, Proposition 3.7] Let T ∈ L(H, K). Then R(T ∗T ) = R(T ∗) and R(T T ∗) = R(T ). Definition 3.4. For any T ∈ L(H, K), let T denote the square root of T ∗T . That is, T = (T ∗T ) 2 and T ∗ = (T T ∗) 1 1 2 . Lemma 3.5. [16, Proposition 15.3.7] Let T ∈ L(E). Then the following state- ments are equivalent: (i) E = N (T ) ⊕ R(T ) and E = N (T ∗) ⊕ R(T ); (ii) Both R(T ) and R(T ) are orthogonally complemented; (iii) T has the polar decomposition T = UT , where U ∈ L(E) is a partial isometry such that N (U) = N (T ), N (U ∗) = N (T ∗), R(U) = R(T ), R(U ∗) = R(T ∗). (3.1) Lemma 3.6. Let T ∈ L(H, K) be such that R(T ∗) is orthogonally complemented, and let U ∈ L(H, K) be a partial isometry such that T = UT and U ∗U = PR(T ∗). (3.2) Then R(T ) is also orthogonally complemented, and all equations in (3.1) are satisfied. Furthermore, the following equations are also valid: T ∗ = U ∗T ∗ and U U ∗ = PR(T ), T ∗ = UT U ∗ and UT = T ∗U. Proof. By Proposition 2.9 and Lemma 3.3, we have R(T ) = R(T ∗T ) = R(T ∗) = R(U ∗U), which gives by Proposition 2.7 that R(T ) = R(UT ) = R(U U ∗U) = R(U U ∗), (3.3) (3.4) (3.5) hence R(T ) is orthogonally complemented such that the second equation in (3.3) is satisfied. Furthermore, T T ∗ = UT · T U ∗ = (UT U ∗)2, hence the first equation in (3.4) is satisfied. As a result, U ∗T ∗ = (U ∗UT )U ∗ = T U ∗ = (UT )∗ = T ∗, UT = T = (T ∗)∗ = (U ∗T ∗)∗ = T ∗U. This completes the proof of (3.3) and (3.4). Finally, equations stated in (3.1) can be derived directly from the second equations in (3.2) and (3.3), respectively. (cid:3) Lemma 3.7. [17, Theorem 3.1] Let T ∈ L(H, K) be such that R(T ∗) is or- thogonally complemented. If R(T ∗) ⊆ R(T ), then the following statements are valid: (i) R(T ∗) = R(T ); (ii) R(T ) = R(T ∗); 6 N. LIU, W. LUO, Q. XU (iii) R(T ) is orthogonally complemented. Theorem 3.8. Let T ∈ L(H, K). Then the following statements are equivalent: (i) R(T ) and R(T ∗) are both orthogonally complemented; (ii) R(T ∗) is orthogonally complemented and (3.2) is satisfied for some partial isometry U ∈ L(H, K); (iii) R(T ∗) is orthogonally complemented, R(T ) = R(T ∗) and R(T ∗) = R(T ). Proof. The implications of (ii)=⇒(i) and (iii)=⇒(i) follow from Lemmas 3.6 and 3.7, respectively. "(i) =⇒ (ii)": Let E = H ⊕ K and eT =(cid:18) 0 0 T 0 (cid:19) ∈ L(E). Then both R(eT ) and R(eT ∗) are orthogonally complemented, hence by Lemma 3.5 there exists a partial isometry eU =(cid:18) U11 U12 U U22 (cid:19) ∈ L(E) such that eT = eU eT , R(eU) = R(eT ) = {0} ⊕ R(T ) and R(eU ∗) = R(eT ∗) = R(T ∗) ⊕ {0}, which leads to eU =(cid:18) 0 0 U 0 (cid:19), hence U is a partial isometry satisfying (3.2). "(ii) =⇒ (iii)": By (3.2) -- (3.4), we have T ∗ = (UT )∗ = T U ∗ and T ∗U = U ∗T ∗ · U = U ∗ · UT U ∗ · U = T , which obviously lead to R(T ) = R(T ∗). Replacing T, U with T ∗, U ∗, we obtain R(T ∗) = R(T ). (cid:3) Lemma 3.9. Let T ∈ L(H, K) be such that R(T ∗) is orthogonally complemented. If U, V ∈ L(H, K) are given such that UT = V T and U ∗U = V ∗V = PR(T ∗), then U = V . Proof. The equation UT = V T together with (3.5) yields U PR(T ∗) = V PR(T ∗), hence U = U(U ∗U) = U PR(T ∗) = V PR(T ∗) = V (V ∗V ) = V. (cid:3) Definition 3.10. The polar decomposition of T ∈ L(H, K) can be characterized as T = UT and U ∗U = PR(T ∗), (3.6) where U ∈ L(H, K) is a partial isometry. Remark 3.11. It follows from Theorem 3.8 and Lemma 3.9 that T ∈ L(H, K) has the (unique) polar decomposition if and only if R(T ∗) and R(T ) are both orthogonally complemented. In this case, T ∗ = U ∗T ∗ is the polar decomposition of T ∗. A slight generalization of (3.4) is as follows: Lemma 3.12. Let T = UT be the polar decomposition of T ∈ L(H, K). Then for any α > 0, the following statements are valid: THE POLAR DECOMPOSITION AND CENTERED OPERATORS 7 (i) UT αU ∗ = (UT U ∗)α = T ∗α; (ii) UT α = T ∗αU ; (iii) U ∗T ∗αU = (U ∗T ∗U)α = T α. Proof. (i) Since U ∗UT = T , we have (cid:0)UT U ∗(cid:1)n = UT nU ∗ for any n ∈ N. (3.7) Let f (t) = tα and choose any sequence {Pm}∞ 0 (∀m ∈ N), and Pm(t) → f (t) uniformly on the interval (cid:2)0,(cid:13)(cid:13)T (cid:13)(cid:13)(cid:3). Then from (3.4) and (3.7), we have m=1 of polynomials such that Pm(0) = UT αU ∗ = U f (T )U ∗ = lim m→∞ U Pm(T )U ∗ = lim m→∞ = f(cid:0)UT U ∗(cid:1) = (UT U ∗)α = T ∗α. (ii) By Proposition 2.9 and Lemma 3.3, we have Pm(cid:0)UT U ∗(cid:1) R(T α) = R(T ∗T ) = R(T ∗), and thus U ∗UT α = T α. Taking ∗-operation, we get T α = T αU ∗U. It follows from (i) that UT α = U(cid:0)T αU ∗U(cid:1) =(cid:0)UT αU ∗(cid:1)U = T ∗αU. (iii) Since T ∗ = U ∗T ∗ is the polar decomposition of T ∗, the conclusion follows (cid:3) immediately from (i) by replacing the pair (U, T ) with (U ∗, T ∗). Before ending this section, we provide a criteria for the polar decomposition as follows: Theorem 3.13. Let T ∈ L(H, K) be such that R(T ∗) is orthogonally comple- mented. Let U ∈ L(H, K) be a partial isometry which satisfies T = UT and N (T ) ⊆ N (U). (3.8) Then T = UT is the polar decomposition of T . Proof. By assumption, Q = U ∗U is a projection. For any x ∈ H, we have hT x, T xi = hT x, T xi = hUT x, UT xi = hQT x, T xi, and thus (cid:13)(cid:13)(I − Q)T x(cid:13)(cid:13)2 =(cid:13)(cid:13)h(I − Q)T x, T xi(cid:13)(cid:13) = 0, hence QT = T . It follows that R(T ∗) = R(T ) ⊆ R(Q). On the other hand, by assumption we have R(T ∗) = N (T )⊥ ⊇ N (U)⊥ = N (Q)⊥ = R(Q), hence R(T ∗) = R(Q) and thus Q = PR(T ∗). (cid:3) Remark 3.14. Let T ∈ L(H, K), where H and K are both Hilbert spaces. In this case R(T ∗) is always orthogonally complemented, so if U is a partial isometry such that (3.8) is satisfied, then UT is exactly the polar decomposition of T . Unlike the assertion given in [10, P. 3400], the same is not true for general Hilbert C ∗-modules H and K, since R(T ∗) can be not orthogonally complemented Indeed, there exist a Hilbert C ∗-module H, and an for some T ∈ L(H, K). 8 N. LIU, W. LUO, Q. XU adjointable T and a partial isometry U on H such that (1.1) is satisfied, whereas T has no polar decomposition. Such an example is as follows: Example 3.15. Let H be any countably infinite-dimensional Hilbert space, L(H) and C(H) be the set of bounded linear operators and compact operators on H, respectively. Given any orthogonal normalized basis {en : n ∈ N} for H, let S ∈ C(H) be defined by S(en) = 1 n en, for any n ∈ N. Clearly, S is a positive element in C(H). Let K = A = L(H). With the inner product given by K is a Hilbert A-module. (cid:10)X, Y(cid:11) = X ∗Y for any X, Y ∈ K, Let T : K → K be defined by T (X) = SX for any X ∈ K. Clearly, T ∈ L(K)+ and R(T ) ⊆ C(H). Given any n ∈ N, let Pn be the projection from H onto the linear subspace spanned by {e1, e2, · · · , en}. Let Xn ∈ K be defined by Xn(ej) =(cid:26) jej, 0, if 1 ≤ j ≤ n, otherwise. . ⊥ It is obvious that T (Xn) = Pn, which implies that R(T ) = C(H), hence = {0}, therefore R(T ) fails to be orthogonally complemented. By Re- R(T ) mark 3.11, we conclude that T has no polar decomposition. Furthermore, given any X ∈ K such that T (X) = SX = 0, then X = 0 since S is injective. It follows that N (T ) = {0}. Now, let U be the identity operator on K. Then since T is positive, we have T = UT and N (U) = N (T ), whereas T has no polar decomposition. 4. Characterizations of centered operators In this section, we study centered operators in the general setting of Hilbert C ∗-modules. Definition 4.1. [12] An element T ∈ L(H) is said to be centered if the following sequence · · · , T 3(T 3)∗, T 2(T 2)∗, T T ∗, T ∗T, (T 2)∗T 2, (T 3)∗T 3, · · · consists of mutually commuting operators. We began with a cancelation law introduced in [8, Lemma 3.7]. Let T = UT be the polar decomposition of T ∈ L(H). Suppose that n ∈ N is given such that Then by Proposition 2.6 we have (cid:2)U kT (U k)∗, T (cid:3) = 0 for 1 ≤ k ≤ n. (cid:2)U kT (U k)∗, U ∗U(cid:3) = 0 for 1 ≤ k ≤ n, hence for any s, t ∈ N with 1 ≤ t ≤ s ≤ n + 1, we have U sT (U s)∗U t = U sT (U s−t)∗ and (U t)∗U sT (U s)∗ = U s−tT (U s)∗. (4.3) (4.1) (4.2) THE POLAR DECOMPOSITION AND CENTERED OPERATORS 9 Indeed, by (4.1) and (4.2) we have U sT (U s)∗U t = U · U s−1T (U s−1)∗ · U ∗U · U t−1 = U · U ∗U · U s−1T (U s−1)∗ · U t−1 = U · U s−1T (U s−1)∗ · U t−1 = U 2 · U s−2T (U s−2)∗ · U t−2 = · · · = U sT (U s−t)∗. Taking ∗-operation, we get the second equation in (4.3). Now we are ready to state the technical lemma of this section, which is a modification of [8, Lemma 4.2]. Lemma 4.2. Suppose that T = UT is the polar decomposition of T ∈ L(H). Let n ∈ N be given such that (cid:2)U kT (U k)∗, T l(cid:3) = 0 Then the following statements are equivalent: for any k, l ∈ N with k + l ≤ n + 1. (4.4) (i) (cid:2)U sT (U s)∗, T t(cid:3) = 0 for some s, t ∈ N with s + t = n + 2; (ii) (cid:2)U sT (U s)∗, T t(cid:3) = 0 for any s, t ∈ N with s + t = n + 2. Proof. Let s, t ∈ N be such that s + t = n + 2. Put At = T t2 · U sT (U s)∗ and Bt = U sT (U s)∗ · T t2. Then clearly, (4.5) (4.6) (4.7) Substituting k = 1 and UT U ∗ = T ∗ into (4.4) yields (cid:2)U sT (U s)∗, T t(cid:3) = 0 ⇐⇒ At = Bt. (cid:2)T ∗, T l(cid:3) = 0 for 1 ≤ l ≤ n, (cid:2)U U ∗, T l(cid:3) = 0 for 1 ≤ l ≤ n. which leads by Proposition 2.6 to Note that t = n + 2 − s ≤ n + 1, so if t ≥ 2, then by (4.4), T t2 = T ∗ · (T t−1)∗T t−1 · T = T U ∗ · T t−12 · UT = U ∗ · UT U ∗ · T t−12 · UT = U ∗ · T t−12 · UT U ∗ · UT = U ∗ · T t−12 · UT 2. (4.8) Assume now that t ≥ 2. Then s + 1 = (n + 2 − t) + 1 ≤ n + 1, hence by (4.5), (4.8), (4.4) with l = 1 and (4.3), we have At = U ∗T t−12U · T 2 · U sT (U s)∗ = U ∗T t−12U · U sT (U s)∗ · T 2 = U ∗ · At−1 · UT 2. (4.9) (4.10) 10 Similarly, N. LIU, W. LUO, Q. XU Bt = U sT (U s)∗ · U ∗U U ∗T t−12UT 2 = U ∗U · U sT (U s)∗U ∗T t−12UT 2 = U ∗ · U s+1T (U s+1)∗T t−12 · UT 2 = U ∗ · Bt−1 · UT 2. (4.11) It follows from (4.10) and (4.11) that At = Bt whenever At−1 = Bt−1. Suppose on the contrary that At = Bt. Then by (4.9), (4.7) and (4.5), we have U At = U U ∗ · T t−12U · U sT (U s)∗ · T 2 = T t−12 · U U ∗U · U sT (U s)∗T 2 = T t−12 · U · U sT (U s)∗ · T 2 = T t−12 · U s+1T (U s+1)∗ · UT 2 = At−1 · UT 2 Furthermore, it can be deduced directly from (4.11) and (4.5) that As a result, we obtain which gives U Bt = Bt−1 · UT 2. At−1 · UT 2 = Bt−1 · UT 2, At−1 = At−1U U ∗ = Bt−1U U ∗ = Bt−1, since R(UT 2) = R(U U ∗) and (cid:2)T t−12, U U ∗(cid:3) = 0. Letting t = 2, 3, · · · , n + 1, respectively, we conclude that A1 = B1 ⇐⇒ A2 = B2 ⇐⇒ · · · ⇐⇒ An+1 = Bn+1. In view of (4.6), the proof of the equivalence of (i) and (ii) is complete. (cid:3) In view of Lemma 4.2, we introduce the terms of restricted sequence and the commutativity of an operator along a restricted sequence as follows: Definition 4.3. A sequence {tn}∞ n=1 is called restricted if tn ∈ {1, 2, · · · , n} for each n ∈ N, and an operator T ∈ L(H) is called commutative along this restricted sequence if T has the polar decomposition T = UT such that hU tnT (cid:0)U tn(cid:1)∗ ,(cid:12)(cid:12)T n+1−tn(cid:12)(cid:12)i = 0 for any n ∈ N. A direct application of Lemma 4.2 and Definition 4.3 gives the following corol- lary: Corollary 4.4. Let T ∈ L(H) have the polar decomposition T = UT . Then the following statements are equivalent: (i) (cid:2)U sT (U s)∗, T t(cid:3) = 0 for any s, t ∈ N; (ii) T is commutative along any restricted sequence; (iii) T is commutative along some restricted sequence. Lemma 4.5. [8, Lemma 4.3] Suppose that T = UT is the polar decomposition of T ∈ L(H). Let n ∈ N be given such that (4.1) is satisfied. Then for 1 ≤ k ≤ n+1, (T k)∗ = UT U ∗ · U 2T (U 2)∗ · · · · · U kT (U k)∗. (4.12) THE POLAR DECOMPOSITION AND CENTERED OPERATORS 11 Proof. This lemma was given in [8, Lemma 4.3], where H is a Hilbert space and T ∈ B(H). Checking the proof of [8, Lemma 4.3] carefully, we find out that the same is true for an adjointable operator on a Hilbert C ∗-module. (cid:3) The main result of this section is as follows: Theorem 4.6. (cf. [8, Theorem 4.1]) Let T ∈ L(H) have the polar decomposition T = UT . Then the following statements are equivalent: (i) T is a centered operator; (iv) T is commutative along any restricted sequence; (v) T is commutative along some restricted sequence; (ii) (cid:2)T n, (T m)∗(cid:3) = 0 for any m, n ∈ N; (iii) (cid:2)T n, T ∗(cid:3) = 0 for any n ∈ N; (vi) (cid:2)U mT (U m)∗, T n(cid:3) = 0 for any m, n ∈ N; (vii) (cid:2)U nT (U n)∗, T (cid:3) = 0 for any n ∈ N; (viii) (cid:2)(T n)∗, T (cid:3) = 0 for any n ∈ N; (ix) (cid:2)(U n)∗T ∗U n, T ∗(cid:3) = 0 for any n ∈ N; (x) (cid:2)(U m)∗T ∗U m, (T n)∗(cid:3) = 0 for any m, n ∈ N; (xi) The operators in {T , UT U ∗, U ∗T U, U 2T (U 2)∗, (U 2)∗T U 2, · · · } com- mute with one another. Proof. The proof of (i)⇐⇒(ii) is the same as that given in [8, Theorem 4.1]. "(ii)=⇒(iii)" is clear by putting m = 1 in (ii). "(iii)⇐⇒(vii)": Putting tn = 1 and sn = n for any n ∈ N. Then T is commuta- tive along {tn} ⇐⇒ (iii) is satisfied, and T is commutative along {sn} ⇐⇒ (vii) is satisfied. The equivalence of (iii) -- (vii) then follows from Corollary 4.4. "(vi)=⇒(ii)": Let m and n be any in N. From (vii) and Lemma 4.5 we know that (T m)∗ has the form (4.12) with k therein be replaced by m. Now, each term in (4.12) commutes with T n by (vi), so (cid:2)T n, (T m)∗(cid:3) = 0. The proof of the equivalence of (i) -- (vii) is therefore complete. Since T is centered if and only if T ∗ is centered, the equivalent conditions (viii), (ix) and (x) are then obtained by replacing T and U with T ∗ and U ∗, respectively. It is obvious that (xi)=⇒(vii). "(vii)+(ix)=⇒(xi)": From (vii), (ix) and Proposition 2.6, we get We prove that the operators in (cid:2)U kT (U k)∗, U ∗U(cid:3) =(cid:2)(U k)∗T ∗U k, U U ∗(cid:3) = 0 for any k ∈ N. Ω =(cid:8)T , UT U ∗, U ∗T U, U 2T (U 2)∗, (U 2)∗T U 2, · · ·(cid:9) commute with one another. That is, [A, B] = 0 for any A, B ∈ Ω. To this end, four cases are considered as follows: Case 1: A = U tT (U t)∗ and B = U sT (U s)∗ with 1 ≤ t < s. In this case, we have AB = U t · T · U s−tT (U s−t)∗ · (U t)∗ = U t · U s−tT (U s−t)∗ · T · (U t)∗ = BA. (4.13) 12 N. LIU, W. LUO, Q. XU Case 2: A = (U t)∗T U t and B = (U s)∗T U s with 1 ≤ t < s. In this case, we have [A, B] = 0 as shown in Case 1 by replacing U, T with U ∗, T ∗, since A, B can be expressed alternately as A = (U t+1)∗T ∗U t+1, B = (U s+1)∗T ∗U s+1. Case 3: A = U tT (U t)∗ and B = (U s)∗T U s with t, s ∈ N. In this case, we have BA = (U s)∗T U s+tT (U t)∗ = (U s)∗ · T · U s+tT (U s+t)∗ · U s = (U s)∗ · U s+tT (U s+t)∗ · T · U s = U tT (U s+t)∗ · T · U s = AB. Case 4: A = T and B = (U s)∗T U s with s ∈ N. In this case, we have AB = U ∗ · UT U ∗ · (U s−1)∗T U s−1 · U = U ∗ · (U s−1)∗T U s−1 · UT U ∗ · U = (U s)∗T U s · T U ∗U = BA. This completes the proof that any two elements in Ω are commutative. (cid:3) Acknowledgments. The authors thank the referee for very helpful comments and suggestions. References 1. M. Embry-Wardrop and A. Lambert, Measurable transformations and centered composition operators, Proc. Royal Irish Acad. 90A (1990), no. 2, 165 -- 172. 2. M. Frank and K. Sharifi, Generalized inverses and polar decomposition of unbounded regular operators on Hilbert C ∗-modules, J. Operator Theory 64 (2010), no. 2, 377 -- 386. 3. T. Furuta, On the polar decomposition of an operator, Acta Sci. Math. 46 (1983), no. 1-4, 261 -- 268. 4. R. Gebhardt and K. Schmudgen, Unbounded operators on Hilbert C ∗-modules, Internat. J. Math. 26 (2015), no. 11, 197 -- 255. 5. F. Gesztesy, M. Malamud, M. Mitrea and S. Naboko, Generalized polar decompositions for closed operators in Hilbert spaces and some applications, Integral Equations Operator Theory 64 (2009), no. 1, 83 -- 113. 6. P. R. Halmos, A Hilbert space problem book, Van Nostrand, Princeton, N.J., 1967. 7. W. Ichinose and K. Iwashita, On the uniqueness of the polar decomposition of bounded operators in Hilbert spaces, J. Operator Theory 70 (2013), no. 1, 175 -- 180. 8. M. Ito, T. Yamazaki and M. Yanagida, On the polar decomposition of the Aluthge trans- formation and related results, J. Operator Theory 51 (2004), no. 2, 303 -- 319. 9. M. Ito, T. Yamazaki and M. Yanagida, On the polar decomposition of the product of two operators and its applications, Integral Equations Operator Theory 49 (2004), no. 4, 461 -- 472. 10. M. M. Karizaki, M. Hassani and M. Amyari, Moore-Penrose inverse of product operators in Hilbert C ∗-modules, Filomat 30 (2016), no. 13, 3397 -- 3402. 11. E. C. Lance, Hilbert C ∗-modules-A toolkit for operator algebraists, Cambridge University Press, Cambridge, 1995. 12. B. B. Morrel and P. S. Muhly, Centered operators, Studia Math. 51 (1974), 251 -- 263. 13. V. Paulsen, C. Pearcy and S. Petrovi´c, On centered and weakly centered operators, J. Funct. Anal. 128 (1995), no. 1, 87 -- 101. 14. G. K. Pedersen, C ∗-algebras and their automorphism groups, Academic Press, New York, 1979. 15. J. Stochel and F. H. Szafraniec, The complex moment problem and subnormality: A polar decomposition approach, J. Funct. Anal. 159 (1998), no. 2, 432 -- 491. THE POLAR DECOMPOSITION AND CENTERED OPERATORS 13 16. N. E. Wegge-Olsen, K-theory and C ∗-algebras: A friendly approach, Oxford Univ. Press, Oxford, England, 1993. 17. Q. Xu and X. Fang, A note on majorization and range inclusion of adjointable operators on Hilbert C ∗-modules, Linear Algebra Appl. 516 (2017), 118 -- 125. Department of Mathematics, Shanghai Normal University, Shanghai 200234, PR China. E-mail address: [email protected]; [email protected]; qingxiang [email protected]
1611.08525
3
1611
2018-10-11T13:34:24
Nica-Toeplitz algebras associated with right tensor $C^*$-precategories over right LCM semigroups
[ "math.OA" ]
We introduce and analyze the full $\mathcal{NT}_{\mathcal{L}}(\mathcal{K})$ and the reduced $\mathcal{NT}_{\mathcal{L}}^r(\mathcal{K})$ Nica-Toeplitz algebra associated to an ideal $\mathcal{K}$ in a right tensor $C^*$-precategory $\mathcal{L}$ over a right LCM semigroup $P$. Our main results are uniqueness theorems in the spirit of classical Coburn's theorem, generalizing uniqueness results for Toeplitz-type $C^*$-algebras associated to single $C^*$-correspondences, quasi-lattice ordered semigroups, and crossed products twisted by product systems of $C^*$-correspondences obtained by Fowler, Laca and Raeburn. We formulate geometric conditions on a representation $\Phi$ of $\mathcal{K}$ so that the $C^*$-algebra it generates, $C^*(\Phi(\mathcal{K}))$, naturally lies between $\mathcal{NT}_{\mathcal{L}}^r(\mathcal{K})$ and $\mathcal{NT}_{\mathcal{L}}(\mathcal{K})$. Under suitable amenability hypotheses, $C^*(\Phi(\mathcal{K}))$ and $\mathcal{NT}_{\mathcal{L}}(\mathcal{K})$ are isomorphic. The geometric conditions are necessary for our uniqueness result when the right tensoring preserves $\mathcal{K}$ and in general they capture uniqueness of the $C^*$-algebra generated by a natural extension of $\Phi$ to $\mathcal{L}$. In particular, the latter algebra could be viewed as a Doplicher-Roberts version of $\mathcal{NT}_{\mathcal{L}}(\mathcal{K})$.
math.OA
math
NICA-TOEPLITZ ALGEBRAS ASSOCIATED WITH RIGHT TENSOR C ∗-PRECATEGORIES OVER RIGHT LCM SEMIGROUPS BARTOSZ K. KWAŚNIEWSKI AND NADIA S. LARSEN Abstract. We introduce and analyze the full N T L(K) and the reduced N T r L(K) Nica- Toeplitz algebra associated to an ideal K in a right tensor C ∗-precategory L over a right LCM semigroup P . These C ∗-algebras unify cross-sectional C ∗-algebras associated to Fell bundles over discrete groups and Nica-Toeplitz C ∗-algebras associated to product systems. They also allow a study of Doplicher-Roberts versions of the latter. A new phenomenon is that when P is not right cancellative then the canonical condi- tional expectation takes values outside the ambient algebra. Our main result is a uniqueness theorem that gives sufficient conditions for a representation of K to generate a C ∗-algebra naturally lying between N T L(K) and N T r L(K). We also characterise the situation when N T L(K) ∼= N T r L(K). Unlike previous results for quasi-lattice monoids, P is allowed to contain nontrivial invertible elements, and we accommodate this by identifying an assump- tion of aperiodicity of an action of the group of invertible elements in P . One prominent condition for uniqueness is a geometric condition of Coburn's type, exploited in the work of Fowler, Laca and Raeburn. Here we shed new light on the role of this condition by relating it to a C ∗-algebra associated to L itself. 1. Introduction Tensor C ∗-categories (or monoidal C ∗-categories) and all the more right-tensor C ∗-categories (also called semitensor C ∗-categories) arise naturally in quantum field theory and duality the- ory of compact (quantum) groups [10], [11]. In particular, these structures play a fundamental role in numerous recent results with a flavor of geometric group theory, see e.g. [24], [30], [34] and references therein. Right tensor C ∗-(pre)categories proved also to be a very natural framework allowing efficient description of the structure of Cuntz-Pimsner and Nica-Toeplitz algebras associated to product systems, see [19], [22], [20]. In the present paper we initiate a systematic study of C ∗-algebras associated to right tensor C ∗-precategories inspired by this last class of examples. In fact, already in the context of Nica-Toeplitz algebras associated to product systems, our results extend substantially the existing theory, see [20]. We believe that the "categorial language" is well suited to the complicated analysis of C ∗-algebras over semigroup structures, and has a potential to be used, for instance, in the study of Cuntz- Nica-Pimsner algebras [36] or Doplicher-Roberts algebras [10], [11] and their generalizations. Our initial data consists of an ideal K in a right-tensor C ∗-precategory L over a discrete, left cancellative and unital semigroup P . A C ∗-precategory, as introduced in [19], is a non- unital version of a C ∗-category. The important example that the reader may keep in mind is that of Banach spaces L(Xp, Xq) of adjointable operators between Hilbert A-modules Xp, Xq for p, q ∈ P , that form a product system X = Fp∈P Xp over P . Then the right tensoring structure {⊗1r}r∈P on L = {L(Xp, Xq)}p,q∈P is given by tensoring on the right with the unit 1r in L(Xr). The Banach spaces K(Xp, Xq), p, q ∈ P , of generalized compacts form a C ∗-precategory K, in fact an ideal in L, which need not be preserved by the 'functors' ⊗1r, Date: 25 November 2016. Revised 5 June 2018. 2010 Mathematics Subject Classification. Primary 46L55; Secondary 46L05. 1 2 BARTOSZ K. KWAŚNIEWSKI AND NADIA S. LARSEN r ∈ P . In special cases such structures were considered in [19, Example 3.2], [22, Subsection 3.1]. We give a detailed analysis of this example in [20] where we also explain how the results of the present paper give a new insight to C ∗-algebras associated with X. Another somewhat trivial but important and instructive example is when P = G is a group. Then as we explain (see Section 12) our framework is equivalent to the theory of Fell bundles over discrete groups. In general, there is a natural notion of a right-tensor representation of K which allows us to define the Toeplitz algebra TL(K) of K as the universal C ∗-algebra for these representations. We construct a canonical injective right-tensor representation T of K on an appropriate Fock module FK. We call T the Fock representation of K. We wish to study C ∗-algebras which in general are quotients of TL(K) obtained by considering additional relations coming from the Fock representation. It was Nica [31], who first identified and used explicitly such relations in the context of C ∗-algebras associated to positive cones in quasi-lattice ordered groups. Fowler [14] generalized these conditions to product systems over semigroups studied by Nica. In particular, he introduced notions of compact-alignment and Nica-covariance for such objects. Recently, definitions of Nica covariant representations and the corresponding Nica-Toeplitz algebras were generalized to product systems over right LCM semigroups in [6]. In order to define Nica covariance we work under the assumption that P is a right LCM semigroup, a terminology introduced in [7]. Such semigroups appear also under the name of semigroups satisfying Clifford's condition, see [27] and [32]. We emphasize that passing from positive cones in quasi-lattice ordered groups to right LCM semigroups is not straightforward, and has a number of important consequences. First, it allows to develop a theory independent of the ambient group. In fact, the semigroups we consider need not be (right) cancellative, and hence they might not be embeddable into any group. Second, LCM semigroups allow invertible elements. This makes a number of problems much more delicate, but also allows us to cover a larger class of interesting examples. In particular LCM semigroups could be viewed as a unification of quasi-lattice ordered semigroups and groups. Given a right-tensor C ∗-precategory L over an LCM semigroup P we say that an ideal K in L is well-aligned if for every two 'morphisms' in K that can be tensored so that they can be composed, the composition is again in K. This generalises the notion of compactly aligned product systems. For a well-aligned ideal K in a right-tensor C ∗-precategory L we introduce representations which we call Nica covariant. We show that the Fock representation of K is Nica covariant. Two C ∗-algebras are then naturally associated to K: a Nica-Toeplitz algebra N T L(K) universal for Nica covariant representations, and a reduced Nica-Toeplitz algebra N T r L(K) which is generated by the Fock representation of K. L(K), and a faithful completely positive map ET from N T r One important tool to study Nica-Toeplitz type C ∗-algebras is a conditional expectation onto a natural core subalgebra. While core subalgebras of both N T L(K) and N T r L(K) are easy to write down, conditional expectations onto the respective cores do not obviously exist. In the generality of a right LCM semigroup, which need not be right cancellative, we find new ingredients, namely a self-adjoint operator space BK, which in general is not a subspace of N T r L(K) onto BK. We refer to BK as a transcendental core and to ET as a transcendental conditional expectation. The map ET becomes a genuine conditional expectation onto the core subalgebra of N T r L(K) if and only if the semigroup P is cancellative. We note that a similar phenomenon was recently discovered in the context of C ∗-algebras associated to actions of inverse semigroups of Hilbert bimodules in [8], where a notion of a weak conditional expectation is introduced (it is a genuine conditional expectation if and only if the space of units is closed in the dual transformation groupoid). We define an exotic Nica-Toeplitz algebra to be the C ∗-algebra C ∗(Φ(K)) generated by a Nica covariant representation Φ such that there is a compatible transcendental conditional NICA-TOEPLITZ ALGEBRAS FOR RIGHT TENSOR C∗-PRECATEGORIES 3 expectation from C ∗(Φ(K)) onto BK. This is equivalent to existence of a ∗-homomorphism Φ∗ making the following diagram commute: N T L(K) Φ⋊P / C ∗(Φ(K)) Φ∗ / N T r L(K) T ⋊P The uniqueness theorems we aim at require studying two somewhat independent problems, which are interesting in their own right. The first problem is to identify ideals K for which the regular representation T ⋊ P is injective. The second one is to find conditions on a Nica covariant representation Φ so that C ∗(Φ(K)) is an exotic Nica-Toeplitz algebra. Having these two ingredients we infer that both Φ ⋊ P and Φ∗ are isomorphisms. Amenability as a characterization of injectivity of the regular representation of the universal C ∗-algebra for Nica covariant representations is prominent in work of Nica [31], Laca-Raeburn [25] and Fowler [14]. Motivated by this, we say that a well aligned ideal K of L is amenable if the regular representation T ⋊ P is an isomorphism from N T L(K) onto N T r In Theorem 8.4 we prove a far-reaching generalization of [25, Proposition 4.2] and [14, Theorem 8.1]. Our result establishes necessary and sufficient conditions for amenability of K in terms of amenability of a Fell bundle that arises in a canonical way whenever there is a controlled semigroup homomorphism from P to another right LCM semigroup that sits inside a group G. In particular, if G can be chosen to be amenable, then any well-aligned ideal in a right-tensor C ∗-precategory over P is amenable. As we show in Corollary 8.6 this applies to a large class of semigroups obtained from free products of right LCM semigroups. L(K). A novel ingredient in our study is that we identify an algebraic condition which character- izes when a representation Φ ⋊ P of N T L(K) is injective on the core: we call this Toeplitz covariance. Such a condition was previously considered only in the case P = N, in the context of relative Cuntz-Pimsner and relative Doplicher-Roberts algebras, cf. [19]. In Corollary 6.4 we prove that T is Toeplitz covariant, which equivalently means that T ⋊ P is injective on the core of N T L(K). The main technical result needed to achieve these characterizations is Theorem 6.1. It says that any controlled semigroup homomorphism from P to an arbitrary right LCM semigroup induces a C ∗-subalgebra of N T L(K) containing the core, and gives conditions characterising when Φ ⋊ P is injective on the induced C ∗-algebra. This result is inspired by [25, Lemma 4.1] and the proof of [14, Theorem 8.1]. Nevertheless, since we deal here with much more general situation our proof requires some new non-trivial steps. Another new ingredient in our approach is related to a potential existence of invertible elements in a right LCM semigroup P . We show that for any well-aligned ideal K in a right- tensor C ∗-precategory L the restriction of right tensoring to the group of invertible elements P ∗ preserves K. In Definition 10.8 we introduce the aperiodicity condition for this action of P ∗ on K. If P ∗ = {e} is trivial, this condition becomes vacuous. If P ∗ = P , that is, when P is a group, then K can be viewed as a Fell bundle B over P , and aperiodicity of K is equivalent to aperiodicity of B (see Section 12). By [21], if Be is separable or of Type I, apperiodicity of B is equivalent to topological freeness of the dual partial action. Non-trivial examples where the aperiodicity condition holds come for instance from wreath products, see [20]. Our main results are the uniqueness theorems in Section 10 and Section 11. For a Nica covariant representation Φ of a well-aligned ideal K in a right-tensor C ∗-precategory L we identify two sources from which to extract characterizations of injectivity of Φ ⋊ P . In Definition 10.1 we introduce a geometric condition on Φ, which we call condition (C). It is closely related to the condition describing injectivity of representations of the Toeplitz algebra of a single C ∗-correspondence, see [17, Theorem 2.1] or the condition for semigroup 5 5 / / 4 BARTOSZ K. KWAŚNIEWSKI AND NADIA S. LARSEN crossed products twisted by a product system, see [14, Equation (7.2)]. Theorem 10.12 is the main technical result on Φ that links condition (C), injectivity of Φ ⋊ P on the core, Toeplitz covariance, and generation of an exotic Nica-Toeplitz algebra. Our first uniqueness result, Corollary 10.14, says that when K is amenable and the action of P ∗ on K is aperiodic, then condition (C) on a representation Φ of K implies injectivity of Φ ⋊ P . The converse holds if K is right-tensor invariant. The most satisfactory uniqueness result is contained in Section 11. Here we reveal the true nature of condition (C). Under natural assumptions we show that the C ∗-algebra N T L(K) associated to K is a subalgebra of the C ∗-algebra N T (L) := N T L(L) associated to L (the latter can be thought of as Doplicher-Roberts version of the former). Moreover, every Nica covariant representation Φ of K extends uniquely to a Nica covariant representation Φ of L. We prove, see Corollary 11.5, that condition (C) for Φ is equivalent to injectivity of the representation Φ ⋊ P of the (Doplicher-Roberts) C ∗-algebra N T (L). This also implies injectivity of Φ ⋊ P on N T L(K), as Φ ⋊ P is a restriction of Φ ⋊ P . In addition condition (C) is equivalent to Φ being Toeplitz covariant and injective, cf. Theorem 10.15. It seems that in general, the geometric condition (C) is responsible for uniqueness of the C ∗-algebra associated to L while uniqueness of C ∗-algebra associated to K should be related with the algebraic condition of Toeplitz covariance, cf. also [20]. 1.1. Acknowledgements. The research leading to these results has received funding from the European Union's Seventh Framework Programme (FP7/2007-2013) under grant agree- ment number 621724. B.K. was partially supported by the NCN (National Centre of Science) grant number 2014/14/E/ST1/00525. This work was finished while he participated in the Simons Semester at IMPAN - Fundation grant 346300 and the Polish Government MNiSW 2015-2019 matching fund. Part of the work was done during the participation of both authors in the program "Classification of operator algebras: complexity, rigidity, and dynamics" at the Mittag-Leffler Institute (Sweden), and during the visit of N.L. at the University of Victoria (Canada). She thanks Marcelo Laca and the department at UVic for hospitality. 2. Preliminaries 2.1. LCM semigroups. We refer to [7] and [5] and the references therein for general facts about LCM semigroups. Throughout this paper P is a left cancellative semigroup with the identity element e. We let P ∗ be the group of units, or invertible elements, in P , where x ∈ P is invertible if there exists (a necessarily unique) x−1 ∈ P such that xx−1 = x−1x = e. A principal right ideal in P is a right ideal in P of the form pP = {ps : s ∈ P } for some p ∈ P . Occasionally, we will write hpi := pP . The relation of inclusion on the principal right ideals induces a left invariant preorder on P given by p ≤ q def ⇐⇒ qP ⊆ pP ⇐⇒ ∃r ∈ P q = pr. This preorder is a partial order if and only if P ∗ = {e}. Left cancellation in P implies that for fixed p, q ∈ P , pr = q determines r ∈ P uniquely, motivating the notation: p−1q := r if q = pr. The following property of semigroups is sometimes called Clifford's condition [27], [32]. Definition 2.1. A semigroup P is a right LCM semigroup if it is left cancellative and the family {pP }p∈P of principal right ideals extended by the empty set is closed under intersec- tions, that is if for every pair of elements p, q ∈ P we have pP ∩ qP = ∅ or pP ∩ qP = rP for some r ∈ P . NICA-TOEPLITZ ALGEBRAS FOR RIGHT TENSOR C∗-PRECATEGORIES 5 In the case that pP ∩ qP = rP , the element r is a right least common multiple (LCM) of p and q. Note that a right LCM is determined by p and q up to multiplication from the right by an invertible element. Namely, if pP ∩ qP = rP , then pP ∩ qP = tP if and only if there is x ∈ P ∗ such that t = rx. If P is a right LCM semigroup we will refer to J(P ) := {pP }p∈P ∪ {∅} as the semilattice of principal right ideals of P , see [7] and [28]. Example 2.2. One of the most known and studied examples of right LCM semigroups are positive cones in quasi-lattice ordered groups, introduced by Nica [31]. More precisely, suppose that P is a subsemigroup of a group G such that P ∩ P ∗ = {e}. Then the partial order we defined on P extends to a left-invariant partial order on G where g ≤ h if and only if g−1h ∈ P for all g, h ∈ G. The pair (G, P ) is a quasi-lattice ordered group if every pair of elements g, h ∈ G that has an upper bound in P admits a least upper bound in P . Note that if the upper bound exists, it is unique since P ∗ = {e}. In [13, Definition 32.1], Exel calls the pair (G, P ) weakly quasi-lattice ordered if for each pair of elements g, h ∈ P with an upper bound in P there exists a (necessarily unique) least upper bound in P . Every positive cone P in a weakly quasi-lattice ordered group (G, P ) is an LCM semigroup with P ∗ = {e}. Conversely, if P is an LCM subsemigroup of a group G such that P ∗ = {e} then (G, P ) is a weakly quasi-lattice ordered group. LCM semigroup with semilattice of principal right ideals isomorphic to the direct sum of i∈I Pi is a right LCM semigroup. This follows from the proof of the following proposition, which is a generalization of [25, Proposition 4.3]. It links the two aforementioned constructions by a useful homomorphism. Let Pi, i ∈ I, be a family of right LCM semigroups. The direct sum Li∈I Pi is a right semilattices J(Pi), i ∈ I. It is slightly less obvious that also the free product Q∗ Proposition 2.3. Let Pi, i ∈ I, be a family of right LCM semigroups. Put P := Q∗ and P :=Li∈I Pi, and let θ : P → P be the homomorphism which is the identity on each Pi, Proof. Note that P ∗ = Q∗ i ∈ I. Then θ(P ∗) = P∗ and for any s, t, r ∈ P with sP ∩ tP = rP we have θ(s) = θ(t) implies s = t. i . Hence θ(P ∗) = P∗. Now, let s = pi1 · · · pin ∈ P be in reduced form, that is pik ∈ Pik and ik 6= ik+1 for all k = 1, . . . , n − 1. Similarly, write t = qj1 · · · qjm ∈ P . Without loss of generality we may assume that m ≥ n. Suppose that sP ∩ tP 6= ∅. This implies that pi1 · · · pin−1 = qi1 · · · qin−1, in = jn and either i and P∗ = Li∈I P ∗ i∈I P ∗ i∈I Pi (2.1) θ(s)P ∩ θ(t)P = θ(r)P and (1) pin ≤ qjn, if m > n , in which case sP ∩ tP = tP , or (2) m = n and pinPin ∩qjnPin = rinPin for some rin ∈ Pin, in which case sP ∩tP = srinP . Clearly, in both cases, we have θ(s)P ∩ θ(t)P = θ(r)P. Moreover, θ(s) = θ(t) implies that m = n and pin = qjn, that is, s = t. (cid:3) The property identified in (2.1) is important. We will formalise it in a definition. To indicate the analogy with similar concepts introduced in [25] and [9], we borrow Crisp and Laca's terminology of controlled map, see [9, Definition 4.1] and in particular conditions (C2), (C3) and (C5), but add the tag of right LCM semigroup. Definition 2.4. A controlled map of right LCM semigroups is an identity preserving homo- morphism θ : P → P between right LCM semigroups P, P such that θ(P ∗) = P∗ and for all s, t ∈ P with sP ∩ tP 6= ∅ we have (2.2) θ(s)P ∩ θ(t)P = θ(r)P whenever r is a right LCM for s, t 6 BARTOSZ K. KWAŚNIEWSKI AND NADIA S. LARSEN and (2.3) θ(s) = θ(t) =⇒ s = t. Remark 2.5. (a) If θ : P → P is a controlled map of right LCM semigroups and P is right cancellative, then so is P . Indeed, if pr = qr in P then θ(p)θ(r) = θ(q)θ(r), so right cancellation in P implies θ(p) = θ(q) giving p = q by condition (2.3). In our applications, P will be contained in a group so it will necessarily be cancellative. (b) A controlled map of right LCM semigroups seems to be different from a homomorphism of right LCM semigroups as considered in [6, Equation (3.1)]. This is because a homomorphism of right LCM semigroups need not preserve the group of units and, besides, it keeps track of pairs of elements in P and their images in P that do not have a right common upper bound as well as those who do. 2.2. C ∗-precategories. We recall some background on C ∗-precategories from [19]. C ∗- precategories should be viewed as non-unital versions of C ∗-categories, cf. [18], [10]. A precategory L consists of a set of objects Ob(L) and a collection {L(σ, ρ)}σ,ρ∈Ob(L) of sets of morphisms endowed with an associative composition. Explicitly, L(σ, ρ) stands for the space of morphisms from ρ to σ, and the composition L(τ, σ) × L(σ, ρ) → L(τ, ρ), (a, b) → ab must satisfy (ab)c = a(bc) whenever the compositions of morphisms a, b, c are allowable. A morphism in L(σ, ρ) may be regarded as an arrow from ρ to σ. One can equip (if necessary) L(σ, σ) with identity morphisms in such a way that a given precategory L becomes a category. In the sequel, we will identify L with the collection of morphisms {L(σ, ρ)}σ,ρ∈Ob(L). Definition 2.6. ([19, Definition 2.2]) A C ∗-precategory is a precategory L = {L(σ, ρ)}σ,ρ∈Ob(L) together with an operation ∗ : L → L such that the following hold: (p1) each set of morphisms L(σ, ρ), σ, ρ ∈ L is a complex Banach space; (p2) composition gives a bilinear map L(τ, σ) × L(σ, ρ) ∋ (a, b) → ab ∈ L(τ, ρ), which satisfies kabk ≤ kak · kbk; (p3) for each σ, ρ ∈ L, a → a∗ is an antilinear map from L(σ, ρ) to L(ρ, σ) such that (a∗)∗ = a and (ab)∗ = b∗a∗ for all b ∈ L(ρ, τ ) and all τ ∈ L; (p4) ka∗ak = kak2 for every a ∈ L(σ, ρ); and (p5) for each a ∈ L(σ, ρ), we have a∗a = b∗b for some b ∈ L(ρ, ρ). We say that a C ∗-precategory L is C ∗-category if L is a category. Note that (p3) says that the operation ∗ is antilinear, involutive and contravariant. More- over, (p1) -- (p4) imply that L(ρ, ρ) is a C ∗-algebra, and (p5) says that a∗a is positive as an element of L(ρ, ρ). Condition (p4) implies that the operation ∗ is isometric on every space L(σ, ρ). A C ∗-precategory L is a C ∗-category if and only if every C ∗-algebra L(ρ, ρ), ρ ∈ Ob(L), is unital. Lemma 2.7. Given a C ∗-precategory L and objects σ, ρ we have L(σ, ρ) = L(σ, σ)L(σ, ρ) = L(σ, ρ)L(ρ, ρ). Proof. In a natural way, L(σ, ρ) is a left Banach L(σ, σ)-module and a right Banach L(ρ, ρ)- module. By [19, Lemma 2.5], these module actions are non-degenerate. Hence the assertion follows by the Cohen-Hewitt factorization theorem. (cid:3) NICA-TOEPLITZ ALGEBRAS FOR RIGHT TENSOR C∗-PRECATEGORIES 7 Definition 2.8 (Definition 2.4 in [19]). An ideal in a C ∗-precategory L is a collection K = {K(σ, ρ)}σ,ρ∈Ob(L) of closed linear subspaces K(σ, ρ) of L(σ, ρ), ρ, σ ∈ Ob(L), such that L(τ, σ)K(σ, ρ) ⊆ K(τ, ρ) and K(τ, σ)L(σ, ρ) ⊆ K(τ, ρ), for all σ, ρ, τ ∈ Ob(L). An ideal K in a C ∗-precategory L is automatically selfadjoint in the sense that K(σ, ρ)∗ = [18, Proposition 1.7]. Hence K is a C ∗-precategory. Each K(ρ, σ), for all σ, ρ ∈ Ob(L), cf. space K(ρ, ρ) is a closed two-sided ideal in the C ∗-algebra L(ρ, ρ). A useful fact is that K is uniquely determined by these diagonal ideals. Proposition 2.9 (Theorem 2.6 in [19]). If K is an ideal in a C ∗-precategory L, then for all σ and ρ the space K(σ, ρ) coincides with (2.4) {a ∈ L(σ, ρ) : a∗a ∈ K(ρ, ρ)} = {a ∈ L(σ, ρ) : aa∗ ∈ K(σ, σ)}. Conversely, a collection of ideals K(ρ, ρ) in L(ρ, ρ), for ρ ∈ Ob(L), satisfying (2.4) gives rise to an ideal K in L. We generalize the notion of an essential ideal in a C ∗-algebra as follows. Definition 2.10. An ideal K in a C ∗-precategory L is an essential ideal in L if (2.5) K(ρ, ρ) is an essential ideal in L(ρ, ρ) for every ρ ∈ Ob(L). Homomorphisms between C ∗-precategories are defined in a natural way. We recall from [19, Definition 2.8] that a homomorphism Φ from a C ∗-precategory L to a C ∗-precategory consists of a map Ob(L) ∋ σ 7→ Φ(σ) ∈ Ob() and linear operators L(σ, ρ) ∋ a 7→ Φ(a) ∈ (Φ(σ), Φ(ρ)), σ, ρ ∈ Ob(L), such that Φ(a)Φ(b) = Φ(ab) and Φ(a∗) = Φ(a)∗ for all a ∈ L(τ, σ), b ∈ L(σ, ρ) and all σ, ρ, τ ∈ Ob(L). An endomorphism of L is a homomorphism from L to L. Note that composition of homomorphisms is again a homomorphism. If Φ : L → is a homo- morphism between C ∗-precategories, then Φ : L(ρ, ρ) → (Φ(ρ), Φ(ρ)) are ∗-homomorphisms of C ∗-algebras. Using this observation one gets the following fact, cf. [19, Proposition 2.9]. Lemma 2.11. For any homomorphism Φ : L → of C ∗-precategories the operators (2.6) Φ : L(σ, ρ) → (Φ(σ), Φ(ρ)), σ, ρ ∈ Ob(L), are contractions. Moreover, all the maps in (2.6) are isometric if and only if all the maps in (2.6) with σ = ρ are injective. Definition 2.12. A representation of a C ∗-precategory L in a C ∗-algebra B is a homomor- phism Φ : L → B where B is considered as a C ∗-precategory with a single object. Equiva- lently, Φ may be viewed as a collection {Φσ,ρ}σ,ρ∈Ob(L) of linear operators Φσ,ρ : L(σ, ρ) → B such that Φσ,ρ(a)∗ = Φρ,σ(a∗), and Φτ,ρ(ab) = Φτ,σ(a)Φσ,ρ(b), for all a ∈ L(τ, σ), b ∈ L(σ, ρ). A representation of L on a Hilbert space H is a representation of L in the C ∗-algebra B(H) of all bounded operators on H. We say that a representation {Φσ,ρ}σ,ρ∈Ob(L) is injective if all Φρ,ρ for ρ ∈ Ob(L) are injective (then all the maps Φσ,ρ, σ ∈ Ob(L), are isometric by Lemma 2.11). Let {Φσ,ρ}σ,ρ∈L be a representation of a C ∗-precategory L. If K is an ideal in L then {Φσ,ρK(σ,ρ)}σ,ρ∈L is a representation of K. In the converse direction we have the following result, cf. [19, Proposition 2.13]. 8 BARTOSZ K. KWAŚNIEWSKI AND NADIA S. LARSEN Proposition 2.13. Suppose that K is an ideal in a C ∗-precategory L and let Φ = {Φσ,ρ}σ,ρ∈L be a representation of K on a Hilbert space H. There is a unique extension Φ = {Φσ,ρ}σ,ρ∈L of Φ to a representation of L such that the essential subspace of Φσ,ρ is contained in the essential subspace of Φσ,ρ, for every σ, ρ ∈ Ob(L). Namely, (2.7) Φσ,ρ(a)(Φρ,ρ(K(ρ, ρ))H)⊥ = 0, and Φσ,ρ(a)Φρ,ρ(b)h = Φσ,ρ(ab)h for all a ∈ L(σ, ρ), b ∈ K(ρ, ρ), h ∈ H. Moreover, (2.8) (ker Φ)(σ, ρ) = {a ∈ L(σ, ρ) : aK(ρ, ρ) ⊆ ker Φσ,ρ}. In particular, Φ is injective if and only if Φ is injective and K is an essential ideal in L. Proof. The existence and uniqueness of Φ satisfying (2.7) are guaranteed by [19, Proposition 2.13]. Let a ∈ L(σ, ρ), σ, ρ ∈ Ob(L). By (2.7), we have a ∈ ker Φσ,ρ ⇐⇒ Φσ,ρ(aK(ρ, ρ)) = {0} =⇒ aK(ρ, ρ) ⊆ ker Φσ,ρ, which proves the second part of the assertion. (cid:3) Example 2.14. The prototypical examples of C ∗-precategories arise from adjointable maps between Hilbert modules. Specifically, let X = {Xp}p∈S be a family of right Hilbert modules over a C ∗-algebra A, indexed by a set S. Then the families K(p, q) := K(Xq, Xp), L(p, q) := L(Xq, Xp) for p, q ∈ S with operations inherited from the corresponding spaces form C ∗-precategories. In fact, L is a C ∗-category and K is an essential ideal in L. 3. Nica covariant representations and Nica-Toeplitz algebra Right-tensor C ∗-precategories over the semigroup N were introduced in [19], where they were shown to provide a good framework for studying Pimsner and Doplicher-Roberts type C ∗-algebras. Here we notice that [19, Definition 3.1] makes sense for an arbitrary semigroup P , and we set out to study associated C ∗-algebras. Definition 3.1. A right-tensor C ∗-precategory is a C ∗-precategory L = {L(p, q)}p,q∈P whose objects form a semigroup P with identity e and which is equipped with a semigroup {⊗1r}r∈P of endomorphisms of L such that ⊗1r acts on P by sending p to pr, for all p, r ∈ P , and ⊗1e = id. For a morphism a ∈ L(p, q) we denote the value of ⊗1r on a by a⊗1r, and note that it belongs to L(pr, qr). We refer to {⊗1r}r∈P as to a right tensoring on L = {L(p, q)}p,q∈P . Note in particular that for all a ∈ L(p, q), b ∈ L(q, s), and p, q, r, s ∈ P we have ((a ⊗ 1r) ⊗ 1s) = a ⊗ 1rs, (a ⊗ 1r)∗ = a∗ ⊗ 1r, (a ⊗ 1r)(b ⊗ 1r) = (ab) ⊗ 1r. The following definition is a semigroup generalization of [19, Definition 3.6]. Definition 3.2. Let K be an ideal in a right-tensor C ∗-precategory T . We say that a rep- resentation Φ : K → B of K in a C ∗-algebra B is a right-tensor representation if for all a ∈ K(p, q) and b ∈ K(s, t) such that sP ⊆ qP we have (3.1) Φ(a)Φ(b) = Φ(cid:0)(a ⊗ 1q−1s)b(cid:1) . We let C ∗(Φ(K)) be the C ∗-algebra generated by the spaces Φ(K(p, q)), p, q ∈ P . We call it the C ∗-algebra generated by Φ. NICA-TOEPLITZ ALGEBRAS FOR RIGHT TENSOR C∗-PRECATEGORIES 9 Remark 3.3. Since K is an ideal the right hand side of (3.1) makes sense. Furthermore, by taking adjoints one gets the symmetrized version of this equation: where a ∈ K(p, q), b ∈ K(s, t), qP ⊆ sP , p, q, s, t ∈ P . Φ(a)Φ(b) = Φ(a(b ⊗ 1s−1q)), Standard arguments coupled with our proof of existence of an injective Nica covariant representation, see Proposition 5.2 below, show that the following proposition holds. Proposition 3.4. Let K be an ideal in a right-tensor C ∗-precategory L. There are a C ∗- algebra TL(K) and an injective right-tensor representation tK : K → TL(K), such that (a) for every right-tensor representation Φ of K there is a homomorphism Φ × P of TL(K) such that (Φ × P ) ◦ tK = Φ; and (b) TL(K) = C ∗(tK(K)). The C ∗-algebra TL(K) is unique up to canonical isomorphism. Proof. Since every representation Φ : K → B is automatically contractive, cf. Lemma 2.11, a direct sum of right-tensor representations of K is a right-tensor representation. Thus existence and uniqueness of TL(K) follow from [3, Section 1]. Injectivity of tK follows from Proposition 5.2 that we prove below. (cid:3) Definition 3.5. Given an ideal K in a right-tensor C ∗-precategory (L, {⊗1r}r∈P ), the C ∗- algebra TL(K) described in Proposition 3.4 is the Toeplitz algebra of K. The Toeplitz algebra TL(K) in general is very large. It lacks a version of 'Wick ordering' and therefore its structure is hardly accessible. This is the main reason why in the present paper we will study C ∗-algebras generated by representations satisfying a condition of Nica type, which is stronger than (3.1). Since such conditions are (so far) established only for right LCM semigroups, from now on (with the exception of Section 4) we will always assume that P is a right LCM semigroup. Definition 3.6. Let (L, {⊗1r}r∈P ) be a right-tensor C ∗-precategory over a right LCM semi- group P . An ideal K in L is well-aligned in (L, {⊗1r}r∈P ) if for all a ∈ K(p, p), b ∈ K(q, q) we have (3.2) (a ⊗ 1p−1r)(b ⊗ 1q−1r) ∈ K(r, r) whenever pP ∩ qP = rP. An ideal K in L is ⊗1-invariant if K(p, p) ⊗ 1r ⊆ K(pr, pr) for all p, r ∈ P . We denote this property of K as K ⊗ 1 ⊆ K. Note that by Proposition 2.9, if K ⊗ 1 ⊆ K, then K(p, q) ⊗ 1r ⊆ K(pr, qr) for all p, q, r ∈ P . Plainly, if K is a ⊗1-invariant ideal in a right-tensor C ∗-precategory L, then K is itself a right- tensor C ∗-precategory, and K is well-aligned both in L and in K. The condition in (3.2) is a generalization of the notion of compact alignment for product systems of C ∗-correspondences from [14, Definition 5.7], cf. [6], and see [20] for details. The next lemma shows that (3.2) captures more than just diagonal fibres K(p, p) for p ∈ P . Lemma 3.7. Let K be a well-aligned ideal in a right-tensor C ∗-precategory L. For all a ∈ K(p, q), b ∈ K(s, t) we have (a ⊗ 1q−1r)(b ⊗ 1s−1r) ∈ K(pq−1r, ts−1r) (3.3) Proof. By Lemma 2.7 we have a = a′a′′ and b = b′b′′ where a′ ∈ K(p, q), b′ ∈ K(s, t) and a′′ ∈ K(q, q), b′′ ∈ K(s, s). If qP ∩ sP = rP then by (3.2) we have (a′′ ⊗ 1q−1r)(b′′ ⊗ 1s−1r) ∈ K(r, r). Composing this from the left by (a′ ⊗ 1q−1r) ∈ L(pq−1r, r) and from the right by (b′ ⊗ 1s−1r) ∈ L(r, ts−1r) and using that K is an ideal in L gives (3.3). (cid:3) whenever qP ∩ sP = rP. 10 BARTOSZ K. KWAŚNIEWSKI AND NADIA S. LARSEN The notion of Nica covariance for a representation of a compactly aligned product system of C ∗-correspondences was introduced in [14] in the context of quasi-lattice ordered groups, and was extended to right LCM semigroups in [6]. Lemma 3.7 allows us to extend this concept to C ∗-precategories over right LCM semigroups. In our generalization below, Nica covariance will be imposed in subspaces K(p, q) for p, q ∈ P that are not necessarily diagonal, in the sense that p need not equal q. Definition 3.8. Let K be a well-aligned ideal in a right-tensor C ∗-precategory L. A repre- sentation Φ : K → B of K in a C ∗-algebra B is Nica covariant if for all a ∈ K(p, q), b ∈ K(s, t) we have Φ(a)Φ(b) =(Φ(cid:0)(a ⊗ 1q−1r)(b ⊗ 1s−1r)(cid:1) 0 if qP ∩ sP = rP for some r ∈ P, otherwise. (3.4) (3.5) Remark 3.9. If Φ is a Nica covariant representation of a well-aligned ideal K, then Φ is a right-tensor representation and moreover the space C ∗(Φ(K))0 := span{ [p,q∈P Φ(K(p, q))} is a dense ∗-subalgebra of C ∗(Φ(K)); it is clearly closed under taking adjoints and it is closed under multiplication by (3.4). Hence C ∗(Φ(K)) = span{Sp,q∈P Φ(K(p, q))}. Since a right-tensor C ∗-precategory L is well-aligned in itself we may always talk about Nica covariant representations of L. For every Nica covariant representation Φ : L → B of L and every well-aligned ideal K in L the restriction Φ : K → B is a Nica covariant representation of K. Moreover, (3.4) readily implies that C ∗(Φ(K)) is a C ∗-subalgebra of C ∗(Φ(L)), and if K ⊗ 1 ⊆ K, then C ∗(Φ(K)) is in fact an ideal in C ∗(Φ(L)). Since the element r in the right hand side of (3.4) is determined only up to invertible elements in P ∗, Nica covariant representations behave in a special way with respect to ⊗1x, x ∈ P ∗. Lemma 3.10. Let K be a well-aligned ideal in a right-tensor C ∗-precategory L. For every x ∈ P ∗, ⊗1x is an automorphism of L which maps K(p, q) onto K(px, qx) for every p, q ∈ P , thus it restricts to an automorphism of K. Moreover, if x ∈ P ∗ and Φ : K → B is a Nica covariant representation of K, then Φ(a) = Φ(a ⊗ 1x) for all a ∈ K(p, q), p, q ∈ P . Proof. Let x ∈ P ∗. Clearly, ⊗1x is an automorphism of L since ⊗1x−1 acts as an inverse. Take a ∈ K(p, q). As in the proof of Lemma 3.7, write a = a′a′′ where a′ ∈ K(p, q) and a′′ ∈ K(q, q). Since qP = qxP , by (3.3) we get a ⊗ 1x = (a′a′′) ⊗ 1x = (a′ ⊗ 1x)(a′′ ⊗ 1x) ∈ K(px, qx). If Φ : K → B is a Nica covariant representation of K, using (3.4) we get Φ(a) = Φ(a′)Φ(a′′) = Φ((a′ ⊗ 1x)(a′′ ⊗ 1x)) = Φ(a ⊗ 1x). (cid:3) Existence (and uniqueness) of the universal C ∗-algebra described in the following propo- sition can be shown as in the proof of Proposition 3.4, or by considering a quotient of the Toeplitz algebra TL(K). Injectivity of the universal Nica covariant representation follows from Proposition 5.2 below. Proposition 3.11. Let K be a well-aligned ideal in a right-tensor C ∗-precategory L. There are a C ∗-algebra N T L(K) and an injective Nica covariant representation iK : K → N T L(K), such that (a) for every Nica covariant representation Φ of K there is a homomorphism Φ ⋊ P of N T L(K) such that (Φ ⋊ P ) ◦ iK = Φ; and NICA-TOEPLITZ ALGEBRAS FOR RIGHT TENSOR C∗-PRECATEGORIES 11 (b) N T L(K) = C ∗(iK(K)). The C ∗-algebra N T L(K) is unique up to canonical isomorphism. Definition 3.12. Given a well-aligned ideal K in a right-tensor C ∗-precategory L, the C ∗- algebra N T L(K) described in Proposition 3.11 is called the Nica-Toeplitz algebra of K. We write N T (L) for the Nica-Toeplitz algebra N T L(L) associated to L, viewed as a well-aligned ideal in itself. Remark 3.13. The universal property of N T L(K) ensures existence of a homomorphism (3.6) ι : N T L(K) 7−→ N T (L), iK(a) 7−→ iLK(a), for a ∈ K(p, q) and p, q ∈ P . We note that ι is injective whenever the universal representation iK : K → N T L(K) can be extended to a Nica covariant representation iK : L → B where B is a C ∗-algebra containing N T L(K). Indeed, in this case we have (iK ⋊ P ) ◦ iL = iK, from which it follows that (iK ⋊ P ) ◦ ι = idN T L(K), showing the claimed injectivity. We will explore these issues in more detail in Section 11. Remark 3.14. If K ⊗ 1 ⊆ K, then N T L(K) does not depend on L. in this case, the definition of Nica covariant representations involves only elements of K. There- fore, with N T (K) denoting the Nica-Toeplitz algebra of the right-tensor C ∗-category K, we have N T (K) = N T L(K). In general, N T L(K) depends only on the C ∗-precategory structure of K equipped with a family of mappings {Nr}r∈P where Nr, r ∈ P , is defined for quadruples p, q, s, t ∈ P such that qP ∩ sP = rP by the formula Indeed, K(p, q) × K(s, t) ∋ (a, b) 7−→ Nr(a, b) := (a ⊗ 1q−1r)(b ⊗ 1s−1r) ∈ K(pq−1r, ts−1r). Note that Nr(a, b)∗ = Nr(a∗, b∗) and Ne(a, b) = ab if q = s. Moreover, for any c ∈ K(u, w) where uP ∩ ts−1rP = zP , for some z ∈ P , we have Nz(Nr(a, b), c) = Ny−1z(a, Ny(b, c)) for any y ∈ P such that tP ∩uP = yP (we then necessarily have qP ∩st−1yP = y−1zP ). Nev- ertheless, in this paper we will not pursue this more general intrinsic description of N T L(K) for three reasons. Firstly, such a theory would be technically more involved. Secondly, we do not have good examples that require such an approach. Thirdly, the relationship between N T L(K) and N T (L) is interesting in its own right (in the context of Doplicher-Roberts al- gebras such a relationship was studied, for instance, in [11], [15], [19]). The latter problem, in our setting, will be addressed in Section 11. Obviously, the Nica-Toeplitz algebra N T L(K) may be viewed as a quotient of the Toeplitz algebra TL(K). If every two elements in P are comparable, then every ideal K in a right-tensor C ∗-precategory L over P is automatically well-aligned and right-tensor representations of K coincide with Nica covariant representations. Hence N T L(K) ∼= TL(K) in this case. Example 3.15 (The case when P = N). Let L be a right-tensor C ∗-precategory over N and K an ideal in L. Due to above discussion K is automatically well-aligned, and a representation Φ of K is Nica covariant if and only if Φ is a right-tensor representation. For any ideal J in J(K) := ⊗1−1(K) ∩ K, C ∗-algebras OL(K, J ) were introduced in [19] as universal C ∗-algebras with respect to right-tensor representations Φ of K satisfying Φn,m(a) = Φn+1,m+1(a ⊗ 1) for all a ∈ J (n, m) and n, m, ∈ N. Therefore, and every C ∗-algebra OL(K, J ) is a quotient of N T L(K). TL(K) ∼= N T L(K) ∼= OL(K, {0}) 12 BARTOSZ K. KWAŚNIEWSKI AND NADIA S. LARSEN Example 3.16 (Product systems). Let X =Fp∈P Xp be a product system as defined in [14]. In [20, Section 2.1], cf. the introduction, we associate to X the right-tensor C ∗-precategory LX. Then KX = {K(Xp, Xq)}p,q∈P is an essential ideal in LX . By [20, Proposition 2.8] we have a natural isomorphism TLX (KX ) ∼= T (X) where T (X) is the Toeplitz algebra of X defined in [14]. Assume that P is a right LCM semigroup. Then KX is well-aligned if and only if X is compactly aligned. In this case, [20, Proposition 2.10] gives N T LX (KX ) ∼= N T (X), where N T (X) is the Nica-Toeplitz algebra associated to X, see [6], [14]. In [20] we also analyze a Doplicher-Roberts version DR(N T (X)) of N T (X), which by definition is N T (LX). In view of the following lemma we may always assume that a well-aligned ideal K in a right-tensor C ∗-precategory L generates L as a right-tensor C ∗-precategory. Lemma 3.17. For every well-aligned ideal K in a right-tensor C ∗-precategory L the spaces LK(p, q) = span{K(s, t) ⊗ 1r : sr = p, tr = q, for s, t, r ∈ P }, p, q ∈ P, define the minimal right-tensor sub-C ∗-precategory of L containing K. In particular, we have N T L(K) ∼= N T LK(K). Proof. Plainly, the family {LK(p, q)}p,q∈P is closed under right-tensoring ⊗1 and under taking adjoints. Suppose that a ⊗ 1r ∈ LK(p, q) and b ⊗ 1w ∈ K(q, z) where a ∈ K(s, t), b ∈ K(u, v), sr = p, tr = q, uw = q, vw = z. Then tP ∩ uP = yP for some y ∈ P , which implies that (t−1y)−1r = y−1q and (u−1y)−1w = y−1q. Since K is well-aligned we have (a ⊗ 1t−1y)(b ⊗ 1u−1y) ∈ K(st−1y, vu−1y). Using this we obtain (a ⊗ 1r)(b ⊗ 1w) =(cid:16)(a ⊗ 1t−1y) ⊗ 1y−1q(cid:17)(cid:16)(b ⊗ 1u−1y) ⊗ 1y−1q(cid:17) =(cid:16)(a ⊗ 1t−1y)(b ⊗ 1u−1y)(cid:17) ⊗ 1y−1q ∈ LK(p, z). Hence {LK(p, q)}p,q∈P is a right-tensor sub-C ∗-precategory of L. Clearly, it is the smallest sub-C ∗-precategory of L containing K and invariant under ⊗1. (cid:3) An important role in the theory is played by the following core C ∗-algebra. Definition 3.18. Let K be a well-aligned ideal in a right-tensor C ∗-precategory L. For an arbitrary Nica covariant representation Φ of K the space BΦ e := spann[p∈P Φ(K(p, p))o is a C ∗-algebra. We call BΦ e the core C ∗-subalgebra of C ∗(Φ(K)). Remark 3.19. For any Nica covariant representation Φ : K → B the core C ∗-algebra BΦ e is a non-degenerate subalgebra of C ∗(Φ(K)). Indeed, by Lemma 2.7, every a ∈ K(p, q) can be written as a = apa′aq where ap ∈ K(p, p), a′ ∈ K(p, q), aq ∈ K(q, q). Hence Φ(a) ∈ BΦ e extends via the formula mΦ(a) := (mΦ(ap))Φ(a′aq) to a multiplier of C ∗(Φ(K)). We will use this embedding in the sequel to identify M (BΦ e . In particular, every multiplier m ∈ M (BΦ e ) as a subalgebra of M (C ∗(Φ(K))). e C ∗(Φ(K))BΦ e ) of BΦ NICA-TOEPLITZ ALGEBRAS FOR RIGHT TENSOR C∗-PRECATEGORIES 13 4. Fock representation Only for the purposes of this section we fix an arbitrary left cancellative semigroup P , a right-tensor C ∗-precategory (L, {⊗1r}r∈P ) and an arbitrary ideal K in L. We will construct a canonical right-tensor representation of L associated to K, whose restriction to K is injective. This construction will proceed in two steps. First we associate a representation to each t ∈ P , regarded as a fixed source, and then we consider the direct sum of these representations as we vary t. We fix t ∈ P . For each s ∈ P the space Xs,t := K(s, t) is naturally equipped with a structure of a right Hilbert module over At := K(t, t) given by x · a := xa, hx, yi := x∗y, x, y ∈ Xs,t, a ∈ At. Thus we may consider the following direct sum right Hilbert At-module: F t K :=Ms∈P Xs,t. We will construct representations of L using the maps defined in the following lemma. Lemma 4.1. Let a ∈ L(p, q) for p, q ∈ P . For each s ∈ qP there is a well defined operator T s,t p,q(a) ∈ L(Xs,t, Xpq−1s,t) given by T s,t p,q(a)x := (a ⊗ 1q−1s)x, x ∈ Xs,t. The adjoint is T pq−1s,t q,p (a∗) ∈ L(Xpq−1s,t, Xs,t). Proof. Let x ∈ Xs,t and y ∈ Xpq−1s,t. Clearly, (a ⊗ 1q−1s)x ∈ Xpq−1s,t = K(pq−1s, t), and hT s,t p,q(a)x, yi = h(a ⊗ 1q−1s)x, yi = x∗(a∗ ⊗ 1q−1s)y = hx, (a∗ ⊗ 1q−1s)yi = hx, T pq−1s,t q,p (a∗)yi. (cid:3) Let a ∈ L(p, q) for p, q ∈ P . By Lemma 4.1, under the obvious identifications of the p,q(a) ∈ p,q(a)k ≤ kak and the map qP ∋ s → pq−1s ∈ pP is a Hilbert modules Xs,t, s ∈ P , with the corresponding submodules of F t L(Xs,t, Xpq−1s,t) ⊆ L(F t bijection, the direct sum K, we have T s,t K). Since kT s,t (4.1) T t p,q(a) := Ms∈qP T s,t p,q(a) is an operator in L(F t mapping T t p,q : L(p, q) → L(F t K), which satisfies K) with norm bounded by kak. In other words, we get a contractive (4.2) T t p,q(a)x =((a ⊗ 1q−1s)x 0 if s ∈ qP, otherwise, for every a ∈ L(p, q), s ∈ P and x ∈ Xs,t. Lemma 4.2. For each t ∈ P , the family of maps T right-tensor representation T t : L → L(F t K). t = {T t p,q}p,q∈P given by (4.2) is a 14 BARTOSZ K. KWAŚNIEWSKI AND NADIA S. LARSEN Proof. By Lemma 4.1 we have (T s,t from (4.1) that T any a ∈ L(r, p), b ∈ L(p, q), x ∈ Xs,t, where r, p, q ∈ P with s ∈ qP , we have (a∗) for a ∈ L(p, q), and therefore it follows t p,q, p, q ∈ P are linear. Moreover, for t q,p(a∗). Clearly, the maps T t p,q(a)∗ = T p,q(a))∗ = T pq−1s,t q,p T pq−1s,t r,p (a)T s,t p,q(b)x = (a ⊗ 1q−1s)(b ⊗ 1q−1s)x = (ab ⊗ 1q−1s)x = T s,t r,q (ab)x. t p,q(b) = T t r,q(ab) and thus T t : L → L(F t K) is a representation of C ∗- t r,p(a)T Hence T precategories. t To see that T is a right-tensor representation let a ∈ L(p, q) and b ∈ L(s, l) for p, q, s, l ∈ P t such that sP ⊆ qP . Note that if w /∈ lP then both T pq−1s,l((a ⊗ 1q−1s)b) act as zero on Xw,t. Assume then that w ∈ lP and let x ∈ Xw,t. Then (b ⊗ 1l−1w)x ∈ Xsl−1w,t and sl−1w ∈ qP . Thus we have t p,q(a)(b ⊗ 1l−1w)x = (a ⊗ 1q−1sl−1w)(b ⊗ 1l−1w)x t s,l(b) and T t s,l(b)x = T t p,q(a)T t p,q(a)T T =(cid:16)(cid:0)(a ⊗ 1q−1s)b(cid:1) ⊗ 1l−1w(cid:17)x = T t s,l(b) and T t pq−1s,l((a ⊗ 1q−1s)b) coincide, and T t pq−1s,l((a ⊗ 1q−1s)b)x. t : L → L(FK) is a (cid:3) Accordingly, T right-tensor representation. t p,q(a)T Note that for each t ∈ P , we may view F t A :=Lp∈P Ap, where multiplication on the right by an element of the summand Ap for p 6= t is defined to be zero. We define the Fock module of K to be the direct sum Hilbert A-module of F t K as a right Hilbert module over the C ∗-algebra K as t ∈ P : FK :=Mt∈P F t K = Ms,t∈P Xs,t. Accordingly, FK consists of elementsLs,t∈P xs,t where xs,t ∈ K(s, t), s, t ∈ P , and the element Lt∈P(cid:16)Ps∈P x∗ s,txs,t(cid:17) belongs to the C ∗-algebraic direct sumLt∈P K(t, t). We will treat the C ∗-algebraic direct productQt∈P L(F t Proposition 4.3. The direct sum of representations from Lemma 4.2 as t varies in P yields a right-tensor representation T : L → L(FK) determined by the formula K) as a C ∗-subalgebra of L(FK). (4.3) T p,q(a)x =((a ⊗ 1q−1s)x 0 if s ∈ qP, otherwise, for a ∈ L(p, q), x ∈ Xs,t and p, q, s, t ∈ P . Furthermore, the restriction of T to K yields an injective right-tensor representation T : K → L(FK). Proof. It is immediate that the direct sum of right-tensor representations: T p,q :=Lt∈P T Ls∈qP,t∈P T s,t t p,q = p,q, p, q ∈ P, yields a right-tensor representation T : L → L(FK) which sat- isfies (4.3). Let T : K → L(FK) be its restriction to K. For every p ∈ P the map T p,p p,p : K(p, p) → L(Xp,p) is injective and hence Tp,p : K(p, p) → L(FK) is injective. (cid:3) Definition 4.4. We call the right-tensor representation T : K → L(FK) from Proposition 4.3 the Fock representation of the ideal K in the right-tensor C ∗-precategory L. Remark 4.5. For each t ∈ P , we may view the restriction T t : K → L(F t to K as a Fock representation of K with fixed source t. However, T t is injective if and only if for every a ∈ K(p, p) and p ∈ P there is r ∈ P such that (a ⊗ 1r)K(pr, t) 6= 0; and this may fail. K) of T t NICA-TOEPLITZ ALGEBRAS FOR RIGHT TENSOR C∗-PRECATEGORIES 15 Remark 4.6. The Fock representation is the direct sum T = Lt∈P T t of t-th Fock repre- sentations T t : K → L(F t K), t ∈ P . So by projecting, for each t ∈ P , we get a surjective homomorphism ht : C ∗(T (K)) → C ∗(T t(K)), where ht ◦ T = T t. We will show that he is an isomorphism for Fell bundles (cf. Proposition 7.6 below) and for right-tensor C ∗-precategories arising from compactly aligned product systems, see [20]. Thus our Fock representation gen- eralizes those for product systems and Fell bundles. that Qw projects onto the subspace of FK of fixed range w. The grading of the Fock Hilbert module FK yields natural conditional expectations. To K) be the make this explicit we introduce some notation. For w, t ∈ P we let Qt projection onto Xw,t, and Qw :=Ls∈P Qs (a) For each t ∈ P , the space Dt := (cid:8)S ∈ L(F t w ∈ L(FK) be the projection ontoLs∈P Xw,s. Note w for every w ∈ P(cid:9) is a Lemma 4.7. Let T : L → L(FK) be the Fock representation of K. C ∗-subalgebra of L(F t w ∈ L(F t wS = SQt K) : Qt K) and the map Et : L(F t wSQt w, K) 7→ Dt given by S ∈ L(F t Qt K), (4.4) Et(S) = Xw∈P E(S) = Xw∈P is a faithful conditional expectation. (b) The space D := {S ∈ L(FK) : QwS = SQw for every w ∈ P } is a C ∗-subalgebra of L(FK) and the map E : L(FK) 7→ D given by (4.5) QwSQw, S ∈ L(FK), K) to L(Xw,t) ⊆ L(F t is a faithful conditional expectation. Furthermore, EQt∈P L(F t Proof. For part (a), clearly Dt is a C ∗-subalgebra of L(F t L(F t direct sum of these maps is a well defined contractive completely positive map Et : L(F t Dt ⊆ L(F t claimed. That Et is faithful follows because if a ∈ L(F t every x ∈ Xw,t, w ∈ P , w from K) is a contractive completely positive map with range in Dt. The K) → K) which is the identity on Dt. Hence (4.4) defines a conditional expectation as K) is such that Et(a∗a) = 0 then for K). The map a 7→ Qt waQt K) =Lt∈P Et. kaxk2 = khax, axik = khx, a∗axik = khx, Et(a∗a)xik = 0. Since the elements x ∈ Xw,t, w ∈ P , span F t analogous to (a) and is left to the reader. K it follows that a = 0. The proof of part (b) is (cid:3) 5. Reduced Nica-Toeplitz algebra and the transcendental core In this section, we come back to our standing assumption that P is a right LCM semigroup. We fix a right-tensor C ∗-precategory (L, {⊗1r}r∈P ) and a well-aligned ideal K in L. Under these assumptions, the Fock representation is Nica covariant: Lemma 5.1. For each t ∈ P , the family of maps T covariant representation T t : L → L(F t K). t = {T t p,q}p,q∈P given by (4.2) is a Nica Proof. By Lemma 4.2 we only need to show that T b ∈ L(s, l) for p, q, s, l ∈ P . Note that if w ∈ lP and x ∈ Xw,t, then T w,t Since T u,t is Nica covariant. Let a ∈ L(p, q) and s,l (b) is in Xsl−1w,t. t p,q (a) acts in Xu,t, we have p,q (a)T w,t T u,t (5.1) s,l (b) 6= 0 =⇒ u = sl−1w and u ∈ qP. 16 BARTOSZ K. KWAŚNIEWSKI AND NADIA S. LARSEN p,q (a)T w,t In particular, T u,t Hence, if qP ∩ sP = ∅, then T the fact that ls−1rP ∋ w → sl−1w ∈ rP is a bijection, we get s,l (b) 6= 0 implies that qP ∩ sP = rP for some r ∈ P such that u ∈ rP . t s,l(b) = 0. Assume now qP ∩ sP = rP . Using (5.1) and t p,q(a)T T t p,q(a)T T u,t t s,l(b) = Mu∈qP = Mw∈ls−1rP p,q (a) Mw∈lP T w,t s,l (b) = Mu∈rP T u,t p,q (a) Mw∈ls−1rP T w,t s,l (b) T sl−1w,t p,q (a)T w,t s,l (b). Moreover, for every w ∈ ls−1rP and every x ∈ Xw,t we have T sl−1w,t p,q (a)T w,t s,l (b)x = (a ⊗ 1q−1sl−1w)(b ⊗ 1l−1w)x Accordingly, T Nica covariant. t p,q(a)T t s,l(b) = T t = (a ⊗ 1q−1r(ls−1r)−1w)(b ⊗ 1s−1r(ls−1r)−1w)x = T w,t pq−1r,ls−1r(cid:0)(a ⊗ 1q−1r)(b ⊗ 1s−1r)(cid:1) x. pq−1r,ls−1r(cid:0)(a ⊗ 1q−1r)(b ⊗ 1s−1r)(cid:1). Thus T t : L → L(FK) is (cid:3) Proposition 5.2. The right-tensor representation T : L → L(FK) given by (4.3) is Nica covariant. In particular, the Fock representation T : K → L(FK) is an injective Nica covariant representation. Proof. Since a direct sum of Nica covariant representations is a Nica covariant representation, the first part follows from Lemma 5.1. The second part follows from Proposition 4.3 and the fact that restriction of a Nica covariant representation to a well-aligned ideal is Nica covariant. (cid:3) Definition 5.3. We define the reduced Nica-Toeplitz algebra of the well-aligned ideal K in the right-tensor C ∗-precategory L to be the C ∗-algebra N T r L(K) := C ∗(T (K)), and we call T ⋊ P : N T L(K) → N T r sition 3.11. When K = L, we also write N T r(L) := N T r L(K) the regular representation of N T L(K), cf. Propo- L(L). It turns out that if P is right cancellative, in particular if it is a group, then the conditional L(K) onto the L(K). If P is not right cancellative, this is no longer true and expectation (4.5) restricts to a conditional expectation of the C ∗-algebra N T r core C ∗-subalgebra BT in particular E may not preserve the C ∗-algebra N T r e of N T r L(K). Proposition 5.4. The conditional expectation defined by (4.5) restricts to a faithful contrac- tive completely positive map ET : N T r L(K) → L(FK) given by (5.2) (5.3) ET(cid:16) Xp,q∈F T (ap,q)(cid:17) = Xp,q∈F Mw∈pP ∩qP,t∈P p−1w=q−1 w T w,t p,q (ap,q) for F ⊆ P finite, ap,q ∈ K(p, q), p, q ∈ F . The range of ET is the following self-adjoint operator space: BK := span(cid:26) Mw∈pP ∩qP,t∈P p−1 w=q−1 w T w,t p,q (a) : a ∈ K(p, q), p, q ∈ P(cid:27) ⊆ L(FK). NICA-TOEPLITZ ALGEBRAS FOR RIGHT TENSOR C∗-PRECATEGORIES 17 (5.4) (5.5) If P is cancellative, then BK = spannSp∈P Tp,p(K(p, p))o equals the core C ∗-subalgebra BT L(K) and ET is a faithful conditional expectation onto BK given by the formula of N T r e ET(cid:16) Xp,q∈F T (ap,q)(cid:17) =Xp∈F T (ap,p), for all ap,q ∈ K(p, q), p, q ∈ F and F ⊆ P finite. Proof. The crucial observation is the following claim: for a ∈ L(p, q), w ∈ qP and p, q, t ∈ P we have QwT w,t p,q (a)Qw =(T w,t 0 p,q (a) if w ∈ pP ∩ qP and p−1w = q−1w otherwise. To see this, recall that T w,t p,q (a) acts as an adjointable operator from Xw,t to Xpq−1w,t. Since Qw is the projection onto fixed range w, the only possibility to have QwT w,t p,q (a)Qw nonzero is that w = pq−1w, in which case it equals T w,t p,q (a). Thus to prove (5.5) it remains to see that if p, q ∈ P with w ∈ qP , then w = pq−1w is equivalent to w ∈ pP ∩ qP and p−1w = q−1w. However, the non-trivial left to right implication follows since w ∈ pP implies w = ps for some s ∈ P and so left cancellation gives s = q−1w = p−1w. Let now t ∈ P , F ⊆ P finite, ap,q ∈ K(p, q) for p, q ∈ F . By (5.5), E(Tp,q(ap,q)) = {0} when pP ∩ qP = ∅, and if pP ∩ qP 6= ∅, then E(Tp,q(ap,q)) = E Ms∈qP,t∈P T s,t p,q(ap,q) = Mw∈pP ∩qP,t∈P p−1 w=q−1 w T w,t p,q (ap,q). Thus ET = EN T r L(K) maps N T r L(K) onto BK according to the formula (5.2). Now suppose that P is right cancellative. Note that w ∈ pP ∩ qP and p−1w = q−1w if and only if p = q and w ∈ pP = qP , where the non-trivial left to right implication follows upon invoking right cancellation in w = p(p−1w) = q(q−1w). By using this observation one sees that (5.2) reduces to (5.4) and BK = BT e . Clearly, ET is an idempotent map and therefore a conditional expectation onto BK. (cid:3) Definition 5.5. Let K be a well-aligned ideal in a right-tensor C ∗-precategory L. We call the space BK given by (5.3) the transcendental core for K, and the map ET given by (5.2) the transcendental conditional expectation from N T r Remark 5.6. The transcendental core BK always contains the core C ∗-subalgebra BT Definition 3.18. For semigroups P that are not right cancellative, we may have BT and then it is not clear whether there is a conditional expectation from N T r e , see e ( BK, L(K) onto BK. L(K) onto BT e . Remark 5.7. In view of Lemma 4.7 and Proposition 5.4 one sees that for each t ∈ P , the conditional expectation defined by (4.4) restricts to a faithful contractive completely positive map ET,t on C ∗(T t(K)) with range the subspace of BK where t is fixed. In fact, we have ET =Lt∈P ET,t. The semilattice of projections introduced in the following lemma is one of the key tools to analyze the structure of the reduced Nica-Toeplitz algebra. Lemma 5.8. For each p ∈ P , let QT hpi ∈ L(FK) be the projection QT hpi(cid:16) Ms,t∈P xs,t(cid:17) := Ms∈pP,t∈P xs,t, xs,t ∈ FK. Ms,t∈P 18 BARTOSZ K. KWAŚNIEWSKI AND NADIA S. LARSEN hpi and ∅ 7→ 0 forms a semilattice homomorphism J(P ) 7−→ Proj(L(FK)), The assignment pP 7→ QT meaning that (5.6) QT hpiQT hqi =(QT 0, hri, if pP ∩ qP = rP for some r ∈ P if pP ∩ qP = ∅ for all p, q ∈ P . In particular, we have J (P ) ∼= {QT hpi : p ∈ P } ∪ {0}. Proof. The proof is immediate from the definition of QT hpi. (cid:3) Lemma 5.9. Let T : L → L(FK) be the representation given by (4.3). Then the projections introduced in Lemma 5.8 satisfy the relation: (5.7) T (a)QT hpi =(T (a ⊗ 1q−1w) 0 if qP ∩ pP = wP for some w ∈ P, otherwise, for all a ∈ L(r, q), r, p, q ∈ P . Proof. Let a ∈ L(r, q) and xu,v ∈ Xu,v where u, v, r, q ∈ P . Let p ∈ P . If T (a)QT hpixu,v 6= 0 then by the definition of QT hpi and (4.3) we necessarily have that u ∈ pP and u ∈ qP , which implies that qP ∩ pP = wP for some w ∈ P where u ∈ wP . Assume then that qP ∩ sP = wP for some w ∈ P . Note that T (a ⊗ 1q−1w)xu,v 6= 0 implies that u ∈ wP . Thus both sides of (5.7) are zero when u /∈ wP . It remains to verify that equality holds when u ∈ wP . This follows from two applications of (4.3): T (a)xu,v (4.3) = (a ⊗ 1q−1u)xu,v = (a ⊗ 1q−1(w(w−1u)))xu,v =(cid:16)(a ⊗ 1q−1w) ⊗ 1w−1u(cid:17)xu,v (4.3) = T (a ⊗ 1q−1w)xu,v. (cid:3) Lemma 5.10. The projections {QT BT of both N T r L(K) and BT e . e . If K ⊗ 1 ⊆ K in the sense of Definition 3.6, then {QT hpi}p∈P may be treated as multipliers of both C ∗(T (L)) and hpi}p∈P may be treated as multipliers L(K), C ∗(T (L)) and BT Proof. By Lemma 2.7 we have T (K(p, p))Xp,t = Xp,t for all p, t ∈ P , which implies that BT e and therefore also N T r e act on FK in a non-degenerate way. Using (5.7) we see that C ∗(T (L))QT e . Thus, since QT hpi is self- adjoint, we may treat QT e , cf. [26, Proposition 2.3]. If K ⊗ 1 ⊆ K, then (5.7) implies that N T r e , which finishes the proof. (cid:3) hpi ⊆ BT hpi as a multiplier of both C ∗(T (L)) and BT hsi ⊆ C ∗(T (L)) and BT hpi ⊆ N T r L(K) and BT e QT hpi ⊆ BT L(K)QT e QT We end this section by establishing certain norm formulas for elements in the transcendental core BK which can be considered far reaching generalizations of similar formulas obtained in [14, Lemma 7.4]. We begin by introducing some notation, which is inspired by [25, Remark 1.5] and [16, Remark 5.2], where quasi-ordered groups were considered, cf. also [14] and [5]. Suppose that C is a finite subset of P . We put σ(C) := e if C = ∅. If C is non-empty, then either Tc∈C cP = ∅ orTc∈C cP = c′P for an element c′ ∈ P (determined by C up to multiplication from the right by elements of P ∗). In the latter case we write σ(C) = c′. Let F be a finite subset of P . A subset C of F is an initial segment of F if σ(C) exists in P and NICA-TOEPLITZ ALGEBRAS FOR RIGHT TENSOR C∗-PRECATEGORIES 19 C = {t ∈ F : t ≤ σ(C)}. We denote by In(F ) the collection of all initial segments of F . For each C ∈ In(F ), the set PF,C := {t ∈ P : σ(C) ≤ t and f 6≤ t for all f ∈ F \ C} is non-empty. Note that neither definition of initial segment nor of the set PF,C depends on the choice of σ(C). Moreover, {PF,C : C ∈ In(F )} form a decomposition of the set P ; in particular, s ∈ PF,C if and only if C = {t ∈ F : t ≤ s} (the latter set is in In(F ) for every s ∈ P ). Now, for every finite set F ⊆ P and any C ∈ In(F ), QT F,C := QT (1 − QT hsi) hσ(C)i Ys∈F \C are mutually orthogonal projections that sum up to the identity in L(FK) as C varies. We use this partition of the identity to prove the following lemma. Lemma 5.11. Suppose that K is a well-aligned ideal in a right-tensor C ∗-precategory L and consider an element T w,t p,q (ap,q) ∈ BK, p−1w=q−1 w Z = Xp,q∈F Mw∈pP ∩qP,t∈P w∈PF,C(cid:13)(cid:13)(cid:13)T w,w hri = Mw∈pP ∩qP ∩rP,t∈P p−1 w=q−1 w sup where F ⊆ P is a finite set and ap,q ∈ K(p, q) for p, q ∈ F . Then kZk = max C∈In(F ) w,w ( Xp,q∈C p−1w=q−1 w Proof. For any projection QT hri, r ∈ P , and every a ∈ K(p, q), p, q ∈ P , we have T w,t p,q (a). T w,t p,q (a) = QT T w,t p,q (a)QT ap,q ⊗ 1q−1w)(cid:13)(cid:13)(cid:13). hri Mw∈pP ∩qP,t∈P p−1w=q−1 w Mw∈pP ∩qP,t∈P p−1 w=q−1 w Thus the projections QT of identity and ZQT (5.10) we get (5.8) (5.9) (5.11) F,C, where C ∈ In(F ), commute with Z. Since they form a partition p,q (ap,q) = Mw∈PF,C,t∈P Xp,q∈C p−1w=q−1 w T w,t p,q (ap,q) T w,t p−1w=q−1 w F,C = Xp,q∈C Mw∈PF,C ,t∈P kZk = maxn(cid:13)(cid:13)(cid:13) Mw∈PF,C,t∈P Xp,q∈C w∈PF,C,t∈P(cid:13)(cid:13)(cid:13) Xp,q∈C w∈PF,C,t∈P(cid:13)(cid:13)(cid:13)T w,t kZk = max = max C∈In(F ) C∈In(F ) sup sup p−1w=q−1 w p−1w=q−1 w T w,t p,q (ap,q)(cid:13)(cid:13)(cid:13) : C ∈ In(F )o p,q (ap,q)(cid:13)(cid:13)(cid:13). T w,t w,w(cid:16) Xp,q∈C p−1 w=q−1 w ap,q ⊗ 1q−1w(cid:17)(cid:13)(cid:13)(cid:13). Now, in the summation above over p, q ∈ C we have ap,q ⊗ 1q−1w ∈ L(w, w), and therefore Next note that for any w, t ∈ P and a ∈ L(w, w) we have ker T w,w (5.12) w,w ⊆ ker T w,t w,w ⇐⇒ aK(w, w) = {0} =⇒ aK(w, t)K(t, w) = {0} a ∈ ker T w,w w,w due to ⇐⇒ aK(w, t) = {0} ⇐⇒ a ∈ ker T w,t w,w. 20 BARTOSZ K. KWAŚNIEWSKI AND NADIA S. LARSEN Therefore, the supremum in (5.11) is attained for t = w. This proves (5.9). (cid:3) Corollary 5.12. With the assumptions from Lemma 5.11, if K is essential in the right-tensor sub-C ∗-precategory LK of L generated by K, cf. Lemma 3.17, then (5.9) reduces to (5.13) kZk = max C∈In(F ) sup p−1 w=q−1 w ap,q ⊗ 1q−1w(cid:13)(cid:13)(cid:13). w∈PF,C(cid:13)(cid:13)(cid:13) Xp,q∈C ap,p ⊗ 1p−1σ(C)(cid:13)(cid:13)(cid:13) : C ∈ In(F )o. e , then (5.14) If, additionally, Z =Pp∈F Tp,p(ap,p) ∈ BT kZk = maxn(cid:13)(cid:13)(cid:13)Xp∈C Proof. Note that elements of the formP p,q∈C If, in addition, Z = Pp∈F Tp,p(ap,p) ∈ BT C ∈ In(F ) the supremum supw∈σ(C)P kPp∈C ap,p ⊗ 1p−1wk is attained at w = σ(C). w,w are injective on LK(w, w), and therefore (5.9) reduces to (5.13). e , then (5.13) reduces to (5.14) because for every (cid:3) 6. Faithfulness of representations on core C ∗-subalgebras ap,q ⊗ 1q−1w belong to LK. If K is essential in LK then the maps T w,w p−1w=q−1 w The goal of this section is to prove Theorem 6.1, which is inspired by certain results used to obtain amenability criteria, cf. [25, Lemma 4.1], [16, Theorem 6.1], [14, Theorem 8.1]. In this section we use it to detect necessary and sufficient conditions for injectivity of representations Φ ⋊ P on the core C ∗-subalgebra BiK e ⊆ N T L(K). We will apply Theorem 6.1, in its full force, in Section 8 where we discuss the problem of amenability. Theorem 6.1. Let K be a well-aligned ideal in a right-tensor C ∗-precategory L over a right LCM semigroup P . Suppose that θ : P → P is a controlled map of right LCM semigroups, cf. Definition 2.4. (a) The subspace of N T L(K) = C ∗(iK(K)) defined as (6.1) spanniK(K(p, q)) : p, q ∈ P, θ(p) = θ(q)o is a C ∗-subalgebra of N T L(K) on which the regular representation T ⋊ P is faithful. (b) If Φ is a Nica covariant representation of K, the representation Φ ⋊ P is faithful on (6.1) if and only if Φ is injective and satisfies (6.2) span{Φ(K(p, q)) : θ(p) = θ(q) = u} ∩ span{Φ(K(s, t)) : θ(s) = θ(t) ∈ F } = {0} for all u ∈ P and all finite sets F ⊆ P such that u 6≥ v for every v ∈ F . Before we pass to the proof of Theorem 6.1 let us derive some consequences in the case when the homomorphism θ is the identity map. To facilitate the discussion we introduce a name for the condition in (6.2) in the case θ is injective. Definition 6.2. A Nica covariant representation Φ : K → B is Toeplitz covariant if (6.3) for every p ∈ P and q1, ..., qn ∈ P such that p 6≥ qi for all i = 1, ..., n, n ∈ N, we have Φ(K(p, p)) ∩ span{Φ(K(qi, qi)) : i = 1, ..., n} = {0}. In short we will call such representations Nica-Toeplitz covariant representations of K. Corollary 6.3. Let Φ : K → B be a Nica covariant representation. The representation Φ ⋊ P of N T L(K) restricts to an isomorphism (Φ ⋊ P ) iK B e : BiK e ∼=−→BΦ e NICA-TOEPLITZ ALGEBRAS FOR RIGHT TENSOR C∗-PRECATEGORIES 21 if and only if Φ is injective and Toeplitz covariant. Proof. Applying Theorem 6.1 with θ = id, we see that the C ∗-algebra in (6.1) equals BiK e , and condition (6.2) collapses to (6.3). (cid:3) Corollary 6.4. The Fock representation T : K → L(FK) is an injective Nica-Toeplitz co- variant representation of K. Equivalently, the regular representation T ⋊ P restricts to an isomorphism of core subalgebras (T ⋊ P ) ∼=−→BT e . : BiK e iK B e Proof. Apply Theorem 6.1 with θ = id. (cid:3) Corollary 6.5. If P is cancellative, then the formula E(cid:16) Xp,q∈F iK(ap,q)(cid:17) =Xp∈F (6.4) iK(ap,p), ap,q ∈ K(p, q), p, q ∈ F ⊆ P, defines a conditional expectation E : N T L(K) → BiK e ⊆ N T L(K). e onto BT Proof. Since P is cancellative, BK = BT e . By Corollary 6.4, T ⋊ P restricts to an isomorphism e . Denote by (T ⋊ P )−1 the inverse to this isomorphism. Taking into account of BiK (5.4) one sees that E = (T ⋊ P )−1 ◦ ET ◦ (T ⋊ P ) is a conditional expectation satisfying (6.4). (cid:3) The overall strategy of the proof of Theorem 6.1 is comparable to that behind the proofs of the quoted results in [25], [16], [14]. Nevertheless, we deal here with a much more general situation, which will require new insight. One of the new difficulties is the presence of invertible elements in the semigroup P . To deal with them we consider the following equivalence relation on P : p ∼ q ⇐⇒ p = qx for some x ∈ P ∗. We denote by [p] the equivalence class of p ∈ P in P/∼. Note that ∼ might not be a congruence and therefore P/∼ might not inherit the semigroup structure from P , cf. [5, Proposition 2.7]. However, P/∼ inherits the preorder. In fact, [p] ≤ [q] ⇐⇒ q ∈ pP yields a partial order on P/∼. Moreover, [p] and [q] have a common upper bound if and only if pP ∩ qP = rP for some r ∈ P , and if this holds then [r] is the unique least common upper bound of [p] and [q]. In the latter situation we write [p] ∨ [q] := [r]. The following two lemmas play a key role in the proof of Theorem 6.1. Lemma 6.6. Retain the assumptions of Theorem 6.1 (possibly without condition (2.3)). For every subset F ⊆ P/∼ such that [u] ∨ [v] ∈ F whenever [u], [v] ∈ F , the space (6.5) is a C ∗-subalgebra of N T L(K) such that for any F0 ⊆ P with F = {[u] : u ∈ F0} we have KF := span(cid:8)iK(K(p, q)) : p, q ∈ P, θ(p) = θ(q), [θ(p)] ∈ F(cid:9) KF = span(cid:8)iK(K(p, q)) : p, q ∈ P, θ(p) = θ(q) ∈ F0(cid:9). Proof. To see that KF is a C ∗-algebra, it suffices to show that KF is closed under multipli- cation. Let u, v ∈ P with [u], [v] ∈ F and suppose θ(p) = θ(q) = u and θ(s) = θ(t) = v. Let a ∈ K(p, q) and b ∈ K(s, t). Since iK is Nica covariant the product iK(a)iK(b) is either zero or qP ∩ sP = rP , for some r ∈ P , and then (6.6) iK(a)iK(b) = iK(cid:0)(a ⊗ 1q−1r)(b ⊗ 1s−1r)(cid:1). 22 BARTOSZ K. KWAŚNIEWSKI AND NADIA S. LARSEN In the latter case we have (a ⊗ 1q−1r)(b ⊗ 1s−1r) ∈ K(pq−1r, ts−1r) and using (2.2) we get [θ(r)] = [u] ∨ [v] ∈ F . Since θ(q)θ(q−1r) = θ(r) we have θ(q−1r) = θ(q)−1θ(r) and therefore θ(pq−1r) = θ(p)θ(q−1r) = θ(p)θ(q)−1θ(r) = θ(r). Similarly, θ(ts−1r) = θ(r). Thus iK(a)iK(b) ∈ KF . Hence KF is a C ∗-algebra. To prove (6.6), let F0 be a transversal for F and suppose that θ(p) = θ(q) with [θ(p)] ∈ F for p, q ∈ P . Since θ(P ∗) = P∗ there is x ∈ P ∗ such that θ(px) ∈ F0. By Lemma 3.10 we have iK(K(px, qx)) = iK(K(p, q)). This finishes the proof. (cid:3) We will prove injectivity in Theorem 6.1 by induction over the size of finite subsets of P/∼. In particular, it is crucial to establish the claim on sets with one element. As a matter of notation, if F = {[u]} ⊂ P/∼, we write K[u] instead of KF . Lemma 6.7. Retain the assumptions of Theorem 6.1. If Φ is an injective Nica covariant representation of K, then the homomorphism Φ ⋊ P of N T L(K) is faithful on K[u] for every u ∈ P. Suppose that Φ is an injective Nica covariant representation of K. Note that whenever s, t ∈ θ−1(u) with s 6= t, (2.3) implies that sP ∩ tP = ∅, and so by Nica covariance the C ∗- subalgebras iK(K(s, s)) and iK(K(t, t)) of K[u] are orthogonal. Hence we have a direct sum Proof. Let u ∈ P. Then K[u] = spannSp,q∈θ−1(u) iK(K(p, q))o by (6.6). C ∗-subalgebra Du :=Ls∈θ−1(u) iK(K(s, s)) in K[u]. Since Φ ⋊ P is faithful on each summand We claim that for every surjective ∗-homomorphism π : K[u] → B for B a C ∗-algebra, the it is also faithful on Du. formula (6.7) Eπ(cid:16)π(cid:0)Xs,t∈I iK(as,t(cid:1)(cid:17) :=Xs∈I π(cid:0)iK(as,s)(cid:1), where as,t ∈ K(s, t) and I ⊆ θ−1(u) is a finite set, defines a faithful conditional expectation from π(K[u]) onto π(Du). By choosing a faithful and non-degenerate representation of B on a Hilbert space H, we may assume that π : K[u] → B(H) is a non-degenerate representation. For each r ∈ θ−1(u) let Qr ∈ B(H) be the projection onto the essential space π(iK(K(r, r)))H for the C ∗-algebra π(iK(K(r, r))). Note that the projections Qr, r ∈ θ−1(u), are pair- wise orthogonal and their ranges span H. Thus, exactly as in the proof of Lemma 4.7, ful conditional expectation from B(H) onto the commutant of {Qr}r∈θ−1(u). Thus it suf- Indeed, clearly we have fices to check that this map satisfies (6.7). But this is easy. we conclude that the formula Eπ(a) = Pr∈θ−1(u) Qr(a)Qr, for a ∈ B(H), defines a faith- Qsπ(cid:0)iK(as,s)(cid:1)Qs = π(cid:0)iK(as,s)(cid:1) for every s ∈ I. On the other hand, if r ∈ θ−1(u) is such that either r 6= s or r 6= t, then Qrπ(cid:0)iK(as,t)(cid:1)Qr = 0, since, by Lemma 2.7, we have π(cid:0)iK(as,t)(cid:1) ∈ π(cid:0)iK(K(s, s))iK(K(s, t))iK(K(t, t))(cid:1) ⊆ QsB(H)Qt. The above claim applied separately to id and the ∗-homomorphism Φ ⋊ P gives a commu- tative diagram K[u] Eid Du Φ⋊P / (Φ ⋊ P )(K[u]) EΦ⋊P Φ⋊P / (Φ ⋊ P )(Du) in which the vertical arrows and the bottom horizontal arrow are faithful. Therefore, also (Φ ⋊ P ) : K[u] → (Φ ⋊ P )(K[u]) is faithful. (cid:3)   /   / NICA-TOEPLITZ ALGEBRAS FOR RIGHT TENSOR C∗-PRECATEGORIES 23 Proof of Theorem 6.1. We begin by proving that the Fock representation T satisfies (6.2). To this end, let u ∈ P and F ⊆ P be a finite set such that u 6≥ v for every v ∈ F . Suppose that p, q ∈ P satisfy θ(p) = θ(q) = u. By (5.5), the projection Qq in L(FK) ontoLt∈P Xq,t satisfies Qq(cid:0)T (K(s, t))(cid:1)Qq = 0 whenever s, t are in P such that q /∈ sP ∩ tP . Thus, for all s, t ∈ P with θ(s) = θ(t) ∈ F , we have Qq(cid:0)T (K(s, t))(cid:1)Qq = 0 by the choice of u and F . An arbitrary element a in span{T (K(p, q)) : θ(p) = θ(q) = u} has the form T (K(p, q))o a =Xq∈I aq where aq ∈ spann [p∈θ−1(u) where I ⊆ θ−1(u) is finite. Suppose that a ∈ span{T (K(s, t)) : θ(s) = θ(t) ∈ F }. The considerations of the previous paragraph imply that aQq = 0. On the other hand, aQq = (Pr∈I ar)Qq. For each r ∈ I with r 6= q, the assumption (2.3) implies that rP ∩ qP = ∅, and so the considerations above show that arQq = 0. Therefore aQq = aqQq for every q ∈ I. This implies that a∗ qaq ∈ T (K(q, q)) and T (K(q, q)) acts faithfully on ⊕t∈P Xq,t (because K(q, q) acts faithfully on the subspace Xq,q) we get a∗ qaq = 0. Thus aq = 0 for every q ∈ I. Accordingly, a = 0 and (6.2) is proved. Thus the injectivity claim in part (a) of the theorem will follow from part (b). qaqQq = 0. However, since a∗ Sufficiency in part (b). Let Φ : K → B be an injective Nica covariant representation satisfying (6.2). Let F denote the collection of all finite subsets F ⊆ P/∼ such that [u] ∨ [v] ∈ F whenever [u], [v] ∈ F . For any finite F0 ⊆ P/∼ the set F consisting of least upper bounds of all finite sub-collections of elements in F0 is finite, contains F0 and is closed under ∨. Thus F is a directed set. Accordingly, the corresponding C ∗-algebras (6.5) form an inductive system {KF : F ∈ F} with limit SF ∈F KF = spannS p,q∈P, iK(K(p, q))o. Hence (6.1) is a C ∗- algebra, and to prove faithfulness of Φ ⋊ P on SF ∈F KF it suffices to show that Φ ⋊ P is faithful on KF for each F ∈ F, see e.g. [1, Lemma 1.3]. By Lemma 6.7, Φ ⋊ P is faithful on KF if F = {[u]} for some u ∈ P. For the inductive step, let F ∈ F and suppose that Φ ⋊ P is faithful on KF ′ whenever F ′ ∈ F and F ′ < F ; we aim to prove that Φ ⋊ P is faithful on KF . Let Z ∈ KF be a finite sum of the form θ(p)=θ(q) Z = X[u]∈F Z[u] where Z[u] ∈ K[u] for [u] ∈ F. Suppose that Φ ⋊ P (Z) = 0. We will show that Z = 0, giving the desired injectivity. Since F is finite it has a minimal element, that is, there is [u0] ∈ F such that [u] 6≤ [u0] for every [u] ∈ F \ {[u0]}. It is immediate from considerations concerning products of elements in KF , cf. the proof of Lemma 6.6, that the C ∗-algebra KF \{[u0]} is an ideal in KF . Hence (Φ ⋊ P )(KF \{[u0]}) is an ideal in (Φ ⋊ P )(KF ). Let ρ : (Φ ⋊ P )(KF ) → (Φ ⋊ P )(KF )/(Φ ⋊ P )(KF \{[u0]}) be the quotient map. We claim that ρ is injective on (Φ ⋊ P )(K[u0]). Indeed, by (6.6) we have (Φ ⋊ P )(K[u0]) = span{Φ(K(p, q)) : θ(p) = θ(q) = u0} and for any F0 ⊆ P finite set such that F \ {[u0]} = {[u] : u ∈ F0}, we have (Φ ⋊ P )(KF \{[u0]}) = span{Φ(K(s, t)) : θ(s) = θ(t) ∈ F0}. 24 BARTOSZ K. KWAŚNIEWSKI AND NADIA S. LARSEN Note that v 6≤ u0 for every v ∈ F0. Thus condition (6.2) applied to u0 and F0 proves our claim. In particular, ρ is isometric on the C ∗-algebra Φ ⋊ P (K[u0]). Thus This implies that Z[u0] = 0, as Φ ⋊ P is isometric on K[u0] by the first inductive step. Hence k(Φ ⋊ P )(Z[u0])k = kρ(cid:16)(Φ ⋊ P )(Z[u0])(cid:17)k = kρ(cid:16)(Φ ⋊ P )(Z)(cid:17)k = 0. Z =P[u]∈F \{u0} Z[u] ∈ KF \{[u0]}. Therefore, Z = 0 by the inductive hypothesis. Necessity in part (b). Since T satisfies (6.2), we conclude by sufficiency in part (b) that the regular representation T ⋊ P is faithful on the C ∗-algebra in (6.1). It is not difficult to see that this forces the universal representation iK to satisfy (6.2): take u and F as specified for (6.2) and suppose that an element C in N T L(K) satisfies C = iK(C1) = iK(C2) where C1 ∈ K(p, q) is in the closed span determined by u and C2 ∈ K(s, t) is in the closed span determined by F . Then T (C1) = (T ⋊ P ) ◦ iK(C1) = (T ⋊ P ) ◦ iK(C2) = T (C2) and so, first, T (C1) = T (C2) = 0 because T satisfies (6.2), and second, iK(C1) = iK(C2) = 0 because both are in the C ∗-subalgebra (6.1) on which T ⋊ P is faithful. This gives C = 0 in this particular case, and the general case follows from here. Now, since iK satisfies (6.2), then every Nica covariant representation whose integrated form is faithful on (6.1) has to satisfy (6.2). This concludes the proof of the theorem. (cid:3) 7. Exotic Nica-Toeplitz C ∗-algebras We fix a well-aligned ideal K in a right-tensor C ∗-category L. In this section, we introduce Nica-Toeplitz C ∗-algebras of K that sit between N T L(K) and N T r Inspired by the terminology introduced in the context of group C ∗-algebras in [4], we shall call these C ∗- algebras exotic. In view of Example 12.12 below, our exotic Nica-Toeplitz C ∗-algebras are in fact generalizations of exotic crossed products for actions of discrete groups. L(K). Definition 7.1. We say that a Nica covariant representation Φ : K → B generates an exotic Nica-Toeplitz C ∗-algebra if ker(Φ ⋊ P ) ⊆ ker(T ⋊ P ). If this is the case, we refer to C ∗(Φ(K)) as an exotic Nica-Toeplitz C ∗-algebra of K. Remark 7.2. Clearly, Φ generates an exotic Nica-Toeplitz C ∗-algebra if and only if there is a homomorphism Φ∗ : C ∗(Φ(K)) → N T r L(K) making the following diagram commute: (7.1) N T L(K) Φ⋊P / C ∗(Φ(K)) Φ∗ / N T r L(K) . T ⋊P Proposition 7.3. If Φ : K → B is a Nica covariant representation that generates an exotic Nica-Toeplitz C ∗-algebra, then Φ is injective and Toeplitz covariant. Proof. The diagram (7.1) restricts to a commuting diagram of core subalgebras (7.2) BiK e Φ⋊P / BΦ e T ⋊P Φ∗ / BT e . By Corollary 6.4, the map T ⋊ P restricted to BiK e BiK e is an isomorphism and Corollary 6.3 implies the assertion. e → BΦ Nica covariant representations generating exotic Nica-Toeplitz algebras can be characterized e . Thus Φ ⋊ P : (cid:3) is an isomorphism onto BT as representations which admit a transcendental conditional expectation: Proposition 7.4. Let BK be the self-adjoint operator subspace of L(FK) introduced in (5.2). For a Nica covariant representation Φ : K → B the following conditions are equivalent: 5 5 / / 7 7 / / NICA-TOEPLITZ ALGEBRAS FOR RIGHT TENSOR C∗-PRECATEGORIES 25 (i) Φ : K → B generates an exotic Nica-Toeplitz C ∗-algebra of K, (ii) There is a bounded map EΦ : C ∗(Φ(K)) → BK satisfying (7.3) EΦ(cid:16) Xp,q∈F Φ(ap,q)(cid:17) = Xp,q∈F Mw∈pP ∩qP,t∈P p−1w=q−1 w T w,t p,q (ap,q) for every ap,q ∈ K(p, q) and finite F ⊆ P . If (i) and (ii) hold then EΦ is a contractive completely positive map making the diagram (7.4) C ∗(Φ(K)) $❏❏❏❏❏❏❏❏❏ EΦ / N T r L(K) z✉✉✉✉✉✉✉✉✉ ET Φ∗ BK commute. Moreover, Φ∗ is an isomorphism if and only if EΦ is faithful. Proof. The implication (i)⇒(ii) follows by letting EΦ := ET ◦ Φ∗. In particular, EΦ is a contractive completely positive map and the diagram (7.4) commutes. Moreover, since ET is faithful, Φ∗ is faithful if and only if EΦ is faithful. We prove next the implication (ii)⇒(i). Let us put EiK := ET ◦ (T ⋊ P ). Then ker(T ⋊ P ) = {a ∈ N T L(K) : EiK(a∗a) = 0}. Indeed, inclusion ker(T ⋊P ) ⊆ {a ∈ N T L(K) : EiK(a∗a) = 0} is trivial. The reverse inclusion follows, because if a ∈ N T L(K) is such that EiK(a∗a) = 0, then ET(cid:0)(T ⋊P )(a)∗(T ⋊P )(a)(cid:1) = 0 and therefore (T ⋊ P )(a) = 0 by faithfulness of ET . The assumption in (ii) implies that we have a commutative diagram N T L(K) $■■■■■■■■■ EiK Φ⋊P / C ∗(Φ(K)) zttttttttt EΦ BK Indeed, this diagram commutes when restricted to the dense C ∗-subalgebra N T L(K)0 = span{Sp,q∈P iK(K(p, q))} of N T L(K), cf. (3.5). Hence by continuity the diagram commutes on N T L(K). Now, a ∈ ker(Φ ⋊ P ) =⇒ EΦ(cid:0)(Φ ⋊ P )(a∗a)(cid:1) = 0 =⇒ EiK (a∗a) = 0. Thus ker(Φ ⋊ P ) ⊆ {a ∈ N T L(K) : EiK (a∗a) = 0} = ker(T ⋊ P ). (cid:3) Remark 7.5. Suppose that P is cancellative. Then the right-hand side of (7.3) reduces to the right-hand side of (5.4) and BK = BT e , cf. Proposition 5.4. Accordingly, Proposition 7.4 reduces to the following statement: a Nica covariant representation Φ of K generates an exotic Nica-Toeplitz C ∗-algebra if and only if the formula EΦ(cid:16) Xp,q∈F Φ(ap,q)(cid:17) =Xp∈F Φ(ap,p), ap,q ∈ K(p, q), p, q ∈ F ⊆ P, F finite, $ / z $ / z 26 BARTOSZ K. KWAŚNIEWSKI AND NADIA S. LARSEN defines a genuine conditional expectation from C ∗(Φ(K)) onto its core subalgebra BΦ e . fact, if Φ generates an exotic Nica-Toeplitz C ∗-algebra, we have the commutative diagram In (7.5) N T L(K) Φ⋊P / C ∗(Φ(K)) E BiK e Φ⋊P EΦ / BΦ e Φ∗ Φ∗ / N T r L(K) ET / BT e where the bottom horizontal arrows are isomorphisms and EΦ = (Φ∗BΦ Φ∗ : C ∗(Φ(K)) → N T r L(K) is an isomorphism if and only if EΦ is faithful. e )−1 ◦ EΦ. Moreover, As a first application of Proposition 7.4, we show that in order to study N T r L(K), one may in some cases use the t-th Fock subrepresentation T t of T (usually T e will work, see [20]). Proposition 7.6. Let t ∈ P and suppose that the well-aligned ideal K satisfies the condition: (7.6) ∀p∈P ∃x∈P ∗ K(px, t)K(t, px) is an essential ideal in the C ∗-algebra LK(px, px). Then the t-th Fock representation T t : K → L(F t that there is an isomorphism ht : N T r K) generates a copy of N T r L(K), in the sense L(K) → C ∗(T t(K)) such that ht ◦ T = T t. Proof. In view of Remark 5.7 and the last part of Proposition 7.4, it suffices to show that we have an isometry (7.7) BK ∈ Z = Xp,q∈F Mw∈pP ∩qP,s∈P p−1w=q−1 w T w,s p,q (ap,q) 7−→ Z t := Xp,q∈F Mw∈pP ∩qP p−1w=q−1 w p,q (ap,q) ∈ Bt T w,t K, This map is a well defined contraction, as it is the restriction of the projection fromLs∈P L(F s K). Now, the argument leading to (5.11) gives us that to L(F t K) kZ tk = max C∈In(F ) sup w∈PF,C(cid:13)(cid:13)(cid:13)T w,t w,w( Xp,q∈C p−1 w=q−1 w ap,q ⊗ 1q−1w)(cid:13)(cid:13)(cid:13). Let us fix w ∈ PF,C. Note that the sum under T w,t w,w belongs to LK(px, px). By (7.6) there is x ∈ P ∗ such that the homomorphism T wx,t wx,wx : LK(wx, wx) → L(Xwx,t) is injective (and hence isometric), cf. (5.12). Since LK ⊗ 1 ⊆ LK and the homomorphism ⊗1x is isometric we get (cid:13)(cid:13)(cid:13) Xp,q∈C p−1 w=q−1 w ap,q ⊗ 1q−1w(cid:13)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:13) Xp,q∈C =(cid:13)(cid:13)(cid:13)T wx,t wx,wx(cid:16) Xp,q∈C p−1 wx=q−1 wx ap,q ⊗ 1q−1wx(cid:13)(cid:13)(cid:13) p−1wx=q−1 wx ap,q ⊗ 1q−1wx(cid:17)(cid:13)(cid:13)(cid:13). Clearly, wx ∈ PF,C and therefore kZ tk is not greater than kZk by (5.9). Hence (7.7) is an isometry. (cid:3) 8. Amenability and Fell Bundles We fix a well-aligned ideal K in a right-tensor C ∗-precategory (L, {⊗1r}r∈P ) over a right In this section, we study the properties under which all exotic Nica- LCM semigroup P . Toeplitz algebras of K are naturally isomorphic.   /   /   / / NICA-TOEPLITZ ALGEBRAS FOR RIGHT TENSOR C∗-PRECATEGORIES 27 Definition 8.1. The well-aligned ideal K in (L, {⊗1r}r∈P ) is called amenable if the regu- L(K) is an isomorphism. A right-tensor C ∗- lar representation T ⋊ P : N T L(K) −→ N T r precategory L is amenable if it is amenable as an ideal in itself. We have the following simple characterization of amenability in terms of the natural map on the transcendental core BK. Lemma 8.2. A well-aligned ideal K in a right-tensor C ∗-precategory L is amenable if and only if the map EiK = ET ◦ (T ⋊ P ) : N T L(K) → BK is faithful. If P is cancellative, then K is amenable if and only if the conditional expectation E from N T L(K) onto BiK e given by (6.4) is faithful. Proof. Apply the last part of Proposition 7.4. (cid:3) We obtain more efficient amenability criteria using the theory of Fell bundles. We first recall the definition of a full coaction of G on a C ∗-algebra A. An unadorned tensor product of C ∗-algebras will denote the minimal tensor product. We write g 7→ iG(g) for the canonical inclusion of G as unitaries in the full group C ∗-algebra C ∗(G). There is a homomorphism δG : C ∗(G) → C ∗(G) ⊗ C ∗(G) given by δG(g) = iG(g) ⊗ iG(g). A full coaction of G on A is an injective, non-degenerate homomorphism δ : A → A ⊗ C ∗(G) that satisfies the coaction identity (δ ⊗ idC∗(G)) ◦ δ = (idA ⊗ δG) ◦ δ. Proposition 8.3. Suppose that θ : P → G is a unital semigroup homomorphism from P into a group G. There is a full coaction δ of G on N T L(K) such that δ(iK(a)) = iK(a) ⊗ iG(θ(p)θ(q)−1) for every a ∈ K(p, q), p, q ∈ P. The Fell bundle Bθ = {Bθ g }g∈G associated to δ has fibers given by Bθ g = span(cid:8)iK (K(p, q)) : p, q ∈ P, g = θ(p)θ(q)−1(cid:9) if g ∈ θ(P )θ(P )−1, g = {0} if g /∈ θ(P )θ(P )−1. In particular, N T L(K) ∼= C ∗(Bθ) with the isomorphism and Bθ which is identity on the spaces Bθ g , g ∈ G. Proof. We claim that the maps K(p, q) ∋ a 7−→ iK(a) ⊗ iG(θ(p)θ(q)−1) yield a Nica covariant representation Φ : K → N T L(K) ⊗ C ∗(G). Indeed, let a ∈ K(p, q) and b ∈ K(s, t), p, q, s, t ∈ P . If qP ∩ sP = ∅, then iK(a)iK(b) = 0 and therefore Φ(a)Φ(b) = 0. Assume that qP ∩ sP = rP for some r ∈ P . Writing for example qq′ = ss′ = r for some s′, q′ ∈ P shows that θ(p)θ(q)−1θ(s)θ(t)−1 = θ(pq′)θ(ts′)−1. With q′ = q−1r and s′ = s−1r, we therefore have Φ(a)Φ(b) =(cid:16)iK(a)iK(b)(cid:17) ⊗(cid:16)iG(θ(p)θ(q)−1)iG(θ(s)θ(t)−1)(cid:17) = iK(cid:16)(a ⊗ 1q−1r)(b ⊗ 1s−1r)(cid:17) ⊗(cid:16)θ(pq−1r)θ(ts−1r)−1)(cid:17) = Φ(cid:16)(a ⊗ 1q−1r)(b ⊗ 1s−1r)(cid:17). Thus Φ integrates to a homomorphism δ = Φ ⋊ P : N T L(K) → N T L(K) ⊗ C ∗(G). By Lemma 2.7 and the Hewitt-Cohen factorization theorem every element a ∈ K(p, q), p, q ∈ P , can be written as a = a′a′′ where a′ ∈ K(p, p) and a′′ ∈ K(p, q). Thus iK(a) ⊗ iG(θ(p)θ(q)−1) =(cid:16)iK(a′) ⊗ iG(e)(cid:17)(cid:16)iK(a′′) ⊗ iG(θ(p)θ(q)−1)(cid:17) ∈ N T L(K) ⊗ C ∗(G). 28 BARTOSZ K. KWAŚNIEWSKI AND NADIA S. LARSEN This implies that δ is non-degenerate. For every a ∈ K(p, q), p, q ∈ P we have (cid:16)(δ ⊗ idC∗(G)) ◦ δ(cid:17)(iK(a)) = (δ ⊗ idC∗(G))(cid:0)iK(a) ⊗ iG(θ(p)θ(q)−1)(cid:1) = iK(a) ⊗ iG(cid:0)θ(p)θ(q)−1) ⊗ iG(θ(p)θ(q)−1(cid:1) =(cid:16)(idA ⊗ δG) ◦ δ(cid:17)(iK(a)). Hence δ is a full coaction. It is readily seen that the spectral subspaces Bθ g , g ∈ G, for the coaction δ are of the claimed form. To see that N T L(K) ∼= C ∗(Bθ) it suffices to note that C ∗(Bθ) is generated by a universal Nica covariant representation. Now C ∗(Bθ) is generated by the spaces iK(K(p, q)), p, q ∈ G, and for a Nica covariant representation Φ of K the map g → C ∗(Φ(K)) given by (Φ ⋊ G)(⊕g∈Gbg) := Pg∈G(Φ ⋊ P )(bg) is a ∗- Φ ⋊ G : Lg∈G Bθ homomorphism. Since C ∗(Bθ) is the completion of Lg∈G Bg in the maximal C ∗-norm, the homomorphism Φ ⋊ G extends to the epimorphism Φ ⋊ G : C ∗(Bθ) → C ∗(Φ(K)). (cid:3) Theorem 8.4. Suppose that θ : P → P ⊆ G is a controlled map of right LCM semigroups such that P is a subsemigroup of a group G. Let Bθ be the Fell bundle described in Proposition 8.3. We have a commutative diagram C ∗(Bθ) Λ ∼= / C ∗ r (Bθ) ∼= N T L(K) T ⋊P / N T r L(K) where vertical arrows are isomorphisms and horizontal ones are regular representations. In particular, Bθ is amenable if and only if K is amenable as an ideal of L. r (Bθ) → Bθ r (Bθ). There are canonical conditional expectations Eδ : C ∗(Bθ) → Bθ Proof. Let Π : N T L(K) → C ∗(Bθ) be the isomorphism given by Proposition 8.3. It is easy to see that Φ := Λ ◦ Π ◦ iK is a Nica covariant representation of K such that C ∗(Φ(K)) is equal to C ∗ e and Eδ,r : C ∗ e . The existence of a controlled map into a semigroup P that is right cancellative (being a subsemigroup of G) guarantees that P is cancellative, see Remark 2.5 (a). By Proposition 5.4, the transcendental core BK is equal to BT e , hence it is a subspace of (T ⋊ P )(Bθ T ⋊ P : Bθ completely positive map from C ∗ (7.3). Note first that e ) = span(cid:8)T (K(p, q)) : p, q ∈ P, θ(p) = θ(q)(cid:9). By Theorem 6.1 the ∗-homomorphism e ) is an isomorphism. Hence EΦ := ET ◦ (T ⋊ P ) ◦ Eδ,r is a faithful e . We claim that EΦ satisfies equation e → (T ⋊ P )(Bθ r (Bθ) onto BK = BT (8.1) Eδ ◦ Π(iK(a)) :=(Π(iK(a)) 0 if θ(p) = θ(q), if θ(p) 6= θ(q), for every a ∈ K(p, q), p, q ∈ P.   /   / NICA-TOEPLITZ ALGEBRAS FOR RIGHT TENSOR C∗-PRECATEGORIES 29 Then for every choice of finite family ap,q in K(p, q), where p, q ∈ F finite, we have EΦ(Xp,q∈F Φ(ap,q)) = ET ◦ (T ⋊ P ) ◦ Eδ,r(cid:0)Xp,q∈F = ET ◦ (T ⋊ P ) ◦ Eδ(cid:0)Xp,q∈F = ET(cid:0) X{p,q∈F :θ(p)=θ(q)} T (ap,q)(cid:1), Λ ◦ Π ◦ iK(ap,q)(cid:1) Π ◦ iK(ap,q)(cid:1) which is the term in the right-hand side of (7.3) (which in this case reduces to the right-hand side of (5.4)). Thus Proposition 7.4 implies that there is a ∗-homomorphism Φ∗ from C ∗ r (Bθ) L(K) such that T ⋊ P = Φ∗ ◦ (Φ ⋊ P ). Since EΦ is faithful, the same proposition onto N T r shows that in fact Φ∗ is an isomorphism. This implies the assertion. (cid:3) Remark 8.5. Under the assumptions of Theorem 8.4, the Fell bundle Bθ can be constructed using any injective Nica covariant representation Φ satisfying (6.2). More specifically, given such a representation Φ we define g BΦ,θ := {0} for g /∈ θ(P )θ(P )−1. Then BΦ,θ := {BΦ,θ := span(cid:8)Φ (K(p, q)) : p, q ∈ P, g = θ(p)θ(q)−1(cid:9) if g ∈ θ(P )θ(P )−1 g }g∈G is a Fell bundle isomorphic is an isomorphism by Theorem 6.1, and hence by the g and BΦ,θ to Bθ. Indeed, Φ ⋊ P : Bθ C ∗-equality all the mappings Φ ⋊ P : Bθ e → BΦ,θ e g → BΦ,θ g , g ∈ G, are isomorphisms. The condition of amenability of Bθ in Theorem 8.4 is satisfied for instance when G is In particular, we get the following amenable or Bθ has the approximation property, [12]. generalization of [14, Corollary 8.2]. i∈I Pi is the free product of a family of right LCM semigroups Pi, i ∈ I, where each Pi is a subsemigroup of an amenable group Gi. Any well- aligned ideal K in a right-tensor C ∗-precategory L over P is amenable. Corollary 8.6. Suppose that P := Q∗ Proof. The direct sum G := Li∈I Gi is an amenable group. By Proposition 2.3, we have a homomorphism θ : P → G which satisfies the assumptions of Theorem 8.4. Hence the assertion follows from Theorem 8.4. (cid:3) 9. Projections associated to Nica covariant representations p }p∈P and {QΦ In this section we fix a Nica covariant representation Φ : K → B(H) of a well-aligned ideal K in a right-tensor C ∗-precategory L. We let Φ : L → B(H) be the extension of Φ from Proposition 2.13. Our goal is to investigate two families of projections associated to Φ: {QΦ hpi}p∈P . The former are projections onto the essential spaces we used to define Φ. The latter can be considered analogues of projections we associated to the Fock representation in Lemma 5.8. In general, the family {QΦ p }p∈P has some good properties that {QΦ hpi}p∈P lacks and vice versa. Thus the case when these families coincide is desirable and we give a natural condition implying that. Definition 9.1. For each p ∈ P , we denote by QΦ hpi projections in B(H) defined by p and QΦ (9.1) p H =(Φ(K(p, p))H if p ∈ P \ P ∗, C ∗(Φ(K))H if p ∈ P ∗. QΦ 30 BARTOSZ K. KWAŚNIEWSKI AND NADIA S. LARSEN and QΦ hpiH = span{Φ(K(w, w))H : w ∈ pP } (so we have QΦ Lemma 9.2. There is a well defined mapping J(P ) 7−→ Proj(B(H)) given by the assignment hpi =Ww≥p QΦ w). (9.2) pP 7−→ QΦ p and ∅ 7−→ 0 which sends comparable ideals (in the sense of inclusion) to commuting projections and has the property that QΦ q = 0 whenever pP ∩ qP = ∅ for p, q ∈ P . Moreover, for every p, q, s ∈ P and a ∈ K(p, q) we have p QΦ (9.3) Φ(a)QΦ s =(Φ(a ⊗ 1q−1r)QΦ 0 s if sP ∩ qP = rP for some r ∈ P, if sP ∩ qP = ∅. If additionally d ∈ L(s, s) and t ≥ s, then Φ(d)QΦ t = QΦ t Φ(d) and (9.4) Φ(a)Φ(d) =(Φ(a ⊗ 1q−1r)Φ(d) 0 if sP ∩ qP = rP for some r ∈ P, if sP ∩ qP = ∅. Proof. Note first that for every p ∈ P \ P ∗ the projection QΦ {Φ(µp λ}λ∈Λ is an approximate unit in K(p, p). λ)}λ∈Λ, where {µp p is the strong limit of the net If pP = qP for some p, q ∈ P then q = px for some x ∈ P ∗ and therefore Φ(K(q, q)) = Φ(K(p, p) ⊗ 1x) = Φ(K(p, p)) by Lemma 3.10. Thus QΦ q , so the map in (9.2) is well defined. If pP ∩ qP = ∅, then Φ(K(p, p))Φ(K(q, q)) = 0 by Nica covariance. Thus QΦ q = 0, as claimed. Now suppose that p ≤ q for some p, q ∈ P , and let a ∈ K(p, p). Then the equality Φp,p(a)Φq,q(b) = Φq,q((a ⊗ 1p−1q)b) implies that Φp,p(a)QΦ q . By passing to q ∈ Φp,p(K(p, p))′. It follows from the adjoints we get QΦ definition of Φp,p, cf. (2.7), that Φp,p(K(p, p))′ ⊆ Φp,p(L(p, p))′. Hence q Φp,p(a) = QΦ q . Thus QΦ q Φp,p(a)QΦ q Φp,p(a)QΦ q = QΦ p = QΦ p QΦ (9.5) QΦ q ∈ Φp,p(L(p, p))′, and in particular, QΦ p QΦ q = QΦ q QΦ p . Let now p, q, s ∈ P and a ∈ K(p, q). If sP ∩ qP = ∅, then Φ(a)QΦ s = Φ(a)QΦ q QΦ s = 0 by Lemma 2.7. Assume then that sP ∩ qP = rP . For any b ∈ K(s, s) we have Φ(a)Φ(b) = Φ(cid:0)(a ⊗ 1q−1r)(b ⊗ 1s−1r)(cid:1) = Φ(a ⊗ 1q−1r)Φ(b ⊗ 1s−1r) = s- lim Φ(a ⊗ 1q−1r)Φ(µr = s- lim Φ(a ⊗ 1q−1r)Φ(µr = s- lim Φ(a ⊗ 1q−1r)Φ(µr = Φ(a ⊗ 1q−1r)Φ(b) λ)Φ(b ⊗ 1s−1r) λ(b ⊗ 1s−1r)) λ)Φ(b) by construction of Φ in (2.7), and Nica covariance of Φ. Claim (9.3) follows. It implies (9.4) because Φ(d) = QΦ (cid:3) s Φ(d). If t ≥ s, then Φ(d)QΦ t Φ(d) by (9.5). t = QΦ Lemma 9.3. The assignment pP 7→ QΦ of semilattices, i.e. QΦ In particular, QΦ hqi = QΦ hpiQΦ hpi and ∅ 7→ 0 is a homomorphism J(P ) 7−→ Proj(B(H)) hqi = 0 for pP ∩ qP = ∅. hri when pP ∩ qP = rP and QΦ hpiQΦ hpi for p ∈ P are mutually commuting projections. Moreover, (9.6) QΦ hqi ∈ Φ(K(p, p))′ for all p, q ∈ P. NICA-TOEPLITZ ALGEBRAS FOR RIGHT TENSOR C∗-PRECATEGORIES 31 If Φ : L → B(H) is Nica covariant (which happens for instance when K = L) then (9.7) Φ(a)QΦ hsi =(Φ(a ⊗ 1q−1r) 0 if sP ∩ qP = rP for some r ∈ P, if sP ∩ qP = ∅ for every p, q, s ∈ P and a ∈ L(p, q). In particular, {QΦ hpi}p∈P ⊆ M (BΦ e ) ⊆ M (C ∗(Φ(L))). Proof. By Nica covariance of Φ, C ∗ each p ∈ P . We claim that, for every p, q ∈ P we have p (Φ) := span{Φ(K(w, w)) : w ∈ pP } is a C ∗-algebra, for (9.8) C ∗ p (Φ)C ∗ q (Φ) = C ∗ p (Φ) ∩ C ∗ q (Φ) =(C ∗ 0 r (Φ) if pP ∩ qP = rP for some r ∈ P, otherwise. Indeed, if pP ∩ qP = ∅, then for any w ∈ pP and v ∈ qP we have wP ∩ vP = ∅ and hence Φ(K(w, w))Φ(K(v, v)) = {0} by Nica covariance. Accordingly, in this case C ∗ q (Φ) = {0}. Assume next that pP ∩ qP = rP for some r ∈ P . Then C ∗ p (Φ) ⊆ C ∗ r (Φ) readily follows from Nica covariance of Φ. p (Φ). The reverse inclusion C ∗ p (Φ)C ∗ q (Φ) ∩ C ∗ r (Φ) ⊆ C ∗ p (Φ) ⊆ C ∗ q (Φ)C ∗ q (Φ)C ∗ Plainly, for every p ∈ P , the essential space C ∗ p (Φ)H for C ∗ hpi is a strong limit of any approximate unit in C ∗ p (Φ) is equal to QΦ In particular, QΦ p (Φ), and (9.8) justifies the claim about the semilattice homomorphism. Morever, (9.8) implies that for every p, q ∈ P and a ∈ K(p, p) we have Φ(a)QΦ hqiΦ(a) = QΦ hqiΦ(a)QΦ Suppose now that Φ : L → B(H) is Nica covariant. Let p, q, s ∈ P and a ∈ L(p, q). If sP ∩ qP = ∅, then Φ(a)C ∗ hsi = 0. Assume that sP ∩ qP = rP . Let w ∈ sP and b ∈ K(w, w). Since rP ∩ wP = (sP ∩ qP ) ∩ wP = sP ∩ (qP ∩ wP ) = sP ∩ tP for some t ∈ P , using Nica covariance of Φ twice we get s (Φ) = {0} by Nica covariance, and therefore Φ(a)QΦ hqi. By passing to adjoints we get QΦ hqi and therefore Φ(a)QΦ hqiΦ(a) which proves (9.6). hqiΦ(a)QΦ hqi = QΦ hqi = QΦ hpiH. Φ(a)Φ(b) = Φ(cid:0)(a ⊗ 1q−1t)(b ⊗ 1w−1t)(cid:1) = Φ(a ⊗ 1q−1r)Φ(b). hsi = Φ(a ⊗ 1q−1r)QΦ hsi. Since Φ(a ⊗ 1q−1r) = Φ(a ⊗ 1q−1r)QΦ hri and This implies that Φ(a)QΦ QΦ hsi, we get Φ(a)QΦ hri ≤ QΦ Relation (9.7) readily implies BΦ hsi = Φ(a ⊗ 1q−1r). This proves (9.7). Thus assuming the standard identification M (BΦ we conclude that QΦ hpi ∈ M (BΦ e ). e QΦ hpi ⊆ BΦ e . By taking adjoints we obtain QΦ e ) = {a ∈ QΦ heiB(H)QΦ hei : aBΦ e , BΦ hpiBΦ e ⊆ BΦ e . e a ⊆ BΦ e } Proposition 9.4. The following conditions are equivalent: (cid:3) p = QΦ hpi for every p ∈ P ; (i) QΦ (ii) The map (9.2) is a semilattice homomorphism; (iii) The map (9.2) is a pre-order homomorphism (pP ⊆ qP implies QΦ (iv) The extension Φ : L → B(H) of Φ is Nica covariant, and p ≤ QΦ q ); (9.9) Φ(a)QΦ s =(Φ(a ⊗ 1q−1r) 0 if sP ∩ qP = rP for some r ∈ P, if sP ∩ qP = ∅, for every p, q, s ∈ P and a ∈ L(p, q). In particular, if the above equivalent conditions hold, then QΦ p = QΦ hpi ∈ Φ(L(q, q))′ for all p, q ∈ P , and we have {QΦ p }p∈P = {QΦ hpi}p∈P ⊆ M (BΦ e ) ⊆ M (C ∗(Φ(L))). 32 BARTOSZ K. KWAŚNIEWSKI AND NADIA S. LARSEN Proof. Implication (i)⇒(ii) follows from Lemma 9.3. Implication (ii)⇒(iii) is trivial. We prove that (iii)⇒(iv). Let p, q, s ∈ P and a ∈ L(p, q). If sP ∩ qP = ∅, then QΦ by Lemma 9.2, and hence Φ(a)QΦ q QΦ that Φ(a) is a strong limit of the net Φ(aµq λ) = Φ(a)Φ(µq unit in K(q, q). Thus (9.9) follows from the calculations s = Φ(a)QΦ s = 0 s = 0. Assume then that sP ∩ qP = rP . Note λ}λ∈Λ is an approximate λ), where {µq q QΦ Φ(a)QΦ s = s- lim Φ(aµq λ)QΦ s = s- lim Φ(a ⊗ 1q−1r)Φ(µq = Φ(a ⊗ 1q−1r)QΦ r QΦ q QΦ s (9.3) = s- lim Φ(cid:0)(aµq λ) ⊗ 1q−1r(cid:1)QΦ s λ ⊗ 1q−1r)QΦ (iii) = Φ(a ⊗ 1q−1r)QΦ s (9.3) = s- lim Φ(a ⊗ 1q−1r)Φ(µq λ)QΦ s r = Φ(a ⊗ 1q−1r). To prove Nica covariance, let a ∈ L(p, q), b ∈ L(s, t). By (i), if qP ∩ sP = rP , then Φ(a)Φ(b) = Φ(a)QΦ q QΦ s Φ(b) = Φ(a)QΦ r Φ(b), which is Φ(cid:0)(a ⊗ 1q−1r)(b ⊗ 1s−1r)(cid:1) by the previous paragraph. The calculations above also show that Φ(a)Φ(b) = 0 when qP ∩ sP = ∅. To see that (iv)⇒(i), note that (9.7) and (9.9) imply that for every p, q ∈ P and a ∈ K(p, p) p are zero on the orthogonal complement of we have QΦ C ∗(Φ(K))H, this implies that QΦ p Φ(a). Since QΦ hpi and QΦ hpi = QΦ p . hpiΦ(a) = QΦ This proves the equivalence of (i)-(iv). Applying (9.9) and its adjoint to a ∈ L(q, q) we get p ∈ Φ(L(p, p))′, for all p, q ∈ P . The remaining part follows from Lemma 9.3. (cid:3) QΦ It is possible to cook up an example where the above equivalent conditions fail: Example 9.5. Let L = K = {K(n, m)}n,m∈N be a C ∗-precategory where K(n, m) = {0} for all n 6= m and K(n, n) are arbitrary (non-zero) C ∗-algebras. The multiplication in L is zero. We equip L with a right tensoring which is also zero. Then N T L(K) = N T (L) = T (L) is naturally isomorphic with the direct sum Ln∈N K(n, n). In particular, taking any faithful representations Φn,n : K(n, n) → B(Hn) and setting H :=Ln Hn and Φn,m := 0 for n 6= m, hni is the projection onto Lk≥n Hk. Note that Φ = Φ is Nica n, m ∈ N, we get an injective Nica covariant representation Φ of K. For each n ∈ N, QΦ the projection onto Hn and QΦ covariant and (9.7) is satisfied. However, (9.9) fails. n is In order to avoid situations as in Example 9.5, a nondegeneracy condition was introduced in [19, Definition 3.8]1, in the case P = N. We generalize this notion to arbitrary LCM semigroups. Virtually all examples considered in [20] will satisfy this condition. Definition 9.6. An ideal K in a right-tensor C ∗-precategory L is ⊗1-nondegenerate if (K(p, p) ⊗ 1r)K(pr, pr) = K(pr, pr) for every p ∈ P \ P ∗ and r ∈ P. Proposition 9.7. If K is an ⊗1-nondegenerate ideal in L then for every Nica covariant representation Φ of K the equivalent conditions in Proposition 9.4 hold true. p = QΦ hpi for all p ∈ P \ P ∗ Proof. We show condition (i) in Proposition 9.4. By definitions, QΦ hpi for all p ∈ P ∗. By nondegeneracy of K, for any p ∈ P \ P ∗ and any w ∈ pP and QΦ we have K(w, w) = (K(p, p) ⊗ 1p−1w)K(w, w). Thus by Nica covariance we get Φ(K(w, w)) = Φ(K(p, p) ⊗ 1p−1w)K(w, w)) = Φ(K(p, p))Φ(K(w, w)), which implies that QΦ (cid:3) p ≤ QΦ p ≥ QΦ hpi. 1We note that there are typos in [19, Definition 3.8], one should put there m = n. NICA-TOEPLITZ ALGEBRAS FOR RIGHT TENSOR C∗-PRECATEGORIES 33 10. Representations generating exotic C ∗-algebras - uniqueness theorem We fix a well-aligned ideal K in a right-tensor C ∗-precategory (L, {⊗1r}r∈P ). In this section we study conditions implying that a Nica covariant representation of K generates an exotic Nica-Toeplitz C ∗-algebra of K. In the presence of amenability this will lead to isomorphism theorems for the universal Nica-Toeplitz algebra N T L(K) as well. We start by introducing the key condition that we call (C). Here the letter C stands for both compression and Coburn. Definition 10.1. Let Φ : K → B(H) be a Nica covariant representation of K on a Hilbert space, and let {QΦ hpi}p∈P ⊆ B(H) be the projections introduced in Definition (9.1). We say that Φ satisfies condition (C) if for every p ∈ P and q1, ..., qn ∈ P such that p 6≥ qi, for i = 1, ..., n, n ∈ N, (10.1) the representation K(p, p) ∋ a 7−→ Φ(a)Qn Remark 10.2. Since projections {QΦ hqii). By (9.6), (1 − QΦ i=1(1 − QΦ hqii) is indeed a representation of the C ∗-algebra K(p, p). hpi}p∈P mutually commute, we have (1 −Wn hqii) ∈ Φ(K(p, p))′ and hence K(p, p) ∋ a 7−→ Φ(a)Qn i=1 QΦ Qn QΦ hqii) = i=1(1 − i=1(1 − QΦ hqii) is faithful. Every well-aligned ideal K admits a representation satisfying condition (C). Proposition 10.3. Let π be a representation of N T r FK from a faithful representation π0 :Lt∈P K(t, t) → B(H). Then π ◦ T : K → B(FK ⊗π0 H) is a Nica covariant representation satisfying condition (C). L(K) induced by the Fock Hilbert module Proof. Recall that π : L(FK) → B(FK ⊗π0 H) is a faithful representation given by the formula a(x ⊗π0 h) := (ax) ⊗π0 h, a ∈ L(FK), x ∈ FK, h ∈ H. {QT i=1 introduced in Lemma 5.8. Since the representation K(p, p) ∋ a 7−→ T p,p Let p ∈ P and q1, ..., qn ∈ P be such that p 6≥ qi, for i = 1, ..., n. Consider projections hqii}n p,p (a) ∈ hqii). With the definition of pro- hqii), for i = 1, ..., n. Hence jections from Lemma 9.3, it readily follows that Qπ◦T L(Xp,p) is faithful, so is K(p, p) ∋ a 7−→ T (a)Qn π(Qn hqii ) and therefore hqii ≤ π(QT i=1(1 − QT i=1(1 − QT i=1(1 − Qπ◦T hqii)) ≤Qn hqii) 6= 0 =⇒ π(cid:16)T (a) nYi=1 T (a) (1 − QT nYi=1 (1 − QT hqii)(cid:17) 6= 0 =⇒ π(T (a)) (1 − Qπ◦T hqii ) 6= 0. nYi=1 (cid:3) Thus K(p, p) ∋ a 7−→ π(T (a))Qn i=1(1 − Qπ◦T hqii ) is faithful. Plainly, every Nica covariant representation satisfying condition (C) is injective and Toeplitz covariant in the sense of (6.3). In the converse direction we have the following: Proposition 10.4. If K ⊗ 1 ⊆ K, then a Nica covariant representation Φ : K → B(H) satisfies condition (C) if and only if Φ is injective and Toeplitz covariant. Proof. The 'only if' part is clear. Suppose than that Φ is injective and Toeplitz covariant. By Corollaries 6.3 and 6.4 there is an isomorphism Φ∗ : BΦ e making the diagram (7.2) commute. Denote by Φ∗ : M (BΦ e ) the strictly continuous extension of Φ∗. Let p ∈ P and q1, ..., qn ∈ P be such that p 6≥ qi. We noticed in the proof of Proposition 10.3 that hqii) is faithful. By Lemma 5.10 we may e ) → M (BT the representation K(p, p) ∋ a 7−→ Tp,p(a)Qn k=1(1 − QT e → BT 34 BARTOSZ K. KWAŚNIEWSKI AND NADIA S. LARSEN hqii as elements of M (BT view QT that we may treat projections QΦ e ). Similarly, by Lemma 9.3 applied to L = K, one concludes hqii as elements of M (BΦ hqii. Thus e ), and then Φ∗(QΦ hqii) = QT Tp,p(a) (1 − QT nYk=1 Therefore K(p, p) ∋ a 7−→ Φp,p(a)Qn hqii) = Φ∗(cid:16)Φp,p(a) nYk=1 (1 − QΦ hqii)(cid:17) for all a ∈ K(p, p). k=1(1 − QΦ hqii) is faithful, and Φ satisfies (C). (cid:3) Corollary 10.5. Let Φ : K → B(H) be a Nica covariant representation. Suppose that K is essential in L and that the extended representation Φ : L → B(H) is Nica covariant (which is automatic when K is ⊗1-non-degenerate). The following conditions are equivalent: (i) Φ satisfies condition (C); (ii) Φ satisfies condition (C); (iii) Φ is injective and Toeplitz covariant. Proof. Equivalence (ii)⇔(iii) follows from Proposition 10.4 applied to L and Φ. To see that (i)⇔(ii) let p ∈ P and q1, ..., qn ∈ P such that p 6≥ qi, for i = 1, ..., n. Note that QΦ hqi = QΦ hqi, for q ∈ P . Hence since K(p, p) is an essential ideal in the C ∗-algebra L(p, p), the representation hqii) is faithful if and only if the representation L(p, p) ∋ (cid:3) i=1(1 − QΦ hqii) is faithful. K(p, p) ∋ a 7−→ Φ(a)Qn a 7−→ Φ(a)Qn i=1(1 − QΦ Next we introduce an auxiliary condition which describes properties of the (not necessarily Nica covariant) representation of L that extends a Nica covariant representation of K. Definition 10.6. Let Φ : K → B(H) be a Nica covariant representation on a Hilbert space, and let Φ : L → B(H) be the extension from Proposition 2.13. Let {QΦ p }p∈P ⊆ B(H) be the projections given by (9.1). We say that Φ satisfies condition (C ′) if (10.2) for every p ∈ P and q1, ..., qn ∈ P such that p 6≥ qi, for i = 1, ..., n, qi)k = kΦ(a)k for all a ∈ L(p, p). i=1 QΦ i=1 QΦ k(1 −Wn qi)Φ(a)(1 −Wn Proposition 10.7. Let Φ : K → B(H) be a Nica covariant representation of K and let Φ : L → B(H) be the extended representation of L. Consider the following assertions: (i) Φ satisfies condition (C); (ii) Φ is injective and Φ satisfies condition (C ′); Then (i)⇒(ii). If K is ⊗1-non-degenerate or K ⊗ 1 ⊆ K, then (i)⇔(ii). QΦ Proof. (i)⇒(ii). Plainly, condition (C) implies that Φ is injective. Let p ∈ P and q1, ..., qn ∈ P be such that p 6≥ qi. Let T p,p p,p : L(p, p) → L(Xp,p) = L(K(p, p)) be the homomorphism given by multiplication from left, cf. Lemma 4.1. By (2.8) and the construction of T , we have that ker T p,p p,p = ker Φp,p. Similarly, using (10.1), we see that the kernel of the representation i=1(1 − qi), this implies that qi)k ≥ kΦ(a)k for all a ∈ L(p, p). The reverse inequality is p,p . Thus kΦ(a)Qn hqii) coincides with the kernel of T p,p hqii) ≤ (1 −Wn i=1(1 − QΦ i=1 QΦ i=1 QΦ i=1(1 − QΦ L(p, p) ∋ a 7−→ Φ(a)Qn hqii)k = kΦ(a)k for all a ∈ L(p, p). SinceQn k(1 −Wn qi)Φ(a)(1 −Wn j = i, . . . , n, we have k(1 −Wn qi)Φ(a)(1 −Wn i=1 QΦ (ii)⇒(i). Let p ∈ P and q1, ..., qn ∈ P be such that p 6≥ qi. For every a ∈ K(qi, qi), qi)k = 0. Hence condition (C ′) implies that Φ is Toeplitz covariant, and therefore if K ⊗ 1 ⊆ K then Φ satisfies condition (C) by i=1 QΦ i=1 QΦ clear. NICA-TOEPLITZ ALGEBRAS FOR RIGHT TENSOR C∗-PRECATEGORIES 35 Proposition 10.4. Suppose then that K is ⊗1-non-degenerate. By Proposition 9.7 we have i=1 QΦ i=1(1 − QΦ hqii). Hence for any a ∈ K(p, p) we get (1 −Wn qi) =Qn nYi=1 kΦ(a) Thus Φ satisfies condition (C). (1 − QΦ hqii)k = k(1 − QΦ qi)Φ(a)(1 − n_i=1 n_i=1 QΦ qi)k = kΦ(a)k = kak. (cid:3) In order to get a uniqueness result in the case when the group P ∗ ⊆ P is non-trivial, we need to impose certain additional conditions on K and L. It seems that there is no single candidate for such a condition available on the scene. However, there are several natural conditions that can be applied in different situations. For instance, if P can be embedded into a group G via a monomorphism θ : P → G, and Bθ is the Fell bundle described in Proposition 8.3, then Theorem 8.4 and results of [21], [23] indicate that a natural condition is aperiodicity of the Fell bundle Bθ. We recall from [23, Definition 4.1] that a Fell bundle B = {Bg}g∈G is aperiodic if for each t ∈ G \ {e}, each bt ∈ Bt and every non-zero hereditary subalgebra D of Be, inf{kdbtdk : d ∈ D+, kdk = 1} = 0. For general semigroups P we may consider the following modification (and in fact general- ization) of this notion expressed in terms of L and K. We recall, see Lemma 3.10, that the group P ∗ of invertible elements acts on K by automorphisms. Definition 10.8. We say that the group {⊗1x}x∈P ∗ of automorphisms of L is aperiodic on K if for every p ∈ P , every non-zero hereditary C ∗-subalgebra D ⊆ K(p, p) and every b ∈ K(px, p) where x ∈ P ∗ \ {e} we have (10.3) inf{k(a ⊗ 1x)bak : a ∈ D+, kak = 1} = 0. Remark 10.9. For a fixed p ∈ P the spaces B(p) x multiplication and involution defined by bx · by := bx ⊗ 1yby and (bx)∗ := (b∗ bx ∈ B(p) {⊗1x}x∈P ∗ is aperiodic on K if and only if for each p ∈ P the Fell bundle B(p) is aperiodic. := K(px, p), x ∈ P ∗, equipped with x) ⊗ 1x−1, for x }x∈P ∗ over P ∗. In particular, the group x , by ∈ B(p) y , form a Fell bundle B(p) := {B(p) The following lemma can be proved using a standard argument, cf. the proof of [33, Lemma 5.2]. In view of Remark 10.9, it is a corollary of [23, Lemma 4.2]. Lemma 10.10. Suppose that the group {⊗1x}x∈P ∗ of automorphisms of K is aperiodic. Take any p ∈ P and let F ⊆ P ∗ be a finite set containing e. For every family of elements ax ∈ K(px, p) for x ∈ F with ae = a∗ e, and every ε > 0 there is an element d ∈ aeK(p, p)ae, kdk = 1, such that k(d ⊗ 1x)axdk < ε for every x ∈ F \ {e} and kdaedk > kaek − ε. Proof. By passing to −ae if necessary, we may assume that kaek is a spectral value of ae. Then applying to ae a non-decreasing continuous function that vanishes on (−∞, 0] and is identity on a neighborhood of kaek we may assume ae is positive. Then the assertion follows from [23, Lemma 4.2] applied to the Fell bundle B(p) described in Remark 10.9. (cid:3) We will also need the following fact. Lemma 10.11. Suppose that K is an essential ideal in L. Then the group {⊗1x}x∈P ∗ of automorphisms is aperiodic on K if and only if it is aperiodic on L. Proof. In view of Remark 10.9 the assertion follows from [21, Corollary 6.9]. (cid:3) 36 BARTOSZ K. KWAŚNIEWSKI AND NADIA S. LARSEN Our main results relate the properties of a Nica covariant representation of K on a Hilbert space that have been introduced so far. Recall that Toeplitz covariance was defined in (6.3); we think of it as being an algebraic condition. By contrast, condition (C) can be viewed as a geometric condition. Theorem 10.12. Let K be a well-aligned ideal in a right-tensor C ∗-precategory (L, {⊗1r}r∈P ). Consider the following conditions on a Nica covariant representation Φ : K → B(H): (i) Φ satisfies condition (C); (ii) Φ is injective and Φ : L → B(H) satisfies condition (C ′); (iii) Φ generates an exotic Nica-Toeplitz C ∗-algebra C ∗(Φ(K)) of K; (iv) Φ ⋊ P is injective on the core C ∗-subalgebra BiK (v) Φ is injective and Toeplitz covariant. e of N T L(K); Then (i) ⇒ (ii) and (iii) ⇒ (iv) ⇔ (v). If one of the following conditions holds: (1) P ∗ = {e}, (2) the group {⊗1x}x∈P ∗ is aperiodic on K and K is essential in LK, cf. Lemma 3.17, (3) there is a unital monomorphism θ : P → G into a group G and the Fell bundle Bθ from Proposition 8.3 is aperiodic, then (ii) ⇒ (iii). Further, if K ⊗ 1 ⊆ K, then all five conditions are equivalent. Proof. Implications (i)⇒(ii) and (iii)⇒(v)⇔(iv) follow respectively from Propositions 10.7 and 7.3 and Corollary 6.3. If K ⊗ 1 ⊆ K we also have (v)⇒(i) by Proposition 10.4. Thus it remains to show that (ii)⇒(iii), provided one of the conditions (1)-(3) holds. Suppose therefore that Φ is injective and Φ : L → B(H) satisfies condition (C ′). In view of Proposition 7.4, it suffices to show that (7.3) defines a bounded map EΦ on the dense ∗- subalgebra C ∗(Φ(K))0 of C ∗(Φ(K)), cf. (3.5). Moreover, it suffices to define a bounded map EΦ on the R-linear subspace C ∗(Φ(K))0 sa of C ∗(Φ(K))0 consisting of elements of the form (10.4) a = Xp,q∈F Φ(ap,q), where ap,q ∈ K(p, q), a∗ p,q = aq,p, for F ⊆ P finite and p, q ∈ F . Indeed, such map on C ∗(Φ(K))0 EΦ(a) = EΦ( a+a∗ ) + iEΦ( a−a∗ 2i ) to a bounded map EΦ on C ∗(Φ(K))0 satisfying (7.3). sa extends via the formula Let us then fix a self-adjoint element a given by (10.4). Denote by Z the right hand side of (7.3), cf. also (5.8). Note that Z ∗ = Z. Our strategy is to show (by consecutive compressions of a and its compressions) that for every ε > 0 we have 2 (10.5) kZk − ε ≤ k ak. To this end, we fix ε > 0, and note that by Lemma 5.11 we may find an initial segment C of F and w ∈ σ(C)P with w ∈ PF,C such that As noticed in the proof of Proposition 10.7, faithfulness of Φ implies that ker T w,w Thus we obtain w,w = ker Φw,w. pq−1 w=w w,w(cid:16) Xp,q∈C kZk − ε/2 ≤(cid:13)(cid:13)(cid:13)T w,w kZk − ε/2 ≤(cid:13)(cid:13)(cid:13)Φ(cid:16) Xp,q∈C pq−1 w=w ap,q ⊗ 1q−1w(cid:17)(cid:13)(cid:13)(cid:13). ap,q ⊗ 1q−1w(cid:17)(cid:13)(cid:13)(cid:13). (10.6) NICA-TOEPLITZ ALGEBRAS FOR RIGHT TENSOR C∗-PRECATEGORIES 37 Clearly, (10.5) will follow from (10.6) if we prove that (10.7) We fix the above C and w. Put (cid:13)(cid:13)(cid:13)Φ(cid:16) Xp,q∈C pq−1 w=w ap,q ⊗ 1q−1w(cid:17)(cid:13)(cid:13)(cid:13) − ε 4 ≤ kak. Frest := {r ∈ P : tP ∩ wP = rP, t ∈ F \ C}, F>w := {pq−1w : pq−1w = wx, x ∈ P \ P ∗, p, q ∈ C}, Fw := {pq−1w : pq−1w = wx, x ∈ P ∗, p, q ∈ C}. Note that for s ∈ Fw we have s ≥ w ≥ s and for s ∈ F>w we have s ≥ w and w 6≥ s. Since ω /∈ St∈F \C tP , for r ∈ Frest we have r ≥ w and w 6≥ r. Now suppose that d ∈ L(w, w). Using (9.4) (and its adjoint) and that Φ is a representation we get Φ(d)aΦ(d) = Xp,q∈C + Xp,q∈C pq−1 w∈Fw pq−1 w∈F>w Φ(d)Φ(ap,q ⊗ 1q−1w)Φ(d) Φ(d)Φ(ap,q ⊗ 1q−1w)Φ(d) + + Φ(d)Φ(ap,q ⊗ 1q−1r)Φ(d) Φ(d)Φ(ap,q ⊗ 1p−1r)Φ(d). qP ∩wP =rP,r∈Frest Xp∈F,q∈F \C Xp∈F \C,q∈C pP ∩wP =rP,r∈Frest We put F0 := Frest ∪ F>w and consider the projection QΦ F0 := _s∈F0 QΦ s . Note that (QΦ therefore Φ(d) and QΦ F0)⊥ := I − QΦ F0 is a nonzero projection. For any s ∈ F0 we have s ≥ w and s commute, by the last part of Lemma 9.2. This implies that (QΦ F0)⊥Φ(d)QΦ s = 0 and QΦ s Φ(d)(QΦ F0 )⊥ = 0 for every s ∈ F0. Applying this observation to QΦ waQΦ w we get (10.8) (QΦ F0)⊥Φ(d)aΦ(d)(QΦ F0 )⊥ = (QΦ F0)⊥ Xp,q∈C pq−1w∈Fw Φ(d)Φ(ap,q ⊗ 1q−1w)Φ(d)(QΦ F0 )⊥. To deduce (10.7) from here we consider the cases (1)-(3). Case (1). Assume that P ∗ = {e}. Then Fw = {w}. In particular, allowing d in (10.8) to run through an approximate unit in K(w, w) and taking strong limit we get (QΦ F0)⊥QΦ waQΦ w(QΦ F0)⊥ = (QΦ F0)⊥Φ(cid:16) Xp,q∈C pq−1 w=w ap,q ⊗ 1q−1w(cid:17)(QΦ F0)⊥. 38 BARTOSZ K. KWAŚNIEWSKI AND NADIA S. LARSEN Employing this equality and condition (C ′) we have kak ≥ k(QΦ F0)⊥QΦ waQΦ w(QΦ F0)⊥k = k(QΦ F0)⊥Φ(cid:16) Xp,q∈C pq−1 w=w ap,q ⊗ 1q−1w(cid:17)(QΦ F0)⊥k (10.2) = kΦ(cid:16) Xp,q∈C pq−1w=w ap,q ⊗ 1q−1w(cid:17)k. This proves (10.7) and finishes the proof under hypothesis (1). Case (2). Suppose that the group {⊗1x}x∈P ∗ is aperiodic on K and that K is essential in LK. Then {⊗1x}x∈P ∗ is also aperiodic on LK, by Lemma 10.11. Since Z is self-adjoint, it follows from (5.10) thatP{p,q∈C:pq−1w=w} ap,q ⊗ 1q−1w ∈ LK(w, w) is also self-adjoint. Hence by Lemma 10.10 there is d ∈ LK(w, w), kdk = 1, such that (10.9) k(d ⊗ 1x)(ap,q ⊗ 1q−1w)dk < ε 8C2 for every p, q ∈ C and x ∈ P ∗ \ {e} where pq−1w = wx, and (10.10) kd(cid:16) Xp,q∈C pq−1 w=w ap,q ⊗ 1q−1wk − ε 8 . We now return to the computation (10.8) with d chosen above, and note that (QΦ F0)⊥Φ(d)aΦ(d)(QΦ F0)⊥ = (QΦ pq−1 w=w ap,q ⊗ 1q−1w(cid:17)dk > k Xp,q∈C F0)⊥Φ(cid:16)d(cid:0) Xp,q∈C F0)⊥ Xp,q∈C,x∈P ∗\{e} + (QΦ pq−1 w=wx pq−1 w=w F0)⊥ ap,q ⊗ 1q−1w(cid:1)d(cid:17)(QΦ Φ(cid:16)(d ⊗ 1x)(ap,q ⊗ 1q−1w)d(cid:17)(QΦ F0)⊥. Since K is essential in LK and Φ is injective, Φ is isometric on LK, by the last part of Proposition 2.13. Note that d(cid:0)P p,q∈C estimates pq−1w=w ap,q ⊗ 1q−1w(cid:1)d ∈ LK(w, w). Thus we obtain the kak ≥ k(QΦ F0)⊥Φ(d)aΦ(d)QΦ F0)⊥k (10.9) > k(QΦ F0)⊥Φ(cid:16)d(cid:0) Xp,q∈C pq−1w=w ap,q ⊗ 1q−1w(cid:1)d(cid:17)(QΦ F0)⊥k − ε 8 ap,q ⊗ 1q−1w(cid:1)d(cid:17)k − ε 8 pq−1w=w (10.2) = kΦ(cid:16)d(cid:0) Xp,q∈C = kd(cid:0) Xp,q∈C ≥ kΦ(cid:16) Xp,q∈C pq−1 w=w (10.10) pq−1 w=w ε 8 ap,q ⊗ 1q−1w(cid:1)dk − ap,q ⊗ 1q−1w(cid:17)k − ε 4 . This proves (10.7) and finishes the proof under hypothesis (2). Case (3). Suppose that θ : P → G is a unital monomorphism and the Fell bundle Bθ is aperiodic. Then θ : P → θ(P ) ⊆ G is a controlled map of right LCM semigroups, P is NICA-TOEPLITZ ALGEBRAS FOR RIGHT TENSOR C∗-PRECATEGORIES 39 e = BiK e . Injectivity of Φ and condition (C ′) imply that Φ is a Toeplitz cancellative, and Bθ covariant representation, see the proof of Proposition 10.7. Hence by Corollary 6.3 (and the C ∗-equality) the maps Φ ⋊ P : Bθ g → C ∗(Φ(K)), g ∈ G, are injective. Putting FG := {θ(p)θ(q)−1 : p, q ∈ F } we have a = (Φ ⋊ P )(⊕g∈FGbg) where iK(ap,q), g ∈ FG. bg = Xp,q∈F, θ(p)θ(q)−1=g By (10.4), be is self-adjoint. Hence by [23, Lemma 4.2] there is d ∈ Be such that kdk = 1, kdbed − dadk < ε/2 and kdbedk > kbek − ε/2. Thus we get (10.11) kbek − ε/2 < kdbedk = kΦ ⋊ P (dbed)k ≤ kΦ ⋊ P (dad)k + ε/2 ≤ kak + ε/2. But since P is cancellative and θ injective we also have Z =Xp∈F Tp,p(ap,p) = Xp,q∈F θ(p)θ(q)−1=e Tp,q(ap,q) = T ⋊ P (be). Hence kZk = kbek and (10.11) implies (10.5). (cid:3) Remark 10.13. Under condition (3) in Theorem 10.12, also the implication (v) ⇒ (iii) holds. Indeed, given Nica covariant Φ that is injective and Toeplitz covariant, assume θ : P → G is an injective controlled homomorphism such that the associated Fell bundle Bθ is aperiodic. Corollary 6.3 implies that Φ ⋊P is injective on BiK e and, therefore ker(Φ ⋊P ) ⊆ ker Λ by e Proposition 12.10 below. Thus Φ generates an exotic Nica-Toeplitz C ∗-algebra by Theorem 8.4. ∼= Bθ Corollary 10.14 (Uniqueness Theorem I). Suppose that K is an amenable well-aligned ideal in a C ∗-precategory L and that one of conditions (1) − (3) in Theorem 10.12 holds. Consider the following properties of a Nica covariant representation Φ : K → B(H): (i) Φ satisfies condition (C); (ii) The map Φ ⋊ P is an isomorphism N T L(K) ∼= C ∗(Φ(K)). Then (i) ⇒ (ii). If K ⊗ 1 ⊆ K, then (i) ⇔ (ii). Proof. Since K is amenable, the regular representation T ⋊ P is injective. If any of the conditions (1) − (3) in Theorem 10.12 is satisfied, we infer from that result that Φ generates an exotic Nica-Toeplitz C ∗-algebra C ∗(Φ(K)). Now the very definition of an exotic Nica- Toeplitz algebra means that Φ ⋊ P is injective. (cid:3) By Proposition 9.7, if the ideal K is ⊗1-nondegenerate, then for every Nica covariant representation Φ of K the extended representation Φ of L is Nica covariant. Moreover, if K is essential in L and Φ is injective, then Φ is injective, see Proposition 2.13. This observation leads us to the following reformulation of Theorem 10.12: Theorem 10.15. Suppose that the well-aligned ideal K in L is essential and ⊗1-nondegenerate. Assume also that either P ∗ = {e} or the group {⊗1x}x∈P ∗ is aperiodic on K. For every Nica covariant representation Φ : K → B(H) the extended representation Φ of L is Nica covariant and the following statements are equivalent: (i) Φ satisfies condition (C); (ii) Φ generates an exotic Nica-Toeplitz C ∗-algebra C ∗(Φ(K)) of K, and Φ generates an exotic Nica-Toeplitz C ∗-algebra C ∗(Φ(L)) of L; (iii) Φ is injective and Toeplitz covariant; (iv) The map Φ ⋊ P is injective on the core C ∗-subalgebra BiL e of N T (L). 40 BARTOSZ K. KWAŚNIEWSKI AND NADIA S. LARSEN Proof. By Theorem 10.12 condition (i) implies that Φ generates an exotic Nica-Toeplitz C ∗- algebra C ∗(Φ(K)) of K. By Lemma 10.11, if the group {⊗1x}x∈P ∗ is aperiodic on K, then it is also aperiodic on L. By Corollary 10.5, Φ satisfies condition (C) if and only if Φ satisfies this condition. Hence we may apply Theorem 10.12 to the extended representation Φ : L → B(H). Then we get that each of conditions (i), (iii) and (iv) is equivalent to that Φ generates an exotic Nica-Toeplitz C ∗-algebra C ∗(Φ(L)) of L. (cid:3) The above theorem explains why in general condition (C) is stronger then the "uniqueness" for Nica-Toeplitz algebras: (C) implies uniqueness not only for a representation of K but also for the extended representation of L. We make this comment more formal in next section, see Corollary 11.5. 11. On the relationship between Nica-Toeplitz algebras of K and L Let K be a well-aligned ideal in a right-tensor C ∗-precategory L. In this section we collect results which reflect relationship between the full and reduced Nica-Toeplitz algebras asso- ciated to K and L. Recall from (3.6) the existence of a homomorphism ι from N T L(K) to N T (L). We write N T L(K) ֒→ N T (L) whenever ι is injective. Lemma 11.1. If K is ⊗1-nondegenerate, then N T L(K) ֒→ N T (L). Proof. Let iK : K → N T L(K) be the universal covariant representation of K. Suppose that N T L(K) ⊆ B(H) (for example in the usual universal representation as a C ∗-algebra). Then Proposition 9.7 and Proposition 9.4(ii) imply that the extension iK : L → B(H) of iK is Nica covariant. Then injectivity of ι follows as explained in Remark 3.13. (cid:3) Proposition 11.2. If K is amenable, then N T L(K) ֒→ N T (L). Proof. The Fock representation T of K is the restriction of a Nica covariant representation T : L → L(FK), cf. Proposition 4.3. The hypothesis on K means that T ⋊P is an isomorphism from N T L(K) onto N T r L(K). With η denoting the composition of the inclusion map from N T r L(K) into L(FK) with T ⋊ P , it follows that iK : K → N T L(K) admits the extension T : L → η(N T L(K)), which is a Nica covariant representation of L. Then injectivity follows by Remark 3.13. (cid:3) Theorem 11.3. Let K be a well-aligned ideal in a right-tensor C ∗-precategory (L, {⊗1r}r∈P ). Suppose that either P is cancellative or K is essential in the right-tensor C ∗-precategory LK generated by K. There is an embedding of C ∗-algebras: ιr : N T r L(K) ֒→ N T r(L) determined by N T r L(K) ∋ T (a) → S(a) ∈ N T r(L), a ∈ K(p, q), p, q ∈ P , where T and S are Fock representations of K and L, respectively. Moreover, ιr intertwines the conditional L(K) → BK and ES : N T r(L) → BL, in the sense that ES ◦ ιr = expectations ET : N T r ιrBK ◦ ET . Proof. Suppose first that P is cancellative. Then ET and ES are faithful conditional ex- pectations onto BK = BT e respectively, and are given by (5.4). Since S is an injective Nica-Toeplitz representation of L, it restricts to an injective Nica-Toeplitz rep- resentation Φ : K → L(FL), and ES restricts to a faithful conditional expectation EΦ : e ⊆ C ∗(Φ(K)). Applications of Corollaries 6.3 and 6.4 yield an isomorphism C ∗(Φ(K)) → BΦ h : BK → BΦ e ⊆ BL such that Φ = h ◦ T on each space K(p, p), p ∈ P . The composition EΦ := h−1 ◦ EΦ is a faithful completely positive map from C ∗(Φ(K)) onto BK satisfying equa- tion (7.3). Hence Proposition 7.4 gives an isomorphism Φ∗ from C ∗(Φ(K)) onto N T r K(L). e and BL = BS NICA-TOEPLITZ ALGEBRAS FOR RIGHT TENSOR C∗-PRECATEGORIES 41 Composing (Φ∗)−1 with the embedding of C ∗(Φ(K)) = C ∗(S(K)) into N T r(L) yields an em- bedding ιr. Since ES ◦ιr = EΦ ◦(Φ∗)−1 and ιrBK ◦ET = h◦ET , the claim about intertwining conditional expectations follows. Suppose now that K is essential in LK. Let Z ∈ BK be as in (5.8) and put ZS := Xp,q∈F Mw∈pP ∩qP,t∈P p−1w=q−1 w Sw,t p,q (ap,q) ∈ BL, where the adjointable maps Sw,t p,q are as defined in Lemma 4.1 (for K = L). By Corollary 5.12, kZk = kZSk, so the map Z → ZS extends to an isometry h : BK → h(BK) ⊆ BL. Let Φ : K → L(FL) be the restriction of S. Then ES restricts to a faithful map EΦ : C ∗(Φ(K)) → h(BK) ⊆ C ∗(Φ(K)). The composition EΦ := h−1 ◦ EΦ is a faithful completely positive map satisfying (7.3) and the proof if completed as in the case of P cancellative. (cid:3) Corollary 11.4. Suppose that L is amenable and that either P is cancellative or K is essential in the right-tensor C ∗-precategory LK generated by K. If N T L(K) ֒→ N T (L), then K is amenable. Proof. By our assumptions and Theorem 11.3 we have a commutative diagram N T L(K) ι / N T (L) T ⋊P S⋊P L(K) where ι, ιr and S ⋊ P are injective. Hence T ⋊ P is injective. L(K) / N T r N T r ιr (cid:3) Corollary 11.5 (Uniqueness Theorem II). Suppose that the well-aligned ideal K in L is essen- tial and ⊗1-nondegenerate. Assume also that either P ∗ = {e} or that the group {⊗1x}x∈P ∗ is aperiodic on K. If L is amenable, then K is also amenable and for any Nica covariant representation Φ : K → B(H) the following statements are equivalent: (i) Φ satisfies condition (C); (ii) Φ generates the universal Nica-Toeplitz algebra N T (L), i.e. C ∗(Φ(L)) ∼= N T (L). Under the embedding N T L(K) ֒→ N T (L), the isomorphism in (ii) restricts to an isomor- phism C ∗(Φ(K)) ∼= N T L(K). Proof. Lemma 11.1 implies that N T L(K) ֒→ N T (L), hence K is amenable by Corollary 11.4. The claim follows from Theorem 10.15. (cid:3) 12. Fell bundles and right-tensor C ∗-precategories over groups In this section we show that the theory of right-tensor C ∗-precategories over groups is equivalent to the theory of Fell bundles over (discrete) groups. On one hand, we explain how results about C ∗-algebras associated to Fell bundles follow from our results. On the other hand, we will see the origin of some of our assumptions. Throughout this section we assume that P equals a (discrete) group G. Proposition 12.1. If (L, {⊗1r}r∈G) is a right-tensor C ∗-precategory over G, then the Banach spaces Bg := L(g, e) for g ∈ G equipped with the operations (12.1) (12.2) ◦ : Bg × Bh ∋ (bg, bh) −→ bg ◦ bh := (bg ⊗ 1h)bh ∈ Bgh ⋆ : Bg ∋ bg −→ b⋆ g := b∗ g ⊗ 1g−1 ∈ Bg−1,   /   / 42 BARTOSZ K. KWAŚNIEWSKI AND NADIA S. LARSEN for g, h ∈ G, give rise to a Fell bundle BL := {Bg}g∈G isomorphic to the Fell bundle Bid associated to L and id as in Proposition 8.3, and so (12.3) r (BL) ∼= N T r(L), C ∗ C ∗(BL) ∼= N T (L). Conversely, for any Fell bundle B = {Bg}g∈G over a discrete group G the spaces (12.4) L(g, h) := Bgh−1, g, h ∈ G, with composition and involution inherited from B and right-tensoring given by ⊗ 1x : L(g, h) = Bgh−1 ∋ b −→ b ⊗ 1x := b ∈ Bgx(hx)−1 = L(gx, hx) (12.5) for g, h, x ∈ G, yield a right-tensor C ∗-precategory L := {L(g, h)}g,h∈G such that B = BL. Proof. Let (L, {⊗1r}r∈G) be a right-tensor C ∗-precategory over G. By Lemma 3.10, for every g, p, q ∈ G the map ⊗1g : L(p, q) → L(pg, qg) is an isometric isomorphism and for every a ∈ L(p, q) we have iL(a ⊗ 1g) = iL(a). This readily implies that the maps Bg = L(g, e) ∋ a 7→ iL(a) ∈ Bid g , g ∈ G, are isometric isomorphisms. Under these maps, the operations in the Fell bundle Bid = {Bid g }g∈G translate to (12.1), (12.2). Hence, B is a Fell bundle isomorphic to Bid. The isomorphisms (12.3) follow from Theorem 8.4. The remaining claims of the proposition are straightforward and left to the reader. (cid:3) Remark 12.2. To clarify relationships between the constructions in Proposition 12.1, let BL = {L(g, e)}g∈G denote the Fell bundle associated to a right-tensor C ∗-precategory L and let LB = {Bgh−1}g,h∈G be the right-tensor C ∗-precategory associated to a Fell bundle B. Then BLB = B and LBL ∼= L as right-tensor C ∗-precategories: the maps ⊗1h : L(gh−1, e) → L(g, h) for g, h ∈ G implement an isomorphism LBL ∼= L of C ∗-precategories which intertwines right- tensorings. Remark 12.3. When P = G is a group and L is any right-tensor C ∗-precategory over G, then for every t ∈ G condition (7.6) is trivially satisfied (one may take x = p−1t). Thus, by Proposition 7.6, N T r L) denoting the t-th Fock representation on L. Therefore, if B = {Bg}g∈G is a Fell bundle and L = {Bgh−1}g,h∈G is the associated C ∗-precategory from Proposition 12.1, then the Fock representation T e of L, viewed as a representation of B, coincides with the usual Fock representation of B introduced in [12]. L(K) ∼= C ∗(T t(K)), for all t ∈ G, with T t : L → L(F t For right-tensor C ∗-precategories over groups well-aligned ideals have nice structure. Lemma 12.4. Let K be a well-aligned ideal in a right-tensor C ∗-precategory (L, {⊗1r}r∈G) over a group G. Then K is automatically ⊗1-invariant and ⊗1-nondegenerate. Moreover, K is essential in L if and only if K(e, e) is essential in L(e, e). Proof. Lemma 3.10 gives directly that K is ⊗1-invariant and ⊗1-nondegenerate. If K(e, e) is essential in L(e, e), then K(p, p) = K(e, e) ⊗ 1p is essential in L(p, p) = L(e, e) ⊗ 1p, which proves the second part of the assertion. (cid:3) Lemma 12.5. Let (L, {⊗1r}r∈G) be a right-tensor C ∗-precategory and BL = {L(g, e)}g∈G the associated Fell bundle. Then there is a bijective correspondence between ideals I = {Ig}g∈G in B and well-aligned ideals K in L, given by Ig = K(g, e) for all g ∈ G. Proof. In view of Remark 12.2, we may assume that L = {Bgh−1}g,h∈G and the right tensoring is given by identity maps (12.5). Now it is clear that for any ⊗1-invariant ideal K in L, Ig = K(g, e) defines an ideal in BL. Conversely, let I be an ideal in BL and put K(g, h) := Igh−1, for all g, h ∈ G. Clearly, K is an ideal in L. It is also the smallest ⊗1-invariant ideal in B NICA-TOEPLITZ ALGEBRAS FOR RIGHT TENSOR C∗-PRECATEGORIES 43 satisfying Ig = K(g, e) for all g. However, by Lemma 12.4 well-alignment and ⊗1-invariance coincide. Hence this proves the assertion. (cid:3) As a first application of our results we obtain embedding of cross-sectional algebras of ideals in Fell bundles. The latter problem for sub-bundles is studied in [13, Section 21]. Proposition 12.6. Given a right-tensor C ∗-precategory (L, {⊗1r}r∈G) over a group G and a well-aligned ideal K in L, there are natural embeddings N T r(K) ֒→ N T r(L) and N T (K) ֒→ N T (L). Equivalently, for any ideal I in a Fell bundle B over a discrete group G we have C ∗ r (I) ֒→ C ∗ r (B) and C ∗(I) ֒→ C ∗(B). Proof. Theorem 11.3 gives N T r(K) ֒→ N T r(L). By Lemmas 12.4 and 11.1 we obtain N T (K) ֒→ N T (L). In view of Proposition 12.1 and Lemma 12.5, the second part of the assertion is equivalent to the first one. (cid:3) Remark 12.7. It is shown in [13, Theorem 21.13] that the embedding C ∗(I) ֒→ C ∗(B) is valid, not only for ideals but also for hereditary sub-bundles I of C ∗(B). In fact, Proposition 12.6 could be deduced from [13, Theorem 21.13 and Proposition 21.3] As a next application we get a correct version of [23, Proposition 3.15], see Remark 12.9 below. Proposition 12.8. Let B = {Bg}g∈G be a Fell bundle and let Ψ : C ∗(B) → B(H) be a representation injective on Be. The following conditions are equivalent (i) ker Ψ ⊆ ker Λ where Λ : C ∗(B) → C ∗ r (B) is the regular representation; (ii) Ψ(Lg∈G Bg) is topologically graded, that is we have kbek ≤ kΨ(Lg∈G bg)k for any Lg∈G bg ∈Lg∈G Bg; (iii) condition (ii) is satisfied for any positiveLg∈G bg ∈Lg∈G Bg. Proof. Let L = {Bgh−1}g,h∈G be the right-tensor C ∗-precategory associated to B, as in Propo- sition 12.1. It is straightforward that the maps Φg,h := ΨBgh−1 for g, h ∈ G form a repre- sentation Φ of L such that Φgr,hr(a ⊗ 1r) = Φg,h(a) for a ∈ L(g, h) and all g, h, r ∈ G. Identifying N T (L) with C ∗(B), via (12.3), the core of Accordingly, Φ is Nica covariant. N T (L) is L(e, e) = Be and Ψ = Φ ⋊ P . Thus Proposition 7.4 applied to Φ gives the equiv- alence of (i) and (ii). Equivalence between (ii) and (iii) is explained in the beginning of the proof of Theorem 10.12. (cid:3) Remark 12.9. Retain the notation of Proposition 12.8. Unless B is amenable, condition (i) is stronger than condition (i'): Λ(ker Ψ) ∩ Be = {0}. Unfortunately, [23, Proposition 3.15] says that counterparts of conditions (i'), (ii), (iii) are equivalent. In particular, the proof of implication (i)⇒(ii) in [23, Proposition 3.15] is incorrect. This mistake does not affect the remaining results of [23]. Let B be a Fell bundle over G and L the associated right-tensor C ∗-precategory, given by (12.4) and (12.5). It is immediate that the action of {⊗1g}g∈G on L is aperiodic if and only if B is aperiodic. By (12.3), amenability of B is equivalent to amenability of L. Moreover, since g ≤ h holds for any g, h in G, conditions (C) and (C') are void. Thus Theorem 10.12 reduces to the following uniqueness-type result, which in view of Remark 12.9 is stronger than [23, Corollary 4.3]. 44 BARTOSZ K. KWAŚNIEWSKI AND NADIA S. LARSEN Proposition 12.10. Let B = {Bg}g∈G be an aperiodic Fell bundle. For any representation Ψ : C ∗(B) → B(H) injective on Be we have ker Ψ ⊆ ker Λ. Remark 12.11. By [21, Theorem 9.11], when G = Z or G = Zn for a square free number n > 0, aperiodicity of B is not only sufficient but also necessary for Proposition 12.10 to hold, at least when Be contains an essential ideal which is separable or of Type I. Example 12.12 (Crossed products by group actions). Let α : G → Aut(A) be an action of a discrete group G on a C ∗-algebra A. Consider the C ∗-precategory L = {L(g, h)}g,h∈G where for each g, h ∈ G, L(g, h) equals A as a Banach space, and composition and involution in L are given by multiplication and involution in A. Then a −→ a ⊗ 1r := αr(a) for a ∈ A = L(g, h) and g, h, r ∈ G defines a right-tensoring on L. The corresponding Fell bundle BL coincides with the semi-direct product bundle associated to α, cf. [13]. Hence N T r(L) ∼= A ⋊r α G and N T (L) ∼= A ⋊α G. We have a natural bijective correspondence between α-invariant ideals I in A and well-aligned (⊗1-invariant) ideals K in L. Moreover, assuming that A contains an essential ideal which is separable or of Type I, by [21, Theorems 2.13, 8.1 and Corollary 9.10], the following conditions are equivalent: (i) the action of {⊗1g}g∈G on L is aperiodic; (ii) α is pointwise properly outer, i.e. each αg for g ∈ G \ {e} is properly outer; the setTn Using this, one recovers [2, Theorem 1] from Proposition 12.10. (iii) the dual actionbα : G → Homeo(bA) is topologically free, i.e. for any g1, . . . , gn ∈ G\{e} i=1{π ∈ bA :bαgi(π) = π} has empty interior. References [1] S. Adji, M. Laca, M. Nilsen and I. Raeburn, Crossed products by semigroups of endomorphisms and the Toeplitz algebras of ordered groups, Proc. Amer. Math. Soc. 122 (1994), 1133-1141. [2] R. J. Archbold, J. S. Spielberg, Topologically free actions and ideals in discrete C ∗-dynamical systems, Proc. Edinb. Math. Soc. (2) 37 (1993), 119 -- 124. [3] B. Blackadar, Shape theory for C ∗-algebras, Math. Scand. 56 (1985), 249 -- 275. [4] N. P. Brown and E. P. Guentner, New C ∗-completions of discrete groups and related spaces, Bull. London. Math. Soc. 45 (2013), 1181 -- 1193. [5] N. Brownlowe, N. S. Larsen and N. Stammeier, On C ∗-algebras associated to right LCM semigroups, Trans. Amer. Math. Soc. 369 (2017), no.1, 31 -- 68. [6] N. Brownlowe, N. S. Larsen and N. Stammeier, C ∗-algebras of algebraic dynamical systems and right LCM semigroups, Indiana Univ. Math. J., preprint arXiv:1503.01599v1. [7] N. Brownlowe, J. Ramagge, D. Robertson and M. F. Whittaker, Zappa-Szép products of semigroups and their C ∗-algebras, J. Funct. Anal. 266 (2014), 3937 -- 3967. [8] A. Buss, R. Exel, R. Meyer, Reduced C*-algebras of Fell bundles over inverse semigroups, Isr. J. Math. 220 (2017), 225 -- 274. [9] J. Crisp and M. Laca, Boundary quotients and ideals of Toeplitz C ∗-algebras of Artin groups, J. Funct. Anal. 242 (2007), 127 -- 156. [10] S. Doplicher, J. E. Roberts, A new duality theory for compact groups, Invent. Math. 98 (1989), 157 -- 218. [11] S. Doplicher. C. Pinzari, R. Zuccante, The C ∗-algebra of a Hilbert bimodule. Boll Unione Mat. Ital. Sez. B Artc. Ric. Mat. 1 (1998), 263 -- 281. [12] R. Exel, Amenability for Fell bundles, J. reine angew. Math. 492 (1997), 41 -- 73. [13] R. Exel, Partial dynamical and applications, systems, Fell bundles book available at: mtm.ufsc.br/ exel/papers/pdynsysfellbun.pdf [14] N. J. Fowler, Discrete product systems of Hilbert bimodules, Pacific J. Math. 204 (2002), 335 -- 375. [15] N. J. Fowler, P. S. Muhly, I. Raeburn, Representations of Cuntz-Pimsner algebras, Indiana Univ. Math. J. 52 (2003), 569 -- 605. NICA-TOEPLITZ ALGEBRAS FOR RIGHT TENSOR C∗-PRECATEGORIES 45 [16] N. J. Fowler and I. Raeburn, Discrete product systems and twisted crossed products by semigroups, J. Funct. Anal., 155 (1998), 171 -- 204. [17] N. J. Fowler and I. Raeburn, The Toeplitz algebra of a Hilbert bimodule, Indiana Univ. Math. J. 48 (1999), 155 -- 181. [18] P. Ghez, R. Lima, J.E. Roberts, W ∗-categories, Pacific J. Math. 120 (1985), 79 -- 109. [19] B. K. Kwaśniewski, C ∗-algebras generalizing both relative Cuntz-Pimsner and Doplicher-Roberts algebras, Trans. Amer. Math. Soc. 365 (2013), 1809 -- 1873. [20] B. K. Kwaśniewski and N. S. Larsen, Nica-Toeplitz algebras associated with product systems over right LCM semigroups, accepted in J. Math. Anal. Appl., DOI : 10.1016/j.jmaa.2018.10.020, arXiv:1706.04951 [21] B. K. Kwaśniewski, R. Meyer, Aperiodicity, topological freeness and pure outerness: from group actions to Fell bundles, Studia Math. 241 (2018), 257 -- 302. [22] B. K. Kwaśniewski, W. Szymański, Topological aperiodicity for product systems over semigroups of Ore type, J. Funct. Anal. 270 (2016), no. 9, 3453-3504. [23] B. K. Kwaśniewski, W. Szymański, Pure infiniteness and ideal structure of C ∗-algebras associated to Fell bundles, J. Math. Anal. Appl. 445 (2017), no. 1, 898-943. [24] D. Kyed, S. Raum, S. Vaes, M. Valvekens, L2-Betti numbers of rigid C ∗-tensor categories and discrete quantum groups, Anal. PDE 10 (2017), 1757-1791. [25] M. Laca and I. Raeburn, Semigroup crossed products and the Toeplitz algebras of nonabelian groups, J. Funct. Anal., 139 (1996), 415 -- 440. [26] E.C. Lance, Hilbert C ∗-Modules: A Toolkit for Operator Algebraists. Cambridge University Press, Cam- bridge (1995). [27] M.V. Lawson, Non-commutative Stone duality: inverse semigroups, topological groupoids and C∗-algebras, Internat. J. Algebra Comput. 22 (2012), no. 6, 1250058, 47 pp. [28] X. Li, Semigroup C ∗-algebras and amenability of semigroups, J. Funct. Anal. 262 (2012), no. 10, 4302 -- 4340. [29] G.J. Murphy, Crossed products of C ∗-algebras by semigroups of automorphisms, Proc. Lond. Math. Soc. 3 (1994), 423 -- 448. [30] S. Neshveyev, M. Yamashita, Drinfeld center and representation theory for monoidal categories, Comm. Math. Phys. 345 (2016), 385 -- 434 [31] A. Nica, C ∗-algebras generated by isometries and Wiener-Hopf operators, J. Operator Theory, 27 (1992), 17 -- 52. [32] M.D. Norling, Inverse semigroup C ∗-algebras associated with left cancellative semigroups, Proc. Edinb. Math. Soc. 57 (2014), no. 2, 533 -- 564. [33] P. Muhly and B. Solel, On the Morita equivalence of tensor algebras, Proc. London Math Soc. 81 (2000), 113 -- 168. [34] S. Popa and S. Vaes, Representation theory for subfactors, λ-lattices and C ∗-tensor categories, Commun. Math. Phys. 340 (2015), 1239-1280. [35] M. V. Pimsner, A class of C ∗-algebras generalizing both Cuntz-Krieger algebras and crossed products by Z, in Free probability theory (Waterloo, ON, 1995), Amer. Math. Soc., Providence, RI, 1997, 189 -- 212. [36] A. Sims and T. Yeend, C ∗-algebras associated to product systems of Hilbert bimodules, J. Operator Theory 64 (2010), 349 -- 376. E-mail address: [email protected] Institute of Mathematics, University of Bialystok ul. Ciolkowskiego 1M, 15-245 Bialystok, Poland // Department of Mathematics and Computer Science, The University of Southern Denmark, Campusvej 55, DK-5230 Odense M, Denmark E-mail address: [email protected] Department of Mathematics, University of Oslo, PO Box 1053 Blindern, 0316 Oslo, Norway
1002.4390
2
1002
2010-10-14T18:55:24
A characterization of freeness by invariance under quantum spreading
[ "math.OA", "math.PR", "math.QA" ]
We construct spaces of quantum increasing sequences, which give quantum families of maps in the sense of Soltan. We then introduce a notion of quantum spreadability for a sequence of noncommutative random variables, by requiring their joint distribution to be invariant under taking quantum subsequences. Our main result is a free analogue of a theorem of Ryll-Nardzewski: for an infinite sequence of noncommutative random variables, quantum spreadability is equivalent to free independence and identical distribution with respect to a conditional expectation.
math.OA
math
A CHARACTERIZATION OF FREENESS BY INVARIANCE UNDER QUANTUM SPREADING STEPHEN CURRAN Abstract. We construct spaces of quantum increasing sequences, which give quantum families of maps in the sense of So ltan. We then introduce a notion of quantum spreadability for a sequence of noncommutative random variables, by requiring their joint distribution to be invariant under taking quantum subsequences. Our main result is a free analogue of a theorem of Ryll-Nardzewski: for an infinite sequence of noncommutative random variables, quantum spreadability is equivalent to free independence and identical distribution with respect to a conditional expectation. Introduction The study of random objects with distributional symmetries is an important subject in modern probability. Consider a sequence (ξ1, ξ2, . . . ) of random variables. Such a sequence is called exchangeable if its distribution is invariant under finite permutations, and spreadable if it is invariant under taking subsequences, i.e., if (ξ1, . . . , ξk) d∼ (ξl1, . . . , ξlk) for all k ∈ N and l1 < · · · < lk. In the 1930's, de Finetti gave his famous characterization of infinite exchangeable sequences of random variables taking values in {0, 1} as conditionally i.i.d. This was extended to variables taking values in a compact Hausdorff space by Hewitt and Savage [6]. It was later discovered by Ryll-Nardzewski that de Finetti's theorem in fact holds under the apparently weaker condition of spreadability [12]. For a comprehensive treatment of distributional symmetries in classical probability, the reader is referred to the recent text of Kallenberg [8]. Free probability, developed by Voiculescu in the 1980's, is based on the notion of free in- dependence for random variables with the highest degree of noncommutativity. Remarkably, there is a deep parallel between the theories of classical and free probability. However, it is only quite recently that this parallel has been extended to the study of distributional sym- metries. The breakthrough came with the work of Kostler and Speicher [10], who discovered that, roughly speaking, in free probability one should consider quantum distributional sym- metries. More specifically, they defined the notion of quantum exchangeability for a sequence (x1, x2, . . . ) of noncommutative random variables by requiring that for each n ∈ N, the joint distribution of (x1, . . . , xn) is invariant under the natural action of the quantum permutation group As(n) of Wang [17]. They then gave a free analogue of de Finetti's theorem: for an in- finite sequence of noncommutative random variables, quantum exchangeability is equivalent to free independence and identical distribution with respect to a conditional expectation. 2010 Mathematics Subject Classification. 46L54 (46L65, 60G09). Key words and phrases. Free probability, quantum increasing sequence, quantum spreadability. 1 2 STEPHEN CURRAN This has since been extended to more general sequences [4], and to sequences invariant un- der actions of other compact quantum groups [5, 3]. (See also [9] for a detailed analysis of exchangeability and spreadability for sequences of noncommutative random variables). The purpose of the present paper is to develop a notion of quantum spreadability for sequences of noncommutative random variables. The first problem is to find a suitable quantum analogue of an increasing sequence. The answer which we suggest here is similar to Wang's notion of a quantum permutation. For natural numbers k ≤ n we construct certain universal C∗-algebras Ai(k, n), which we call quantum increasing sequence spaces, whose spectrum is naturally identified with the space of increasing sequences 1 ≤ l1 < · · · < lk ≤ n. These objects form quantum families of maps, in the sense of So ltan [13], from {1, . . . , k} into {1, . . . , n}. Quantum spreadability is naturally defined as invariance under these familes of quantum transformations. This approach is justified by our main result, which is a free analogue of the Ryll-Nardzewski theorem for quantum spreadable sequences (see Sections 1 and 3 for definitions and motivating examples): Theorem 1. Let (ρi)i∈N be an infinite sequence of unital ∗-homomorphisms from a unital ∗-algebra C into a tracial W∗-probability space (M, τ ). Then the following are equivalent: (i) (ρi)i∈N is quantum exchangeable. (ii) (ρi)i∈N is quantum spreadable. (iii) (ρi)i∈N is freely independent and identically distributed with respect to the conditional expectation E onto the tail algebra B = \n≥1 W ∗(cid:0){ρi(c) : c ∈ C, i ≥ n}(cid:1). The equivalence of (i) and (iii) is the main result of [10] in the case C = C[t], and was shown for general C in [4]. Observe that Theorem 1 holds only for infinite sequences. In [4], we have given an ap- proximation to how far a finite quantum exchangeable sequence is from being free with amalgamation. As in the classical case, finite quantum spreadable sequences are more diffi- cult, and we will not attempt an analysis here. For a treatment of classical finite spreadable sequences, see [7]. Our paper is organized as follows. Section 1 contains notations and preliminaries. We recall the basic notions from free probability, and introduce Wang's quantum permutation group As(n). In Section 2, we introduce the algebras Ai(k, n) and prove some basic results. In particular we show that Ai(k, n) is a quotient of As(n). In Section 3, we introduce the notions of quantum exchangeability and spreadability, and prove the implications (i) ⇒ (ii) and (iii) ⇒ (i) of Theorem 1. These implications hold in fact for finite sequences, and in a purely algebraic context. We complete the proof of Theorem 1 in Section 4, by showing the implication (ii) ⇒ (iii). 1.1. Notations. Let C be a unital ∗-algebra. Given an index set I, we let 1. Background and notations CI = ∗ i∈I C (i) denote the free product (with amalgamation over C), where for each i ∈ I, C (i) is an isomorphic copy of C. For c ∈ C and i ∈ I we denote the image of c in C (i) as c(i). The A CHARACTERIZATION OF FREENESS BY INVARIANCE UNDER QUANTUM SPREADING 3 universal property of the free product is that given a unital ∗-algebra A and a family (ρi)i∈I of unital ∗-homomorphisms from C to A, there is a unique unital ∗-homomorphism from CI to A, which we denote by ρ, such that ρ(c(i)) = ρi(c) for c ∈ C and i ∈ I. We will mostly be interested in the case that I = {1, . . . , n}, in which case we denote CI by Cn, and I = N in which case we denote CI = C∞. 1.2. Free Probability. We begin by recalling some basic notions from free probability, the reader is referred to [16],[11] for further information. Definition 1.3. (1) A noncommutative probability space is a pair (A, ϕ), where A is a unital ∗-algebra and ϕ is a state on A. (2) A W∗-probability space is a pair (M, τ ), where M is a von Neumann algebra and τ is a faithful normal state which is tracial, i.e., τ (xy) = τ (yx) for x, y ∈ M. Definition 1.4. Let C be a unital ∗-algebra, (A, ϕ) a noncommutative probability space and (ρi)i∈I a family of unital ∗-homomorphisms from C to A. The joint distribution of the family (ρi)i∈I is the state ϕρ on CI defined by ϕρ = ϕ ◦ ρ. ϕρ is determined by the moments k where c1, . . . , ck ∈ C and i1, . . . , ik ∈ I. ϕρ(c(i1) 1 · · · c(ik) ) = ϕ(ρi1(c1) · · · ρik(ck)), 1.5. Examples. (1) Let (Ω, F , P ) be a probability space, let (S, S) be a measure space and (ξ)i∈I a family of S-valued random variables on Ω. Let A = L∞(Ω), and let ϕ : A → C be the expectation functional ϕ(f ) = E[f ]. Let C be the algebra of bounded, complex-valued, S-measurable functions on S. For i ∈ I, define ρi : C → A by ρi(f ) = f ◦ ξi. Then ϕρ is determined by ϕρ(f (i1) 1 · · · f (ik) k ) = E[f1(ξi1) · · · fk(ξik)] for f1, . . . , fk ∈ C and i1, . . . , ik ∈ I. (2) Let C = C[t], and let (xi)i∈I be a family of self-adjoint random variables in A. Define ρi : C → A to be the unique unital ∗-homomorphism such that ρi(t) = xi. Then CI = Chti : i ∈ Ii, and we recover the usual definitions of the joint distribution and moments of the family (xi)i∈I. Remark 1.6. These definitions have natural "operator-valued" extensions given by replacing C by a more general algebra of scalars. This is the right setting for the notion of freeness with amalgamation, which is the analogue of conditional independence in free probability. Definition 1.7. A B-valued probability space (A, E) consists of a unital ∗-algebra A, a ∗- subalgebra 1 ∈ B ⊂ A, and a conditional expectation E : A → B, i.e., E is a linear map such that E[1] = 1 and for all b1, b2 ∈ B and a ∈ A. E[b1ab2] = b1E[a]b2 Definition 1.8. Let C be a unital ∗-algebra, (A, E) a B-valued probability space and (ρi)i∈I a family of unital ∗-homomorphisms from C into A. 4 STEPHEN CURRAN (1) We let C B I denote the free product over i ∈ I, with amalgamation over B, of C (i) ∗ B, which is naturally isomorphic to CI ∗ B. For each i ∈ I, we extend ρi to a unital ∗-homomorphism eρi : C ∗ B → A by setting eρi = ρi ∗ id. We then let eρ denote the induced unital ∗-homomorphism from C B I ρ ∗ id. Explicitly, we have into A, which is naturally identified with 1 b1 · · · c(ik) k bk) = b0ρi1(c1)b1 · · · ρik (ck)bk for b0, . . . , bk ∈ B, c1, . . . , ck ∈ C and i1, . . . , ik ∈ I. eρ(b0c(i1) (2) The B-valued joint distribution of the family (ρi)i∈I is the linear map Eρ : CI ∗B → B Eρ[b0c(i1) 1 · · · c(ik) k defined by Eρ = E ◦eρ. Eρ is determined by the B-valued moments (3) The family (ρi)i∈I is called identically distributed with respect to E if E ◦eρi = E ◦eρj for all i, j ∈ I. This is equivalent to the condition that for c1, . . . , ck ∈ C, b0, . . . , bk ∈ B and i1, . . . , ik ∈ I. E[b0ρi(c1) · · · ρi(ck)bk] = E[b0ρj(c1) · · · ρj(ck)bk] bk] = E[b0ρi1(c1) · · · ρik (ck)bk] for any i, j ∈ I and c1, . . . , ck ∈ C, b0, . . . , bk ∈ B. (4) The family (ρi)i∈I is called freely independent with respect to E, or free with amalga- mation over B, if E[eρi1(β1) · · ·eρik (βk)] = 0 whenever i1 6= · · · 6= ik ∈ I, β1, . . . , βk ∈ C ∗ B and E[eρil(βl)] = 0 for 1 ≤ l ≤ k. Remark 1.9. Voiculescu introduced the notion of freeness with amalgamation and developed its basic theory in [15]. Freeness with amalgamation also has a rich combinatorial structure developed by Speicher [14]. The basic objects, which we will now recall, are non-crossing set partitions and free cumulants. For further information on the combinatorial aspects of free probability, the reader is referred to the text [11]. Definition 1.10. (i) A partition π of a set S is a collection of disjoint, non-empty sets V1, . . . , Vr such that V1 ∪ · · · ∪ Vr = S. V1, . . . , Vr are called the blocks of π, and we set π = r. The collection of partitions of S will be denoted P(S), or in the case that S = {1, . . . , k} by P(k). (ii) If S is ordered, we say that π ∈ P(S) is non-crossing if whenever V, W are blocks of π and s1 < t1 < s2 < t2 are such that s1, s2 ∈ V and t1, t2 ∈ W , then V = W . The set of non-crossing partitions of S is denoted by N C(S), or by N C(k) in the case that S = {1, . . . , k}. (iii) The non-crossing partitions can also be defined recursively, a partition π ∈ P(S) is non-crossing if and only if it has a block V which is an interval, such that π \ V is a non-crossing partition of S \ V . (iv) Given π, σ ∈ P(S), we say that π ≤ σ if each block of π is contained in a block of σ. (v) Given i1, . . . , ik in some index set I, we denote by ker i the element of P(k) whose blocks are the equivalence classes of the relation Note that if π ∈ P(k), then π ≤ ker i is equivalent to the condition that whenever s and t are in the same block of π, is must equal it. s ∼ t ⇔ is = it. A CHARACTERIZATION OF FREENESS BY INVARIANCE UNDER QUANTUM SPREADING 5 Definition 1.11. Let (A, E) be a B-valued probability space. (i) For each k ∈ N, let ρ(k) : A⊗Bk → B be a linear map (the tensor product is with respect to the natural B − B bimodule structure on A). For n ∈ N and π ∈ N C(n), we define a linear map ρ(π) : A⊗B n → B recursively as follows. If π has only one block, we set ρ(π)[a1 ⊗ · · · ⊗ an] = ρ(n)(a1 ⊗ · · · ⊗ an) for any a1, . . . , an ∈ A. Otherwise, let V = {l + 1, . . . , l + s} be an interval of π. We then define, for any a1, . . . , an ∈ A, ρ(π)[a1 ⊗ · · · ⊗ an] = ρ(π\V )[a1 ⊗ · · · ⊗ al · ρ(s)(al+1 ⊗ · · · ⊗ al+s) ⊗ · · · ⊗ an]. For example, if π = {{1, 5, 8}, {2, 4}, {3}, {6, 7}, {9, 10}} ∈ N C(10), 1 2 3 4 5 6 7 8 9 10 then ρ(π)[a1 ⊗ · · · ⊗ a10] is given by ρ(3)(a1 · ρ(2)(a2 · ρ(1)(a3) ⊗ a4) ⊗ a5 · ρ(2)(a6 ⊗ a7) ⊗ a8) · ρ(2)(a9 ⊗ a10). (ii) For k ∈ N, define the B-valued moment functions E(k) : A⊗B k → B by E(k)[a1 ⊗ · · · ⊗ ak] = E[a1 · · · ak]. (iii) The B-valued cumulant functions κ(k) E : A⊗B k → B are defined recursively for π ∈ N C(k), k ≥ 1, by the moment-cumulant formula: for each n ∈ N and a1, . . . , an ∈ A we have E[a1 · · · an] = Xπ∈N C(n) κ(π) E [a1 ⊗ · · · ⊗ an]. 1.12. The cumulant functions can be solved for in terms of the moment functions by the following formula: for each n ∈ N and a1, . . . , an ∈ A, κ(π) E [a1 ⊗ · · · ⊗ an] = Xσ∈N C(n) σ≤π µn(σ, π)E(σ)[a1 ⊗ · · · ⊗ an], where µn is the Mobius function on the partially ordered set N C(n). The key relation between B-valued cumulant functions and free independence with amalga- mation is that freeness can be characterized in terms of the "vanishing of mixed cumulants". Theorem 1.13. ([14]) Let C be a unital ∗-algebra, (A, E) be a B-valued probability space and (ρi)i∈I a family of unital ∗-homomorphisms from C into A. Then the family (ρi)i∈I is free with amalgamation over B if and only if whenever i1, . . . , ik ∈ I, β1, . . . , βk ∈ C ∗ B and π ∈ N C(k) is such that π 6≤ ker i. κ(π) E [eρi1(β1) ⊗ · · · ⊗eρik(βk)] = 0 6 STEPHEN CURRAN Corollary 1.14. Let C be a unital ∗-algebra, (A, E) a B-valued probability space and (ρi)i∈N a family of unital ∗-homomorphisms from C into A. Then (ρi)i∈N is freely independent and identically distributed with respect to E if and only if κ(π) E[eρi1(β1) · · ·eρik (βk)] = Xπ∈N C(k) π≤ker i E [eρ1(β1) ⊗ · · · ⊗eρ1(βk)] for every k ∈ N, β1, . . . , βk ∈ C ∗ B and i1, . . . , ik ∈ I. (cid:3) 1.15. Quantum Permutation Group. Wang introduced the following noncommutative analogue of Sn in [17], and showed that it is the quantum automorphism group of a set with n points. For further information see [1],[2]. Definition 1.16. A matrix (uij)1≤i,j,≤n ∈ Mn(A), where A is a unital C∗-algebra, is called a magic unitary if (1) uij is a projection for each 1 ≤ i, j ≤ n. (2) uikuil = 0 and ukjulj = 0 if 1 ≤ i, j, k, l ≤ n and k 6= l. (3) For each 1 ≤ i, j ≤ n, uik = 1, nXk=1 ukj = 1. nXk=1 Note that the second condition in fact follows from the third. The quantum permutation group As(n) is defined as the universal C∗-algebra generated by elements {uij : 1 ≤ i, j ≤ n} such that (uij) is a magic unitary. As(n) is a compact quantum group in the sense of Woronowicz [18], with comultiplication, counit and antipode given by ∆(uij) = uik ⊗ ukj nXk=1 ǫ(uij) = δij S(uij) = uji. The existence of these maps is given by the universal property of As(n). 2. Quantum increasing sequences In this section we introduce objects Ai(k, n) which we call quantum increasing sequence spaces. As with Wang's quantum permutation group, the idea is to find a natural family of coordinates on the space of increasing sequences 1 ≤ l1 < · · · < lk ≤ n and "remove commutativity". Definition 2.1. For k, n ∈ N with k ≤ n, we define the quantum increasing sequence space Ai(k, n) to be the universal unital C∗-algebra generated by elements {uij : 1 ≤ i ≤ n, 1 ≤ j ≤ k} such that (1) uij is an orthogonal projection: u∗ (2) each column of the rectangular matrix u = (uij) forms a partition of unity: ij = uij = u2 ij. for 1 ≤ j ≤ k we have uij = 1. nXi=1 A CHARACTERIZATION OF FREENESS BY INVARIANCE UNDER QUANTUM SPREADING 7 (3) increasing sequence condition: uijui′j ′ = 0 if j < j′ and i ≥ i′. Remark 2.2. We note that the algebra Ai(k, n), together with the morphism α : Cn → Ck ⊗ Ai(k, n) defined by gives a quantum family of maps from {1, . . . , k} to {1, . . . , n}, in the sense of So ltan [13]. α(ei) = ej ⊗ uij, kXj=1 The motivation for the above definition is as follows. Consider the space Ik,n of increasing sequences l = (1 ≤ l1 < · · · < lk ≤ n). For 1 ≤ i ≤ n, 1 ≤ j ≤ k, define fij : Ik,n → C by fij(l) =(1, 0, lj = i lj 6= i . The functions fij generate C(Ik,n) by the Stone-Weierstrass theorem, and clearly satisfy the defining relations among the uij above. Moreover, it can be seen from the Gelfand theory that C(Ik,n) is the universal commutative C∗-algebra generated by {fij : 1 ≤ i ≤ n, 1 ≤ j ≤ k} satisfying these relations. In other words, C(Ik,n) is the abelianization of Ai(k, n). Remark 2.3. A first question is whether Ai(k, n) can be larger than C(Ik,n), i.e., "do quantum increasing sequences exist"? Clearly Ai(k, n) is commutative and hence equal to C(Ik,n) for k = 1. Using Lemma 2.4 below, it is not hard to see that Ai(k, n) is also commutative at k = n and n − 1. In particular we have Ai(k, n) = C(Ik,n) whenever n ≤ 3. However, if p, q are arbitrary projections in any unital C∗-algebra then the following gives a representation of Ai(2, 4): uijui′j ′ = uij(cid:18) mYl=1 nXil=1 uil(j+l)(cid:19)ui′j ′ = X1≤i1,...,im≤n uijui1(j+1) · · · uim(j+m)ui′(j+m+1). 1 − p p 0 0  0 0 q 1 − q  In particular, the free product C(Z2) ∗ C(Z2) is a quotient of Ai(2, 4) and hence Ai(2, 4) is infinite-dimensional. Observe that if (1 ≤ l1 < · · · < lk ≤ n) then we must have lj ′−lj ≥ j′−j for 1 ≤ j ≤ j′ ≤ k. In terms of the coordinates fij on C(Ik,n), this means that fijfi′j ′ = 0 if i′ − i < j′ − j. This relation also holds for the coordinates uij on Ai(k, n), which will be useful to our further analysis. Lemma 2.4. Fix k, n ∈ N with k ≤ n, and let {uij : 1 ≤ i ≤ n, 1 ≤ j ≤ k} be the standard generators of Ai(k, n). Then (1) uijui′j ′ = 0 if 1 ≤ j ≤ j′ ≤ k and i′ − i < j′ − j. (2) uij = 0 unless j ≤ i ≤ n − k + j, or equivalently k + i − n ≤ j ≤ i. Proof. (1) is trivial for j = j′, so fix 1 ≤ j < j′ ≤ k and set m = j′ − j − 1 ≥ 0. Then we have 8 STEPHEN CURRAN From the increasing sequence condition, each term in the sum is zero unless i < i1 < · · · < im < i′, which implies i′ − i ≥ m + 1 = j′ − j. For (2), note that from (1) we have ul1uij = 0 if i − l < j − 1, or equivalently l > i − j + 1. So if i < j then ul1uij = 0 for l = 1, . . . , n and we then have uij =(cid:18) nXl=1 uij = uij ·(cid:18) nXl=1 ul1(cid:19) · uij = 0. ulk(cid:19) = 0, Likewise we have uijulk = 0 if l < k + i − j, so if i > n − k + j then this holds for l = 1, . . . , n and which completes the proof. (cid:3) Now observe that any increasing sequence 1 ≤ l1 < · · · < lk ≤ n can be extended to a permutation in Sn which sends j to lj for 1 ≤ j ≤ k. One way to create such an extension is to set π(j) = lj for 1 ≤ j ≤ k, then inductively define π(k + m), for m = 1, . . . , n − k, by setting π(k + m) to be the least element of {1, . . . , n} \ {π(1), . . . , π(k + m − 1)}. After a moment's thought, one sees that m ≤ π(k + m) ≤ m + k and that π(k + m) = m + p exactly when lp < m + p but lp+1 > m + p for 1 ≤ m ≤ n − k and 0 ≤ p ≤ k, where we set l0 = −∞, lk+1 = ∞. This gives an inclusion of the space Ik,n of increasing sequences into Sn, which dualizes to a unital ∗-homomorphism C(Sn) → C(Ik,n). Consider the natural coordinates {fij : 1 ≤ i, j ≤ n} on Sn and {gij : 1 ≤ i ≤ n, 1 ≤ j ≤ k} on Ik,n. Clearly this map sends fij to gij for 1 ≤ i ≤ n, 1 ≤ j ≤ k. From the remark at the end of the previous paragraph, it follows that fi(k+m) is sent to 0 unless i = m + p for some 0 ≤ p ≤ k, and that f(m+p)(k+m) 7→ gip − g(i+1)(p+1), m+p−1Xi=0 where we set g00 = 1 and gi0 = g0i = gi(k+1) = 0 for i ≥ 1. For example, when k = 2 and n = 4 the matrix (fij) is as follows: 0 1 − g11 g11 g21 g22 g11 − g22 g31 g32 0 g42 g22 0 0 1 − g11 − g21 g11 + g21 − g22 − g32 g22 + g32   We can now use this formula to define a ∗-homomorphism from As(n) to Ai(k, n), which we might think of as "extending quantum increasing sequences to quantum permutations". Proposition 2.5. Fix natural numbers k < n. Let {vij : 1 ≤ i ≤ n, 1 ≤ j ≤ k}, {uij : 1 ≤ i, j ≤ n} be the standard generators of Ai(k, n), As(n), respectively. Then there is a unique unital ∗-homomorphism from As(n) to Ai(k, n) determined by • uij 7→ vij for 1 ≤ i ≤ n, 1 ≤ j ≤ k. • ui(k+m) 7→ 0 for 1 ≤ m ≤ n − k and i < m or i > m + k. • For 1 ≤ m ≤ n − k and 0 ≤ p ≤ k, u(m+p)(k+m) 7→ m+p−1Xi=0 vip − v(i+1)(p+1), A CHARACTERIZATION OF FREENESS BY INVARIANCE UNDER QUANTUM SPREADING 9 where we set v00 = 1 and vi0 = v0i = vi(k+1) = 0 for i ≥ 1. Proof. Let (vij) be the standard generators of Ai(k, n), and define {uij : 1 ≤ i, j ≤ n} in Ai(k, n) by • uij = vij for 1 ≤ i ≤ n, 1 ≤ j ≤ k. • ui(k+m) = 0 for 1 ≤ m ≤ n − k and i < m or i > m + k. • For 1 ≤ m ≤ n − k and 0 ≤ p ≤ k, u(m+p)(k+m) = m+p−1Xi=0 vip − v(i+1)(p+1), where we set v00 = 1 and vi0 = v0i = vi(k+1) = 0 for i ≥ 1. We need to show that (uij)1≤i,j≤n satisfies the magic unitary condition, and the result will then follow from the universal property of As(n). First let us check that uij is an orthogonal projection for 1 ≤ i, j ≤ n. The only non-trivial case is u(m+p)(k+m) for 1 ≤ m ≤ n − k and 0 ≤ p ≤ k. Here we just need to check that vl(p+1) ≤ m+p−1Xi=0 vip for 1 ≤ l ≤ m + p. The cases p = 0, k are trivial, so let 0 < p < k. We have vl(p+1) = vl(p+1) · vip = vl(p+1) · nXi=1 vip, l−1Xi=1 where we have applied the increasing sequence condition vl(p+1)vip = 0 for i ≥ l. So we have vl(p+1) ≤ vip ≤ l−1Xi=1 m+p−1Xi=0 vip as desired. Now we need to check that the sum along any row or column of (uij) gives the identity. For the first k columns, this follows from the defining relations of vij. For m = 1, . . . , n − k, the sum along column k + m gives ul(k+m) = nXl=1 = kXp=0 kXp=0 u(m+p)(k+m) m+p−1Xi=0 vip − v(i+1)(p+1) 10 STEPHEN CURRAN Now since vip = v(i+1)(p+1) = 0 if i < p, we continue with kXp=0 m+p−1Xi=p vip − v(i+1)(p+1) = = = kXp=0 m−1Xi=0 m−1Xi=0 = 1, v(i+p)p − v(i+p+1)(p+1) v(i+p)p − v(i+p+1)(p+1) m−1Xi=0 kXp=0 vi0 − v(i+k+1)(k+1) since the only nonzero term in the last sum is v00 = 1. It now remains only to show that the sum along any row of (uij) gives the identity. We have nXj=1 uij = = = = = kXj=1 kXj=1 kXj=1 kXj=1 kXj=1 m+p=i uij + vij + uij + ui(k+m) ui(k+m) n−kXm=1 X0≤p≤k min{i,n−k}Xm=max{i−k,1} min{i,n−k}Xm=max{i−k,1} i−1Xl=0 vij +(cid:18) min{i,n−k}Xm=max{i−k,1} v0(i−m) − vi(i−m+1)(cid:19) + vij +(cid:18) min{i,n−k}Xm=max{i−k,1} v0(i−m) − vi(i−m+1)(cid:19) + vl(i−m) − v(l+1)(i−m+1) min{i,n−k}Xm=max{i−k,1} vl(i−m) − vl(i−m+1) vl max{0,k+i−n} − vl min{k+1,i}. i−1Xl=1 i−1Xl=1 Now note that if 1 ≤ l ≤ i − 1 then vl min{k+1,i} = 0, indeed this is true by definition if min{k + 1, i} = k + 1, and if min{k + 1, i} = i then vli = 0 since l < i. Also we have vij = 0 unless k + i − n ≤ j ≤ i. Plugging this in above and rearranging terms, we have min{k,i}Xj=max{1,k+i−n} vij − min{i,n−k}Xm=max{i−k,1} vi(i−m+1) + min{i,n−k}Xm=max{i−k,1} v0(i−m) + i−1Xl=1 vl max{0,k+i−n}. After reindexing the second sum and combining with the first, we obtain max{1,k+i+1−n}−1Xj=max{1,k+i−n} vij + min{i,n−k}Xm=max{i−k,1} v0(i−m) + i−1Xl=1 vl max{0,k+i−n}. A CHARACTERIZATION OF FREENESS BY INVARIANCE UNDER QUANTUM SPREADING 11 Now if i ≤ n − k, then the first and third sums are zero while the second is 1. If i > n − k then the second sum is zero and the first and third combine as Now since vl(k+i−n) = 0 if l > n − k + (k + i − n) = i, we have vl(k+i−n). iXl=1 vl(k+i−n) = vl(k+i−n) = 1. iXl=1 So (uij) does indeed satisfy the magic unitary condition, which completes the proof. (cid:3) 3. Quantum invariant sequences of random variables In this section we introduce the notions of quantum exchangeability and quantum spread- ability for sequences of noncommutative random variables, and prove the implications (i)⇒ (ii) and (iii) ⇒ (i) in Theorem 1. First let us recall the notion of quantum exchangeability from [10] (see also [4]). Let C be a unital ∗-algebra. For each n ∈ N there is a unique unital ∗-homomorphism αn : Cn → Cn ⊗ As(n) determined by αn(c(j)) = c(i) ⊗ uij nXl=1 nXi=1 for c ∈ C and 1 ≤ j ≤ n, indeed this follows from the relations in As(n) and the universal property of the free product Cn = C (1) ∗ · · · ∗ C (n). Moreover αn is a right coaction of As(n) in the sense that (αn ⊗ id) ◦ αn = (id ⊗ αn) ◦ αn (id ⊗ ǫ) ◦ αn = id, see [4] for details. The coaction αn may be regarded as "quantum permuting" the n copies of C inside Cn. Definition 3.1. Let C be a unital ∗-algebra, (A, ϕ) a noncommutative probability space and (ρ1, . . . , ρn) a sequence of unital ∗-homomorphisms of C into A. We say that the distribution ϕρ is invariant under quantum permutations, or that the sequence is quantum exchangeable, if ϕρ is invariant under the coaction αn, i.e., for any c ∈ Cn. This is extended to infinite sequences (ρi)i∈N by requiring that (ρ1, . . . , ρn) is quantum (ϕρ ⊗ id)αn(c) = ϕρ(c)1As(n) exchangeable for each n ∈ N. 3.2. Remarks. (1) More explicitly, this amounts to the condition that ϕ(ρi1(c1) · · · ρik (ck))ui1j1 · · · uikjk = ϕ(ρj1(c1) · · · ρjk(ck)) · 1 X1≤i1,...,ik≤n for any c1, . . . , ck ∈ C and 1 ≤ j1, . . . , jk ≤ n, where uij are the standard generators of As(n). 12 STEPHEN CURRAN (2) By the universal property of As(n), the sequence (ρ1, . . . , ρn) is quantum exchangeable if and only if the equation in (1) holds for any family {uij : 1 ≤ i, j ≤ n} of projections in a unital C∗-algebra B such that (uij) ∈ Mn(B) is a magic unitary matrix. (3) For 1 ≤ i, j ≤ n, define fij ∈ C(Sn) by fij(π) = δiπ(j). The matrix (fij) is a magic unitary, and the equation in (1) becomes ϕ(ρj1(c1) · · · ρjk (ck))1C(Sn) = X1≤i1,...,ik≤n Evaluating both sides at π ∈ Sn, we find ϕ(ρi1(c1) · · · ρin(cn))fi1j1 · · · fikjk. ϕ(ρj1(c1) · · · ρjk (ck)) = ϕ(ρπ(j1)(c1) · · · ρπ(jk)(ck)), so that quantum exchangeability implies invariance under classical permutations. It is shown in [10] that any sequence (ρ1, . . . , ρn) which is freely independent and iden- tically distributed with respect to a conditional expectation which preserves ϕ is quantum exchangeable. For the convenience of the reader we include a sketch of the proof, and re- fer to that paper for details. Note that the implication (iii) ⇒ (i) in Theorem 1 follows immediately. Proposition 3.3. Let C be a unital ∗-algebra and (ρ1, ρ2, . . . , ρn) a sequence of unital ∗- homomorphisms from C into a noncommutative probability space (A, ϕ). Let B ⊂ A be a unital ∗-subalgebra and suppose that there is a ϕ-preserving conditional expectation E : A → B such that (ρ1, . . . , ρn) is freely independent and identically distributed with respect to E. Then (ρ1, . . . , ρn) is quantum exchangeable. Proof. Let c1, . . . , ck ∈ C and 1 ≤ j1, . . . , jk ≤ n. We have ϕ(E[ρi1(c1) · · · ρik(ck)])ui1j1 · · · uikjk X1≤i1,...,ik≤n ϕ(ρi1(c1) · · · ρik(ck))ui1j1 · · · uikjk = X1≤i1,...,ik≤n = X1≤i1,...,ik≤n Xπ∈N C(k) = Xπ∈N C(k) ϕ(κ(π) ϕ(κ(π) π≤ker i E [ρ1(c1) ⊗ · · · ⊗ ρ1(ck)])ui1j1 · · · uikjk E [ρ1(c1) ⊗ · · · ⊗ ρ1(ck)]) X1≤i1,...,ik≤n π≤ker i ui1j1 · · · uikjk, where in the second line we have applied Corollary 1.14. It can be seen from induction on the number of blocks of π that ui1j1 · · · uikjk =(1As(n), π ≤ ker j otherwise 0, , X1≤i1,...,ik≤n π≤ker i and it follows that X1≤i1,...,ik≤n ϕ(ρi1(c1) · · · ρik(ck))ui1j1 · · · uikjk = Xπ∈N C(k) π≤ker j ϕ(κ(π) E [ρ1(c1) ⊗ · · · ⊗ ρ1(ck)])1As(n) where again we have applied Corollary 1.14. (cid:3) = ϕ(ρj1(c1) · · · ρjk (ck))1As(n), A CHARACTERIZATION OF FREENESS BY INVARIANCE UNDER QUANTUM SPREADING 13 We will now introduce the quantum spreadability condition. Let C be a unital ∗-algebra, then for any natural numbers k ≤ n there is a unique unital ∗-homomorphism αk,n : Ck → Cn ⊗ Ai(k, n) determined by αk,n(c(j)) = c(i) ⊗ uij nXi=1 for c ∈ C and 1 ≤ j ≤ k, indeed this follows as above from the relations in Ai(k, n) and the universal property of Ck. Definition 3.4. Let C be a unital ∗-algebra and (ρ1, ρ2, . . . , ρn) a sequence of unital ∗- homomorphisms from C into a noncommutative probability space (A, ϕ). We say that the distribution ϕρ is invariant under quantum spreading, or that the sequence is quantum spread- able, if for each k = 1, . . . , n the distribution ϕρ is invariant under αk,n in the sense that (ϕρ ⊗ id)αk,n(c) = ϕρ(c)1Ai(k,n) for any c ∈ Ck. An infinite sequence (ρ1, ρ2, . . . ) is called quantum spreadable if (ρ1, . . . , ρn) is quantum spreadable for each n. Remark 3.5. (1) Explicitly, the condition is that for each k = 1, . . . , n we have ϕ(ρj1(c1) · · · ρjm(cm)) · 1 = X1≤i1,...,im≤n ϕ(ρi1(c1) · · · ρim(cm)) · ui1j1 · · · uimjm for all 1 ≤ j1, . . . , jm ≤ k and c1, . . . , cm ∈ C, where (uij) denote the standard generators of Ai(k, n). (2) From the universal property of Ai(k, n), the sequence (ρ1, . . . , ρn) is quantum spread- able if and only if for each 1 ≤ k ≤ n, equation (1) holds for any family {uij : 1 ≤ i ≤ n, 1 ≤ j ≤ k} of projections in a unital C∗-algebra B which satisfy the definining relations of Ai(k, n). (3) Let (fij) denote the generators of C(Ik,n) introduced in Section 2. Plugging fij into equation (1) and applying both sides to l = (1 ≤ l1 < · · · < lk ≤ n), we have ϕ(ρj1(c1) · · · ρjm(cm)) = X1≤i1,...,im≤n = ϕ(ρlj1 (c1) · · · ρljm (cm)) ϕ(ρi1(c1) · · · ρim(cm)fi1j1(l) · · · fimjm(l) for any 1 ≤ j1, . . . , jm ≤ k. So (ρ1, . . . , ρk) has the same distribution as (ρl1, . . . , ρlk), and hence quantum spreadability implies classical spreadability. In particular, quan- tum spreadable sequences are identically distributed. We can now prove the implication (i) ⇒ (ii) of Theorem 1, this holds in fact for finite sequences and in a purely algebraic context: Proposition 3.6. Let C be a unital ∗-algebra and (ρ1, . . . , ρn) be a sequence of unital ∗- homomorphisms from C into a noncommutative probability space (A, ϕ). If the sequence (ρ1, . . . , ρn) is quantum exchangeable, then it is quantum spreadable. 14 STEPHEN CURRAN Proof. Fix 1 ≤ k ≤ n and let {vij : 1 ≤ i ≤ n, 1 ≤ j ≤ k} and {uij : 1 ≤ i, j ≤ n} be the standard generators of Ai(k, n) and As(n), respectively. Assume (ρ1, . . . , ρn) is quantum exchangeable, and fix 1 ≤ j1, . . . , jm ≤ k and c1, . . . , cm ∈ C. We have ϕ(ρj1(c1) · · · ρjm(cm))1As(n) = X1≤i1,...,im≤n ϕ(ρi1(c1) · · · ρim(cm)) · ui1j1 · · · uimjm. By Proposition 2.5, there is a unital ∗-homomorphism from As(n) to Ai(k, n) which sends uij to vij for 1 ≤ i ≤ n, 1 ≤ j ≤ k. Applying this map to both sides of the above equation, we obtain ϕ(ρj1(c1) · · · ρjm(cm))1Ai(k,n) = X1≤i1,...,im≤n ϕ(ρi1(c1) · · · ρim(cm)) · vi1j1 · · · vimjm, so that (ρ1, . . . , ρn) is quantum spreadable as desired. (cid:3) 4. Quantum spreadability implies freeness with amalgamation 4.1. In this section we will complete the proof of Theorem 1. Throughout this section we will assume that C is a unital ∗-algebra, and that (ρi)i∈N is an infinite sequence of unital ∗-homomorphisms from C into a W∗-probability space (M, τ ). B will denote the tail algebra: B = \n≥1 W ∗(cid:0){ρi(c) : c ∈ C, i ≥ n}(cid:1). 1 eαk,n(b0c(j1) eαk,n(b0c(j) 1 L2(M) will denote the Hilbert space given by the GNS-representation for τ . Since τ is a trace, there is a unique conditional expectation E : M → B given my E[m] = P (m), where P is the orthogonal projection of L2(M) onto L2(B). We will assume without loss of generality that M is generated by ρ∞(C∞), i.e., M = W ∗(cid:0){ρi(c) : i ∈ I, c ∈ C}(cid:1). Observe that if the sequence (ρi)i∈N is spreadable and hence stationary, the linear map determined by U(ρi1(c1) · · · ρim(cm)) = ρi1+1(c1) · · · ρim+1(cm) for i1, . . . , im ∈ N and c1, . . . , cm ∈ C, is well-defined and extends to an isometry U : L2(M) → L2(M). also invariant under quantum spreading. By this we mean that the joint distribution Eρ is Recall from Definition 1.8 that we set eρi = ρi ∗ id : C ∗ B → M. We will begin by showing that if (ρi)i∈N is quantum spreadable, then the B-valued distribution of (eρi)i∈N is invariant under the ∗-homomorphisms eαk,n : Ck ∗ B → (Cn ∗ B) ⊗ Ai(k, n) determined by m bm ⊗ ui1j1 · · · uimjm b0c(i1) 1 b1 · · · c(im) b1 · · · c(jm) m bm) = X1≤i1,...,im≤n for all k ≤ n, 1 ≤ j1, . . . , jm ≤ k, b0, . . . , bm ∈ B and c1, . . . , cm ∈ C. Note that if 1 ≤ j ≤ k, b0, . . . , bm ∈ B and c1, . . . , cm ∈ C then · · · c(j) m bm) = X1≤i1,...,im≤n b0c(i1) 1 · · · c(im) m bm ⊗ ui1j · · · uimj = b0c(i) 1 · · · c(i) m bm ⊗ uij, nXi=1 A CHARACTERIZATION OF FREENESS BY INVARIANCE UNDER QUANTUM SPREADING 15 from which it follows that if β ∈ C ∗ B then Proposition 4.2. Suppose that the sequence (ρi)i∈N is quantum spreadable. Then the joint eαk,n(β(j)) = nXi=1 β(i) ⊗ uij. where the equality holds in B ⊗ Ai(k, n). for each k ≤ n, 1 ≤ j1, . . . , jm ≤ k and β1, . . . , βm ∈ C ∗ B we have Proof. We need to show that if 1 ≤ j1, . . . , jm ≤ k, b0, . . . , bm ∈ B and c1, . . . , cm ∈ C then E[eρi1(β1) · · ·eρim(βm)] ⊗ ui1j1 · · · uimjm, distribution of (eρi)i∈N with respect to E is invariant under quantum spreading. Explicitly, E[eρj1(β1) · · ·eρjm(βm)] ⊗ 1Ai(k,n) = X1≤i1,...,im≤n E[b0ρj1(c1) · · · ρjm(cm)bm] ⊗ 1 = X1≤i1,...,im≤n τ (b0ρj1(c1) · · · ρjm(cm)bm) ⊗ 1 = X1≤i1,...,im≤n Since E preserves the faithful state τ , it suffices to show that E[b0ρi1(c1) · · · ρim(cm)bm] ⊗ ui1j1 · · · uimjm. τ (b0ρi1(c1) · · · ρim(cm)bm) ⊗ ui1j1 · · · uimjm. We will show that this in fact holds for b0, . . . , bm in W ∗({ρi(c) : i > k, c ∈ C}). By Kaplansky's density theorem, it suffices to consider the case that b0, . . . , bm are elements of the form ρl1(d1) · · · ρlr (dr) for k < l1, . . . , lr ≤ N and d1, . . . , dr ∈ C. To show this, we extend (uij) to a (n + N) × (k + N) matrix by setting uij, δ(i−n)(j−k), 0, 1 ≤ i ≤ n, 1 ≤ j ≤ k i > n, j > k otherwise vij = Observe if b = ρl1(d1) . . . ρlr (dr) is as above, then X1≤i1,...,ir≤n+N ρi1(d1) · · · ρim(dr) ⊗ vi1l1 · · · virlr = ρl1+(n−k)(d1) · · · ρlr+(n−k)(dr) ⊗ 1Ai(k,n) = U (n−k)(b) ⊗ 1Ai(k,n). Now it is clear that (vij) satisfies the defining relations of Ai(k + N, n + N), so applying the quantum spreadability condition with (vij), we have τ (b0ρj1(c1) · · · ρjm(cm)bm) But since (ρi)i∈N is spreadable, the right hand side is equal to = X1≤i1,...,im≤n X1≤i1,...,im≤n τ(cid:0)U (n−k)(b0)ρi1(c1) · · · ρim(cm)U (n−k)(bm)(cid:1) ⊗ ui1j1 · · · uimjm. τ(cid:0)b0ρi1(c1) · · · ρim(cm)bm(cid:1) ⊗ ui1j1 · · · uimjm, which completes the proof. (cid:3) 16 STEPHEN CURRAN 4.3. The key ingredient in our proof that an infinite quantum spreadable sequence is free with amalgamation is a "measure" on the space of quantum increasing sequences, i.e., a state on Ai(k, n). Unlike in the classical case, there does not appear to be a good notion of "uniform" measure on this quantum space. Instead, we will use the measures induced by a certain representation of Ai(k, k · n). Proposition 4.4. Fix k, n ∈ N. Then there is a state ψk,n : Ai(k, k · n) → C such that: (1) (2) unless (jr − 1) · n < lr ≤ jr · n for r = 1, . . . , m. ψk,n(ul1j1 · · · ulmjm) = 0 ψk,n(u((j1−1)·n+i1)j1 · · · u((jm−1)·n+im)jm) = Xπ∈N C(m) π≤ker j Xσ∈N C(m) σ≤π∧ker i µm(σ, π)n−σ for all 1 ≤ j1, . . . , jm ≤ k and 1 ≤ i1, . . . , im ≤ n. Proof. Let {pij : 1 ≤ i ≤ n, 1 ≤ j ≤ k} be projections in a C∗-probability space (A, ϕ) such that (1) The families ({pi1 : 1 ≤ i ≤ n}, . . . , {pik : 1 ≤ i ≤ n}) are freely independent. (2) For j = 1, . . . , k, we have and ϕ(pij) = n−1 for 1 ≤ i ≤ n. pij = 1, nXi=1 Define {ulj : 1 ≤ l ≤ kn, 1 ≤ j ≤ k} by ulj = 0 unless (j − 1) · n < l ≤ j · n, and for 1 ≤ i ≤ n, so that (ulj) is given by the following matrix: u((j−1)·n+i)j = pij  Clearly (ulj) satisfies the defining relations of Ai(k, k · n) and so we obtain a unital ∗- homorphism from Ai(k, k · n) into A. Composing with ϕ gives a state ψk,n : Ai(k, k · n) → C, and we need only show that (ulj) in (A, ϕ) has the distribution appearing in the statement. p11 ... p1n 0 ... 0 0 ... 0 ... 0 0 ... 0 p21 ... p2n 0 ... 0 ... 0 · · · . . . · · · · · · . . . · · · · · · . . . · · · . . . · · · 0 ... 0 0 ... 0 0 ... pk1 ... pkn  A CHARACTERIZATION OF FREENESS BY INVARIANCE UNDER QUANTUM SPREADING 17 (1) is trivial, as ul1j1 · · · ulmjm = 0 unless (jr − 1) · n < lr ≤ jr · n for r = 1, . . . , m. For (2), we need to show that ϕ(pi1j1 · · · pimjm) = Xπ∈N C(m) π≤ker j Xσ∈N C(m) σ≤π∧ker i µm(σ, π)n−σ. Now by freeness, we have µm(σ, π)ϕ(σ)[pi1j1 ⊗ · · · ⊗ pimjm]. π≤ker j κ(π)[pi1j1 ⊗ · · · ⊗ pimjm] ϕ(pi1j1 · · · pimjm) = Xπ∈N C(m) π≤ker j Xσ∈N C(m) = Xπ∈N C(m) ϕ(pl1j · · · plsj) =(n−1, ϕ(σ)[pi1j1 ⊗ · · · ⊗ pimjm] =(n−σ, σ ≤ ker i l1 = · · · = ls otherwise σ 6≤ ker i , . σ≤π 0, 0, Now since for 1 ≤ j, l1, . . . , ls ≤ n we have it follows that if σ ≤ ker j then Combining this with the previous equation yields the desired result. (cid:3) Remark 4.5. Observe that the formula in (2) above has a very similar structure to the highest order expansion of the Weingarten formula for evaluating integrals over the quantum permutation group As(n) with respect to its Haar state, see [2, 4]. The final tool which we require to complete the proof of Theorem 1 is von Neumann's mean ergodic theorem. This will allow us to give a formula for the expectation functionals E(σ) as certain weighted averages. We note that the unpleasant indices which appear are chosen as to correspond to the formula in Proposition 4.4. Lemma 4.6. Suppose that the sequence (ρi)i∈N is quantum spreadable. Then for any j ∈ N and β ∈ C ∗ B, we have with convergence in 2. E[eρ1(β)] = lim n→∞ 1 n nXi=1 eρ(j−1)·n+i(β), Proof. Since (ρi)i∈N is spreadable, we have τ (m1m2) = τ (m1U(m2)) whenever m1 ∈ W ∗({ρi(c) : 1 ≤ i ≤ n, c ∈ C}) and m2 ∈ W ∗({ρi(c) : i > n, c ∈ C}). It follows that for m ∈ M and b ∈ B, hence b = U(b). It follows easily that τ (mb) = τ (mU(b)) U(eρi(β)) =eρi+1(β) 18 STEPHEN CURRAN for any i ∈ N and β ∈ C ∗ B. Since it is clear that any vector fixed by U must lie in L2(B), we have in fact the equality L2(B) = {ξ ∈ L2(M) : U ξ = ξ}. By von Neumann's mean ergodic theorem, we have lim n→∞ 1 n n−1Xi=0 U i = P, where P is the orthogonal projection of L2(M) onto L2(B) and the limit holds in the strong operator topology. Therefore for any m ∈ M we have E[m] = P (m) = lim n→∞ U i(m), 1 n n−1Xi=0 with the limit holding in 2. Since U is contractive in 2, we have also for any j ∈ N that lim n→∞ 1 n n−1Xi=0 U (j−1)·n+i(m) = lim n→∞ = lim n→∞ U (j−1)·n(cid:18) 1 n U (j−1)·nP (m) U i(m)(cid:19) n−1Xi=0 = E[m], since U · P = P . Applying this to m =eρ1(β) gives the desired result. Proposition 4.7. Suppose that the sequence (ρi)i∈N is quantum spreadable. Fix j1, . . . , jm ∈ N and choose σ ∈ N C(m) such that σ ≤ ker j. Then for any β1, . . . , βm ∈ C ∗ B, we have (cid:3) E(σ)[eρ1(β1) ⊗ · · · ⊗eρ1(βm)] = lim n→∞ n−σ X1≤i1,...,im≤n σ≤ker i eρ(j1−1)·n+i1(β1) · · ·eρ(jm−1)·n+im(βm), with convergence in 2. Proof. We will use induction on the number of blocks of σ. If σ = 1m has only one block, then σ ≤ ker j implies j1 = · · · = jm and we have lim n→∞ n−σ X1≤i1,...,im≤n σ≤ker i eρ(j1−1)·n+i1(β1) · · ·eρ(jm−1)·n+im(βm) = lim n→∞ 1 n nXi=1 eρ(j1−1)·n+i(β1β2 · · · βm). By Lemma 4.6, this converges in 2 to E[eρ1(β1β2 · · · βm)] = E(σ)[eρ1(β1) ⊗ · · · ⊗eρ1(βm)]. A CHARACTERIZATION OF FREENESS BY INVARIANCE UNDER QUANTUM SPREADING 19 Now let σ ∈ N C(m) and let V = {l + 1, . . . , l + s} be an interval of σ, and let j be the common value of jl+1, . . . , jl+s. We have n σ≤ker i il+s+1,...,im≤n eρ(j1−1)·n+i1(β1) · · ·eρ(jm−1)·n+im(βm) σ\V ≤ker i eρ(j1−1)·n+i1(β1) · · ·(cid:16) 1 n−σ X1≤i1,...,im≤n nXi=1 eρ(j−1)·n+i(βl+1 · · · βl+s)(cid:17) · · ·eρ(jm−1)·n+im(βm) = n−σ\V X1≤i1,...,il, As above, the interior sum converges to E[eρ1(βl+1 · · · βl+s)] in 2 as n → ∞. Now for any β ∈ C ∗ B, since the variables eρi(β) are identically ∗-distributed with respect to the faithful trace τ , it follows that keρi(β)k is independent of i. Therefore there is a constant D such that nXi=1 eρj(βl+1 · · · βl+s)(cid:17) · · ·eρ(jm−1)·n+im(βm) σ\V ≤ker i eρ(j1−1)·n+i1(β1) · · · E[eρ1(βl+1 · · · βl+s)] · · ·eρ(jm−1)·n+im(βm). eρi1(β1) · · ·eρil(βl) · ξ ·eρil+s+1(βl+s+1) · · ·eρim(βm)2 ≤ Dξ2 σ\V ≤ker i eρ(j1−1)·n+i1(β1) · · ·(cid:16) 1 n−σ\V X1≤i1,...,il, for any ξ ∈ L2(M) and i1, . . . , im ∈ N. It follows that n−σ\V X1≤i1,...,il, = lim n→∞ il+s+1,...,im≤n il+s+1,...,im≤n lim n→∞ n By induction, this converges in 2 to E(σ\V )[eρ1(β1) ⊗ · · · ⊗eρ1(βl) · E[eρ1(βl+1 · · · βl+s)] ⊗ · · · ⊗eρ1(cm)], which is precisely E(σ)[eρ1(β1) ⊗ · · · ⊗eρ1(βm)], as desired. We can now complete the proof of Theorem 1. (cid:3) Proof of (ii)⇒(iii). Fix β1, . . . , βm ∈ C ∗ B and 1 ≤ j1, . . . , jm ≤ k. By Proposition 4.2, for each n ∈ N we have Applying (id ⊗ ψk,n), with ψk,n from Proposition 4.4, to each side of the above equation, we obtain E[eρl1(β1) · · ·eρlm(βm)] ⊗ ul1j1 · · · ulmjm. E[eρj1(β1) · · ·eρjm(βm)] ⊗ 1Ai(k,k·n) = X1≤l1,...,lm≤kn E[eρj1(β1) · · ·eρjm(βm)] = X1≤i1,...,im≤n = Xπ∈N C(m) π≤ker j Xσ∈N C(m) µm(σ, π)Ehn−σ X1≤i1,...,im≤n E[eρ(j1−1)·n+i1(β1) · · ·eρ(jm−1)·n+im(βm)] Xπ∈N C(m) σ≤ker i σ≤π µm(σ, π)n−σ π≤ker j Xσ∈N C(m) eρ(j1−1)·n+i1(β1) · · ·eρ(jm−1)·n+im(βm)i. σ≤π∧ker i 20 STEPHEN CURRAN Letting n → ∞ and applying Proposition 4.7, we have π≤ker j Xσ∈N C(m) E[eρj1(β1) · · ·eρjm(βm)] = Xπ∈N C(m) = Xπ∈N C(m) κ(π) π≤ker j µm(σ, π)E(σ)[eρ1(β1) ⊗ · · · ⊗eρ1(βm)] σ≤π E [eρ1(β1) ⊗ · · · ⊗eρ1(βm)], and the result now follows from Corollary 1.14. (cid:3) Acknowledgement. I would like to thank my thesis advisor, Dan-Virgil Voiculescu, for his continued guidance and support while completing this project. References [1] T. Banica, J. Bichon, and B. Collins, Quantum permutation groups: a survey, in Noncommutative harmonic analysis with applications to probability, vol. 78 of Banach Center Publ., Polish Acad. Sci. Inst. Math., Warsaw, 2007, 13 -- 34. [2] T. Banica and B. Collins, Integration over quantum permutation groups, J. Funct. Anal., 242 (2007), 641 -- 657. [3] T. Banica, S. Curran, and R. Speicher, De Finetti theorems for easy quantum groups, Ann. Probab., to appear. arXiv:0907.3314 [math.OA], 2009. [4] S. Curran, Quantum exchangeable sequences of algebras, Indiana Univ. Math. J. 58 (2009), 1097 -- 1126. [5] S. Curran, Quantum rotatability, Trans. Amer. Math. Soc. 362 (2010), 4831 -- 4851. [6] E. Hewitt and L. J. Savage, Symmetric measures on Cartesian products, Trans. Amer. Math. Soc. 80 (1955), 470 -- 501. [7] O. Kallenberg, Spreading-invariant sequences and processes on bounded index sets, Probab. Theory Related Fields 118 (2000), 211 -- 250. [8] O. Kallenberg, Probabilistic symmetries and invariance principles, Probability and its Applications, Springer, New York, 2005. [9] C. Kostler, A noncommutative extended de Finetti theorem, J. Funct. Anal. 258 (2010), 1073 -- 1120. [10] C. Kostler and R. Speicher, A noncommutative de Finetti theorem: invariance under quantum permu- tations is equivalent to freeness with amalgamation, Comm. Math. Phys. 291 (2009), 473 -- 490. [11] A. Nica and R. Speicher, Lectures on the combinatorics of free probability, vol. 335 of London Mathe- matical Society Lecture Note Series, Cambridge University Press, Cambridge, 2006. [12] C. Ryll-Nardzewski, On stationary sequences of random variables and the de Finetti's equivalence, Colloq. Math. 4 (1957), 149 -- 156. [13] P. M. So ltan, Quantum families of maps and quantum semigroups on finite quantum spaces, J. Geom. Phys. 59 (2009), 354 -- 368. [14] R. Speicher, Combinatorial theory of the free product with amalgamation and operator-valued free probability theory, Mem. Amer. Math. Soc., 132 (1998), x+88. [15] D. Voiculescu, Symmetries of some reduced free product C ∗-algebras, in Operator algebras and their connections with topology and ergodic theory, vol. 1132 of Lecture Notes in Math., Springer, Berlin, 1985, 556 -- 588. [16] D. Voiculescu, K. Dykema, and A. Nica, Free random variables, vol. 1 of CRM Monograph Series, American Mathematical Society, Providence, RI, 1992. [17] S. Wang, Quantum symmetry groups of finite spaces, Comm. Math. Phys. 195 (1998), 195 -- 211. [18] S. L. Woronowicz, Compact matrix pseudogroups, Comm. Math. Phys. 111 (1987), 613 -- 665. Department of Mathematics, University of California, Los Angeles, CA 90095, USA. E-mail address: [email protected] URL: http://www.math.ucla.edu/~curransr
1707.05940
1
1707
2017-07-19T05:44:53
Semigroup C*-algebras
[ "math.OA" ]
We give an overview of some recent developments in semigroup C*-algebras.
math.OA
math
SEMIGROUP C*-ALGEBRAS XIN LI Abstract. We give an overview of some recent developments in semigroup C*-algebras. Contents Introduction 1. 2. C*-algebras generated by left regular representations 3. Examples 3.1. The natural numbers 3.2. Positive cones in totally ordered groups 3.3. Monoids given by presentations 3.4. Examples from rings in general, and number theory in particular 3.5. Finitely generated abelian cancellative semigroups 4. Preliminaries 4.1. Embedding semigroups into groups 4.2. Graph products 4.3. Krull rings 5. C*-algebras attached to inverse semigroups, partial dynamical systems, and groupoids Inverse semigroups ´Etale groupoids 5.1. 5.2. Partial dynamical systems 5.3. 5.4. The universal groupoid of an inverse semigroup 5.5. 5.6. C*-algebras of partial dynamical systems as C*-algebras of partial Inverse semigroup C*-algebras as groupoid C*-algebras transformation groupoids 5.7. The case of inverse semigroups admitting an idempotent pure partial homomorphism to a group 6. Amenability and nuclearity 6.1. Groups and groupoids 6.2. Amenability for semigroups 6.3. Comparing reduced C*-algebras for left cancellative semigroups and their left inverse hulls 6.4. C*-algebras generated by semigroups of projections 6.5. The independence condition 6.6. Construction of full semigroup C*-algebras 6.7. Crossed product and groupoid C*-algebra descriptions of reduced semigroup C*-algebras 6.8. Amenability of semigroups in terms of C*-algebras 1 2 3 4 4 4 5 8 9 9 9 11 14 16 16 22 26 29 30 33 36 37 37 40 42 47 54 61 63 66 2 XIN LI 6.9. Nuclearity of semigroup C*-algebras and the connection to amenability 68 7. Topological freeness, boundary quotients, and C*-simplicity 69 80 8. The Toeplitz condition 86 9. Graph products 87 9.1. Constructible right ideals 9.2. The independence condition 91 94 9.3. The Toeplitz condition 10. K-theory 97 11. Further developments, outlook, and open questions References 99 101 1. Introduction A semigroup C*-algebra is the C*-algebra generated by the left regular represen- tation of a left cancellative semigroup. In the case of groups, this is the classical construction of reduced group C*-algebras, which received great interest and serves as a motivating class of examples in operator algebras. For semigroups which are far from being groups, we encounter completely new phenomena which are not visible in the group case. It is therefore a natural and interesting task to try to understand and explain these new phenomena. This challenge has been taken up by several authors in many pieces of work, and our present goal is to give a unified treatment of this endeavour. We point out that particular classes of semigroups played a predominant role in the development, as they serve as our motivation and guide us towards important properties of semigroups which allow for a systematic study of their C*-algebras. The examples include positive cones in totally ordered groups, semigroups given by particular presentations and semigroups coming from rings of number-theoretic origin. Important properties that isolate from the general and wild class of all left cancellative semigroups a manageable subclass were first given by Nica's quasi- lattice order [Nic92] and later on by the independence condition [Li12] and the Toeplitz condition [Li13]. Aspects of semigroup C*-algebras which we would like to discuss in the following include descriptions as crossed products and groupoid C*-algebras, the connec- tion between amenability and nuclearity, boundary quotients, and the classification problem for semigroup C*-algebras. The first three topics are discussed in detail, and we give a more or less self-contained presentation. The last topic puts to- gether many results. In particular, it builds on the K-theory computations that are explained in detail by S. Echterhoff in [Ech17]. Since a detailed account of clas- sification results would take too much space, we just briefly summarize the main SEMIGROUP C*-ALGEBRAS 3 results, and refer the interested reader to the relevant papers for more details and complete proofs. Our discussion of semigroup C*-algebras builds on previous work of J. Renault on groupoids and their C*-algebras [Ren80], and the work of R. Exel on C*-algebras of inverse semigroups, their quotients corresponding to tight representations of inverse semigroups, and on partial actions [Exe08, Exe09, Exe15]. Inevitably, certain interesting aspects of semigroup C*-algebras are not covered in this book. This includes a discussion of C*-algebras of semigroups which do not embed into groups such as general right LCM semigroups (see [Sta15b]) or Zappa- Sz´ep products (see [BRRW14]), or C*-algebras of certain topological semigroups (see [RS15, Sun14]). Moreover, we do not discuss KMS-states in detail, but we refer the reader to [LR10, BaHLR11, CDL13, CaHR15] for more information. We also mention that in [Cun17a], J. Cuntz describes KMS-states for particular examples. We apologize for these omissions and try to make up for them by pointing the interested reader to the relevant literature. To this end, we have included a long (but not complete) list of references. 2. C*-algebras generated by left regular representations Let P be a semigroup. We assume that P is left cancellative, i.e., for all p, x, y ∈ P , px = py implies x = y. In other words, the map P → P, x 7→ px given by left multiplication with p ∈ P is injective for all p ∈ P . The left regular representation of P is given as follows: The Hilbert space ℓ2P comes with a canonical orthonormal basis {δx : x ∈ P}. Here δx is the delta-function in x ∈ P , defined by δx(y) = 1 if y = x and δx(y) = 0 if y 6= x. For every p ∈ P , the map is injective by left cancellation, so that the mapping P → P, x 7→ px extends (uniquely) to an isometry δx 7→ δpx (x ∈ P ) The assignment Vp : ℓ2P → ℓ2P. represents our semigroup P as isometries on ℓ2P . This is called the left regular representation of P . It generates the following C*-algebra: p 7→ Vp (p ∈ P ) Definition 2.1. C∗λ(P ) := C∗({Vp : p ∈ P}) ⊆ L(ℓ2P ). 4 XIN LI By definition, C∗λ(P ) is the smallest subalgebra of L(ℓ2P ) containing {Vp : p ∈ P} which is invariant under forming adjoints and closed in the operator norm topology. We call C∗λ(P ) the semigroup C*-algebra of P , or more precisely, the left reduced semigroup C*-algebra of P . Note that left cancellation is a crucial assumption for our construction. In general, without left cancellation, the mapping δx 7→ δpx does not even extend to a bounded linear operator on ℓ2P . Moreover, we point out that we view our semigroups as discrete objects. Our construction, and some of the analysis, carries over to certain topological semigroups (see [RS15, Sun14]). Finally, C∗λ(P ) will be separable if P is countable. This helps to exclude pathological cases. Therefore, for convenience, we assume from now on that all our semigroups are countable, although this is not always necessary in our discussion. 3. Examples We have already pointed out the importance of examples. Therefore, it is ap- propriate to start with a list of examples of semigroups where we can apply our construction. All our examples are actually semigroups with an identity, so that they are all monoids. 3.1. The natural numbers. Our first example is given by P = N = {0, 1, 2, . . .}, the set of natural numbers including zero, viewed as an additive monoid. By con- struction, V1 is the unilateral shift. Since N is generated by 1 as a monoid, it is clear that C∗λ(N) is generated as a C*-algebra by the unilateral shift. This C*-algebra has been studied by Coburn (see [Cob67, Cob69]). It turns out that it is the universal C*-algebra generated by one isometry, i.e., C∗λ(N) ∼= C∗(v v∗v = 1), V1 7→ v. C∗λ(N) is also called the Toeplitz algebra. This name comes from the observation that C∗λ(N) can also be described as the C*-algebra of Toeplitz operators on the Hardy space, defined on the circle. This interpretation connects our semigroup C*-algebra C∗λ(N) with index theory and K-theory. 3.2. Positive cones in totally ordered groups. Motivated by connections to index theory and K-theory, several authors including Coburn and Douglas studied the following examples in [CD71, CDSS71, Dou72, DH71]: Let G be a subgroup of (R, +), and consider the additive monoid P = [0,∞) ∩ G. The case G = Z gives our previous example P = N. The case where G = Z[λ, λ−1] for some positive real number λ is discussed in [CPPR11, Li15]. These examples belong to the bigger class of positive cones in totally ordered groups. A left invariant total order on a group G is a relation ≤ on G such that SEMIGROUP C*-ALGEBRAS 5 • For all x, y ∈ G, we have x = y if and only if x ≤ y and y ≤ x. • For all x, y ∈ G, we always have x ≤ y or y ≤ x. • For all x, y, z ∈ G, x ≤ y and y ≤ z imply x ≤ z. • For all x, y, z ∈ G, x ≤ y implies zx ≤ zy. Given a left invariant total order ≤ on G, define P := {x ∈ G : e ≤ x}. Here e is the identity in G. P is called the positive cone in G. It is a monoid satisfying G = P ∪ P −1 and P ∩ P −1 = {e} . (1) Conversely, every submonoid P ⊆ G of a group G satisfying (1) gives rise to a left invariant total order ≤ by setting, for x, y ∈ G, x ≤ y if y ∈ xP . Here xP = {xp : p ∈ P} ⊆ G. In the examples mentioned above of subgroups of (R, +), we have canonical left invariant total orders given by restricting the canonical order on (R, +). The study of left invariant total orders on group is of great interest in group theory. For instance, the existence of a left invariant total order on a group G implies the Kaplansky conjecture for G. This conjecture says that for a torsion-free group G and a ring R, the group ring RG does not have zero-divisors if R does not have zero-divisors. We refer to [MR77, DNR14] for more details. While it is known that every torsion-free nilpotent group admits a left invariant total order, it is an open conjecture that lattices in simple Lie groups of rank at least two have no left invariant total order. It is also an open question whether an infinite property (T) group can admit a left invariant total order (see [DNR14] for more details). 3.3. Monoids given by presentations. Another source for examples of monoids comes from group presentations. One way to define a group is to give a presentation, i.e., generators and relations. For instance, the additive group of integers is the group generated by one element with no relation, Z = hai. The non-abelian free group on two generators is the group generated by two elements with no relations, F2 = Z ∗ Z = ha, bi. And Z × Z is the group generated by two elements which commute, Z2 = Z × Z = ha, b ab = bai. If we look at the semigroups (or rather monoids) defined by the same presentations, we get N = hai+, N∗2 = N∗N = ha, bi+, N2 = N×N = ha, b ab = bai+. Here, we write h· ·i+ for the universal monid given by a particular presentation, while we write h· ·i for the universal group given by a particular representation. This is to distinguish between group presentations and monoid presentations. Of course, in general, it is not clear whether this procedure of taking generators and relations from group presentations to define monoids leads to interesting semi- groups, or whether we can apply our C*-algebraic construction to the resulting semigroups. For instance, it could be that the monoid given by a presentation ac- tually coincides with the group given by the same presentation. Another problem that might arise is that the canonical homomorphism from the monoid to the group 6 XIN LI given by the same presentation, sending generator to generator, is not injective. In that case, our monoid might not even be left cancellative. However, there are condi- tions on our presentations which ensure that these problems do not appear. There is for instance the notion of completeness (see [Deh03]), explained in § 6.5. Now let us just give a list of examples. The presentations for Z, F2 and Z2 all have in common that two generators either commute or satisfy no relation (i.e., they are free), and these are the only relations we impose. This can be generalized. Let Γ = (V, E) be an undirected graph, where we connect two vertices by at most one edge and no vertex to itself. This means that we can think of E as a subset of V × V . We then define AΓ := h{σv : v ∈ V } σvσw = σwσv for all (v, w) ∈ Ei , Γ := h{σv : v ∈ V } σvσw = σwσv for all (v, w) ∈ Ei+ . A+ For instance, the graph for Z only consists of one vertex and no edge, the graph for F2 consists of two vertices and no edges, and the graph for Z2 consists of two vertices and one edge joining them. The groups AΓ are called right-angled Artin groups and the monoids A+ Γ are called right-angled Artin monoids. Their C*-algebras are discussed in [CL02, CL07, Iva10, ELR16]. Right-angled Artin monoids and the corresponding groups are special cases of graph products. Let Γ = (V, E) be a graph as above, with E ⊆ V × V . Assume that for every v ∈ V , Gv is a group containing a submonoid Pv. Then let Γv∈V Gv be the group obtained from the free product ∗v∈V Gv by introducing the relations xy = yx for all x ∈ Gv and y ∈ Gw with (v, w) ∈ E. Similarly, define Γv∈V Pv as the monoid obtained from the free product ∗v∈V Pv by introducing the relations xy = yx for all x ∈ Pv and y ∈ Pw with (v, w) ∈ E. It is explained in [CL02] (see also [Gre90, HM95]) that the embeddings Pv ֒→ Gv induce an embedding Γv∈V Pv ֒→ Γv∈V Gv. In the case that Pv ⊆ Gv is given by N ⊆ Z for all v ∈ V , we obtain right-angled Artin monoids and the corresponding groups. We will have more to say about general graph products in § 4.2 and § 9. As the name suggests, there is a more general class of Artin groups which contains right-angled Artin groups. Let I be a countable index set, {mij ∈ {2, 3, 4, . . .} ∪ {∞} : i, j ∈ I, i 6= j} SEMIGROUP C*-ALGEBRAS 7 For mij = ∞, there is no relation involving σi and σj, i.e., σi and σj are free. And define be such that mij = mji for all i and j. Then define mij G :=*{σi : i ∈ I} σiσjσiσj ··· {z } P :=*{σi : i ∈ I} σiσjσiσj ··· } {z mij mji {z mji {z } } = σjσiσjσi ··· for all i, j ∈ I, i 6= j+ . = σj σiσjσi ··· for all i, j ∈ I, i 6= j++ . If mij ∈ {2,∞} for all i and j, then we get right-angled Artin groups and monoids. To see some other groups, take for instance I = {1, 2} and m1,2 = m2,1 = 3. We get the (third) Braid group and the corresponding Braid monoid B3 := hσ1, σ2 σ1σ2σ1 = σ2σ1σ2i , 3 := hσ1, σ2 σ1σ2σ1 = σ2σ1σ2i+ . B+ In general, for n ≥ 1, the braid group Bn and the corresponding braid monoid B+ are given by n Bn :=(cid:28)σ1, . . . , σn−1 n :=(cid:28)σ1, . . . , σn−1 B+ σiσi+1σi = σi+1σiσi+1 for 1 ≤ i ≤ n − 2, σiσj = σjσi for i − j ≥ 2 σiσi+1σi = σi+1σiσi+1 for 1 ≤ i ≤ n − 2, σiσj = σj σi for i − j ≥ 2 (cid:29) , (cid:29)+ . This corresponds to the case where I = {1, . . . , n − 1} and mi,i+1 = mi+1,i = 3 for all 1 ≤ i ≤ n − 2 and mi,j = mj,i = 2 for all i − j ≥ 2. These Artin groups form an interesting class of examples which is of interest for group theorists. Another family of examples is given by Baumslag-Solitar groups and their presen- tations: For k, l ≥ 1, define the group and the monoid and the monoid Also, again for k, l ≥ 1, define the group . B+ Bk,l :=(cid:10)a, b abk = bla(cid:11) k,l :=(cid:10)a, b abk = bla(cid:11)+ B−k,l :=(cid:10)a, b a = blabk(cid:11) −k,l :=(cid:10)a, b a = blabk(cid:11)+ B+ . These are the Baumslag-Solitar groups and the Baumslag-Solitar monoids. The reader may find more about the semigroup C*-algebras attached to Baumslag- Solitar monoids in [Spi12, Spi14]. Finally, let us mention the Thompson group and the Thompson monoid. The Thompson group is given by F := hx0, x1, . . . xnxk = xkxn+1 for k < ni . 8 XIN LI This is just one possible presentation defining the Thompson group. There are other, for instance The first presentation however has the advantage that it leads naturally to the definition of the Thompson monoid as F =(cid:10)A, B [AB−1, A−1BA] = [AB−1, A−2BA2] = e(cid:11) . F + := hx0, x1, . . . xnxk = xkxn+1 for k < ni+ . The Thompson group is of great interest in group theory, in particular the question whether it is amenable or not is currently attracting a lot of attention. Therefore, it would be very interesting to study the Thompson monoid and its semigroup C*-algebra. 3.4. Examples from rings in general, and number theory in particular. Let us present another source for examples. This time, our semigroups come from rings. Let R be a ring without zero-divisors (x 6= 0 is a zero-divisor if there exists 0 6= y ∈ R with xy = 0). Then R× = R \ {0} is a cancellative semigroup with respect to multiplication. We can also construct the ax + b-semigroup R ⋊ R×. The underlying set is R× R×, and multiplication is given by (d, c)(b, a) = (d + cb, ca). It is a semidirect product for the canonical multiplicative action of R× on R. Another possibility would be to take an integral domain R, i.e., a commutative ring with unit not containing zero-divisors, and form the semigroup Mn(R)× of n × n- matrices over R with non-vanishing determinant. We could also form the semidirect product Mn(R) ⋊ Mn(R)× for the canonical multiplicative action of Mn(R)× on Mn(R). In particular, rings from number theory are interesting. Let K be a number field, i.e., a finite extension of Q. Then the ring of algebraic integers R in K is given by (cid:8)x ∈ K : There are n ≥ 1, an−1, . . . , a0 ∈ Z with xn + an−1xn−1 + . . . + a0 = 0(cid:9) . For instance, for the classical case K = Q, the ring of algebraic integers is given by the usual integers, R = Z. For the number field of Gaussian numbers, K = Q[i], the ring of algebraic integers are given by the Gaussian integers, R = Z[i]. More generally, for the number field K = Q[ζ] generated by a root of unity ζ, the ring of algebraic integers is given by R = Z[ζ]. For the real quadratic number field K = Q[√2], the ring of algebraic integers is given by Z[√2], while for the real quadratic number field K = Q[√5], the ring of algebraic integers is given by R = Z[ 1+√5 ]. 2 Let us briefly mention an interesting invariant of number fields. Let K be a number field with ring of algebraic integers R. We introduce an equivalence relation for non- zero ideals I of R by saying that a ∼ b if there exist a, b ∈ R× with ba = ab. It turns out that with respect to multiplication of ideals, I/∼ becomes a finite abelian group. This is the class group ClK of K. An outstanding open question in number theory is how to compute ClK, or even just the class number hK = #ClK, in a systematic and efficient way. It is not even known whether there are infinitely many SEMIGROUP C*-ALGEBRAS 9 (non-isomorphic) number fields with trivial class group (i.e., class number one). We refer the interested reader to [Neu99] for more details. It is possible to consider more general semidirect products, in the more flexible setting of semigroups acting by endomorphisms on a group. Particular cases are discussed in [Cun17a]. We also refer to [CV13, BLS14, BS16, Sta15a] and the ref- erences therein for more examples and for results on the corresponding semigroup C*-algebras. 3.5. Finitely generated abelian cancellative semigroups. Finally, one more class of examples which illustrates quite well that the world of semigroups can be much more complicated than the world of groups: Consider finitely generated abelian cancellative semigroups, or monoids. For groups, we have a well under- stood structure theorem for finitely generated abelian groups. But for semigroups, this class of examples is interesting and challenging to understand. For instance, particular examples are given by numerical semigroups, i.e., semigroups of the type P = N \ F , where F is a finite subset of N such that N\ F is additively closed. For instance, we could take F = {1} or F = {1, 3}. We refer the interested reader to [RGS09] and the references therein for more about numerical semigroups, and also to [Cun17b]. 4. Preliminaries 4.1. Embedding semigroups into groups. As we mentioned earlier, we need left cancellation for semigroups in our construction of semigroup C*-algebras. One way to ensure cancellation is to embed our semigroups into groups, i.e., to find an injective semigroup homomorphism from our semigroup into a group. In general, the question which semigroups embed into groups is quite complicated. Cancella- tion is necessary but not sufficient. Malcev gave the complete answer. He found an infinite list of conditions which are necessary and sufficient for group embeddability, and showed that any finite subset of his list is no longer sufficient. His list includes cancellation, which means both left cancellation and right cancellation. The latter means that for every p, x, y ∈ P , xp = yp implies x = y. But Malcev's list also consists of conditions like the following: For every a, b, c, d, u, v, x, y ∈ P , xa = yb, xc = yd, ua = vb implies uc = vd. We refer to [CP67, § 12] for more details. As explained in [CP67, § 12], if a semigroup P embeds into a group, then there is a universal group embedding P ֒→ Guniv, meaning that for every homomorphism P → G of the semigroup P to a group G, there is a unique homomorphism Guniv → 10 XIN LI G which makes the diagram P  commutative. Guniv P P P P P P P P P P P P P P 'P G Group embeddability is in general a complicated issue. Therefore, whenever it is convenient, we will simply assume that our semigroups can be embedded into groups. Verifying this assumption might be a challenge, for instance in the case of Artin monoids (compare [Par02]). However, we would like to mention one sufficient condition for group embeddability. Let P be a cancellative semigroup, i.e., P is left and right cancellative. Furthermore, assume that P is right reversible, i.e., for every p, q ∈ P , we have P p∩P q 6= ∅. Here P p = {xp : x ∈ P}. Then P embeds into a group. Actually, the universal group in the universal group embedding of P is given by an explicit construction as the group G of left quotients. This means that G consists of formal quotients of the form q−1p, for all q ∈ P and p ∈ P . We say that two such formal expressions q−1 p and q−1p represent the same element in G if there is r ∈ P with q = rq and p = rp. To multiply elements in G, we make use of right reversibility: Given p, q, r, s ∈ P , suppose we want to multiply s−1r with q−1p. As P q ∩ P r 6= ∅, there exist x and y in P with q = yr. Thus q−1p = (xq)−1(xp) = (yr)−1(xp). Let us now make the following formal computation: (s−1r)(q−1p) = (s−1r)(yr)−1(xp) = s−1rr−1y−1(xp) = s−1y−1(xp) = (ys)−1(xp). Motivated by this computation, we set (s−1r)(q−1p) := (ys)−1(xp). It is now straightforward to check that this indeed defines a group G = P −1P , and that P → G, p 7→ e−1p is an embedding of our semigroup P into our group G. Here e is the identity of P . We can always arrange that P has an identity by simply adjoining one if necessary. It is easy to see that this group embedding which we just constructed is actually the universal group embedding for P . We refer the reader to [CP61, § 1.10] for more details. Obviously, by symmetry, we also obtain that a cancellative semigroup P embeds into a group, if P is left reversible, i.e., if for every p, q ∈ P , we have pP ∩ qP 6= ∅. In that case, P embeds into its group G of right quotients, G = P P −1, and this is the universal group embedding for P . For instance, both of these necessary conditions for group embeddability are sat- isfied for cancellative abelian semigroups. They are also satisfied for the Braid monoids B+ n introduced above. The ax + b-semigroup R ⋊ R× over an integral domain R is right reversible, but if R is not a field, then R ⋊ R× is not left reversible.  / / '   SEMIGROUP C*-ALGEBRAS 11 The Thompson monoid is left reversible but not right reversible. Finally, the non-abelian free monoid N∗n is neither left nor right reversible. 4.2. Graph products. We collect some basic facts about graph products which we will use later on in § 9. Basically, we follow [CL02, § 2]. Let Γ = (V, E) be a graph with vertices V and edges E. Two vertices in V are connected by at most one edge, and no vertex is connected to itself. Hence we view E as a subset of V × V . For every v ∈ V , assume that we are given a submonoid Pv of a group Gv. We can then form the graph products P := Γv∈V Pv and G := Γv∈V Gv. As we explained, the group G is obtained from the free product ∗v∈V Gv by introducing the relations xy = yx for all x ∈ Gv and y ∈ Gw with (v, w) ∈ E. Similarly, P is defined as the monoid obtained from the free product ∗v∈V Pv by introducing the relations xy = yx for all x ∈ Pv and y ∈ Pw with (v, w) ∈ E. As explained in [CL02], it turns out that for every v, the monoid Pv sits in a canonical way as a submonoid inside the monoid Γv∈V Pv. Similarly, for each v ∈ V , the group Gv sits in a canonical way as a subgroup inside the group Γv∈V Gv. Moreover, the monoid P = Γv∈V Pv can be canonically embedded as a submonoid of the group G = Γv∈V Gv. A typical element g of G = Γv∈V Gv is a product of the form x1x2 ··· xl, where xi ∈ Gvi are all non-trivial. (To obtain the identity, we would have to allow the empty word, i.e., the case l = 0.) We distinguish between words like x1x2 ··· xl and the element g they represent in the graph product G by saying that x1x2 ··· xl is an expression for g. Let us now explain when two words are expressions for the same group element. First of all, for a word like x1x2 ··· xl, we call the xis the syllables and l the length of the word. We write v(xi) for the vertex vi ∈ V with the property that xi lies in Gvi . Given a word x1 ··· xixi+1 ··· xl with the property that (v(xi), v(xi+1)) ∈ E, we can replace the subword xixi+1 by xi+1xi. In this way, we transform the original word x1 ··· xixi+1 ··· xl to the new word This procedure is called a shuffle. Two words are called shuffle equivalent if one can be obtained from the other by performing finitely many shuffles. x1 ··· xi+1xi ··· xl. Moreover, given a word x1 ··· xixi+1 ··· xl with the property that v(xi) = v(xi+1), then we say that our word admits an amalgamation. In that case, we can replace the subword xixi+1 by the product xi · xi+1 ∈ Gvi , where vi = v(xi) = v(xi+1). Furthermore, if xi · xi+1 = e in Gvi , 12 XIN LI then we delete this part of our word. In this way, we transform the original word to the new word if xi · xi+1 6= e in Gvi and x1 ··· xixi+1 ··· xl x1 ··· (xi · xi+1)··· xl x1 ··· xi−1xi+2 ··· xl if xi · xi+1 = e in Gvi . This procedure is called an amalgamation. Finally, we say that a word is reduced if it is not shuffle equivalent to a word which admits an amalgamation. We have the following Lemma 4.1 (Lemma 1 in [CL02]). A word x1 ··· xl is reduced if and only if for all 1 ≤ i < j ≤ l with v(xi) = v(xj ), there exists 1 ≤ k ≤ l with i < k < j such that (v(xi), v(xk)) /∈ E. Suppose that we are given two words, and we can transform one word into the other by finitely many shuffles and amalgamations. Then it is clear that these two words are expressions for the same element in our group G. The converse is also true, this is the following result due to Green (see [Gre90]): Theorem 4.2 (Theorem 2 in [CL02]). Any two reduced words which are expressions for the same group element in G are shuffle equivalent. In other words, two words which are expressions for the same group element in G can be transformed into one another by finitely many shuffles and amalgamations. This is because, with the help of Lemma 4.1, it is easy to see that every word can be transformed into a reduced one by finitely many shuffles and amalgamations. Because of Theorem 4.2, we may introduce the notion of length: Definition 4.3. The length of an element g in our graph product G is the length of a reduced word which is an expression for g. We also introduce the following Definition 4.4. Suppose we are given a reduced word x = x1 ··· xl. Then we call xi an initial syllable and v(xi) an initial vertex of our word, if for every 1 ≤ h < i, (v(xh), v(xi)) ∈ E. The set of all initial vertices of x is denoted by V i(x) (in [CL02], the notation ∆(x) is used). SEMIGROUP C*-ALGEBRAS 13 Similarly, we call xj a final syllable and v(xj ) a final vertex of our word, if for every j < k ≤ l, (v(xj ), v(xk)) ∈ E. The set of all final vertices of x is denoted by V f (x) (it is denoted by ∆r(x) in [CL02]). The following is an easy observation: Lemma 4.5 (Lemma 3 in [CL02]). Let be a reduced word. x = x1 ··· xl If xi is an initial syllable of x, then x is shuffle equivalent to xix1 ··· xi−1xi+1 ··· xl. For all v, w ∈ V i(x), we have (v, w) ∈ E. For every v ∈ V i(x), there is a unique initial syllable xi of x with v(xi) = v. Let us denote this syllable by Si v(x). If x′ is shuffle equivalent to x, then V i(x) = V i(x′) and for every v ∈ V i(x) = V i(x′), Si v(x) = Si v(x′). The last three statements are also true for final vertices and final syllables. So we denote for a reduced word x with final syllable v the unique final syllable xj of x with v(xj ) = v by Sf v (x). Definition 4.6. Let g be an element in our graph product G, and let x be a reduced word which is an expression for g. Then we set V i(g) := V i(x), and for v ∈ V , Similarly, we define and for v ∈ V , We need the following v(g) :=(Si e Si v(x) if v ∈ V i(g) if v /∈ V i(g). V f (g) := V f (x), v (g) :=(Sf e Sf v (x) if v ∈ V f (g) if v /∈ V f (g). Lemma 4.7 (Lemma 5 in [CL02]). Given g and h in our graph product G, let and suppose that W := V f (g) ∩ V i(h), zw := Sf w(g)Si w(h) 6= e 14 XIN LI for all w ∈ W . Define in any order. z := Yw∈W zw, reduced expression for h, then x · z · y is a reduced expression for g · h. w(g) is a reduced expression for g and Qw∈W Si Then, if x ·Qw∈W Sf w(h) · y is a 4.3. Krull rings. Since we want to study ax + b-semigroups over integral domains and their semigroup C*-algebras later on, we collect a few basic facts in this context. Let R be an integral domain. Definition 4.8. The constructible (ring-theoretic) ideals of R are given by I(R) :=(c−1 n\i=1 aiR! : a1, . . . , an, c ∈ R×) . Here, for c ∈ R× and an ideal I of R, we set c−1I := {r ∈ R : cr ∈ I} . Now let Q be the quotient field of R. (2) Note that for c ∈ R× and X ⊆ R, we set I(R ⊆ Q) :=(cid:8)(x1 · R) ∩ . . . ∩ (xn · R) : xi ∈ Q×(cid:9) . c−1X = {r ∈ R : cr ∈ X} , but c−1 · X =(cid:8)c−1x : x ∈ X(cid:9) . Moreover, note that I(R) = {J ∩ R : J ∈ I(R ⊆ Q)}. By construction, the family I(R) consists of integral divisorial ideals of R, and I(R ⊆ Q) consists of divisorial ideals of R. By definition, a divisorial ideal of an integral domain R is a fractional ideal I that satisfies I = (R : (R : I)), where (R : J) = {q ∈ Q : qJ ⊆ R}. Equivalently, divisorial ideals are non-zero intersections of some non-empty family of principal fractional ideals (ideals of the form qR, q ∈ Q). Let D(R) be the set of divisorial ideals of R. In our situation, we only consider finite intersections of principal fractional ideals (see (2)). So in general, our family I(R ⊆ Q) will only be a proper subset of D(R). However, for certain rings, the set I(R ⊆ Q) coincides with D(R). For instance, this happens for noetherian rings. It also happens for Krull rings. The latter have a number of additional favourable properties which are very helpful for our purposes. Let us start with the following Definition 4.9. An integral domain R is called a Krull ring if there exists a family of discrete valuations (vi)i∈I of the quotient field Q of R such that (K1) R = {x ∈ Q : vi(x) ≥ 0 for all i ∈ I}, SEMIGROUP C*-ALGEBRAS 15 (K2) for every 0 6= x ∈ Q, there are only finitely many valuations in (vi)i such that vi(x) 6= 0. The following result gives us many examples of Krull rings. Theorem 4.10. [Bou06, Chapitre VII, § 1.3, Corollaire] A noetherian integral domain is a Krull ring if and only if it is integrally closed. Let us collect some basic properties of Krull rings: [Bou06, Chapitre VII, § 1.5, Corollaire 2] yields Lemma 4.11. For a Krull ring R, I(R ⊆ Q) = D(R) and I(R) is the set of integral divisorial ideals. Moreover, the prime ideals of height 1 play a distinguished role in a Krull ring. Theorem 4.12. [Bou06, Chapitre VII, § 1.6, Th´eor`eme 3 and Chapitre VII, § 1.7, Th´eor`eme 4] Let R be a Krull ring. Every prime ideal of height 1 of R is a divisorial ideal. Let P(R) = {p ⊳ R prime : ht(p) = 1} . For every p ∈ P(R), the localization Rp = (R \ p)−1R is a principal valuation ring. Let vp be the corresponding (discrete) valuations of the quotient field Q of R. Then the family (vp)p∈P(R) satisfies the conditions (K1) and (K2) from Definition 4.9. Proposition 4.13. [Bou06, Chapitre VII, § 1.5, Proposition 9] Let R be a Krull ring and (vp)p∈P(R) be the valuations from the previous theorem. Given finitely many integers n1, ..., nr and finitely many prime ideals p1, ..., pr in P(R), there exists x in the quotient field Q of R with vpi (x) = ni for all 1 ≤ i ≤ r and vp(x) ≥ 0 for all p ∈ P(R) \ {p1, . . . , pr} . Moreover, given a fractional ideal I of R, we let I∼ := (R : (R : I)) be the divisorial closure of I. I∼ is the smallest divisorial ideal of R which contains I. We can now define the product of two divisorial ideals I1 and I2 to be the divisorial closure of the (usual ideal-theoretic) product of I1 and I2, i.e., I1 • I2 := (I1 · I2)∼. D(R) becomes a commutative monoid with this multiplication. Theorem 4.14. [Bou06, Chapitre VII, § 1.2, Th´eor`eme 1; Chapitre VII, § 1.3, Th´eor`eme 2 and Chapitre VII, § 1.6, Th´eor`eme 3] For a Krull ring R, (D(R),•) is a group. It is the free abelian group with free generators given by P(R), the set of prime ideals of R which have height 1. 1 r This means that every I ∈ I(R ⊆ Q) (Q is the quotient field of the Krull ring R) • ··· • p(nr) is of the form I = p(n1) , with ni ∈ Z. Here for p ∈ P(R) and n ∈ N, we write , and p(−n) for p−1 • ··· • p−1 p(n) for p • ··· • p {z } } {z n times , n times 16 XIN LI where p−1 = (R : p). We set for p ∈ P(R): vp(I) :=(ni if p = pi, 0 if p /∈ {p1, . . . , pr} . With this notation, we have I = Qp∈P(R) p(vp(I)), where the product is taken in D(R). In addition, we have for I ∈ I(R ⊆ Q) that I ∈ I(R) if and only if vp(I) ≥ 0 for all p ∈ P(R). And combining the last statement in [Bou06, Chapitre VII, § 1.3, Th´eor`eme 2] with [Bou06, Chapitre VII, § 1.4, Proposition 5], we obtain for every I ∈ I(R ⊆ Q): (3) Finally, the principal fractional ideals F (R) form a subgroup of (D(R),•) which is isomorphic to Q×. Suppose that R is a Krull ring. Then the quotient group C(R) := D(R)/F (R) is called the divisor class group of R. I = {x ∈ Q : vp(x) ≥ vp(I) for all p ∈ P(R)} . These were basic properties of Krull rings. We refer the interested reader to [Bou06, Chapitre VII] or [Fos73] for more information. 5. C*-algebras attached to inverse semigroups, partial dynamical systems, and groupoids We refer the interested reader to [Ren80, Exe08, Exe15, Pat99] for more references for this section. 5.1. Inverse semigroups. Inverse semigroups play an important role in the study of semigroup C*-algebras. Definition 5.1. An inverse semigroup is a semigroup S with the property that for every x ∈ S, there is a unique y ∈ S with x = xyx and y = yxy. We write y = x−1 and call y the inverse of x. Definition 5.2. An inverse semigroup S is called an inverse semigroup with zero if there is a distinguished element 0 ∈ S satisfying 0 · s = 0 = s · 0 for all s ∈ S. Usually, if we write "inverse semigroup", we mean an inverse semigroup with or without zero. Sometimes we write "inverse semigroups without zero" for ordinary inverse semigroups which do not have a distinguished zero element. Every inverse semigroup can be realized as partial bijections on a fixed set. Multi- plication is given by composition. However, a partial bijection is only defined on its domain. Therefore, if we want to compose the partial bijection s : dom(s) → im (s) with another partial bijection t : dom(t) → im (t), we have to restrict t to dom(t) ∩ t−1(dom(s)) SEMIGROUP C*-ALGEBRAS 17 to make sure that the image of the restriction of t lies in the domain of s. Only then we can form s◦ t. The inverse of a partial bijection is the usual inverse, in the category of sets. Inverse semigroups can also be realized as partial isometries on a Hilbert space. To make sure that the product of two partial isometries is again a partial isometry, we have to require that the source and range projections of our partial isometries com- mute. Then multiplication in the inverse semigroup is just the usual multiplication of operators on a fixed Hilbert space, i.e., composition of operators. The inverse in our inverse semigroup is given by the adjoint operation for operators in general or partial isometries in our particular situation. Let us explain how to attach an inverse semigroup to a left cancellative semigroup. Assume that P is a left cancellative semigroup. Its left inverse hull Il(P ) is the inverse semigroup generated by the partial bijections P → pP, x 7→ px, whose domain is P and whose image is pP = {px : x ∈ P}. Its inverse is given by pP → P, px 7→ x. So Il(P ) is the smallest semigroup of partial bijections on P which is closed under inverses and contains Given p ∈ P , we denote the partial bijection {P → P, x 7→ px : p ∈ P} . P → pP, x 7→ px by p. In this way, we obtain an embedding of P into Il(P ) by sending p ∈ P to the partial bijection p ∈ Il(P ). This allows us to view P as a subsemigroup of Il(P ). We say that Il(P ) is an inverse semigroup with zero if the partial bijection which is nowhere defined, ∅ → ∅, is in Il(P ). In that case, ∅ → ∅ is the distinguished zero element 0. Alternatively, we can also describe Il(P ) as the smallest inverse semigroup of partial isometries on ℓ2P generated by the isometries {Vp : p ∈ P}. This means that Il(P ) can be identified with the smallest semigroup of partial isometries on ℓ2P containing the isometries {Vp : p ∈ P} and their adjoints (cid:8)V ∗p : p ∈ P(cid:9) and which is closed under multiplication. In this picture, Il(P ) is an inverse semigroup with zero if and only if the zero operator is in Il(P ). An important subsemigroup of an inverse semigroup S is its semilattice of idempo- tents. Definition 5.3. The semilattice E of idempotents in an inverse semigroup S is given by E :=(cid:8)x−1x : x ∈ S(cid:9) =(cid:8)xx−1 : x ∈ S(cid:9) =(cid:8)e ∈ S : e = e2(cid:9) . Define an order on E by setting, for e, f ∈ E, e ≤ f if e = ef . 18 XIN LI If S is an inverse semigroup with zero, E becomes a semilattice with zero, and the distinguished zero element of S becomes the distinguished zero element of E. In the case of partial bijections, the semilattice of idempotents is given by all domains and images. Multiplication in this semilattice is intersection of sets, and ≤ is ⊆ for sets, i.e., containment. Definition 5.4. For the left inverse hull Il(P ) attached to a left cancellative semi- group P , the semilattice of idempotents is denoted by JP . It is easy to see that JP is given by JP =(cid:8)pn ··· q−1 1 p1(P ) : qi, pi ∈ P(cid:9) ∪(cid:8)q−1 Here, for X ⊆ P and p, q ∈ P , we write p(X) = {px : x ∈ X} n pn ··· q−1 1 p1(P ) : qi, pi ∈ P(cid:9) . and q−1(X) = {y ∈ P : qy ∈ X} . n pn ··· q−1 1 p1(P ) or q−1 Subsets of the form pn ··· q−1 1 p1(P ) are right ideals of P . Here, we call X ⊆ P a right ideal if for every x ∈ X and r ∈ P , we always have xr ∈ X. Definition 5.5. The elements in JP are called constructible right ideals of P . We will work out the set of constructible right ideals explicitly for classes of exam- ples in § 6.5. There is a duality between semilattices, i.e., abelian semigroups of idempotents, and totally disconnected locally compact Hausdorff spaces. Given a semilattice E, bE = {χ : E → {0, 1} non-zero semigroup homomorphism} . latter set is equipped with the usual multiplication when we view it as a subspace of R (or C). In addition, we require that these multiplicative maps must take the value 1 for some element e ∈ E. If our semilattice E is a semilattice with zero, and we construct its space of characters bE as follows: In other words, elements in bE are multiplicative maps from E to {0, 1}, where the 0 is its distinguished zero element, then we require that χ(0) = 0 for all χ ∈ bE. The topology on bE is given by pointwise convergence. Every χ ∈ bE is uniquely determined by χ−1(1) = {e ∈ E : χ(e) = 1} . χ−1(1) is an E-valued filter (which we simply call filter from now on), i.e., a subset of E satisfying: • χ−1(1) 6= ∅. • For all e, f ∈ E with e ≤ f , e ∈ χ−1(1) implies f ∈ χ−1(1). SEMIGROUP C*-ALGEBRAS 19 • For all e, f ∈ E with e, f ∈ χ−1(1), ef lies in χ−1(1). never an element of a filter. Conversely, every filter, i.e., every subset F ∈ E satisfying these three conditions If E is a semilattice with zero, and 0 is the distinguished zero element, then we determines a unique χ ∈ bE with χ−1(1) = F . Therefore, we have a one-to-one correspondence between characters χ ∈ bE and filters. have χ(0) = 0 for all χ ∈ bE. In terms of filters, this amounts to saying that 0 is structible right ideals JP and the space of characters cJP for the non-abelian free potents E and the space of its characters bE for the inverse semigroup S = Il(N∗N). semigroup on two generators P = N∗ N, or in other words, the semilattice of idem- As an illustrative example, the reader is encouraged to work out the set of con- Now assume that we are given a subsemigroup P of a group G. We define Il(P )× := Il(P ) \ {0} if Il(P ) is an inverse semigroup with zero, and 0 is its distinguished zero element, and Il(P )× := Il(P ) otherwise. Now it is easy to see that for every partial bijection s in Il(P )×, there exists a unique σ(s) ∈ G such that s is of the form Here we view P as a subset of the group G and make use of multiplication in G. s(x) = σ(s) · x for x ∈ dom(s). In the alternative picture of Il(P ) as the inverse semigroup of partial isometries on ℓ2P generated by the isometries {Vp : p ∈ P}, Il(P )× is given by all non-zero partial isometries in Il(P ). Every element in Il(P )× is of the form Vq1 ··· V ∗pn , V ∗p1 Vq1 ··· V ∗pn , Vq1 ··· V ∗pn Vqn , or V ∗p1 Vq1 ··· V ∗pn Vqn . The map σ which we introduced above is then given by σ(Vq1 ··· V ∗pn ) = q1 ··· p−1 σ(V ∗p1 Vq1 ··· V ∗pn ) = p−1 σ(Vq1 ··· V ∗pn Vqn ) = q1 ··· p−1 n ∈ G, n qn ∈ G, n ∈ G, 1 q1 ··· p−1 or σ(V ∗p1 Vq1 ··· V ∗pn Vqn ) = p−1 1 q1 ··· p−1 n qn ∈ G. To see that σ is well-defined, note that, similarly as above, every partial isometry V ∈ Il(P )× has the property that there exists a unique g ∈ G such that for every x ∈ P , either V δx = 0 or V δx = δg·x. And σ is defined in such a way that σ(V ) = g. 20 XIN LI It is easy to see that the map σ : Il(P )× → G satisfies σ(st) = σ(s)σ(t) for all s, t ∈ Il(P )×, as long as the product st lies in Il(P )×, i.e., is non-zero. Moreover, setting if JP is a semilattice with zero, and 0 is the distinguished zero element, and J ×P := JP \ {0} J ×P := JP otherwise, it is also easy to see that Here e is the identity in our group G. σ−1(e) = J ×P . We formalize this in the next definition: Let S be an inverse semigroup and E the semilattice of idempotents of S. We set S× := S \ {0} if S is an inverse semigroup with zero, and 0 is the distinguished zero element, and S× := S otherwise. Similarly, let E× := E \ {0} if E is a semilattice with zero, and 0 is the distinguished zero element, and E× := E otherwise. Moreover, let G be a group. Definition 5.6. A map σ : S× → G is called a partial homomorphism if σ(st) = σ(s)σ(t) for all s, t ∈ S× with st ∈ S×. A map σ : S× → G is called idempotent pure if σ−1(e) = E×. The existence of an idempotent pure partial homomorphism will allow us to describe C*-algebras attached to inverse semigroups as crossed products of partial dynamical systems later on. The following is a useful observation which we need later on. Lemma 5.7. Assume that S is an inverse semigroup and σ : S× → G is an idempotent pure partial homomorphism to a group G. Whenever two elements s and t in S× satisfy s−1s = t−1t and σ(s) = σ(t), then we must have s = t. Proof. It is clear that st−1 lies in S×. Since σ(st−1) = e, we must have st−1 ∈ E. Hence st−1 = ts−1st−1 = tt−1, and therefore s = ss−1s = st−1t = tt−1t = t. (cid:3) Let us now explain the construction of reduced and full C*-algebras for inverse semigroups. Let S be an inverse semigroup, and define S× as above. For s ∈ S, define λs : ℓ2S× → ℓ2S× SEMIGROUP C*-ALGEBRAS 21 by setting λs(δx) := δsx if s−1s ≥ xx−1, and λs(δx) := 0 otherwise. Note that we require s−1s ≥ xx−1 because on (cid:8)x ∈ S : s−1s ≥ xx−1(cid:9) , the map x 7→ sx given by left multiplication with s is injective. This is because we can reconstruct x from sx due to the computation x = xx−1x = s−1sxx−1x = s−1(sx). Therefore, for each s, we obtain a partial isometry λs by our construction. The assignment s 7→ λs is a *-representation of S by partial isometries on ℓ2S×. It is called the left regular representation of S. The star in *-representation indicates that we have λs−1 = λ∗s. Definition 5.8. We define C∗λ(S) := C∗({λs : s ∈ S}) ⊆ L(ℓ2S×). C∗λ(S) is called the reduced inverse semigroup C*-algebra of S. The full C*-algebra of an inverse semigroup S is given by a universal property. Definition 5.9. We define C∗(S) := C∗(cid:0){vs}s∈S vsvt = vst, v∗s = vs−1 , v0 = 0 if 0 ∈ S(cid:1) . C∗(S) is the full inverse semigroup C*-algebra of S. Here, 0 ∈ S is short for "S is an inverse semigroup with zero, and 0 is the distinguished zero element". This means that C∗(S) is uniquely determined by the property that given any C*-algebra B with elements {ws : s ∈ S} satisfying the above relations, i.e., then there exists a unique *-homomorphism from C∗(S) to B sending vs to ws. wswt = wst, w∗s = ws−1 , w0 = 0 if 0 ∈ S, In other words, C∗(S) is the C*-algebra universal for *-representation of S by partial isometries (in a C*-algebra, or on a Hilbert space). Note that we require that if 0 ∈ S, then the zero element of S should be represented by the partial isometry 0. That is why v0 = 0 in case 0 ∈ S. This is different from the definition in [Pat99, § 2.1], where the partial isometry representing 0 in the full C*-algebra of S is a non-zero, minimal and central projection. We will come back to this difference in the definitions later on. By construction, there is a canonical *-homomorphism λ : C∗(S) → C∗λ(S), vs 7→ λs. It is called the left regular representation (of C∗(S)). We refer the reader to [Pat99] for more about inverse semigroups and their C*- algebras. 22 XIN LI 5.2. Partial dynamical systems. Whenever we have a semigroup embedded into a group, or an inverse semigroup with an idempotent pure partial homomorphism to a group, we can construct a partial dynamical system. Let us first present the general framework. In the following, our convention will be that all our groups are discrete and count- able, and all our topological spaces are locally compact, Hausdorff and second countable. Definition 5.10. Let G be a group with identity e, and let X be a topological space. A partial action α of G on X consists of • a collection {Ug}g∈G of open subspaces Ug ⊆ X, • a collection {αg}g∈G of homeomorphisms αg : Ug−1 → Ug, x 7→ g.x such that – Ue = X, αe = idX ; – for all g1, g2 ∈ G, we have g2.(U(g1g2)−1 ∩ Ug−1 2 ) = Ug2 ∩ Ug−1 1 , and (g1g2).x = g1.(g2.x) for all x ∈ U(g1g2)−1 ∩ Ug−1 2 . We call such a triple (X, G, α) a partial dynamical system, and denote it by α : G y X or simply G y X. Let α : G y X be a partial dynamical system. The dual action α∗ of α is the partial action (in the sense of [McC95]) of G on C0(X) given by α∗g : C0(Ug−1 ) → C0(Ug), f 7→ f (g−1.⊔). We set out to describe a canonical partial action attached to a semigroup P em- bedded into a group G. Let C∗λ(P ) be the reduced semigroup C*-algebra of P . It contains a canonical commutative subalgebra Dλ(P ), which is given by Dλ(P ) := C∗({1X : X ∈ JP}) ⊆ C∗λ(P ). It is clear that Dλ(P ) coincides with span(σ−1(e)). Recall that the map σ : Il(P )× → G is given as follows: Every partial isometry V ∈ Il(P )× has the prop- erty that there exists a unique g ∈ G such that for every x ∈ P , either V δx = 0 or V δx = δg·x. And σ is defined in such a way that σ(V ) = g. Let us now describe the canonical partial action G y Dλ(P ). We will think of it as a dual action α∗. For g ∈ G, let Dg−1 := span((cid:8)V ∗V : V ∈ Il(P )×, σ(V ) = g(cid:9)). By construction, we have that De = Dλ(P ). Moreover, it is easy to see that Dg−1 is an ideal of Dλ(P ). Here is the argument: Suppose we are given V ∈ Il(P )× with σ(V ) = g, and W ∈ Il(P )× with σ(W ) = e. Then W must be a projection since for every x ∈ P , either W δx = 0 or W δx = δe·x = δx. Moreover, W and V ∗V commute SEMIGROUP C*-ALGEBRAS 23 as both of these are elements in the commutative C*-algebra ℓ∞(P ). Hence W V ∗V is non-zero if and only if V ∗V W is non-zero, and if that is the case, we obtain W V ∗V = V ∗V W = W V ∗V W = (V W )∗(V W ). As σ(V W ) = g, this implies that both W V ∗V and V ∗V W lie in Dg−1 . Therefore, as we claim, Dg−1 is an ideal of Dλ(P ). We then define α∗g as α∗g : Dg−1 → Dg, V ∗V → V V ∗ for V ∈ I×V with σ(V ) = g. This is well-defined: If we view ℓ2P as a subspace ℓ2G and let λ be the left regular representation of G, then every V ∈ I×V with σ(V ) = g satisfies V = λgV ∗V . Therefore, V V ∗ = λgV ∗V λ∗g. This shows that α∗g is just conjugation with the unitary λg. This also explains why α∗g is an isomorphism. Of course, we can also describe the dual action α. Set and for every g ∈ G, let It is easy to see that ΩP := Spec (Dλ(P )) Ug−1 := [Dg−1 . Ug−1 =(cid:8)χ ∈ ΩP : χ(V ∗V ) = 1 for some V ∈ I×V with σ(V ) = g(cid:9) . We then define αg by setting αg(χ) := χ ◦ α∗g−1 . These αg, g ∈ G, give rise to the canonical partial dynamical system G y ΩP attached to a semigroup P embedded into a group G. Our next goal is to describe a canonical partial dynamical system attached to inverse semigroups equipped with a idempotent pure partial homomorphism to a group. Let S be an inverse semigroup and E the semilattice of idempotents of S. Let G be a group. Assume that σ is a partial homomorphism S× → G which is idempotent pure. later (see Corollary 5.23) that the reduced C*-algebra C∗λ(S) of S is canonically In this situation, we describe a partial dynamical system G y bE, and we will show isomorphic to C0(bE) ⋊r G. Consider the sub-C*-algebra C∗(E) := C∗({λe : e ∈ E}) ⊆ C∗λ(S). As we will see, we have a canonical isomorphism Spec (C∗(E)) ∼= bE, so that C0(bE) ∼= C∗(E). Now let us describe the partial action G y C∗(E). For g ∈ G, define a sub-C*- algebra of C∗(E) by C∗(E)g−1 := span((cid:8)λs−1s : s ∈ S×, σ(s) = g(cid:9)). 24 XIN LI As σ is idempotent pure, we have C∗(E)e = C∗(E). For every g ∈ G, we have a C*-isomorphism α∗g : C∗(E)g−1 → C∗(E)g, λs−1s 7→ λss−1 . The corresponding dual action is given as follows: We identify Spec (C∗(E)) with bE. Then, for every g ∈ G, we set It is easy to see that Ug = Spec (C∗(E)g) ⊆ bE. Ug−1 =nχ ∈ bE : χ(s−1s) = 1 for some s ∈ S× with σ(s) = go . For every g ∈ G, the homeomorphism αg : Ug−1 → Ug defining the partial dynam- ical system G y bE is given by αg(χ) = χ ◦ α∗g−1 . More concretely, given χ ∈ Ug−1 and s ∈ S× with σ(s) = g and χ(s−1s) = 1, we have αg(χ)(e) = χ(s−1es). These αg, g ∈ G, give rise to the canonical partial dynamical system G y bE attached to an inverse semigroup S equipped with an idempotent pure partial homomorphism to a group G. At this point, a natural question arises. Assume we are given a semigroup P embedded into a group G. We have seen above that this leads to an idempotent pure partial homomorphism on the left inverse hull Il(P ) to our group G. How is the partial dynamical system G y ΩP related to the partial dynamical system G y cJP ? We will see the answer in § 6.7. Let us now recall the construction, originally defined in [McC95], of the reduced and full crossed products C0(X) ⋊α∗,r G and C0(X) ⋊α∗ G attached to our partial dynamical system α : G y X. We usually omit α∗ in our notation for the crossed products for the sake of brevity. First of all, and involution becomes a *-algebra under component-wise addition, multiplication given by C0(X) ⋊ℓ1 Xg G :=(Xg fgδg ∈ ℓ1(G, C0(X)) : fg ∈ C0(Ug)) fgδg! · Xh fhδh! :=Xg,h Xg fgδg!∗ :=Xg α∗g(α∗g−1 (fg) fh)δgh α∗g(f∗g−1 )δg. As in [McC95], we construct a representation of C0(X) ⋊ℓ1 discrete set, we define ℓ2X and the representation G. Viewing X as a M : C0(X) → L(ℓ2X), f 7→ M (f ), SEMIGROUP C*-ALGEBRAS 25 where M (f ) is the multiplication operator M (f )(ξ) := f · ξ for ξ ∈ ℓ2X. M is obviously a faithful representation of C0(X). Every g ∈ G leads to a twist of M , namely Mg : C0(X) → L(ℓ2X) given by Mg(f )ξ := fUg (g.⊔) · ξUg−1 . Here we view fUg (g.⊔) as an element in Cb(Ug−1 ), and Cb(Ug−1) acts on ℓ2Ug−1 just by multiplication operators. Given ξ ∈ ℓ2X, we set ξUg−1 (x) := ξ(x) if x ∈ Ug−1 and ξUg−1 (x) := 0 if x /∈ Ug−1 . In other words, ξUg−1 is the component of ξ in ℓ2Ug−1 with respect to the decom- position So we have ℓ2X = ℓ2Ug−1 ⊕ ℓ2U c g−1 . Mg(f )ξ(x) = f (g.x)ξ(x) if x ∈ Ug−1 and Mg(f )ξ(x) = 0 if x /∈ Ug−1 . Consider now the Hilbert space and define the representation H := ℓ2(G, ℓ2X) ∼= ℓ2G ⊗ ℓ2X, µ : C0(X) → L(H) given by µ(f )(δg ⊗ ξ) := δg ⊗ Mg(f )ξ. For g ∈ G, let Eg be the orthogonal projection onto µ(C0(Ug−1 ))H. Moreover, let λ denote the left regular representation of G on ℓ2G, and set Vg := (λg ⊗ I) · Eg. Here I is the identity operator on H. We can now define the representation µ × λ : C0(X) ⋊ℓ1 G → L(H),Xg fgδg 7→Xg µ(fg)Vg. Following the original definition in [McC95], we set Definition 5.11. C0(X) ⋊r G := C0(X) ⋊ℓ1 Gk·kµ×λ . To define the full crossed product C0(X) ⋊ G attached to our partial dynamical system G y X, recall that we have already introduced the *-algebra C0(X) ⋊ℓ1 G. Definition 5.12. Let C0(X) ⋊ G be the universal enveloping C*-algebra of the *-algebra C0(X) ⋊ℓ1 G. This means that C0(X) ⋊ G is universal for *-representations of C0(X) ⋊ℓ1 G as bounded operators on Hilbert spaces or to C*-algebras. To construct this universal C*-algebra, we follow the usual procedure of completing C0(X)⋊ℓ1 G with respect to the maximal C*-norm on C0(X)⋊ℓ1 G. Usually, we only obtain a C*-seminorm and have to divide out vectors with trivial seminorm, but because the *-representation µ × λ constructed above is faithful, we get a C*-norm. So there is an embedding C0(X) ⋊ℓ1 G ֒→ C0(X) ⋊ G, and the universal property of C0(X) ⋊ G means that 26 XIN LI whenever we have a *-homomorphism C0(X) ⋊ℓ1 there is a unique *-homomorphism C0(X) ⋊ G → B which makes the diagram G → B to some C*-algebra B, C0(X) ⋊ℓ1 G  C0(X) ⋊ G ❚ ❚ ❚ ❚ ❚ ❚ ❚ ❚ ❚ ❚ ❚ ❚ ❚ ❚ ❚ ❚ ❚ *❚ B commutative. By construction, there is a canonical *-homomorphism C0(X) ⋊ G → C0(X) ⋊r G extending the identity on C0(X) ⋊ℓ1 G. The reader may consult [McC95, Exe15] for more information about partial dynam- ical systems and their C*-algebras. 5.3. ´Etale groupoids. Groupoids play an important role in operator algebras in general and for our topic of semigroup C*-algebras in particular. This is because many C*-algebras can be written as groupoid C*-algebras. This also applies to many semigroup C*-algebras. Let us first introduce groupoids. In the language of categories, a groupoid is simply a small category with inverses. Very roughly speaking, this means that a groupoid is a group where multiplication is not globally defined. Roughly speaking, a groupoid G is a set, whose elements γ are arrows r(γ) ←− s(γ). Here r(γ) and s(γ) are elements in G(0), the set of units. r stands for range and s stands for source. For idu←− u in our groupoid G. This every u ∈ G(0), there is a distinguished arrow u allows us to define an embedding G(0) ֒→ G, u 7→ idu, which in turn allows us to view G(0) as a subset of G. G comes with a multiplication {(γ, η) ∈ G × G : s(γ) = r(η)} −→ G, (γ, η) 7→ γη. We think of this multiplication as concatenation of arrows. With this picture in mind, the condition s(γ) = r(η) makes sense. Also, G comes with an inversion G → G, γ → γ−1. We think of this inversion as reversing arrows. The picture of arrows, with concate- nation as multiplication and reversing as inversion, leads to obvious axioms, which, once imposed, give rise to the formal definition of a groupoid. Let us present the details.  / / *   SEMIGROUP C*-ALGEBRAS 27 Definition 5.13. A groupoid is a set G, together with a bijective map G → G, γ 7→ γ−1, a subset G ∗ G ⊆ G × G, and a map G ∗ G → G, (γ, η) 7→ γη, such that (γ−1)−1 = γ for all γ ∈ G, (γη)ζ = γ(ηζ) for all (γ, η), (η, ζ) ∈ G ∗ G, γ−1γη = η, γηη−1 = γ for all (γ, η) ∈ G ∗ G. Note that we implicitly impose conditions on G ∗ G so that these equations make sense. For instance, the second equation implicitly requires that for all (γ, η) and (η, ζ) in G ∗ G, ((γη), ζ) and (γ, (ηζ)) must lie in G ∗ G as well. Elements in G ∗ G are called composable pairs. The set of units is now defined by it is also given by Moreover, we define the source map by setting G(0) :=(cid:8)γ−1γ : γ ∈ G(cid:9) , G(0) =(cid:8)γγ−1 : γ ∈ G(cid:9) . s : G → G(0), γ 7→ γ−1γ and the range map by setting r : G → G(0), γ 7→ γγ−1. It is now an immediate consequence of the axioms that G ∗ G = {(γ, η) ∈ G × G : s(γ) = r(η)} . A groupoid G is called a topological groupoid if the set G comes with a topology such that multiplication and inversion become continuous maps. A topological groupoid is called ´etale if r and s are local homeomorphisms. A topological groupoid is called locally compact if it is locally compact (and Hausdorff) as a topological space. As an example, let us describe the partial transformation groupoid attached to the partial dynamical system α : G y X. It is denoted by G α⋉ X and is given by with source map s(g, x) = x, range map r(g, x) = g.x, composition G α⋉ X :=(cid:8)(g, x) ∈ G × X : g ∈ G, x ∈ Ug−1(cid:9) , and inverse (g1, g2.x)(g2, x) = (g1g2, x) (g, x)−1 = (g−1, g.x). We equip G α⋉ X with the subspace topology from G× X. Usually, we write G⋉ X for G α⋉ X if the action α is understood. The unit space of G ⋉ X coincides with X. Since G is discrete, G ⋉ X is an ´etale groupoid. Actually, if we set Gx :=(cid:8)g ∈ G : x ∈ Ug−1(cid:9) and Gx := {g ∈ G : x ∈ Ug} 28 XIN LI for x ∈ X, then we have canonical identifications s−1(x) ∼= Gx, (g, x) 7→ g and r−1(x) ∼= Gx, (g, g−1.x) 7→ g. Let G be an ´etale locally compact groupoid. For x ∈ G(0), let Gx = s−1(x) and Gx = r−1(x). Cc(G) is a *-algebra with respect to the multiplication (f ∗ g)(γ) = Xβ∈Gs(γ) f (γβ−1)g(β) and the involution f∗(γ) = f (γ−1). For every x ∈ G(0), define a *-representation πx of Cc(G) on ℓ2Gx by setting πx(f )(ξ)(γ) = (f ∗ ξ)(γ) = Xβ∈Gx f (γβ−1)ξ(β). Alternatively, if we want to highlight why these representations play the role of the left regular representation, attached to left multiplication, we could define πx by setting πx(f )δγ = Xα∈Gr(γ) f (α)δαγ. Here {δγ : γ ∈ Gx} is the canonical orthonormal basis of ℓ2Gx. With these definitions, we are ready to define groupoid C*-algebras. Definition 5.14. Let kfkC ∗ r (G) := sup x∈G(0) kπx(f )k for f ∈ Cc(G). We define C∗r (G) := Cc(G)k·kC∗ C∗r (G) is called the reduced groupoid C*-algebra of G. r (G) . Alternatively, we could set and π = Mx∈G(0) πx C∗r (G) = π(Cc(G)) ⊆ L(Mx ℓ2Gx). Let us now define the full groupoid C*-algebra. Let G be an ´etale locally compact groupoid. Then G(0) is a clopen subspace of G. Therefore, we can think of Cc(G(0)) as a subspace of Cc(G) simply by extending functions on G(0) by 0 to functions on G. This allows us to define the full groupoid C*-algebra. SEMIGROUP C*-ALGEBRAS 29 Definition 5.15. For f ∈ Cc(G), let kfkC ∗(G) = sup π kπ(f )k , ). where the supremum is taken over all *-representations of Cc(G) which are bounded on Cc(G(0)) (with respect to the supremum norm k·k∞ We then set C∗(G) := Cc(G)k·kC∗ (G) . C∗(G) is called the full groupoid C*-algebra of G. Remark 5.16. We will only deal with second countable locally compact ´etale groupoids. In that case, [Ren80, Chapter II, Theorem 1.21] tells us that every *- representation of Cc(G) on a separable Hilbert space is automatically bounded. In other words, the full groupoid C*-algebra of G is the universal enveloping C*-algebra of Cc(G). This notion has been explained after Definition 5.12. By construction, there is a canonical *-homomorphism C∗(G) → C∗r (G) extending the identity on Cc(G). It is called the left regular representation. 5.4. The universal groupoid of an inverse semigroup. We attach groupoids to inverse semigroups so that full and reduced C*-algebras coincide. The groupoids we construct are basically Paterson's universal groupoid, as in [Pat99, § 4.3] or [MS14]. There is however a small difference. In case of inverse semigroups with zero, our construction differs from Paterson's because we want the distinguished zero element to be represented by zero in the reduced and full C*-algebras. Let us first explain our construction. We start with an inverse semigroup S with semilattice of idempotents denoted by E. Set Σ :=n(s, χ) ∈ S × bE : χ(s−1s) = 1o . Note that in case 0 ∈ S, we must have s 6= 0 since χ(0) = 0 by our convention. We introduce an equivalence relation on Σ. Given (s, χ) and (t, ψ) in Σ, we define (s, χ) ∼ (t, ψ) if there exists e ∈ E with se = te and χ(e) = 1. The equivalence class of (s, χ) ∈ Σ with respect to ∼ is denoted by [s, χ]. We set G(S) := Σ/∼, i.e., G(S) = {[s, χ] : (s, χ) ∈ Σ} . To define a multiplication on G(S), we need to introduce the following notation: Let s ∈ S and χ ∈ bE be such that χ(s−1s) = 1. Then we define a new element s.χ of bE by setting Then we say that [t, ψ] and [s, χ] are composable if ψ = s.χ. In that case, we define their product as (s.χ)(e) := χ(s−1es). [t, ψ][s, χ] := [ts, χ]. we define U ⊆nχ ∈ bE : χ(s−1s) = 1o , D(s, U ) := {[s, χ] : χ ∈ U} . 30 XIN LI The inverse map is given by [s, χ]−1 := [s−1, s.χ]. It is easy to see that multiplication and inverse are well-defined, and they give rise to a groupoid structure on G(S). Moreover, we introduce a topology on G(S) by choosing a basis of open subsets. Given s ∈ S and an open subspace We equip G(S) with the topology which has as a basis of open subsets D(s, U ), for s ∈ S and U ⊆nχ ∈ bE : χ(s−1s) = 1o open. It is easy to check that with this topology, G(S) becomes a locally compact ´etale In all our examples, S will be countable, in which case G(S) will be groupoid. second countable. Let us explain the difference between our groupoid G(S) and the universal groupoid attached to S in [Pat99, § 4.3]. Assume that S is an inverse semigroup with zero, the space of semi-characters X introduced in [Pat99, § 2.1] and [Pat99, § 4.3] do not coincide. They are related by and 0 is the distinguished zero element. The starting point is that our space bE and Here χ0 is the semi-character on E which sends every element of E to 1, even 0. The disjoint union above is not only a disjoint union of sets, but also of topological spaces, i.e., χ0 is an isolated point in X (it is open and closed). X = bE ⊔ {χ0} . Now it is easy to see that our G(S) is the restriction of the universal groupoid Gu attached to S in [Pat99, § 4.3] to bE. This means that Actually, the only element in Gu which does not have range and source in bE is χ0 G(S) =nγ ∈ Gu : r(γ) ∈ bE, s(γ) ∈ bEo . itself. It follows that (4) Gu = G(S) ⊔ {χ0} . 5.5. Inverse semigroup C*-algebras as groupoid C*-algebras. We begin by identifying the full C*-algebras. Given an inverse semigroup S with semilattice of idempotents E, let us introduce the notation that for e ∈ E, we write Ue :=nχ ∈ bE : χ(e) = 1o . SEMIGROUP C*-ALGEBRAS 31 Theorem 5.17. For every inverse semigroup S, there is a canonical isomorphism C∗(S) ∼=−→ C∗(G(S)) sending the generator vs ∈ C∗(S) to the characteristic function on D(s, Us−1s), viewed as an element in Cc(G) ⊆ C∗(G). Recall that D(s, Us−1s) = {[s, χ] : χ ∈ Us−1s} . Proof. If case of inverse semigroups without zero, our theorem is just [Pat99, Chap- ter 4, Theorem 4.4.1]. Now let us assume that 0 ∈ S. Then the full C*-algebra attached to S in [Pat99, § 2.1] is canonically isomorphic to C∗(S) ⊕ Cv0, where C∗(S) is our full inverse semigroup C*-algebra in the sense of Definition 5.9, and v0 is a (non-zero) projection. For the full groupoid C*-algebra of the universal groupoid Gu attached to S in [Pat99, § 4.3], we get because of (4): C∗(Gu) ∼= C∗(G(S)) ⊕ C1χ0. Here 1χ0 is the characteristic function of the one-point set {χ0}, and it is easy to see that 1χ0 is a (non-zero) projection. With these observations in mind, it is easy to see that the identification in [Pat99, Chapter 4, Theorem 4.4.1] of the full C*-algebra attached to S in [Pat99, § 2.1] with the full groupoid C*-algebra C∗(Gu) respects these direct sum decompositions, i.e., it sends C∗(S) in the sense of Definition 5.9 to C∗(G(S)). Finally, it is also easy to see that the identification we get in this way really sends vs ∈ C∗(S) to the characteristic function on D(s, Us−1s). (cid:3) Next, we identify the reduced C*-algebras. Theorem 5.18. For every inverse semigroup S, there is a canonical isomorphism C∗λ(S) ∼=−→ C∗r (G(S)) sending the generator λs ∈ C∗λ(S) to the characteristic function on D(s, Us−1s), viewed as an element in Cc(G) ⊆ C∗r (G). We could give a proof of this result in complete analogy to the case of the full C*-algebras, using [Pat99, Chapter 4, Theorem 4.4.2] instead of [Pat99, Chapter 4, Theorem 4.4.1]. Instead, since all these C*-algebras are defined using concrete representations, we give a concrete proof identifying certain representations. 32 XIN LI Proof. For e ∈ E×, define It is then easy to see that S×e :=(cid:8)x ∈ S× : x−1x = e(cid:9) . This yields the direct sum decomposition S× = Ge∈E× ℓ2S× = Me∈E× S×e . ℓ2S×e . The left regular representation of S respects this direct sum decomposition. This is because given s ∈ S and x ∈ S×e with s−1s ≥ xx−1, we have that sx ∈ S×e since (sx)−1(sx) = x−1(s−1s)x = x−1(s−1sxx−1)x = x−1(xx−1)x = x−1x = e. Therefore, for every s ∈ S, we have λs = Me∈E× . λs(cid:12)(cid:12)ℓ2S× e Now define for every e ∈ E× the character χe ∈ bE by setting χe(f ) = 1 if e ≤ f, χe(f ) = 0 if e (cid:2) f. The map S×e −→ G(S)χe , x 7→ [x, χe] is surjective as every (x, χe) ∈ Σ is equivalent to (xe, χe), and xe lies in S×e as χe(x−1x) = 1 implies e ≤ x−1x. It is also injective as [x, χe] = [y, χe] for x, y ∈ S×e implies that xf = yf for some f ∈ E× with e ≤ f , and thus x = y. Therefore, the map above is a bijection. It induces a unitary U : ℓ2S×e ∼=−→ ℓ2G(S)χe , δx 7→ δ[x,χe]. Now let 1D(s,Us−1s) be the characteristic function on D(s, Us−1s), viewed as an element in Cc(G). Then we have (5) = πχe (1D(s,Us−1s)) ◦ U. This is because and e U ◦ λs(cid:12)(cid:12)ℓ2S× (U ◦ λs(cid:12)(cid:12)ℓ2S× e )(δx) = U (δsx) = [sx, χe] (πχe ◦ 1D(s,Us−1s) ◦ U )(δx) = πχe (1D(s,Us−1 s))([x, χe]) = [sx, χe] if s−1s ≥ xx−1, and both sides of (5) are zero if s−1s (cid:3) xx−1. Hence it follows that the left regular representation of C∗(S) is unitarily equivalent to under the isomorphism from Theorem 5.17. Me∈E× πχe SEMIGROUP C*-ALGEBRAS 33 Thus, all we have to show in order to conclude our proof is that (6) for all f ∈ Cc(G(S)). To show this, we first need to observe that e∈E× kπχe (f )k , sup χ∈ bE kπχ(f )k = sup (cid:8)χe : e ∈ E×(cid:9) is dense in bE. This is because a basis of open subsets for the topology of bE are given by U (e; e1, . . . , en) :=nχ ∈ bE : χ(e) = 1; χ(e1) = . . . = χ(en) = 0o , for e, e1, . . . , en ∈ E× with ei (cid:2) e. It is then clear that χe lies in U (e; e1, . . . , en). Because of density, (6) follows from [Pat99, Chapter 3, Proposition 3.1.2]. (cid:3) Remark 5.19. It is clear that the explicit isomorphisms provided by Theorem 5.17 and Theorem 5.18 give rise to a commutative diagram C∗(S) ∼= C∗(G(S)) C∗λ(S) ∼= / C∗r (G(S)) where the horizontal arrows are the left regular representations and the vertical arrows are the identifications provided by Theorem 5.17 and Theorem 5.18. 5.6. C*-algebras of partial dynamical systems as C*-algebras of partial transformation groupoids. Our goal is to identify the full and reduced crossed products attached to partial dynamical systems with full and reduced groupoid C*-algebras for the corresponding partial transformation groupoids. Given a partial dynamical system G y X, we have constructed its partial trans- formation groupoid G ⋉ X in § 5.3. The following result is [Aba04, Theorem 3.3]: Theorem 5.20. The canonical homomorphism Cc(G ⋉ X) → C0(X) ⋊ℓ1 G, θ 7→Xg θ(g, g−1.⊔)δg, where θ(g, g−1.⊔) is the function Ug−1 → C, x 7→ θ(g, g−1.x), extends to an iso- morphism C∗(G ⋉ X) ∼=−→ C0(X) ⋊ G. Here we use the same notation for partial dynamical systems and their crossed products as in § 5.2. Let us now identify reduced crossed products. / /     / 34 XIN LI Theorem 5.21. The canonical homomorphism (7) Cc(G ⋉ X) → C0(X) ⋊ℓ1 θ(g, g−1.⊔)δg, G, θ 7→Xg Ug−1 → C, x 7→ θ(g, g−1.x), where θ(g, g−1.⊔) is the function extends to an isomorphism C∗r (G ⋉ X) ∼=−→ C0(X) ⋊r G. We include a proof of this result. It is taken from [Li16b]. Proof. We use the same notation as in the construction of the reduced crossed product in § 5.2. As above, let µ × λ be the representation C0(X) ⋊ℓ1 G → L(H) which we used to define C0(X) ⋊r G. Our first observation is (8) To see this, observe that for all g ∈ G, δh ⊗ ℓ2Uh−1. im (µ × λ)(H) =Mh∈G im (Eg) ⊆Mh x /∈ h−1.(Uh ∩ Ug−1 ) = U(gh)−1 ∩ Uh−1, δh ⊗ ℓ2(Uh−1 ∩ U(gh)−1 ). This holds since for fUh(h.x) = 0 for f ∈ C0(Ug−1 ). Therefore, Hence and thus, π(C0(Ug−1 ))(δh ⊗ ℓ2X) ⊆ δh ⊗ ℓ2(Uh−1 ∩ U(gh)−1 ). im (Eg) ⊆Mh δh ⊗ ℓ2(Uh−1 ∩ U(gh)−1 ), δgh ⊗ ℓ2(Uh−1 ∩ U(gh)−1) ⊆Mh im (Vg) ⊆Mh δh ⊗ ℓ2Uh−1. This shows "⊆" in (8). For "⊇", note that for f ∈ C0(X), (µ × λ)(f δe) = µ(f )Ee, and for ξ ∈ ℓ2Uh−1, µ(f )Ee(δh ⊗ ξ) = δh ⊗ fUh (h.⊔)ξ. So (µ × λ)(f δe)(H) contains δh ⊗ f · ξ for all f ∈ C0(Uh−1) and ξ ∈ ℓ2Uh−1, hence also δh ⊗ ℓ2Uh−1. This proves "⊇". For x ∈ X, let Gx =(cid:8)g ∈ G : x ∈ Ug−1(cid:9) as before. Our second observation is that for every x ∈ X, the subspace Hx := ℓ2Gx ⊗ δx is (µ × λ)-invariant. It is clear that µ(f ) leaves Hx invariant for all f ∈ C0(X). For g, h ∈ G, Eg(δh ⊗ δx) = δh ⊗ δx SEMIGROUP C*-ALGEBRAS 35 if x ∈ Uh−1 ∩ U(gh)−1, and if that is the case, then Vg(δh ⊗ δx) = δgh ⊗ δx ∈ Hx. H = Mx∈X Hx! ⊕ (µ × λ)(C0(X) ⋊ℓ1 G)(H)⊥ is a decomposition of H into µ × λ-invariant subspaces. For x ∈ X, set Therefore, Then ρx := (µ × λ)Hx . C0(X) ⋊r G = C0(X) ⋊ℓ1 Gk·kLx ρx . Moreover, we have for x ∈ Uh−1, fgδg! (δh ⊗ δx) =Xg µ(fg)(δgh ⊗ δx) = Xg: x∈U(gh)−1 µ(fg)Vg(δh ⊗ δx) δgh ⊗ fg(gh.x)δx ρx Xg = Xg: x∈U(gh)−1 = Xk∈Gx (9) δk ⊗ fkh−1(k.x)δx. Let us compare this construction with the construction of the reduced groupoid C*-algebra of G ⋉ X. Obviously, (7) is an embedding of Cc(G ⋉ X) as a subalgebra which is k·kℓ1-dense in C0(X) ⋊ℓ1 G. Therefore, C0(X) ⋊r G = Cc(G ⋉ X)k·kLx ρx . Now, to construct the reduced groupoid C*-algebra C∗r (G ⋉ X), we follow our explanations in § 5.3 and construct for every x ∈ X the representation by setting πx : Cc(G ⋉ X) → L(ℓ2(s−1(x))) πx(θ)(ξ)(ζ) := Xη ∈ s−1(x) θ(ζη−1)ξ(η). In our case, using s−1(x) = Gx × {x}, we obtain for ξ = δh ⊗ δx with h ∈ Gx: πx(θ)(δh ⊗ δx)(k, x) = θ((k.x)(h, x)−1) = θ(kh−1, h.x). πx(θ)(δh ⊗ δx)(k, x) = Xk∈Gx θ(kh−1, h.x)δk ⊗ δx. Thus, (10) By definition, C∗r (G ⋉ X) = Cc(G ⋉ X)k·kLx πx . coincide on Cc(G ⋉ X), it Therefore, in order to show that k·kLx ρx suffices to show that for every x ∈ X, πx and the restriction of ρx to Cc(G ⋉ X) and k·kLx ρx ρx(θ)(δh ⊗ δx) θ(g, g−1.⊔)δg)(δh ⊗ δx) δk ⊗ θ(kh−1, h.x)δx (10) = πx(θ)(δh ⊗ δx). (7) = ρx(Xg = Xk∈Gx (9) 36 XIN LI are unitarily equivalent. Given x ∈ X, using s−1(x) = Gx × {x}, we obtain the canonical unitary ℓ2(s−1(x)) ∼= Hx = ℓ2(Gx) ⊗ δx, so that we may think of both ρx and πx as representations on ℓ2(Gx)⊗ δx. We then have for x ∈ X, θ ∈ Cc(G ⋉ X) and h ∈ Gx: This yields the canonical identification as desired. (cid:3) C0(X) ⋊r G ∼= C∗r (G ⋉ X), 5.7. The case of inverse semigroups admitting an idempotent pure partial homomorphism to a group. We would like to show that in the case of inverse semigroups which admit an idempotent pure partial homomorphism to a group, all our constructions above coincide. Let S be an inverse semigroup and E the semilattice of idempotents of S. Let G be a group. Assume that σ is a partial homomorphism S× → G which is idempotent pure. In this situation, we constructed a partial dynamical system G y bE in § 5.2. Our first observation is that the partial transformation groupoid of G y bE can be canonically identified with the groupoid G(S) we attached to S in § 5.4. Lemma 5.22. In the situation described above, we have a canonical identification of topological groupoids. G(S) ∼=−→ G ⋉ bE, [s, χ] 7→ (σ(s), χ). Proof. We use the notations from § 5.2 and § 5.4. To see that the mapping [s, χ] 7→ (σ(s), χ) is well-defined, suppose that (s, χ) and (t, χ) in Σ are equivalent. Then there exists e ∈ E× such that se = te, and se (or te) cannot be zero in case 0 ∈ S. Therefore, σ(s) = σ(se) = σ(te) = σ(t). To see that [s, χ] 7→ (σ(s), χ) is a morphism of groupoids, note that [s, χ]−1 = [s−1, s.χ] is sent to (σ(s−1), s.χ) = (σ(s), χ)−1. Hence our mapping respects in- verses. For multiplication, observe that and s([s, χ]) = χ = s(σ(s), χ) r([s, χ]) = s.χ = σ(s).χ = r(σ(s), χ). SEMIGROUP C*-ALGEBRAS 37 Moreover, [t, s.χ] · [s, χ] = [ts, χ] is mapped to [σ(ts), χ] = [σ(t), s.χ] · [σ(s), χ]. Hence it follows that our mapping is a groupoid morphism. We now set out to construct an inverse. Define the map G ⋉ bE −→ G(S), (g, χ) 7→ [s, χ] where for every g in G, we choose s ∈ S with σ(s) = g and χ(s−1s) = 1. This is well-defined: Given t ∈ S with σ(t) = g and χ(t−1t) = 1, set e := s−1st−1t. Then χ(e) = 1. Moreover, se = st−1t and te = ts−1s. As σ(se) = σ(s) = g = σ(t) = σ(te) and (se)−1(se) = e = (te)−1(te), we deduce by Lemma 5.7 that se = te. Hence (s, χ) ∼ (t, χ). It is easy to see that we have just constructed the inverse of Moreover, it is also easy to see that both our mappings are open, so that they give rise to the desired identification of topological groupoids. (cid:3) G(S) −→ G ⋉ bE, [s, χ] 7→ (σ(s), χ). Combining Theorem 5.17 with Theorem 5.20 and Theorem 5.18 with Theorem 5.21, we obtain the following Corollary 5.23. Let S be an inverse semigroup and E the semilattice of idempo- tents of S. Let G be a group. Assume that σ is a partial homomorphism S× → G which is idempotent pure. In this situation, we have canonical isomorphisms and C∗(S) → C∗(E) ⋊ G, vs 7→ λss−1 δσ(s) C∗λ(S) → C∗(E) ⋊r G, λs 7→ λss−1 Vσ(s). 6. Amenability and nuclearity Amenability is an important structural property for groups and groupoids, while nuclearity plays a crucial role in the structure theory for C*-algebras, in particular in the classification program. In the case of groups and groupoids, it is known that amenability and nuclearity of C*-algebras are closely related. Moreover, there are further alternative ways to characterize amenability in terms of C*-algebras. Our goal now is to explain to what extent analogous results hold true in the semigroup context. 6.1. Groups and groupoids. Let us start by reviewing the case of groups and groupoids. Let G be a discrete group. We recall three conditions. 38 XIN LI Definition 6.1. Our group G is said to be amenable if there exists a left invariant state on ℓ∞(G). This means that we require the existence of a state µ : ℓ∞(G) → C with the property that µ(f (s⊔)) = µ(f ) for every f ∈ ℓ∞(G) and s ∈ G. Here f (s⊔) is the function G → C, x 7→ f (sx). Definition 6.2. Our group G is said to satisfy Reiter's condition if there exists a net (θi)i of probability measures on G such that for all g ∈ G. lim i→∞ kθi − gθik = 0 Here gθ is the pushforward of θ under G ∼= G, x 7→ gx. Definition 6.3. Our group G is said to satisfy Følner's condition if for every finite subset E ⊆ G and every ε > 0, there exists a non-empty finite subset F ⊆ G with (sF )△F / F < ε for all s ∈ E. Here sF = {sx : x ∈ F}, and △ stands for symmetric difference. It turns out that a group is amenable if and only if it satisfies Reiter's condition if and only if it satisfies Følner's condition. We refer the reader to [BO08, Chapter 2, § 6] for more details. All abelian, nilpotent and solvable groups are amenable, to mention some examples. Non-abelian free groups are not amenable. We now turn to groupoids. Definition 6.4. An ´etale locally compact groupoid G is amenable if there is a net (θi)i of continuous systems of probability measures θi = (θx i )x∈G(0) with i − γθs(γ) i lim i→∞(cid:13)(cid:13)(cid:13)θr(γ) (cid:13)(cid:13)(cid:13) = 0 for all γ ∈ G. Here θx is a probability measure on G with support contained in Gx. "Continuous" means that for every f ∈ Cc(G), the function is continuous. As above, γθ is the pushforward of θ under G(0) → C, x 7→Z f dθx Gs(γ) → Gr(γ), η 7→ γη. SEMIGROUP C*-ALGEBRAS 39 Note that what we call amenability of groupoids is really Reiter's condition for groupoids. Moreover, we may require that the convergence in our definition happens uniform on compact subsets of G. This is because of [Ren15]. For instance, if G is an amenable group, and G y Ω is a partial dynamical system on a locally compact Hausdorff space Ω, then the partial transformation groupoid G ⋉ Ω is amenable by [Exe15, Theorem 20.7 and Theorem 25.10]. But we can get amenable partial transformation groupoids even if G is not amenable. Let us now introduce nuclearity for C*-algebras. Definition 6.5. A C*-algebra A is nuclear if there exists a net of contractive : A → Fi and ψi : Fi → A, where Fi are finite completely positive maps ϕi dimensional C*-algebras, such that for all a ∈ A. lim i→∞kψi ◦ ϕi(a) − ak = 0 For instance, all commutative C*-algebras are nuclear, and all finite dimensional C*-algebras are nuclear. The reader may find more about nuclearity for C*-algebras for example in [BO08, Chapter 2]. Let us now relate amenability and nuclearity. Let us start with the case of groups. Recall that the full group C*-algebra C∗(G) of a discrete group G is the C*-algebra universal for unitary representations of G. This means that C∗(G) is generated by unitaries ug, g ∈ G, satisfying and whenever we find unitaries vg, g ∈ G, in another C*-algebra B satisfying ugh = uguh for all g, h ∈ G, vgh = vgvh for all g, h ∈ G, then there exists a (unique) *-homomorphism C∗(G) → B sending ug to vg. The reduced group C*-algebra C∗λ(G) of a discrete group G is the C*-algebra gen- erated by the left regular representations of G. The left regular representation is exactly what we get when we apply the construction at the beginning of § 2 to G. Therefore, C∗λ(G) is the C*-algebra we get when we apply Definition 2.1 to G in place of P . By construction, we have a canonical *-homomorphism λ : C∗(G) → C∗λ(G), ug → λg. It is called the left regular representation (of C∗(G)). 40 XIN LI Here are a couple of C*-algebraic characterizations of amenability for groups. We refer the reader to [BO08, Chapter 2, § 6] for details and proofs. Theorem 6.6. Let G be a discrete group. The following are equivalent: • G is amenable. • C∗(G) is nuclear. • C∗λ(G) is nuclear. • The left regular representation λ : C∗(G) → C∗λ(G) is an isomorphism. • There exists a character on C∗λ(G). Here, by a character on a unital C*-algebra A, we simply mean a unital *-homomorphism from A to C. We now turn to groupoids and C*-algebraic characterizations of amenability for them. We already introduced full and reduced groupoid C*-algebras in § 5.3. We also introduced the left regular representation (of the full groupoid C*-algebra) Theorem 6.7. Let G be an ´etale locally compact groupoid. Consider the statements λ : C∗(G) → C∗r (G). (i) G is amenable. (ii) C∗(G) is nuclear. (iii) C∗λ(G) is nuclear. (iv) λ : C∗(G) → C∗λ(G) is an isomorphism. Then (i) ⇔ (ii) ⇔ (iii) ⇒ (iv). We refer to [BO08, Chapter 5, § 6] and [ADR00] for more details. It was an open question whether statement (iv) implies the other statements. But Rufus Willett gave a counterexample in [Wil15]. There are, however, results say- ing that statement (iv) does imply the other statements for particular classes of groupoids. For instance, we mention [Mat14]. 6.2. Amenability for semigroups. Let us now turn to amenability for semi- groups. As in the group case, we have the following definitions: Definition 6.8. A discrete semigroup P is called left amenable if there exists a left invariant mean on ℓ∞(P ), i.e. a state µ on ℓ∞(P ) such that for every p ∈ P and f ∈ ℓ∞(P ), µ(f (p⊔)) = µ(f ). Here f (p⊔) is the function P → C, x 7→ f (px). For instance, every abelian semigroup is left amenable. SEMIGROUP C*-ALGEBRAS 41 Definition 6.9. A discrete semigroup P is said to satisfy Reiter's condition if there is a net (θi)i of probability measures on P with the property that i kθi − pθik = 0 for all p ∈ P. lim Here pθ is the pushforward of θ under P → P, x 7→ px. Definition 6.10. A discrete semigroup P satisfies the strong Følner condition if for every finite subset E ⊆ P and every ε > 0, there exists a non-empty finite subset F ⊆ P such that (pF )△F / F < ε for all p ∈ C. Here pF = {px : x ∈ F} and △ stands for symmetric difference. As in the group case, a discrete left cancellative semigroup is left amenable if and only if it satisfies Reiter's condition if and only if it satisfies the strong Følner condition. The reader may consult [Li12] for a proof, and we also refer to [Pat88] for more details. Our goal now is to find the analogues of Theorem 6.6 and Theorem 6.7 in the context of semigroups and their C*-algebras. The motivation is to understand and explain – in a conceptual way – the following two observations: Let P = N × N, the universal monoid generated by two commuting elements. This is an abelian semigroup, so it is left amenable. So far, we have not discussed the question how to construct full semigroup C*-algebras. But a natural candidate for the full semigroup C*-algebra of N × N would be C∗ (va, vb v∗ava = 1, v∗b vb = 1, vavb = vbva) . In other words, this is the universal C*-algebra generated by two commmuting isometries. It is the C*-algebra universal for isometric representations of our semi- group. This is a very natural candidate for the full semigroup C*-algebra. But Murphy showed that this C*-algebra is not nuclear in [Mur96, Theorem 6.2]. Next, consider P = N ∗ N, the non-abelian free monoid on two generators. As in the group case, non-abelian free semigroups are examples of semigroups which are not left amenable. But it is easy to see that C∗λ(N∗ N) is generated as a C*-algebra by two isometries Va and Vb with orthogonal range projections, i.e., (VaV ∗a ) · (VbV ∗b ) = 0. Therefore, C∗λ(N∗ N) is isomorphic to the canonical extension of the Cuntz algebra O2, as introduced in [Cun77, § 3]. It fits into an exact sequence 0 → K → C∗λ(N ∗ N) → O2 → 0, where K is the C*-algebra of compact operators on a infinite dimensional and separable Hilbert space. Hence it follows that C∗λ(N ∗ N) is nuclear. Moreover, 42 XIN LI C∗λ(N ∗ N) can be described as a universal C*-algebra, because C∗λ(N ∗ N) ∼= C∗ (va, vb v∗ava = 1, v∗b vb = 1, vav∗avbv∗b = 0) . So this is a hint that for the semigroup N ∗ N, the full and reduced semigroup C*-algebras are isomorphic. But, as we remarked above, N∗ N is not left amenable. Our goal now is to explain these phenomena, to clarify the relation between amenabil- ity and nuclearity, and to obtain analogues of Theorem 6.6 and Theorem 6.7 in the context of semigroups. The first step for us will be to find a systematic and rea- sonable way to define full semigroup C*-algebras. It turns out that left inverse hulls attached to left cancellative semigroups, as introduced in § 5.1, give rise to an approach to this problem. However, before we come to the construction of full semigroup C*-algebras, we first need to compare the reduced C*-algebras of left cancellative semigroups and their left inverse hulls. 6.3. Comparing reduced C*-algebras for left cancellative semigroups and their left inverse hulls. Let P be a left cancellative semigroup and Il(P ) the left inverse hull attached to P , as in § 5.1. As we explained in § 5.1, we have a canonical embedding of P into Il(P ), denoted by It gives rise to the isometry P ֒→ Il(P ), p 7→ p. Thus, we may think of ℓ2P as a subspace of ℓ2S×. I : ℓ2P → ℓ2S×, δp 7→ δp. The following observation appears in [Nor14, § 3.2]. Lemma 6.11. Assume that P is a left cancellative semigroup with left inverse hull Il(P ). Then the subspace ℓ2P of ℓ2Il(P )× is invariant under C∗λ(Il(P )). Moreover, we obtain a well-defined surjective *-homomorphism sending λp to Vp for every p ∈ P . C∗λ(Il(P )) → C∗λ(P ), T 7→ I∗T I Proof. We first claim that every s ∈ Il(P ) has the following property: (11) For every x ∈ dom(s) and every r ∈ P, xr lies in dom(s), and s(xr) = s(x)r. To prove our claim, first observe that for every p ∈ P , the partial bijection p ∈ Il(P ) certainly has this property, as it is just given by left multiplication with p. Moreover, p−1 is the partial bijection Certainly, for every px ∈ pP and every r ∈ P , pxr lies in pP , and p−1(pxr) = xr = p−1(px)r. pP → P, px 7→ x. Hence p−1 has the desired property as well. To conclude the proof of our claim, suppose that s, t ∈ Il(P ) both have the desired property. Choose x ∈ dom(st). Then for every r ∈ P , xr lies in dom(t), and t(xr) = t(x)r. Since t(x) lies in SEMIGROUP C*-ALGEBRAS 43 dom(s), t(x)r lies in dom(s) as well. The conclusion is that xr lies in dom(st), and we have (st)(xr) = s(t(x)r) = s(t(x))r = (st)(x)r. As every element in Il(P ) is a finite product of partial bijections in this proves our claim. {p : p ∈ P} ∪(cid:8)p−1 : p ∈ P(cid:9) , The second step is to show that for every s ∈ Il(P ) and x ∈ P with s−1s ≥ pp−1, we must have sx = s(x) ∈ P . This is because we have, for every y ∈ P : (sx)(y) = s(x(y)) = s(xy) = s(x)y = (s(x))(y). Here we used our first claim from above. Now let s ∈ Il(P ) be arbitrary. We want to show that λs(ℓ2P ) ⊆ ℓ2P . Given x ∈ P , we have λs(δx) = 0 if s−1s (cid:3) pp−1. If s−1s ≥ pp−1, then what we showed in the second step implies that λs(δx) = δs(x) lies in ℓ2P . As s was arbitrary, this shows that C∗λ(Il(P ))(ℓ2P ) ⊆ ℓ2P. Therefore, every T ∈ C∗λ(Il(P )) satisfies T II∗ = II∗T II∗, and since C∗λ(Il(P )) is *-invariant, we even obtain that every T ∈ C∗λ(Il(P )) satisfies T II∗ = II∗T . This shows that the map is a *-homomorphism. Its image is C∗λ(P ) because we have, for p ∈ P and x ∈ P : C∗λ(Il(P )) → L(ℓ2P ), T 7→ I∗T I λp(δx) = δpx = Vp(δx), so that I∗λpI = Vp for all p ∈ P . (cid:3) Recall that we denote the semilattice of idempotents in Il(P ) by JP , and we iden- tified this semilattice with the constructible right ideals of P (see § 5.1). Moreover, we also introduced in § 5.2 the sub-C*-algebra of C∗λ(Il(P )) generated by JP : C∗(JP ) = C∗({λX : X ∈ JP}). It is easy to see that for every X ∈ JP , we get I∗λX I = 1X, where 1X is the characteristic function of X, viewed as an element in ℓ∞(P ). Hence, restricting the *-homomorphism C∗λ(Il(P )) → C∗λ(P ) from Lemma 6.11 to C∗(JP ), we obtain a *-homomorphism from C∗(JP ) onto the sub-C*-algebra Dλ(P ) = C∗({1X : X ∈ JP}) of C∗λ(P ), which is generated by {1X : X ∈ JP}, C∗(JP ) ։ Dλ(P ), T 7→ I∗T I. 44 XIN LI Obviously, if the *-homomorphism from Lemma 6.11 is an isomorphism, then its restriction to C∗(JP ) must be an isomorphism (onto its image) as well. Let us now discuss a situation when the converse holds. We need the following Lemma 6.12. Let X be a set. There exists a faithful conditional expectation ΘX : L(ℓ2X) ։ ℓ∞(X) such that, for every T ∈ L(ℓ2X), we have (12) for all x, y ∈ X. hΘX(T )δx, δyi = δx,y hT δx, δyi Proof. Let ex,x be the rank one projection onto Cδx ⊆ ℓ2X, given by ex,x(ξ) = hξ, δxi δx for all ξ ∈ ℓ2X. Consider the linear map We have (13) span({δx : x ∈ X}) → span({δx : x ∈ X}),Xx (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xx αx(ex,x ◦ T )(δx)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 2 αxδx 7→Xx αx(ex,x ◦ T )(δx). = *Xx = Xx ≤ kTk2Xx αx(ex,x ◦ T )(δx),Xx αx2 h(ex,x ◦ T )(δx), (ex,x ◦ T )(δx)i αx2 = kTk2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xx αx(ex,x ◦ T )(δx)+ αxδx(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 2 So the linear map in (13) extends to a bounded linear operator ℓ2X → ℓ2X, which we denote by ΘX(T ). Our computation shows that kΘX(T )k ≤ kTk . By definition, This shows that ΘX(T ) lies in ℓ∞(X). It also shows that ΘX (T ) satisfies (12). ΘX(T )(δx) = hT δx, δxi δx. Moreover, by construction, ΘX(T ) = T for all T ∈ ℓ∞(X). Therefore, the map ΘX : L(ℓ2X) → ℓ∞(X), T 7→ ΘX (T ) is a projection of norm 1. Hence it follows by [Bla06, Theorem II.6.10.2] that ΘX is a conditional expectation. Finally, ΘX is faithful because given T ∈ L(ℓ2X), ΘX (T ∗T ) = 0 implies that 0 = hT ∗T δx, δxi = kT δxk2 , so that T δx = 0 for all x ∈ X, and hence T = 0. (cid:3) SEMIGROUP C*-ALGEBRAS 45 Applying Lemma 6.12 to X = Il(P )× and X = P , we obtain faithful conditional expectations and ΘIl(P ) : L(ℓ2Il(P )×) ։ ℓ∞(Il(P )×) They fit into the following commutative diagram: ΘP : L(ℓ2P ) ։ ℓ∞(P ). (14) L(ℓ2Il(P )×) ΘIl(P ) L(ℓ2Il(P )×) I∗ ⊔ I I∗ ⊔ I L(ℓ2P ) ΘP / ℓ∞(P ) Here I∗ ⊔ I is our notation for the map sending T to I∗T I. Commutativity of the diagram above follows from the following computation: ΘP (I∗T I) δx = hI∗T Iδx, δxi δx = hT δx, δxi δx = (I∗ΘIl(P )(T )I) δx. This leads us to Corollary 6.13. Assume that (15) Then the *-homomorphism ΘIl(P )(C∗λ(Il(P ))) = C∗(JP ). from Lemma 6.11 is an isomorphism if and only if its restriction to C∗(JP ), C∗λ(Il(P )) → C∗λ(P ), T 7→ I∗T I C∗(JP ) ։ Dλ(P ), T 7→ I∗T I, is an isomorphism. Proof. Take the commutative diagram (14) and restrict the upper left corner to As I∗C∗λ(Il(P ))I = C∗λ(P ) by Lemma 6.11, and because of (15), we obtain the commutative diagram C∗λ(Il(P )) ⊆ L(ℓ2Il(P )×). (16) C∗λ(Il(P )) ΘIl(P ) C∗(JP ) I∗ ⊔ I I∗ ⊔ I C∗λ(P ) ΘP / Dλ(P ) As the vertical arrows are faithful, it is now easy to see that if the lower horizontal arrow is faithful, the upper horizontal arrow has to be faithful as well. This proves our corollary. (cid:3) Remark 6.14. The condition (15), i.e., implies that ΘIl(P )(C∗λ(Il(P ))) = C∗(JP ), C∗(JP ) = C∗λ(Il(P )) ∩ ℓ∞(Il(P )×), / /     / / /     / 46 and XIN LI This is because we always have Dλ(P ) = C∗λ(P ) ∩ ℓ∞(P ). (17) and (18) C∗(JP ) ⊆ C∗λ(Il(P )) ∩ ℓ∞(Il(P )×) ⊆ ΘIl(P )(C∗λ(Il(P ))), Dλ(P ) ⊆ C∗λ(P ) ∩ ℓ∞(P ) ⊆ ΘP (C∗λ(P )), and (15) implies that all these inclusions are equalities in (17), and also in (18) because Dλ(P ) = I∗ C∗(JP ) I = ΘP (C∗λ(P )). (15) = I∗ ΘIl(P )(C∗λ(Il(P ))) I = ΘP (I∗ C∗λ(Il(P )) I) Here we used commutativity of the diagram in (16) and Lemma 6.11. It remains to find out when condition (15) holds. We follow [Nor14, § 3.2]. Let us introduce the following Definition 6.15. An inverse semigroup S is called E*-unitary if for every s ∈ S, we must have s ∈ E if there exists x ∈ S× with sx = x. Remark 6.16. If there exists an idempotent pure partial homomorphism σ : S× → G to some group G, then S is E*-unitary. This is because if we are given s ∈ S, and there exists x ∈ S× with sx = x, then σ(x) = σ(s)σ(x), so that σ(s) = e, where e is the identity element in G. Since σ is idempotent pure, s must lie in E. Now we apply Lemma 6.12 to X = S×. Then we get a faithful conditional expec- tation ΘS× : L(ℓ2S×) ։ ℓ∞(S×), and we may apply it to elements in C∗λ(S). Lemma 6.17. In the situation above, our inverse semigroup S is E*-unitary if and only if for every s ∈ S, we always have or ΘS×(λs) = 0 s ∈ E and ΘS×(λs) = λs. Proof. For "⇒", assume that ΘS×(λs) 6= 0. This is equivalent to saying that there exists x ∈ S× with sx = x. But since S is E*-unitary, this implies s ∈ E. And since λs lies in ℓ∞(S) for all s ∈ E, we must have ΘS×(λs) = λs. Conversely, for "⇐", take s ∈ S and suppose that there is x ∈ S× with sx = x. Then sxx−1 = xx−1, so that sxx−1 is idempotent, and we conclude that s−1sxx−1 = (xx−1s−1)(sxx−1) = sxx−1 = xx−1, SEMIGROUP C*-ALGEBRAS 47 i.e., s−1s ≥ xx−1. Hence λs(δx) = δsx = δx. Hence it follows that ΘS×(λs) 6= 0, and this implies, by assumption, that s lies in E. (cid:3) In particular, we can draw the following conclusion Corollary 6.18. If S is an E*-unitary inverse semigroup, then ΘS×(C∗λ(S)) = C∗(E). Combining Corollary 6.13, Remark 6.14, Corollary 6.18, Remark 6.16 and the ob- servation that Il(P ) admits an idempotent pure partial homomorphism to a group if P embeds into a group (see § 5.1), we obtain Corollary 6.19. Assume that P is a semigroup which embeds into a group G. Then condition (15) holds, i.e., and the *-homomorphism ΘIl(P )(C∗λ(Il(P ))) = C∗(JP ), from Lemma 6.11 is an isomorphism if and only if its restriction to C∗(JP ), C∗λ(Il(P )) → C∗λ(P ), T 7→ I∗T I C∗(JP ) ։ Dλ(P ), T 7→ I∗T I, is an isomorphism. Moreover, and (19) C∗(JP ) = C∗λ(Il(P )) ∩ ℓ∞(Il(P )×), Dλ(P ) = C∗λ(P ) ∩ ℓ∞(P ). Corollary 6.19 prompts the question when the *-homomorphism C∗(JP ) ։ Dλ(P ), T 7→ I∗T I, is an isomorphism. Note that both C∗(JP ) and Dλ(P ) are generated by a family of commuting projections, closed under multiplication, and our *-homomorphism sends generator to generator, i.e., λX to 1X for all X ∈ JP . Let us now investigate when such a *-homomorphism is an isomorphism. 6.4. C*-algebras generated by semigroups of projections. We basically fol- low [Li12, § 2.6] in this subsection. If we think of elements of an inverse semigroup as partial isometries on a Hilbert space, then the semilattice of idempotents is a family of commuting projections, closed under multiplication, or in other words, a semigroup of projections. 48 XIN LI Let us consider the general setting of a semilattice E of idempotents, i.e., E is an abelian semigroup consisting of idempotents. Suppose that D is a C*-algebra generated by a multiplicatively closed family {de : e ∈ E} of projections such that E → D, e 7→ de is a semigroup homomorphism. We make the following easy observation df , f ∈ F . Lemma 6.20. For every finite subset F of E, there exists a projection in D, denoted byWf∈F df , which is the smallest projection dominating all the projections Moreover, with E(F ) denoting the subsemigroup of E generated by F ,Wf∈F df lies in span({de : e ∈ E(F )}). Just to be clear, the projectionWf∈F df is uniquely characterized by df ≤ _f∈F df for all f ∈ F, and whenever a projection d ∈ D satisfies then we must have df ≤ d for all f ∈ F, _f∈F df ≤ d. Proof. We proceed inductively on the cardinality of F . The case F = 1 is trivial. Now assume that our claim holds for a finite subset F , and take an arbitrary element f ∈ E. We want to check our claim for F ∪n fo. Consider the element (20) _f∈F df + d f −_f∈F df · d f . It is easy to see that this is projection in D, which dominates all the df , f ∈ F , as well as d f . Moreover, if d is a projection in D which dominates all the df , f ∈ F , and also d f , then d obviously also dominates the projection in (20). Furthermore, sinceWf∈F df lies in span({de : e ∈ E(F )}) by induction hypothesis, the projection in (20) lies in span(nde : e ∈ E(F ∪n fo)o). (cid:3) SEMIGROUP C*-ALGEBRAS 49 As above, let E be a semilattice of idempotents. Suppose that D is a C*-algebra generated by projections {de : e ∈ E} such that d0 = 0 if 0 ∈ E and def = dedf for all e, f ∈ E. We prove the following result about *-homomorphisms out of D. Proposition 6.21. Let B be a C*-algebra containing a semigroup of projections {be : e ∈ E} such that b0 = 0 if 0 ∈ E and bef = bebf for all e, f ∈ E. There exists a *-homomorphism D → B sending de to be for all e ∈ E if and only of for every e ∈ E and every finite subset F ⊆ E such that f (cid:12) e for all f ∈ F , the equation implies that de = _f∈F be = _f∈F df in D bf in B. In that case, the kernel of the *-homomorphism D → B, de → be is generated by de − _f∈F df ∈ D : e ∈ E, F ⊆ {f ∈ E : f (cid:12) e} finite, be = _f∈F  bf in B . Proof. Let us start with the first part. Our condition is certainly a necessary condition for the existence of a *-homomorphism D → B, de → be. To prove that it is also sufficient, write E as an increasing union of finite subsemigroups Ei, i.e., is a family of pairwise orthogonal projections which generates Di. Moreover, it is also easy to see that Let Di := C∗({de : e ∈ Ei}). Obviously, For every e ∈ Ei, let Fe := {f ∈ Ei : f (cid:12) e}. Then, by Lemma 6.20, is a projection in Di. It is easy to see that Ei. Di. df E =[i D =[i de − _f∈Fe  df : e ∈ Ei de − _f∈Fe  bf : e ∈ Ei be − _f∈Fe 50 XIN LI is a family of pairwise orthogonal projections in B. Hence it follows that there exists a *-homomorphism Di → B sending to for all e ∈ Ei if and only if implies de − _f∈Fe be − _f∈Fe df bf de − _f∈Fe be − _f∈Fe df = 0 in Di bf = 0 in B, for all e ∈ Ei. But this is precisely the condition in the first part of our proposition. Moreover, it is easy to see that the *-homomorphism Di → B we just constructed sends de to be for all e ∈ Ei. Hence these *-homomorphisms, taken together for all B. i, are compatible and give rise to the desired *-homomorphism from D =Si Di to For the second part of the proposition, let I be the ideal of D generated by de − _f∈F df ∈ D : F ⊆ E finite, be = _f∈F  bf in B . Obviously, I is contained in the kernel of D → B, de 7→ be. It remains to show that the induced *-homomorphism D/I → B is injective. With the Dis as above, set Ii := I ∩ Di. Obviously, we have Hence it suffices to prove that the restriction Di/Ii → B is injective, or in other words, that the *-homomorphism Di → B we constructed above has kernel equal to Ii. But we have seen that Di/Ii. I =[i  de − _f∈Fe Ii and D/I =[i df : e ∈ Ei de − _f∈Fe df is a family of pairwise orthogonal projections which generates Di. So the kernel is generated by those projections for which we have Therefore, the kernel is Ii, as required. be − _f∈Fe bf = 0 in B. (cid:3) SEMIGROUP C*-ALGEBRAS 51 As before, let D be a C*-algebra generated by a semigroup {de : e ∈ E} of projec- tions such that d0 = 0 if 0 ∈ E and def = dedf for all e, f ∈ E. We set E× := E if E is a semilattice without zero, and E× := E \ {0} if 0 ∈ E. Proposition 6.22. The following are equivalent: (i) Our C*-algebra D is universal for representations of E by projections, i.e., we have an isomorphism D ∼=−→ C∗({ve : e ∈ E} v∗e = ve = v2 e , v0 = 0 if 0 ∈ E, vef = vevf ) sending de to ve. (ii) For every e ∈ E and every finite subset F ⊆ E with f (cid:12) e for all f ∈ F , we have (iii) The projections {de : e ∈ E×} are linearly independent in D. Proof. Obviously, (iii) implies (ii). df (cid:12) de. _f∈F de = _f∈F df in D Moreover, (ii) implies (i) by Proposition 6.21, because if (ii) holds, we can never have for any finite subset F ⊆ E with f (cid:12) e for all f ∈ F . It remains to prove that (i) implies (iii). First of all, consider the left regular representation λ on ℓ2E× as in § 5.1. It is given by λeδx = δx if e ≥ x and λeδx = 0 if e (cid:3) x. By universal property of D, there is a *-homomorphism D → L(ℓ2E×) sending de to λe. But it is easy to see that λe = λf if and only if e = f . Hence it follows that de = df if and only if e = f . Furthermore, again by universal property of D, there exists a *-homomorphism Let D → D ⊗ D, de 7→ de ⊗ de. D = span({de : e ∈ E}) ⊆ D. Restricting the *-homomorphism D → D ⊗ D from above to D, we obtain a ho- momorphism ∆ : D → D ⊙ D which is determined by de 7→ de ⊗ de for every e ∈ E. We now deduce from the existence of such a homomorphism ∆ that {de : e ∈ E×} is a C-basis of D. As {de : e ∈ E×} generates D as a C-vector space, we can always find a subset S of E× such that {de : e ∈ S} is a C-basis for D. It then follows that {de ⊗ df : e, f ∈ S} is a C-basis of D ⊙ D. 52 XIN LI Now take e ∈ E×. We can find finitely many ei ∈ S and αi ∈ C with de =Pi αidei . Applying ∆ yields Xi,j αiαjdei ⊗ dej = de ⊗ de = ∆(de) =Xi αi∆(dei ) =Xi αidei ⊗ dei . Hence it follows that among the αis, there can only be one non-zero coefficient which must be 1. The corresponding vector dei must then coincide with de. This implies e = ei ∈ S, i.e. {de : e ∈ E×} is a C-basis of D. This proves (iii). (cid:3) Now let S be an inverse semigroup with semilattice of idempotents E, and let C∗λ(S) be its reduced C*-algebra. Recall that we defined Lemma 6.23. The C*-algebra C∗(E) is universal for representations of E by pro- jections. C∗(E) := {λe : e ∈ E} . Proof. By Proposition 6.22, all we have to show is that for every e ∈ E and every finite subset F ⊆ E with f (cid:12) e for all f ∈ F , we have λf (cid:12) λe. _f∈F But this follows from λf (δe) = 0 for all f ∈ E with f (cid:12) e, while λe(δe) = δe for all e ∈ E×. (cid:3) It turns out that C∗(E) can be identified with the corresponding sub-C*-algebra of the full C*-algebra of S. Corollary 6.24. We have an isomorphism C∗(E) ∼=−→ C∗({ve : e ∈ E}) ⊆ C∗(S) sending λe to ve for all e ∈ E. Proof. By Lemma 6.23, there is a *-homomorphism C∗(E) ∼=−→ C∗({ve : e ∈ E}) ⊆ C∗(S) sending λe to ve for all e ∈ E. It is an isomorphism because the inverse is given by restricting the left regular representation C∗(S) → C∗λ(S) to C∗({ve : e ∈ E}) ⊆ C∗(S). (cid:3) This justifies why we denote the sub-C*-algebra C∗({λe : e ∈ E}) of C∗λ(S) by C∗(E). Corollary 6.25. We have a canonical identification bE ∼= Spec (C∗(E)). Proof. This is because by universal property of C∗(E) (see Lemma 6.23), there is a one-to-one correspondence between non-zero *-homomorphisms C∗(E) → C and non-zero semigroup homomorphisms E → {0, 1} (sending 0 to 0 if 0 ∈ E). (cid:3) SEMIGROUP C*-ALGEBRAS 53 Now suppose that we have an inverse semigroup S with semilattice of idempotents E, and that we have a surjective *-homomorphism C∗(E) → D sending λe → de. Then D is a commutative C*-algebra, and we can describe its spectrum as follows: Corollary 6.26. Viewing Spec (D) as a closed subspace of bE, Spec (D) is given by the subspace of all χ ∈ bE with the property that whenever we have e ∈ E de =Wf∈F df in D, then we must have χ(f ) = 1 for some f ∈ F . with χ(e) = 1 and a finite subset F ⊆ E with f (cid:12) e for every f ∈ F satisfying Proof. This is an immediate consequence of Proposition 6.21. (cid:3) Now let us suppose that we have a left cancellative semigroup P . We now apply Corollary 6.26 and Proposition 6.22 to the situation where S = Il(P ), E = JP and D = Dλ(P ) ⊆ C∗λ(P ). First, we make the following easy observation: Lemma 6.27. Suppose that we are given finitely many Xi ∈ JP . Then we have _i 1Xi = 1Si Xi in Dλ(P ) ⊆ ℓ∞(P ). The following follows immediately from Corollary 6.26: Corollary 6.28. The spectrum ΩP = Spec (Dλ(P )) is given by the closed subspace of cJP consisting of all χ ∈ cJP with the property that for all X ∈ JP with χ(X) = 1 and all X1, . . . , Xn ∈ JP with X = Sn i=1 Xi in P , we must have χ(Xi) = 1 for some 1 ≤ i ≤ n. Proposition 6.22 yields in our situation: Corollary 6.29. The following are equivalent: • We have an isomorphism Dλ(P ) ∼=−→ C∗{vX : X ∈ JP} • We have an isomorphism v∗X = vX = v2 X , v0 = 0 if 0 ∈ JP , vX∩Y = vX vY )  , 1X 7→ vX . C∗(E) ∼=−→ Dλ(P ), λX 7→ 1X. • For every X ∈ JP and all X1, . . . , Xn ∈ JP , X = Xi n[i=1 implies that X = Xi for some 1 ≤ i ≤ n. • The projections (cid:8)1X : X ∈ J ×P(cid:9) are linearly independent in Dλ(P ). 54 XIN LI 6.5. The independence condition. Corollary 6.29 justifies the following Definition 6.30. We say that our left cancellative semigroup P satisfies the inde- pendence condition (or simply independence) if for every X ∈ JP and all X1, . . . , Xn ∈ JP , X = implies that X = Xi for some 1 ≤ i ≤ n. Xi n[i=1 Let us now discuss examples of left cancellative semigroups which satisfy indepen- dence, and also some examples which do not. We start with the following Lemma 6.31. Suppose that P is a left cancellative semigroup with identity e. If every non-empty constructible right ideal of P is principal, i.e., J ×P = {pP : p ∈ P} , then P satisfies independence. Proof. Suppose that pP = piP n[i=1 for some p, p1, . . . , pn ∈ P . Then, since P has an identity, the element p lies in pP , hence we must have p ∈ piP for some 1 ≤ i ≤ n. But then, since piP is a right ideal, we conclude that pP ⊆ piP . Hence it follows that pP = piP , since we always have pP ⊇ piP . (cid:3) When are all non-empty constructible right ideals principal? Here is a necessary and sufficient condition: Lemma 6.32. For a left cancellative semigroup P (with or without identity), we have if and only if the following criterion holds: J ×P = {pP : p ∈ P} For all p, q ∈ P with pP ∩ qP 6= ∅, there exists r ∈ P with pP ∩ qP = rP . Proof. Our criterion is certainly necessary, since JP is a semilattice, hence closed under intersections. To show that our condition is also sufficient, we first observe that JP can be characterized as the smallest family of subsets of P containing P itself and closed under left multiplication, i.e., as well as pre-images under left multiplication, i.e., X ∈ JP , p ∈ P ⇒ p(X) ∈ cJP , X ∈ JP , q ∈ P ⇒ q−1(X) ∈ cJP . SEMIGROUP C*-ALGEBRAS 55 Now {pP : p ∈ P} is obviously closed under left multiplication. Hence it suffices to prove that principal right ideals are also closed under pre-images under left multiplication, up to ∅. Take p, q ∈ P . We always have q−1(pP ) = q−1(pP ∩ qP ). Therefore, if pP ∩ qP = ∅, then q−1(pP ) = ∅. If pP ∩ qP 6= ∅, then by our criterion, there exists r ∈ P with pP ∩ qP = rP . As rP ⊆ qP , we must have r ∈ qP , so that we can write r = qx for some x ∈ P . Therefore, we conclude that q−1(pP ) = q−1(pP ∩ qP ) = q−1(rP ) = q−1(qxP ) = xP. (cid:3) For instance, positive cones in totally ordered groups (as in § 3.2) always satisfy independence. This is because if P is such a positive cone, then for p, q ∈ P , we have pP ∩ qP = pP if p ≥ q and pP ∩ qP = qP if p ≤ q. Hence, all constructible right ideals are principal by Lemma 6.32. Moreover, right-angled Artin monoids (see § 3.3) satisfy independence. Actu- ally, all non-empty constructible right ideals are principal, because the criterion of Lemma 6.32 is true. This will come out of our general discussion of graph prod- ucts in § 9. To discuss more examples, let us explain a general method for verifying the criterion in Lemma 6.32. This is based on [Deh03]. Suppose that we are given a monoid P defined by a presentation, i.e., generators Σ and relations R, so that P = hΣ Ri+. Assume that all the relations in R are of the form w1 = w2, where w1 and w2 are formal words in Σ. Now we introduce formal symbols and look at formal words in Σ and Σ−1. For two such words w and w′, we write w yR w′ if w can be transformed into w′ be finitely many of the following two possible steps: (cid:8)σ−1 : σ ∈ Σ(cid:9) =: Σ−1, • Delete σ−1σ. • Replace σ−1 i σj by uv−1 if σiu = σj v is a relation in R. We then say that our presentation (Σ, R) is complete for yR if for two formal words u and v in Σ, we have u−1v yR ε (where ε is the empty word) if and only if u and v define the same element in our monoid P = hΣ Ri+. There are criteria on (Σ, R) which ensure completeness for yR (see [Deh03]). If completeness for yR is given, then we can read of properties of our monoid P = hΣ Ri+ from the presentation (Σ, R). We refer the reader to [Deh03] for a 56 XIN LI general and more complete discussion. For our purposes, the following observation is important: If (Σ, R) is complete for yR, then P = hΣ Ri+ has the property that for all p, q ∈ P with pP ∩ qP 6= ∅, there exists r ∈ P with pP ∩ qP = rP for all σi, σj ∈ Σ, there is at most one relation of the form σiu = σjv in R. if and only if Coming back to examples, it turns out that the presentations for Artin monoids, dis- cussed in § 3.3, are complete for yR. Also, the presentations for Baumslag-Solitar monoids B+ k,l, for k, l ≥ 1, are complete for yR. Furthermore, the presentation for the Thompson monoid F + is complete for yR. Following our discussion above, it is now easy to see that for Artin monoids, the Baumslag-Solitar monoids B+ k,l, for k, l ≥ 1, and the Thompson monoid F +, all non-empty constructible right ideals are principal. In particular, all these examples satisfy independence. For semigroups coming from rings, we have the following result: Lemma 6.33. Let R be a principal ideal domain. For both semigroups M×n (R) and Mn(R) ⋊ M×n (R), every non-empty constructible right ideal is principal. For the proof, we need the following Lemma 6.34. For every a, c in M×n (R), there exists x ∈ M×n (R) such that aMn(R) ∩ cMn(R) = xMn(R) and aM×n (R) ∩ cM×n (R) = xM×n (R). Proof. For brevity, we write M for Mn(R) and M× for M×n (R). We will use the observation that for every z ∈ M×, there exist u and v in GLn(R) such that uzv is a diagonal matrix (see for instance [Kap49]). To prove our lemma, let us first of all define x. Let c ∈ M× satisfy cc = cc = det(c) · 1n (1n is the identity matrix). Choose u and v in GLn(R) with ca = u · diag(α1, . . . , αn) · v, where diag(α1, . . . , αn) is the diagonal matrix with α1, ..., αn on the diagonal. For all 1 ≤ i ≤ n, set βi := lcm(αi, det(c)) and γi := det(c)−1βi. Then our claim is that we can choose x as x = c · u · diag(γ1, . . . , γn). In the following, we verify our SEMIGROUP C*-ALGEBRAS 57 claim: aM ∩ cM = c−1(caM ∩ (det(c) · 1n)M ) = c−1((u · diag(α1, . . . , αn) · v)M ∩ (det(c) · 1n)M ) = c−1u(diag(α1, . . . , αn)M ∩ (det(c) · 1n)M ) = c−1 · u · diag(β1, . . . , βn)M = c−1(det(c) · 1n) · u · diag(γ1, . . . , γn)M = c · u · diag(γ1, . . . , γn)M. Thus we have shown aM ∩ cM = xM . Exactly the same computation shows that aM× ∩ cM× = xM×. (cid:3) Proof of Lemma 6.33. For M×n (R), our claim is certainly a consequence of the Lemma 6.34. For Mn(R)⋊M×n (R), first note that given (b, a) and (d, c) in Mn(R)⋊ M×n (R), we have (b, a)(Mn(R) ⋊ M×n (R)) = (b + aMn(R)) × (aM×n (R)), (d, c)(Mn(R) ⋊ M×n (R)) = (d + cMn(R)) × (cM×n (R)). Moreover, the intersection is either empty or of the form (b + aMn(R)) ∩ (d + cMn(R)) y + (aMn(R) ∩ cMn(R)) for some y ∈ Mn(R). Now Lemma 6.34 provides an element x ∈ M×n (R) with aMn(R) ∩ cMn(R) = xMn(R) and aM×n (R) ∩ cM×n (R) = xM×n (R). Thus either (b, a)(Mn(R) ⋊ M×n (R)) ∩ (d, c)(Mn(R) ⋊ M×n (R)) is empty or we obtain (b, a)(Mn(R) ⋊ M×n (R)) ∩ (d, c)(Mn(R) ⋊ M×n (R)) = (y, x)(Mn(R) ⋊ M×n (R)). (cid:3) In general, however, given an integral domain R, the semigroups R× and R ⋊ R× do not have the property that all non-empty constructible right ideals are principal. For example, just take a number field with non-trivial class number, and let R be its ring of algebraic integers. The property that all non-empty constructible right ideals are principal, for R× or R ⋊ R×, translates to the property of the ring R of being a principal ideal domain. But this is not the case if the class number is bigger than 1. However, for all rings of algebraic integers, and more generally, for all Krull rings R, the semigroups R× and R ⋊ R× do satisfy independence. Let R be an integral domain. Recall that we introduced the set I(R) of constructible ideals in § 4.3. It is now easy to see that JR× =(cid:8)I× : I ∈ I(R)(cid:9) 58 and XIN LI JR⋊R× =(cid:8)(r + I) × I× : r ∈ R, a, I ∈ I(R)(cid:9) , where I× = I \ {0}. Let us make the following observation about the relationship between the indepen- dence condition for multiplicative semigroups and ax + b-semigroups: Lemma 6.35. Let R be an integral domain. Then R× satisfies independence if and only if R ⋊ R× satisfies independence. Proof. If JR⋊R× is not independent, then we have a non-trivial equation of the form (r + I) ⋊ I× = n[i=1 (ri + Ii) × I×i with (ri + Ii) × I×i ( (r + I) ⋊ I×. It is clear that (ri + Ii) × I×i ( (r + I) ⋊ I× implies that Ii ( I, for all 1 ≤ i ≤ n. Projecting onto the second coordinate of R × R×, we obtain This means that R× does not satisfy independence. I× = I×i . n[i=1 Conversely, assume that R× does not satisfy independence, so that we have a non- trivial equation of the form I× = with I×i ( I×. Hence it follows that I = I×i Ii, n[i=1 n[i=1 and Ii ( I for all 1 ≤ i ≤ n. By [Got94, Theorem 18], we may assume without loss of generality that But then we have [I : Ii] < ∞ for all 1 ≤ i ≤ n. I × I× = n[i=1 [r+Ii∈I/Ii (r + Ii) × I×i . This shows that JR⋊R× does not satisfy independence. Lemma 6.36. For a Krull ring R, both semigroups R× and R ⋊ R× satisfy inde- pendence. (cid:3) SEMIGROUP C*-ALGEBRAS 59 Proof. We use the same notations as in § 4.3. Let Q be the quotient field of R, and let I, I1, ..., In be ideals in I(R) with Ii ( I for all 1 ≤ i ≤ n. Then for every 1 ≤ i ≤ n, there exists pi ∈ P(R) with vpi (Ii) > vpi (I). By Proposition 4.13, there exists x ∈ Q× with vpi(x) = vpi(I) for all 1 ≤ i ≤ n and vp(x) ≥ vp(I) for all p ∈ P(R) \ {p1, . . . , pr} . Thus x lies in I, but does not lie in Ii for any 1 ≤ i ≤ n. Therefore, and thus Ii ( I, I×i ( I×. n[i=1 n[i=1 This shows that R× satisfies independence. By Lemma 6.35, R ⋊ R× must satisfy independence as well. (cid:3) Let us present an example of a semigroup coming from a ring which does not satisfy independence. Consider the ring R := Z[i√3]. Its quotient field is given by 2 (1 + i√3). α is a primitive Q = Q[i√3]. R is not integrally closed in Q. Let α := 1 sixth root of unity. It is clear that α /∈ R. But 2α = 1 + i√3 lies in R. The integral closure of R is given by ¯R := Z[α]. We claim that 2 ¯R = 2−1(2αR) = 2−1(1 + i√3)R. To prove "⊆", observe that ¯R = Z · 1 + Z · α. Now 2 · (2 · 1) = 4 = (1 + i√3) · (1 − i√3) ∈ (1 + i√3)R, and 2 · (2α) = 2 · (1 + i√3) ∈ (1 + i√3)R. For "⊇", let x = m + n · i√3 be in R such that 2x ∈ 2αR. As 2αR = (1+i√3)R = Z·(1+i√3)+Z·((1+i√3)i√3) = Z·(1+i√3)+Z·(−3+i√3), there exist k, l ∈ Z with 2x = 2m + 2n · i√3 = k(1 + i√3) + l(−3 + i√3) = (k − 3l) + (k + l)(i√3), It follows that 2n = 2m + 4l, and thus so that 2m = k − 3l and 2n = k + l. n = m + 2l or m = n − 2l. We conclude that x = −2l + n · (1 + i√3) ∈ 2 ¯R. This shows that 2 ¯R = 2−1(1 + i√3)R. Hence it follows that 2 ¯R is a constructible (ring-theoretic) ideal of R. 60 XIN LI We have ¯R = R ∪ αR ∪ α2R in Q. This is because R = Z + Z(2α), αR = Zα + Z(2α2) = Zα + Z(2α − 2) and α2R = Z(α − 1) + Z2. Now take x = m + nα ∈ ¯R with m, n ∈ Z. If n is even, then x is contained in R. If n is odd and m is even, then write l = m 2 . We have Finally, if n is odd and m is odd, we write k = m+n x = (n + m) · α + (−l) · (2α − 2) ∈ αR. . Then 2 x = n · (α − 1) + k · 2 ∈ α2R. This shows ¯R = R ∪ αR ∪ α2R. Therefore, 2 ¯R = 2R ∪ 2αR ∪ 2α2R = 2R ∪ (1 + i√3)R ∪ (−1 + i√3)R. But 2R ( 2 ¯R, (1 + i√3)R ( 2 ¯R and (−1 + i√3)R ( 2 ¯R. This means that R× does not satisfy independence. By Lemma 6.35, R ⋊ R× does not satisfy independence, either. Let us present another example of a left cancellative semigroup not satisfying in- dependence. Consider P = N \ {1}. Clearly, P is a semigroup under addition. We have the following constructible right ideals 2 + P = {2, 4, 5, 6, . . .} and 3 + P = {3, 5, 6, 7, . . .} . Hence 5 + N = {5, 6, 7, . . .} = (2 + P ) ∩ (3 + P ) is also a constructible right ideal of P . Moreover, it is clear that 5 + N = (5 + P ) ∪ (6 + P ). But since 5 + P ( 5 + N and 6 + P ( 5 + N, it follows that P does not satisfy independence. A similar argument shows that for every numerical semigroup of the form N \ F , where F is a non-empty finite subset of N such that N \ F is still closed under addition, the independence condition does not hold. The reader may also compare [Cun17b] for more examples of a similar kind (which are two-dimensional versions), where the independence condition typically fails. Now let us come back to the comparison of reduced C*-algebras for left cancella- tive semigroups and their left inverse hulls. Combining Corollary 6.19 and Propo- sition 6.29, we get Proposition 6.37. Let P be a subsemigroup of a group. The *-homomorphism C∗λ(Il(P )) → C∗λ(P ), λp 7→ Vp is an isomorphism if and only if P satisfies independence. SEMIGROUP C*-ALGEBRAS 61 6.6. Construction of full semigroup C*-algebras. Proposition 6.37 explains when we can identify C∗λ(Il(P )) and C∗λ(P ) in a canonical way, in case P embeds into a group. Motivated by this result, we construct full semigroup C*-algebras. Definition 6.38. Let P be a left cancellative semigroup, and Il(P ) its left inverse hull. We define the full semigroup C*-algebra of P as the full inverse semigroup C*-algebra of Il(P ), i.e., C∗(P ) := C∗(Il(P )). Recall that C∗(Il(P )) is the C*-algebra universal for *-representations of the inverse semigroup Il(P ) by partial isometries (see § 5.1). As we saw in § 5.1, there is a canonical *-homomorphism C∗(Il(P )) → C∗λ(Il(P )), vp 7→ λp. Composing with the *-homomorphism C∗λ(Il(P )) → C∗λ(P ), λp 7→ Vp, we obtain a canonical *-homomorphism We call it the left regular representation of C∗(P ). C∗(P ) → C∗λ(P ), vp 7→ Vp. Remark 6.39. It is clear that if the left regular representation of C∗(P ) is an isomorphism, then P must satisfy independence. This is because the restriction of C∗(P ) → C∗λ(P ) to C∗({vX : X ∈ JP}) is the composition C∗({vX : X ∈ JP}) → C∗(E) → Dλ(P ), and we know that the first *-homomorphism is always an isomorphism (see Corol- lary 6.24), while the second one is an isomorphism if and only if P satisfies inde- pendence (see Corollary 6.29). Given a concrete left cancellative semigroup P , it is usually possible to find a natural and simple presentation for C∗(P ) as a universal C*-algebra generated by isometries and projections, subject to relations. Let us discuss some examples. For the example P = N, the full semigroup C*-algebra C∗(N) is the universal unital C*-algebra generated by one isometry, C∗(N) ∼= C∗(v v∗v = 1). For P = N × N, C∗(N × N) is the universal unital C*-algebra generated by two isometries which *-commute, i.e., C∗(N × N) ∼= C∗(va, vb v∗ava = 1 = v∗b vb, vavb = vbva, v∗avb = vbv∗a). Note that this C*-algebra is a quotient of C∗(va, vb v∗ava = 1 = v∗b vb, vavb = vbva). 62 XIN LI As we remarked in § 6.2, the latter C*-algebra is not nuclear by [Mur96, Theo- rem 6.2]. However, as we will see in § 6.8, this quotient, and hence C∗(N × N), is nuclear. For the non-abelian free monoid on two generators P = N ∗ N, C∗(N ∗ N) is the universal unital C*-algebra generated by two isometries with orthogonal range pro- jections, i.e., C∗(N ∗ N) ∼= C∗(va, vb v∗ava = 1 = v∗b vb, vav∗avbv∗b = 0). More generally, for a right-angled Artin monoid P , a natural and simple presenta- tion for C∗(P ) has been established in [CL02] (see also [ELR16]). Let us also mention that for a class of left cancellative semigroups, full semigroup C*-algebras can be identified in a canonical way with semigroup crossed products by endomorphisms. Let P be a left cancellative semigroup with constructible right ideals JP . We then have a natural action α of P by endomorphisms on D(P ) := C∗({vX : X ∈ JP}) ⊆ C∗(P ), where p ∈ P acts by the endomorphism αp : D(P ) → D(P ), vX 7→ vpX . If P is right reversible, i.e., P p ∩ P q 6= ∅ for all p, q ∈ P , or if every non-empty constructible right ideal of P is principal, i.e., J ×P = {pP : p ∈ P}, then we have a canonical isomorphism We refer to [Li12, § 3] for more details. Writing out the definition of the crossed product, we get the following presentation: C∗(P ) ∼= D(P ) ⋊α P. {eX : X ∈ JP} ∪ {vp : p ∈ P} e∗X = eX = e2 X ; v∗pvp = 1; e∅ = 0 if ∅ ∈ JP , eP = 1, eX∩Y = eX · eY ; vpq = vpvq; vpeX v∗p = epX C∗(P ) ∼= C∗  ∼= C∗ C∗(R ⋊ R×)   In particular, for an integral domain R, we obtain the following presentation for the full semigroup C*-algebra of R ⋊ R×: {eI : I ∈ I(R)} ∪{sa : a ∈ R×} ∪(cid:8)ub : b ∈ R(cid:9) e∗I = eI = e2 I ; ub(ub)∗ = 1 = (ub)∗ub; v∗ava = 1 eR = 1, eI∩J = eI · eJ ; sac = sasc, ub+d = ubud, saub = uabsa; saeIs∗a = eaI ; ubeI = eIub if b ∈ I, eI ubeI = 0 if b /∈ I We refer to [CDL13, § 2] as well as [Li12, § 2.4]. In order to explain how this definition of full semigroup C*-algebras is related to previous constructions in the literature, we mention first of all that our definition SEMIGROUP C*-ALGEBRAS 63 generalizes Nica's construction in the quasi-lattice ordered case [Nic92]. Moreover, in the case of ax + b-semigroups over rings of algebraic integers (or more generally Dedekind domains), our definition includes the construction in [CDL13]. In the case of subsemigroups of groups, our definition coincides with the construction, denoted by C∗s (P ), in [Li12, Definition 3.2]. Last but not least, we point out that in comparison with another construction in [Li12, Definition 2.2], our definition is always a quotient of the construction in [Li12, Definition 2.2], and in certain cases (see [Li12, § 3.1] for details), our definition is actually isomorphic to the construction in [Li12, Definition 2.2]. 6.7. Crossed product and groupoid C*-algebra descriptions of reduced semigroup C*-algebras. We now specialize to the case where our semigroup P embeds into a group G. To explain the connection between amenability and nuclearity, we would like to write the reduced C*-algebra C∗λ(P ) of P as a reduced crossed product attached to a partial dynamical system, and hence as a reduced groupoid C*-algebra. Let us start with the underlying partial dynamical system. We already saw that ΩP = Spec (Dλ(P )) may be identified with the subspace of Corollary 6.28). i=1 Xi, χ(X) = 1 implies that χ(Xi) = 1 for some 1 ≤ i ≤ n (see cJP given by the characters χ with the property that for all X, X1, . . . , Xn in JP with X = Sn Moreover, we introduced the partial dynamical system G y cJP in § 5.1. It is given as follows: Every g ∈ G acts on Ug−1 =nχ ∈ cJP : χ(s−1s) = 1 for some s ∈ Il(P )× with σ(s) = go , and for χ ∈ Ug−1 , g.χ = χ(s−1 ⊔ s) where s ∈ Il(P )× is an element satisfying χ(s−1s) = 1 and σ(s) = g. We now claim: Lemma 6.40. ΩP is an G-invariant subspace of cJP . Proof. Take g ∈ G and χ ∈ Ug−1 ∩ ΩP , and suppose that s ∈ Il(P )× satisfies χ(s−1s) = 1 and σ(s) = g. We have to show that g.χ = χ(s−1 ⊔ s) lies in ΩP . Suppose that X, X1, . . . , Xn in JP satisfy X = Sn i=1 Xi. Then, identifying s−1s with dom(s), we have s−1Xs = (g−1X) ∩ dom(s) = n[i=1 (g−1Xi) ∩ dom(s) = s−1Xis. n[i=1 Hence, if g.χ(X) = 1, then χ(s−1Xs) = 1, and hence g.χ(Xi) = χ(s−1Xis) = 1 for some 1 ≤ i ≤ n. This shows that g.χ lies in ΩP . (cid:3) 64 XIN LI Hence we obtain a partial dynamical system G y ΩP by restricting G y cJP to ΩP . A moment's thought shows that this partial dynamical system coincides with the one introduced in § 5.2. If our group G were exact, then this observation, together with Corollary 5.23, would immediately imply that C∗λ(P ) ∼= C(ΩP ) ⋊r G with respect to the G-action G y ΩP . However, it turns out that we do not need exactness here. Theorem 6.41. There is a canonical isomorphism C∗λ(P ) ∼= C(ΩP ) ⋊r G deter- mined by Vp 7→ Wp. Here Wg denote the canonical partial isometries in C(ΩP )⋊rG. Proof. We work with the dual action G y Dλ(P ) as described in § 5.2. Our strategy is to describe both C∗λ(P ) and Dλ(P ) ⋊r G as reduced (cross sectional) algebras of Fell bundles, and then to identify the underlying Fell bundles. Let us start with C∗λ(P ). As in § 5.1, we think of Il(P ) as partial isometries. Recall that we defined the partial homomorphism σ : Il(P )× → G in § 5.1. Now we set Bg := span(σ−1(g)) for every g ∈ G. We want to see that (Bg)g∈G is a grading for C∗λ(P ), in the sense of [Exe97, Definition 3.1]. Conditions (i) and (ii) are obviously satisfied. For (iii), we use the faithful conditional expectation ΘP : C∗λ(P ) ։ Dλ(P ) = Be from § 6.3. Given a finite sum xg ∈ C∗λ(P ) x =Xg 0 = x∗x =Xg,h 0 = ΘP (x∗x) =Xg x∗gxh, x∗gxg. of elements xg ∈ Bg such that x = 0, we conclude that and hence Here we used that ΘPBg = 0 if g 6= e. This implies that xg = 0 for all g. Therefore, the subspaces Bg are independent. It is clear that the linear span of all the Bg is dense in C∗λ(P ). This proves (iii). If we let B be the Fell bundle given by (Bg)g∈G, then [Exe97, Proposition 3.7] implies C∗λ(P ) ∼= C∗r (B) because ΘP : C∗λ(P ) ։ Dλ(P ) = Be is a faithful conditional expectation satisfying ΘPBe = idBe and ΘPBg = 0 if g 6= e. Let us also describe Dλ(P ) ⋊r G as a reduced algebra of a Fell bundle. We denote by Wg the partial isometry in Dλ(P ) ⋊r G corresponding to g ∈ G, and we set B′g := DgWg. Recall that we defined Dg−1 = span((cid:8)V ∗V : V ∈ Il(P )×, σ(V ) = g(cid:9)) It is easy to check that (B′g)g∈G satisfy (i), (ii) and (iii) in [Exe97, in § 5.2. Definition 3.1]. Moreover, B′e = De = Dλ(P ), and it follows immediately from the construction of the reduced partial crossed product that there is a faithful conditional expectation Dλ(P ) ⋊r G ։ Dλ(P ) = B′e which is identity on B′e and SEMIGROUP C*-ALGEBRAS 65 0 on B′g for g 6= e. Hence if we let B′ be the Fell bundle given by (B′g)g∈G, then [Exe97, Proposition 3.7] implies Dλ(P ) ⋊r G ∼= C∗r (B′). To identify C∗λ(P ) and Dλ(P ) ⋊r G, it now remains to identify B with B′. We claim that the map αiViV ∗i Wg . and αiVi 7→Xi span({V : σ(V ) = g}) → span({V V ∗Wg : σ(V ) = g}),Xi is well-defined and extends to an isometric isomorphism Bg → B′g, for all g ∈ G. All we have to show is that our map is isometric. We have =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) αiαjViV ∗j(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Dλ(P ) Xi,j αiαjViV ∗i Vj V ∗j(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Dλ(P ) =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xi,j αiViV ∗i Wg(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xi (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xi αiVi(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) αiViV ∗i Wg(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) αiVi(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xi Since Vi = ViV ∗i λg and V ∗j = λg−1 VjV ∗j , we have ViV ∗j = ViV ∗i λgλg−1 Vj V ∗j = ViV ∗i VjV ∗j . (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xi 2 2 2 2 , Hence, indeed, and we are done. All in all, we have proven that C∗λ(P ) ∼= C∗r (B) ∼= C∗r (B′) ∼= Dλ(P ) ⋊r G. Our isomorphism sends Vp to VpV ∗p Wp, but a straightforward computation shows that actually, VpV ∗p Wp = Wp for all p ∈ P . Thus the isomorphism we constructed is given by Vp 7→ Wp for all p ∈ P . (cid:3) In particular, in combination with Theorem 5.21, we get an isomorphism (21) C∗λ(P ) ∼=−→ C∗r (G ⋉ ΩP ), Vp 7→ 1{p}×ΩP . Together with Remark 5.19 and Lemma 5.22, we see that we obtain a commutative diagram (22) C∗(P ) = C∗(Il(P )) C∗λ(Il(P )) C∗λ(P ) ∼= ∼= ∼= C∗(G ⋉ cJP ) C∗r (G ⋉ cJP ) / C∗r (G ⋉ ΩP ) / /     / /     / 66 XIN LI Here the upper left vertical arrow is the left regular representation of C∗(Il(P )). The lower left vertical arrow is the *-homomorphism provided by Lemma 6.11. The lower right vertical arrow is the canonical projection map; it corresponds to The upper right vertical arrow is the left regular representation of C∗(G ⋉ cJP ). the canonical map C(cJP ) ⋊r G ։ C(Ω) ⋊r G under the identification from The- orem 5.21. The first horizontal arrow is the identifications from Theorem 5.20. The second horizontal arrow is the isomorphism from Theorem 5.21. For both of these horizontal arrows, we also need Lemma 5.22. The third horizontal arrow is provided by the isomorphism (21). Now we are ready to discuss the relationship between amenability and nuclearity and thereby explain the strange phenomena mentioned at the beginning of § 6.2. 6.8. Amenability of semigroups in terms of C*-algebras. Let us start by explaining how to characterize amenability of semigroups in terms of their C*- algebras. Theorem 6.42. Let P be a cancellative semigroup, i.e., P is both left and right cancellative. Assume that P satisfies the independence condition. Then the follow- ing are equivalent: 1) P is left amenable. 2) C∗(P ) is nuclear and there is a character on C∗(P ). 3) C∗λ(P ) is nuclear and there is a character on C∗(P ). 4) The left regular representation C∗(P ) → C∗λ(P ) is an isomorphism and there is a character on C∗(P ). 5) There is a character on C∗λ(P ). By a character, we mean a unital *-homomorphism to C. For the proof, we need the following Lemma 6.43. Let P be a left cancellative semigroup. The following are equivalent: 1. There is a character on C∗(P ). 2. P is left reversible, i.e., pP ∩ qP 6= ∅ for all p, q ∈ P . 3. Il(P ) does not contain ∅ → ∅, the partial bijection which is nowhere defined. Recall that in the convention we introduced in § 5.1, if ∅ → ∅ lies in Il(P ), then we say that Il(P ) is an inverse semigroup with zero, and let ∅ → ∅ be its distinguished zero element, which we denote by 0. Proof. 1. ⇒ 2.: If χ is a character on C∗(P ), then for every p, q ∈ P , we have χ(1pP∩qP ) = χ(1pP )χ(1qP ) = χ(VpV ∗p )χ(VqV ∗q ) = χ(Vp)2 χ(Vq)2 = 1. Hence pP ∩ qP 6= ∅. SEMIGROUP C*-ALGEBRAS 67 2. ⇒ 3.: Every partial bijection in Il(P ) is a finite product of elements in {p : p ∈ P} ∪(cid:8)q−1 : q ∈ P(cid:9) . Hence, by an inductive argument, it suffices to show that if s ∈ Il(P ) is not ∅ → ∅, then for all p, q ∈ P , ps and q−1s are not ∅ → ∅. For ps, this is clear. For q−1s, choose x ∈ dom(s). Then xP ⊆ dom(s) and s(xr) = s(x)r for all r ∈ P by property (11). As P is left reversible, there exists y ∈ P with y ∈ qP ∩ s(x)P . Hence y = s(x)r = qz for some r, z ∈ P . Therefore, (q−1s)(xz) = q−1(s(xr)) = q−1(s(x)r) = q−1(qz) = z. Hence q−1s is not ∅ → ∅, as desired. 3. ⇒ 1.: Since Il(P ) does not contain ∅ → ∅, we have by definition that C∗(P ) = C∗(Il(P )) = C∗({vs : s ∈ Il(P )} vst = vsvt, vs−1 = v∗s ). Obviously, by universal property, we obtain a character C∗(P ) → C, vs → 1. (cid:3) Proof of Theorem 6.42. 1) ⇒ 2): If P is left amenable, then there exists a left invariant state µ on ℓ∞(P ) by definition. Hence, for every p ∈ P , we have µ(1pP ) = µ(1pP (p⊔)) = µ(1P ) = 1. Now, if there were p, q ∈ P with pP ∩ qP = ∅, then 1pP + 1qP would be a projection in ℓ∞(P ) with 1pP + 1qP ≤ 1P , so that 1 = µ(1P ) ≥ µ(1pP + 1qP ) = µ(1pP ) + µ(1qP ) = 1 + 1 = 2. This is a contradiction. Therefore, P must be left reversible. By Lemma 6.43, it follows that C∗(P ) has a character. In addition, by our discussion of group embeddability in § 4.1, we see that P embeds into its group G of right quotients. Moreover, as P is left amenable, G must be amenable by [Pat88, Proposition (1.27)]. Hence, statement 2) follows from Theorem 6.44 (see also Corollary 6.45). 2) ⇒ 3) is obvious. 3) ⇒ 4) follows again from Theorem 6.44. 4) ⇒ 5) is obvious. 5) ⇒ 1): We follow [Li12, § 4.2]. Let χ : C∗λ(P ) → C be a non-zero character. Viewing χ as a state, we can extend it by the theorem of Hahn-Banach to a state on L(ℓ2(P )). We then restrict the extension to ℓ∞(P ) ⊆ L(ℓ2(P )) and call this restriction µ. The point is that by construction, µC ∗ λ(P ) = χ is multiplicative, hence C∗λ(P ) is in the multiplicative domain of µ. Thus we obtain for every f ∈ ℓ∞(P ) and p ∈ P µ(f (p⊔)) = µ(V ∗p f Vp) = µ(V ∗p )µ(f )µ(Vp) = µ(Vp)∗µ(Vp)µ(f ) = µ(f ). Thus µ is a left invariant mean on ℓ∞(P ). This shows "5) ⇒ 1)". (cid:3) 68 XIN LI Theorem 6.42 tells us that for the example P = N × N discussed in § 6.2, our definition of full semigroup C*-algebras leads to a full C*-algebra C∗(N× N) which is nuclear and whose left regular representation is an isomorphism. This explains and resolves the strange phenomenon described in § 6.2. At the same time, we see why it is not a contradiction that N ∗ N is not amenable while its C*-algebra behaves like those of amenable semigroups. The point is that there is no character on C∗(N ∗ N) because N ∗ N is not left reversible. However, we still need an explanation why the semigroup C*-algebra of N ∗ N behaves like those of amenable semigroups. This leads us to our next result. 6.9. Nuclearity of semigroup C*-algebras and the connection to amenabil- ity. Theorem 6.44. Let P be a semigroup which embeds into a group G. Consider (i) C∗(P ) is nuclear. (ii) C∗λ(P ) is nuclear. (iii) G ⋉ ΩP is amenable. (iv) The left regular representation C∗(P ) → C∗λ(P ) is an isomorphism. We always have (i) ⇒ (ii) ⇔ (iii), and (iv) implies that P satisfies independence. If P satisfies independence, then we also have (iii) ⇒ (i) and (iii) ⇒ (iv). Note that the ´etale locally compact groupoid G ⋉ ΩP really only depends on P , not on the embedding P ֒→ G. This follows from Lemma 5.22. Proof. The first claim follows from the description of C∗(P ) = C∗(Il(P )) as a full groupoid C*-algebra (see Theorem 5.17), the description of C∗λ(P ) as a reduced groupoid C*-algebra (see § 6.7 and the isomorphism (21)), the commutative dia- gram (22), and Theorem 6.7. That (iv) implies that P satisfies independence was explained in Remark 6.39. The second claim follows from the observation that if P satisfies independence, then ΩP = cJP (see Corollary 6.28 and equation (19)), so that the partial dynamical systems G y ΩP and G y cJP , and hence their partial transformation groupoids coincide, and Theorem 6.7. (cid:3) Corollary 6.45. If P is a subsemigroup of an amenable group G, then statements (i), (ii) and (iii) from Theorem 6.44 hold, and (iv) holds if and only if P satisfies independence. SEMIGROUP C*-ALGEBRAS 69 Proof. This is because if G is amenable, the partial transformation groupoid G⋉ΩP is amenable by [Exe15, Theorem 20.7 and Theorem 25.10]. (cid:3) This explains the second strange phenomenon mentioned at the beginning of § 6.2, that the semigroup C*-algebra of N∗ N behaves like those of amenable semigroups. The underlying reason is that N∗ N embeds into an amenable group: Let F2 be the free group on two generators. By [Hoc69], we have an embedding N ∗ N ֒→ F2/F′′2, where F′′2 is the second commutator subgroup of F2. But F2/F′′2 is solvable, in particular amenable. Moreover, N ∗ N satisfies independence (see § 6.5). This is why statements (i) to (iv) from Theorem 6.44 are all true for the semigroup P = N ∗ N. Remark 6.46. If we modify the definition of full semigroup C*-algebras, then we can get the same results as in Theorem 6.42, Theorem 6.44 and Corollary 6.45 without having to mention the independence condition. Simply define C∗(P ) as the full groupoid C*-algebra of the restriction G(Il(P ))ΩP = {γ ∈ G(Il(P )) : r(γ), s(γ) ∈ ΩP} of the universal groupoid G(Il(P )) of Il(P ) to ΩP . This means that we would set C∗(P ) := C∗(G(Il(P ))ΩP ). Then, in Theorem 6.44, we would have (i) ⇔ (ii) ⇔ (iii), and all these statements imply (iv). Corollary 6.45 would say that statements (i) to (iv) from Theorem 6.44 hold whenever G is amenable. Moreover, Theorem 6.42 would be true without the assumption that P satisfies independence. We have chosen not to follow this route and keep the definition of full semigroup C*- algebras as full C*-algebras of left inverse hulls because the C*-algebras C∗(Il(P )) usually have a nicer presentation, i.e., a nicer and simpler description as universal C*-algebras given by generators and relations. Moreover, in the case of semigroups embeddable into groups, we know that these two definitions of full semigroup C*- algebras differ precisely by the (failure of the) independence condition. 7. Topological freeness, boundary quotients, and C*-simplicity Given a semigroup P which embeds into a group G, we have constructed a partial dynamical system G y ΩP and identified the reduced semigroup C*-algebra C∗λ(P ) with the reduced crossed product C(ΩP ) ⋊r G. Let us now present a criterion for topological freeness of G y ΩP . First recall (compare [ELQ02] and [Li16b]) that a partial dynamical system G y X is called topologically free if for every e 6= g ∈ G, is dense in Ug−1 . Here, we use the same notation as in § 5.2. (cid:8)x ∈ Ug−1 : g.x 6= x(cid:9) We first need the following observation: Let P be a monoid. For p ∈ P , let is a monoid, χpP lies in ΩP for all p ∈ P . χpP ∈ cJP be defined by χpP (X) = 1 if and only if pP ⊆ X, for X ∈ JP . Since P 70 XIN LI Lemma 7.1. The subset {χpP : p ∈ P} is dense in ΩP . Proof. Basic open sets in ΩP are of the form U (X; X1, . . . , Xn) = {χ ∈ ΩP : χ(X) = 1, χ(Xi) = 0 for all 1 ≤ i ≤ n} . empty if X =Sn may choose p ∈ X such that p /∈Sn Here X, X1, . . . , Xn are constructible ideals of P . Clearly, U (X; X1, . . . , Xn) is i=1 Xi. Thus, for a non-empty basic open set U (X; X1, . . . , Xn), we i=1 Xi, and then χpP ∈ U (X; X1, . . . , Xn). (cid:3) Theorem 7.2. Let P be a monoid with identity e which embeds into a group G. If P has trivial units P ∗ = {e}, then G y ΩP is topologically free. Proof. For p ∈ P , let χpP ∈ cJP be defined as in Lemma 7.1, i.e., χpP (X) = 1 if and only if pP ⊆ X, for X ∈ JP . Assume that g ∈ G satisfies g.χpP = χpP for some p ∈ P . This equality only makes sense if χp ∈ Ug−1, i.e., there exists s ∈ Il(P ) with σ(s) = g and χp(s−1s) = 1. The latter condition is equivalent to pP ⊆ dom(s). Then g.χpP (X) = χpP (s−1Xs) = χpP (s−1(X ∩ im (s))) = χpP (g−1(X ∩ im (s))). So for X ∈ JP , if and only if g.χpP (X) = 1 pP ⊆ g−1(X ∩ im (s)) = g−1X ∩ dom(s). But since pP ⊆ dom(s) holds, we have that g.χpP (X) = 1 if and only if pP ⊆ g−1X if and only if gpP ⊆ X. Therefore, χpP = g.χpP means that for X ∈ JP , we have pP ⊆ X if and only if gpP ⊆ X. Note that gpP = s(pP ) lies in JP . Hence, for X = pP , we obtain gpP ⊆ pP , and for X = gpP , we get pP ⊆ gpP . Hence there exist x, y ∈ P with gp = px and p = gpy. So p = gpy = pxy and gp = px = gpyx. Thus xy = yx = e. Hence x, y ∈ P ∗. Since P ∗ = {e} by assumption, we must have x = y = e, and hence gp = p. This implies g = e. In other words, for every e 6= g ∈ G, we have g.χpP 6= χpP for all p ∈ P such that χpP ∈ Ug−1. Hence it follows that (cid:8)χ ∈ Ug−1 : g.χ 6= χ(cid:9) contains (cid:8)χpP ∈ Ug−1 : p ∈ P(cid:9) , and the latter set is dense in Ug−1 as {χpP : p ∈ P} is dense in ΩP . (cid:3) Note that G y ΩP can be topologically free if P ∗ 6= {e}. For instance, partial dynamical systems attached to ax + b-semigroups over rings of algebraic integers in number fields are shown to be topologically free in [EL13]. A generalization of this result is obtained in [Li16c, Proposition 5.8]. By [ELQ02, Theorem 2.6] and because of Theorem 6.41, we obtain the following SEMIGROUP C*-ALGEBRAS 71 Corollary 7.3. Suppose that P is a monoid with trivial units which embeds into a group. Let I be an ideal of C∗λ(P ). If I ∩ Dλ(P ) = (0), then I = (0). In other words, a representation of C∗λ(P ) is faithful if and only if it is faithful on Dλ(P ). Let us now discuss boundary quotients. We start with general inverse semigroups (with or without zero). In many situations, we are not only interested in the reduced C*-algebra of an inverse semigroup, but also in its boundary quotient. This is a notion going back to Exel (see [Exe08, Exe09, Exe15, EGS12]). Let us recall the namely the character χ satisfying χ(e) = 1 for all e ∈ E. For later purposes, we make the following observation: construction. Given a semilattice E, let bEmax be the subset of bE consisting of those χ ∈ bE such that {e ∈ E : χ(e) = 1} is maximal among all characters χ ∈ bE. Note that if E is a semilattice without zero, then bEmax consists of only one element, element. Suppose that χ ∈ bEmax satisfies χ(e) = 0 for some e ∈ E×. Then there Lemma 7.4. Let E be a semilattice with zero, and let 0 be its distinguished zero exists f ∈ E× with χ(f ) = 1 and ef = 0. Proof. If every f ∈ E× with χ(f ) = 1 satisfies ef 6= 0, then we can define a filter F by defining, for every f ∈ E×, It is obvious that F is a filter, so that there exists a character χF ∈ bE with χ−1 f ∈ F if there exists f ∈ E× with χ(f ) = 1 and ef ≤ f . By construction, F = F . (cid:8)f ∈ E× : χ(f ) = 1(cid:9) ⊆(cid:8)f ∈ E× : χF (f ) = 1(cid:9) , but χF (e) = 1 while χ(e) = 0. This contradicts maximality of {f ∈ E : χ(f ) = 1} . (cid:3) We define Now let E be the semilattice of idempotents in an inverse semigroup S. As ∂bE ⊆ bE is closed, we obtain a short exact sequence ∂bE := bEmax ⊆ bE. Now there are two options. We could view I as a subset of C∗λ(S) and form the ideal hIi of C∗λ(S) generated by I. The boundary quotient in Exel's sense (see [Exe08, Exe09, Exe15, EGS12]) is given by 0 → I → C0(bE) → C0(∂bE) → 0. ∂C∗λ(S) := C∗λ(S)/ hIi . 72 XIN LI Alternatively, we could take the universal groupoid G(S) of our inverse semigroup, form its restriction to ∂bE, G(S) ∂bE :=nγ ∈ G(S) : r(γ), s(γ) ∈ ∂bEo , and form the reduced groupoid C*-algebra As the canonical homomorphism C∗r (G(S) ∂bE). contains hIi in its kernel, we obtain canonical projections C∗λ(S) ∼= C∗r (G(S)) ։ C∗r (G(S) ∂bE) C∗λ(S) ։ C∗λ(S)/ hIi ։ C∗r (G(S) ∂bE). Under an exactness assumption, the second *-homomorphism actually becomes an isomorphism, so that our two alternatives for the boundary quotient coincide. For our purposes, it is more convenient to work with C∗r (G(S) ∂bE) because it is, by its very definition, a reduced groupoid C*-algebra, so that groupoid techniques apply. Now let us assume that our inverse semigroup S admits an idempotent pure partial homomorphism σ : S× → G to a group G. In that situation, we can define the Lemma 7.5. Let S be an inverse semigroup with an idempotent pure partial ho- partial dynamical system G y bE (see § 5.2) and identify G(S) with the partial transformation groupoid G ⋉ bE (see Lemma 5.22). We have the following momorphism to a group G. Let G y bE be its partial dynamical system. Then ∂bE is G-invariant. Proof. Let us first show that for every g ∈ G, g.(Ug−1 ∩ bEmax) ⊆ Ug ∩ bEmax. Take χ ∈ bEmax with χ(s−1s) = 1 for some s ∈ S with σ(s) = g. Then g.χ(e) = χ(s−1es). Assume that g.χ /∈ bEmax. This means that there is ψ ∈ bEmax such that ψ(e) = 1 for all e ∈ E with g.χ(e) = 1, and there exists f ∈ E with ψ(f ) = 1 but χ(s−1f s) = 0. Then ψ ∈ Ug since g.χ(ss−1) = 1, which implies ψ(ss−1) = 1. Con- sider g−1.ψ given by g−1.ψ(e) = ψ(ses−1). Then for every e ∈ E, χ(e) = 1 implies χ(s−1ses−1s) = 1, hence χ(s−1(ses−1)s) = 1, so that g−1.ψ(e) = ψ(ses−1) = 1. But χ(s−1f s) = 0 and g−1.ψ(s−1f s) = ψ(ss−1f ss−1) = ψ(f ) = 1. This contra- To see that dicts χ ∈ bEmax. Hence g.(Ug−1 ∩ bEmax) ⊆ Ug ∩ bEmax. g.(Ug−1 ∩ ∂bE) ⊆ Ug ∩ ∂bE, let χ ∈ Ug−1 ∩∂bE and choose a net (χi)i in bEmax with limi χi = χ. As Ug−1 is open, we may assume that all the χi lie in Ug−1. Then g.χi ∈ bEmax, and limi g.χi = g.χ. This implies g.χ ∈ ∂bE. (cid:3) SEMIGROUP C*-ALGEBRAS 73 Corollary 7.6. In the situation of Lemma 7.5, we have canonical isomorphisms and G(S) ∂bE ∼= G ⋉ ∂bE C∗r (G(S) ∂bE) ∼= C0(∂bE) ⋊r G. Proof. The first identification follows immediately from Lemma 7.5, while the sec- ond one is a consequence of the first one and Theorem 5.21. (cid:3) Let us now specialize to the case where S is the left inverse hull of a left cancellative semigroup P . First, we observe the following: Lemma 7.7. We have ∂cJP ⊆ ΩP . Proof. Let X, X1, . . . , Xn ∈ JP satisfy X = Sn i=1 Xi. Then for χ ∈ (cJP )max, χ(Xi) = 0 implies that there exists X′i ∈ J with χ(X′i) = 1 and Xi ∩ X′i = ∅ (see Lemma 7.4). Thus if χ(Xi) = 0 for all 1 ≤ i ≤ n, then let X′i, 1 ≤ i ≤ n be as above. Then for X′ =Tn i=1 X′i, χ(X′) = 1 and X ∩ X′ = ∅. Thus χ(X) = 0. This shows (cJP )max ⊆ ΩP . As ΩP is closed, we conclude that ∂cJP ⊆ ΩP . Definition 7.8. We write ∂ΩP := ∂cJP . For simplicity, let us now restrict to semigroups which embed into groups. Definition 7.9. We call C∗r (G(Il(P )) ∂ΩP ) the boundary quotient of C∗λ(P ), and denote it by ∂C∗λ(P ). (cid:3) Note that by Corollary 7.6, given a semigroup P embedded into a group G, we have a canonical isomorphism ∂C∗λ(P ) ∼= C(∂ΩP ) ⋊r G. Let us discuss some examples. Assume that our semigroup P is cancellative, and that it is left reversible, i.e., pP ∩ qP 6= ∅ for all p, q ∈ P . This is for instance the case for positive cones in totally ordered groups. Given such a semigroup, we know because of Lemma 6.43 that JP is a semilattice without zero, so that (cJP )max degenerates to a point. Therefore, ∂ΩP degenerates to a point. Hence it follows that the boundary quotient ∂C∗λ(P ) coincides with the reduced group C*-algebra of the group of right quotients of P . For the non-abelian free monoid N ∗ N on two generators, the boundary quotient ∂C∗λ(N ∗ N) is canonically isomorphic to the Cuntz algebra O2. More generally, boundary quotients for right-angled Artin monoids are worked out and studied in [CL07]. 74 XIN LI Given an integral domain R, the boundary quotient ∂C∗λ(R ⋊ R×) of the ax + b- semigroup over R is canonically isomorphic to the ring C*-algebra Ar[R] of R (see [CL10, CL11, Li10]). It is given as follows: Consider the Hilbert space ℓ2R with canonical orthonormal basis {δx : x ∈ R}. For every a ∈ R×, define Sa(δx) := δax, and for every b ∈ R, define U b(δx) := δb+x. Then the ring C*-algebra of R is the C*-algebra generated by these two families of operators, i.e, Ar[R] := C∗((cid:8)Sa : a ∈ R×(cid:9) ∪(cid:8)U b : b ∈ R(cid:9)) ⊆ L(ℓ2R). We refer to [CL10, CL11, Li10] and also [Li13, § 8.3] for details. Let us now establish structural properties for boundary quotients. From now on, let us suppose that our semigroup P embeds into a group G. arbitrary, and choose X ∈ JP with χ(X) = 1. Choose p ∈ X and χ ∈ C. As χ(P ) = 1, so that p.χ(X) = 1 as p ∈ X implies pP ⊆ X (X is a right ideal). Set χX := p.χ. Consider the net (χX )X indexed by X ∈ J with χ(X) = 1, ordered by inclusion. Passing to a convergent subnet if necessary, we may assume that limX χX Lemma 7.10. ∂ΩP is the minimal non-empty closed G-invariant subspace of cJP . Proof. Let C ⊆ cJP be non-empty, closed and G-invariant. Let χ ∈ (cJP )max be Up−1 = cJP , we can form p.χ, and we know that p.χ ∈ C. We have p.χ(pP ) = exists. But it is clear because of χ ∈ (cJP )max that limX χX = χ. As χX ∈ C for all X, we deduce that χ ∈ C. Thus (cJP )max ⊆ C, and hence ∂ΩP ⊆ C. In particular, ∂ΩP is the minimal non-empty closed G-invariant subspace of ΩP . Another immediate consequence is (cid:3) Corollary 7.11. The transformation groupoid G ⋉ ∂ΩP is minimal. To discuss topological freeness of G y ∂ΩP , let as in [Li13, § 7.3]. Clearly, G0 =(cid:8)g ∈ G : X ∩ gP 6= ∅ 6= X ∩ g−1P for all ∅ 6= X ∈ JP(cid:9) , G0 =(cid:8)g ∈ G : pP ∩ gP 6= ∅ 6= pP ∩ g−1P for all p ∈ P(cid:9) . Furthermore, we have the following Lemma 7.12. G0 is a subgroup of G. Proof. Take g1, g2 in G0. Then for all ∅ 6= X ∈ J , we have ((g1g2)P ) ∩ X = g1((g2P ) ∩ (g−1 1 X)) ⊇ g1((g2P ) ∩ (g−1 1 X)) ∩ (g1P ) = g1((g2P ) ∩ ((g−1 1 X) ∩ P )). Now (g−1 1 X) ∩ P = g−1 1 (X ∩ (g1P )) 6= ∅. SEMIGROUP C*-ALGEBRAS 75 Thus there exists x ∈ P such that x ∈ (g−1 1 X)∩ P . Hence xP ⊆ (g−1 1 X)∩ P . Thus ∅ 6= g1((g2P ) ∩ (xP )) ⊆ ((g1g2)P ) ∩ X. (cid:3) Proposition 7.13. G y ∂ΩP is topologically free if and only if G0 y ∂ΩP is topologically free. Proof. "⇒" is clear. For "⇐", assume that G0 y ∂ΩP is topologically free, and suppose that G y ∂ΩP is not topologically free, i.e., there exists g ∈ G and U ⊆ Ug−1 ∩ ∂ΩP such that g.χ = χ for all χ ∈ U . As (cJP )max = ∂ΩP , we can find χ ∈ Ug−1 ∩ (cJP )max with g.χ = χ. For every X ∈ JP with χ(X) = 1, choose x ∈ X and ψX ∈ (cJP )max with ψX (xP ) = 1, so that ψX (X) = 1. Consider the net (ψX )X indexed by X ∈ JP with χ(X) = 1, ordered by inclusion. Passing to a convergent subnet if necessary, we may assume that limX ψX = χ. As U is open, we may assume that ψX ∈ U for all X. Then ψX (xP ) = 1 implies that ψX ∈ Ux ∩ U . Hence for sufficiently small X ∈ JP with χ(X) = 1, there exists x ∈ X such that x−1.(Ux ∩ U ) is a non-empty open subset of ∂ΩP . We conclude that (x−1gx).ψ = ψ for all ψ ∈ x−1.(Ux ∩ U ). This implies that x−1gx /∈ G0 as G0 y ∂ΩP is topologically free. So there exists p ∈ P with pP ∩ x−1gxP = ∅ or pP ∩ x−1g−1xP = ∅. Let χX ∈ (cJP )max satisfy χX (xpP ) = 1. If pP ∩ x−1gxP = ∅, then xpP ∩ gxP = ∅, so that xpP ∩ g−1xpP = ∅. Hence g.χX 6= χX if χX ∈ Ug−1 . If pP ∩ x−1g−1xP = ∅, then xpP ∩ g−1xP = ∅, so that xpP ∩ g−1xpP = ∅. Again, g.χX 6= χX if χX ∈ Ug−1. For every sufficiently small X ∈ JP with χ(X) = 1, we can find x ∈ X and χX as above. Hence we can consider the net (χX )X as above, and assume after passing to a convergent subnet that limX χX = χ. As χ ∈ U ⊆ Ug−1 ∩ ∂ΩP , it follows that χX ∈ U ⊆ Ug−1 ∩ ∂ΩP for sufficiently small X. So we obtain g.χX 6= χX , although g acts trivially on U . This is a contradiction. (cid:3) Corollary 7.14. If G0 y ∂ΩP is topologically free, then ∂C∗λ(P ) is simple. Proof. This follows from Lemma 7.10, Proposition 7.13 and [Ren80, Chapter II, Proposition 4.6]. (cid:3) We present a situation where Corollary 7.14 applies. Recall that we introduced the notion of "completeness for yR" for presentations after Lemma 6.32. Moreover, 76 XIN LI a pair P ⊆ G consisting of a monoid P embedded into a group G is called quasi- lattice ordered (see [Nic92]) if P has trivial units P ∗ = {e} and for every g ∈ G with gP ∩ P 6= ∅, we can find an element p ∈ P such that gP ∩ P = pP . Theorem 7.15. Let P = hΣ, Ri+ be a monoid given by a presentation (Σ, R) which is complete for yR, in the sense of [Deh03]. Assume that for all u ∈ Σ, there is v ∈ Σ such that there is no relation of the form u ··· = v ··· in R. Also, suppose that P embeds into a group G such that P ⊆ G is quasi-lattice ordered in the sense of [Nic92]. Then G0 = {e} and ∂C∗λ(P ) is simple. Proof. In view of Corollary 7.14, it suffices to prove G0 = {e}. Let g ∈ G0. Assume that gP ∩ P 6= P . Then g ∈ G0 implies that this intersection is not empty. Hence, we must have gP ∩ P = pP for some p ∈ P because P ⊆ G is quasi-lattice ordered. If p 6= e, then there exists u ∈ Σ with pP ⊆ uP . By assumption, there exists v ∈ Σ such that no relation in R is of the form u ··· = v ··· . Because (Σ, R) is complete for yR, we know that uP ∩ vP = ∅ (see [Deh03, Proposition 3.3]), so that gP ∩ vP = ∅. This contradicts g ∈ G0. Hence, we must have gP ∩ P = P , and similarly, g−1P ∩ P = P . These two equalities imply g ∈ P ∗. But P ∗ = {e} because P ⊆ G is quasi-lattice ordered. Thus g = e. (cid:3) Theorem 7.15 implies that for every right-angled Artin monoid A+ the property that (AΓ, A+ quotient ∂C∗λ(A+ Γ (see § 3.3) with Γ ) is graph-irreducible in the sense of [CL07], the boundary Γ ) is simple. Moreover, assume that we have a cancellative semigroup. By going over to the opposite semigroup, the left regular representation becomes the right regular rep- resentation. In this way, our discussion about C*-algebra generated by left regular representations applies to C*-algebras of right regular representations. In partic- ular, we can define boundary quotients for C*-algebras generated by right regular representations of semigroups. For instance, for the Thompson monoid F + = hx0, x1, . . . xnxk = xkxn+1 for k < ni+ , it is easy to see that Theorem 7.15 applies to the opposite monoid, so that the boundary quotient of the C*-algebra generated by the right regular representation of F + is simple. We now turn to the property of pure infiniteness. As we mentioned, the boundary quotient ∂C∗λ(N∗ N) is isomorphic to O2, a purely infinite C*-algebra. We will now see that this is not a coincidence. First of all, it is easy to see that for a partial dynamical system G y X, the transformation groupoid G ⋉ X is purely infinite in the sense of [Mat15] if and only if every compact open subset of X is (G,CO)-paradoxical in the sense of [GS14, Definition 4.3], where CO is the set of compact open subsets of X. We recall that a non-empty subset V ⊆ X is called (G,CO)-paradoxical in [GS14, Definition 4.3] if there exist V1, . . . , Vn+m ∈ O and t1, . . . , tn+m ∈ G SEMIGROUP C*-ALGEBRAS 77 such that and Vi = V = n[i=1 m[i=n+1 Vi, Vi ∈ Ut−1 i , ti.Vi ⊆ V, and ti.Vi ∩ tj.Vj = ∅ for all i 6= j. Theorem 7.16. The groupoid G ⋉ ∂ΩP is purely infinite if and only if there exist p, q ∈ P with pP ∩ qP = ∅. Proof. Obviously, if pP ∩ qP 6= ∅ for all p, q ∈ P , then ∂ΩP degenerates to a point. Let us prove the converse. Every compact open subset of cJP can be written as a disjoint union of basic open sets U = {ψ ∈ ∂ΩP : ψ(X) = 1, ψ(X1) = . . . = ψ(Xn) = 0} , for some X, X1, . . . , Xn ∈ JP . Hence it suffices to show that U is (G,CO)- paradoxical. Since (cJP )max is dense in ∂ΩP , there exists χ ∈ (cJP )max with χ ∈ U . As χ lies in (cJP )max, χ(Xi) = 0 implies that there exists Yi ∈ J with Xi ∩ Yi = ∅ and χ(Yi) = 1 (see Lemma 7.4). Let Y := X ∩ Yi. n\i=1 Certainly, Y 6= ∅ as χ(Y ) = 1. Moreover, for every ψ ∈ ∂ΩP , ψ(Y ) = 1 implies ψ ∈ U . Now choose x ∈ Y . By assumption, we can find p, q ∈ P with pP ∩ qP = ∅. For ψ ∈ ∂ΩP , xp.ψ(xpP ) = ψ(P ) = 1. Similarly, for all ψ ∈ ∂ΩP , we have xq.ψ(xqP ) = 1. Thus xp.U ⊆ xp.∂ΩP ⊆ U, xq.U ⊆ xq.∂ΩP ⊆ U and (xp.U ) ∩ (xq.U ) ⊆ (xp.∂ΩP ) ∩ (xq.∂ΩP ) = ∅ since xpP ∩ xqP = ∅. Corollary 7.17. If P is not the trivial monoid, P 6= {e}, and if G0 y ∂ΩP is topologically free, then the boundary quotient ∂C∗λ(P ) is a purely infinite simple C*-algebra. (cid:3) Proof. First of all, by Corollary 7.14, the boundary quotient is simple. Furthermore, we observe that our assumptions that P 6= {e} and that G0 acts topologically freely on ∂ΩP imply that P is not left reversible: If P were left reversible, then ∂ΩP would consist of only one point. Also, if P were left reversible, then we would have P ⊆ G0. Since every element in P obviously leaves ∂ΩP fixed, and by our assumption that P 6= {e}, we conclude that G0 cannot act topologically freely on ∂ΩP if P were left reversible. Hence Theorem 7.16 implies that the groupoid G ⋉ ∂ΩP is purely infinite. This, together with [GS14, Theorem 4.4], ∂C∗λ(P ) is purely infinite. This completes our proof. implies that the boundary quotient (cid:3) 78 XIN LI Corollary 7.18. If P is not the trivial monoid, P 6= {e}, if G ⋉ ∂ΩP is amenable, and if G0 y ∂ΩP is topologically free, then the boundary quotient ∂C∗λ(P ) is a unital UCT Kirchberg algebra. Proof. By assumption, our semigroups are countable, so that all the C*-algebras we construct are separable. Clearly, the boundary quotient ∂C∗λ(P ) is unital. Since G ⋉ ∂ΩP is amenable, the boundary quotient ∂C∗λ(P ) is nuclear and satisfies the UCT. Now our claim follows from Corollary 7.17 (cid:3) Note that this shows that [Li13, Corollary 7.23] holds without the independence and the Toeplitz condition. Let us now study simplicity of reduced semigroup C*-algebras. Let P be a semi- group which embeds into a group. If C∗λ(P ) is simple, then the groupoid G ⋉ ΩP must be minimal, as C∗λ(P ) ∼= C∗r (G ⋉ ΩP ) (see the isomorphism (21)). In partic- ular, we must have ΩP = ∂ΩP . This equality can be characterized in terms of the semigroup as follows: Lemma 7.19. Let P be a monoid. We have ΩP = ∂ΩP if and only if for every X1, . . . , Xn ∈ JP with Xi ( P for all 1 ≤ i ≤ n, there exists p ∈ P with pP ∩Xi = ∅ for all 1 ≤ i ≤ n. Proof. Let χP be the character in ΩP determined by χP (X) = 1 if and only if X = P , for all X ∈ JP . Such a character exists in cJP , and our assumption that P has an identity element ensures that χP lies in ΩP . This is because an equation of the form P = Xi n[i=1 for some Xi ∈ JP implies that Xi = P for some 1 ≤ i ≤ n, as one of the Xi must contain the identity element. First, we claim that ΩP = ∂ΩP holds if and only if χP lies in ∂ΩP . This is certainly necessary. It is also sufficient as ∂ΩP is G-invariant, and is dense in ΩP (see Lemma 7.1). {p.χP = χpP : p ∈ P} Now basic open subsets containing χP are of the form U (P ; X1, . . . , Xn) = {χ ∈ ΩP : χ(X1) = . . . = χ(Xn) = 0} , for X1, . . . , Xn ∈ JP with Xi ( P for all 1 ≤ i ≤ n. SEMIGROUP C*-ALGEBRAS 79 with χ ∈ U (P ; X1, . . . , Xn) if and only if there exists p ∈ P with pP ∩ Xi = ∅ for all 1 ≤ i ≤ n. χP ∈ ∂ΩP if and only if χP ∈ (cJP )max if and only if for all X1, . . . , Xn ∈ JP with Xi ( P for all 1 ≤ i ≤ n, there is χ ∈ (cJP )max with χ ∈ U (P ; X1, . . . , Xn). Hence it follows that our proof is complete once we show that there exists χ ∈ (cJP )max For "⇒", assume that χ ∈ (cJP )max lies in χ ∈ U (P ; X1, . . . , Xn). Then χ(Xi) = 0 for all 1 ≤ i ≤ n. But this means that there must exist Yi ∈ JP , for 1 ≤ i ≤ n, such that χ(Yi) = 1 and Xi ∩ Yi = ∅ (see Lemma 7.4). Take the intersection Y := Yi. n\i=1 As χ(Y ) = 1, Y is not empty. Therefore, we may choose some p ∈ Y . Obviously, pP ⊆ Y as Y is a right ideal. Moreover, for every 1 ≤ i ≤ n, we have Xi ∩ pP ⊆ Xi ∩ Y ⊆ Xi ∩ Yi = ∅. For "⇐", suppose that there exists p ∈ P with pP∩Xi = ∅ for all 1 ≤ i ≤ n. An easy χ(Xi) = ∅ for all 1 ≤ i ≤ n, and it follows that χ lies in U (P ; X1, . . . , Xn). application of Zorn's Lemma yields a character χ ∈ (cJP )max with χ(pP ) = 1. Hence (cid:3) Let us derive some immediate consequences. Corollary 7.20. If G y ΩP is topologically free, then C∗λ(P ) is simple if and only if for every X1, . . . , Xn ∈ JP with Xi ( P for all 1 ≤ i ≤ n, there exists p ∈ P with pP ∩ Xi = ∅ for all 1 ≤ i ≤ n. Proof. This follows immediately from Corollary 7.11 and Lemma 7.19 (see also Corollary 7.14). (cid:3) Corollary 7.21. Let P be a monoid with identity e, and suppose that P embeds into a group. Suppose that P has trivial units P ∗ = {e}. Then C∗λ(P ) is simple if and only if for every X1, . . . , Xn ∈ JP with Xi ( P for all 1 ≤ i ≤ n, there exists p ∈ P with pP ∩ Xi = ∅ for all 1 ≤ i ≤ n. Proof. This follows from Corollary 7.20 and Lemma 7.2. (cid:3) As an example, the countable free product P = ∗∞i=1N satisfies the criterion in Lemma 7.19. Moreover, we obviously have P ∗ = {e}. Hence Corollary 7.21 ap- plies, and we deduce that C∗λ(∗∞i=1N) is simple. Actually, C∗λ(∗∞i=1N) is canonically isomorphic to the Cuntz algebra O∞. 80 XIN LI 8. The Toeplitz condition So far, we were able to derive all our results about semigroup C*-algebras just using descriptions as partial crossed products. However, it turns out that when we want to compute K-theory or the primitive ideal space, we need descriptions (at least up to Morita equivalence) as ordinary crossed products, attached to globally defined dynamical systems. Let us now introduce a criterion which guarantees such descriptions as ordinary crossed products. Definition 8.1. Let P ⊆ G be a semigroup embedded into a group G. We say that P ⊆ G satisfies the Toeplitz condition (or simply that P ⊆ G is Toeplitz) if for every g ∈ G with g−1P ∩ P 6= ∅, the partial bijection lies in the inverse semigroup Il(P ). g−1P ∩ P → P ∩ gP, x 7→ gx We can also think of Il(P ) as partial isometries on ℓ2P . In this picture, we can give an equivalent characterization of the Toeplitz condition. First, using the embedding P ⊆ G, we pass to the bigger Hilbert space ℓ2G. Let 1P be the characteristic function of P , viewed as an element in ℓ∞(G). Moreover, let λ be the left regular representation of G on ℓ2G. Then P ⊆ G is Toeplitz if and only if for every g ∈ G with 1P λg1P 6= 0, we can write 1P λg1P as a finite product of isometries and their adjoints from the set {Vp : p ∈ P} ∪(cid:8)V ∗q : q ∈ P(cid:9) . Let us now explain why the reduced semigroup C*-algebra C∗λ(P ) is a full corner in an ordinary crossed product if P ⊆ G is Toeplitz. In terms of the partial dynamical system G y ΩP , this amounts to showing that if P ⊆ G is Toeplitz, then G y ΩP has an enveloping action, in the sense of [Aba03], on a locally compact Hausdorff space. This is because if P ⊆ G is Toeplitz, then g−1P ∩ P lies in the semilattice JP . Hence, for every g ∈ G, Ug−1 =(cid:8)χ ∈ ΩP : χ(g−1P ∩ P ) = 1(cid:9) , since among all s ∈ Il(P )× with σ(s) = g, g−1P ∩ P is the maximal domain. This means that for every g ∈ G, the subspace Ug−1 is clopen. Whenever this is the case, our partial dynamical system will have an enveloping action on a locally compact Hausdorff space. This follows easily from [Aba03]. In the following, we give a direct argument describing C∗λ(P ) as a full corner in an ordinary crossed product in a very explicit way. First, we introduce some notation. Fix an embedding P ⊆ G of a semigroup P into a group G. Definition 8.2. We let JP⊆G be the smallest G-invariant semilattice of subsets of G containing JP . SEMIGROUP C*-ALGEBRAS 81 Lemma 8.3. We have (23) If P ⊆ G is Toeplitz, then giP : gi ∈ G) . JP⊆G =( n\i=1 J ×P =n∅ 6= Y ∩ P : Y ∈ J ×P⊆Go . Proof. Clearly, giP : gi ∈ G) ( n\i=1 is a G-invariant semilattice of subsets of G. It remains to show that it contains JP . It certainly includes the subset P of G. Moreover, for every subset X ∈ P and all p, q ∈ P , we have p(X) = pX and q−1(X) = q−1X ∩ P . Here, pX and q−1X are products taken in G. Therefore, we see that JP⊆G is closed under left multiplication and pre-images under left multiplication. But JP may be characterized as the smallest semilattice of subsets of P containing P and closed under left multiplication and pre-images under left multiplication. Therefore, JP is contained in JP⊆G. Our argument above also shows that we always have Now let us assume that P ⊆ G is Toeplitz, and let us prove "⊇". By assumption, the partial bijection J ×P ⊆n∅ 6= Y ∩ P : Y ∈ J ×P⊆Go . g−1P ∩ P → P ∩ gP, x 7→ gx lies in Il(P ) as long as g−1P ∩ P 6= ∅. Therefore, as long as g−1P ∩ P 6= ∅, the image of this partial bijection, P ∩ gP , lies in JP . Hence it follows, because of (23), that n∅ 6= Y ∩ P : Y ∈ J ×P⊆Go (cid:3) is contained in J ×P . Definition 8.4. We define DP⊆G := C∗({1Y : Y ∈ JP⊆G}) ⊆ ℓ∞(G). Obviously, DP⊆G is G-invariant with respect to the canonical action of G on ℓ∞(G) by left multiplication. Therefore, we can form the crossed product DP⊆G ⋊r G. It is easy to see, and explained in [CEL15, § 2.5], that we can identify this crossed product DP⊆G ⋊r G with the C*-algebra C∗({1Y λg : Y ∈ JP⊆G, g ∈ G}) ⊆ L(ℓ2G) concretely represented on ℓ2G. Proposition 8.5. In the situation above, 1P is a full projection in DP⊆G ⋊r G. If P ⊆ G is Toeplitz, then (24) C∗λ(P ) = 1P (DP⊆G ⋊r G)1P . 82 XIN LI In particular, C∗λ(P ) is a full corner in DP⊆G ⋊r G. Equation (24) is meant as an identity of sub-C*-algebras of L(ℓ2G). Proof. As the linear span of elements of the form 1Y λg, Y ∈ JP⊆G, g ∈ G is dense in DP⊆G ⋊r G, it suffices to show that, for all Y ∈ JP⊆G and g ∈ G, 1Y λg ∈ (DP⊆G ⋊r G) 1P (DP⊆G ⋊r G) in order to show that 1P is a full projection. Let Then lies in Y = giP. n\i=1 1Y λg = (λg1 1P ) 1P(cid:0)λ∗g1 1Tn i=2 giP λg(cid:1) (DP⊆G ⋊r G) 1P (DP⊆G ⋊r G) . Let us prove that C∗λ(P ) = 1P (DP⊆G ⋊r G)1P if P ⊆ G is Toeplitz. First, observe that "⊆" always holds as for all p ∈ P , we have Vp = 1P λp1P . Conversely, it suffices to show that for every Y ∈ JP⊆G and g ∈ G, 1P 1Y λg1P lies in C∗λ(P ). But 1P 1Y λg1P = (1P 1Y 1P ) (1P λg1P ) , and 1P 1Y 1P lies in C∗λ(P ) as P ∩ Y lies in JP as long as it is not empty by Lemma 8.3, and 1P λg1P lies in C∗λ(P ) because P ⊆ G is Toeplitz. (cid:3) Let us discuss some examples. First, assume that P is cancellative, and right reversible, i.e., P p ∩ P q 6= ∅ for all p, q ∈ P . Then P embeds into its group G of left quotients. We have G = P −1P . We claim that P ⊆ G is Toeplitz in this case: Take g ∈ G, and write g = q−1p for some p, q ∈ P . Then the partial bijection g−1P ∩ P → P ∩ gP, x 7→ gx is the composition of q−1 : qP → P, qx 7→ x and p : P → pP, x 7→ px. This is because g−1P ∩ P = p−1qP ∩ P = p−1(qP ) ∩ P = p−1(dom(q−1)) = dom(q−1p), and for x ∈ g−1P ∩ P = dom(q−1p), we have gx = q−1px = (q−1p)(x). In particular, if P is the positive cone in a totally ordered group G, then P ⊆ G is Toeplitz. Also, the inclusion B+ n ⊆ Bn of the Braid monoid into the corresponding Braid group is Toeplitz. Furthermore, if R is an integral domain with quotient SEMIGROUP C*-ALGEBRAS 83 field Q, then for the ax + b-semigroup R ⋊ R×, we have that R ⋊ R× ⊆ Q ⋊ Q× is Toeplitz. Let us discuss a second class of examples. Suppose that we have a monoid P with identity e, and that P ⊆ G is an embedding of P into a group G. Furthermore, we assume that In this situation, we claim that P ⊆ G is Toeplitz. J ×P⊆G = {gP : g ∈ G} . If g−1P ∩ P 6= ∅, then we can find p ∈ P such that To see this, take g ∈ G. g−1P ∩ P = pP . This is because we have J ×P⊆G = {gP : g ∈ G} by assumption. Here, we used the hypothesis that P has an identity element. Therefore, we can find q ∈ P with g−1q = p. We now claim that the partial bijection g−1P ∩ P → P ∩ gP, x 7→ gx is the composition of p : P → pP, x 7→ px and q−1 : qP → P, qx 7→ x. This is because g−1P ∩ P = qP = dom(pq−1), and for x ∈ g−1P ∩ P = dom(pq−1), we have gx = pq−1x = (pq−1)(x). In particular, for every graph Γ as in § 3.3, the inclusion A+ Γ ⊆ AΓ of the right- angled Artin monoid in the corresponding right-angled Artin group is Toeplitz. For instance, the canonical embedding N ∗ N ֒→ F2 is Toeplitz. Also, the inclusion B+ k,l ⊆ Bk,l of the Baumslag-Solitar monoid into the corresponding Baumslag- Solitar group is Toeplitz, for k, l ≥ 1. Moreover, the inclusion F + ⊆ F of the Thompson monoid into the Thompson group is Toeplitz. We make the following observation, which is an immediate consequence of our preceding discussion and Lemma 8.3: Remark 8.6. Suppose that P is a monoid which is embedded into a group G. If J ×P = {pP : p ∈ P} , then P ⊆ G is Toeplitz if and only if J ×P⊆G = {gP : g ∈ G} . Let us present two examples of semigroup embeddings into groups which are not Toeplitz. In both of our examples, the semigroup will be given by the non-abelian free monoid N ∗ N on two generators. First, consider the canonical homomorphism N ∗ N → F2/F′′2. Here, F′′2 is the second commutator subgroup of the non-abelian free group F2 on two generators. By [Hoc69], this canonical homomorphism N ∗ N → F2/F′′2 is injective. We want to see that N ∗ N ֒→ F2/F′′2 is not Toeplitz. 84 XIN LI Let us denote both the canonical generators of N ∗ N and F2 by a and b. We use the notation [g, h] = ghg−1h−1 for commutators. Obviously, [(ab)−1, (ba)−1][ba, bab][(ab)−1, (ba)−1]−1[ba, bab]−1 lies in F′′2 . Thus (ba)(ab)[(ab)−1, (ba)−1][ba, bab][(ab)−1, (ba)−1]−1[ba, bab]−1(ab)−1(ba)−1 lies in F′′2 . Now set Then pq−1yx−1 p = (ab)(ba)(ba)(bab) q = (ab)(ba)(bab)(ba) x = (ba)(ab)(bab)(ba) y = (ba)(ab)(ba)(bab). = (ba)(ab)[(ab)−1, (ba)−1][ba, bab][(ab)−1, (ba)−1]−1[ba, bab]−1(ab)−1(ba)−1 lies in F′′2 . Therefore, we have pq−1 = xy−1 in F2/F′′2 . Now we consider g = pq−1. Obviously, P ∩ gP 6= ∅ as p ∈ gP . Moreover, we know that for P = N ∗ N, the non-empty constructible right ideals are given by J ×P = {pP : p ∈ P}. Hence by Remark 8.6, if N∗ N ֒→ F2/F′′2 were Toeplitz, we would have P ∩ gP = zP for some z ∈ P , as P ∩ gP must lie in J ×P . We already know that p lies in P ∩ gP . Moreover, x lies in P ∩ gP as x = gy in F2/F′′2. But the only element z ∈ P with p ∈ zP and x ∈ zP is the identity element z = e. This is because p starts with a while x starts with b. Hence, if N ∗ N ֒→ F2/F′′2 were Toeplitz, we would have P ∩ gP = P , or in other words, P ⊆ gP . In particular, the identity element e ∈ P must be of the form e = pq−1r for some r ∈ P . Hence it would follow that q = rp in F2/F′′2, and therefore in N ∗ N. But this is absurd as p 6= q while p and q have the same word length with respect to the generators a and b. All in all, this shows that N ∗ N ֒→ F2/F′′2 is not Toeplitz. Our second example is given as follows: Again, we take P = N ∗ N. But this time, we let our group be the Thompson group F := hx0, x1, . . . xnxk = xkxn+1 for k < ni . Let a and b be the canonical free generators of N∗ N. Consider the homomorphism N ∗ N → F, a 7→ x0, b 7→ x1. This is an embedding. For instance, this follows from uniqueness of the normal form in [BG84, (1.3) in § 1]. We claim that this embedding N ∗ N ֒→ F is not Toeplitz. SEMIGROUP C*-ALGEBRAS 85 To simplify notations, let us identify N∗ N with the monoid hx0, x1i+ generated by x0 and x1 in F . Consider q = x4 0x1 and p = x3 0. Set g := pq−1. Then we have p ∈ P ∩ gP . But we also have that x0x1x2 P ∩ gP because 0 lies in x3 0x1x0x1 = x4 0x2x1 = x4 0x1x3 in F, so that pq−1x3 0x1x0x1 = pq−1x4 0x1x3 = px3 = x3 0x3 = x0x1x2 0 in F. If N∗ N ֒→ F were Toeplitz, we would have that P ∩ gP is of the form zP for some z ∈ P . The argument is the same as in the previous example. But as we saw that x3 0 and x0x1x2 0 both lie in P ∩ gP , our element z can only be either the identity element e or the generator x0. If z = e, then we would have P ∩ gP = P , hence the identity e must lie in gP . This means that there exists r ∈ P with e = gr = pq−1r and therefore q = rp. But this is absurd. If z = x0, then we would have P ∩ gP = x0P , hence x0 ∈ gP . Thus there must exist an element r ∈ P with x0 = gr = pq−1r, and thus qp−1x0 = r. We conclude that r = qp−1x0 = x4 0x1x−3 0 x0 = x4 0x1x−2 0 so that But this is again absurd. x4 0x1 = rx2 0. All in all, this shows that N ∗ N ֒→ F is not Toeplitz. Looking at the preceding two examples, and comparing with our observation above that the canonical embedding N ∗ N ֒→ F2 is Toeplitz, we get the feeling that it is easier for the universal group embedding of a semigroup to satisfy the Toeplitz condition than for any other group embedding. Indeed, this is true. Let us explain the reason. We need the following equivalent formulation of the Toeplitz condition: Lemma 8.7. Let P be a semigroup, and suppose that P ⊆ G is an embedding of P into a group G. The inclusion P ⊆ G satisfies the Toeplitz condition if and only if for all p, q ∈ P , there exists a partial bijection s ∈ Il(P ) with s(q) = p and the intersection P ∩ qp−1P , taken in G, is contained in the domain dom(s). Proof. If g ∈ G satisfies g−1P ∩ P 6= ∅, then there exists p, q in P with g−1p = q, i.e., g = pq−1. This shows that P ⊆ G is Toeplitz if and only if for all p, q ∈ P , the partial bijection qp−1P ∩ P → P ∩ pq−1P, x 7→ pq−1x lies in Il(P ). But this is precisely what our condition says. (cid:3) 86 XIN LI Corollary 8.8. Suppose that we have a semigroup P with two group embeddings P ֒→ G and P ֒→ G. Furthermore, assume that there is a group homomorphism G → G such that the diagram (25) G P  ◆ ◆ ◆ ◆ ◆ ◆ ◆ ◆ ◆ ◆ ◆ ◆ &◆ G commutes. Then if P ֒→ G is Toeplitz, then the inclusion P ֒→ G must be Toeplitz as well. Proof. In our equivalent formulation of the Toeplitz condition (see Lemma 8.7), the only part which depends on the group embedding of our semigroup is the intersection P ∩ qp−1P . In our particular situation, the intersection P ∩ qp−1P taken in G is given by while the intersection P ∩ qp−1P taken in G is given by nx ∈ P : pq−1x ∈ P in Go , (cid:8)x ∈ P : pq−1x ∈ P in G(cid:9) . Because of the commutative diagram (25), the condition pq−1x ∈ P in G implies the condition pq−1x ∈ P in G. Hence the intersection P ∩qp−1P , taken in G, is con- tained in the intersection P ∩ qp−1P , taken in G, where we view both intersections as subsets of P . Our claim follows. (cid:3) As an immediate consequence, we obtain Corollary 8.9. Let P be a semigroup which embeds into a group, and assume that P ֒→ Guniv is its universal group embedding. If P ֒→ Guniv does not satisfy the Toeplitz condition, then for any other embedding P ֒→ G of our semigroup into a group G, we must have that P ֒→ G does not satisfy the Toeplitz condition either. 9. Graph products We discuss the independence condition and the Toeplitz condition for graph prod- ucts. Let Γ = (V, E) be a graph with vertices V and edges E. Assume that two vertices in V are connected by at most one edge, and no vertex is connected to itself. Hence we view E as a subset of V × V . For every v ∈ V , let Pv be a submonoid of a group Gv. We then form the graph products and P := Γv∈V Pv G := Γv∈V Gv,  / / &   SEMIGROUP C*-ALGEBRAS 87 as in § 4.2. As explained in § 4.2, we can think of P as a submonoid of G in a canonical way. Our goal is to prove that if each of the individual semigroups Pv, for all v ∈ V , satisfy the independence condition, then the graph product P also satisfies the independence condition. Similarly, if each of the pairs Pv ⊆ Gv, for all v ∈ V , are Toeplitz, then the pair P ⊆ G satisfies the Toeplitz condition as well. Along the way, we give an explicit description for the constructible right ideals of P . We use the same notation as in § 4.2. 9.1. Constructible right ideals. Let us start with some easy observations. Lemma 9.1. Let x1 . . . xs be a reduced expression for x ∈ G, with xi ∈ Gvi . Assume that v1, . . . , vj ∈ V i(x). Then for all 1 ≤ i ≤ j, x1 . . . xi−1xi+1 . . . xs is a reduced expression (for x−1 i x). Similarly, if vs−j , . . . , vs ∈ V f (x), then for all 1 ≤ i ≤ j, x1 . . . xs−i−1xs−i+1 . . . xs is a reduced expression (for xx−1 s−i). Proof. By assumption, the expressions x1 . . . xs and xix1 . . . xi−1xi+1 . . . xs are shuf- fle equivalent. In particular, the latter expression is reduced. Our first claim follows. The second assertion is proven analogously. (cid:3) Lemma 9.2. For w ∈ V , let g be an element in Gw. Then for every x ∈ G, we have gSi w(x) = Si w(gx). Proof. Let x1 . . . xs be a reduced expression for x. If w /∈ V i(x), then Lemma 4.7 implies that gx1 . . . xs is a reduced expression for gx, and our claim follows. If w ∈ V i(x), we may assume that x1 = Si w(x). If gx1 6= e, then obviously (gx1)x2 . . . xs is a reduced expression for gx, and we are done. If gx1 = e, then x2 . . . xs is a reduced expression for gx by Lemma 9.1. Clearly, w /∈ V i(gx), and our claim follows. (cid:3) Definition 9.3. Let W ⊆ V be a subset with W × W ⊆ E, i.e., for every w1, w2 in W , we have (w1, w2) ∈ E. Given constructible right ideals Xw ∈ JPw for every w ∈ W , we set Yw∈W Xw! · P :=(cid:8)x ∈ P : Si w(x) ∈ Xw for all w ∈ W(cid:9) . If for some w ∈ W , we have Xw = ∅, then we set(cid:0)Qw∈W Xw(cid:1) · P = ∅. If W = ∅, we set(cid:0)Qw∈W Xw(cid:1) · P = P . Yw∈W Xw! · P = \w∈W By construction, we clearly have (Xw · P ). 88 XIN LI Lemma 9.4. Assume that Xw = p−1 we have Xw · P = p−1 the canonical embedding Pw ⊆ P ). 1 q1 . . . p−1 1 q1 . . . p−1 n qn(Pw) for some pi, qi ∈ Pw. Then n qn(P ). Here we view pi, qi as elements of P (via 2 q2 . . . p−1 w(x) ∈ q1(Yw)(cid:9) 1 (x) ∈ Yw · P(cid:9) Proof. We proceed inductively on n. The case n = 0 is trivial. Let pi, qi be elements of Pw, for 1 ≤ i ≤ n + 1. Set Yw := p−1 n+1qn+1(Pw). We compute (q1(Yw)) · P = (cid:8)x ∈ P : Si = (cid:8)x ∈ q1P : q−1 = (cid:8)x ∈ P : x ∈ q1p−1 1 q1(Yw)) · P = (cid:8)x ∈ P : Si 1 q1(Yw)(cid:9) w(x) ∈ p−1 w(x) ∈ q1(Yw)(cid:9) = (cid:8)x ∈ P : p1Si w(p1x) ∈ q1(Yw)(cid:9) = (cid:8)x ∈ P : Si = {x ∈ q1P : x ∈ q1(Yw · P )} 2 q2 . . . p−1 n+1qn+1(P )(cid:9) . by Lemma 9.2 Finally, (p−1 = {x ∈ P : p1x ∈ (q1(Yw)) · P} = p−1 1 (q1(Yw)) · P = p−1 1 q1 . . . p−1 n+1qn+1(P ). (cid:3) Lemma 9.5. Assume that we are given p ∈ P and W , {Xw : w ∈ W} as in Def- inition 9.3. Assume that ∅ 6= p(cid:0)Qw∈W Xw(cid:1) · P 6= P . Then there exist p in P , W ⊆ V with W × W ⊆ E, Xw ∈ JPw for w ∈ W with • W 6= ∅ and ∅ 6= Xw 6= Pw for every w ∈ W , • either p = e or for all v ∈ V f (p), there exists w ∈ W with (v, w) /∈ E, such that p Yw∈W Xw! · P = pYw∈ W Xw · P. Proof. We proceed inductively on the length l(p) of p. If l(p) = 0, i.e., p = e, then for all w ∈ W , we must have Xw 6= ∅, and there must exist w ∈ W with Xw 6= Pw. Thus, we can set W := {w ∈ W : Xw 6= Pw} and Xw := Xw for w ∈ W . Now assume that l(p) > 0. Without changing p(cid:0)Qw∈W Xw(cid:1)· P , we can replace W by {w ∈ W : Xw 6= Pw}. So we may just as well assume that for every w ∈ W , we have ∅ 6= Xw 6= Pw. If for every v ∈ V f (p), there exists w ∈ W with (v, w) /∈ E, then we can just set W = W and Xw = Xw for all w ∈ W . If not, then we choose v ∈ V f (p) with (v, w) ∈ E for every w ∈ W . Let p1 . . . pr be a reduced expression Xw! · P = (p1 ··· pr−1)(prXv) · Yv6=w∈W Xw · P. Now our claim follows once we apply the induction hypothesis with p1 ··· pr−1 in place of p. (cid:3) Definition 9.6. Assume that we are in the situation of Lemma 9.5, i.e., we are given p ∈ P and W , {Xw : w ∈ W} as in Definition 9.3. Assume that Thus Then ∅ 6= p Yw∈W p Yw∈W Xw! · P 6= P. Xw! · P SEMIGROUP C*-ALGEBRAS 89 for p, with pr ∈ Pv. Set Xv := Pv if v /∈ W . Using Lemma 9.1 and Lemma 9.2, we deduce Xw! · P =(y ∈ P : y = prx for some x ∈ Yw∈W v(y) ∈ prXv(cid:9) w(y) ∈ Xw for all v 6= w ∈ W and Si Xw! · P) pr Yw∈W = (cid:8)y ∈ P : Si = (prXv) · Yv6=w∈W p Yw∈W Xw · P. is said to be in standard form if both conditions from the Lemma 9.5 are satisfied, i.e., • W 6= ∅ and ∅ 6= Xw 6= Pw for every w ∈ W , • either p = e or for all v ∈ V f (p), there exists w ∈ W with (v, w) /∈ E. Lemma 9.7. Assume that p(cid:0)Qw∈W Xw(cid:1) · P is in standard form. Given reduced expressions p1 ··· pr for p and x1 ··· xs for x ∈(cid:0)Qw∈W Xw(cid:1) · P , p1 ··· prx1 ··· xs is a reduced expression for px. In particular, if in addition p 6= e, then for every v ∈ V i(p), we have Si v(px) = Si v(p). Proposition 9.8. The non-empty constructible right ideals of P are precisely given Proof. For our first claim, the case p = e is trivial. So let us assume p 6= e. As Xw 6= Pw for all w ∈ W , we know that W ⊆ V i(x), so that V f (p) ∩ V i(x) = ∅ because reduced follows from Lemma 4.7. p(cid:0)Qw∈W Xw(cid:1) · P is in standard form. Then our assertion that p1 ··· prx1 ··· xs is by all the non-empty subsets of P of the form p(cid:0)Qw∈W Xw(cid:1) · P , with p ∈ P and Proof. First, we prove that p(cid:0)Qw∈W Xw(cid:1) · P is constructible. It certainly suffices to check that(cid:0)Qw∈W Xw(cid:1)· P is constructible. But Lemma 9.4 tells us that Xw · P W , {Xw : w ∈ W} as in Definition 9.3. (cid:3) 90 XIN LI constructible itself. Secondly, we show that every non-empty constructible right ideal is of the form is constructible for every w ∈ W . Therefore,(cid:0)Qw∈W Xw(cid:1) · P =Tw∈W (Xw · P ) is p(cid:0)Qw∈W Xw(cid:1)·P . For this purpose, let J ′ be the set of all non-empty constructible right ideals which are of the form p(cid:0)Qw∈W Xw(cid:1) · P . Clearly, P lies in J ′. Also, if ∅ 6= X ∈ J ′ and p ∈ P , then obviously pX lies in J ′. It remains to prove that for ∅ 6= X ∈ J ′ and q ∈ P , we have q−1(X) ∈ J ′ if q−1(X) 6= ∅. Since the set J ×P of non-empty constructible right ideals of P is minimal with respect to these properties, this would then show that J ×P ⊆ J ′, as desired. By induction on l(q), we may assume that q ∈ Pv, and it even suffices to consider the case q ∈ Pv \ P ∗v . For X = p(cid:0)Qw∈W Xw(cid:1)· P , we want to show that q−1(X) = ∅ or q−1(X) ∈ J ′. We distinguish between the following cases: 1.) p = e: 1.a) There exists w ∈ W with (v, w) /∈ E. Without loss of generality we may assume that Xw 6= Pw for all w ∈ W . Then for every x ∈ P , w /∈ V i(qx) since v ∈ V i(qx). Thus Si w(qx) = e /∈ Xw. Therefore, q−1 Yw∈W Xw! · P = ∅. 1.b) We have (v, w) ∈ E for all w ∈ W and v /∈ W . Then q−1 Yw∈W Xw! · P = (cid:8)x ∈ P : Si = (cid:8)x ∈ P : Si = Yw∈W w(qx) ∈ Xw for all w ∈ W(cid:9) w(x) ∈ Xw for all w ∈ W(cid:9) Xw! · P ∈ J ′. 1.c) We have (v, w) ∈ E for all w ∈ W and v ∈ W . Then q−1 Yw∈W Xw! · P = (cid:8)x ∈ P : Si w(qx) ∈ Xw for all w ∈ W(cid:9) Xw · P ∈ J ′. = (q−1(Xv)) · Yv6=w∈W Xw! · P 6= P. ∅ 6= p Yw∈W 2.) p 6= e: We can clearly assume that SEMIGROUP C*-ALGEBRAS 91 because we have already finished the case p = e, we can in addition assume that p 6= e. Without loss of generality, we may assume v ∈ V i(p), as we would otherwise Lemma 9.7 gives Si By Lemma 9.5, we may assume that p(cid:0)Qw∈W Xw(cid:1) · P is in standard form. And have q−1(cid:2)p(cid:0)Qw∈W Xw(cid:1) · P(cid:3) = ∅ or q−1(cid:2)p(cid:0)Qw∈W Xw(cid:1) · P(cid:3) = p(cid:0)Qw∈W Xw(cid:1) · P . v(p) for every x ∈ (cid:0)Qw∈W Xw(cid:1) · P . Now y lies in q−1(cid:2)p(cid:0)Qw∈W Xw(cid:1) · P(cid:3) if and only if there exists x ∈ (cid:0)Qw∈W Xw(cid:1) · P such that qy = px. Hence if there exists y ∈ q−1(cid:2)p(cid:0)Qw∈W Xw(cid:1) · P(cid:3), we must have v(px) = Si qSi by Lemma 9.2. Thus p ∈ Si v(qy) = Si v(y) = Si v(p) v(p)P ⊆ qP . This implies that v(px) = Si q−1"p Yw∈W Xw! · P# = (q−1p) Yw∈W Xw! · P ∈ J ′. 9.2. The independence condition. (cid:3) Lemma 9.9. Assume that are in standard form, with p 6= e. If Xw · P 6= P Xw! · P 6= P and ∅ 6= pYw∈ W ∅ 6= p Yw∈W Xw · P ⊆ p Yw∈W pYw∈ W Xw! · P, then p ∈ pP . Take x ∈ (cid:16)Qw∈ W Proof. First of all, let us show that p 6= e. Namely, assume the contrary, i.e., p = e. Xw(cid:17) · P . By assumption, we can find x ∈ (cid:0)Qw∈W Xw(cid:1) · P Xw(cid:17) · P v(p). Thus we have proven that every x ∈ (cid:16)Qw∈ W so that x = px. Moreover, choose v ∈ V i(p). By Lemma 9.7, it follows that Si v(x) = Si must satisfy Si must have p 6= e. v(p). But this is obviously a wrong statement. Thus we v(px) = Si v(x) = Si Now we proceed inductively on l(p). We start with the case l(p) = 1, i.e., p ∈ Pv. v(p), we can always find x ∈(cid:0)Qw∈W Xw(cid:1)· v(px) = Si v(px) = Si v(p). v(px) = Si v(p)P = pP . Xw(cid:17)·P with Si P so that px = px. By Lemma 9.7, we deduce that p = Si Therefore, p ∈ Si For x ∈(cid:16)Qw∈ W For the induction step, take v ∈ V i(p). For x ∈(cid:16)Qw∈ W v(p), again choose x ∈(cid:0)Qw∈W Xw(cid:1) · P so that px = px. Then Si Si Xw(cid:17) · P with Si v(p) = Si v(px) = v(px) = 92 XIN LI Si v(px) = Si v(p). This shows that both p and p lie in Si v(p)P . We deduce that (Si v(p)−1 p)Yw∈ W Xw · P ⊆ (Si v(p)−1p) Yw∈W Xw! · P. Since l(Si done. v(p)−1p) < l(p), we can now apply the induction hypothesis, and we are (cid:3) Lemma 9.10. As above, let then ∅ 6= Yw∈W be in standard form, this time with p 6= e (and p = e). If Xw · P 6= P Xw! · P 6= P and ∅ 6= pYw∈ W Xw · P ⊆ Yw∈W pYw∈ W Xw! · P, Xw! · P. p ∈ Yw∈W Xw(cid:17) · P with Si Proof. For x ∈(cid:16)Qw∈ W v(p), px lies in(cid:0)Qw∈W Xw(cid:1) · P in Xw. Thus p lies in(cid:0)Qw∈W Xw(cid:1) · P . Xw · P 6= P Xw! · P 6= P and ∅ 6=Yw∈ W ∅ 6= Yw∈W Xw · P ⊆ Yw∈W Yw∈ W Xw! · P if and only if W ⊆ W and Xw ⊆ Xw for every w ∈ W . by assumption. Hence, Lemma 9.7 tells us that for all w ∈ W , Si be in standard form. Then Lemma 9.11. Let v(px) = Si w(p) = Si w(px) lies (cid:3) Si Proof. The direction "⇐" is obvious. To prove the reverse direction, first assume that W * W . Choose for every w ∈ W an element x w ∈ X w. Then the prod- Xw(cid:17) · P . But for w ∈ W \ W , we have uct Q w∈ W x w obviously lies in (cid:16)Qw∈ W Xw(cid:17) · P ⊆ w(Q w∈ W x w) = e /∈ Xw as Xw 6= Pw. This contradicts (cid:16)Qw∈ W (cid:0)Qw∈W Xw(cid:1)· P . So we must have W ⊆ W . If for some w ∈ W , we have Xw * Xw, then choose xw ∈ Xw \ Xw. For all remaining w ∈ W \{w}, choose x w ∈ X w. Then the product Q w∈ W x w lies in (cid:16)Qw∈ W w(Q w∈ W x w) = xw /∈ Xw. This again contradicts(cid:16)Qw∈ W Xw(cid:17) · P . But Si Xw(cid:17) · P ⊆(cid:0)Qw∈W Xw(cid:1) · P . (cid:3) SEMIGROUP C*-ALGEBRAS 93 Proposition 9.12. If for every v ∈ V , the semigroup Pv satisfies independence, then the graph product P satisfies independence. be finitely many constructible right ideals of P in standard form. If Proof. Let be in standard form, and let X (i) Xw! · P 6= P w ! · P 6= P pi Yw∈Wi (p−1pi) Yw∈Wi pi Yw∈Wi ∅ 6= p Yw∈W ∅ 6= pi Yw∈Wi Xw! · P =[i p Yw∈W Xw! · P =[i Yw∈W Xw! · P =[i Yw∈W p′i = Yw∈W w(pi). Si X (i) w ! · P, w ! · P w ! · P. X (i) then either p = e or pi ∈ pP for all i by Lemma 9.9. Hence X (i) Therefore, we may without loss of generality assume that p = e, i.e. Let I = {i : pi 6= e} and J = {i : pi = e}. By Lemma 9.10, we have for all i ∈ I and w ∈ W that Si w(pi) ∈ Xw. We define for every i ∈ I: (26) Therefore, For each i ∈ I, we obviously have pi Yw∈Wi Yw∈W Xw! · P. w ! · P. X (i) w ! · P ⊆ piP ⊆ p′iP ⊆ Yw∈W Xw! · P =[i∈I (p′iP ) ∪[i∈J Yw∈Wi w :=(Si w(pi)Pw X (i) w if i ∈ I, w ∈ Wi, if i ∈ J, w ∈ Wi. X (i) X (i) Set Wi := W if i ∈ I, Wi := Wi for i ∈ J and Since(cid:16)Qw∈ Wi (27) X (i) w (cid:17) · P = p′iP for all i ∈ I, we obviously again have Yw∈W Xw! · P =[i w  · P.  Yw∈ Wi X (i) Moreover, X (i) w 6= Pw for all i and w ∈ Wi. This is possible since JPv is independent for every v ∈ V , so that X (i) w(i). xw ∈ Xw \ [{i : w(i)=w} Xw \ [{i : w(i)=w} X (i) w(i) 6= ∅. 94 XIN LI By Lemma 9.11, we must have X (i) all i with Wi = W , there exists w(i) ∈ W with X (i) w ∈ {w(i)}i an element w ⊆ Xw for all i and w ∈ Wi. Assume that for w(i) ( Xw(i). Choose for every X (i) w(i)(x) does not lie in X (i) For all remaining w ∈ W , just choose some xw ∈ Xw. Then x := Qw∈W xw lies in (cid:0)Qw∈W Xw(cid:1) · P , but for all i with Wi = W , Si Therefore, x does not lie in(cid:16)Qw∈ Wi w (cid:17) · P whenever i satisfies Wi = W . For i with Wi 6= W , take w ∈ Wi \ W . Then Si w (cid:17) · P . Since this contradicts (27), there must Wi 6= W , we have x /∈(cid:16)Qw∈ Wi Xw! · P = Yw∈ Wi Yw∈W exist an index i with Wi = W and X (i) that index i, we must have w = Xw for all w ∈ W . In particular, for w  · P. w(x) = e /∈ X (i) w . Thus also for i with X (i) w(i). X (i) right ideal, and we are done. If this index i lies in J, then we have proven that If this index i lies in I, then we have shown that (cid:0)Qw∈W Xw(cid:1) · P is a principal (cid:0)Qw∈W Xw(cid:1) · P coincides with one of the (constructible right) ideals on the right hand side of (26) (since pi = e for i ∈ J), and we are also done. (cid:3) 9.3. The Toeplitz condition. Definition 9.13. Let x ∈ G, and assume that x1 ··· xs is a reduced expression for x. We set S(x) := {x1, . . . , xs}. Note that this is well-defined by Theorem 4.2. Lemma 9.14. Let g, x ∈ G, v ∈ V f (g), and assume Sf Sf v (g)Si v(x) lies in S(gx). v (g)Si v(x) 6= e. Then Proof. Let g1 ··· gr be a reduced expression for g, with gr = Sf First of all, if V f (g) ∩ V i(x) = ∅, then Lemma 4.7 tells us that for every reduced expression x1 ··· xs for x, g1 ··· grx1 ··· xs is a reduced expression for gx. Hence gr = Sf v(x) lies in S(gx). v (g)Si v (g). Secondly, assume that V f (g)∩V i(x) = {v}. If x1 ··· xs is a reduced expression for x with x1 = Si is a reduced expression for gx. Again, our claim follows. v(x), then since grx1 6= e, Lemma 4.7 tells us that g1 ··· gr−1(grx1)x2 ··· xs SEMIGROUP C*-ALGEBRAS 95 Finally, it remains to treat the case ∅ 6= V f (g) ∩ V i(x) 6= {v} . We proceed inductively on l(g). The cases l(g) = 0 and l(g) = 1 are taken care of by the previous cases. As ∅ 6= V f (g) ∩ V i(x) 6= {v} , we can choose w ∈ V f (g) ∩ V i(x) with w 6= v. If v lies in V f (g) ∩ V i(x), then choose a reduced expression g1 ··· gr for g with gr−1 ∈ Gw and gr ∈ Gv, and let x1 ··· xs be a reduced expression for x with x1 ∈ Gv and x2 ∈ Gw. Then gx = g1 ··· gr−2(grx1)(gr−1x2)x3 ··· xs. Set g′ := g1 ··· gr−2gr and x′ := x1(gr−1x2)x3 ··· xs. By Lemma 9.1, we know that g1 ··· gr−2gr is a reduced expression, so that gr = v (g′). Also, x1(gr−1x2)x3 ··· xs is a reduced expression. This is clear if gr−2x2 6= e, Sf v(x′). So we again and it follows from Lemma 9.1 in case gr−2x2 = e. Thus x1 = Si have v (g′)Si Sf v(x′) = Sf v (g)Si v(x) 6= e. v(x′) Since l(g′) < l(g), induction hypothesis tells us that Sf lies in S(g′x′) = S(gx). The case v /∈ V f (g) ∩ V i(x) is treated similarly. Just set x1 = e. v(x) = Sf v (g′)Si v (g)Si (cid:3) For g ∈ G, let us denote the partial bijection g−1P ∩ P → P ∩ gP, x 7→ gx by gP . Lemma 9.15. Let g1 ··· gr be a reduced expression for g ∈ G. Then gP = (g1)P ··· (gr)P . v(x) lies in S(gx) or grSi Proof. We proceed inductively on l(g). The case l(g) = 1 is trivial. First, we show that for x ∈ P , gx ∈ P implies grx ∈ P . Let gr ∈ Gv. Then by Lemma 9.14, grSi v(x) = e. Since gx ∈ P , we conclude that in any case, we have grSi grx ∈ P . Therefore, we compute v(x) ∈ Pv. Obviously, S(grx) ⊆(cid:8)grSi v(x)(cid:9) ∪ S(x). So we obtain dom(gP ) = {x ∈ P : gx ∈ P} = {x ∈ P : gx ∈ P and grx ∈ P} = {x ∈ P : grx ∈ P and (g1 . . . gr−1)(grx) ∈ P} = dom((g1 ··· gr−1)P gP ). Hence it follows that gP = (g1 ··· gr−1)P gP . By induction hypothesis, (g1 ··· gr−1)P = (g1)P . . . (gr−1)P , and we are done. (cid:3) 96 XIN LI Lemma 9.16. For g ∈ Gv, we have g−1Pv ∩ Pv 6= ∅ if and only if g−1P ∩ P 6= ∅. Assume that this is the case, and that there are pi, qi in Gv with gPv = p−1 1 q1 . . . p−1 n qn in Il(Pv). Then in Il(P ). gP = p−1 1 q1 . . . p−1 n qn Proof. Let us start proving the first claim. Since Pv ⊆ P , the implication "⇒" is obvious. For the reverse direction, assume that g−1P ∩ P 6= ∅, i.e., there exists x ∈ P with gx ∈ P . Then obviously, Si v(gx) lies in Pv (here we used Lemma 9.2), so g−1Pv ∩ Pv 6= ∅. n pn . . . q−1 Secondly, we show g−1P ∩ P = q−1 v(x) ∈ Pv, and gSi v(x) = Si 1 p1(P ): v(gx) ∈ Pv(cid:9) g−1P ∩ P = {x ∈ P : gx ∈ P} =(cid:8)x ∈ P : Si v(x) ∈ Pv(cid:9) by Lemma 9.2 v(x) ∈ g−1Pv ∩ Pv(cid:9) 1 p1(Pv)(cid:9) v(x) ∈ q−1 1 p1(Pv)(cid:1) · P 1 q1 ··· p−1 = (cid:8)x ∈ P : gSi = (cid:8)x ∈ P : Si = (cid:8)x ∈ P : Si = (cid:0)q−1 n pn . . . q−1 n pn . . . q−1 1 p1(P ) by Lemma 9.4. dom(gP ) = dom(p−1 n pn . . . q−1 = q−1 n qn) Therefore, we have as subsets of P . Hence it follows that in Il(P ) because we have p−1 products of p−1 Proposition 9.17. If for all v ∈ V , Pv ⊆ Gv is Toeplitz, then P ⊆ G is Toeplitz. and qi as group elements in Gv and G. 1 q1 . . . p−1 n qn = g in Gv ⊆ G. Here we are taking gP = p−1 1 q1 ··· p−1 n qn (cid:3) i Proof. Let g1 ··· gr be a reduced expression for g ∈ G, with gi ∈ Gvi . Assume that g−1P ∩ P 6= ∅. By Lemma 9.15, we know that gP = (g1)P . . . (gr)P . In particular, g−1 i P ∩ P 6= ∅ for all 1 ≤ i ≤ r. By Lemma 9.16, we conclude that g−1 i Pvi ∩ Pvi 6= ∅ for all 1 ≤ i ≤ r. Since for all 1 ≤ i ≤ r, the embedding Pvi ⊆ Gvi is Toeplitz, we can find pi,j, qi,j in Pvi (for 1 ≤ j ≤ ni) with qi,ni in Il(Pvi ). i,1 qi,1 . . . p−1 i,ni = p−1 (gi)Pvi Lemma 9.16 implies that (gi)P = p−1 i,1 qi,1 . . . p−1 i,ni qi,ni in Il(P ) for all 1 ≤ i ≤ r. Thus we have, in Il(P ): gP = (g1)P . . . (gr)P =(cid:0)p−1 1,n1q1,n1(cid:1) . . .(cid:0)p−1 1,1q1,1 . . . p−1 r,1qr,1 . . . p−1 r,nr qr,nr(cid:1) . SEMIGROUP C*-ALGEBRAS 10. K-theory 97 (cid:3) Let us apply the K-theory results from [Ech17] to semigroups and their reduced semigroup C*-algebras. Let P be a semigroup which embeds into a group. Assume that P satisfies indepen- dence, and that we have an embedding P ⊆ G into a group G such that P ⊆ G is Toeplitz. Furthermore, suppose that G satisfies the Baum-Connes conjecture with coefficients. As J ×P⊆G = G.J ×P by Lemma 8.3, we can choose a set of representatives X ⊆ JP for the G-orbits G\J ×P⊆G. For every X ∈ X, let and let GX := {g ∈ G : gX = X} , ιX : C∗λ(GX ) → C∗λ(P ), λg 7→ λg1X . Here we identify C∗λ(P ) with the crossed product DP⊆G ⋊r G as in Proposition 8.5. This is possible because of our assumption that P ⊆ G is Toeplitz. Theorem 10.1. In the situation above, we have that MX∈X (ιX )∗ : MX∈X K∗(C∗λ(GX )) ∼=−→ K∗(C∗λ(P )) is an isomorphism. To see how Theorem 10.1 follows from [Ech17, Corollary 5.19], we explain how to choose Ω, I and ei, i ∈ I (in the notation of [Ech17, Corollary 5.19]). Let Ω be the spectrum of DP⊆G (DP⊆G was introduced in Definition 8.4), so that our semigroup C*-algebra is a full corner in C0(Ω) ⋊r G by Proposition 8.5. Moreover, let I be J ×P⊆G, and let eX be given by 1X for all X ∈ J ×P⊆G. Applying [Ech17, Corol- lary 5.19], with coefficient algebra A = C, to this situation yields Theorem 10.1. If, in addition, P is a monoid and we have J ×P = {pP : p ∈ P}, then we must have J ×P⊆G = {gP : g ∈ P}, so that we may choose X = {P}. Then the stabilizer group GP = P ∗ becomes the group of units in P . The theorem above then says that the *-homomorphism induces an isomorphism ι : C∗λ(P ∗) ∼=−→ C∗λ(P ), λg 7→ Vg ι∗ : K∗(C∗λ(P ∗)) ∼=−→ K∗(C∗λ(P )). In particular, if we further have that P has trivial unit group, then we obtain that the unique unital *-homomorphism C → C∗λ(P ) induces an isomorphism K∗(C) ∼=−→ K∗(C∗λ(P )). 98 XIN LI This applies to positive cones in total ordered groups, as long as the group satisfies the Baum-Connes conjecture with coefficients. It also applies to right-angled Artin monoids, to Braid monoids, to Baumslag-Solitar monoids of the type B+ k,l for k, l ≥ 1, and to the Thompson monoid. Let us also discuss the case of ax + b-semigroups over rings of algebraic integers in number fields. This case is also discussed in detail in [Cun17a]. Let K be a number field with ring of algebraic integers R. We apply our K-theory result to the semigroup P = R ⋊ R×. This semigroup embeds into the ax + b-group K ⋊ K×. All our conditions are satisfied, so that we only need to compute orbits and stabilizers. We have a canonical identification G\JR⋊R×⊆K⋊K× ∼=−→ ClK, [a × a×] 7→ [a]. Moreover, for the stabilizer group Ga×a×, we obtain Ga×a× = a ⋊ R∗. Here, R∗ is the group of multiplicative units in R. Hence, our K-theory formula reads in this case K∗(C∗λ(a ⋊ R∗)) ∼=−→ K∗(C∗λ(R ⋊ R×)). M[a]∈ClK M[a]∈C(R) There is a generalization of this formula to ax + b-semigroups over Krull rings (see [Li16c]). Let us explain this, using the notation from § 4.3. Let R be a countable Krull ring with group of multiplicative units R∗ and divisor class group C(R). Then our K-theory formula gives K∗(C∗λ(a ⋊ R∗)) ∼=−→ K∗(C∗λ(R ⋊ R×)). The reader may also consult [Cun17a]. Building on our discussion of graph products in § 4.2 and § 9, we can also present a K-theory formula for graph products. As in § 4.2 and § 9, let Γ = (V, E) be a graph with vertices V and edges E, such that two vertices in V are connected by at most one edge, and no vertex is connected to itself. So we view E as a subset of V × V . For every v ∈ V , let Pv be a submonoid of a group Gv. We then form the graph products P := Γv∈V Pv and G := Γv∈V Gv. We have a canonical embedding P ⊆ G. For every v ∈ V , choose a system Xv of representatives for the orbits Gv\J ×Pv⊆Gv which do not contain Pv. Moreover, for every non-empty subset W ⊆ V , define SEMIGROUP C*-ALGEBRAS 99 XW :=Qw∈W Xw. Combining Proposition 9.8, Proposition 9.12, Proposition 9.17 and Theorem 10.1, we obtain Theorem 10.2. Assume that for every vertex v in V , our semigroup Pv satisfies independence, and that Pv ⊆ Gv is Toeplitz. Moreover, assume that G satisfies the Baum-Connes conjecture with coefficients. Then the K-theory of the reduced C*-algebra of P is given by K∗(C∗λ(P ∗)) ⊕ M∅6=W⊆V W×W∈E M(Xw)w∈XW K∗(C∗λ(Yw∈W GXw )) ∼=−→ K∗(C∗λ(P )). Proof. We know that P satisfies independence by Proposition 9.12, and we know that P ⊆ G is Toeplitz by Proposition 9.17. Moreover, it is an immediate conse- quence of Proposition 9.8 that G\ J ×P⊆G = {P} ⊔(" Yw∈W As we get for the stabilizer groups Xw! · P# : ∅ 6= W ⊆ V, W × W ⊆ E, (Xw)w ∈ XW) . G(Qw∈W Xw)·P = Yw∈W GXw , our theorem follows from Theorem 10.1. (cid:3) Note that the graph product G satisfies the Baum-Connes conjecture with coeffi- cients if for every vertex v ∈ V , the group Gv has the Haagerup property. This is because, by [AD13], the graph product G has the Haagerup property in this case. 11. Further developments, outlook, and open questions Based on the result we presented, in particular descriptions as partial or ordinary crossed products as well as our K-theory formula, we obtain classification results for semigroup C*-algebras. For instance, the case of positive cones in countable subgroups of the real line, where these groups are equipped with the canonical total order coming from R, have been studied in [Dou72, JX88, CPPR11, Li15]. It turns out that the semigroup C*-algebra of such positive cones remembers the semigroup completely. Actually, we can replace the semigroup C*-algebra by the ideal corresponding to the boundary quotient. It turns out that also these ideals determine the positive cones completely. For right-angled Artin monoids, a complete classification result was obtained in [ELR16], building on previous work in [CL02, CL07, Iva10, LR96]. The final classifi- cation result allows us to decide which right-angled Artin monoids have isomorphic 100 XIN LI semigroup C*-algebras by looking at the underlying graphs defining our right-angled Artin monoids. The invariants of the graphs deciding the isomorphism class of the semigroup C*-algebras are explicitly given, and easy to compute in concrete exam- ples. For Baumslag-Solitar monoids, important structural results about their semigroup C*-algebras were obtained in [Spi12, Spi14]. In the case of ax + b-semigroups over rings of algebraic integers in number fields, partial classification results have been obtained in [Li14], building on previous work in [CDL13, EL13]. It turns out that for two number fields with the same number of roots of unity, if the ax + b-semigroups over their rings of algebraic integers have isomorphic semigroup C*-algebras, then our number fields must have the same zeta function. In other words, they must be arithmetically equivalent (see [Per77,SP95]). In addition to these classification results, another observation is that the canoni- cal commutative sub-C*-algebra (denoted by Dλ(P )) of our semigroup C*-algebra often provides interesting extra information. In many situations, the partial dy- namical system attached to our semigroup (embedded into a group) is topologically free, and then this canonical commutative sub-C*-algebra is a Cartan subalgebra in the sense of [Ren08]. For instance, for rings of algebraic integers in number fields, it is shown in [Li16a] that Cartan-isomorphism for two semigroup C*-algebras of the ax + b-semigroups implies that the number fields are arithmetically equivalent and have isomorphic class groups. This is a strictly stronger statement then just being arithmetically equivalent, as there are examples of number fields which are arithmetically equivalent but have difference class numbers (see [dSP94]). It would be interesting to obtain structural results for semigroup C*-algebras of the remaining examples mentioned in § 3. For instance, for more general totally ordered groups, the semigroup C*-algebras of their positive cones have not been studied and would be interesting to investi- gate. Their boundary quotients are given by the reduced group C*-algebras of our totally ordered groups. It would be interesting to study the structure of the ideals corresponding to these boundary quotients. For Artin monoids which are not right-angled, it would be interesting to find out more about their semigroup C*-algebras. For example, the case of Braid monoids would already be interesting. Here the boundary quotients are given by the re- duced group C*-algebras of Braid groups. Therefore, the semigroup C*-algebras of Braid monoids cannot be nuclear. But what about the ideals corresponding to the boundary quotients? It would also be very interesting to study the semigroup C*-algebra of the Thomp- son monoid. While the boundary quotient of the semigroup C*-algebra attached to the left regular representation is isomorphic to the reduced group C*-algebra of the Thompson group, the boundary quotient of the semigroup C*-algebra generated SEMIGROUP C*-ALGEBRAS 101 by the right regular representation is a purely infinite simple C*-algebra (see our discussion after Theorem 7.15, and also Corollary 7.17). Is it nuclear? In the case of ax + b-semigroups over rings of algebraic integers in number fields, is it possible to find a complete classification result for their semigroup C*-algebras? This means that we want to know when precisely two such ax + b-semigroups have isomorphic semigroup C*-algebras. It would be interesting to find a characterization in terms of the underlying number fields and their invariants. Finally, it seems that not much is known about semigroup C*-algebras of finitely generated abelian cancellative semigroups. However, we remark that it is not diffi- cult to see that all numerical semigroups have isomorphic semigroup C*-algebras. Moreover, subsemigroups of Z2 are discussed in [Cun17b]. Moreover, apart from the issue of classification, we would like to mention a couple of interesting further questions. Given a semigroup P which is cancellative, i.e., both left and right cancellative, we can form the semigroup C*-algebra C∗λ(P ) generated by the left regular representa- tion, and also the semigroup C*-algebra C∗ρ (P ) generated by the right regular repre- sentation. It was observed in [CEL13, Li16c] that these two types of semigroup C*- algebras are completely different. However, strangely enough, they seem to share some properties. For instance, in all the examples we know, our semigroup C*- algebras C∗λ(P ) and C∗ρ (P ) have isomorphic K-theory (see [CEL13, Li16c]). There is even an example when this is the case, where our semigroup does not satisfy independence (see [LN16]). Is this a general phenomenon? Do C∗λ(P ) and C∗ρ (P ) always have isomorphic K-theory? What other properties do C∗λ(P ) and C∗ρ (P ) have in common? For instance, what about nuclearity? Looking at Theorem 6.44, and in particular Corollary 6.45, the following task seems interesting: Find a semigroup P which embeds into a group, whose semigroup C*- algebra is nuclear, such that P does not embed into an amenable group. With our discussion of the Toeplitz condition in mind (see § 8), it would be in- teresting to find a semigroup which embeds into a group, for which the universal group embedding is not Toeplitz. Finally, we remark that it would be an interesting project to try to generalize our K-theory computations to subsemigroups of groups without using the Toeplitz condition. References [Aba03] F. Abadie, Enveloping actions and Takai duality for partial actions, J. Funct. Anal. 197 (2003), 14–67. [Aba04] , On partial actions and groupoids, Proc. Amer. Math. Soc. 132 (2004), 1037– 1047. 102 XIN LI [AD13] Y. Antol´ın and D. Dreesen, The Haagerup property is stable under graph products, preprint, arXiv:1305.6748 (2013). [ADR00] C. Anantharaman-Delaroche and J. Renault, Amenable groupoids, Monographie no. 36, L'Enseignement Math´ematique, Geneva, 2000. [BaHLR11] N. Brownlowe, A. an Huef, M. Laca, and I. Raeburn, Boundary quotients of the Toeplitz algebra of the affine semigroup over the natural numbers, Ergod. Th. Dyn. Sys. 32 (2011), 35–62. [BG84] K.S. Brown and R. Geoghegan, An infinite-dimensional torsion-free FP∞ group, Invent. Math. 77 (1984), 367–382. [Bla06] B. Blackadar, Operator algebras: Theory of C*-algebras and von Neumann algebras, Encyclopaedia of Mathematical Sciences Vol. 122, Springer, Berlin, 2006. [BLS14] N. Brownlowe, N.S. Larsen, and N. Stammeier, C*-algebras associated to right LCM semigroups, to appear in Trans. Amer. Math. Soc., arXiv:1406.5725 (2014). [BO08] N.P. Brown and N. Ozawa, C*-algebras and finite-dimensional approximations, Grad- uate Studies in Mathematics, vol. 88, Amer. Math. Soc., Providence, RI, 2008. [Bou06] N. Bourbaki, Alg`ebre commutative. Chapitres 5 `a 7. ´El´ements de math´ematique. R´eimpression inchang´ee de l'´edition originale de 1975, Springer, Berlin Heidelberg, 2006. [BRRW14] N. Brownlowe, J. Ramagge, D. Robertson, and M.F. Whittaker, Zappa-Sz´ep products of semigroups and their C*-algebras, J. Funct. Anal. 266 (2014), 3937–3967. [BS16] N. Brownlowe and N. Stammeier, The boundary quotient for algebraic dynamical systems, J. Math. Anal. Appl. 438 (2016), 772–789. [CaHR15] L.O. Clark, A. an Huef, and I. Raeburn, Phase transitions on the Toeplitz algebras of Baumslag-Solitar semigroups, preprint, arXiv:1503.04873 (2015). [CD71] L.A Coburn and R.G. Douglas, C*-algebras of operators on a half-space, Publ. Math. Inst. Hautes ´Etudes Sci. 40 (1971), 59–68. [CDL13] J. Cuntz, C. Deninger, and M. Laca, C ∗-algebras of Toeplitz type associated with algebraic number fields, Math. Ann. 355 (2013), 1383–1423. [CDSS71] L.A. Coburn, R.G. Douglas, D.G. Schaeffer, and I.M. Singer, C*-algebras of operators on a half-space II. Index theory, Publ. Math. Inst. Hautes ´Etudes Sci. 40 (1971), 69– 79. [CEL13] J. Cuntz, S. Echterhoff, and X. Li, On the K-theory of crossed products by automor- phic semigroup actions, Q. J. Math. 64 (2013), 747–784. [CEL15] , On the K-theory of the C*-algebra generated by the left regular representation of an Ore semigroup, J. Eur. Math. Soc. 17 (2015), 645–687. [CL02] J. Crisp and M. Laca, On the Toeplitz algebras of right-angled and finite-type Artin groups, J. Austral. Math. Soc. 72 (2002), 223–245. [CL07] , Boundary quotients and ideals of Toeplitz algebras of Artin groups, J. Funct. Anal. 242 (2007), 127–156. [CL10] J. Cuntz and X. Li, The Regular C ∗-algebra of an Integral Domain, Clay Mathematics Proceedings 10 (2010), 149–170. [CL11] , C ∗-algebras associated with integral domains and crossed products by actions on adele spaces, J. Noncommut. Geom. 5 (2011), 1–37. [Cob67] L.A. Coburn, The C*-algebra generated by an isometry I, Bull. Amer. Math. Soc 73 (1967), 722–726. [Cob69] , The C*-algebra generated by an isometry II, Trans. Amer. Math. Soc. 137 (1969), 211–217. [CP61] A.H. Clifford and G.B. Preston, The algebraic theory of semigroups, Mathematical Surveys, vol. I, Amer. Math. Soc., Providence, RI, 1961. [CP67] , The algebraic theory of semigroups, Mathematical Surveys, vol. II, Amer. Math. Soc., Providence, RI, 1967. [CPPR11] A.L. Carey, J. Phillips, I.F. Putnam, and A. Rennie, Families of type III KMS states on a class of C*-algebras containing On and QN , J. Funct. Anal. 260 (2011), 1637– 1681. [Cun17a] J. Cuntz, Algebraic actions and their C ∗-algebras, to appear in: K-theory for group C ∗-algebras and semigroup C ∗-algebras, Oberwolfach Seminars, Birkhauser (2017). SEMIGROUP C*-ALGEBRAS 103 [Cun17b] , Semigroup C ∗-algebras and toric varieties, to appear in: K-theory for group C ∗-algebras and semigroup C ∗-algebras, Oberwolfach Seminars, Birkhauser, arXiv:1703.07103 (2017). [Cun77] , Simple C ∗-algebras generated by isometries, Comm. Math. Phys. 57 (1977), 173–185. [CV13] J. Cuntz and A. Vershik, C ∗-algebras associated with endomorphisms and polymor- phisms of compact abelian groups, Comm. Math. Phys. 321 (2013), 157–179. [Deh03] P. Dehornoy, Complete positive group presentations, J. Alg. 268 (2003), 156–197. [DH71] R.G. Douglas and R. Howe, On the C*-algebra of Toeplitz operators on the quarter- plane, Trans. Amer. Math. Soc. 158 (1971), 203–217. [DNR14] B. Deroin, A. Navas, and C. Rivas, Groups, orders, and dynamics, preprint, arXiv:1408.5805 (2014). [Dou72] G. Douglas R, On the C*-algebra of a one-parameter semigroup of isometries, Acta Math. 128 (1972), 143–151. [dSP94] B. de Smit and R. Perlis, Zeta functions do not determine class numbers, Bull. Amer. Math. Soc. (New Series) 31 (1994), 213–215. [Ech17] S. Echterhoff, Bivariant KK-Theory and the Baum-Connes conjecure, to appear in: K-theory for group C ∗-algebras and semigroup C ∗-algebras, Oberwolfach Seminars, Birkhauser, arXiv:1703.10912 (2017). [EGS12] R. Exel, D. Gon¸calves, and C. Starling, The tiling C*-algebra viewed as a tight inverse semigroup algebra, Semigroup Forum 84 (2012), 229–240. [EL13] S. Echterhoff and M. Laca, The primitive ideal space of the C ∗-algebra of the affine semigroup of algebraic integers, Math. Proc. Cambridge Philos. Soc. 154 (2013), 119– 126. [ELQ02] R. Exel, M. Laca, and J. Quigg, Partial dynamical systems and C*-algebras generated by partial isometrics, J. Operator Th. 47 (2002), 169–186. [ELR16] S. Eilers, X. Li, and E. Ruiz, The Isomorphism Problem for Semigroup C*-algebras of Right-Angled Artin Monoids, Doc. Math. 21 (2016), 309–343. [Exe08] R. Exel, Inverse semigroups and combinatorial C*-algebras, Bull. Braz. Math. Soc. (N.S.) 39 (2008), 191–313. [Exe09] , Tight representations of semilattices and inverse semigroups, Semigroup Fo- rum 79 (2009), 159–182. [Exe15] , Partial dynamical systems, Fell bundles and Applications, preprint, arXiv:1511.04565 (2015). [Exe97] [Fos73] R.M. Fossum, The divisor class group of a Krull domain, Ergebnisse der Mathematik , Amenability for Fell bundles, J. Reine Angew. Math. 492 (1997), 41–73. und ihrer Grenzgebiete, vol. 74, Springer, Berlin Heidelberg New York, 1973. [Got94] C. Gottlieb, On finite unions of ideals and cosets, Commun. Alg. 22 (1994), 3087– 3097. [Gre90] E.R. Green, Graph products of groups, Ph.D. Thesis, 1990. [GS14] T. Giordano and A. Sierakowski, Purely infinite partial crossed products, J. Funct. Anal. 266 (2014), 5733–5764. [HM95] S. Hermiller and J. Meier, Algorithms and geometry for graph products of groups, J. Alg. 171 (1995), 230–257. [Hoc69] M. Hochster, Subsemigroups of amenable groups, Proc. Amer. Math. Soc. 21 (1969), 363–364. [Iva10] N. Ivanov, The K-theory of Toeplitz C*-algebras of right-angled Artin groups, Trans. Amer. Math. Soc. 362 (2010), 6003–6027. [JX88] R. Ji and J. Xia, On the classification of commutator ideals, J. Funct. Anal. 78 (1988), 208–232. [Kap49] I. Kaplansky, Elementary divisors and modules, Trans. Amer. Math. Soc. 66 (1949), 464–491. [Li10] X. Li, Ring C∗-algebras., Math. Ann. 348 (2010), 859–898. [Li12] , Semigroup C ∗-algebras and amenability of semigroups, J. Funct. Anal. 262 (2012), 4302–4340. [Li13] , Nuclearity of semigroup C ∗-algebras and the connection to amenability, Adv. Math. 244 (2013), 626–662. 104 XIN LI [Li14] , On K-theoretic invariants of semigroup C*-algebras attached to number fields, Adv. Math. 264 (2014), 371–395. [Li15] , A new approach to recent constructions of C*-algebras from modular index theory, J. Funct. Anal. 269 (2015), 841–864. [Li16a] , On K-theoretic invariants of semigroup C*-algebras attached to number fields, Part II, Adv. Math. 291 (2016), 1–11. [Li16b] , Partial transformation groupoids attached to graphs and semigroups, to ap- pear in Int. Math. Res. Not., arXiv:1603.09165 (2016). [Li16c] , Semigroup C*-algebras of ax + b-semigroups, Trans. Amer. Math. Soc. 368 (2016), 4417–4437. [LN16] X. Li and M.D. Norling, Independent resolutions for totally disconnected dynamical systems II: C*-algebraic case, J. Operator Th. 75 (2016), 163–193. [LR10] M. Laca and I. Raeburn, Phase transition on the Toeplitz algebra of the affine semi- group over the natural numbers, Adv. Math. 225 (2010), 643–688. [LR96] , Semigroup crossed products and the Toeplitz algebras of nonabelian groups, J. Funct. Anal. 139 (1996), 415–440. [Mat14] M. Matsumura, A characterization of amenability of group actions on C*-algebras, J. Operator Th. 72 (2014), 41–47. [Mat15] H. Matui, Topological full groups of one-sided shifts of finite type, J. Reine Angew. Math. 705 (2015), 35–84. [McC95] K. McClanahan, K-theory for partial crossed products by discrete groups, J. Funct. Anal. 130 (1995), 77–117. [MR77] R.B. Mura and A. Rhemtulla, Orderable groups, Lecture Notes in Pure and Applied Mathematics, vol. 27, Marcel Dekker, New York and Basel, 1977. [MS14] D. Milan and B. Steinberg, On inverse semigroup C*-algebras and crossed products, Groups Geom. Dyn. (2014), 485–512. [Mur96] G.J. Murphy, C*-algebras generated by commuting isometries, Rocky Mountain J. Math. 26 (1996), 237–267. [Neu99] J. Neukirch, Algebraic number theory, Die Grundlehren der mathematischen Wis- senschaften, vol. 322, Springer, Berlin, 1999. [Nic92] A. Nica, C*-algebras generated by isometries and Wiener-Hopf operators, J. Operator Th. 27 (1992), 17–52. [Nor14] M.D. Norling, Inverse semigroup C*-algebras associated with left cancellative semi- groups, Proc. Edinb. Math. Soc. (Series 2) 57 (2014), 533–564. [Par02] L. Paris, Artin monoids inject in their groups, Comment. Math. Helv. 77 (2002), 609–637. [Pat88] A.L.T. Paterson, Amenability, Mathematical Surveys and Monographs, Amer. Math. Soc., Providence, RI, 1988. [Pat99] , Groupoids, inverse semigroups, and their operator algebras, Birkhauser, Boston, 1999. [Per77] R. Perlis, On the equation ζk(s) = ζK′ (s), J. Number Th. 9 (1977), 342–360. [Ren08] J. Renault, Cartan subalgebras in C*-algebras, Irish Math. Soc. Bull. 61 (2008), 29– 63. [Ren15] [Ren80] , Topological Amenability is a Borel Property, Math. Scand. 117 (2015), 5–30. , A groupoid approach to C ∗-algebras, Lecture Notes in Math., vol. 793, Springer, Berlin, 1980. [RGS09] J.C. Rosales and P.A. Garc´ıa-S´anchez, Numerical Semigroups, Developments in Math- ematics, vol. 20, Springer, New York, 2009. [RS15] J. Renault and S. Sundar, Groupoids associated to Ore semigroup actions, J. Operator Th. 73 (2015), 491–514. [SP95] D. Stuart and R. Perlis, A new characterization of arithmetic equivalence, J. Number Th. 53 (1995), 300–308. [Spi12] J. Spielberg, C*-algebras for categories of paths associated to the Baumslag-Solitar groups, J. London Math. Soc. 86 (2012), 728–754. [Spi14] , Groupoids and C*-algebras for categories of paths, Trans. Amer. Math. Soc. 366 (2014), 5771–5819. [Sta15a] N. Stammeier, On C*-algebras of irreversible algebraic dynamical systems, J. Funct. Anal. 269 (2015), 1136–1179. SEMIGROUP C*-ALGEBRAS 105 [Sta15b] C. Starling, Boundary quotients of C*-algebras of right LCM semigroups, J. Funct. Anal. 268 (2015), 3326–3356. [Sun14] S. Sundar, C*-algebras associated to topological Ore semigroups, to appear in Munster J. Math., arXiv:1408.4242 (2014). [Wil15] R. Willett, A non-amenable groupoid whose maximal and reduced C*-algebras are the same, to appear in Munster J. Math., arXiv:1504.05615 (2015).
1909.12787
2
1909
2019-11-09T16:03:16
Edwards' condition for quasitraces on C*-algebras
[ "math.OA" ]
We prove that Cuntz semigroups of C*-algebras satisfy Edwards' condition with respect to every quasitrace. This condition is a key ingredient in the study of the realization problem of functions on the cone of quasitraces as ranks of positive elements. In the course of our investigation, we identify additional structure of the Cuntz semigroup of an arbitrary C*-algebra and of the cone of quasitraces.
math.OA
math
EDWARDS' CONDITION FOR QUASITRACES ON C*-ALGEBRAS RAMON ANTOINE, FRANCESC PERERA, LEONEL ROBERT, AND HANNES THIEL Abstract. We prove that Cuntz semigroups of C*-algebras satisfy Edwards' condition with respect to every quasitrace. This condition is a key ingredient in the study of the realization problem of functions on the cone of quasitraces as ranks of positive elements. In the course of our investigation, we identify additional structure of the Cuntz semigroup of an arbitrary C*-algebra and of the cone of quasitraces. 1. Introduction The rank of a positive element a in a C*-algebra A with respect to a trace τ (or, more generally, a quasitrace) is defined as dτ (a) = limn τ (a1/n). In case of a trace, this rank is nothing but the value of the support projection of a in A∗∗ under the canonical extension of τ to a normal trace on A∗∗; see [ORT11]. If A is unital and stably finite, then the set QT1(A) of normalized quasitraces is a nonempty Choquet simplex. Given an extreme quasitrace τ in QT1(A), it was shown in [Thi17, Theorem 4.7] that for any two positive elements a and b in A, the minimum of the ranks of a and b with respect to τ can be approximated by the ranks of positive elements c that are dominated by a and b in the sense of Cuntz: min(cid:8)dτ (a), dτ (b)(cid:9) = sup(cid:8)dτ (c) : c - a, b(cid:9). This property was termed Edwards' condition for τ by the fourth named author due to its relation with the work in [Edw69]. This paper concerns the extension of Edwards' condition to all quasitraces (not necessarily extremal) defined on a general (not necessarily unital) C*-algebra. Edwards' condition for extremal, normalized quasitraces was a crucial ingredient in [Thi17] for the solution of the rank problem for unital, simple C*-algebras of stable rank one. In the same spirit, the general Edwards' condition as developed in this paper is a crucial ingredient in [APRT18] for the solution of the rank problem for general C*-algebras of stable rank one. The rank problem for a C*-algebra A is to determine which functions on the topological cone QT(A) of quasitraces on A arise as the ranks of positive operators in A. Here, the rank of a in A+ is the function that associates to each quasitrace τ the rank of a with respect to τ ; see [DT10], [Thi17], [APRT18]. The rank problem for A is closely connected to the question of whether the set of ranks of elements in A+ is closed under infima, that is, if f, g : QT(A) → [0, ∞] are realized as the ranks of positive elements in A, is the same true for f ∧ g? Loosely speaking, Edward's condition is the requirement that this can at least be done pointwise, that is, the Date: November 12, 2019. The two first named authors were partially supported by MINECO (grant No. MTM2017- 83487-P), and by the Comissionat per Universitats i Recerca de la Generalitat de Catalunya (grant No. 2017-SGR-1725). The fourth named author was partially supported by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) under the SFB 878 (Groups, Ge- ometry & Actions) and under Germany's Excellence Strategy EXC 2044-390685587 (Mathematics Munster: Dynamics-Geometry-Structure). 1 2 RAMON ANTOINE, FRANCESC PERERA, LEONEL ROBERT, AND HANNES THIEL infimum of the ranks of two positive elements a and b can be pointwise approximated by the ranks of elements dominated by a and b; see Definition 4.1. The Cuntz semigroup of a C*-algebra as introduced in [CEI08] satisfies a series of properties (see below for details) denoted (O1)-(O6). In Section 2, we show that Cuntz semigroups satisfy a new property, which we call (O7); see Definition 2.1 and Proposition 2.2. This property allows us to deal naturally with ideals in the semigroup. In particular, it allows us to obtain infima of elements in the Cuntz semigroup under the assumption that one of the elements is idempotent. Note that idempotents in the Cuntz semigroup of a separable C*-algebra are in natural correspondence with the lattice of closed, two-sided ideals. Given a C*-algebra A, the Cuntz semigroup Cu(A) appears naturally in the study of quasitraces on A. Building on results from [Cun78] and [BH82], it was shown in [ERS11] that the cone QT(A) of [0, ∞]-valued 2-quasitraces on A is homeomorphic to the cone F (Cu(A)) of functionals on the Cuntz semigroup of A. Therefore, ranks are naturally viewed as elements in the dual of this cone (or in the second dual of the semigroup), and hence their properties may be obtained from the study of structural properties of both Cu(A) and its cone of functionals. This will be done, respectively, in Sections 2 and 3. In Section 3 we study compact cones and their duals. We apply our results to cones of functionals of Cuntz semigroups, continuing the work in [ERS11] and [Rob13]. In particular, we prove that the cone of quasitraces on a C*-algebra A satisfies Riesz refinement, Proposition 3.3, a result which is significant towards establishing Edwards' condition for Cuntz semigroups of C*-algebras. Section 4 exclusively concerns Edwards' condition in an abstract setting. As noted above, ranks arise as elements in the dual of a suitable monoid satisfying Riesz refinement. We use the Riesz-Kantorovich type description of infima in this setting (see (3.2)) to generalize Edwards's condition for arbitrary functionals. The main result of the section states that, for semigroups satisfying (O7), one can verify Edwards' condition on functionals taking finite values. Finally, in Section 5 we show that Cuntz semigroups of C∗-algebras satisfy Ed- wards' condition; see Theorem 5.3. Our method of proof follows the line of attack developed by the fourth author in [Thi17] combined with the results obtained in the previous sections. 2. Properties of Cuntz semigroups Let A be a C*-algebra. In [Cun78], Cuntz introduced the following relations for positive elements a, b ∈ A: a - b if there is a sequence (xn)n in A such that limn ka − xnbx∗ nk = 0; a ∼ b provided that a - b and b - a. The Cuntz semigroup of A is defined as Cu(A) := (A ⊗ K)+/ ∼, where K denotes the algebra of compact operators on ℓ2(N). We denote the class of a positive element a ∈ A ⊗ K by [a]. Then Cu(A) becomes an ordered abelian semigroup with order induced by the subequivalence - and addition induced by [a] + [b] = [( a 0 0 b )]. 2.1. Properties (O1)-(O6). The Cuntz semigroup of a C*-algebra is known to satisfy a number of order properties, which we now briefly recall. The reader is referred to [CEI08] and [APT18] for background on the Cuntz semigroups of C*- algebras and their abstract counterparts, Cu-semigroups. Let S be an ordered set such that every increasing sequence has a supremum. Given x, y ∈ S we say that x is way-below y if whenever y ≤ supn yn for some increasing sequence (yn)n, then there exists n0 ∈ N such that x ≤ yn0 . We denote this relation by x ≪ y. Suppose now that S is a positively ordered monoid, that is, an ordered monoid such that 0 ≤ x for all x ∈ S. Consider the following properties on S: EDWARDS' CONDITION FOR QUASITRACES ON C*-ALGEBRAS 3 (O1) Every increasing sequence in S has a supremum. (O2) For each x ∈ S there exists an ≪-increasing sequence (xn)n such that x = supn xn. (O3) If x1 ≪ y1 and x2 ≪ y2 then x1 + x2 ≪ y1 + y2. (O4) If (xn)n and (yn)n are increasing sequences then supn(xn + yn) = supn xn + supn yn By a Cu-semigroup we understand a positively ordered monoid satisfying (O1)- (O4). A map between Cu-semigroups is called a Cu-morphism if it is a monoid homomorphism that preserves order, suprema of increasing sequences, and the way- below relation. It was shown in [CEI08] that the Cuntz semigroup of a C*-algebra is a Cu-semigroup, and that a ∗-homomorphism A → B naturally induces a Cu- morphism Cu(A) → Cu(B). The Cuntz semigroup of a C*-algebra also satisfies the following two properties: (O5) For all x′ ≪ x ≤ y and w′ ≪ w such that x + w ≤ y there exists z such that x′ + z ≤ y ≤ x + z and w′ ≪ z. (O6) For all x′ ≪ x ≤ y + z there exist y′, z′ such that x′ ≤ y′ + z′, y′ ≤ x, y and z′ ≤ x, z. That Cuntz semigroups of C*-algebras satisfy (O5) was proved in [APT18, Propo- sition 4.6, p.34]. We will often use a weaker version of (O5) that first appeared in [RW10]: For all x′ ≪ x ≤ y there exists z such that x′ + z ≤ y ≤ x + z. It was shown in [Rob13] that Cuntz semigroups of C*-algebras satisfy (O6). 2.2. Property (O7). We identify a new property that Cuntz semigroups of C*- algebras satisfy. 2.1. Definition. A Cu-semigroup S is said to satisfy (O7) if for all x′ S satisfying 1, x1, x′ 2, x2, w ∈ x′ 1 ≪ x1 ≤ w and x′ 2 ≪ x2 ≤ w, there exists x ∈ S such that x′ 1, x′ 2 ≪ x ≤ w, x1 + x2. 2.2. Proposition. The Cuntz semigroup of every C*-algebra satisfies (O7). Proof. Let A be a C*-algebra. Let xi, x′ i ≪ xi ≤ w for i = 1, 2. Choose positive elements b1, b2, a ∈ A ⊗ K such that x1 = [b1], x2 = [b2], and w = [a]. i, w ∈ Cu(A) satisfy x′ Since x′ 1 ≪ [b1] and x′ 2 ≪ [b2], we may choose ǫ > 0 such that x′ i ≤ [(bi − ǫ)+] for i = 1, 2. Since [bi] ≤ [a], there are positive elements c1, c2 ∈ a(A ⊗ K)a such that (bi − ǫ)+ ∼ ci for i = 1, 2. Set x = [c1 + c2]. Then x′ i ≪ [(bi − ǫ)+] = [ci] ≤ [c1 + c2] = x for i = 1, 2. Also, since c1 + c2 belongs to a(A ⊗ K)a, we have x ≤ [a] = w. Using at the first step that the Cuntz class of the sum of two positive elements is always dominated by the sum of their Cuntz classes ([APT11, Lemma 2.10]), we obtain x ≤ [c1] + [c2] = [(b1 − ǫ)+] + [(b2 − ǫ)+] ≤ x1 + x2, as desired. (cid:3) An ideal of a Cu-semigroup is a downward hereditary subsemigroup closed under suprema of increasing sequences. (See [APT18, Section 5.1, p.37ff] for more details.) The relevance of (O7) when dealing with ideals of a Cu-semigroup is demonstrated in the following result. 2.3. Proposition. Let S be a Cu-semigroup satisfying (O7), let w ∈ S, and let J ⊆ S be an ideal. Then the set {x ∈ S : x ∈ J, x ≤ w} is upward directed. 4 RAMON ANTOINE, FRANCESC PERERA, LEONEL ROBERT, AND HANNES THIEL Proof. Notice that {x ∈ S : x ≤ w, x ∈ J} is a downward hereditary subset closed under suprema of increasing sequences. Thus, by [APRT18, Lemma 3.2], it suffices to show that (cid:8)x′ ∈ S : there exists x ∈ J such that x′ ≪ x, x ≤ w(cid:9) is upward directed. To this end, suppose that x′, y′ ∈ S satisfy that x′ ≪ x, y′ ≪ y for some elements x, y ∈ J such that x, y ≤ w. We deduce by (O7) that there exists z ∈ S such that x′, y′ ≪ z ≤ w, x + y. Since z ≤ x + y, and since J is an ideal, we have z ∈ J. Choose z′ ∈ S with z′ ≪ z, and such that x′ ≤ z′ and y′ ≤ z′. Then z′ is in the set displayed above and, being an upper bound for both x′ and y′, this shows that this set is upward directed, as desired. (cid:3) A Cu-semigroup is called countably based if it contains a countable subset such that every element is the supremum of a ≪-increasing sequence with terms in the said countable subset. It is a standard result that in a countably based Cu-semi- group every directed subset admits a supremum; see [APT18, Remarks 3.1.3, p.21f]. Cuntz semigroups of separable C*-algebras are countably based (see, for example, [APS11, Lemma 1.3]). Let J be an ideal of a countably based Cu-semigroup S. Since ideals of Cu-semi- groups are upward directed, J has a largest element wJ := sup J (see also [APT18, Paragraph 5.1.6, p.39f]). Further, this element is idempotent, that is, 2wJ = wJ . Conversely, given an idempotent w ∈ S, the order ideal generated by w is an ideal of S with supremum w. In light of this correspondence, Proposition 2.3 immediately implies the following result. 2.4. Theorem. Let S be a countably based Cu-semigroup satisfying (O7). Then each x ∈ S and each idempotent element w ∈ S have an infimum x ∧ w in S. Moreover, these infima with idempotent elements are well behaved as the follow- ing results illustrates. 2.5. Theorem. Let S be a countably based Cu-semigroup satisfying (O5)-(O7). Let w ∈ S be an idempotent element. Then the following are satisfied: (i) The map S → S given by x 7→ x ∧ w is a monoid homomorphism preserving the order and the suprema of increasing sequences. (ii) Given x, y ∈ S, we have x ≤ y + w if and only if x + (y ∧ w) ≤ y + (x ∧ w). (iii) We have x ∧ w1 + x ∧ w2 = x ∧ (w1 ∧ w2) + x ∧ (w1 + w2) for all x ∈ S and idempotents w1, w2 ∈ S. Proof. (i): Define w : S → S by w(x) := x ∧ w. It is obvious that w is order preserving. To prove additivity, let x, y ∈ S. Since x ∧ w + y ∧ w ≤ x + y and x ∧ w + y ∧ w ≤ 2w = w, we have w(x) + w(y) = x ∧ w + y ∧ w ≤ (x + y) ∧ w = w(x + y). To show the converse inequality, set z = (x + y) ∧ w, and let z′ ≪ z. Apply (O6) for z′ ≪ z ≤ x + y to obtain x′, y′ ∈ S satisfying z′ ≤ x′ + y′, x′ ≤ x, z, and y′ ≤ y, z. Since x′ ≤ z ≤ w and x′ ≤ x, we have x′ ≤ x ∧ w. Analogously, we deduce that y′ ≤ y ∧ w. Hence, z′ ≤ x ∧ w + y ∧ w. Since this holds for all z′ ≪ z, we obtain w(x + y) = z ≤ x ∧ w + y ∧ w = w(x) + w(y). EDWARDS' CONDITION FOR QUASITRACES ON C*-ALGEBRAS 5 Finally, let us show that w preserves sequential suprema. Let (xn)∞ n=1 be an increasing sequence in S. The inequality sup n w(xn) = sup n (xn ∧ w) ≤ (sup n xn) ∧ w = w(sup n xn) is clear. Set z = (supn xn) ∧ w and let z′ ≪ z. Since z ≤ supn xn, there exists n such that z′ ≤ xn. Also, z′ ≤ z ≤ w. Therefore z′ ≤ xn ∧ w ≤ sup n (xn ∧ w). Since this holds for all z′ ≪ z, we obtain w(sup n xn) = z ≤ sup n (xn ∧ w) = sup w(xn). (ii): Let x, y ∈ S. If x + (y ∧ w) ≤ y + (x ∧ w), then x ≤ x + (y ∧ w) ≤ y + (x ∧ w) ≤ y + w. To show the converse implication, assume that x ≤ y + w, and let y′ ≪ y ∧ w. By (O5), we can choose z such that y′+z ≤ y ≤ y∧w+z. Then x ≤ y∧w+z+w = z+w. Let x′ ≪ x. By (O6), x′ ≤ z+x∧w. Adding y′ on both sides we get x′+y′ ≤ y+x∧w. Passing to the supremum over all x′ ≪ x and y′ ≪ y ∧ w, the result follows. (iii): Let x ∈ S and let w1, w2 ∈ S be idempotents. By (i), w1 ∧ w2 is also an idempotent. It thus makes sense to write x ∧ (w1 ∧ w2) and this agrees with (x ∧ w1) ∧ w2. We first show that x ∧ (w1 + w2) + w1 = x ∧ w2 + w1. The inequality '≥' is clear. On the other hand, applying (ii) to x ∧ (w1 + w2) ≤ w1 + w2 at the first step yields x ∧ (w1 + w2) + w1 ∧ w2 ≤ w1 + x ∧ (w1 + w2) ∧ w2 = w1 + x ∧ w2. Adding w1 to the previous inequality, we obtain the desired reverse inequality x ∧ (w1 + w2) + w1 ≤ x ∧ w2 + w1. Given y, z ∈ S satisfying y + w1 = z + w1, it follows from (ii) that y + z ∧ w1 = z + y ∧ w1. Applying this for y = x ∧ (w1 + w2) and z = x ∧ w2, we get x ∧ (w1 + w2) + (x ∧ w2) ∧ w1 = x ∧ w2 + (x ∧ (w1 + w2)) ∧ w1, which implies the desired equality. (cid:3) 2.6. Remark. Let A be a C∗-algebra, and let J be a σ-unital, closed, two-sided ideal. Then Cu(A) is a Cu-semigroup satisfying (O5)-(O7). We identify Cu(J) with the ideal {[a] ∈ Cu(A) : a ∈ (J ⊗ K)+} of Cu(A). Since J is σ-unital, there exists a largest element in Cu(J), denoted wJ . Recall that Cu(A) can be identified with certain equivalence classes of countably generated, right Hilbert C*-modules over A; see [CEI08], see also [APT11]. If M is a countably generated, right Hilbert C*-module over A, then M J is a countably generated, right Hilbert C*-module over J, and [M J] - the class of M J in Cu(J) - depends only on the class of M , which is the justification to denote [M J] by [M ]J; see [CRS10]. One can show that [M ] ∧ wJ = [M ]J in Cu(A). Hence, Theorem 2.5(i) and (ii) generalize (and recover) Proposition 4.3 and Theorem 1.1 in [CRS10] in the case that A is a separable C*-algebra. 6 RAMON ANTOINE, FRANCESC PERERA, LEONEL ROBERT, AND HANNES THIEL If S is the Cuntz semigroup of a (not necessarily separable) C*-algebra, then the results in [CRS10] show that the infimum of any x ∈ S and any idempotent w ∈ S exist. Thus, for Cuntz semigroups of C*-algebras, Theorem 2.4 holds without the assumption of countable generation. It seems unclear if the same holds for Cu-sem- igroups: 2.7. Question. Let S be a Cu-semigroup satisfying (O7). Do each x ∈ S and each idempotent w ∈ S admit an infimum in S? What, if we additionally assume that S satisfies (O5) and (O6)? 3. Cones and their duals Here we establish a number of results on algebraically ordered, compact cones and their duals. We then apply these results to our main object of study: the cone F (S) of functionals on a Cu-semigroup S. 3.1. Algebraically ordered compact cones. Recall that a cone is a commuta- tive monoid C together with a scalar multiplication by (0, ∞). More specifically, the scalar multiplication is a map (0, ∞) × C → C, denoted (t, a) 7→ ta, that is ad- ditive in each variable, and such that (st)a = s(ta) and 1a = a, for all s, t ∈ (0, ∞) and a ∈ C. Note that we do not define scalar multiplication by 0. A topological cone is a cone together with a topology such that addition and scalar multiplication are jointly continuous. (Here we equip (0, ∞) with the usual Hausdorff topology of real numbers.) The algebraic pre-order on a cone C is defined as a ≤ b if a + c = b for some If the algebraic pre-order is an order then we speak of an algebraically c ∈ C. ordered cone. The following result is standard. It holds more generally in compact, ordered spaces as studied by Nachbin, [Nac65], see [GHK+03, Proposition VI-1.3, p.441]. 3.1. Proposition. Let C be an algebraically ordered, compact cone. Then C is both directed complete and filtered complete. Moreover, given an upward (downward) directed subset D of C, considering D as a net indexed over itself, D converges to sup D (to inf D). Let C be an algebraically ordered, compact cone. We set E(C) :=(cid:8)a ∈ C : 2a = a(cid:9). n a)n is decreasing and therefore converges. We have Given a ∈ C, the sequence ( 1 2 limn 1 n a = limn 1 n a, which justifies to define ε : C → E(C) by ε(a) := lim n 1 n a, for a ∈ C. It is straightforward to verify that ε is additive and order-preserving. Moreover, we have ε(a) + a = a for every a ∈ C. Following Wehrung (Definitions 1.12, 2.10, and 3.1 in [Weh92]), we say that C is pseudo-cancellative if for all a, b, c ∈ C with a + c ≤ b + c there exists d ∈ C such that a ≤ b + d and d + c = c. 3.2. Lemma. Let C be an algebraically ordered, compact cone. Let a, b, c ∈ C satisfy a + c ≤ b + c. Then a + ε(c) ≤ b + ε(c). In particular, C is pseudo-cancellative. Proof. Multiplying by 1 2 in a + c ≤ b + c we get 2 b + 1 2 a + 1 2 c ≤ 1 1 2 c. Adding 1 2 a, and then using the above inequality, we obtain a + 1 2 c ≤ 1 2 a + 1 2 b + 1 2 c ≤ b + 1 2 c. EDWARDS' CONDITION FOR QUASITRACES ON C*-ALGEBRAS 7 Thus, we inductively deduce that a + 1 2n c ≤ b + 1 2n c, for each n ∈ N. It follows that a + ε(c) = lim n (a + 1 2n c) ≤ lim n (b + 1 2n c) = b + ε(c), as desired. (cid:3) Recall that a monoid M is said to satisfy Riesz refinement if for all a1, a2, b1, b2 ∈ M with a1+a2 = b1+b2 there exist xi,j ∈ M , for i, j = 1, 2, such that ai = xi,1+xi,2 for i = 1, 2, and bj = x1,j + x2,j for j = 1, 2. An inf-semilattice ordered monoid is a positively ordered monoid M that is an inf-semilattice and such that addition is distributive over ∧, that is, (3.1) a + (b ∧ c) = (a + b) ∧ (a + c), for all a, b, c ∈ M . Dually, one defines sup-semilattice ordered monoids. A lattice- ordered monoid is a positively ordered monoid M that is a lattice and such that addition is distributive over ∧ and ∨, that is, for all a, b, c ∈ M we have (3.1) and a + (b ∨ c) = (a + b) ∨ (a + c). The following proposition is a consequence of results of Wehrung: 3.3. Proposition. Let C be an algebraically ordered, compact cone. Then the fol- lowing are equivalent: (1) C satisfies Riesz refinement. (2) C is inf-semilattice ordered. (3) C is lattice ordered. Proof. By Lemma 3.2, C is pseudo-cancellative. Therefore, it follows from [Weh92, Proposition 1.23] that (2) implies (1); and it follows from [Weh92, Lemma 1.16] that (1) implies (3). (cid:3) Let C be an algebraically ordered, compact cone satisfying Riesz refinement. Let C∗ denote the collection of linear maps C → [0, ∞], where by a linear map we understand an additive map satisfying f (0) = 0, and such that f (ta) = tf (a) for all t ∈ (0, ∞) and a ∈ C. We equip C∗ with pointwise addition and the algebraic order. In fact, the algebraic order on C∗ agrees with the pointwise order. The property of Riesz refinement of C implies that C∗ is lattice ordered. Further, the infimum and supremum of elements f, g ∈ C∗ are given by the Riesz-Kantorovich formulas: (f ∧ g)(a) = lim n (cid:0)f (a1,n) + g(a2,n)(cid:1). Since C is compact, we can choose convergent subnets such that (a1,n(j))j∈J and (a2,n(j))j∈J converge to some a1 and a2 in C, respectively. Then a = a1 + a2. Using (3.2) (3.3) (f ∧ g)(a) = inf(cid:8)f (a1) + g(a2) : a = a1 + a2(cid:9), (f ∨ g)(a) = sup(cid:8)f (a1) + g(a2) : a = a1 + a2(cid:9), for a ∈ C. (See [Sho90, Lemma 1.12].) A map f ∈ C∗ is said to be lower semicontinuous if for every t ∈ [0, ∞) the set {a ∈ C : f (a) ≤ t} is closed (in the topology of C). We let C′ denote the family of lower semicontinuous maps in C∗. It is easy to see that C′ is closed under addition. The partial order on C′ (induced by C∗) is the pointwise order, and it is usually not the algebraic order, even though C∗ is algebraically ordered. 3.4. Lemma. Let C be an algebraically ordered, inf-semilattice ordered, compact cone, let f, g ∈ C′, and let a ∈ C. Then the infimum in (3.2) is realized. More precisely, there exist a1, a2 ∈ C with a = a1 + a2 and (f ∧ g)(a) = f (a1) + g(a2). Proof. Choose sequences (a1,n)n and (a2,n)n in C such that a = a1,n + a2,n for each n, and such that 8 RAMON ANTOINE, FRANCESC PERERA, LEONEL ROBERT, AND HANNES THIEL this at the last step, and using that f and g are lower semicontinuous at the second step, we obtain (f ∧ g)(a) = lim j (cid:0)f (a1,n(j)) + g(a2,n(j))(cid:1) ≥ f (a1) + g(a2) ≥ (f ∧ g)(a). Thus, we have a1, a2 ∈ C such that (f ∧g)(a) = f (a1)+g(a2) and a = a1+a2. (cid:3) The following result contains analogs of results in [Weh92] for lower semicontin- uous functionals. It can also be considered as an analog of [Rob13, Theorem 4.2.2]. 3.5. Theorem. Let C be an algebraically ordered, inf-semilattice ordered, compact cone. Then C′ ⊆ C∗ is closed under finite infima and directed suprema. Moreover, given f, g, h ∈ C′ and an increasing net (gj)j in C′, we have (3.4) (3.5) f ∧ (sup j gj) = sup (f ∧ gj), j f + (g ∧ h) = (f + g) ∧ (f + h). Proof. By Proposition 3.3, C satisfies Riesz refinement, and thus we obtain that C∗ is lattice-ordered with infimum given by (3.2). We first show that C′ is closed under infima. Let f, g ∈ C′. In order to verify that f ∧ g in C∗ is lower semicontinuous (and thus it belongs to C′), we have to check that the set T := {a ∈ C : (f ∧ g)(a) ≤ t} is closed for any t ∈ [0, ∞). Let (aj)j∈J be a net in T that converges to a in C. For each j ∈ J apply Lemma 3.4 to obtain aj,1, aj,2 ∈ C such that f (aj,1) + g(aj,2) = (f ∧ g)(aj) ≤ t, and aj = aj,1 + aj,2. Using that C is compact, choose a subnet (j(i))i∈I such that (aj(i),1)i∈I and (aj(i),2)i∈I converge to some a1 and a2 in C, respectively. Then a = a1 + a2. Using at the third step that f and g are lower semicontinuous, we deduce (f ∧ g)(a) ≤ f (a1) + g(a2) = f(cid:0) lim i∈I ≤ lim inf i∈I ≤ lim inf i∈I i∈I aj(i),1(cid:1) + g(cid:0) lim aj(i),2(cid:1) f(cid:0)aj(i),1(cid:1) + lim inf g(cid:0)aj(i),2(cid:1) (cid:0)f (aj(i),1) + g(aj(i),2)(cid:1) ≤ t. i∈I Secondly, it is straightforward to verify that lower semicontinuity passes to suprema of upward directed families. Further, (3.5) follows using that C∗ is lattice- ordered. Finally, let us verify (3.4). Let f ∈ C′, and let (gj)j∈J be an increasing net in C′. Set g := supj gj. It is straightforward to verify that f ∧ g ≥ supj(f ∧ gj). To show the converse inequality, let a ∈ C. Given j ∈ J, apply Lemma 3.4 to obtain aj,1, aj,2 ∈ C such that (f ∧ gj)(a) = f (aj,1) + gj(aj,2), and a = aj,1 + aj,2. Using that C is compact, choose a subnet (j(i))i∈I such that (aj(i),1)i∈I and (aj(i),2)i∈I converge to some a1 and a2 in C, respectively. Then a = a1 + a2. For each i0 ∈ I, using at the first step that the net (gj)j is increasing, and using at the second step that gj(i0) is lower semicontinuous, we obtain lim i gj(i)(aj(i),2) ≥ lim i gj(i0)(aj(i),2) ≥ gj(i0)(a2). Since this holds for all i0 ∈ I, we deduce lim i∈I gj(i)(aj(i),2) ≥ sup i0∈I gj(i0)(a2) = g(a2). EDWARDS' CONDITION FOR QUASITRACES ON C*-ALGEBRAS 9 Using this inequality and using that f is lower semicontinuous at the third step, we obtain (f ∧ gj)(a) = sup sup j∈J j∈J(cid:0)f (aj,1) + gj(aj,2)(cid:1) i∈I(cid:0)f (aj(i),1) + gj(i)(aj(i),2)(cid:1) = lim ≥ f (a1) + g(a2) ≥ (f ∧ g)(a), as desired. (cid:3) 3.2. The cone of functionals on a Cu-semigroup. Let S be a Cu-semigroup. A map λ : S → [0, ∞] is called a functional if λ(0) = 0 and if λ preserves addi- tion, order and suprema of increasing sequences. We denote by F (S) the set of functionals on S. This is a cone when endowed with the operations of pointwise addition and pointwise scalar multiplication by positive reals. We also equip F (S) with the topology such that λj → λ, for a given net (λj)j and a functional λ in F (S), provided that lim sup j λj (x′) ≤ λ(x) ≤ lim inf j λj(x) for all x′, x ∈ S with x′ ≪ x. Then F (S) is a compact cone; see [ERS11, Rob13], see also [Kei17, Theorem 3.17]. If S satisfies (O5), then F (S) is algebraically ordered; see [Rob13, Proposition 2.2.3]. Further, if S satisfies (O5) and (O6), then F (S) is an algebraically ordered, lattice ordered, compact cone; see [Rob13, Theorem 4.1.2]. Combined with Proposition 3.3, we deduce the following: 3.6. Theorem. Let S be a Cu-semigroup satisfying (O5) and (O6). Then F (S) satisfies Riesz refinement. Since the Cuntz semigroup of a C*-algebra is a Cu-semigroup satisfying (O5) and (O6), the previous result applies to F (Cu(A)). Moreover, by [ERS11, Theorem 4.4], the cone of functionals F (Cu(A)) is isomorphic (as an ordered topological cone) to the cone of lower semicontinuous 2-quasitraces QT(A) on A via the assignment QT(A) → F (Cu(A)), τ 7→ dτ . Here, dτ ([a]) := limn τ (a1/n) for a ∈ (A ⊗ K)+. We thus obtain the following result, which does not seem to have appeared in the literature before. 3.7. Corollary. Let A be a C∗-algebra. Then the cone QT(A) of lower semicontin- uous 2-quasitraces satisfies Riesz refinement. 3.8. Remark. For unital, simple C∗-algebras, Corollary 3.7 follows from more clas- sical results of Blackadar and Handelman, [BH82]. Indeed, they show that if A is a unital C∗-algebra, then the cone QTb(A) of bounded 2-quasitraces is lattice ordered. Since this cone embeds in a vector space, it follows from the well known equivalence between Riesz interpolation and Riesz refinement in the setting of ordered abelian groups that QTb(A) has Riesz refinement; see, for example, [Goo86, Proposition 2.1]. If A is also simple, then QT(A) = QTb(A) ∪ {τ∞}, where τ∞ : A+ → [0, ∞] is infi- nite on all non-zero elements of A+. It is then straightforward to extend the Riesz refinement from QTb(A) to QT(A). Given a Cu-semigroup S satisfying (O5), recall that F (S)′ denotes the family of linear, lower semicontinuous functions f : F (S) → [0, ∞]. (Note that F (S)′ is denoted by Lsc(F (S)) in [APT18] and [Rob13].) Given x ∈ S, we obtainbx ∈ F (S)′ defined bybx(λ) = λ(x), for λ ∈ F (S). Since F (S) is an algebraically ordered, lattice ordered, compact cone, we may apply Lemma 3.4 and Theorem 3.5 to obtain: 10 RAMON ANTOINE, FRANCESC PERERA, LEONEL ROBERT, AND HANNES THIEL 3.9. Proposition. Let S be a Cu-semigroup satisfying (O5) and (O6). Then F (S)′ is an inf-semilattice-ordered, directed complete cone, with infimum given as in (3.2). (3.6) In particular, given x, y ∈ S, the infimum of bx and by in F (S)′ satisfies (bx ∧by)(λ) = inf(cid:8)λ1(x) + λ2(y) : λ = λ1 + λ2(cid:9), for all λ ∈ F (S), and the infimum is attained. 3.10. Question. Let S be a countably-based Cu-semigroup satisfying (O5) and will be taken up in Section 4. Proposition 3.9, F (S)′ is inf-semilattice ordered. Thus, if S is also inf-semilattice (O6). There is a natural semigroup morphismb: S → F (S)′ given by x 7→ bx. By ordered, it is natural to ask wether [x ∧ y = bx ∧by, for all x, y ∈ S. This question 3.3. Well capped cones. Recall that a subset K of a topological cone C is called a cap if it is compact, convex, and C\K is also convex. The cone C is said to be well capped if it is the union of its caps; see, for example, [Phe01]. Let S be a Cu-semigroup satisfying (O5). In this subsection we show that the cone F (S) contains many well capped subcones; see Proposition 3.11. If S is also countably based, then F (S) naturally decomposes as the disjoint union of well capped, cancellative subcones. Recall that L(F (S)) is defined as a certain subset of F (S)′, which can be iden- tified with the sequential closure of the span of the set {tbx : t ∈ (0, ∞), x ∈ S} in F (S)′; see [Rob13]. Further, we have L(F (S)) ∼= S⊗[0, ∞]; see [APT18, Section 7.5, p.132ff]. Given an ideal J in S, we let λJ ∈ F (S) denote the functional that is 0 on J and ∞ otherwise. Then 2λJ = λJ . Moreover, every idempotent in F (S) arises this way for some ideal, that is, E(F (S)) with the reverse order is naturally order-isomorphic to the lattice of ideals in S. By a subcone of F (S) we understand a subset that is closed under addition and multiplication by strictly positive scalars. Given an ideal J in S, we set FJ (S) := λJ +(cid:8)λ ∈ F (S) : λ(x′) < ∞ whenever x′ ≪ x for some x ∈ J(cid:9). Then FJ (S) is a subcone of F (S) with apex λJ . The cone F (S) decomposes as the disjoint union of the subcones FJ (S), with J ranging over the ideals of S. The support ideal of λ ∈ F (S) is the unique ideal J such that λ ∈ FJ (S). One can show that the support ideal of λ is J if and only if ε(λ) = λJ . 3.11. Proposition. Let S be a Cu-semigroup satisfying (O5), and let J be a count- ably generated ideal of S. Then FJ (S) is well capped. Proof. Since J is countably generated, it contains a largest element; see [APT18, Paragraph 5.1.6, p.39f], and also the comments before Theorem 2.4. Choose a ≪-increasing sequence (xn)n whose supremum is the largest element of J. Let λ ∈ FJ (S). Then (λ(xn))n is an increasing sequence in [0, ∞). Define where we choose the numbers (αn)n in (0, ∞) such that αn → 0 fast enough so that f = ∞Xn=1 αncxn ∈ L(F (S)), f (λ) ≤ 1. Observe thatbx ≤ ∞f for any x ∈ J. We consider Cf :=(cid:8)µ ∈ FJ (S) : f (µ) ≤ 1(cid:9), which contains λ. Let us show that Cf is a cap of FJ (S). Since f is linear, both Cf and its complement in FJ (S) are convex. It remains to show that Cf is compact. EDWARDS' CONDITION FOR QUASITRACES ON C*-ALGEBRAS 11 We show first that if µ ∈ F (S) is such that f (µ) ≤ 1 then λJ + µ ∈ FJ (S). Let x′ ≪ x in J. Using [Rob13, Lemma 2.2.5] at the first step, we get Hence, bx′ ≤ N f for some N ∈ N. Then µ(x′) ≤ N < ∞, which in turn implies that λJ + µ ∈ FJ (S). Thus, Cf agrees with λJ + {µ ∈ F (S) : f (µ) ≤ 1}. This set is closed in F (S) and therefore compact. (cid:3) bx′ ≪ 2bx ≤ ∞f. 4. Edwards' condition for abstract Cuntz semigroups In this section we introduce Edwards' condition for Cu-semigroups; see Definition 4.1. This condition is inspired by a property considered by Edwards [Edw69, Condi- tion (2)], and it has been studied in a more restrictive setting in [Thi17]. In Theorem 5.3 below we show that Edwards' condition is satisfied by Cuntz semi- groups of general C∗-algebras. 4.1. Definition. Let S be a Cu-semigroup and let λ ∈ F (S). We say that S satisfies Edwards' condition for λ if (4.1) inf(cid:8)λ1(x) + λ2(y) : λ = λ1 + λ2(cid:9) = sup(cid:8)λ(z) : z ≤ x, y(cid:9), for all x, y ∈ S. If this holds for all λ ∈ F (S), then we say that S satisfies Edwards' condition. 4.2. Remark. Let S be a Cu-semigroup satisfying (O5) and (O6). It follows from Lemma 3.4 that the infimum in (4.1) is attained (see also the remarks be- fore Theorem 3.6). Further, it follows from Proposition 3.9 that the left hand side only if in (4.1) agrees with (bx ∧by)(λ). Thus, S satisfies Edwards' condition for λ if and (4.2) (bx ∧by)(λ) = sup(cid:8)λ(z) : z ≤ x, y(cid:9), for all x, y ∈ S. Notice that the inequality '≥' always holds. If S is also an inf-semilattice, then we have sup{λ(z) : z ≤ x, y} = [x ∧ y(λ). Therefore, in this setting, Edwards' condition is equivalent to for all x, y ∈ S. (See Question 3.10.) bx ∧by = [x ∧ y, Next, we show that the supremum in (4.1) is achieved. We first need a lemma. 4.3. Lemma. Let S be a Cu-semigroup satisfying (O5) and (O6), let λ ∈ F (S) such that S satisfies Edwards' condition for λ, let z′ ≪ z ≤ x, y in S, and let t ∈ R Proof. If t < λ(z), then we can set z := z. Thus, we may assume that λ(z) ≤ t, and in particular λ(z) is finite. We distinguish two cases. satisfy t < (bx ∧by)(λ). Then there exists z ∈ S such that z′ ≪ z ≤ x, y and t < λ(z). Case 1. Assume that (bx ∧by)(λ) < ∞. t + ε < (bx ∧by)(λ). Since λ(z) < ∞, we can choose z′′ ∈ S such that z′ ≪ z′′ ≪ z, and λ(z) < λ(z′′) + ε/2. In this case, choose ε > 0 such that Applying (O5) to z′′ ≪ z ≤ x and z′′ ≪ z ≤ y, we obtain u, v ∈ S such that z′′ + u ≤ x ≤ z + u, and z′′ + v ≤ y ≤ z + v. Since (bu ∧bv)(λ) < ∞, we can apply Edwards' condition to obtain w ∈ S such that w ≤ u, v, and (bu ∧bv)(λ) ≤ λ(w) + ε/2. 12 RAMON ANTOINE, FRANCESC PERERA, LEONEL ROBERT, AND HANNES THIEL Set z := z′′ + w. Then z′ ≪ z ≤ x, y. Using that F (S)′ is semilattice-ordered (Proposition 3.9) at the second step, we deduce (bx ∧by)(λ) ≤ ( [z + u ∧ [z + v)(λ) =bz(λ) + (bu ∧bv)(λ) ≤ λ(z′′) + ε/2 + λ(w) + ε/2 = λ(z) + ε, which implies t < λ(z). This proves this case of the lemma. Construct u and v as in case 1. Then Case 2. Suppose that (bx ∧by)(λ) = ∞. Choose z′′ ∈ S satisfying z′ ≪ z′′ ≪ z. which implies that (bu ∧bv)(λ) = ∞. Applying Edwards' condition, we obtain w ∈ S ∞ = (bx ∧by)(λ) ≤ ( [z + u ∧ [z + v)(λ) =bz(λ) + (bu ∧bv)(λ), such that w ≤ u, v and t < λ(w). Then, as in Step 1, the element z := z′′ + w has the desired properties. (cid:3) 4.4. Theorem. Let S be a Cu-semigroup satisfying (O5) and (O6), let λ ∈ F (S) such that S satisfies Edwards' condition for λ, and let x, y ∈ S. Then there exists z ∈ S such that z ≤ x, y and Moreover, given also z′ such that z′ 0 ≪ z. 0, z0 ∈ S with z′ 0 ≪ z0 ≤ x, y, the element z may be chosen (bx ∧by)(λ) = λ(z). Proof. Let (tn)n be a strictly increasing sequence in R with supn tn = (bx ∧by)(λ). 0 and z0 are not given, we simply consider z′ 0 = 0 and z0 = 0. We inductively n, zn ∈ S for n ≥ 1 such that If z′ construct z′ z′ n−1 ≪ z′ n ≪ zn ≤ x, y, and tn < λ(z′ n), for n ≥ 1. Given n ≥ 1, assume that z′ n−1, zn−1 with z′ Using Lemma 4.3, we obtain zn ∈ S such that n−1 ≪ zn−1 ≤ x, y have been chosen. z′ n−1 ≪ zn ≤ x, y, and tn < λ(zn). Choose z′ n ∈ S such that z′ n−1 ≪ z′ n ≪ zn, and tn < λ(z′ n). Then z′ n and zn have the claimed properties. We obtain a ≪-increasing sequence (z′ Then z′ 0 ≪ z ≤ x, y and n)n, which allows us to set z := supn z′ n. which implies that z has the desired properties. (cid:3) (bx ∧by)(λ) = sup n tn ≤ sup n λ(z′ n) = λ(z), 4.5. Corollary. Let S be a Cu-semigroup satisfying (O5) and (O6) and Edwards' condition. Then for every λ ∈ F (S) and x, y ∈ S, there exist λ1, λ2 ∈ F (S) and z ∈ S such that λ = λ1 + λ2, z ≤ x, y, and λ1(x) + λ2(y) = λ(z). The following result can be interpreted as the fact that the Edwards' condition implies that its dual version is also satisfied. It is not clear wether or not these conditions are actually equivalent. EDWARDS' CONDITION FOR QUASITRACES ON C*-ALGEBRAS 13 4.6. Proposition. Let S be a Cu-semigroup satisfying (O5) and (O6). Let λ ∈ F (S) be such that S satisfies Edwards' condition for λ. Then (4.3) for all x, y ∈ S. sup(cid:8)λ1(x) + λ2(y) : λ1 + λ2 = λ(cid:9) = inf(cid:8)λ(a) : x, y ≤ a(cid:9), Proof. The inequality '≤' in (4.3) is straightforward to obtain. Let us show the opposite inequality. Let r denote the left side. If r = ∞, we are done. Let us thus suppose that r < ∞. Observe that this implies that λ(x) < ∞ and λ(y) < ∞. Applying Corollary 4.5, we obtain λ1, λ2 ∈ F (S) and z ∈ S such that λ = λ1 + λ2, and z ≤ x, y, and λ1(x) + λ2(y) = λ(z). Let ε > 0. Since λ(z) is finite, we can choose z′ ≪ z such that λ(z) ≤ λ(z′) + ε. Applying (O5) for z′ ≪ z ≤ x and z′ ≪ z ≤ y, we obtain u, v ∈ S such that z′ + u ≤ x ≤ z + u, and z′ + v ≤ y ≤ z + v. Set a := z + u + v which clearly satisfies x, y ≤ a. Then λ(a) + λ1(x) + λ2(y) = λ(a) + λ(z) = λ(z) + λ(u) + λ(z) + λ(v) ≤ λ(z′) + λ(u) + λ(z′) + λ(v) + 2ε ≤ λ(x) + λ(y) + 2ε = λ1(x) + λ2(x) + λ1(y) + λ2(y) + 2ε. Since λ1(x) and λ2(y) are finite, we may cancel them and obtain λ(a) ≤ λ2(x) + λ1(y) + 2ǫ ≤ r + 2ε, which implies the desired inequality. (cid:3) The following result shows that to prove Edwards' condition for S, it suffices to deal with the case where the functional λ is finite on the given elements x, y ∈ S. This reduction will come in handy when we prove Edwards' condition for Cuntz semigroups of C*-algebras. 4.7. Theorem. Let S be a Cu-semigroup satisfying (O5), (O6) and (O7), and let λ ∈ F (S). Then S satisfies Edwards' condition for λ if, for all x, y ∈ S with λ(x), λ(y) < ∞, we have We will prove the theorem using a series of lemmas. Let S be a Cu-semigroup and let J ⊆ S be an ideal. We define λJ : S → [0, ∞] as in Subsection 3.3, and hJ : F (S) → [0, ∞] as follows: (bx ∧by)(λ) ≤ sup(cid:8)λ(z) : z ≤ x, y(cid:9). hJ (λ) =(0 if λ ≤ λJ ∞ if λ (cid:2) λJ . Observe that, if J has a largest element wJ (for example, if J is countably based), then hJ = cwJ . 4.8. Lemma. Let S be a Cu-semigroup and let λ ∈ F (S). Set Then J is an ideal in S and ε(λ) = λJ . J :=(cid:8)x ∈ S : λ(x′) < ∞ for all x′ ≪ x(cid:9). for all λ ∈ F (S). Proof. Let λ ∈ F (S). Recall that F (S) is a complete lattice. This allows us to define (bx ∧ hJ )(λ) = sup(cid:8)bz(λ) : z ∈ J, z ≤ x(cid:9), λ(J) := inf(cid:8)µ ∈ F (S) : λ ≤ µ + λJ(cid:9). 14 RAMON ANTOINE, FRANCESC PERERA, LEONEL ROBERT, AND HANNES THIEL Proof. It is straightforward to verify that J is an ideal. The sequence ( 1 n λ)n converges to ε(λ) in F (S). By definition of the topology in F (S), this means that for all x′, x ∈ S with x′ ≪ x, we have lim sup n 1 n λ(x′) ≤ ε(λ)(x) ≤ lim inf n 1 n λ(x). To show that ǫ(λ) ≥ λJ , let x ∈ S satisfy ǫ(λ)(x) = 0. We need to verify that x ∈ J. Let x′ ≪ x. If λ(x′) = ∞, then lim supn 1 n λ(x′) = ∞, which contradicts lim sup n 1 n λ(x′) ≤ ε(λ)(x) = 0. Thus, λ(x′) < ∞. Since this holds for all x′ ≪ x, we conclude that x ∈ J. To show the converse inequality, let x ∈ J. We need to verify that ε(λ)(x) = 0. Choose a ≪-increasing sequence (xn)n with supremum x. By assumption, we have λ(xn) < ∞ for each n. This implies that ε(λ)(xn) = 0. Using that ε(λ) preserves suprema of increasing sequences, we deduce that ε(λ)(x) = 0. (cid:3) 4.9. Lemma. Let S be a Cu-semigroup satisfying (O5) and (O6), let J ⊆ S be an ideal of S, and let x ∈ S. Then We have λ(J) ≤ λ and λ(J) + λJ = λ + λJ . The result will follow by combining the following two claims. Claim 1 : Given y ∈ S, we have λ(J)(y) = sup(cid:8)λ(z) : z ∈ J, z ≤ y(cid:9). To prove the claim, let z ∈ J satisfy z ≤ y. Then λJ (z) = 0 and therefore λ(z) = λ(z) + λJ (z) = λ(J)(z) + λJ (z) = λ(J)(z) ≤ λ(J)(y), which shows inequality '≥'. To prove the converse inequality, one shows that the function µ : S → [0, ∞] defined by µ(y) := sup(cid:8)λ(z) : z ∈ J, z ≤ y(cid:9), for y ∈ S, is a functional on S satisfying λ ≤ µ + λJ . By definition of λ(J) we obtain λ(J) ≤ µ. This proves the claim. and Theorem 3.6 at the third step, we deduce Claim 2 : We have (bx ∧ hJ )(λ) = λ(J)(x). Indeed, using (3.2) at the first step (bx ∧ hJ )(λ) = inf(cid:8)λ1(x) + hJ (λ2) : λ = λ1 + λ2(cid:9) = inf(cid:8)λ1(x) : λ = λ1 + λ2, λ2 ≤ λJ(cid:9) = inf(cid:8)µ(x) : λ ≤ µ + λJ(cid:9) = λ(J)(x), which proves the claim. (cid:3) Given a Cu-semigroup S and x ∈ S, we let hxi denote the ideal generated by x. 4.10. Lemma. Let S be a Cu-semigroup satisfying (O5), (O6) and (O7), let x, y ∈ S, and let λ ∈ F (S). Then there exist r, s ∈ S such that r ≤ x, s ≤ y, r, s ∈ hxi ∩ hyi, and (bx ∧by)(λ) = (br ∧bs)(λ). Similarly, the set Dy := {s : s ≤ y, s ∈ hxi} is upward directed. We deduce r∈Dxbr) ∧by = sup r∈Dx(cid:0)br ∧by(cid:1). bx ∧by =bx ∧(cid:0)hhyi ∧by(cid:1) =(cid:0)bx ∧ hhyi(cid:1) ∧by = ( sup r∈Dx,s∈Dy(cid:0)br ∧bs(cid:1). n (cid:0)brn ∧csn(cid:1)(λ). bx ∧by = (bx ∧by)(λ) = sup sup Choose sequences (rn)n in Dx and (sn)n in Dy such that EDWARDS' CONDITION FOR QUASITRACES ON C*-ALGEBRAS 15 Proof. Set Dx := {r : r ∈ hyi, r ≤ x}. By Proposition 2.3, Dx is upward directed. By Theorem 3.5, infima commute with directed suprema in F (S)′. Using this at third step, we obtain the last step, and using thatby ≤ hhyi at the first step, and using Lemma 4.9 at the Using that Dx and Dy are upward directed, we may assume that (rn)n and (sn)n are increasing. Then r := supn rn and s := supn sn have the desired properties. (cid:3) 4.11. Lemma. Let S be a Cu-semigroup satisfying (O6), and let x, y ∈ S. Then hxi ∩ hyi is the ideal generated by {z ∈ S : z ≤ x, y}. Proof. Let z ∈ hxi ∩ hyi and z′ ≪ z. Then z′ ≤ nx and z′ ≤ ny for some n ∈ N. Let z′′ ≪ z′. By (O6) used in z′′ ≪ z′ ≤ nx, there exist x1, . . . , xn such that k ≪ xk such k=1 xk and xk ≤ z′, x for all k. Choose for each k an element x′ For each k, applying (O6) again in x′ k ≪ xk ≤ ny, we obtain yk,1, . . . , yk,n such that x′ l=1 yk,l and yk,l ≤ xk, y for all l. It follows that yk,l belongs to the set {z : z ≤ x, y} for all k and l. Hence, z′′ belongs to the ideal of S generated by this set. Since z′′ and z′ can be chosen arbitrarily such that z′′ ≪ z′ ≪ z, we deduce that z belongs to this ideal as well. (cid:3) k=1 x′ k. z′′ ≪Pn that z′′ ≪Pn k ≪Pn Proof of Theorem 4.7. Let x, y ∈ S and λ ∈ F (S). The inequality '≥' in (4.2) is clear (see also (4.1)). We prove the opposite inequality, that is, (bx ∧by)(λ) ≤ sup(cid:8)λ(z) : z ≤ x, y(cid:9). If the right hand side is ∞ we are done. Let us thus assume that x, y, and λ are such that if z ≤ x, y then λ(z) is finite. Let J ⊆ S be the ideal as in Lemma 4.8 such that ǫ(λ) = λJ . Namely, J = {x ∈ S : λ(x′) < ∞ for all x′ ≪ x}. Then z ∈ J whenever z ≤ x, y. Thus, by the previous lemma, we have hxi ∩ hyi ⊆ J. Use Lemma 4.10 to obtain r, s ∈ S such that r ≤ x, s ≤ y, r, s ∈ hxi ∩ hyi ⊆ J, and (bx ∧by)(λ) = (br ∧bs)(λ). Choose ≪-increasing sequences (rn)n and (sn)n with suprema r and s, respectively. Since r, s ∈ J, we have that λ(rn) < ∞ and λ(sn) < ∞ for each n. By assumption, we can choose zn such that n ≤ λ(zn), and zn ≤ rn, sn. Using Theorem 3.5 at the second step, we deduce (brn ∧csn)(λ) − 1 (bx ∧by)(λ) = (br ∧bs)(λ) = sup n as desired. (brn ∧csn)(λ) = sup n λ(zn) ≤ sup(cid:8)λ(z) : z ≤ x, y(cid:9), (cid:3) Jτ =(cid:8)(an)n ∈ ℓ∞(A) : lim ¯τ(cid:0)π((an)n)(cid:1) = lim U τ (a∗ U nan) = 0(cid:9). τ (an), 16 RAMON ANTOINE, FRANCESC PERERA, LEONEL ROBERT, AND HANNES THIEL 5. Edwards' condition for Cuntz semigroups of C*-algebras In this section we prove the main result of this paper, namely that Cuntz semi- groups of C*-algebras satisfy Edwards' condition. To this end, we first recall nec- essary results and constructions from [BH82] and [Haa14]. 5.1. Let A be a C*-algebra, and let τ : A → C be a bounded 2-quasitrace on A. Denote by ℓ∞(A) the C∗-algebra of norm-bounded sequences in A. Given a free ultrafilter U on N, let Jτ ⊆ ℓ∞(A) be defined as Then Jτ is a closed, two-sided ideal and Mτ := ℓ∞(A)/Jτ is an AW ∗-algebra. Moreover, there exists a bounded 2-quasitrace ¯τ : Mτ → C such that where π : ℓ∞(A) → Mτ denotes the quotient map. (See [Haa14, Proposition 4.2] and [BH82, I.4 and II.2].) The 2-quasitrace τ extends to a lower semicontinuous 2-quasitrace on A ⊗ K. As in Subsection 3.2, we denote by dτ ∈ F (Cu(A)) the functional associated to τ . Recall that dτ ([a]) := limn τ (a1/n) for all a ∈ (A ⊗ K)+. Since this is independent of the class [a] of a, we may also write dτ (a) in place of dτ ([a]). Recall also that the assignment τ 7→ dτ allows us to identify the cone of lower semicontinuous 2-quasitraces QT(A) with F (Cu(A)); see [ERS11]. Now, for a ∈ A+, we set pa := π((a1/n)n) ∈ Mτ . Then pa is a projection in Mτ such that ¯τ (pa) = dτ (a). 5.2. Lemma. Let A be a C*-algebra, let τ be a bounded 2-quasitrace on A, and let a, b ∈ A+. Then (5.1) (c[a] ∧ b[b])(dτ ) = max(cid:8)¯τ (q) : q ∈ Mτ is a projection such that q - pa, pb(cid:9). Proof. We identify F (Cu(A)) with QT(A) as explained above. Set λ := dτ . follows from Proposition 3.9 that It (c[a] ∧ b[b])(λ) = inf(cid:8)λ1([a]) + λ2([b]) : λ = λ1 + λ2(cid:9). Let λ1, λ2 ∈ F (Cu(A)) satisfy λ = λ1 + λ2. Let τ1, τ2 ∈ QT(A) be such that λ1 = dτ1 and λ2 = dτ2 Then τ = τ1 + τ2. It follows that τ1 and τ2 are bounded 2- quasitraces that induce bounded 2-quasitraces ¯τ1 and ¯τ2 on Mτ such that ¯τ = ¯τ1+¯τ2. (For projections, Cuntz subequivalence as recalled at the beginning of Section 2 agrees with Murray-von Neumann subequivalence.) Then Let q ∈ Mτ be a projection satisfying q - pa, pb. λ1([a]) + λ2([b]) = dτ1(a) + dτ2 (b) = ¯τ1(pa) + ¯τ2(pb) ≥ ¯τ1(q) + ¯τ2(q) = ¯τ (q). Passing to the infimum over all decompositions λ = λ1 + λ2 and the supremum over all such projections q, we obtain the inequality '≥' in (5.1). Let us show the converse. By [Ber72, Corollary 14.1, p.80], AW ∗-algebras have generalized comparability, that is, given two projections e, f there exists a central projection z such that ze - zf and (1 − z)e % (1 − z)f . Applied to pa, pb ∈ Mτ , we obtain a central projection z ∈ Mτ such that zpa - zpb and (1 − z)pa % (1 − z)pb. Set r := zpa + (1 − z)pb. Then r is a projection satisfying r - pa, pb. Given a projection r′ ∈ Mτ with r′ - pa, pb let us verify r′ - r. Indeed, r′ - pa implies zr′ - zpa and similarly we obtain (1 − z)r′ - (1 − z)pb. Then r′ = zr′ + (1 − z)r′ - zpa + (1 − z)pb = r. EDWARDS' CONDITION FOR QUASITRACES ON C*-ALGEBRAS 17 Thus, the right hand side in (5.1) is equal to ¯τ (r). Define ¯τ1, ¯τ2 : Mτ → C by ¯τ1(y) = ¯τ (zy) and ¯τ2(y) = ¯τ ((1 − z)y) for all y ∈ Mτ . Now regard A embedded in ℓ∞(A) as constant sequences, and let τ1, τ2 : A → C be the induced 2-quasitraces on A, that is, τ1(a) = ¯τ1(π(a)) and τ2(a) = ¯τ2(π(a)) for all a ∈ A. Then τ1, τ2 ∈ QT(A) and τ = τ1 + τ2. Thus, λ = dτ1 + dτ2. It follows that (c[a] ∧ b[b])(λ) = inf(cid:8)λ1([a]) + λ2([b]) : λ = λ1 + λ2 in F (Cu(A))(cid:9) ≤ dτ1 ([a]) + dτ2([b]) = ¯τ1(pa) + ¯τ2(pb) = ¯τ (zpa) + ¯τ ((1 − z)pb) = ¯τ (zpa + (1 − z)pb) = ¯τ (r), which completes the proof. (cid:3) 5.3. Theorem. Let A be a C*-algebra. Then Cu(A) satisfies Edwards' condition. Proof. First, we may assume that A is stable. Recall that Cu(A) is a Cu-semi- group satisfying (O5) and (O6). By Proposition 2.2 it also satisfies (O7). Hence by Theorem 4.7, it is enough to show that (c[a] ∧ b[b])(λ) ≤ sup(cid:8)λ([c]) : [c] ≤ [a], [b](cid:9) for all λ ∈ F (Cu(A)) and [a], [b] ∈ Cu(A) with λ([a]), λ([b]) < ∞. We continue to identify F (Cu(A)) with QT(A), and therefore we consider τ ∈ QT(A) and a, b ∈ A+ with dτ (a), dτ (b) < ∞. Let h = a + b. Observe that a, b ∈ hAh and dτ (h) < ∞. Set B := hAh. The restriction of τ to B is a bounded 2-quasitrace with norm dτ (h). Choose a free ultrafilter U on N and consider the AW ∗-algebra Mτ , with bounded 2-quasitrace ¯τ , associated to the pair (B, τ ) as described in Paragraph 5.1. Set pa = π((a1/n)n) and pb = π((b1/n)n) where π is the quotient map. Apply Lemma 5.2 to obtain a projection q ∈ Mτ satisfying (c[a] ∧ b[b])(λ) = ¯τ (q) and q - pa, pb. We may assume that q ≤ pa. Choose v ∈ Mτ with q = vv∗ and v∗v ≤ pb. Lift v to a contractive element ¯v = (vn)n in ℓ∞(B). For each n, set wn := a1/nvnb1/n. Set w := (wn)n. Then w ∈ ℓ∞(B) and π(w) = pavpb = v. Let t < (ba ∧bb)(λ). Since (c[a] ∧ b[b])(λ) = ¯τ (q) = lim U τ (wnw∗ n), there exists n ∈ N such that t < τ (wnw∗ n). Set c := wnw∗ construction satisfies c - a, b. Therefore t ≤ dτ (c) and the result follows. n) ≤ dτ (wnw∗ n which by (cid:3) References [APRT18] R. Antoine, F. Perera, L. Robert, and H. Thiel, C ∗-algebras of stable rank one [APS11] and their Cuntz semigroups, preprint (arXiv:1809.03984 [math.OA]), 2018. R. Antoine, F. Perera, and L. Santiago, Pullbacks, C(X)-algebras, and their Cuntz semigroup, J. Funct. Anal. 260 (2011), 2844 -- 2880. MR 2774057. [APT18] R. Antoine, F. Perera, and H. Thiel, Tensor products and regularity properties of Cuntz semigroups, Mem. Amer. Math. Soc. 251 (2018), viii+191. MR 3756921. [APT11] P. Ara, F. Perera, and A. S. Toms, K-theory for operator algebras. Classification of C ∗-algebras, in Aspects of operator algebras and applications, Contemp. Math. 534, Amer. Math. Soc., Providence, RI, 2011, pp. 1 -- 71. MR 2767222. Zbl 1219.46053. 18 RAMON ANTOINE, FRANCESC PERERA, LEONEL ROBERT, AND HANNES THIEL [Ber72] [BH82] S. K. Berberian, Baer *-rings, Springer-Verlag, New York-Berlin, 1972, Die Grundlehren der mathematischen Wissenschaften, Band 195. MR 0429975. Zbl 0242.16008. B. Blackadar and D. Handelman, Dimension functions and traces on C ∗-algebras, J. Funct. Anal. 45 (1982), 297 -- 340. MR 650185. Zbl 0513.46047. [CEI08] [DT10] [Cun78] [Edw69] [CRS10] A. Ciuperca, L. Robert, and L. Santiago, The Cuntz semigroup of ideals and quotients and a generalized Kasparov stabilization theorem, J. Operator Theory 64 (2010), 155 -- 169. MR 2669433. Zbl 1212.46084. K. T. Coward, G. A. Elliott, and C. Ivanescu, The Cuntz semigroup as an in- variant for C ∗-algebras, J. Reine Angew. Math. 623 (2008), 161 -- 193. MR 2458043. Zbl 1161.46029. J. Cuntz, Dimension functions on simple C ∗-algebras, Math. Ann. 233 (1978), 145 -- 153. MR 0467332. Zbl 0354.46043. M. Dadarlat and A. S. Toms, Ranks of operators in simple C ∗-algebras, J. Funct. Anal. 259 (2010), 1209 -- 1229. MR 2652186. Zbl 1202.46061. D. A. Edwards, On uniform approximation of affine functions on a compact convex set, Quart. J. Math. Oxford Ser. (2) 20 (1969), 139 -- 142. MR 0250044. Zbl 0177.16303. [ERS11] G. A. Elliott, L. Robert, and L. Santiago, The cone of lower semicontinuous traces on a C ∗-algebra, Amer. J. Math. 133 (2011), 969 -- 1005. MR 2823868. Zbl 1236.46052. [GHK+03] G. Gierz, K. H. Hofmann, K. Keimel, J. D. Lawson, M. Mislove, and D. S. Scott, Continuous lattices and domains, Encyclopedia of Mathematics and its Applications 93, Cambridge University Press, Cambridge, 2003. MR 1975381. Zbl 1088.06001. K. R. Goodearl, Partially ordered abelian groups with interpolation, Mathematical Surveys and Monographs 20, American Mathematical Society, Providence, RI, 1986. MR 845783. Zbl 0589.06008. U. Haagerup, Quasitraces on exact C ∗-algebras are traces, C. R. Math. Acad. Sci. Soc. R. Can. 36 (2014), 67 -- 92. MR 3241179. Zbl 1325.46055. K. Keimel, The Cuntz semigroup and domain theory, Soft Comput. 21 (2017), 2485 -- 2502. Zbl 1391.46066. L. Nachbin, Topology and order, Translated from the Portuguese by Lulu Bechtol- sheim. Van Nostrand Mathematical Studies, No. 4, D. Van Nostrand Co., Inc., Prince- ton, N.J.-Toronto, Ont.-London, 1965. MR 0219042. Zbl 0131.37903. [Goo86] [Haa14] [Kei17] [Nac65] [ORT11] E. Ortega, M. Rørdam, and H. Thiel, The Cuntz semigroup and comparison of open [Phe01] [Rob13] [RW10] [Sho90] [Thi17] [Weh92] projections, J. Funct. Anal. 260 (2011), 3474 -- 3493. MR 2781968. Zbl 1222.46043. R. R. Phelps, Lectures on Choquet's theorem, second ed., Lecture Notes in Mathe- matics 1757, Springer-Verlag, Berlin, 2001. MR 1835574. Zbl 0997.46005. L. Robert, The cone of functionals on the Cuntz semigroup, Math. Scand. 113 (2013), 161 -- 186. MR 3145179. Zbl 1286.46061. M. Rørdam and W. Winter, The Jiang-Su algebra revisited, J. Reine Angew. Math. 642 (2010), 129 -- 155. MR 2658184. Zbl 1209.46031. R. M. Shortt, Duality for cardinal algebras, Forum Math. 2 (1990), 433 -- 450. MR 1067211. Zbl 0717.28003. H. Thiel, Ranks of operators in simple C ∗-algebras with stable rank one, Commun. Math. Phys. (to appear), preprint (arXiv:1711.04721 [math.OA]), 2017. F. Wehrung, Injective positively ordered monoids. I, J. Pure Appl. Algebra 83 (1992), 43 -- 82. MR 1190444. Zbl 0790.06016. Ramon Antoine & Francesc Perera, Departament de Matem`atiques, Universitat Aut`onoma de Barcelona, 08193 Bellaterra, Barcelona, Spain E-mail address: [email protected], [email protected] Leonel Robert, Department of Mathematics, University of Louisiana at Lafayette, Lafayette, LA 70504-1010, USA E-mail address: [email protected] Hannes Thiel, Mathematisches Institut, Universitat Munster, Einsteinstr. 62, 48149 Munster, Germany E-mail address: [email protected]
1806.04638
1
1806
2018-06-12T16:46:42
On the Primitive Ideals of Nest Algebras
[ "math.OA" ]
We show that Ringrose's diagonal ideals are primitive ideals in a nest algebra (subject to the Continuum Hypothesis). This provides for the first time concerete descriptions of enough primitive ideals to obtain the Jacobson radical as their intersection. Separately, we provide a standard form for all left ideals of a nest algebra, which leads to insights into the maximal left ideals. In the case of atomic nest algebras we show how primitive ideals can be categorized by their behaviour on the diagonal, and provide concrete examples of all types.
math.OA
math
ON THE PRIMITIVE IDEALS OF NEST ALGEBRAS JOHN LINDSAY ORR Abstract. We show that Ringrose's diagonal ideals are primitive ideals in a nest algebra (subject to the Continuum Hypothesis). This provides for the first time concerete descriptions of enough primitive ideals to obtain the Jacobson radical as their intersection. Separately, we provide a standard form for all left ideals of a nest algebra, which leads to insights into the maximal left ideals. In the case of atomic nest algebras we show how primitive ideals can be categorized by their behaviour on the diagonal, and provide concrete examples of all types. . A O h t a m [ 1 v 8 3 6 4 0 . 6 0 8 1 : v i X r a 1. Introduction The Jacobson radical has been a frequent object of study in non-selfadjoint al- gebras, and considerable effort has been expended to identify the radical in the context of various classes of non-selfadjoint algebras, e.g., [23, 22, 5, 8, 16, 4, 9, 12]. Why is this? At fist glance it might seem that since many non-selfadjoint algebras are modelled more or less on the algebra of finite-dimensional upper triangular ma- trices, the desire is to obtain Wedderburn-type structure theorems for the algebras. In fact, however, the Jacobson radical is rarely the right ideal for such a decompo- sition, if it is even possible. The Jacobson radical is often too small, and indeed in some cases non-selfadjoint algebras are even semisimple [8, 16, 9]. Thus knowledge about the Jacobson radical rather points towards more general structural informa- tion about the algebra and, in particular, when the radical is small, indicates the presence of a rich supply of irreducible representations, even in algebras which have a strong heuristic connection with the upper triangular matrix algebra. The nest algebras are one such case. Indeed the main result of Ringrose's paper [23], which introduced the class of nest algebras, was to describe the Jacobson radical RN of a nest algebra T (N ) (see Section 2 below for precise definitions of terms). However, except in the trivial case of a finite nest, there is no Wedderburn- type decomposition T (N ) = D(N )⊕RN as the sum of the diagonal algebra and the Jacobson radical. In fact by [18, Theorem 4.1], a decomposition T (N ) = D(N )⊕R for some ideal R is only possible if R is Larson's ideal R∞ N [14], and then only if the nest has no continuous part. At issue here is the fact that unless the nest is finite R∞ N is much bigger than the Jacobson radical; in the case of upper triangular matrixes on ℓ2(N), R∞ N is the collection of all strictly upper triangular operators, while RN is the set of compact strictly upper triangular operators. Thus, the comparatively small Jacobson radical in nest algebras indicates that there must be many irreducible representations other than the trivial ones obtained as the compression to an atom of the nest. 1991 Mathematics Subject Classification. 47L35, 47L75. Key words and phrases. nest algebra, primitive ideals, nets, continuum hypothesis. 1 2 JOHN LINDSAY ORR N and in fact the two coincide when the nest is atomic. However, up to now, the only other primitive ideals which could be identified explicitly were the maximal two-sided ideals. (Maximal two-sided ideals are primi- tive; see Remark 3.1 for a review of this and other ring-theoretic facts.) In [17] we described the maximal two-sided ideals of a continuous nest algebra and in [20] we extended the description to cover all nest algebras. (It should be noted these results rest on deep foundations; between them, they require the similarity theory of nests and the Paving Theorem.) Even so, however, these ideals alone do not account for the small Jacobson radical. Their intersection, called the strong radical, is similar in character to R∞ The goal of this paper is to identify enough examples of primitive ideals of nest algebras to account for the small Jacobson radical, by which we mean that their intersection should equal the Jacobson radical. The key examples have been in plain view all along; they are the "diagonal ideals" which Ringrose used in his original description of the radical [23, Theorem 5.3]. We shall show in Theorem 3.7 that the diagonal ideals are primitive. This answers an open question of Lance [13] (repeated in [2]). Interestingly, this result relies on assuming a positive answer to the Continuum Hypothesis. See the excellent survey paper [24] for other recent results in operator algebras which make use of nonstandard foundational considerations. After this, we turn to an analysis of the left ideals of nest algebras in Section 4. We establish a standard form for all left ideals, and also a stronger form which holds for many norm-closed left ideals, including the maximal left ideals. In Sec- tion 5 we explore the primitive ideals of atomic nest algebras in more depth. We identify three classes of primitive ideals (the smallest, the largest, and the inter- mediate ones), and we show that they are distinguished by their behaviour on the diagonal. Section 6 focusses on the infinite upper-triangular matrices, where we can give concrete examples of all types of primitive ideals, and also applications to quasitriangular algebras. 2. Preliminaries Throughout this paper the underlying Hilbert spaces are always assumed sepa- rable. A nest is a set of projections on a Hilbert space which is linearly ordered, contains 0 and I, and is weakly closed (or, equivalently, order-complete). The nest algebra, T (N ), of a nest N is the set of bounded operators leaving invariant the ranges of N . The diagonal algebra, D(N ), is the set of operators having the ranges of projections in N as reducing subspaces; equivalently, the commutant of N . An interval of N is the difference N − M of two projections N > M in N . Minimal intervals are called atoms and the atoms (if there are any) are pairwise orthogonal. If the join of the atoms is I the nest is called atomic; if there are no atoms it is called continuous. For N ∈ N , define N − :=_{M ∈ N : M < N} and N + :=^{M ∈ N : M > N} Conventionally 0− = 0 and I + = I. If N > N − then N − N − is an atom of N , and all atoms are of this form. Conversely, if N = N − > 0 then there is a strictly increasing sequence of projections in N which converge to N . Similar remarks apply for N +. We shall make continual use of the fact that the rank-1 operator x 7→ hx, fie, which we write as ef ∗, belongs to T (N ) if and only if there is an N ∈ N such that e ∈ ran(N +) and f ∈ ran(N ⊥). See [2] for further properties of nest algebras. ON THE PRIMITIVE IDEALS OF NEST ALGEBRAS 3 Example 2.1. Let H := ℓ2(N) and let {ei}∞ i=1 be the standard basis. For n ∈ N, let Nn be the projection onto the span of {e1, . . . , en} and let N := {Nn : n ∈ N}∪{I}. This is a nest, and T (N ) is the nest algebra of all infinite upper triangular operators with respect to the standard basis. By slight abuse of notation, we write T (N) for this algebra. We now recall Ringrose's description of the Jacobson radical of a nest algebra, in terms of diagonal seminorms and diagonal ideals: Definition 2.2. Let N be a nest and fix N < I in N . The diagonal seminorm function i+ N (X) is defined for X ∈ T (N ) by i+ N (X) := inf{k(M − N )X(M − N ) : M > N in N} Likewise, for N > 0 the diagonal seminorm function i− N (X) is i− N (X) := inf{k(N − M )X(N − M ) : M < N in N} It is straightforward to see that the functions i± N are submultiplicative seminorms on T (N ) and dominated by the norm, and so their kernels are norm closed two- sided ideals of T (N ): Definition 2.3. Let N be a nest. The diagonal ideals are the ideals and N := {X ∈ T (N ) : i+ I + N := {X ∈ T (N ) : i− I− N (X) = 0} N (X) = 0} (for N < I) (for N > 0) The diagonal ideals can be viewed as generalizations of those ideals of upper- triangular n× n matrices consisting of all the matrices which vanish at a particular diagonal entry. Indeed if N > N − then I− N = {X ∈ T (N ) : (N − N −)X(N − N −) = 0} (1) However if N = N −, then I − to N (from below). More precisely, in the case of T (N), I− all N < I and I − I discussion of the primitive ideals in this algebra. N is the set of operators asymptotically vanishing close N is of the form (1) for is the compact operators of T (N). See Section 6 for a detailed Ringrose gave the following description of the Jaconson radical in terms of these diagonal ideals. Theorem 2.4 ([23], Theorem 5.3). The Jacobson radical of T (N ) is the intersec- tion of the diagonal ideals of T (N ). A key point to bear in mind is that although the diagonal ideals are related to the primitive ideals, as the next result quoted shows, they were not known to be primitive. Lance [13] asked whether the diagonal ideals are primitive and, in his study of the diagonal ideals and their quotients and proved a number of results which are entailed by primitivity. In Theorem 3.7 we show that the diagonal ideals are in fact primitive ideals. The following useful result shows that each primitive ideal of a nest algebra is associated with a unique diagonal ideal. Theorem 2.5 ([23], Theorem 4.9). Every primitive ideal of T (N ) contains exactly one diagonal ideal. Based on this result we adopt the following notation: 4 JOHN LINDSAY ORR Definition 2.6. If P is a primitive ideal of the nest algebra T (N ), write IP for the unique diagonal ideal contained in P. Finally we close the section by recalling Larson's ideal [14], R∞ N : Definition 2.7. Let R∞ N be the set of X ∈ T (N ) such that, given ǫ > 0 we can find a collection {Ni − Mi :∈ N} of pairwise orthogonal intervals of N which sum to I and such that k(Ni − Mi)X(Ni − Mi)k < ǫ. Ringrose [23, Theorem 5.4] provides an alternate description of the Jacobson radical which is formally very similar to Larson's ideal. The only difference is the requirement that the collections of pairwise orthogonal intervals must be finite. However this makes an enormous difference to the size of the ideal as the following example shows. Example 2.8. Let N be the canonical nest on ℓ2(N). Then RN is the set of zero- diagonal compact operators in T (N) and R∞ N is the set of all zero-diagonal operators in T (N). Note in particular that T (N) = D(N)⊕R∞ N but that T (N) 6= D(N)⊕RN (for example, the right-hand side fails to contain the unilateral backward shift). 3. The diagonal ideals are primitive The main result of this section is Theorem 3.7, in which we prove that the diagonal ideals of a nest algebra are primitive. We start by recalling some basic facts about primitive ideals which can be found in many standard texts of ring theory or Banach algebras. See, e.g., [1, Chapter III]. Remark 3.1. Let A be a unital Banach algebra. The (left) primitive ideals of A are the annihilators of left A-modules, or, equivalently, the kernels of the irreducible representations of A. If P is any primitive ideal of A then there is a maximal left ideal L of A such that P is the kernel of the left regular representation of A on A/L. Thus P is the largest two-sided ideal of A contained in L, and is equal to {x ∈ A : xA ⊆ L} From this, together with the maximality of L, it follows easily that x 6∈ P if and only if there are a, b ∈ A such that e − axb ∈ L (where e is the unit of A). Finally, of course, the Jacobson radical is, by definition, the intersection of all the primitive ideals of A. Analogously, the right primitive ideals are the kernels of right A- modules and each right primitive ideal is the kernel of the right module action of A on the quotient A/R of A by some maximal right ideal. The intersection of the maximal right primitive ideals is also the (same) Jacobson radical. Lemma 3.3 will enable us to convert arbitrary upper triangular operators to It relies on the following useful technical lemma which we block diagonal form. quote in full. Lemma 3.2. [19, Lemma 2.2] Let X ∈ B(H) and let Pn, Qn (n ∈ N) be sequences of projections such that dist(PnXQn,F4n−4) > 1 for all n, where Fk denotes the set of operators of rank not greater than k. Then there are orthonormal sequences xi ∈ PiH and yi ∈ QiH such that hxi, Xyji = 0 for all i 6= j, and hxi, Xyii is real and greater than 1 for all i ∈ N. ON THE PRIMITIVE IDEALS OF NEST ALGEBRAS 5 Lemma 3.3. Suppose X ∈ T (N ) but X 6∈ I− N for some N = N − > 0 in N . Then there are A, B ∈ T (N ) and a sequence Nk of nest projections strictly increasing to N such that ∞ AXB = (Nk − Nk−1)AXB(Nk − Nk−1) Xk=1 and each of the terms (Nk − Nk−1)AXB(Nk − Nk−1) has norm greater than 1. Proof. Rescaling if necessary, assume i− N (X) > 1. Choose a sequence Nk ∈ N which increases strictly to N . We shall inductively construct a subsequence Nkn such that dist((Nkn − Nkn−1)X(Nkn − Nkn−1),F4n−4) > 1 for all k, and the result will follow from an easy application of Lemma 3.2. Take k1 := 1 and suppose k1 < k2 < ··· < kn−1 to have been chosen with the desired property. for all k > kn−1. Fix an a with 1 < a < i− Fk ∈ F4n−4 such that Suppose for a contradiction that dist((Nk − Nkn−1)X(Nk − Nkn−1),F4n−4) ≤ 1 N (X) and for each k ≥ kn−1 find k(Nk − Nkn−1)X(Nk − Nkn−1) − Fkk < a The sequence Fk is norm-bounded and so has a w∗-convergent subsequence, Fmj → F . But F ∈ F4n−4 since F4n−4 is w∗-closed and, by the the lower semicontinuity of the norm, i− N (X) ≤ k(N − Nkn−1)X(N − Nkn−1) − Fk j→∞ k(Nmj − Nkn−1)X(Nmj − Nkn−1) − Fmjk ≤ lim inf ≤ a ∞ Xn=1 ∞ Xn=1 which is a contradiction. Thus we find kn > kn−1 with which to continue the induction. With Nkn chosen, apply Lemma 3.2 to obtain unit vectors xn, yn in the range of Nkn − Nkn−1 such that hxm, Xyni = 0 for all m 6= n, and hxm, Xymi > 1 for all m ∈ N. Set A := x3nx∗ 3n+1 and B := y3n+1y∗ 3n+2 Then A, B ∈ T (N ) since the terms of both sums are of the form NkmRN ⊥ , and AXB = P∞ n=1(Nk3n+2 − Nk3n−1)AXB(Nk3n+2 − Nk3n−1) and each of the terms of the sum has norm greater than 1. 3n+2, so that AXB = P∞ n=1hx3n+1, Xy3n+1ix3ny∗ km (cid:3) The following, unfortunately rather technical, definition is central to our analysis in this section. Definition 3.4. Fix a nest N and a projection N ∈ N . Say that a set S of operators in B(H) are of Type-S if there exists a strictly increasing sequence Nn in N which converges to N , and a sequence of unit vectors xn = (Nn − Nn−1)xn such that for each X ∈ S both Xxn → 0 and X ∗xn → 0. Clearly if S ⊆ T (N ) is of Type-S, then it lies in both a proper left ideal of T (N ) and in a proper right ideal of T (N ). Note, however that it need not lie in a proper two-sided ideal; for example consider the singleton {I−U} where U is the unilateral backward shift on ℓ2(N). This is Type-S with respect to the sequences N2n and 6 JOHN LINDSAY ORR 1−n 2 P2n−1 i=2n−1 ei but does not lie in a proper two-sided ideal of T (N). xn := 2 In fact this example is the prototype of the analysis which follows and an analogous sequence is at the heart of the proof of the next lemma. Note also that, strictly speaking, "Type-S" is a property which a set has with respect to a particular N and N ∈ N . In the following arguments these will always be easily discerned from the context. Lemma 3.5. Fix a nest N and a projection N = N − > 0 in N and let {Xi : i ∈ N} be a set of Type-S. Let X ∈ T (N ) but X 6∈ I− N . Then there are A, B ∈ T (N ) such that {I − AXB} ∪ {Xi : i ∈ N} is also of Type-S. Proof. Take a sequence Nn ∈ N which increases strictly to N and unit vectors xn = (Nn − Nn−1)xn such that Xixn, X ∗ By Lemma 3.3 there are A, B in T (N ) and a sequence of nest projections strictly increasing to N such that AXB is block diagonal with respect to these projections and each of the blocks has norm greater than 1. Since Nk and xk demonstrate the Type-S property, so does any subsequence of theirs and so, replacing Nk with a subsequence, we may assume that each interval Nk − Nk−1 dominates a block of AXB. Multiplying AXB by a diagonal projection to select only those blocks which are dominated by an Nk− Nk−1 and have norm greater than 1, and replacing X with the resulting operator we may now assume that X is block diagonal with respect to Nk, and that all the blocks (Nk−Nk−1)X(Nk−Nk−1) have norm greater than 1. We shall inductively construct a new sequence of unit vectors yn = (Nkn − i xn → 0 for all i ∈ N. Nkn−1)yn for a subsequence (kn), together with contractions An = (Nkn − Nkn−1)An(Nkn − Nkn−1 ) and Bn = (Nkn − Nkn−1)Bn(Nkn − Nkn−1) in T (N ) such that and max{kXiynk,kX ∗ taking A :=P∞ max{k(I − AnXBn)ynk,k(I − AnXBn)∗ynk} < 1/n n=1 An and B :=P∞ i ynk} < 1/n for all 1 ≤ i ≤ n. The result will then follow by To perform the induction, fix n and suppose km, ym, Am, and Bm have been chosen for all m < n. (To get the induction started when n = 1, define k0 := 0 and observe that no other features of the preceding steps are used in the induction step which follows.) n=1 Bn. Note that, for all sufficiently large m, max{kXixmk,kX ∗ i xmk} < 1/(2n2) for all 1 ≤ i ≤ n. Thus, taking N = 4n2 we can pick m1 < m2 < ··· < mN such that m1 > kn−1 + 1, each mj > mj−1 + 1, and for all 1 ≤ i ≤ n and 1 ≤ j ≤ N , max{kXixmjk,kX ∗ Set kn := mN and yn := N −1/2PN j=1 xmj , which is a unit vector since the xmj are pairwise orthogonal. For each 1 < j ≤ N , the interval Nmj−1 − Nmj−1 dominates a diagonal block of X which has norm greater than 1. Thus, we can choose vectors ej and fj in Nmj −1 − Nmj−1 with kejk ≥ kfjk = 1 and ej = Xfj i xmjk} < 1/(2n2). ON THE PRIMITIVE IDEALS OF NEST ALGEBRAS 7 and set Since An := N Xj=2 kejk−1 xmj−1 e∗ j and Bn := N Xj=2 kejk−1 fjx∗ mj xmj−1 e∗ j = Nmj−1 (xmj−1 e∗ j )N ⊥ nj−1 and fjx∗ mj = Nmj −1(fjx∗ mj )N ⊥ mj −1, j=2 xmj x∗ mj , so that Xj=2 Xj=1 j=1 xmj+1 x∗ N −1 N j=2 xmj−1 x∗ mj , so that N N n=1 An and B :=P∞ and (AnXBn)∗ =PN xmj−1(cid:13)(cid:13)(cid:13) xmj+1(cid:13)(cid:13)(cid:13) each of the terms of the sums are in T (N ), and the ranges and cokernels of the terms are pairwise orthogonal, so that both sums converge strongly. Now clearly for each 1 ≤ i ≤ n, kXiynk ≤ N −1/2PN j=1 kXixmjk < N 1/2/2n2 = 1/n and, likewise i ynk < 1/n. Further, AnXBn =PN kX ∗ k(I − AnXBn)ynk = N −1/2(cid:13)(cid:13)(cid:13) Xj=1 xmj − mj−1 =PN −1 k(I − AnXBn)∗ynk = N −1/2(cid:13)(cid:13)(cid:13) Xj=1 xmj − Note also that each of the ej, fj, xmj for 1 ≤ j ≤ N lie in the range of NmN − Nm1−1 ≤ Nkn − Nkn−1. Thus An = (Nkn − Nkn−1)An(Nkn − Nkn−1), Bn = (Nkn − Nkn−1)Bn(Nkn − Nkn−1), and yn = (Nkn − Nkn−1)yn. Having met all the requirements, the induction proceeds as stated, and we let A :=P∞ n=1 Bn. Clearly for any fixed i ∈ N, max{kXiynk,kX ∗ i ynk} < 1/n for all sufficiently large n and so Xiyn, X ∗ i yn → 0. Moreover since A and B are block diagonal with respect to Nnk , as is X, it follows that (I − AXB)yn = (I − AnXBn)yn → 0 and (I − AXB)∗yn = (I − AnXBn)∗yn → 0 and we are done. (cid:3) Lemma 3.6. Fix a nest N and a projection N ∈ N and let Si (i ∈ N) be a countable collection of countable sets of Type-S which form a chain (i.e. for any i, j, either Si ⊆ Sj or Sj ⊆ Si). Then Si∈N Si is also of Type-S. topology on N is metrizable; let d be a metric for it. Enumerate Si∈N Si and let the sets Cn (n ∈ N) consist of the first n terms of that enumeration. Fix n and suppose Nm and xm have been chosen for m < n so that N1 < N2 < ··· < Nn−1 < N , xm = (Nm − Nm−1)xm, and max{kXxmk,kX ∗xmk} < 1/m for all X ∈ Cm. Each X ∈ Cn belongs to some Si and since Cn is finite and {Si} is a chain, Cn is contained in some Si. Therefore Cn is of Type-S. Using this fact, we can find Nn−1 < Nn < N with d(Nn, N ) < 1/n and xn = (Nn − Nn−1)xn such that max{kXxnk,kX ∗xnk} < 1/n for all X ∈ Cn. Continue this inductively to construct a strictly increasing sequence Nn → N and xn = (Nn − Nn−1) for all n ∈ N such that max{kXxnk,kX ∗xnk} < 1/n for all X ∈ Cn (taking k0 = 0 to get the induction started). Each X ∈Si∈N Si belongs to Cn for all sufficiently large n, and so the result follows with the vectors so chosen. Theorem 3.7. Assume the Continuum Hypothesis and let N be a nest. Then the diagonal ideals of T (N ) are primitive ideals. Proof. The proof is a routine countability argument. Recall that the strong operator = N −1/2 < 1/n = N −1/2 < 1/n. (cid:3) 8 JOHN LINDSAY ORR N . Thus P ⊆ I− Next, let I be a diagonal ideal of T (N ) and suppose that I = I− Proof. The result is trivial when the diagonal ideal is of type I− N with N − < N or I + N with N + > N . For in either case the diagonal ideal is the kernel of the representation X 7→ EXEH where E is an atom of N and whose range is therefore all of B(EH), and so is irreducible. For the remainder of the proof, consider only diagonal ideals which are not of this type. N for some N = N − > 0 in N . It is enough to construct operators AX , BX ∈ T (N ) for each operator X ∈ T (N )\I− N , such that the collection {I − AX XBX : X ∈ T (N )\I− N} generates a proper left ideal of T (N ). For then there is a maximal left ideal L which contains this family of operators and the kernel of the left regular representation of T (N ) on T (N )/ L is a primitive ideal, P, which by Remark 3.1 must exclude all X 6∈ I− N . Since every primitive ideal contains a diagonal ideal [23, Theorem 4.9] and the distinct diagonal ideals are incomparable [23, Lemma 4.7], it follows that I− N for some N = N + < I in N . By the same reasoning, it is enough to find AX , BX ∈ T (N ) such that {I − AX XBX : P ∈ T (N )\I + N} is contained in a proper left ideal of T (N ). To do this, we take adjoints and seek AX , BX ∈ T (N )∗ = T (N ⊥) such that {I − AX XBX : X ∈ T (N ⊥)\I− N ⊥} is contained in a proper right ideal of T (N ⊥). Since N is an arbitrary nest, we can replace N ⊥ with N , to recast this as a second problem about I− N in T (N ): namely, to find AX , BX ∈ T (N ) for each X ∈ T (N ) \ I− N , such that {I − AX XBX : X ∈ T (N ) \ I− N} generates a proper right ideal of T (N ). We shall show in fact that the choice can be made so that the same set of operators {I − AX XBX} serves to generate both a proper left ideal and a proper right ideal. We shall construct these operators using transfinite recursion. N = P and so is primitive. Now consider the case when I = I + The cardinality of T (N ) \ I− N is equal to the cardinality of the contiuum since every operator can be represented as a countable array of complex numbers. Since we are assuming the Continuum Hypothesis, T (N ) \ I− N has cardinality ℵ1 and so it can be put in bijective correspondence with the set of ordinals a < ω1 (where ω1 denotes the first uncountable ordinal). Write this correspondence as Xa (a < ω1). To run the transfinite recursion, we suppose that for some a < ω1 we have operators Ab, Bb in T (N ) for all b < a, and describe how to obtain Aa, Ba. First, if the set {I − AbXbBb : b < a} is of Type-S then observe that {I − AbXbBb : b < a} is a countable collection and use Lemma 3.5 to find Aa, Ba ∈ T (N ) such that {I − AbXbBb : b ≤ a} is also of Type-S. On the other hand, if it happens that {I − AbPbBb : b < a} is not of Type-S then set Aa = Ba = 0. (This is a sink terminal state which we shall prove momentarily is never in fact reached.) Note that formally Lemma 3.5 assumes a countably infinite collection of prede- cessors. However the case of finite a, or even a = 1, can be covered by padding the collection of predecessors with countably many repeated zeros. Note also the recursion step involves an arbitrary choice of operators, which can easily be resolved using the Axiom of Choice. Having described a rule to construct Aa, Ba with (Ab, Bb)b<a given, we apply the principle of transfinite recursion to obtain (Aa, Ba)a<ω1 where the transition rule from the previous paragraph applies for every a < ω1. We next note that for every a < ω1, Sa := {I − AbXbBb : b ≤ a} is of Type-S. For if this were not true, then we could find the least a such that Sa is not Type-S. Thus for each of the countably many b < a, Sb is countable and of Type-S and so by Lemma 3.6, ON THE PRIMITIVE IDEALS OF NEST ALGEBRAS 9 Sb<a Sb is Type-S. But Sb<a Sb = {I − AbXbBb : b < a} and so by the recursion step, Sa = {I − AbXbBb : b ≤ a} is also Type-S. Thus, by contradiction, each Sa is of Type-S and, in particular, generates a proper left ideal of T (N ) and a proper right ideal of T (N ). Now in general, the union of any chain of sets, each of which generates a proper left (resp. right) ideal, will also generate a proper left (resp. right) ideal. Thus, {I − AaXaBa : a < ω1} = Sa<ω1 Sa generates a proper left ideal and a proper right ideal, and the result follows. Corollary 3.8. Assuming the Continuum Hypothesis, the diagonal ideals of T (N ) are also right-primitive ideals, that is to say, the annihilators of simple right mod- ules. Proof. The conjugate-linear anti-isomorphism X 7→ X ∗ maps T (N ) to T (N ⊥), maps diagonal ideals to diagonal ideals, and converts left modules into right mod- ules. (cid:3) (cid:3) We remark in passing that Theorem 3.7 does provide a new proof of Ringrose's characterization of the Jacobson radical of a nest algebra. For in view of Theo- rem 2.5 \I diagonal I ⊆ \P primitive P, and the reverse inclusion follows from Theorem 3.7. Insofar as our result assumes the Continuum Hypothesis and also assumes H is separable, this is, of course, substantially less general than Ringrose's original proof. 4. The left ideals of a nest algebra In this section we study the left ideals of nest algebras. Definition 4.1 gives a method of specifying left ideals and in Theorem 4.4 we shall see that every left ideal can be specified in this way. We then introduce (Definition 4.7) a stronger property which specifies many closed left ideals, including the maximal left ideals. This leads to insights into the structure of left ideals (Proposition 4.18) which we apply in the following sections. Definition 4.1. Let L be a left ideal of T (N ). Say that L is constructible if there is a net indexed by a directed set A consisting of pairs (Nα, xα) of projections Nα ∈ N and vectors xα ∈ H such that L = {X ∈ T (N ) : lim α∈Ak(I − Nα)Xxαk = 0} for every X ∈ T (N ). Lemma 4.2. T (N ) is itself a constructible ideal and, in general, the constructible ideal, L, specified by the net (Nα, xα)α∈A is equal to T (N ) if and only if limα N ⊥ α xα = 0. Proof. If N ⊥ kXkkN ⊥ proper. α xα → 0 then, for any fixed X ∈ T (N ), kN ⊥ α xαk and so X ∈ L. Conversely, if N ⊥ α Xxαk = kN ⊥ α xαk ≤ α xα 6→ 0, then I 6∈ L and so L is α XN ⊥ (cid:3) N ⊥ Note that if (Nα, xα)α∈A is a net in N × H and X ∈ T (N ), then N ⊥ α Xxα = α XN ⊥ α xα. The following interpolation result of Katsoulis, Moore, and Trent enables us to see that all left ideals are constructible. In this context we remark that the results of α xα, for all α and so without loss we can always assume that xα = N ⊥ 10 JOHN LINDSAY ORR [11] have a precursor in Lance's [13, Theorem 2.3], introduced to study the radical and diagonal ideals. Theorem 4.3. [11, Theorem 4]. Let X1, . . . , Xn and Y be in T (N ). Then there are A1, . . . , An in T (N ) such that n AiXi Y = Xi=1 if and only if (2) sup(cid:26) kN ⊥Y xk2 i=1 kN ⊥Xixk2 : N ∈ N , x ∈ H(cid:27) < ∞ Pn (where 0/0 is interpreted as 0). Theorem 4.4. Every left ideal of a nest algebra is constructible. Proof. Let L be a fixed left ideal of the nest algebra T (N ) and take A to be the set of all 4-tuples (F, ǫ, N, x) where F is a finite subset of L, ǫ > 0, N ∈ N , and x ∈ H, subject to the constraint that kN ⊥Xxk < ǫ for all X ∈ F . This is a directed set if we say (F, ǫ, N, x) ≤ (F ′, ǫ′, N ′, x′) when F ⊆ F ′ and ǫ ≥ ǫ′. For the relation is clearly reflexive and transitive, and any pair of members of A, (F, ǫ, N, x) and (F ′, ǫ′, N ′, x′), is dominated by (F ∪ F ′, min{ǫ, ǫ′}, 0, 0). Define a net on A with values in N × H by the mapping which takes α := (F, ǫ, N, x) ∈ A to (Nα, xα) where Nα := N , and xα := x. We shall see that this net specifies L exactly. On the one hand, trivially, if X ∈ L then for any ǫ > 0, the tuple α0 := ({X}, ǫ, 0, 0) belongs to A and so for any α ≥ α0 kN ⊥ α Xxαk < ǫ. So, next, suppose on the other hand that Y ∈ T (N ) \ L. Let an arbitrary α0 := ({X1, . . . , Xn}, ǫ, M, x) in A be given. Since Y 6∈ L, there do not exist any A1, . . . , An in T (N ) such that Pn i=1 AiXi = Y . Thus by Theorem 4.3, the supremum (2) is infinite, and so we can find N ∈ N and y ∈ H such that kN ⊥Xiyk < ǫkN ⊥Y yk (cid:3) β Y xβk = 1. In other words, the net kN ⊥ α Y xαk 6→ 0. for each i = 1, . . . , n. Rescaling y, we obtain N and y such that kN ⊥Xiyk < ǫ and kN ⊥Y yk = 1. Thus β := ({X1, . . . , Xn}, ǫ, N, y) is in A, and we have β ≥ α0 and kN ⊥ α Y xαk is frequently equal to 1, and so kN ⊥ Example 4.5. The set FN of finite rank operators in T (N ) is a two-sided ideal of T (N ) but is not norm-closed. We can specify this with the following net. Let A consist of the set of pairs (F, x) where F is a finite-dimensional subspace of H and x is a vector which is orthogonal to F . For α = (F, x) ∈ A define xα := x and Nα = 0. Say (F, x) ≤ (G, y) in A if F ⊆ G. Clearly T ∈ T (N ) belongs to FN if and only if there is a finite-dimensional space F such that T vanishes on F ⊥. Since the vectors in the pairs are unbounded, the condition kN ⊥ α T xαk < 1 for all α ≥ (F, 0) is equivalent to T vanishing on F ⊥. Example 4.6. The set KN of compact operators in T (N ) is a norm-closed two- sided ideal of T (N ). We can specify it with the following net, which is similar to the previous example. Let A consist of the set of pairs (F, x) where F is a finite- dimensional subspace of H and x is a unit vector which is orthogonal to F . Again for α = (F, x) ∈ A define xα := x and Nα = 0, and say (F, x) ≤ (G, y) in A if ON THE PRIMITIVE IDEALS OF NEST ALGEBRAS 11 F ⊆ G. By the spectral theory, an operator T ∈ T (N ) belongs to KN if and only if for any ǫ > 0 there is a finite-dimensional space F such that kTF ⊥k < ǫ, which is readily seen to be equivalent to kN ⊥ α T xαk → 0. The contrast between the last two examples, in which the net was unbounded in one case and bounded in the other, motivates the following definition. Definition 4.7. Let L be a left ideal of T (N ). Say that L is strongly constructible if it is constructible and a net (Nα, xα)α∈A specifying L can be found in which all the vectors xα = N ⊥ α xα have norm 1. Proposition 4.8. Strongly constructible ideals are norm-closed. Proof. Let L be strongly constructible and specified by (Nα, xα)α∈A, where kxαk = 1 for all α ∈ A. Suppose the sequence of Xn ∈ L converges in norm to X ∈ T (N ). Given ǫ > 0, find a fixed n ∈ N such that kX − Xnk < ǫ/2 and α0 ∈ A such that kN ⊥ α Xnxαk < ǫ/2 for all α ≥ α0. Then kN ⊥ α Xxαk ≤ kX − Xnkkxαk + kN ⊥ α Xnxαk < ǫ (cid:3) Proposition 4.9. The maximal left ideals of T (N ) are strongly constructible. Proof. Let L be a maximal left ideal which we suppose to be specified by the net (Nα, xα)α∈A. Without loss, assume that each xα = N ⊥ α xα. By Lemma 4.2, xα 6→ 0 and so there is an ǫ0 > 0 such that kxαk is frequently at least ǫ0. Let A′ := {α ∈ A : kxαk ≥ ǫ0} and x′ α = xα/kxαk for α ∈ A′. Now A′ is a directed set and (Nα, xα) is a net on it. Again by Lemma 4.2, the net (Nα, x′ α)α∈A′ specifies a proper ideal which, furthermore, contains L since for X ∈ L, α Xx′ kN ⊥ αk ≤ 1 ǫ0 kN ⊥ α Xxαk for all α ∈ A′ and the net on the right converges to zero since (Nα, xα)α∈A′ is a subnet of (Nα, xα)α∈A. By maximality, the ideal which (Nα, x′ α)α∈A′ specifies must equal L. (cid:3) Proposition 4.10. Arbitrary intersections of strongly constructible ideals are strongly constructible. The proof is a consequence of the following simple result about nets. Lemma 4.11. Fix a set X and suppose that we have a family of nets in X indexed by a set K, which we denote by (x(k) α )a∈Ak . Then we can find a net (xα)α∈A in X with the property that for any E ⊆ X, (xα)α∈A is eventually in E if and only if for each k ∈ K, (x(k) α )α∈Ak is eventually in E. Proof. Define A to be set the set of pairs (σ, k) where σ is a section map on the fibre bundle of Ak over K (i.e., for each k ∈ K, σ(k) ∈ Ak), and k is an arbitrary member of K. Put a relation on A by declaring (σ, k) ≤ (τ, l) if σ(i) ≤i τ (i) for all i ∈ K (the relation ≤i is the directed relation defined on Ai). This is a symmetric and transitive relation. Moreover, if (σ, k) and (τ, l) are in A then for each i ∈ K we can find an element of Ai which dominates both σ(i) and τ (i). By the Axiom of Choice there is therefore a section map ρ such that ρ(i) dominates both σ(i) and τ (i) for all i ∈ K. Taking an arbitrary i ∈ K, then (ρ, i) dominates both (σ, k) 12 JOHN LINDSAY ORR and (τ, l) in A. Thus A is a directed set, and we define the net (x(σ,k))(σ,k)∈A by x(σ,k) := x(k) σ(k). Now, on one hand, suppose that (x(σ,k)) is eventually in E ⊆ X. Thus there is a (σ0, k0) ∈ A such that x(σ,k) ∈ E for all (σ, k) ≥ (σ0, k0). Fix k ∈ K and consider α0 := σ0(k) ∈ Ak. If α ≥k α0 then define σ(i) := σ0(i) for all i 6= k and σ(k) := α. Then (σ, k) ≥ (σ0, k0) and so x(k) α = x(σ,k) ∈ E. This shows that for each k ∈ K, (x(k) α )α∈Ak is eventually in E. Conversely, let E ⊆ X and suppose that for every k ∈ K, (x(k) α )α∈Ak is eventually in E. That is to say, for each k ∈ K, we can can find an α0 ∈ Ak such that x(k) α ∈ E for all α ≥k α0 in Ak. Again by the Axiom of Choice we pick one such α0 for each k ∈ K and obtain a section σ0 such that for each k ∈ K and α ≥k σ0(k) in Ak, we have x(k) α ∈ E. Pick an arbitrary k0 ∈ K and then suppose (σ, k) ≥ (σ0, k0). This means that, in particular, σ(k) ≥k σ0(k), so that x(σ,k) = x(k) σ(k) ∈ E. We conclude that the net (x(σ,k))(σ,k)∈A is eventually in E. (cid:3) The proof of Proposition 4.10 now follows straightforwardly. Proof (of Proposition 4.10). Let Lk (k ∈ K) be a collection of strongly constructible left ideals. Writing H1 for the set of unit vectors in H, for each k ∈ K there are directed sets Ak and nets (N (k) α ) ∈ N ×H1 for α ∈ Ak such that an X ∈ T (N ) belongs to Lk if and only if limα∈Ak k(I − N (k) By Lemma 4.11, find a new net (Nα, xα)α∈A in N × H1 which is eventually in a subset of N × H1 if and only if each of the (N (k) α )α∈Ak are eventually in that set. Fix X ∈ T (N ) and let ǫ > 0 be given. Let α k = 0. α )Xx(k) α , x(k) α , x(k) Eǫ := {(N, x) ∈ N × H1 : k(I − N )Xxk < ǫ} constructible. α )α∈Ak is eventually in Eǫ. This happens iff for every ǫ > 0, (Nα, xα)α∈A is eventually in Eǫ, Clearly X ∈ Tk∈K Lk iff for every k ∈ K and every ǫ > 0, (N (k) which in turn happens iff limα∈A k(I − Nα)Xxαk = 0. Thus Tk∈K Lk is strongly L of T (N ) is contained in a smallest Corollary 4.12. Every proper left ideal strongly constructible left ideal, which we shall call the strongly constructible hull of L. α , x(k) (cid:3) Corollary 4.13. The primitive ideals of T (N ) are strongly constructible. Proof. Every primitive ideal is the intersection of the maximal left ideals which contain it [1, §24, Proposition 12 (iv)]. The result follows by By Propositiona 4.9 and 4.10. (cid:3) Example 4.14. In particular, the maximal two-sided ideals of T (N ), being prim- itive, are strongly constructible. Recall that the strong radical of a unital algebra is the intersection of all its maximal two-sided ideals. In [17, Theorem 3.2] we saw that if T (N ) is a continuous nest algebra then any norm-closed, two-sided ideal of T (N ) which contains the strong radical is the intersection of the maximal two- sided ideals which contain it. Thus by Propositions 4.10 and 4.13, all such ideals are strongly constructible. ON THE PRIMITIVE IDEALS OF NEST ALGEBRAS 13 Corollary 4.15. All norm-closed, two-sided ideals of a continuous nest algebra which contain the strong radical are strongly constructible. Question 4.16. Is every norm-closed left ideal of a nest algebra strongly con- structible? Strongly constructible ideals are also characterized by two ostensibly weaker conditions: Proposition 4.17. Let L be a proper left ideal of T (N ). The following are equiv- alent: (1) L is strongly constructible. (2) L can be specified by a net (Nα, xα)α∈A where kxαk ≤ 1 for all α ∈ A. (3) L can be specified by a net (Nα, xα)α∈A where kxαk is bounded. Proof. Clearly (1) ⇒ (2) ⇒ (3) and so it remains to prove (3) ⇒ (1). Suppose (Nα, xα)α∈A specifies L and kxαk is bounded. Since L is proper, by Lemma 4.2 xα 6→ 0 and so there is an ǫ0 such that kxαk is frequently at least ǫ0. For each k ∈ N set Ak := {α ∈ A : kxαk ≥ ǫ0/k} Each Ak is a directed set (with the order relation inherited from A) and the re- stricted net (Nα, xα)α∈Ak defines a left ideal Lk. Since the xα are bounded away from zero on Ak, we can normalize and see each Lk is strongly constructible. It tion 4.10. remains to check that L = Tk∈N Lk and then the result will follow by Proposi- Clearly since each Ak ⊆ A, also L ⊆ Lk and so L ⊆ Tk∈N Lk. Suppose X 6∈ L. Then (I − Nα)Xxα 6→ 0 and so there is an ǫ1 > 0 such that k(I − Nα)Xxαk ≥ ǫ1 frequently. Choose k > ǫ0kXk/ǫ1 so that then whenever k(I − Nα)Xxαk ≥ ǫ1 then kXkkxαk ≥ k(I − Nα)Xxαk ≥ ǫ1 > kXk ǫ0 k and thus α ∈ Ak. It follows that k(I − Nα)Xxαk ≥ ǫ1 frequently on Ak, and so X 6∈ Lk. Proposition 4.18. Let L be a maximal left ideal in T (N ) and let Pn be a sequence of pairwise orthogonal projections in L. There is a subsequence Pnk such that the projection P∞ Proof. By Proposition 4.9, L is strongly constructible, say by a net (Nα, xα) where each xα is a unit vector in the range of Nα. By Kelley's Theorem, this net has a universal subnet, which specifies a proper ideal containing L, hence in fact specifies L itself. Thus we may assume (Nα, xα) is universal. n=1 Pnk belongs to L. (cid:3) The proof now proceeds by means of a fairly routine diagonal argument. For sets, S′ 0 and S′′ 0)xαk and kP (S′′ any S ⊆ N write P (S) := Pn∈S Pn. Take S0 := N and split S0 into two infinite 0 . If kP (S′ 0 )xαk are each eventually greater than 1/√2 then kP (S0)xαk2 = kP (S′ 0 )xαk2 is eventually greater than 1, which is impossible. Since (Nα, xα) is universal that means at least one of 0 )xαk is eventually no greater than 1/√2; without loss suppose kP (S′ that kP (S′ as before, we conclude that at least one of kP (S′ greater than (1/√2)2. Take S2 to be one of S′ 1, S′′ 1 in the same way as the union of infinite subsets and, 1)xαk, kP (S′′ 1 )xαk is eventually no 1 for which this holds. Proceeding 0)xαk ≤ 1/√2 eventually, and set S1 := S′ 0)xαk2 + kP (S′′ Now decompose S1 = S′ 0)xαk, kP (S′′ 1∪S′′ 0. 14 JOHN LINDSAY ORR in this way we obtain a sequence S0 ⊇ S1 ⊇ S2 ⊇ . . . of infinite subsets of N such that for each k, eventually kP (Sk)xαk ≤ (1/√2)k. Now take nk to be the kth element of Sk in order, which is a strictly increasing sequence, and let S := {nk}. Thus S \ Sk is finite for all k. Finally, write P := P (S) and, given ǫ > 0, take k such that (1/√2)k < ǫ. For all sufficiently large α kP xαk = k(P P (Sk) + P P (Sk)⊥)xαk ≤ kP (Sk)xαk + kP (S \ Sk)xαk ≤ ǫ + Xn∈S\Sk kPnxαk α P xαk = kP xαk → 0, so that P ∈ L. But the sum in the last line is finite and so is eventually less than ǫ. We can conclude kN ⊥ (cid:3) Corollary 4.19. Let J be a maximal right ideal in T (N ) and let Pn be a sequence of pairwise orthogonal projections in R. There is a subsequence Pkn such that the projection P∞ Proof. The result follows on taking adjoints and working in T (N ⊥). k=1 Pkn belongs to R. (cid:3) 5. Atomic nest algebras In this section we shall focus on atomic nest algebras and relate the character of primitive ideals to the family of diagonal operators they contain. Observe that if P is a primitive ideal of T (N ) then P ∩ D(N ) is a norm-closed two-sided ideal of the C∗-star algebra D(N ) and is therefore a ∗-ideal. In many interesting cases the nest is multiplicity-free so that D(N ) is an abelian C∗-algebra. Proposition 5.1. Let N be an atomic nest and J a two-sided ideal in T (N ). Then J is a maximal two-sided ideal if and only if J ∩ D(N ) is a maximal two-sided ideal of D(N ). Proof. Suppose J is maximal. Then by [20, Theorem 3.8], J contains R∞ It N . follows that J = (J ∩ D(N )) ⊕ R∞ N . If J ∩ D(N ) is not maximal then there is a larger proper ideal D0 of D(N ). But then D0 ⊕ R∞ N is a proper ideal of T (N ) and strictly larger than J , contrary to fact. Suppose on the other hand that J ∩ D(N ) is maximal. By [6, Theorem 10.2] R∞ N is generated as a two-sided ideal by a generator which is the sum of three commutators [Gi, Pi] (i = 1, 2, 3) where Gi ∈ T (N ) and Pi is a projection in the core C(N ) of T (N ). (Recall that the core of a nest algebra is the abelian von Neumann algebra generated by N .) Now since J ∩ D(N ) is a maximal ideal of D(N ), and the Pi are in the centre of D(N ), it follows that one of Pi, P ⊥ i must lie in J ∩ D(N ) for each i. Thus in any event the commutators [Gi, Pi] = [Gi, P ⊥ i ] belong to J and so J contains R∞ N . Thus, again, J = (J ∩ D(N )) ⊕ R∞ N . If J is not maximal then there is a larger proper ideal J0 of T (N ). But then since J0 also contains R∞ N and so J0 ∩ D(N ) is a proper ideal of D(N ) and larger than J ∩ D(N ), contrary to fact. N , J0 = (J0 ∩ D(N )) ⊕ R∞ (cid:3) The proof of Proposition 5.1 is deceptively straightforward. In fact the result cited from [20] depends on Marcus, Spielman, and Srivastava's proof [15] of the Paving Theorem. Recall (Definition 2.6) that we write IP for the unique diagonal ideal contained by the primitive ideal P. ON THE PRIMITIVE IDEALS OF NEST ALGEBRAS 15 Proposition 5.2. Let N be an atomic nest, let P be a primitive ideal of T (N ), and suppose P 6= IP . Then there are non-zero projections in P \ IP . Proof. We shall prove the result in the case when IP = I− N for some N > 0 in N . If, instead, IP = I + N for some N < I then we take adjoints and apply the result to I− N ⊥ ( P ∗ ⊆ T (N )∗ = T (N ⊥). In this case P is a right primitive ideal of T (N ⊥) and so we shall take care that our proof accommodates the case when P is either left or right primitive. If P is a left primitive ideal, let J be a maximal left ideal such that P is the kernel of the left regular module action of T (N ) on T (N )/J . In the case that P is right primitive, let J be a maximal right ideal such that P is the kernel of the right regular module action of T (N ) on T (N )/J . Suppose that N − < N . Note that rank(N − N −) cannot be finite for if it were then IP = I− N would be a maximal ideal of T (N ) and so P = IP , contrary to hypothesis. If rank(N − N −) = ∞ then the only proper ideal strictly containing I− N is {X ∈ T (N ) : (N − N −)X(N − N −) is compact}, which must therefore equal P. Any finite rank projection of the form P = (N − N −)P (N − N −) will serve to establish the result in this case. For the remainder of the proof, assume that N = N − and take X ∈ P \ I− N . By Lemma 3.3 there are A, B ∈ T (N ) such that AXB is block diagonal with respect to some sequence Mk of nest projections strictly increasing to N and each of the blocks has norm greater than 1. Replacing X with AXB we can assume X = P∞ k=1(Mk − Mk−1)X(Mk − Mk−1) where the norm of each term is greater Consider the sequence of intervals M2k+1 − M2k. These are each in I− N and so in J . By Proposition 4.18 and Corollary 4.19, whether J is assumed to be maximal right or maximal left, there is a subsequence kn such that J contains P∞ n=1 M2kn+1 − M2kn . Then for each n find an atom N + n − Nn ≤ M2kn+1 − M2kn. Choose vectors en, fn, gn such that ene∗ n − Nn and fn and gn are in the range of M2kn+2 − M2kn+1 with kfnk > kgnk = 1 and fn = Xgn. Thus, n! X ∞ Xn=1 n+1 = ∞ Xn=1 kfnk−1enf ∗ kfnk−1gne∗ n+1! n ≤ N + Xn=1 than 1. ene∗ V := ∞ where both of the sums converge strongly and are in T (N ) because enf ∗ n = M2kn+1(enf ∗ n)M ⊥ 2kn+1 and gne∗ n+1 = M2kn+2(gne∗ n+1)M ⊥ 2kn+1 = M2kn+2(gne∗ n+1)M ⊥ 2kn+2 since kn+1 ≥ kn + 1. Thus V ∈ P. Let P :=P∞ 2k − N2k, which is dominated by a projection in J and so is also in J . We shall show that P ∈ P. Suppose for a contradiction that P 6∈ P. It follows, as observed in Remark 3.1, that there are A, B ∈ T (N ) such that I − AP B ∈ J . We can assume that A = AP and B = P B. Write A = A1 + A2 where 2k ≤P∞ k=1 e2ke∗ k=1 N + A1 := ∞ Xk=1 N ⊥ 2k−1A(N + 2k − N2k) and A2 := A − A1 16 JOHN LINDSAY ORR so that A2(N + 2k − N2k). Likewise, write B = B1 + B2 where 2k − N2k) = N2k−1A(N + B1 := (N + ∞ Xk=1 2k − N2k)BN2k+1 2k − N2k)BN ⊥ and B2 := B − B1 so that (N + 2k − N2k)B2 = (N + 2k+1. The sums for A1 and B1 con- verge strongly because the sequences of terms are norm-bounded and have pairwise orthogonal ranges and cokernels. Now set A′ that A′ 2 and B′ 2 := A2V ∗ and B′ 2 are in T (N ) since the terms of the sums are in T (N ): 2 := V ∗B2. From the following computations we see A′ 2 = A2P V ∗ = B′ 2 = V ∗P B2 = A2(N + 2k − N2k)V ∗ = V ∗(N + 2k − N2k)B2 = ∞ ∞ Xk=1 Xk=1 ∞ ∞ Xk=1 Xk=1 Furthermore, since V V ∗ =P∞ we have that k=1 eke∗ A2 = A2P = A2P V ∗V = A′ and N2k−1A(N + 2k − N2k)V ∗N ⊥ 2k−1 N + 2k+1V ∗(N + 2k+1 2k − N2k)BN ⊥ k+1 =P∞ k=1 ek+1e∗ k=2 eke∗ k, k and V ∗V =P∞ 2V ∈ P B2 = P B2 = V V ∗P B2 = V B′ 2 ∈ P Since I − (A1 + A2)P (B1 + B2) ∈ J , it now follows that also I − A1P B1 ∈ J . Now note that A1P B1 = = = = ∞ ∞ ∞ Xk=1 Xk=1 Xk=1 Xk=1 ∞ A1(N + 2k − N2k)B1 N ⊥ 2k−1A(N + 2k − N2k)BN2k+1 (N + 2k − N2k−1)A(N + 2k − N2k)B(N2k+1 − N2k) (N + 2k − N2k−1)Ck(N2k+1 − N2k) where Ck := A(N + 2k − N2k)B. We can decompose A1P B1 in two ways, either as ∞ Xk=1 ∞ or as Xk=1 (N + 2k − N2k)Ck(N2k+1 − N2k) + Xk=1 ∞ (N2k − N2k−1)Ck(N2k+1 − N2k) (N + 2k − N2k−1)Ck(N + 2k − N2k) + (N + 2k − N2k−1)Ck(N2k+1 − N + 2k) ∞ Xk=1 These two cases are of the form P Y + Z and Y P + Z respectively where in both cases Z is nilpotent. Recall that P ∈ J and so, whether J is a maximal left ideal or a maximal right ideal, we conclude that I − Z ∈ J , which is impossible since this is invertible and J is proper. From this contradiction we conclude that P ∈ P. (cid:3) Theorem 5.3. Let N be an atomic nest and let P be a primitive ideal of T (N ). ON THE PRIMITIVE IDEALS OF NEST ALGEBRAS 17 (1) If P ∩ D(N ) is a maximal two-sided ideal of D(N ) then P is a maximal (2) If P ∩ D(N ) is equal to I ∩ D(N ) for some diagonal ideal I then P is a two-sided ideal of T (N ). diagonal ideal and, in fact, P = I. Proof. Case (1) is just Proposition 5.1. To prove Case (2), suppose that P∩D(N ) = I ∩ D(N ) for some diagonal ideal I. First observe that IP ∩ D(N ) ⊆ P ∩ D(N ) = I∩D(N ). Now, distinct diagonal ideals contain complementary projections (see the proof of [23, Lemma 4.8] for this fact) and so I must equal IP . But now if P 6= IP then by Proposition 5.2, P contains projections which are not in IP , contrary to hypothesis. (cid:3) We can now distinguish three classes of primitive ideals based on the diagonal operators they contain. The first class (Πmax) consists of primitive ideals for which P ∩ D(N ) is a maximal ideal of D(N ), and this consists of the maximal two-sided ideals of T (N ). The second class (Πmin) consists of primitive ideals for which P ∩ D(N ) = I ∩D(N ) for some diagonal ideal I and this class consists of diagonal ideals. The third class (Πint) consists of the remaining primitive ideals for which P ∩ D(N ) takes neither its minimal nor its maximal values. The maximal ideals of a general nest algebra were completely described in [20, Corollary 3.10]. In particular when N is atomic the ideals in Πmax are precisely the ideals of the form D0 ⊕ R∞ N where D0 is a maximal two-sided ideal of D(N ). The ideals in Πmin are the primitive ideals which are also diagonal ideals. Trivially all ideals of the form I− N where N < N +) are included in this class. (See the first paragraph of the proof of Theorem 3.7 for details.) By Theorem 3.7, if we assume the Continuum Hypothesis then Πmin consists of all the diagonal ideals. Without the assumption of the Continuum Hypothesis we cannot say which additional diagonal ideals belong to Πmin. The structure of Πint is more delicate. In the following section we will see examples of representatives of all three classes. N where N > N − (or, equivalently, I + 6. The infinite upper triangular operators Throughout this section, let H = ℓ2(N) and consider the algebra T (N) of all upper triangular operators with respect to the standard basis of ℓ2(N). Recall that we write {ei}∞ i=1 for the standard basis and let Nn be the projection onto the span of {e1, . . . , en}, and N := {Nn : n ∈ N}∪{0, I}. Then T (N) := T (N ) is the algebra of infinite upper triangular operators with respect to the ei and R∞ N is simply the ideal of infinite strictly upper triangular operators. Moreover, the diagonal ideals of T (N) are precisely the ideals I1,I2,I3, . . . ;I∞ where In := I− for 1 ≤ n < ∞ and I∞ := I− I . Note that I∞ coincides with the compact operators of T (N), a fact which we shall develop below. Nn 6.1. The quasitriangular algebra. Let K(H) be the set of all compact operators in B(H) and write QT (N) for the quasitriangular algebra T (N)+K(H). By [10] and, in more generality, [7], QT (N) is a norm-closed algebra in B(H) and the canonical isomorphism between QT (N)/K(H) and T (N)/(T (N) ∩ K(H)) is isometric. Corollary 6.1. Assuming the Continuum Hypothesis, T (N)/(T (N) ∩ K(H)) is a left (resp. right) primitive algebra. 18 JOHN LINDSAY ORR Proof. K(H) ∩ T (N) = I∞, which is a left primitive ideal by Theorem 3.7 and a right primitive ideal by Corollary 3.8. (cid:3) Corollary 6.2. Assuming the Continuum Hypothesis, QT (N)/K(H) is a left (resp. right) primitive algebra, and K(H) is a left (resp. right) primitive ideal in QT (N). 6.2. A catalogue of primitive ideals. Clearly Πmin contains {I1,I2, . . .}. As- suming the Continuum Hypothesis then by Theorem 3.7 Πmin = {I1,I2, . . .} ∪ {I∞} By [20, Corollary 3.10] the ideals of Πmax are precisely the ideals of the form D0 ⊕ R∞ N where D0 is a maximal ideal of D(N ). In this case D(N ) is naturally identified with ℓ∞(N) and its maximal ideal space with the sequences vanishing at points of C(βN). The maximal ideals of T (N) corresponding to points of N are precisely the In and so we can write Πmax = {I1,I2, . . .} ∪ {Dx ⊕ R∞ N : x ∈ βN \ N} where Dx is the maximal ideal of D(N ) corresponding to sequences in ℓ∞(N) van- ishing at x ∈ βN. There remains the set Πint of primitive ideals which are neither diagonal ideals nor maximal ideals. These are the primitive ideals P where P ∩ D(N ) is a closed ideal of D(N ) corresponding to an ideal of ℓ∞(N) which strictly contains c0(N) and is not maximal. We cannot give a complete catalogue of these ideals but we can provide a rich set of examples. Consider the following special case of a general construction of epimorphisms between nest algebras, taken from Corollary 5.3 of [3]. Let 0 ≤ mk < nk < +∞ be integers such that the intervals (mk, nk] are pairwise disjoint and let U be a free ultrafilter on N. Suppose that limk∈U nk − mk = +∞. Let Uk : ℓ2(N) → ℓ2(N) be the partial isometry mapping ei to ei−mk when mk < i ≤ nk and zero otherwise. For X ∈ T (N) define φ(X) := lim k∈U UkXU ∗ k where convergence is in the weak operator topology and the limit always exists by WOT-compactness of the unit ball. Then by [3, Corollary 5.3] this map is an epimorphism of T (N) onto T (N). Note also that φ is a *-homomorphism of the diagonal of T (N) onto itself. If φ is such an epimorphism of T (N) onto T (N) and π is an irreducible represen- tation of T (N) then clearly π ◦ φ is also an irreducible representation of T (N). If ker π is in Πmax then so is ker π ◦ φ. However, as we shall see, if ker π ∈ Πmin\ Πmax then ker π◦ φ will be in Πint and this provides a rich supply of examples of primitive ideals in Πint. Assuming the Continuum Hypothesis, I∞ ∈ Πmin \ Πmax, so consider the prim- itive ideal P = φ−1(I∞). Note that φ annihilates I∞ and so I∞ is the unique diagonal ideal in P. Writing ∆(X) for the diagonal expectation P∞ k=1(Nnk − Nmk )X(Nnk − Nmk ), observe that ker ∆ ⊆ ker φ ⊆ P and so P 6= I∞. Thus P 6∈ Πmin. On the other hand, P 6∈ Πmax since, by [20, Theorem 3.8], every max- imal ideal of T (N) contains R∞ N , but P does not contain the unilateral backward shift U since φ(U ) = U 6∈ I∞. Thus P 6∈ Πmin and P 6∈ Πmax, and so P ∈ Πint. In fact this construction readily yields uncountably many incomparable ideals in Πint. For fix projections Pk := Nnk−Nmk where limk→+∞ nk−mk = +∞ and let U ON THE PRIMITIVE IDEALS OF NEST ALGEBRAS 19 be a fixed free ultrafilter. As is well-known we can find an uncountable collection Σ of infinite subsets of N with the property that distinct members of Σ intersect only in finite sets. For σ ∈ Σ, list the elements of σ in order as sk and build an ultrafilter epimorphism φσ : T (N) → T (N) as above, this time employing the intervals Psk and the ultrafilter U. Write ∆σ(X) for the diagonal expectation Pk∈σ PkXPk. As before, ker ∆σ ⊆ ker φσ. Now for any σ 6= σ′, φ−1 σ′ (I∞), for otherwise σ (I∞) 6= φ−1 σ (I∞) = φ−1 φ−1 σ′ (I∞) ⊇ ker ∆σ + ker ∆σ′ + I∞ = T (N) We can also exhibit infinite chains of ideals in Πint for since φ−1(I∞) ) I∞, the ideals Pk := φ−1(φ−1(··· φ−1(I∞)··· )) form a chain of distinct ideals in Πint for any fixed epimorphism φ : T (N ) → T (N ). 6.3. Some properties of ideals in Πint. Although the ultrafilter epimorphism construction of ideals in Πint is not representative, we can prove some properties which all ideals in Πint share with the ultrafilter construction. These results are, however, tightly bound to the case of T (N) (especially Proposition 6.3) and it is unclear how they might be extended. Proposition 6.3. Let P be a primitive ideal of T (N) and suppose P ) I∞. Then there is an increasing sequence of integers nk such that P contains {X ∈ T (N ) : (Nnk − Nnk−1)X(Nnk − Nnk−1) = 0 for all k} Proof. Let L be a maximal left ideal such that P is the kernel of the left-regular representation on T (N )/ L. By Proposition 5.2, P contains a projection P 6∈ I− I . Choose a subsequence of nest projections Nnk such that rank(Nnk+1 − Nnk )P ≥ rank Nnk for all k. We shall show that if S := {X ∈ T (N ) : (Nn2k+2 − Nn2k )X(Nn2k+2 − Nn2k ) = 0 for all k} then S ⊆ P. By Remark 3.1, since S is a two-sided ideal of T (N ), if S ⊆ L then S ⊆ P, so suppose for a contradiction that S 6⊆ L. By maximalty of L, S + L = T (N ) and so there is an X ∈ S such that I − X ∈ L. Decompose X as Y0 + Y1 where ∞ Y0 := and Y1 := X − Y0. Observe that therefore (3) Xk=1 (Nn2k+1 − Nn2k−1 )X(Nn2k+1 − Nn2k−1) (Nnk+2 − Nnk )Y1(Nnk+2 − Nnk ) = 0 for all k. Now take fixed arbitrary M < N < I in N and consider two cases. First, if N − M does not dominate any Nnk+1 − Nnk then there must be a k such that N − M ≤ Nnk+2 − Nnk , and so (N − M )Y1(N − M ) = 0. On the other hand if N − M does dominate some Nnk+1 − Nnk , take k to be the largest possible (which exists since N < I) and observe that, by (3), rank(N − M )Y1(N − M ) = rank Nnk (N − M )Y1(N − M ) ≤ rank Nnk ≤ rank(Nnk+1 − Nnk )P ≤ rank(N − M )P 20 JOHN LINDSAY ORR It follows that in either case rank(N − M )Y1(N − M ) ≤ rank(N − M )P. Since the right-hand side is infinite if N = I, the inequality is valid for all M < N in N . It follows immediately from [21, Theorem 2.6] that Y1 factors through P as Y1 = AP B for some A, B ∈ T (N ), and so Y1 ∈ P ⊆ L, whence I − Y0 ∈ L. However since X ∈ S the terms of the sum for Y0 are (Nn2k+1 − Nn2k−1)X(Nn2k+1 − Nn2k−1 ) = (Nn2k − Nn2k−1 )X(Nn2k+1 − Nn2k ) so that Y0 is nilpotent of order 2. Thus I − Y0 cannot belong to the proper left ideal L, which is a contradiction. (cid:3) Let Ei (i ∈ N) be a set of pairwise orthogonal intervals of N . For σ ⊆ N let Pσ :=Pi∈σ Ei and ∆σ(X) :=Pi∈σ EiXEi. For convenience write ∆ for ∆N. The last result shows that, at least in T (N), primitive ideals which are not in Πmin must contain ker ∆ for suitable {Ei}. The next two lemmas explore the consequences of a primitive ideal containing ker ∆, and hold for general nest algebras. Lemma 6.4. Let P be a primitive ideal of T (N ) and suppose ker ∆ ⊆ P. Then Σ := {σ ⊆ N : ker ∆σ ⊆ P} is an ultrafilter. Proof. Σ itself is non-empty since N ∈ Σ, and the sets in Σ are non-empty since ker ∆∅ = T (N ). If τ ⊇ σ and σ ∈ Σ then ker ∆τ ⊆ ker ∆σ ⊆ P and so τ ∈ Σ. If σ, τ ∈ Σ then ker ∆σ∩τ = ker ∆σ + ker ∆τ ⊆ P, and so σ∩ τ ∈ Σ. Thus Σ is a filter. Let π : T (N ) → L(V ) be an irreducible representation with P = ker π. For any σ ⊆ N and X ∈ T (N ), PσX− XPσ ∈ ker ∆ and so π(Pσ) commutes with π(T (N )). Thus ran(π(Pσ)) is an invariant subspace of π(T (N )) and so π(Pσ) = 0, I. Suppose that Pσ = I and so Pσc = 0. Then for any X ∈ T (N ), X − ∆σ(X) − ∆σc (X) ∈ ker ∆ ⊆ P and so π(X) = π(∆σ(X) + ∆σc (X)) = π(∆σ(X)Pσ + ∆σc (X)Pσc ) = π(∆σ(X)) whence ker ∆σ ⊆ P and σ ∈ Σ. Likewise, if Pσ = 0, then σc ∈ Σ. Thus Σ is an ultrafilter. (cid:3) i + E1 i XEj i . i and the lower endpoint of E1 Lemma 6.5. Let P be a primitive ideal of T (N ) and suppose ker ∆ ⊆ P. Suppose that for each i we can decompose Ei as the sum E0 i of intervals of N . Then P contains one of ker ∆j where ∆j (X) =P∞ i=1 Ej Proof. Each Ei is decomposed into the sum of two intervals which share a common endpoint. Let σ be the set of i for which the shared endpoint is the upper endpoint i . Clearly σc is then the set of i for which of E0 the upper endpoint of E1 i . By Lemma 6.4, P contains one of ker ∆σ, ker ∆σc . Without loss of generality assume ker ∆σ ⊆ P. Let P := Pi∈σ E0 i and observe that for each i ∈ σ there is an Ni ∈ N such that i = N ⊥ i = NiEi and E1 E0 ∆σ(P ⊥XP ) =Xi∈σ i equals the lower endpoint of E0 EiN ⊥ i XNiEi = 0 i Ei, and thus E1 i XE0 i =Xi∈σ ON THE PRIMITIVE IDEALS OF NEST ALGEBRAS 21 If π is an irreducible representation with ker π = P then π(P ⊥XP ) = 0 and so the range of π(P ) is an invariant subspace of π(T (N )), whence, one of P, P ⊥ ∈ P. If P ∈ P then while if P ⊥ ∈ P then ker ∆1 ⊆ ker ∆σ + PT (N ) ⊆ P ker ∆0 ⊆ ker ∆σ + T (N )P ⊥ ⊆ P (cid:3) i , we Theorem 6.6. Let P ∈ Πint in T (N). Then there is a free ultrafilter U and a sequence of pairwise orthogonal finite-rank intervals Ei such that limi∈U rank Ei = +∞ and P contains {X ∈ T (N) : lim i∈U kEiXEik = 0} i + E1 i } or {E1 i }. Moreover, given any decomposition of the Ei as the sums of intervals E0 can replace {Ei} with one of {E0 Proof. The existence of the intervals follows from Proposition 6.3. Let U be the ultrafilter obtained in Lemma 6.4. If limi∈U kEiXEik = 0 then, given ǫ > 0, there is a σ ∈ U such that kEiXEik < ǫ for all i ∈ σ. Thus taking X ′ := X − ∆σ(X), we see that kX − X ′k = k∆σ(X)k ≤ ǫ and that ∆σ(X ′) = 0, whence X ′ ∈ P. Thus X is a limit point of P and since P is norm closed, X ∈ P. i , we know from Proposition 6.5 that one of ker ∆j (j = 0, 1) is in P. Without loss suppose ker ∆0 ⊆ P. Again by Lemma 6.4, U 0 := {σ : ker ∆0 σ ⊆ P} is an ultrafilter. Now let σ ∈ U 0. Since U is an ultrafilter, one of σ, σc ∈ U. But if σc ∈ U then Given a decomposition Ei = E0 i + E1 T (N) = ker ∆0 σ + ker ∆σc ⊆ P which is impossible. Thus σ ∈ U and so, since σ was arbitrary, U 0 ⊆ U. But U0 is also an ultrafilter, so in fact U0 = U. Thus we may replace {Ei} with {E0 i }. Now it follows that the limit of the ranks of the intervals must be +∞, for otherwise after finitely many decompositions we could conclude that P ⊇ R∞ N and so P ∈ Πmax. Similarly if U were not free then P would contain {X : Ei0 XEi0 = 0} for some i0 ∈ N and, after finitely many decompositions if necessary, we would see that P ⊇ In for some n, again contrary to hypothesis. Acknowledgements (cid:3) The author gratefully acknowledges the hospitality of Professor Tony Carbery and the University of Edinburgh Mathematics Department. References [1] Frank F. Bonsall and John Duncan. Complete Normed Algebras, volume 80 of Erg. der Math. Springer, 1973. [2] Kenneth R. Davidson. Nest Algebras, volume 191 of Res. Notes Math. Pitman, Boston, 1988. [3] Kenneth R. Davidson, Kenneth J. Harrison, and John L. Orr. Epimorphisms of nest algebras. Internat. J. Math., 6(5):657 -- 687, 1995. [4] Kenneth R. Davidson, Elias Katsoulis, and David R. Pitts. The structure of free semigroup algebras. J. reine angew. Math., 533:99 -- 125, 2001. [5] Kenneth R. Davidson and John L. Orr. The Jacobson radical of a CSL algebra. Transactions of the Amer. Math. Soc., 334(2):925 -- 947, 1994. [6] Kenneth R. Davidson and John L. Orr. Principal bimodules of nest algebras. J. Func. Anal., 157(2):488 -- 533, 1998. 22 JOHN LINDSAY ORR [7] Kenneth R. Davidson and Stephen C. Power. Best approximation in C∗-algebras. J. reine angew. Math., 368:43 -- 62, 1986. [8] Allan P. Donsig. Semisimple triangular AF algebras. J. Funct. Anal., 111(2):323 -- 349, 1993. [9] Allan P. Donsig, Aristides Katavolos, and Antonios Manoussos. The Jacobson radical for analytic crossed products. J. Func. Anal., 187(1):129 -- 145, 2001. [10] Thomas Fall, William Arveson, and Paul Muhly. Perturbations of nest algebras. J. Operator Theory, 1(1):137 -- 150, 1979. [11] E. G. Katsoulis, R. L. Moore, and T. T. Trent. Interpolation in nest algebras and applications to operator corona theorems. J. Operator Theory, 29(1):115 -- 123, 1993. [12] Elias G. Katsoulis and Christopher Ramsey. Crossed products of operator algebras: applica- tions of takai duality. [13] E. C. Lance. Some properties of nest algebras. Proc. London Math. Soc., 19(3):45 -- 68, 1969. [14] David R. Larson. Nest algebras and similarity transformations. Ann. Math., 121:409 -- 427, 1985. [15] Adam W. Marcus, Daniel A. Spielman, and Nikhil Srivastava. Interlacing families II: Mixed characteristic polynomials and the Kadison-Singer problem. Ann. Math., 182:327 -- 350, 2015. [16] Laura Mastrangelo, Paul S. Muhly, and Baruch Solel. Locating the radical of a triangular operator algebra. Math. Proc. Cambridge Philos. Soc., 115(1):27 -- 38, 1994. [17] John L. Orr. The maximal ideals of a nest algebra. J. Func. Anal., 124:119 -- 134, 1994. [18] John L. Orr. Triangular algebras and ideals of nest algebras. Memoirs of the Amer. Math. Soc., 562(117), 1995. [19] John L. Orr. The stable ideals of a continuous nest algebra. J. Operator Theory, 45:377 -- 412, 2001. [20] John L. Orr. The maximal two-sided ideals of nest algebras. J. Operator Theory, 73:407 -- 416, 2015. [21] John L. Orr and David R. Pitts. Factorization of triangular operators and ideals through the diagonal. Proc. Edin. Math. Soc., 40:227 -- 241, 1997. [22] Justin Peters. Semicrossed products of C ∗-algebras. J. Func. Anal., 59(3):498 -- 534, 1984. [23] John R. Ringrose. On some algebras of operators. Proc. London Math. Soc., 15(3):61 -- 83, 1965. [24] Nik Weaver. Set theory and C ∗-algebras. Bull. Symbolic Logic, 13(1):1 -- 20, 2007. Toll House, Traquair Road, Innerleithen, EH44 6PF, United Kingdom E-mail address: [email protected]
1306.5372
1
1306
2013-06-23T04:01:53
Remarks on free mutual information and orbital free entropy
[ "math.OA", "math.PR" ]
The present notes provide a proof of $i^*(\mathbb{C}P+\mathbb{C}(I-P)\,;\mathbb{C}Q+\mathbb{C}(I-Q)) = -\chi_\mathrm{orb}(P,Q)$ for any pair of projections $P,Q$ with $\tau(P)=\tau(Q)=1/2$. The proof includes new extra observations, such as a subordination result in terms of Loewner equations. A study of the general case is also given.
math.OA
math
REMARKS ON FREE MUTUAL INFORMATION AND ORBITAL FREE ENTROPY MASAKI IZUMI 1 AND YOSHIMICHI UEDA 2 Abstract. The present notes provide a proof of i∗(CP + C(I − P ) ; CQ + C(I − Q)) = −χorb(P, Q) for any pair of projections P, Q with τ (P ) = τ (Q) = 1/2. The proof includes new extra observations, such as a subordination result in terms of Loewner equations. A study of the general case is also given. 1. Introduction There are two quantities which play a role of mutual information in free probability; one is the so-called free mutual information i∗ introduced by Voiculescu [21] in the late 90s and the other is the orbital free entropy χorb due to Hiai, Miyamoto and the second-named author [12],[20] (and its new approaches χorb, etc. due to Biane and Dabrowski [3]). These quantities have many properties in common, but no general relationship between them has been established so far. Any question about i∗ and/or χorb for two projections is known to be a 'commutative one' in essence, that is, can essentially be handled within classical analysis (see [21, §12] and [13]), and a heuristic argument in [15] supports that the identity i∗ = −χorb holds at least for two projections. Hence the question of i∗ = −χorb for two projections seems most tractable in the direction, and can be regarded as a counterpart of the single variable unification between two approaches χ and χ∗ of free entropy, which was already established by Voiculescu (see [23]). Recently Collins and Kemp [5] gave a proof of i∗ = −χorb for two projections with τ (P ) = τ (Q) = 1/2 under a rather restricted assumption, along the lines of the above-mentioned heuristic argument. Here we give an improved assertion of their result (i.e., completion of the analysis when τ (P ) = τ (Q) = 1/2) with a rather short and completely independent proof. Originally the first-named author observed important ideas after the appearance of [15] as a preprint, and then we prepared an essential part of the present short notes some years ago (see e.g. the introduction of [20]). Although the main theorem of the present notes is still an assertion about only the case of τ (P ) = τ (Q) = 1/2, a large part of its proof deals with general two projections and involves new extra observations which also enable us to give a partial result in the case of general trace values τ (P ), τ (Q). Hence the present notes may have some degree of positive significance for future studies in the direction. We should also emphasize that the attempts are important as positive evidence for the conjecture that i∗ = −χorb should hold for general random multivariables, though they have no direct connection with the unification conjecture for free entropy. Throughout the present notes, let (M, τ ) denote a sufficiently large, tracial W ∗-probability space so that all the non-commutative random variables that we will deal with live in (M, τ ). The operator norm is denoted by k − k∞. Let St, t ∈ [0,∞), be a free additive Brownian 1Supported in part by Grant-in-Aid for Scientific Research (B) 22340032. 2Supported in part by Grant-in-Aid for Scientific Research (C) 24540214. AMS subject classification: Primary: 46L54; secondary: 94A17. Keywords: Free mutual information; liberation process; orbital free entropy; free SDE; Loewner equation. 1 2 M. IZUMI AND Y. UEDA motion in (M, τ ) (with S0 = 0). A free unitary multiplicative Brownian motion Ut, t ∈ [0,∞), with U0 = I introduced by Biane [1] is a non-commutative process consisting of unitary random variables determined by the free stochastic differential equation (free SDE for short) dUt = √−1 dSt Ut − (1/2)Ut dt, U0 = I. For given two projections P, Q in M that are freely independent of {Ut}t≥0 the main objective here is to investigate the so-called liberation process t ∈ [0,∞) 7→ (Ut(CP + C(I − P ))U∗t , CQ + C(I − Q)) introduced by Voiculescu [21] in relation with i∗ and χorb. It is known that the liberation process can be understood by looking at the process of self-adjoint random variables Xt := QUtP U∗t Q. Thus we mainly investigate the process Xt in what follows. One can easily derive the free SDE dXt = Ξt ♯ dSt + Yt dt, where Ξt := √−1(Q ⊗ UtP U∗t Q − QUtP U∗t ⊗ Q) and Yt := τ (P )Q − Xt. See [4] for the definitions and the notations concerning free SDE's such as ♯-operation. Note that Ut is operator-norm continuous in t by [1, Lemma 8], and so are Xt, Ξt and Yt too. 2. Free SDE of (zI − Xt)−1 and Cauchy transform of Xt Several ways to investigate the free SDE of the resolvent process R(t, z) := (zI − Xt)−1 and the Cauchy transform of Xt have already been available, see e.g. [7, §6 -- 7],[16, §§3.2],[6, §§3.1] and [5]. However, we do give, for the reader's convenience, a simple proof of their explicit formulas by simple algebraic manipulations based on three naturally expected facts -- (i) the free Ito formula, (ii) the resolvent process becomes again a 'free Ito process' and (iii) every 'free Ito process' has a unique 'Doob -- Meyer decomposition'. In fact, the essential part of our proof will be done in several lines. The above (i) and (ii) were perfectly provided by Biane and Speicher [4], while the above (iii) is the latter half part of Proposition 2.2 below. The proposition (with its lemma) is probably a folklore. Lemma 2.1. Let {Mt}t≥0 be an increasing filtration of von Neumann subalgebras of M, and let t ∈ [0,∞) 7→ Kt be a weakly measurable process such that Kt ∈ Mt and sup0≤s≤t kKsk∞ < +∞ for all t ≥ 0. If t ∈ [0,∞) 7→ Lt := R t 0 Ks ds defines a martingale adapted to {Mt}t≥0, then Lt = 0 for all t ≥ 0. Proof. Since Lt is a martingale, one has, for any division 0 =: t0 < t1 < ··· < tn := t, n Xi=1 (cid:3) τ (L∗t Lt) = τ ((Lti − Lti − 1)∗(Lti − Lti − sup 1≤i≤n (ti − ti−1). 1 )) ≤ t(cid:0) sup 0≤s≤tkKsk∞(cid:1)2 It follows that Lt = 0, since sup1≤i≤n(ti − ti−1) can arbitrarily be small. Proposition 2.2. Let {Mt}t≥0 be as in Lemma 2.1 such that St ∈ Mt for every t ≥ 0. Let t ∈ [0,∞) 7→ Φt, Φ′t ∈ M ⊗alg M be operator-norm continuous biprocesses adapted to {Mt}t≥0 and t ∈ [0,∞) 7→ Kt, K′t ∈ M be weakly measurable processes such that sup0≤s≤t kKsk∞ < +∞ for every t ≥ 0 and the same holds for K′t. Then both Φ 1[0,t] and Φ′ 1[0,t] fall in Ba ∞ (see [4, 0 Φs ♯ dSs +R t §§2.1]) for every t ≥ 0, and hence we have two free stochastic integrals R t 0 Ks ds andR t 0 K′s ds as in [4, §§4.3] for every t ≥ 0. If those free stochastic integrals define Proof. The first part is trivial; hence left to the reader. One has R t 0 (Φ′s − Φs) ♯ dSs =R t 0 (Ks − K′s) ds, which must be zero by Lemma 2.1 and [4, Proposition 3.2.3]. Hence R t 0 Φs ♯ dSs = R t 0 Φ′s ♯ dSs andR t 0 K′s ds hold for every t > 0. The Ito isometry [4, §§3.1] immediately shows that Φ 1[0,t] = Φ′ 1[0,t] holds in Ba 2 for every t > 0, and hence Φ = Φ′ holds. We may and do assume that M has separable predual (with replacing it by its von Neumann subalgebra if necessary); thus one can choose a dense countable subset {ϕn}n∈N of the predual of M. One the same process, then Φ = Φ′ holds and Kt = K′t does almost surely in t. 0 Φ′s ♯ dSs +R t 0 Ks ds =R t FREE MUTUAL INFORMATION AND ORBITAL FREE ENTROPY 3 t1 that Kt = K′t holds almost surely in t. ϕn(Ks − K′s) ds = 0 for every 0 ≤ t1 < t2 (cid:8) ∞ and n ∈ N, which immediately implies has R t2 One can choose, for each z ∈ C+ := {z ∈ C Imz > 0}, a rapidly decreasing function fz on R which coincides with x 7→ (z−x)−1 on a neighborhood of [0, 1], and thus dR(t, z) = d(fz(Xt)) = (∂fz(Xt) ♯ Ξt) ♯ dSt + (∂fz(Xt) ♯ Yt + 1/2∆Ξtfz(Xt)) dt holds by [4, Proposition 4.3.4]. Here we do not recall the definitions of ∂fz(Xt) ♯ Ξt, ∂fz(Xt) ♯ Yt and ∆Ξt fz(Xt) (those can be found in [4, §§4.3], and remark that k∆U f (X)k∞ can be estimated by I2(f )kUk2 ∞ in the same way as in the discussion following [4, Definition 4.1.1]). Here we need only the following trivial fact: (cid:3) sup{k∂fz(Xt) ♯ Ytk∞ + k∆Ξtfz(Xt)k∞ t ≥ 0} < +∞. Write Mt :=R t for short, and let z ∈ C+ be arbitrarily fixed. We have 0 (∂fz(Xs) ♯ Ξs) ♯ dSs, Zt := ∂fz(Xt) ♯ Yt + (1/2)∆Ξtfz(Xt) and Nt :=R t 0 = d(cid:0)R(t, z)(zI − Xt)(cid:1) = dR(t, z) · (zI − Xt) + R(t, z) · d(zI − Xt) − dMt · dNt = dMt · (zI − Xt) + Zt(zI − Xt) dt − R(t, z) · dNt − R(t, z) · Yt dt − dMt · dNt, (2.1) 0 Ξs ♯ dSs and hence dMt · (zI − Xt) − R(t, z) · dNt = R(t, z)Yt dt − Zt(zI − Xt) dt + dMt · dNt. This formal computation can easily be justified by the rigorous formulas in [4, §§4.1]. Note that dMt · dNt = hh∂fz(Xt) ♯ Ξt, Ξtii dt by the free Ito formula (see [4, Definition 4.1.1] for the precise definition of hh−,−ii). Therefore, Proposition 2.2 (which can be used thanks to (2.1)) shows that dMt = R(t, z) · dNt · R(t, z) =(cid:0)(R(t, z) ⊗ R(t, z)) ♯ Ξt(cid:1) ♯ dSt, Zt dt = R(t, z)YtR(t, z) dt + R(t, z) · dNt · R(t, z) · dNt · R(t, z). It is easy to see, by the free Ito formula again, that dNt · R(t, z) · dNt =(cid:0) − 2τ (XtR(t, z))Xt + τ (QR(t, z))Xt + τ (XtR(t, z))Q(cid:1) dt, and hence (the first part of) the next proposition follows. Proposition 2.3. For every z ∈ C+ the resolvent process R(t, z) := (zI − Xt)−1 satisfies: with dR(t, z) =(cid:0)(R(t, z) ⊗ R(t, z)) ♯ Ξt(cid:1) ♯ dSt + Z(t, z) dt Z(t, z) = τ (P )R(t, z)QR(t, z) − R(t, z)XtR(t, z) − 2τ (XtR(t, z))R(t, z)XtR(t, z) + τ (QR(t, z))R(t, z)XtR(t, z) + τ (XtR(t, z))R(t, z)QR(t, z). (2.2) Moreover, the Cauchy transform G(t, z) := τ (R(t, z)), z ∈ C+ satisfies the following partial differential equation (PDE for short): ∂G ∂t = ∂ ∂z (cid:20)(z2 − z)G2 + (2 − τ (P ) − τ (Q) − z)G − (1 − τ (P ))(1 − τ (Q)) z (cid:21) . Proof. The first part has already been obtained. Hence it suffice to show the desired PDE. Remark that Zt = Z(t, z) is operator-norm continuous in t thanks to the fact at the end of §1. By the martingale property, G(t, z) = τ (R(t, z)) = τ (R(0, z)) +R t 0 τ (Zs) ds, and hence, by (2.2), ∂G ∂t = τ (Zt) = τ (P )τ (QR(t, z)2) − τ (XtR(t, z)2) − 2τ (XtR(t, z))τ (XtR(t, z)2) + τ (QR(t, z))τ (XtR(t, z)2)+ τ (XtR(t, z))τ (QR(t, z)2). Note that τ (AR(t, z)2) = − ∂ ∂z τ (AR(t, z)) for any A ∈ M. Since R(t, z) = QR(t, z)Q + z−1(I − Q) and I = (zI − Xt)R(t, z) = zR(t, z) − XtR(t, z), we have τ (QR(t, z)) = G(t, z) − 1−τ (Q) and τ (XtR(t, z)) = zG(t, z) − 1. These altogether imply the desired PDE. (cid:3) z 4 M. IZUMI AND Y. UEDA 3. Analysis of Probability Distribution of Xt 1 Let νt be the probability distribution of Xt, i.e., a unique probability measure on [0, 1] z−x νt(dx), for z ∈ C+. Define c0(t) := τ ((I − UtP U∗t )∧ (I − Q) + determined by G(t, z) =R[0,1] (I − UtP U∗t )∧ Q + UtP U∗t ∧ (I − Q)), c1(t) := τ (UtP U∗t ∧ Q), t ≥ 0. Several facts [21, Corollary 1.7, Proposition 8.7, Corollary 8.6 and Lemma 12.5] on liberation gradients with e.g. [15, (1.3)] altogether show that the projections UtP U∗t , Q are in generic position for every t > 0 and moreover that both c0(t) = 1 − min{τ (P ), τ (Q)} and c1(t) = max{τ (P ) + τ (Q) − 1, 0} hold for every t > 0. (We will give its detailed explanation in Remark 3.5 at the end of this section for the reader's convenience.) By a well-known fact (see e.g. [11, Solution 122]) one easily sees that the functions t 7→ ci(t) are upper semicontinuous, and hence c0(0) ≥ c0(+0) = 1− min{τ (P ), τ (Q)} and c1(0) ≥ c1(+0) = max{τ (P ) + τ (Q) − 1, 0}. Set µt := νt − (1 − min{τ (P ), τ (Q)})δ0 − (max{τ (P ) + τ (Q) − 1, 0})δ1, t ≥ 0, which defines a positive measure on [0, 1], since ci(0) ≥ ci(+0), i = 0, 1. When t > 0, µt agrees with the restriction of νt to (0, 1). Moreover, µ0 agrees with the restriction of ν0 to (0, 1) (or equivalently, both ci(0) = ci(+0), i = 0, 2, hold) if and only if P, Q are in generic position. (See e.g. the proof of [13, Theorem 3.2].) Denote by F (t, z) the Cauchy transform of µt whose domain clearly contains C \ [0, 1]. A tedious computation derives the following PDE from Proposition 2.3: ∂ ∂F ∂t = ∂z (cid:2)(z2 − z)F 2 + a(z − 1)F + bzF(cid:3) (3.1) with a := τ (P ) − τ (Q) and b := τ (P ) + τ (Q) − 1. Similarly to Geronimus's work [10, §30] (based upon the so-called Szego mapping) we trans- form z ∈ C\ [0, 1] 7→ ζ ∈ D, the open unit disk, by z = (2 + ζ + ζ−1)/4 or ζ = 2z − 1 + 2√z2 − z (note that ζ ∈ D determines the branch of √z2 − z with a negative real value at z = 2). Set L(t, ζ) := −√z2 − z F (t, z). Since dζ dz = ζ/√z2 − z, the PDE (3.1) becomes ∂ζ (cid:20)(cid:18)L + a 1 − ζ(cid:19) L(cid:21) = 0. 1 − ζ 1 + ζ 1 + ζ + b ∂ ∂L ∂t + ζ (3.2) Letting µt(dθ) = µt(dx) with x = cos2(θ/2) = 1 2 (1 + cos θ), θ ∈ [0, π], we have L(t, ζ) = 1 1 = ζ − ζ(cid:19)Z[0,π] ζ − ζ(cid:19)Z[0,π] 4(cid:18) 1 4(cid:18) 1 =Z[0,π] −1 + 1 µt(dθ) 1 1 1 4 (2 + e√−1θ + e−√−1θ) ζ(cid:1) − cos2(θ/2) ζ(cid:1) − 1 e−√−1θ e−√−1θ − ζ! µt(dθ), 4(cid:0)2 + ζ + 1 4(cid:0)2 + ζ + 1 e√−1θ e√−1θ − ζ 2 (µt + (µt ↾(0,π)) ◦ j−1) with j : θ ∈ (0, π) 7→ −θ ∈ (−π, 0) µt(dθ) + L(t, ζ) =Z(−π,π] e√−1θ + ζ e√−1θ − ζ µt(dθ). (3.3) and thus the symmetrization µt := 1 satisfies Define H(t, ζ) := (L(t, ζ) + a 1−ζ 1+ζ + b 1+ζ 1−ζ )L(t, ζ), and by (3.2) we have ∂H ∂t + ζ(cid:16)2L(t, ζ) + a 1 − ζ 1 + ζ + b 1 + ζ 1 − ζ(cid:17) ∂H ∂ζ = 0. (3.4) FREE MUTUAL INFORMATION AND ORBITAL FREE ENTROPY 5 As usual, let us consider the ordinary differential equations (ODE's for short) of characteristic curve t 7→(cid:0)gt(ζ), ut(ζ) := H(t, gt(ζ))(cid:1) associated with the PDE (3.4): 1 − gt(ζ)(cid:21) , gt(ζ) = gt(ζ)(cid:20)2L(t, gt(ζ)) + a 1 − gt(ζ) 1 + gt(ζ) 1 + gt(ζ) + b ut(ζ) = 0, u0(ζ) = H(0, ζ). g0(ζ) = ζ, (3.5) (3.6) Here the dot symbol () denotes the differentiation in t. The ODE (3.5) is nothing less than the radial Loewner (or Lowner -- Kufarev) equation (or more precisely radial Loewner ODE) determined by one parameter family of measures t 7→ 2µt + aδπ + bδ0. Note by e.g. [15, (1.3)] that 2µt + aδπ + bδ0 defines a probability measure on T = (−π, π] for every t ≥ 0. (This follows from the fact that UtP U∗t , Q are in generic position for every t > 0 as remarked before and µt → µ0 weakly as t ց 0.) Thus, by a standard fact, see e.g. [19, Theorem 4.14], the radial Loewner ODE (3.5) defines a unique one-parameter family of conformal transformations gt : Dt := {ζ ∈ D Tζ > t} ։ D with gt(0) = 0 and g′t(0) = et (the prime symbol (′) denotes the differentiation in ζ), where Tζ, ζ ∈ D, is the supremum of all T such that a solution of (3.5) exists until time T in such a way that gt(ζ) ∈ D holds for every t ≤ T . It is known, see e.g. [19, Remark 4.15] again, that the inverse ft := g−1 e√−1θ + ζ e√−1θ − ζ ft(ζ) = −ζ f′t(ζ)"Z(−π,π] (2µt + aδπ + bδ0)(dθ)# , : D ։ Dt satisfies f0(ζ) = ζ, (3.7) t a radial Loewner PDE. The ODE (3.6) shows that H(t, gt(ζ)) = ut(ζ) = u0(ζ) = H(0, ζ), and hence H(t, ζ) = H(0, ft(ζ)) holds for all ζ ∈ D. This implies that 1 − ζ(cid:19)2 1 + ζ where √− is the principal branch. The discussions so far are summarized as follows. 1 − ζ(cid:19) + 2s(cid:18)a L(t, ζ) = − + 4H(0, ft(ζ)), 2(cid:18)a 1 − ζ 1 + ζ 1 − ζ 1 + ζ 1 + ζ (3.8) + b + b 1 1 Proposition 3.1. Let νt be the probability distribution of Xt. Define the positive measure µt := νt − (1 − min{τ (P ), τ (Q)})δ0 − (max{τ (P ) + τ (Q) − 1, 0})δ1, and transform it to the positive measure µt(dθ) := µt(dx) on [0, π] by x = cos2(θ/2). Then µt coincides with the restriction of νt to (0, 1) for every t > 0, and moreover so does for t = 0 (or equivalently, µ0 has no atom at both 0 and 1) if and only if the given two projections P, Q are in generic position. − Set L(t, ζ) := R(−π,π) 2 (µt + (µt ↾(0,π) ) ◦ j−1) with j : θ ∈ (0, π) 7→ −θ ∈ (−π, 0). Then the unique one-parameter, subordinate family of conformal self-maps ft on D obtained from the radial Loewner PDE (3.7) driven by the probability measures 2µt + aδπ + bδ0 gives the following subordination relation: µt(dθ), ζ ∈ D, with the symmetrization µt := 1 − e√ e√ 1θ +ζ 1θ−ζ (cid:16)L(t, ζ) + a 1 − ζ 1 + ζ + b 1 + ζ 1 − ζ(cid:17)L(t, ζ) =(cid:16)L(0, ft(ζ)) + a 1 − ft(ζ) 1 + ft(ζ) + b 1 + ft(ζ) 1 − ft(ζ)(cid:17)L(0, ft(ζ)) with a = τ (P ) − τ (Q) and b = τ (P ) + τ (Q) − 1. The next corollary is a specialization of the above proposition. Corollary 3.2. Let L(t, ζ), ft(ζ) be as in Proposition 3.1, set gt(ζ) := f−1 that τ (P ) = τ (Q) = 1/2 or equivalently a = b = 0. Then t (ζ), and suppose • L(t, ζ) = L(0, ft(ζ)), that is, L(t, ζ) is subordinate to L(s, ζ) for s < t, • gt(ζ) = ζe2tL(0,ζ) and ft(ζ) = ζe−2tL(t,ζ), • ReL(t, ζ) = (log ζ − log ft(ζ))/2t, t > 0 and ζ ∈ D \ {0}. 6 M. IZUMI AND Y. UEDA Proof. Under the assumption here the subordination relation in Proposition 3.1 turns out to be the exact subordination L(t, ζ) = L(0, ft(ζ)). This together with (3.5) implies that gt(ζ) = 2gt(ζ)L(t, gt(ζ)) = 2gt(ζ)L(0, ζ) . This ODE can easily be solved as gt(ζ) = ζe2tL(0,ζ), implying ζ = ft(ζ)e2tL(0,ft(ζ)) = ft(ζ)e2tL(t,ζ). The final assertion immediately follows. (cid:3) This allows us to prove some properties of µt by analyzing ft(ζ) and/or gt(ζ) when τ (P ) = τ (Q) = 1/2, but we give a more useful observation as the next proposition. The proposition immediately follows from only (3.2) and (3.3). This means that the proof of the main result of the present notes (Theorem 4.3) needs only a few pages. Proposition 3.3. Under the same assumption as in Corollary 3.2, {2µt/2}t≥0 is identical to the one-parameter semigroup of probability distributions associated with a free unitary multiplicative Brownian motion with initial distribution 2µ0. 1θ − ζe√ 1−ζe√ Proof. Since µt is symmetric, we have ψ(t, ζ) :=R(−π,π] 1θ (2µt/2)(dθ) = L(t/2, ζ) − 1/2, the moment generating function of the measure 2µt/2. The PDE (3.2) can easily be transformed into − ψ + ζ(ψ + 1/2)ψ′ = 0. (3.9) This is the PDE that the moment generating function of a free unitary multiplicative Brownian motion satisfies, see e.g. the proof of [21, Proposition 10.8], and hence the desired assertion follows as seen below. Let U be a unitary random variable with distribution 2µ0, which is freely independent of {Ut}t≥0. Set ψ(t, ζ) := τ ((I − ζUtU )−1 − I), ζ ∈ D, the moment generating function of UtU . Then ψ satisfies the same PDE (3.9). Write ψ(t, ζ) =P∞n=1 cn(t)ζn, ψ(t, ζ) = P∞n=1 cn(t)ζn. Developing (3.9) into power series as above we see that both the coefficients cn fn = − n and cn must satisfy that f1 = − 1 2 f1, k=1 kfkfn−k (n = 2, 3, . . . ) with fn = cn ζe√ 1θ (2µ0)(dθ) = ψ(0, ζ), ζ ∈ D, one has cn(0) = cn(0) for or cn. Since ψ(0, ζ) = R(−π,π) 1−ζe√ every n. Hence one can recursively show that R(−π,π] e√−1nθ (2µt/2)(dθ) = cn(t) = cn(t) = 2 fn −Pn−1 τ ((UtU )n). (cid:3) 1θ − − Remarks 3.4. (1) The above proposition enables us to derive detailed information about µt from many existing results [1],[2, §§4.2],[21, §1] on free unitary multiplicative Brownian motions (with the help of S-transform machinery, see e.g. [22, §3]) when τ (P ) = τ (Q) = 1/2. Moreover, the recent work [24] generalizing Biane's analysis [2, §§4.2] gives more detailed properties of µt and hence those of µt, though we omit to collect any result in the direction here. (2) The above proposition also recaptures, as its specialization, the main theorem of [9]. In fact, the free Jacobi process with parameter (λ, θ) = (1, 1/2) [7] is exactly our Xt (viewed as a random variable in (QMQ, τ (Q) τ )) with P = Q and τ (P ) = τ (Q) = 1/2. Hence the initial distribution 2µ0 is the unit mass at θ = 0, and thus the probability distribution of the free Jacobi process with parameter (λ, θ) = (1, 1/2) is exactly that of the free unitary multiplicative Brownian motion via x = cos2(θ/2). 1 Remark 3.5. The following simple 'liberation theoretic' proof of the fact that UtP U∗t , Q are in generic position for every t > 0 has been available so far: By [21, Corollary 1.7, Proposition 8.7] d∗Ut:C1 ⊗ 1 (see the notation there) exists in L2 for every t > 0, which implies, by [21, Corollary 8.6], that so does the liberation gradient j(Ut(CP + C(I − P ))U∗t : CQ + C(I − Q)). Therefore, by [21, Lemma 12.5] (together with Ut(CP + C(I − P ))U∗t = CUtP U∗t + C(I − UtP U∗t )) we conclude that UtP U∗t , Q are in generic position for every t > 0. This argument indeed shows the following stronger result: U P U∗, Q are in generic position for any unitary U with finite Fisher information F (U ) < +∞ ([21, Definition 8.9]) which is freely independent of P, Q. FREE MUTUAL INFORMATION AND ORBITAL FREE ENTROPY 7 4. Free Mutual Information and Orbital Free Entropy To a given pair of projections P, Q we can associate four quantities: the liberation gradient j(CP + C(I − P ) : CQ + C(I − Q)) (=: j(P : Q) for short), the liberation Fisher information ϕ∗(CP +C(I−P ) : CQ+C(I−Q)) (=: ϕ∗(P : Q)), the mutual free information i∗(CP +C(I−P ) : CQ+C(I−Q)) (=: i∗(P : Q)), all of which are due to Voiculescu [21], and the orbital free entropy χorb(P, Q) [12]. Note that i∗(CP +C(I−P ) ; CQ+C(I−Q)) = i∗(CP +C(I−P ) : CQ+C(I−Q)), see [21, Remarks 10.2 (c)], and hence it suffices to compute the latter quantity for our purpose. According to the change of variables µt µt µt in §3 we need to reformulate Voiculescu's computation of ϕ∗(P : Q), [21, §12], as well as the previous computation of χorb(P, Q) essentially due to Hiai and Petz [13]. For simplicity, write δ := δ CP +C(I−P ) : CQ+C(I−Q), the derivation associated with CP + C(I − P ) and CQ + C(I − Q) [21, §§5.3]. Let µ be the restriction of the probability distribution of QP Q to (0, 1). Note that the measure µ is not changed if QP Q is replaced by P QP and that µ is exactly 1 2 ν in [21, §12]. Write a := τ (P ) − τ (Q) and b := τ (P ) + τ (Q) − 1 for simplicity. If P, Q are in generic position, then by [21, §§12.1 -- 12.6] one has, for n ≥ 1, (τ ⊗ τ ) ◦ δ (P Q)n = 2 PVZ Z(0,1)2 + (a + b)Z(0,1) x − y xn(x − 1) µ ⊗ µ (dx, dy) xn−1(x − 1) µ(dx) + bZ(0,1) 1 xn−1 µ(dx). Here 'PV' is the sign of Cauchy principal value. With θ ∈ (0, π) 7→ x = cos2(θ/2) ∈ (0, 1) and µ(dθ) := µ(dx) as in §3 we have, for n ≥ 1, (τ ⊗ τ ) ◦ δ (P Q)n = −2 PVZ Z(0,π)2 cos2n−1(α/2) sin(α/2) µ ⊗ µ (dα, dβ) sin α (4.1) cos α − cos β cos2(n−1)(θ/2) sin2(θ/2) µ(dθ) + bZ(0,π) − aZ(0,π) cos2n(θ/2) µ(dθ). Here we further suppose that µ has a density function h, i.e., µ(dx) = h(x) dx. Set h(θ) := h(cos2(θ/2)) sin(θ/2) cos(θ/2), and thus µ(dθ) = h(θ) dθ. Then the symmetrization µ := 1 2 (µ + µ ◦ j−1) with j : θ ∈ (0, π) 7→ −θ ∈ (−π, 0) also has a density function, that is, µ(dθ) = h(θ) dθ with h(θ) = (h(cos2(θ/2)) sin θ)/4 = (h(cos2(θ/2)) sin(θ/2) cos(θ/2))/2, θ ∈ (−π, π). The Hilbert transform (or the harmonic conjugate) of h is defined by (Hh)(θ) := 1 2π PVZ h(φ) tan((θ − φ)/2) dφ, θ ∈ T = [−π, π), which exists a.e., see [18, III.C.2]. As in [17, §6.7, (6.86)] the restriction of Hh to (0, π) can be re-written in terms of h as follows. (Hh)(θ) = − sin θ 2π PVZ(0,π) h(φ) cos θ − cos φ dφ, θ ∈ (0, π). Under the equivalent assumptions h ∈ L2(−π, π) if and only if h ∈ L2(0, π), or equivalently Z(0,1)px(1 − x)h(x)2 dx < +∞ (4.2) (4.3) ϕ∗(CP + C(I − P ) : CQ + C(I − Q)) =Z(0,π) 2(cid:12)(cid:12)4π(Hh)(θ) − a tan(θ/2) + b cot(θ/2)(cid:12)(cid:12) =Z(−π,π)(cid:12)(cid:12)2π(H(2h))(θ) − a tan(θ/2) + b cot(θ/2)(cid:12)(cid:12) 2 µ(dθ) 2 (2µ)(dθ) (4.4) 8 M. IZUMI AND Y. UEDA the Cauchy principal value in (4.2) converges in L2-norm by [18, I.E.4]. Define a function ξ : (0, π) → M2(C) by ξ(θ) :=(cid:0)4π(Hh)(θ) − a tan(θ/2) + b cot(θ/2)(cid:1)(cid:20)0 −1 0 (cid:21) 1 With these preliminaries we have: Lemma 4.1. Assume that P, Q are in generic position. If µ(dx) = h(x) dx such that h satisfies (4.3), then ξ gives the liberation gradient j(CP + C(I − P ) : CQ + C(I − Q)) as long as θ 7→ 4π(Hh)(θ) − a tan(θ/2) + b cot(θ/2) is integrable with respect to µ, and moreover possibly to be +∞ under the same integrability assumption. Proof. By the computation (4.1) together with (4.2) and the hypotheses (4.3), one can easily see that τ (ξ(P Q)n) = (τ ⊗ τ )◦ δ (P Q)n for n ≥ 1 (whose proof is just an translation of the proof of [21, Proposition 12.7] into the present context), and conclude j(P : Q) = ξ by its definition (see [21, Definition 5.4]) under the integrability assumption. Then the first equality in (4.4) is immediate, and the second one follows from the fact that θ 7→ 4π(Hh)(θ)−a tan(θ/2)+b cot(θ/2) is an odd function. (cid:3) Keep the notations µ, µ, µ, and a, b above. If P, Q are in generic position, then χorb(P, Q) =Z Z(0,1)2 log x − y µ ⊗ µ (dx, dy) log x µ(dx) + bZ(0,1) + aZ(0,1) log(1 − x) µ(dx) + C; otherwise −∞, where C is a unique constant determined by χorb(P, Q) = 0 when P, Q are freely independent with keeping prescribed values of τ (P ), τ (Q). In particular, C = (log 2)/2 when τ (P ) = τ (Q) = 1/2. See e.g. [15, Lemma 1.1],[12, Lemma 2.4]. In what follows we assume that P, Q are in generic position, and, in particular, µ((0, 1)) = (1 − a − b)/2 by [15, (1.3)]. Since cos α − cos β = (e√−1α − e√−1β · e√−1α − e−√−1β)/2, with x = cos2(α/2), y = cos2(β/2) we have log x − y µ ⊗ µ (dx, dy) =Z Z(0,π)2(cid:0) log cos α − cos β − log 2(cid:1) µ ⊗ µ (dα, dβ) √−1α − e √−1α − e √−1β + log e √−1β µ ⊗ µ (dα, dβ) − √−1α − e−√−1β − 2 log 2(cid:1) µ ⊗ µ (dα, dβ) log 2 (1 − a − b)2. 2 Z Z(0,1)2 =Z Z(0,π)2(cid:0) log e = 2Z Z(−π,π)2 log e Here we used the fact that µ(0, 1) = µ(0, π) = µ(−π, π) = (1 − a − b)/2. With x = cos2(θ/2) we have Z(0,1) log x µ(dx) = 2Z(−π,π) log 1 + e √−1θ µ(dθ) − (1 − a − b) log 2, FREE MUTUAL INFORMATION AND ORBITAL FREE ENTROPY 9 Z(0,1) log(1 − x) µ(dx) = 2Z(−π,π) log 1 − e √−1θ µ(dθ) − (1 − a − b) log 2. Therefore, we conclude: Lemma 4.2. If P, Q are in generic position, then χorb(P, Q) = 2nZ Z(−π,π)2 + aZ(−π,π) log e √−1α − e √−1β µ ⊗ µ (dα, dβ) √−1θ µ(dθ) + bZ(−π,π) √−1θ µ(dθ)o + Z with a universal constant Z = Zτ (P ),τ (Q) depending only on τ (P ), τ (Q); otherwise −∞. In particular, if τ (P ) = τ (Q) = 1/2, then the above formula of χorb(P, Q) simply becomes log 1 + e log 1 − e χorb(P, Q) = 2Z Z(−π,π)2 log e √−1α − e √−1β µ ⊗ µ (dα, dβ). Let us return to the original situation; thus we use the notations in §3. We can now reduce our question to [21, Corollary 10.9] when τ (P ) = τ (Q) = 1/2. Theorem 4.3. For any two projections P, Q with τ (P ) = τ (Q) = 1/2 one has i∗(CP + C(I − P ) ; CQ + C(I − Q)) = −χorb(P, Q) possibly with +∞ = +∞. Proof. By Proposition 3.3 and [21, Corollary 1.7] 2µt/2 has an L∞-density 2h(t/2, θ) for every t > 0. By [21, Corollary 10.9] we have −Z Z(−π,π]2 By Lemma 4.1 √−1α − e log e 2Z +∞ 1 √−1β (2µ0) ⊗ (2µ0) (dα, dβ) Z(−π,π](cid:0)2πH(2h(t/2,−))(θ)(cid:1)2 = 0 2h(t/2, θ) dθ dt. (4.5) holds for every t > 0 so that the right-hand side of (4.6) is identical to (2h(t/2,−))(θ) dθ = ϕ∗(Ut/2P U∗t/2 : Q) Z(−π,π)(cid:0)2πH(2h(t/2,−))(θ)(cid:1)2 2Z +∞ ϕ∗(Ut/2P U∗t/2 : Q) dt = 1 0 1 2Z +∞ 0 ϕ∗(UtP U∗t : Q) 2 dt = 2 i∗(P : Q) = 2 i∗(P ; Q). Assume first that P, Q are in generic position. By Lemma 4.2 the left-hand side of (4.5) is identical to −2 χorb(P, Q). Thus the desired identity follows. Assume next that P, Q are not in generic position. By what we have done in §3 µ0 must have at least one atom at either 0 or π with weight c1(0) − c1(+0) (cid:9) 0 or c0(0) − c0(+0) (cid:9) 0, respectively. Thus the left-hand side of (4.5) must be +∞, and therefore, so is i∗(P ; Q). By definition χorb(P, Q) = −∞ in this case, and hence the desired identity holds as +∞ = +∞. (cid:3) In closing we illustrate how one can use the subordination relation in Proposition 3.1. Lemma 4.4. If H(t, ζ) (see §3) defines a function in ζ of Hardy class with exponent 3/2 (see [18, IV.B.2]) at each t > 0, then i∗(CP + C(I − P ) ; CQ + C(I − Q)) = −χorb(P, Q) holds. 10 M. IZUMI AND Y. UEDA ∂t (ζ) = ∂ReL Proof. Let L(t, ζ) be as in §3, and write L(t, ζ) := L(t, ζ) + a 1−ζ has ∂Re L Σs(p, q) := 2R R(−π,π]2 log 1 − se√−1(α−β) p(e√−1α) q(e√−1β) dα ∂θ(cid:0)Re L(ζ) · ImL(ζ) + Im L(ζ) · ReL(ζ)(cid:1) with ζ = re√−1θ. Write Poisson kernel Pr(θ) the same trick as in the proof of [21, Proposition 10.8] shows that 1−ζ . By the PDE (3.1) one dβ 2π for simplicity. With the ∂t (ζ) = − ∂ 1+ζ + b 1+ζ 2π 1 Σs(Pr ∗ (µt2 + aδπ + bδ0), Pr ∗ µt2 ) − Σs(Pr ∗ (µt1 + aδπ + bδ0), Pr ∗ µt1 ) 1 + sre√−1θ 2Z t2 t1 Z(−π,π] 1 − sre√−1θ(cid:17) = √−1θ) × Im(cid:16)2L(t, re +Z t2 t1 hZ(−π,π] 1 − sre√−1θ 1 + sre√−1θ 1 + re√−1θ 1 − re√−1θ(cid:17) 2ReL(t, re √−1θ) + a 1 − re√−1θ 1 + re√−1θ Im(cid:16)2L(t, sre √−1θ) + a Im(L + L)(t, sre √−1θ) + b + b dθ 2π(cid:17) dt (4.6) √−1θ)ImL(t, re 1 − re√−1θ 1 + re√−1θ × Re(cid:16)a + b 1 + re√−1θ 1 − re√−1θ(cid:17) dθ 2πi dt for every 0 < t1 < t2 < ∞. Since ReL(t, ζ)2 ≤ H(t, ζ), the assumption here implies that µt has an L3-density h(t, θ), i.e., µt(dθ) = h(t, θ) dθ, for every t > 0 (see [18, p.15]). We fix arbitrary 0 < t1 < t2 < +∞ for a while. Set Mt1 := supr<1 kH(t, re√−1(−))k3/2 < +∞ by assumption, where k − kp denotes the usual Lp-norm with respect to dθ rather than dθ/2π following [18]. By the subordination relation in Proposition 3.1 with Littlewood's subordination principle (see [8, Theorem 1.7]) one has kReL(t, re √−1(−))k3 ≤ kH(t, re √−1(−))k1/2 3/2 ≤ M 1/2 t1 ; hence kh(t,−)k3 ≤ M 1/2 t1 /2π (4.7) 1 + re√−1(−) for every t ≥ t1 and 0 ≤ r < 1. Note that 1 − re√−1(−) 1 + re√−1(−) ImL(t, re (cid:13)(cid:13)(cid:13) √−1(−))Re(cid:16)a √−1(−))k3/2 ≤ Mt1 for every t ≥ t1 and 0 ≤ r < 1 by the subordination relation in Proposition 3.1 with Little- wood's subordination principle again. Using the Cauchy -- Schwarz inequality (with respect to Re(··· ) dθ/2π dt) and then the Holder inequality (with respect to dθ and exponents 3, 3/2) with the help of M. Riesz's theorem (see [18, p.91]) we see that the absolute value of the second term of the right-hand side of (4.6) is not greater than 1 − re√−1(−)(cid:17)(cid:13)(cid:13)(cid:13)3/2 ≤ kH(t, re + b nZ t2 t1 hZ(−π,π] Im(L + L)(t, sre √−1θ)2Re(cid:16)a 1 − re√−1θ 1 + re√−1θ + b 1 + re√−1θ 2πi dto1/2 1 − re√−1θ(cid:17) dθ ×M 3/4 t1 pC3(t2 − t1)/2π with a universal constant C3 > 0 (that comes from M. Riesz's theorem) and moreover that this converges to 0 as r ր 1 thanks to [18, p.7 -- 8], (4.7), the continuity of Im(L + L)(t, ζ) in (t, ζ) and Im(L + L)(t,±s) = 0 (due to h(−θ) = h(θ)). By the subordination relation in Proposition 3.1 with Littlewood's subordination principle again √−1(−)) + a 1 − re√−1(−) 1 + re√−1(−) + b Im(cid:16)2L(t, re (cid:13)(cid:13)(cid:13) 1 + re√−1(−) 1 − re√−1(−)(cid:17) 2ReL(t, re √−1(−))(cid:13)(cid:13)(cid:13)3/2 √−1(−))k3/2 ≤ 4Mt1 ≤ 4kH(t, re (4.8) FREE MUTUAL INFORMATION AND ORBITAL FREE ENTROPY 11 for every t ≥ t1 and 0 ≤ r < 1, and we can easily confirm, with the help of facts in [18, p.9; p.88 -- 89], that the first term of the right-hand side of (4.6) converges to 1 + se√−1θ 1 2Z t2 t1 Z(−π,π] 1 − se√−1θ 1 + se√−1θ √−1θ) + a Im 2L(t, se ×(cid:16)2πH(2h(t,−))(θ) − a tan(θ/2) + b cot(θ/2)(cid:17) 2h(t, θ) dθ dt 1 − se√−1θ! + b log s we have as r ր 1. Consequently, letting Z(s) := Zτ (P ),τ (Q) − 1−(a+b)2 4 −n2Z Z(−π,π]2 2Z t2 t1 Z(−π,π] = 1 −n2Z Z(−π,π]2 √−1(α−β) (µt1 + aδπ + δ0)(dα) µt1 (dβ) + Z(s)o √−1θ) + a log 1 − se Im 2L(t, se ×(cid:16)2πH(2h(t,−))(θ) − a tan(θ/2) + b cot(θ/2)(cid:17) 2h(t, θ) dθ dt √−1(α−β) (µt2 + aδπ + δ0)(dα) µt2 (dβ) + Z(s)o. 1 − se√−1θ! 1 − se√−1θ 1 + se√−1θ 1 + se√−1θ log 1 − se + b (4.9) Write k(t, θ) := 2πH(2h(t,−))(θ) − a tan(θ/2) + b cot(θ/2) for simplicity. By (4.8) with [18, p.9; p.88 -- 89] one has (4.10) for every t ≥ t1. By Lemma 4.1 with the aid of (4.7) and (4.10), and moreover by [21, Proposition 10.11 (a)] kk(t,−) 2h(t,−)k3/2 ≤ 2Mt1/π t1 Since Z t2 t1 Z(−π,π] By the Holder inequality, (4.7), (4.10) and M. Riesz's theorem ϕ∗(UtP U∗t : Q) dt ≤ 2i∗(Ut1 P U∗t1 : Q) < +∞. k(t, θ)2 2h(t, θ) dθ dt =Z t2 Z t2 t1 Z(−π,π] 2πH(2h(t,−))(θ)k(t, θ) 2h(t, θ) dθ dt < +∞. Im(cid:16) 1 − se√−1θ Im(cid:16) 1 + se√−1θ a(cid:12)(cid:12)(cid:12) 1 − se√−1θ(cid:17)(cid:12)(cid:12)(cid:12) ≤ a tan(θ/2) + b cot(θ/2) 1 + se√−1θ(cid:17)(cid:12)(cid:12)(cid:12) + b(cid:12)(cid:12)(cid:12) ≤ 2a tan(π/4) + 2b cot(π/4) + k(t, θ) + 2πH(2h(t,−))(θ), 2R t2 −χorb(Ut2 P U∗t2) + ϕ∗(UtP U∗t : Q) dt = −χorb(Ut1P U∗t1 , Q). we easily see, by [18, p.9; p.88 -- 89] again, that the first term of the right-hand side of (4.9) converges to 1 ϕ∗(UtP U∗t : Q) dt as s ր 1. By Lemma 4.2 with the aid of the first 5 lines of [21, p.147] we finally get t1 By [15, Theorem 2.1], [21, Proposition 10.11 (a)] and [12, Proposition 4.6] one has 1 2Z t2 t1 −χorb(Ut2P U∗t2, Q) dt2 ≤Z +∞ Z +∞ t1 t1 ϕ∗(Ut2P U∗t2 : Q) dt2 = 2 i∗(Ut1P U∗t1 : Q) < +∞, implying limt2ր+∞ χorb(Ut2P U∗t2, Q) = 0. (This trick originates in a preprint version of [15].) By [12, Proposition 2.5 (4), Proposition 4.6] one has limt1ց0 χorb(Ut1 P U∗t1, Q) = χorb(P, Q). Hence we are done. (cid:3) 12 M. IZUMI AND Y. UEDA By (3.2) H(t, ζ) becomes the constant (1− (a + b)2)/4 in the time stationary case; hence the assumption of Lemma 4.4 is not strange. Here is a sample of application of Lemma 4.4. Corollary 4.5. Assume that the measure µ0 (see §3) has an L3-density with respect to x(1 − x) dx on [0, 1] and is supported in [α, β] such that α (cid:9) 0 if τ (P ) 6= τ (Q) and β (cid:8) 1 if τ (P ) + τ (Q) 6= 1. Then i∗(CP + C(I − P ) ; CQ + C(I − Q)) = −χorb(P, Q) holds. Proof. For simplicity, assume both a = τ (P ) − τ (Q) 6= 0 and b = τ (P ) + τ (Q) − 1 6= 0. It is easy to see that µ0 has an L3-density h(0, θ) (with respect to dθ); hence L(0, ζ) is a function in ζ of Hardy class with exponent 3 by M. Riesz's theorem with a standard fact (see [18, p.9; p.88 -- 89]). Moreover, the assumption here implies that L(0, ζ) has analytic continuation across both ζ = ±1. Since limζ→±1 L(0, ζ) = 0, L(0, ζ) admits a power series expansion without constant term around ζ = ±1. Thus H(0, ζ) is bounded in some neighborhoods at both ζ = ±1. It is plain to show that H(0, ζ) is a function in ζ of Hardy class with exponent 3/2. Hence the assertion follows thanks to the subordination relation in Proposition 3.1 with Littlewood's subordination principle (see [8, Theorem 1.7]). (cid:3) The above fact suggests that the question should be affirmative without assuming τ (P ) = τ (Q) = 1/2. Only missing piece in our attempt is apparently a more detailed study of H(t, ζ) and/or the conformal transformations ft(ζ); thus the question comes down to a study of Loewner -- Kufarev equations. We thank Fumio Hiai for discussions on this subject matter and comments to a draft of the present notes. Acknowledgement References [1] P. Biane, Free Brownian motion, free stochastic calculus and random matrices, Free probability theory (Waterloo, ON, 1995), 1 -- 19, Fields Inst. Commun, Amer. Math. Soc., Providence, RI, 1997. [2] P. Biane, Segal -- Bargmann transform, functional calculus on matrix spaces and the theory of semi-circular and circular systems, Jour. Funct. Anal., 144 (1997), 232 -- 286. [3] Ph. Biane and Y. Dabrowski, Concavification of free entropy, Adv. Math., 234 (2013), 667 -- 696. [4] P. Biane and R. Speicher, Stochastic calculus with respect to free Brownian motion and analysis on Wigner space, Probab. Theory Related Fields, 112 (1998), 373 -- 409. [5] B. Collins and T. Kemp, Liberating projections, Preprint 2012, arXiv:1211.6037. [6] Y. Dabrowski, A free stochastic partial differential equation, Preprint 2010, arXiv:1008.4742. [7] N. Demni, Free Jacobi process, J. Theor. Probab., 21 (2008), 118 -- 143. [8] P.L. Duren, Theory of H p Spaces, Dover, 2000. [9] N. Demni, T. Hamdi and T. Hmidi, Spectral distribution of the free Jacobi process, Preprint 2012, arXiv:1204.6227. [10] Ya. Geronimus, Polynomials orthogonal on a circle and their applications, Amer. Math. Soc. Transl., Ser. I 3 (1962) 1 -- 78. [11] P. Halmos, A Hilbert Space Problem Book, Second Edition, Springer, 1982. [12] F. Hiai, T. Miyamoto and Y. Ueda, Orbital approach to microstate free entropy, Internat. J. Math., 20 (2009), 227 -- 273. [13] F. Hiai and D. Petz, Large deviations for functions of two projection matrices, Acta Sci. Math. (Szeged), 72 (2006), 581 -- 609. [14] F. Hiai and Y. Ueda, Notes on microstate free entropy of projections, Publ. Res. Inst. Math. Sci., 44 (2008), 49 -- 89. [15] F. Hiai and Y. Ueda, A log-Sobolev type inequality for free entropy of two projections, Ann. Inst. Henri Poincar´e Probab. Stat., 45 (2009), 239 -- 249. (A detailed version is available at math.OA/0601171.) [16] V. Kargin, On free stochastic differential equations, J. Theor. Probab., 24 (2011), 821 -- 848. [17] F. King, Hilbert Transforms, Vol.1, Encyclopedia of Mathematics and its applications, 124, Cambridge, 2009. FREE MUTUAL INFORMATION AND ORBITAL FREE ENTROPY 13 [18] P. Koosis, Introduction to Hp Spaces, Second edition, Cambridge Tracts in Mathematics, 115, Cambridge, 1998. [19] G.F. Lawler, Conformally Invariant Processes in the Plane, Mathematical Surveys and Monographs, 114. American Mathematical Society, Providence, RI, 2005. [20] Y. Ueda, Orbital free entropy, revisited, Indiana Univ. Math. J., to appear. arXiv:1210.6421. [21] D. Voiculescu, The analogue of entropy and of Fisher's information measure in free probability theory VI: Liberation and mutual free information, Adv. Math., 146 (1999) 101 -- 166. [22] D. Voiculescu, Lectures on free probability theory, in Lectures on probability theory and statistics (Saint- Flour, 1998), 279 -- 349, Lecture Notes in Math., 1738, Springer, Berlin, 2000. [23] D. Voiculescu, Free entropy, Bull. London Math. Soc. 34 (2002), 257 -- 278. [24] P. Zhong, On the free convolution with a free multiplicative analogue of the normal distribution, Preprint 2013, arXiv:1211.3160. (M.I.) Department of Mathematics, Graduate School of Science, Kyoto University, Sakyo-ku, Kyoto 606-8502, Japan E-mail address: [email protected] (Y.U.) Graduate School of Mathematics, Kyushu University, Fukuoka, 819-0395, Japan E-mail address: [email protected]
1412.7478
4
1412
2015-05-10T20:31:03
Liberations and twists of real and complex spheres
[ "math.OA", "math.QA" ]
We study the 10 noncommutative spheres obtained by liberating, twisting, and liberating+twisting the real and complex spheres $S^{N-1}_\mathbb R,S^{N-1}_\mathbb C$. At the axiomatic level, we show that, under very strong axioms, these 10 spheres are the only ones. Our main results concern the computation of the quantum isometry groups of these 10 spheres, taken in an affine real/complex sense. We formulate as well a proposal for an extended formalism, comprising 18 spheres.
math.OA
math
5 LIBERATIONS AND TWISTS OF REAL AND COMPLEX SPHERES 1 0 2 TEODOR BANICA Abstract. We study the 10 noncommutative spheres obtained by liberating, twisting, and liberating+twisting the real and complex spheres SN −1 . At the axiomatic level, we show that, under very strong axioms, these 10 spheres are the only ones. Our main results concern the computation of the quantum isometry groups of these 10 spheres, taken in an affine real/complex sense. We formulate as well a proposal for an extended formalism, comprising 18 spheres. , SN −1 R C Introduction A remarkable discovery, due to Goswami [31], is that each noncommutative compact Riemannian manifold X in the sense of Connes [21], [22], [23] has a quantum isometry group G+(X). While the classical, connected manifolds cannot have genuine quantum isometries [29], for the non-classical or non-connected manifolds the quantum isometry group G+(X) can be bigger than the usual isometry group G(X), containing therefore “non-classical” symmetries, worth to be investigated. As a motivating example, the symmetries of the finite noncommutative manifold coming from the Standard Model, axiomatized by Chamseddine and Connes in [18], [19], were studied by Bhowmick, D’Andrea, Dabrowski and Das in [11], [12]. One of their findings is that the usual gauge group component P U3 becomes replaced in this way by the quantum group P U + N , U + N are the quantum groups constructed by Wang in [41], [42], and the twisting result P O+ 9 . Here O+ 3 = P O+ 3 = ¯S+ N , S+ 9 comes from [3]. 3 = ¯S+ At a theoretical level, one interesting question is about adapting the various classical computations of isometry groups. Perhaps the most basic such computation is G(SN −1 ) = ON , where SN −1 R ⊂ RN is the standard sphere. When enlarging attention to disconnected manifolds, the computation G(XN ) = SN , where XN = {e1, . . . , eN} ⊂ RN is the simplex, e1, . . . , eN being the standard basis vectors of RN , should be included as well. Such results are of course quite trivial, but their noncommutative extensions, not al- ways. N , but more complicated computations, such as G(YN ) = HN , where YN = {±e1 . . . ± eN} ⊂ RN is the hypercube, and HN = Z2 ≀ SN , lead to some interesting questions. See [2], [32]. In the discrete manifold case we have G+(XN ) = S+ R 2000 Mathematics Subject Classification. 46L65 (46L54, 46L87). Key words and phrases. Quantum isometry, Noncommutative sphere. 1 2 TEODOR BANICA In the continuous manifold case, which is the one that we are interested in here, the extensions of the basic computation G(SN −1 ) = ON lead to interesting questions as well. This is well-known for instance in the context of the Podle´s spheres [35], and we refer here to [15], [27], [39]. More advanced examples of noncommutative spheres, having more intricate algebraic and differential geometry, come from [24], [25]. R R R,∗ , and the free sphere SN −1 In our joint work with Goswami [5] we introduced two basic generalizations of SN −1 , namely the half-liberated sphere SN −1 R,+ . These spheres appear by definition as dual objects to certain universal C ∗-algebras, inspired by the easy quantum group philosophy [8]. More precisely, the surjections at the C ∗-algebra level produce inclusions SN −1 R,+ , which are related, via the quantum isometry group construction, to the basic inclusions ON ⊂ O∗ R,∗ ⊂ SN −1 Our purpose here is three-fold: (1) We will review the work in [5], with a new axiomatization of these 3 spheres, less N ⊂ O+ N from [8], [9]. R ⊂ SN −1 relying on the structure of the corresponding quantum isometry groups. (2) We will present a unitary extension of [5], based on G(SN −1 C ) = UN , with the isometry group being taken in an affine complex sense. (3) We will present as well a twisting extension of [5], in both the real and complex cases, involving the group ¯ON from [2], and a number of related objects. We will construct in this way 10 noncommutative spheres, as follows: SN −1 C / SN −1 C,∗∗ / SN −1 C,+ ¯SN −1 C,∗∗ ¯SN −1 C SN −1 R SN −1 R,∗ SN −1 R,+ ¯SN −1 R,∗ ¯SN −1 R Here all the maps are inclusions. The spheres in [5] are those at bottom left, their complex analogues are on top left, and the whole right part of the diagram appears from the left part via twisting, with the middle spheres being equal to their own twists. We will prove then that the associated quantum isometry groups, all taken in an affine real/complex sense, in the spirit of [32], are as follows: UN / U ∗∗ N / U + N ON O∗ N O+ N ¯U ∗∗ N ¯O∗ N ¯UN ¯ON We believe that our 10 spheres are “smooth” and “Riemannian”, in some strong sense, which is yet to be determined. Some questions here, still open, were raised in [5]. / / o o o o / / O O / / O O O O O O o o O O o o / / o o o o / / O O / / O O O O O O o o O O o o LIBERATIONS AND TWISTS OF REAL AND COMPLEX SPHERES 3 At the axiomatic level, we will have results and conjectures stating that, under very strong axioms, our 10 spheres (or “geometries”, in a large sense) are the only ones. Our axioms exclude however many interesting objects, like the half-liberated geometry CN ∗ from [11]. Our third contribution will be a proposal, in order to fix this problem. We will show that the 10-geometry formalism has a natural 18-geometry extension, as follows: #●●●●●● ;✇✇✇✇✇✇ CN # CN ∗∗ ;✇✇✇✇✇✇ #●●●●●● :✉✉✉✉✉✉ CN ∗ / CN + ¯CN ∗ / RN ∗ RN + ;✇✇✇✇✇✇ $❍❍❍❍❍❍ CN ◦ CN CN − RN ¯CN # ¯CN ∗∗ ¯RN ∗ ¯CN ◦ ¯CN ¯CN − ¯RN Here the geometries CN inserting the geometry CN extension, however, requires a lot of work, and we have only partial results here. − → CN # are new, and appear when ∗ from [11] and its twist into the 10-geometry framework. This ◦ → CN # and ¯CN − → ¯CN ◦ → ¯CN We refer to the body of the paper for the precise statements of our results, and to the final section below for a summary of questions raised by the present work. The paper is organized as follows: in 1-2 we construct and axiomatize/classify the main 10 spheres, in 3-4 we study the corresponding quantum isometry groups, and in 5-6 we state and prove our main results, and we discuss the extended formalism. Formalism and notations. We use the “noncommutative compact space” framework coming from operator algebras. More precisely, the category of noncommutative compact spaces is by definition the category of unital C ∗-algebras, with the arrows reversed. According to the Gelfand theorem, the category of usual compact spaces embeds co- variantly into the category of noncommutative compact spaces, via X → C(X), the image is formed by the spaces coming from the commutative C ∗-algebras, and the inverse correspondence is obtained by taking the spectrum, X = {χ : C(X) → C}. We denote such noncommutative spaces by X, Y, Z, . . ., with the corresponding C ∗- algebras being denoted C(X), C(Y ), C(Z), . . . A morphism X → Y is by definition injec- tive if the corresponding morphism C(Y ) → C(X) in surjective, and vice versa. Our spaces will be for the most of algebraic geometric nature, coming in families {XNN ∈ N}, with each C(XN ) having N privileged generators x1, . . . , xN (the “co- ordinates”), subject to uniform relations, not depending on N. We will often refer to XN as the “specialization” of the abstract object X = (XN ), at a particular N ∈ N. # { { # ; / o o c c { { ; $ ; c c c c z z : d d O O / / / O O O O o o O O o o O O 4 TEODOR BANICA Acknowledgements. I would like to thank Julien Bichon, Uwe Franz, Adam Skalski, Georges Skandalis and Roland Speicher for useful discussions. This work was partly supported by the “Harmonia” NCN grant 2012/06/M/ST1/00169. 1. Noncommutative spheres We are interested in the noncommutative, undeformed analogues of RN , CN . At the pure algebra level, of the corresponding ∗-algebras of polynomial functions, these ana- logues can be introduced by “liberating” and “twisting” the various commutativity rela- tions ab = ba appearing in the following ∗-algebra presentation results: P ol(RN ) = Dx1, . . . , xN(cid:12)(cid:12)(cid:12) P ol(CN ) = Dz1, . . . , zN(cid:12)(cid:12)(cid:12) xi = x∗ zizj = zjzi, ziz∗ i , xixj = xjxiE j zjE j = z∗ However, if we want to have norms on our universal ∗-algebras, we must restrict atten- tion to compact submanifolds X ⊂ RN , Z ⊂ CN . And, the most natural candidates for such submanifolds are the corresponding spheres, SN −1 Looking at spheres is in fact not very restrictive, because many interesting manifolds . For instance, after a 1/√N rescaling of the coordinates, . In addition, many homogeneous appear as X ⊂ SN −1 any compact Lie group appears as G ⊂ UN ⊂ SN 2−1 spaces G → X appear as well naturally as submanifolds of spheres. analogues of SN −1 To summarize this discussion, we are interested in the noncommutative, undeformed . Our starting point will be the following result: R ⊂ RN and SN −1 , Z ⊂ SN −1 C ⊂ CN . , SN −1 R C C R C Proposition 1.1. The algebras of continuous functions on SN −1 R , SN −1 C are given by xi = x∗ C(SN −1 R C(SN −1 C ) = C ∗ x1, . . . , xN(cid:12)(cid:12)(cid:12) ) = C ∗ z1, . . . , zN(cid:12)(cid:12)(cid:12) x2 i , xixj = xjxi,Xi j zi,Xi j = z∗ i = 1! i = 1! ziz∗ zizj = zjzi, ziz∗ where at right we have universal C ∗-algebras. Proof. This is a well-known consequence of the Gelfand and Stone-Weierstrass theorems. Indeed, the univeral algebras on the right being commutative, they are of the form C(X), C(Z). The coordinate functions xi, zi provide us with embeddings X ⊂ RN , Z ⊂ CN , and then the quadratic conditions give X = SN −1 , Z = SN −1 , as claimed. (cid:3) R C The idea now is to replace the commutation relations ab = ba between the standard coordinates by some well-chosen relations. A first choice is that of using the anticom- mutation relations ab = −ba. A second choice, coming from the easy quantum group philosophy [8], is that of using the half-commutation relations abc = cba. A third choice, LIBERATIONS AND TWISTS OF REAL AND COMPLEX SPHERES 5 coming from the general liberation philosophy in free probability [8], [10], [34], [40], and which is perhaps the most straightforward, is that of using no relations at all. So, let us first construct the free analogues of SN −1 R , SN −1 C : Definition 1.2. The free versions of SN −1 R , SN −1 C are defined by C(SN −1 C(SN −1 R,+ ) = C ∗ x1, . . . , xN(cid:12)(cid:12)(cid:12) C,+ ) = C ∗ z1, . . . , zN(cid:12)(cid:12)(cid:12)Xi xi = x∗ i = 1! i zi = 1! z∗ x2 i ,Xi i =Xi ziz∗ where at right we have universal C ∗-algebras. Here the fact that the norms are bounded, and hence the universal algebras exist indeed, Observe that our definition of SN −1 comes from the quadratic conditions, which give xi ≤ 1,zi ≤ 1. Pi z∗ i zi = 1, instead of just a single one. There are several reasons for this choice: (1) We would like, as in usual projective geometry, the matrix p = (pij) formed by j to satisfy p = p∗ = p2, T r(p) = 1. And, the verification of C,+ involves both the equalities Pi ziz∗ the elements pij = ziz∗ i = 1 and (2) We would like as well, once again in analogy with the classical case, the generators zi to satisfy same the algebraic relations as the variables γi = u1i over the quantum group U + these conditions requires both Pi ziz∗ i = 1 and Pi z∗ N . And, these latter variables satisfy Pi γiγ∗ We will be back later on to these topics, with concrete results justifying our choice, and i =Pi γ∗ i γi = 1. i zi = 1. with some axiomatization results as well, once again relying on this choice. Let us construct now the twisted versions of SN −1 . These are well-known objects, appearing as q = −1 specializations of the Podle´s spheres [35]. In our generators and relations framework, these two spheres are best introduced as follows: , SN −1 R C Definition 1.3. The twisted versions of SN −1 R , SN −1 C are defined by C( ¯SN −1 R C( ¯SN −1 C ) = C(SN −1 ) = C(SN −1 R,+ ).hab = −ba,∀a, b ∈ {xi} distincti C,+ ).hαβ = −βα,∀a, b ∈ {zi} distinct, αβ = βα otherwisei where we use the notations α = a, a∗ and β = b, b∗. In other words, the defining relations for ¯SN −1 R C those for ¯SN −1 are xixj = −xjxi for any i 6= j, and Regarding the free spheres in Definition 1.2, these cannot be twisted. This is well- i zi for any i, and zizj = −zjzi, ziz∗ R,+ , ¯SN −1 j = −z∗ C,+ = SN −1 j zi for any i 6= j. C,+ , where needed. known, and we will use the conventions ¯SN −1 R,+ = SN −1 are ziz∗ i = z∗ Let us discuss now the half-liberation operation. In the real case this is obtained by using the relations abc = cba. In the complex case there are several choices, as explained in 6 TEODOR BANICA [11], [17]. We will use here the “minimal” choice, from [17]. The other choices, including the “maximal” one from [11], will be discussed later on. So, let us construct four more spheres, as follows: Definition 1.4. The half-liberations of SN −1 R , SN −1 C , ¯SN −1 R , ¯SN −1 C are defined by C(SN −1 R,∗ ) = C(SN −1 C(SN −1 C,∗∗ ) = C(SN −1 C( ¯SN −1 R,∗ ) = C(SN −1 R,+ ).habc = cba,∀a, b, c ∈ {xi}i C,+ ).habc = cba,∀a, b, c ∈ {zi, z∗ i }i R,+ ).habc = −cba,∀a, b, c ∈ {xi} distinct, abc = cba otherwisei C,+ ).hαβγ = −γβα,∀a, b, c ∈ {zi} distinct, αβγ = γβα otherwisei C( ¯SN −1 C,∗∗ ) = C(SN −1 where we use the notations α = a, a∗, β = b, b∗ and γ = c, c∗. We have so far 2 + 2 + 2 + 4 = 10 spheres, and we will temporarily stop here, because we will see in the next section that, under strong axioms, these spheres are the only ones. We will be back to more complicated examples later on, in section 6 below. As a first result about these 10 spheres, we have: Proposition 1.5. We have the following diagram, SN −1 C / SN −1 C,∗∗ / SN −1 C,+ ¯SN −1 C,∗∗ ¯SN −1 C SN −1 R SN −1 R,∗ SN −1 R,+ ¯SN −1 R,∗ ¯SN −1 R with all the maps being inclusions. Proof. In the untwisted case all the inclusions are clear from definitions. In the twisted case most of the inclusions are clear too, and we just have to check the two horizontal inclusions at right. Regarding the inclusion ¯SN −1 R,∗ , here the statement is that ab = −ba for a 6= b implies abc = −cba for a, b, c distinct, and abc = cba otherwise. The first claim follows from abc = −bac = bca = −cba. Regarding now the second claim, in the case a = b = c we have aaa = aaa, in the case a = b 6= c we have aac = −aca = caa, in the case a = c 6= b we have aba = aba, and in the case b = c 6= a we have abb = −bab = bba, and this finishes the proof. C,∗∗ , the proof here is similar, by replacing a, b, c with variables α, β, γ, given by α = a, a∗, β = b, b∗ and γ = c, c∗. (cid:3) Regarding the remaining inclusion, ¯SN −1 C ⊂ ¯SN −1 ⊂ ¯SN −1 R We investigate now the properness of the inclusions in the above diagram. A simple criterion for comparing spheres is by looking at the classical versions. We have here: / / o o o o / / O O / / O O O O O O o o O O o o LIBERATIONS AND TWISTS OF REAL AND COMPLEX SPHERES 7 Proposition 1.6. The classical versions of the 10 spheres are SN −1 C / SN −1 C / SN −1 C SN −1,1 C T⊕N SN −1 R SN −1 R SN −1 R SN −1,1 R Z⊕N 2 where SN −1,1 K Proof. The assertions for the untwisted spheres are clear by definition. is a union of (cid:0)N 2(cid:1) copies of S1 ∩ ¯SN −1 K, which is not smooth at N ≥ 3. C = T⊕N , because the relations for ¯SN −1 C C C Observe that we have SN −1 , applied to , read ab = 0, for any a, b ∈ {zi} distinct. We conclude that such the points z ∈ SN −1 points z are those having all but one coordinates vanishing, z ∈ T⊕N . R = Z⊕N By restricting now to the real case, we obtain SN −1 R ∩ ¯SN −1 Regarding the intersections SN −1,1 C,∗∗ , observe precisely when its coordinates satisfy zizjzl = 0, is the union of (cid:0)N that a point z ∈ SN −1 for any i, j, l distinct. Thus SN −1,1 Finally, the non-smoothness assertion is clear. 2(cid:1) copies of S1 ∩ ¯SN −1 R,∗ and SN −1,1 as well. C ∩ ¯SN −1 belongs to SN −1,1 K, as claimed. = SN −1 = SN −1 (cid:3) K K K 2 R R C Now back to the properness question, we have here: Theorem 1.7. The inclusions in Proposition 1.5 are as follows: (1) At N ≥ 3, all these inclusions are proper. (2) At N = 2 we have S1 R,∗ = ¯S1 R,∗ = S1 R,+, and the other inclusions are proper. Proof. We first discuss the general case, N ≥ 2. Here the 5 vertical inclusions are all proper, by Proposition 1.6. In order to check that the 4 horizontal inclusions at left and at right are proper, we can use a trick from [17]. Consider indeed one of the spheres SN −1 , with coordinates denoted z1, . . . , zN , and let us set: / ¯SN −1 C C Xi =(cid:18) 0 z∗ i zi 0(cid:19) It follows that the inclusions SN −1 sum up to 1, so they produce a representation of SN −1 R,∗ and ¯SN −1 don’t commute/anticommute, SN −1 ⊂ SN −1 These matrices are self-adjoint, they half-commute/half-anticommute, and their squares R,∗ . Now since these matrices R,∗ are proper. ⊂ ¯SN −1 It remains to investigate the 4 horizontal inclusions in the middle: (1) Case N ≥ 3. By intersecting everything with S2 because these inclusions appear from the above ones, by intersecting with SN −1 R+ . R,∗ / ¯SN −1 R ⊂ ¯SN −1 C,∗∗ are proper as well, R ⊂ SN −1 C,∗∗ and ¯SN −1 R,+, it is enough to prove that the R,+, this follows from the inclusions S2 fact that the inclusion of corresponding classical versions S2,1 R,+ are proper. For ¯S2 R,+ and ¯S2 R,∗ ⊂ S2 R,∗ ⊂ S2 R,∗ ⊂ S2 R is proper. C C R ⊂ S2 / / o o o o / / O O / / O O O O O O o o O O o o 8 TEODOR BANICA For the inclusion S2 R,∗ ⊂ S2 matrices in M2(C), which are of the following form: R,+, we can use a trick from [5]. Consider indeed the positive R,∗ ⊂ S2 Y =(cid:18)r z ¯z s(cid:19) Here r, s ∈ R and z ∈ C must be chosen such that both eigenvalues are positive, and this happens for instance when r, s > 0 and z ∈ C is small enough. Let us fix some numbers ri, si > 0 with i = 1, 2, 3, satisfyingPi ri =Pi si = 1. For any choice of small complex numbers zi ∈ C satisfying Pi zi = 0, the corresponding elements Yi constructed as above will be positive, and will sum up to 1. Moreover, by carefully choosing the zi’s, we can arrange as for Y1, Y2, Y3 not to pairwise commute. Consider now the matrices Xi = √Yi. These are all self-adjoint, and their squares sum R,+) → M2(C) mapping xi → Xi. Observe that Now since the relations abc = cba imply aabb = baab = bbaa, the squares of the standard up to 1, so we get a representation C(S2 this representation maps x2 i → Yi, and the elements Yi don’t commute. coordinates on S2 R,∗ commute. We conclude that S2 R,+ is indeed proper. (2) Case N = 2. Here we must prove that, among the 4 horizontal inclusions in the middle, the two upper ones are proper, and the two lower ones are isomorphisms: S1 C S1 R / S1 C,∗∗ / S1 C,+ ¯S1 C,∗∗ S1 R,+ S1 R,+ S1 R,+ ¯S1 C ¯S1 R R,∗ = ¯S1 R,∗ = S1 In order to prove S1 It remains to prove that the inclusions S1 R,+ observe that, since we have only two coordinates x, y, the half-commutation relations abc = ±bca reduce to the commutation relations xy2 = y2x, x2y = yx2. But these relations hold over S1 C,∗∗ ⊂ S1 C,∗∗ ⊂ S1 C,+ are proper. In order to do so, we can use a free complexification trick, cf. [1]. Let indeed z, t be the standard coordinates on S1 C, let u be a unitary free from both z, t, and set Z = uz, T = ut. Then ZZ ∗ +T T ∗ = Z ∗Z +T ∗T = 1, and since the relations Z 2T = ±T Z 2 are not satisfied, we conclude that S1 R,+, because x2 + y2 = 1. C,+ and ¯S1 C,+ are both proper. (cid:3) C,∗∗ ⊂ S1 C,+ and ¯S1 C,∗∗ ⊂ S1 Summarizing, we have constructed so far 10 basic examples of underformed noncom- mutative spheres. We will study in detail these spheres in sections 2-5 below, and we will come back to more complicated examples in section 6 below. 2. Axiomatization, classification In this section we prove that, under a suitable axiomatization for the undeformed noncommutative spheres, the 10 spheres constructed above are the only ones. / / o o o o / / O O / / O O O O O O o o O O o o LIBERATIONS AND TWISTS OF REAL AND COMPLEX SPHERES 9 Our axioms will be of course very strong. In order to introduce them, let us begin with some heuristics. The common features of our 10 spheres can be summarized as: Proposition 2.1. The 10 spheres all appear from SN −1 C,+ , by imposing the relations α = α∗, αβ = βα, αβ = ±βα, αβγ = γβα, αβγ = ±γβα with α = a, a∗, β = b, b∗, γ = c, c∗, and where the signs come from anticommutation. Proof. This is clear from the definition of the 10 spheres in section 1 above, with the sign claim coming from the computations in the proof of Proposition 1.5. (cid:3) The point now is that the above 5 types of relations, all coming from certain permuta- tions in S1, S2, S3, can be represented by suitable diagrams, as follows: ◦ • ◦ ◦ ✼✼✼✼✼✼✼✼ ✞✞✞✞✞✞✞✞ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ●●●●●●●●●●● ✇✇✇✇✇✇✇✇✇✇✇ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ◦ More precisely, associated to such a diagram is the relation obtained by putting coor- dinates on the legs, such as each string joins equal coordinates, and then by stating that the product on top equals the product on the bottom. And this, with the convention that the empty/full circles represent symbols of type α = a, a∗, and their conjugates, and that the dotted diagrams bring ± signs, coming from anticommutation. Before doing so, however, there are two important remarks to be made: In short, we can develop a diagrammatic approach to the axiomatization problem. I. We know from Proposition 1.6 that non-smooth manifolds can appear when inter- C,∗∗ is not secting twisted and untwisted spheres, and more specifically that SN −1 smooth. Thus, we do not want to mix usual diagrams with dotted diagrams: C ∩ ¯SN −1 ◦ ◦ ✼✼✼✼✼✼✼✼ ✞✞✞✞✞✞✞✞ ◦ ◦ + ◦ ◦ ◦ ◦ ◦ ◦ =⇒ ∅ II. We do not want to mix either the real and complex cases. Indeed, this would amount in labelling “black and white” all the legs of our diagrams, and the problem is that this would produce many many spheres, some of which are pathological. As an example, consider the “sphere” obtained from SN −1 C,+ by assuming that the coordinates z1, . . . , zN satisfy ab = ba. We would like later on this sphere to have a geometry, and a quantum 10 TEODOR BANICA isometry group. But, at the quantum group level, by using the formalism in [8]: • • ◦ ◦ ✹✹✹✹✹✹✹ ✡✡✡✡✡✡✡ ◦ ◦ • • = ◦ • ✼✼✼✼✼✼✼✼ ✞✞✞✞✞✞✞✞ • ◦ In other words, for a unitary quantum group the relations ab = ba between the standard coordinates imply the relations ab∗ = b∗a, and so the quantum group is classical. Thus, the above “sphere”, while being bigger than SN −1 , would have the same quantum isometry group as SN −1 . And this is a pathology, and so this sphere must be excluded. Summarizing, we have to discuss separately the cases R, C, ¯R, ¯C. Let us begin with: C C Definition 2.2. Let K = R, C, ¯R, ¯C be one of the fields R, C, with the bar standing for the fact that the associated sphere is by definition the twisted one. (1) A monomial relation over K is a formula of type ai1 . . . aik = ±aiσ(1) . . . aiσ(k), where σ ∈ Sk is a permutation, and where the ± sign is the one making the formula zi1 . . . zik = ±ziσ(1) . . . ziσ(k) hold, over the sphere SN −1 K,+ defined via a formula of type C(S) = C(SN −1 K,+ )/ < R >, where R comes from a set of monomial relations, each applied to all the variables γi = xi at K = R, and γi = zi, z∗ (2) A monomial sphere over K is a quantum subspace S ⊂ SN −1 i at K = C. . K Observe that our 10 spheres are all monomial, coming from the relations ab = ±ba and We agree to represent all permutations by diagrams, acting by definition downwards. abc = ±cba, corresponding to the permutations (21) ∈ S2 and (321) ∈ S3. As an example, the permutations (21) ∈ S2 and (321) ∈ S3 are represented as follows: Observe that each monomial sphere over K contains the sphere SN −1 monomial relation is satisfied by definition by the standard coordinates of K , because each SN −1 . K The monomial spheres are best parametrized by groups, as follows: Proposition 2.3. Given a set of permutations E ⊂ S∞, denote by SN −1 K,E the associated monomial sphere over K, with the relations R coming from the elements σ ∈ E. Then any monomial sphere is of the form SN −1 Proof. Consider indeed the set G ⊂ S∞ consisting of elements σ ∈ S∞ such that the relations ai1 . . . aik = aiσ(1) . . . aiσ(k) hold, in our monomial sphere. It is clear then that G is stable by composition, because X = Y, Y = Z implies X = Z. Also clear is the fact that G is stable by inversion, because X = Y implies Y = X, and the fact that G contains the unit permutation. Thus, G is indeed a group. (cid:3) K,G , with G ⊂ S∞ being a group. ◦ ◦ ✼✼✼✼✼✼✼✼ ✞✞✞✞✞✞✞✞ ◦ ◦ ◦ ◦ ◦ ●●●●●●●●●●● ✇✇✇✇✇✇✇✇✇✇✇ ◦ ◦ ◦ LIBERATIONS AND TWISTS OF REAL AND COMPLEX SPHERES 11 As an illustration for this result, by using the convention ∗ = ∗∗, in order to denote the half-liberation operation by ∗ in both the real and complex cases, we have: Proposition 2.4. The basic spheres SN −1 K,+ come from the groups where S∗ ∞ = (S∗ n)n≥1 is such that S∗ K,∗ ⊂ SN −1 K ⊂ SN −1 S∞ ⊃ S∗ ∞ ⊃ {1} 2n ≃ Sn × Sn, S∗ 2n+1 ≃ Sn × Sn+1. Proof. The assertions regarding SN −1 being insensitive to the value of K, we can assume that we are dealing with SN −1 R,∗ . K,+ are trivial. Regarding now SN −1 , SN −1 K K,∗ , the result We use the fact, from [9], that the relations abc = cba imply the relations of type ai1 . . . aik = aiσ(1) . . . aiσ(k), for any σ ∈ Sk having the property that when labelling cycli- cally the legs • ◦ • ◦ . . ., each string joins a black leg to a white leg. In addition, these relations imply the original relations abc = cba, because the permutation (321) ∈ S3 implementing these relations has indeed the “black-to-white” joining property: ◦ ◦ • ●●●●●●●●●●● ✇✇✇✇✇✇✇✇✇✇✇ ◦ • comes from the group S∗ • We conclude that SN −1 R,∗ having the black-to-white joining property. Now observe that S∗ ∞ consisting of permutations σ ∈ S∞ 3 , S∗ 4 are given by: ◦ • ◦ • ◦ ◦ • ◦ • • • ●●●●●●●●●●● ✇✇✇✇✇✇✇✇✇✇✇ ◦ ◦ • • • ◦ • ◦ ◦ • • ◦ ❆❆❆❆❆❆❆❆❆ • ⑥⑥⑥⑥⑥⑥⑥⑥⑥ • ◦ ◦ Thus we have S∗ 3 = S1 × S2 and S∗ 4 = S2 × S2, with the first component of each product coming from dotted permutations, and with the second component coming from the solid line permutations. In the general case, the proof is similar. (cid:3) ❆❆❆❆❆❆❆❆❆ • ⑥⑥⑥⑥⑥⑥⑥⑥⑥ ◦ • ◦ ◦ • ◦ ◦ • • ◦ ◦ • • ◦ ◦ • We call depth of a monomial sphere the smallest k ∈ N ∪ {∞} such that our sphere K,E , as in Proposition 2.3, with E ⊂ Sk. In other words, a monomial can be written as SN −1 sphere is of depth ≤ k when the relations defining it come from permutations σ ∈ Sk. With this convention, we have the following result: 12 TEODOR BANICA Theorem 2.5. The 10 fundamental spheres, which can be written as SN −1 C,S∞ / SN −1 C,S ∗ ∞ / SN −1 C,{1} ¯SN −1 C,S ∗ ∞ ¯SN −1 C,S∞ SN −1 R,S∞ SN −1 R,S ∗ ∞ SN −1 R,{1} ¯SN −1 R,S ∗ ∞ ¯SN −1 R,S∞ ◦ ◦ ◦ ◦ are precisely the monomial spheres having depth k ≤ 3. Proof. The first assertion follows from Proposition 2.4. In order to prove the uniqueness, we have to examine the 6 elements of S3. These are as follows: ◦ ❁❁❁❁❁❁❁❁ ◦ ◦ ✏✏✏✏✏✏✏ ✏✏✏✏✏✏✏ ◦ ◦ ◦ ◦ According to our diagrammatic conventions, the identity produces the 2 free spheres, the basic crossing, which appears twice, produces the 4 classical + twisted spheres, and the last diagram produces the 4 half-liberated spheres. Our claim now, which will finish the proof, is that the 3-cycles produce the same spheres as the basic crossing. ✳✳✳✳✳✳✳ ◦ ✳✳✳✳✳✳✳ ◦ ✂✂✂✂✂✂✂✂ ◦ ❁❁❁❁❁❁❁❁ ◦ ✂✂✂✂✂✂✂✂ ◦ ◦ ✳✳✳✳✳✳✳ ◦ ✏✏✏✏✏✏✏ ◦ ◦ ◦ ✳✳✳✳✳✳✳ ◦ ✏✏✏✏✏✏✏ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ◦ Let us first discuss the case K = R. Here the 3-cycle produce the “sphere” given by abc = cab. The point now is that, by using these relations, we obtain: (ab − ba)2 = abab − abba − baab + baba = aabb − aabb − aabb + baab = aabb − aabb − aabb + aabb = 0 Thus the sphere collapses to SN −1 In the case K = ¯R, the proof is similar. Indeed, the 3-cycle produces relations of type abc = ±cab, the precise formulae being: (1) abc = −acb = cab for a, b, c distinct, (2) aac = −aca = caa for a 6= c, (3) aba = −aab for a 6= b, (4) abb = −bab for a 6= c. With these relations in hand, we have the following computation: , and we are done. R (ab + ba)2 = abab + abba + baab + baba = −aabb + aabb + aabb − baab = −aabb + aabb + aabb − aabb = 0 Thus the sphere collapses to ¯SN −1 R , and we are done. / / o o o o / / O O / / O O O O O O o o O O o o LIBERATIONS AND TWISTS OF REAL AND COMPLEX SPHERES 13 Finally, in the remaining two cases K = C, ¯C the proof of the extra needed formula, (cid:3) namely ab∗ = ±b∗a, is similar, by adding ∗ exponents where needed. The above result is complementary to those in [5]. Let us recall indeed from there that the spheres SN −1 R,+ are precisely those whose corresponding quantum isometry group is easy. This is of course quite a sophisticated result, and Theorem 2.5 above, formulated directly in terms of the spheres themselves, is in a certain sense “better”. However, unifying Theorem 2.5 with [5] remains an open question. R,∗ ⊂ SN −1 R ⊂ SN −1 Let us discuss now what happens at depth 4: Proposition 2.6. There are no new monomial spheres at depth 4. ◦ ◦ ◦ ◦ ◦ Proof. We must study the 24 elements of S4. These are as follows: ◦ ◦ ◦ ✹✹✹✹✹✹✹ ◦ ✕✕✕✕✕✕ ✕✕✕✕✕✕ ◦ ◦ ◦ ◦ • ✹✹✹✹✹✹✹ • ✮✮✮✮✮✮ • • ✡✡✡✡✡✡✡ ✕✕✕✕✕✕ • • • • • • ◦ ◦ ✮✮✮✮✮✮ ◦ ✕✕✕✕✕✕ ◦ ◦ ◦ ◦ ◦ ✮✮✮✮✮✮ ◦ ✮✮✮✮✮✮ ◦ ◦ ✡✡✡✡✡✡✡ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ✮✮✮✮✮✮ ◦ ◦ ✕✕✕✕✕✕ ◦ ◦ ◦ ◦ ◦ ◦ ✹✹✹✹✹✹✹ ◦ ◦ ✕✕✕✕✕✕ ✕✕✕✕✕✕ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ✮✮✮✮✮✮ ◦ ✕✕✕✕✕✕ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ✮✮✮✮✮✮ ◦ ✕✕✕✕✕✕ ◦ ◦ ◦ ◦ ✮✮✮✮✮✮ • • ✹✹✹✹✹✹✹ • • ✡✡✡✡✡✡✡ ✕✕✕✕✕✕ • • • • ◦ ◦ ❃❃❃❃❃❃❃❃◦ ◦ ✕✕✕✕✕✕ ◦ ◦ ◦ ◦ ◦ ◦ ✮✮✮✮✮✮ ◦ ✮✮✮✮✮✮ ◦ ✡✡✡✡✡✡✡ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ✮✮✮✮✮✮ ◦ ◦ ◦ ◦ ◦ ✮✮✮✮✮✮ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ✹✹✹✹✹✹✹ ◦ ◦ ✡✡✡✡✡✡✡ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ✹✹✹✹✹✹✹ ◦ ◦ ✡✡✡✡✡✡✡ ◦ ◦ ◦ ◦ ✮✮✮✮✮✮ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ✕✕✕✕✕✕ ◦ ◦ ◦ ◦ ◦ ✹✹✹✹✹✹✹ • ✹✹✹✹✹✹✹ • ✡✡✡✡✡✡✡ ✡✡✡✡✡✡✡ • • • • ✮✮✮✮✮✮ ◦ ◦ ❃❃❃❃❃❃❃❃◦ ◦ ◦ ◦ ◦ ◦ ❃❃❃❃❃❃❃❃◦ ◦ ◦ ✕✕✕✕✕✕ ◦ ◦ ◦ ◦ Here the dotted lines correspond either to outer (left or right) strings, or to pairs of adjacent strings, and our claim is that all these dotted strings can be deleted. Indeed, for outer strings, this follows from the following computation, by summing over a: ❃❃❃❃❃❃❃❃◦ ◦ ◦ ◦ ◦ ◦ ◦ ❃❃❃❃❃❃❃❃◦ ◦ ◦ ✕✕✕✕✕✕ ◦ ◦ ◦ ◦ ❃❃❃❃❃❃❃❃◦ ◦ ◦ ◦ ◦ ◦ ◦ aX = aY =⇒ a∗aX = a∗aY =⇒ X = Y Xa = Y a =⇒ Xaa∗ = Y aa∗ =⇒ X = Y As for the adjacent string claim, this follows from a similar computation: XabY = ZabT =⇒ Xaa∗Y = Zaa∗T =⇒ XY = ZT XabY = ZbaT =⇒ Xaa∗Y = Za∗aT =⇒ XY = ZT 14 TEODOR BANICA Now since all the diagrams containing dotted strings correspond to depth 3 spheres, we have just to study the diagrams having no dotted strings. And there are 3 such diagrams, namely those having solid circles, with the corresponding relations being as follows: abcd = cadb, abcd = bdac, abcd = cdab The first two relations are equivalent, the corresponding diagrams being related by upside-down turning, and produce the usual sphere SN −1 K . Indeed, we have: abcd = cadb =⇒ abcd = cadb = dcba =⇒ abb∗d = db∗ba =⇒ ad = da As for the last relations, these produce the sphere SN −1 K,∗ , because we have: abc = cba =⇒ abcd = cbad = cdab abcd = cdab =⇒ abcde = cdabe = cbeda =⇒ abb∗de = b∗beda =⇒ ade = eda Thus, we have no new monomial sphere at depth 4, as claimed. (cid:3) We conjecture that the 10 monomial spheres in Theorem 2.5 are the only ones, regardless of the depth. Solving this conjecture would of course fully clarify our axiomatization. 3. Unitary quantum groups In this section we construct 10 compact quantum groups. We will show later on, in sections 5-6 below, that these are the quantum isometry groups of our 10 spheres. We use the formalism of compact matrix quantum groups, developed by Woronowicz in [43], [44]. For a detailed presentation of the theory, we refer to [33]. We begin with the following key definition, due to Wang [41]: N , U + uij = u∗ Definition 3.1. The compact quantum groups O+ N are defined by C(U + C(O+ N ) = C ∗(cid:16)(uij)i,j=1,...,N(cid:12)(cid:12)(cid:12) N ) = C ∗(cid:16)(uij)i,j=1,...,N(cid:12)(cid:12)(cid:12) ij, ut = u−1(cid:17) u∗ = u−1, ut = ¯u−1(cid:17) with Hopf algebra maps ∆(uij) =Pk uik ⊗ ukj, ε(uij) = δij, S(uij) = u∗ so the underlying spaces O+ embeddings ON ⊂ O+ As shown in [41], the above two algebras satisfy the axioms of Woronowicz in [43], [44], N are indeed compact quantum groups. We have proper N , at any N ≥ 2. See [33], [41]. N , U + N , UN ⊂ U + We have as well the following key examples, coming from [8], [17]: ji. Definition 3.2. The half-liberations of ON , UN are defined as C(O∗ N ) = C(O+ C(U ∗∗ N ) = C(U + N ).habc = cba,∀a, b, c ∈ {uij}i N ).(cid:10)abc = cba,∀a, b, c ∈ {uij, u∗ ij}(cid:11) with Hopf algebra maps ∆, ε, S obtained by restriction. LIBERATIONS AND TWISTS OF REAL AND COMPLEX SPHERES 15 We refer to [8], [9] for details regarding O∗ N . As already mentioned, and known since [11], in the unitary case the half-liberation operation is not unique. We will be back to more complicated examples in section 6 below. N , and to [17] for details regarding U ∗∗ Now let us twist the 2 + 2 classical and half-classical quantum groups. We agree that N , obtained by all objects to be constructed appear by definition as subspaces of O+ imposing certain extra relations on the standard coordinates uij. We first have: Proposition 3.3. We have quantum groups ¯ON ⊂ O+ N , ¯UN ⊂ U + N defined via N , U + αβ =(−βα for a, b ∈ {uij} distinct, on the same row or column otherwise βα with the usual convention α = a, a∗ and β = b, b∗. Proof. These quantum groups are well-known, see [2]. The idea indeed is that the existence of ε, S is clear. Regarding now ∆, set Uij =Pk uik ⊗ ukj. For j 6= k we have: uisuis ⊗ usjusk uisuit ⊗ usjutk +Xs −uituis ⊗ utkusj +Xs uisuis ⊗ (−uskusj) UijUik = Xs6=t = Xs6=t = −UikUij Also, for i 6= k, j 6= l we have: UijUkl = Xs6=t = Xs6=t = UklUij uisukt ⊗ usjutl +Xs uktuis ⊗ utlusj +Xs uisuks ⊗ usjusl (−uksuis) ⊗ (−uslusj) This finishes the proof in the real case. In the complex case the remaining relations can (cid:3) be checked in a similar way, by putting ∗ exponents in the middle. It remains to twist the half-liberated quantum groups O∗ N . In order to do so, given three coordinates a, b, c ∈ {uij}, let us set span(a, b, c) = (r, c), where r, c ∈ {1, 2, 3} In other words, if we write are the number of rows and columns spanned by a, b, c. a = uij, b = ukl, c = upq then r = #{i, k, p} and l = #{j, l, q}. We have then: Proposition 3.4. We have quantum groups ¯O∗ N defined via N , ¯U ∗∗ N , U ∗∗ N ⊂ O+ N ⊂ U + αβγ =(−γβα for a, b, c ∈ {uij} with span(a, b, c) = (≤ 2, 3) or (3,≤ 2) otherwise γβα with the usual conventions α = a, a∗, β = b, b∗ and γ = c, c∗. 16 TEODOR BANICA Proof. The commutation/anticommutation signs in the statement are as follows: 2 r\c 1 3 1 + + − 2 + + − 3 − − + We first prove the result for ¯O∗ N . The construction of the counit, ε(uij) = δij, requires the Kronecker symbols δij to commute/anticommute according to the above table. Equiv- alently, we must prove that the situation δijδklδpq = 1 can appear only in a case where the above table indicates “+”. But this is clear, because δijδklδpq = 1 implies r = c. The construction of the antipode S is clear too, because this requires the choice of our ± signs to be invariant under transposition, and this is true, the table being symmetric. We are therefore left with the construction of ∆. With Uij =Pk uik ⊗ ukj, we have: UiaUjbUkc = Xxyz = Xxyz = ±UkcUjbUia uixujyukz ⊗ uxauybuzc ±ukzujyuix ⊗ ±uzcuybuxa We must prove that, when examining the precise two ± signs in the middle formula, their product produces the correct ± sign at the end. The point now is that both these signs depend only on s = span(x, y, z), and for s = 1, 2, 3 respectively: – For a (3, 1) span we obtain +−, +−, −+, so a product − as needed. – For a (2, 1) span we obtain ++, ++, −−, so a product + as needed. – For a (3, 3) span we obtain −−, −−, ++, so a product + as needed. – For a (3, 2) span we obtain +−, +−, −+, so a product − as needed. – For a (2, 2) span we obtain ++, ++, −−, so a product + as needed. Together with the fact that our problem is invariant under (r, c) → (c, r), and with the fact that for a (1, 1) span there is nothing to prove, this finishes the proof. For ¯U ∗∗ N the proof is similar, by putting ∗ exponents in the middle. Regarding the inclusions between these quantum groups, we have: (cid:3) Proposition 3.5. We have the following diagram of quantum groups, UN / U ∗∗ N / U + N ON O∗ N O+ N with all inclusions being proper at N ≥ 3. ¯U ∗∗ N ¯O∗ N ¯UN ¯ON / / o o o o / / O O / / O O O O O O o o O O o o LIBERATIONS AND TWISTS OF REAL AND COMPLEX SPHERES 17 Proof. The inclusions are clear, as in the proof of Proposition 1.5. For the properness assertion, we first compute the classical versions. Our claim is that these are as follows, with the 6 compact groups at right being different at N ≥ 3: UN / UN / UN YN KN ON ON ON XN HN Indeed, regarding the groups HN = ON ∩ ¯ON and KN = UN ∩ ¯UN , these appear respectively from ON , UN by assuming that the standard coordinates satisfy the relations ab = 0, for any a 6= b on the same row or the same column of u. We recognize here the hyperoctahedral group HN = Z2 ≀ SN , and its complex version KN = T ≀ SN . Regarding now XN , YN , these are certain compact groups, appearing respectively from ON , UN by assuming that the coordinates satisfy abc = 0, under the span conditions producing anticommutation in Proposition 3.4. Since these groups are different, and different as well from HN , KN at N ≥ 3, this finishes the proof of our claim. properness of the inclusions on the left, this is well-known from [5], [9]. We deduce that the inclusions on the right in the statement are all proper. As for the (cid:3) At N = 2 the situation is similar to the one for the spheres, the diagram of inclusions between the 10 quantum groups being: U2 O2 / U ∗∗ 2 / U + 2 ¯U ∗∗ 2 O+ 2 O+ 2 O+ 2 ¯U2 ¯ON This can be indeed deduced by using the same arguments as in the sphere case. Regarding now the relation with our 10 spheres, let us first recall: Definition 3.6. A quantum group action G y X consists in having a morphism of C ∗-algebras Φ : C(X) → C(G) ⊗ C(X) satisfying the following conditions: (1) Coassociativity: (id ⊗ Φ)Φ = (∆ ⊗ id)Φ. (2) Counitality: (ε ⊗ id)Φ = id. (3) Faithfulness: (Im Φ)(C(G) ⊗ 1) is dense in C(G) ⊗ C(X). The morphism in the statement is called coaction. See [5], [33]. Consider now one of our 10 quantum groups, denoted U × the corresponding sphere, with the correspondence between quantum groups and spheres being obtained by superposing the diagrams in Proposition 1.5 and Proposition 3.5. N . We denote by SN −1 × / / o o o o / / O O / / O O O O O O o o O O o o / / o o o o / / O O / / O O O O O O o o O O o o 18 TEODOR BANICA We denote the spherical coordinates by zi, in both the real and complex cases. We have the following result, that we will further improve in section 5 below: Theorem 3.7. We have an action U × N y SN −1 × , with the corresponding coaction map being given by Φ(zi) =Pj uij ⊗ zj. Proof. As a first observation, assuming that the formula Φ(zi) = Pj uij ⊗ zj produces indeed a morphism of algebras, the axioms in Definition 3.6 are clear, because they come from the fact that u = (uij) is a fundamental corepresentation for U × N . See [5]. In order to prove now that we have a morphism of algebras, we must check the fact that the following elements satisfy the defining relations for our spheres: Zi =Xj uij ⊗ zj We have 10 spheres to be investigated, and the proof goes as follows: 1-2. SN −1 C,+ . The result for SN −1 C,+ follows from: R,+ , SN −1 Xi Xi ZiZ ∗ i = Xijk i Zi = Xijk Z ∗ (uij ⊗ zj)(u∗ (u∗ ik ⊗ z∗ ik ⊗ z∗ k) =Xjk k)(uij ⊗ zj) =Xjk (ut ¯u)jk ⊗ zjz∗ (u∗u)kj ⊗ z∗ k =Xj kzj =Xj 1 ⊗ zjz∗ j = 1 1 ⊗ z∗ j zj = 1 Regarding now SN −1 R,+ , the result here follows by restriction, because when assuming zi = z∗ i , the relations Zi = Z ∗ i for any i are equivalent to uij = u∗ ij for any i, j. 3-6. SN −1 R , SN −1 C , ¯SN −1 R , ¯SN −1 C . The results in the classical cases are clear, because the actions in the statement are the usual ones, ON y SN −1 R and UN y SN −1 C . For the sphere ¯SN −1 R this follows from the following computation, with i 6= k: ZiZk = Xjl = Xj6=l = −Xjl uijukl ⊗ zjzl =Xj6=l ukluij ⊗ (−zlzj) +Xj ukluij ⊗ zlzj = −ZkZi uijukl ⊗ zjzl +Xj (−ukjuij) ⊗ z2 j uijukj ⊗ z2 j R For the sphere ¯SN −1 7-10. SN −1 the proof is similar, by adding ∗ exponents where needed. R,∗ , ¯SN −1 C,∗∗ . We only prove here the result in the twisted cases, the proof in the untwisted cases being similar, by removing all signs. Let us first discuss the sphere ¯SN −1 R,∗ . We have two sets of conditions to be checked, as follows: C,∗∗ , ¯SN −1 R,∗ , SN −1 The point now is that we can use the half-commutation relations for both the u and the z variables, and we obtain the formula of ZkZjZi, with the signs in front of the 5 sums being respectively +−,−+,−+,−+,−+. Thus we have ZiZjZk = −ZkZjZi. – For i 6= k we must have ZiZiZk = ZkZiZi. We have: uiauibukc ⊗ zazbzc +Xa6=c ZiZiZk = Xa,b,c distinct uiauiaukc ⊗ zazazc uiauibukb ⊗ zazbzb uiauibuka ⊗ zazbza +Xa6=b uiauiauka ⊗ zazaza +Xa6=b +Xa LIBERATIONS AND TWISTS OF REAL AND COMPLEX SPHERES 19 – For i, j, k distinct, we must have ZiZjZk = −ZkZjZi. We have: ZiZjZk = Xa,b,c distinct uiaujbukc ⊗ zazbzc +Xa6=c uiaujaukc ⊗ zazazc uiaujbuka ⊗ zazbza +Xa6=b uiaujauka ⊗ zazaza +Xa6=b +Xa uiaujbukb ⊗ zazbzb Once again, we can use the half-commutation relations for both the u and the z vari- ables, and we obtain the formula of ZkZiZi, with the signs in front of the 5 sums being respectively −−, ++, ++, ++, ++. Thus we have ZiZiZk = ZkZiZi. C,∗∗ is similar, by adding ∗ exponents where needed. The proof for ¯SN −1 (cid:3) Summarizing, the 10 quantum groups that we have constructed here act on the 10 spheres constructed in section 1. Improving Theorem 3.7, with a universality result for the actions constructed there, will be our main goal in what follows. 4. Schur-Weyl duality In order to get more insight into the structure of our 10 spheres, and into the structure of the actions constructed in Theorem 3.7 above, we need a number of new ingredients, and notably the Schur-Weyl theory for the 10 quantum groups. As in [8], we use several types of partitions, as follows: Definition 4.1. We let P (k, l) be the set of partitions between an upper row of k points and a lower row of l points, and consider the following subsets of P (k, l): (1) P2(k, l) ⊂ Peven(k, l): the pairings, and the partitions with blocks having even size. (2) NC2(k, l) ⊂ NCeven(k, l) ⊂ NC(k, l): the subsets of noncrossing partitions. (3) P erm(k, k) ⊂ P2(k, k): the pairings having only up-to-down strings. 20 TEODOR BANICA Observe that the elements of P erm(k, k) correspond to the permutations in Sk, with the usual convention that the permutation diagrams act downwards. See [8], [9]. Given a partition τ ∈ P (k, l), we call “switch” the operation which consists in switching two neighbors, belonging to different blocks, either in the upper row, or in the lower row. By performing a number of such switches, we can always transform τ into a certain noncrossing partition τ ′ ∈ NC(k, l), having the same block structure as τ . operation, in the particular case of the partitions having even blocks: Proposition 4.2. There is a signature map ε : Peven → {−1, 1}, given by ε(τ ) = (−1)c, where c is the number of switches needed to make τ noncrossing. In addition: We will need the following standard result, regarding the behavior of this switching (1) For τ ∈ P erm(k, k), this is the usual signature. (2) For τ ∈ P2 we have (−1)c, where c is the number of crossings. (3) For τ ≤ π ∈ NCeven, the signature is 1. Proof. In order to show that ε is well-defined, we must prove that the number c in the statement is well-defined modulo 2. It is enough to perform the verification for the non- crossing partitions. More precisely, given τ, τ ′ ∈ NCeven having the same block structure, we must prove that the number of switches c required for the passage τ → τ ′ is even. In order to do so, observe that any partition τ ∈ P (k, l) can be put in “standard form”, by ordering its blocks according to the appearence of the first leg in each block, counting clockwise from top left, and then by performing the switches as for block 1 to be at left, then for block 2 to be at left, and so on. Here the required switches are also uniquely determined, by the order coming from counting clockwise from top left. Here is an example of such an algorithmic switching operation, with block 1 being first put at left, by using two switches, then with block 2 left unchanged, and then with block 3 being put at left as well, but at right of blocks 1 and 2, with one switch: ◦ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ◦ → ✒✒✒✒✒✒✒✒✒ ◦ ◦ ◦ ◦ ☎☎☎☎☎☎☎☎☎☎☎ ✒✒✒✒✒✒✒✒✒ ◦ ◦ ◦ ◦ → ◦ ◦ ☎☎☎☎☎☎☎☎☎☎☎ ☎☎☎☎☎☎☎☎☎☎☎ ◦ ◦ ◦ ◦ ◦ ✒✒✒✒✒✒✒✒✒ → ✒✒✒✒✒✒✒✒✒ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ◦ The point now is that, under the assumption τ ∈ NCeven(k, l), each of the moves required for putting a leg at left, and hence for putting a whole block at left, requires an even number of switches. Thus, putting τ is standard form requires an even number of switches. Now given τ, τ ′ ∈ NCeven having the same block structure, the standard form coincides, so the number of switches c required for the passage τ → τ ′ is indeed even. Regarding now the remaining assertions, these are all elementary: (1) For τ ∈ P erm(k, k) the standard form is τ ′ = id, and the passage τ → id comes by composing with a number of transpositions, which gives the signature. δπ(cid:18)i1 . . . ik where δπ ∈ {−1, 0, 1} is constructed, in terms of τ = ker(i Tπ(ei1 ⊗ . . . ⊗ eik) = Xj1...jl j1 . . . jl(cid:19) ej1 ⊗ . . . ⊗ ejl j), as follows: LIBERATIONS AND TWISTS OF REAL AND COMPLEX SPHERES 21 (2) For a general τ ∈ P2, the standard form is of type τ ′ = . . .∪...∪ τ → τ ′ requires c mod 2 switches, where c is the number of crossings. (3) Assuming that τ ∈ Peven comes from π ∈ NCeven by merging a certain number of blocks, we can prove that the signature is 1 by proceeding by recurrence. Indeed, we can first assume that we have only 3 blocks, and then we can further use a recurrence on the number of legs, until we reach to the situation where the block in the middle, which crosses the merged outer blocks, is a semicircle, and where the result is clear. (cid:3) ∩...∩, and the passage With the above notion in hand, we can formulate: Definition 4.3. Associated to a pair-partition π ∈ P2(k, l) are the linear maps (1) In the untwisted case, we set δ = 1 if τ ≤ π, and δ = 0 otherwise. (2) In the twisted case, we set ¯δ = ε(τ ) if τ ≤ π, and ¯δ = 0 otherwise. In the untwisted case we recognize here the usual Brauer intertwiners for ON , discussed for instance in [4], [8], and whose formula is simply: Tπ(ei1 ⊗ . . . ⊗ eik ) = Xj:ker(i j )≤π ej1 ⊗ . . . ⊗ ejl In the twisted case the formula is similar, but requiring this time some signs, constructed according to Proposition 4.2 above. More precisely, we have: ¯Tπ(ei1 ⊗ . . . ⊗ eik ) =Xτ ≤π ε(τ ) Xj:ker(i j )=τ ej1 ⊗ . . . ⊗ ejl Let us work now out a few basic examples of such linear maps, which are of particular interest for the considerations to follow: Proposition 4.4. The linear map associated to the basic crossing is: ¯T/\(ei ⊗ ej) =(−ej ⊗ ei ej ⊗ ei for i 6= j otherwise The linear map associated to the half-liberating permutation (321) ∈ S3 is: ¯T/\ (ei ⊗ ej ⊗ ek) =(−ek ⊗ ej ⊗ ei ek ⊗ ej ⊗ ei for i, j, k distinct otherwise Also, for any noncrossing pairing π ∈ NC2, we have ¯Tπ = Tπ. 22 TEODOR BANICA Proof. We have to compute the signature of the various partitions involved, and we can use here (1,2,3) in Proposition 4.2. We make the convention that the strings which cross and which are of the same type (e.g. dotted) correspond to the same block. Regarding the basic crossing and its collapsed version, the signatures are: ◦ ◦ ✹✹✹✹✹✹✹ ◦ ◦ → −1 ◦ ◦ ✹✹✹✹✹✹✹ ✡✡✡✡✡✡✡ ◦ ◦ → 1 But this gives the first formula in the statement. Regarding now the second formula, this follows from the following signature computations, obtained by counting the crossings (in the first case), by switching twice as to put the partition in noncrossing form (in the next 3 cases), and by observing that the partition is noncrossing (in the last case): ◦ ◦ ◦ ◦ ◦u ◦ → −1 ◦ ◦ ◦ ◦ ◦ ◦ → 1 ◦ ◦ ◦ ◦ ◦u ◦ → 1 ◦ ◦ ◦ ◦ ◦ ◦ → 1 ◦ ◦ ◦ ◦ ◦ ◦ → 1 Finally, the last assertion follows from Proposition 4.2 (3). (cid:3) The relation with the 10 quantum groups comes from: Proposition 4.5. For an orthogonal quantum group G, the following hold: (1) T/\ ∈ End(u⊗2) precisely when G ⊂ ON . (2) T/\ ∈ End(u⊗3) precisely when G ⊂ O∗ N . Proof. These results are well-known in the untwisted case, see [8], [9]. (1) By using the formula of ¯T/\ in Proposition 4.4, we obtain: ( ¯T/\ ⊗ 1)u⊗2(ei ⊗ ej ⊗ 1) =Xk ek ⊗ ek ⊗ ukiukj −Xk6=l u⊗2( ¯T/\ ⊗ 1)(ei ⊗ ej ⊗ 1) =(Pkl el ⊗ ek ⊗ uliuki −Pkl el ⊗ ek ⊗ uljuki el ⊗ ek ⊗ ukiulj if i = j if i 6= j 5 w 7 y 9 { ; } = @  B 5 w 7 y 9 { ; } = @  B LIBERATIONS AND TWISTS OF REAL AND COMPLEX SPHERES 23 For i = j the conditions are u2 ki for any k, and ukiuli = −uliuki for any k 6= l. For i 6= j the conditions are ukiukj = −ukjuki for any k, and ukiulj = uljuki for any k 6= l. Thus we have exactly the relations between the coordinates of ¯ON , and we are done. ki = u2 (2) By using the formula of ¯T/\ in Proposition 4.4, we obtain: ( ¯T/\ ⊗ 1)u⊗2(ei ⊗ ej ⊗ ek ⊗ 1) = Xabc not distinct − Xa,b,c distinct ec ⊗ eb ⊗ ea ⊗ uaiubjuck ec ⊗ eb ⊗ ea ⊗ uaiubjuck The product in the other sense is given by: u⊗2( ¯T/\ ⊗ 1)(ei ⊗ ej ⊗ ek ⊗ 1) =(Pabc ec ⊗ eb ⊗ ea ⊗ uckubjuai −Pabc ec ⊗ eb ⊗ ea ⊗ uckubjuai for i, j, k not distinct for i, j, k distinct For i, j, k not distinct the conditions are uaiubjuck = uckubjuai for a, b, c not distinct, and uaiubjuck = −uckubjuai for a, b, c distinct. For i, j, k distinct the conditions are uaiubjuck = −uckubjuai for a, b, c not distinct, and uaiubjuck = uckubjuai for a, b, c distinct. Thus we have exactly the relations between the coordinates of ¯O∗ N , and we are done. (cid:3) We prove now that the usual categorical operations on the linear maps Tπ, namely the composition, tensor product and conjugation, are compatible with the usual categorical operations on the partitions from [8], namely the composition (π, σ) → [σ π], the horizontal concatenation (π, σ) → [πσ], and the upside-down turning π → π∗. We have: Proposition 4.6. The assignement π → Tπ is categorical, in the sense that Tπ ⊗ Tσ = T[πσ], Tπ Tσ = N c(π,σ) T[σ π], T ∗ π = Tπ∗ where c(π, σ) are certain positive integers. Proof. By using the definition of π → Tπ, we just have to understand the behaviour of the generalized Kronecker symbol construction π → δπ, under the various categorical operations on the partitions π. We have to check three conditions, as follows: 1. Concatenation. In the untwisted case, this follows from the following formula: In the twisted case, it is enough to check the following formula: δπ(cid:18)i1 . . . ip j1 . . . jq(cid:19) δσ(cid:18)k1 . . . kr j1 . . . jq(cid:19)(cid:19) ε(cid:18)ker(cid:18)k1 . . . kr l1 . . . ls(cid:19) = δ[πσ](cid:18)i1 . . . ip k1 . . . kr l1 . . . ls(cid:19) l1 . . . ls(cid:19)(cid:19) = ε(cid:18)ker(cid:18)i1 . . . ip k1 . . . kr j1 . . . jq j1 . . . jq l1 . . . ls(cid:19)(cid:19) ε(cid:18)ker(cid:18)i1 . . . ip Let us denote by τ, ν the partitions on the left, so that the partition on the right is of the form ρ ≤ [τ ν]. Now by switching to the noncrossing form, τ → τ ′ and ν → ν′, the 24 TEODOR BANICA partition on the right transforms into ρ → ρ′ ≤ [τ ′ν′]. Now since [τ ′ν′] is noncrossing, we can use Proposition 4.2 (3), and we obtain the result. 2. Composition. In the untwisted case, this follows from the following formula from [8], where c(π, σ) is the number of closed loops obtained when composing: δπ(cid:18)i1 . . . ip j1 . . . jq(cid:19) δσ(cid:18) j1 . . . jq k1 . . . kr(cid:19) = N c(π,σ)δ[π σ](cid:18) i1 . . . ip k1 . . . kr(cid:19) Xj1...jq In order to prove now the result in the twisted case, it is enough to check that the signs match. More precisely, we must establish the following formula: ε(cid:18)ker(cid:18)i1 . . . ip j1 . . . jq(cid:19)(cid:19) ε(cid:18)ker(cid:18) j1 . . . jq k1 . . . kr(cid:19)(cid:19) = ε(cid:18)ker(cid:18) i1 . . . ip k1 . . . kr(cid:19)(cid:19) Let τ, ν be the partitions on the left, so that the partition on the right is of the form ρ ≤ [τ ν]. Our claim is that we can jointly switch τ, ν to the noncrossing form. Indeed, we can first switch as for ker(j1 . . . jq) to become noncrossing, and then switch the upper legs of τ , and the lower legs of ν, as for both these partitions to become noncrossing. Now observe that when switching in this way to the noncrossing form, τ → τ ′ and ν ′] is ν → ν′, the partition on the right transforms into ρ → ρ′ ≤ [τ ′ noncrossing, we can apply Proposition 4.2 (3), and we obtain the result. ν ′]. Now since [τ ′ 3. Involution. Here we must prove the following formula: But this is clear, both in the untwisted and twisted cases, and we are done. δπ(cid:18)i1 . . . ip j1 . . . jq(cid:19) = δπ∗(cid:18)j1 . . . jq i1 . . . ip(cid:19) (cid:3) In order to formulate the duality result, we use words α, β, . . . over the symbols u, ¯u. Given such a word α, we denote by u⊗α the corepresentation obtained by performing the corresponding tensor product, by inserting ⊗ signs between the u, ¯u symbols. cyclically the legs • ◦ • ◦ . . ., each string joins a black leg to a white leg. See [9]. 2 ⊂ P2 the set of pairings having the property that when labelling With these conventions, the Schur-Weyl duality result is as follows: Also, we denote by P ∗ Theorem 4.7. We have Hom(u⊗α, u⊗β) = span( Tππ ∈ P×(α, β)), where the sets of diagrams for the the 10 quantum groups, with inclusions between them, are P2 P2 P ∗ 2 P ∗ 2 NC2 NC2 / P ∗ 2 / P ∗ 2 P2 / P2 with the convention P×(α, β) = P×(α,β), where . is the word length, and where the upper subsets P×(α, β) ⊂ P×(α, β) consist of partitions with strings joining u, ¯u.   o o   o o   / / /     o o o o / / LIBERATIONS AND TWISTS OF REAL AND COMPLEX SPHERES 25 Proof. In the real untwisted case, all the diagrams are already known, see [8], [9]. In the complex untwisted case, the proof is similar. In the twisted case now, the result for ¯ON follows as in [4], by using Proposition 4.5 (1), Proposition 4.6, and Tannakian duality [44]. For the other twisted quantum groups, the result follows by functoriality, as in [8], [9], by using Proposition 4.5, and by adding ∗ exponents where needed. (cid:3) 5. Affine isometries In this section we go back to Theorem 3.7, and improve the result found there. It is known since [13] that proving universality results for quantum group actions re- quires a good knowledge of the linear relations satisfied by the various products of coor- dinates. And we can deal now with such problems, by using Schur-Weyl duality. We will need the Weingarten integration formula. We begin with: Definition 5.1. Let P2(l) = P2(0, l). For π ∈ P2(l) we set: δπ(j1 . . . jl)ej1 ⊗ . . . ⊗ ejl ξπ = Xj1...jl In other words, we denote by ξπ the vector Tπ constructed in Definition 4.3. In the classical case, we recognize the usual Brauer fixed vectors for ON . In the twisted case, the formula is similar, this time making appear some signatures: ¯ξπ =Xτ ≤π ε(τ ) Xj:ker j=τ ej1 ⊗ . . . ⊗ ejl Here are a few examples of such vectors, coming from the computations in Proposition 4.4 above. First, the vector associated to the basic crossing is: ¯ξ∩∩ =Xi ei ⊗ ei ⊗ ei ⊗ ei −Xi6=j ei ⊗ ej ⊗ ei ⊗ ej Also, the vector associated to the half-liberating pairing (123123) is: ¯ξ∩∩∩ = Xijk not distinct Finally, observe that for any noncrossing pairing π ∈ NC2(l), we have ¯ξπ = ξπ. We will need the following simple fact: ei ⊗ ej ⊗ ek ⊗ ei ⊗ ej ⊗ ek − Xi,j,k distinct ei ⊗ ej ⊗ ek ⊗ ei ⊗ ej ⊗ ek Proposition 5.2. The scalar products between the vectors ξπ are given by and hence coincide in the twisted and the untwisted cases. < ξπ, ξσ >= N π∨σ 26 TEODOR BANICA Proof. In the twisted case, we have the following computation: < ¯ξπ, ¯ξσ > = * Xj:ker j≤π = Xj:ker j≤(π∨σ) ε(ker j)ej1 ⊗ . . . ⊗ ejl, Xj:ker j≤σ ε(ker j)2 = Xj:ker j≤(π∨σ) 1 = N π∨σ ε(ker j)ej1 ⊗ . . . ⊗ ejl+ In the untwisted case the computation is similar, with the signs dissapearing right from (cid:3) the beginning. Thus, in both cases we obtain the formula in the statement. Given one of our quantum groups U × N , and an exponent vector α = (α1, . . . , αk) ∈ {1,∗}k, we denote by P×(α) = P×(∅, α) the set of pairings found in Theorem 4.7 above for U × N , having no upper points, and having the lower points labelled by the entries of α, according to the identifications u → 1, ¯u → ∗. With this convention, we have: Proposition 5.3. We have the Weingarten type formula i1j1 . . . uαk uα1 ikjk Z U × N δπ(i1 . . . ik) δσ(j1 . . . jk)W α = Xπ,σ∈P×(α) kN (π, σ) = N π∨σ, for π, σ ∈ P×(α). kN (π, σ) where W α kN = (Gα kN )−1, with Gα Proof. This follows indeed as in [4], by using Theorem 4.7 and Proposition 5.2. Observe that the Weingarten matrix is the same in the twisted and the untwisted cases. (cid:3) Now back to the spheres, we first have the following result: Lemma 5.4. The linear relations satisfied by the variables rij = zizj are as follows: (1) For SN −1 (2) For the remaining 8 spheres, these elements are linearly independent. we have rij = ±rji, and no other relations. , ¯SN −1 R R In addition, a similar result holds for the variables cij = ziz∗ j . Proof. We first prove the assertion regarding the variables rij = zizj. We have 10 spheres to be investigated, and the proof goes as follows: 1-2. SN −1 3-4. ¯SN −1 R R C , SN −1 , ¯SN −1 C . The results here are clear. . We prove first the result for ¯SN −1 R where uij are the standard coordinates on ¯ON . We have: . We use the model zi → Zi = u1i, ¯δσ(i, j, l, k)W4N (π, σ) < ZiZj, ZkZl >=Z ¯ON u1iu1ju1lu1k = Xπ,σ∈P2(4) Since P2(4) = {∩∩, ⋓,∩∩}, the Weingarten matrix on the right is given by: N(N − 1)(N + 2) N + 1 −1 −1 −1 N + 1 −1  −1 N N N 2  W4N =  −1 N + 1  N 2 N N N N 2 N = −1 1 LIBERATIONS AND TWISTS OF REAL AND COMPLEX SPHERES 27 We conclude that we have the following formula: < ZiZj, ZkZl >= 1 N(N + 2) Xσ∈P2(4) ¯δσ(i, j, l, k) The matrix on the right, taken with indices i ≤ j and k ≤ l, is then invertible. Thus For the sphere ¯SN −1 the variables ZiZj are linearly independent, and so must be the variables zizj. , a similar computation, using now a ¯UN model, gives: C < ZiZj, ZkZl >=Z ¯UN u1iu1ju∗ 1lu∗ 1k = Xπ,σ∈P2(11∗∗) ¯δσ(i, j, l, k)W 11∗∗ 4N (π, σ) We have P2(11 ∗ ∗) = {⋓,∩∩}, and the corresponding Weingarten matrix is: W 11∗∗ 4N =(cid:18)N 2 N N N 2(cid:19)−1 = 1 N(N 2 − 1)(cid:18) N −1 −1 N(cid:19) We therefore obtain the following formula: 1 < ZiZj, ZkZl >= N(N + 1) Xσ∈P2(11∗∗) ¯δσ(i, j, l, k) Once again, since the matrix on the right is invertible, we obtain the result. 5-6. SN −1 R,∗ . We can use here a 2 × 2 matrix trick from [17]. Consider indeed one R,∗ , ¯SN −1 of the spheres SN −1 C / ¯SN −1 C , with coordinates denoted y1, . . . , yN , and let us set: As explained in the proof of Theorem 1.7 above, these matrices produce models for R,∗ , ¯SN −1 SN −1 R,∗ . Now observe that the elements rij = zizj map in this way to: Zi =(cid:18) 0 y∗ i yi 0(cid:19) Rij = ZiZj =(cid:18) 0 y∗ i yi 0(cid:19)(cid:18) 0 y∗ j yj 0(cid:19) =(cid:18)yiy∗ 0 j 0 y∗ i yj(cid:19) Thus, the result follows from the result for ¯SN −1 7-10. SN −1 C,∗∗ , ¯SN −1 R,+ , SN −1 C,+ , SN −1 R C,∗∗ . The results here follow simply by functoriality, from , ¯SN −1 C , established above. those established above, for the smaller spheres SN −1 R,∗ , ¯SN −1 R,∗ . Finally, the proof of the last assertion is similar, with no new computations needed in the real case, where rij = cij, and with the same Weingarten matrix, this time coming from the set P2(1 ∗ 1∗) = {∩∩, ⋓}, appearing in the complex case. (cid:3) We can improve now Theorem 3.7 above. First, we have: Proposition 5.5. ON , UN , ¯ON , ¯UN , O+ ing on their respective spheres, by leaving span(zi) invariant. N , U + N are the biggest compact quantum groups act- 28 TEODOR BANICA Proof. The fact that span(zi) is left invariant means that the coaction must be of the 1-2. SN −1 R,+ , SN −1 C,+ . Let us go back to the proof of Theorem 3.7. As explained there, the coaction axioms are equivalent to the fact that u = (uij) is a corepresentation. Also, with form Φ(zi) =Pj uij ⊗ zj. We have six situations to be investigated, as follows: the notation Zi =Pj uij ⊗ zj, we know from there that we have: i Zi =Xjk k, Xi (u∗u)kj ⊗ z∗ kzj (ut¯u)jk ⊗ zjz∗ i =Xjk ZiZ ∗ Z ∗ Now by using Lemma 5.4 above for the free spheres, we deduce that the conditions i Zi = 1 are equivalent to the conditions ut¯u = u∗u = 1. Now since u is already known to be a corepresentation, by the results of Woronowicz in [43] it follows that u must be a biunitary corepresentation, and we are done. 3-4. SN −1 R , SN −1 C . For the sphere SN −1 R this is done in [13]. We reproduce here the proof, in view of some further extensions and modifications. First, we have: Xi i =Pi Z ∗ Pi ZiZ ∗ Φ(zizj) = Xkl = Xk = Xk≤l uikujl ⊗ zkzl uikujk ⊗ z2 (uikujl + uilujk) ⊗(cid:18)1 − k +Xk<l δkl 2 (cid:19) zkzl (uikujl + uilujk) ⊗ zkzl We deduce from this that Φ maps the commutators [zi, zj] as follows: Φ([zi, zj]) = Xk≤l = Xk≤l (uikujl + uilujk − ujkuil − ujluik) ⊗(cid:18)1 − ([uik, ujl] − [ujk, uil]) ⊗(cid:18)1 − 2 (cid:19) zkzl δkl δkl 2 (cid:19) zkzl Now since the variables {zkzlk ≤ l} are linearly independent, we obtain from this [uik, ujl] = [ujk, uil], for any i, j, k, l. Moreover, if we apply now the antipode we further obtain [ulj, uki] = [uli, ukj], and by relabelling, [uik, ujl] = [uil, ujk]. We therefore conclude that we have [uik, ujl] = 0 for any i, j, k, l, and this finishes the proof. See [13]. For SN −1 C the beginning of the proof is similar, and gives [uik, ujl] = [ujk, uil]. Now if we apply the antipode followed by the involution we obtain as before [ulj, uki] = [uli, ukj], then [uik, ujl] = [uil, ujk], and finally [uik, ujl] = 0. Thus the coordinates are subject to the commutation relations ab = ba, and by using a standard categorial trick, mentioned before Definition 2.2 above, we have as well ab∗ = b∗a, and we are done. LIBERATIONS AND TWISTS OF REAL AND COMPLEX SPHERES 29 5-6. ¯SN −1 R , ¯SN −1 C Lemma 5.4, and by adding signs where needed. First, for ¯SN −1 . The proof here is similar to the above proof for SN −1 we have: R R , SN −1 C , by using Φ(zizj) =Xk uikujk ⊗ z2 k +Xk<l (uikujl − uilujk) ⊗ zkzl We deduce that with [[a, b]] = ab + ba we have the following formula: Φ([[zi, zj]]) =Xk [[uik, ujk]] ⊗ z2 k +Xk<l ([uik, ujl] − [uil, ujk]) ⊗ zkzl Now assuming i 6= j, we have [[zi, zj]] = 0, and we therefore obtain [[uik, ujk]] = 0 for any k, and [uik, ujl] = [uil, ujk] for any k < l. By applying the antipode and then by relabelling, the latter relation gives [uik, ujl] = 0, and we are done. The proof for ¯SN −1 to deduce from the relations ab = ±ba the remaining relations ab∗ = ±b∗a. is similar, by using the above-mentioned categorical trick, in order (cid:3) C In order to deal with the half-liberated cases, we will need: C,∗∗ , ¯SN −1 Lemma 5.6. Consider one of the spheres SN −1 R,∗ , SN −1 R,∗ , ¯SN −1 C,∗∗ . (1) The variables zazbzc with a < c and a, b, c distinct are linearly independent. (2) These variables are independent as well from any zazbzc with a, b, c not distinct. In addition, a similar result holds for the variables of type zaz∗ b zc. Proof. We use the same method as in the proof of Lemma 5.4, with models coming from the quantum groups O∗ N . For the quantum groups O∗ N , we have: N , ¯U ∗∗ N , ¯O∗ N , U ∗∗ N , ¯O∗ < ZaZbZc, ZiZjZk >=Z O∗ N u1au1bu1cu1ku1ju1i = Xπ,σ∈P ∗ 2 (6) ¯δσ(a, b, c, k, j, i)W6N (π, σ) The set P ∗ 2 (6) ≃ P ∗ 2 (3, 3) is by definition formed by the following pairings: • ◦ ◦ ◦ • ◦ ◦ ◦ • ◦ • Now observe that the scalar products of each of these pairings with all the 6 pairings are always, up to a permutation of the terms, N 3, N 2, N 2, N 2, N, N. Thus the Gram matrix is stochastic, G6N ξ = ξ, where ξ = (1, . . . , 1)t is the all-one vector. Thus we have • ◦ ◦ • • ◦ • ◦ • ◦ • ◦ ◦ • ●●●●●●●●●●● ◦ • • • ✇✇✇✇✇✇✇✇✇✇✇ ◦ • ●●●●●●●●●●● ✇✇✇✇✇✇✇✇✇✇✇ ◦ • ◦ • • 30 TEODOR BANICA W6N ξ = W6N G6N ξ = ξ, and so the Weingarten matrix is stochastic too. We conclude that, up to a universal constant depending only on N, we have: < ZaZbZc, ZiZjZk >∼ Xσ∈P ∗ 2 (6) ¯δσ(a, b, c, k, j, i) Now by computing the rank of this matrix, we obtain the result. Regarding now the last assertion, this follows from the same computation. comparing the products of type ZaZ ∗ the pairings in P ∗ Indeed, b Zc leads to the same formula and conclusion, because (cid:3) 2 (6) are all compatible with the leg labelling 1 ∗ 1 ∗ 1∗. We have now all ingredients for fully improving Theorem 3.7 above, with the remark that we will further process this result in section 6 below: Theorem 5.7. Each quantum group U × acting on its respective sphere SN −1 × , by leaving span(zi) invariant. N is the biggest (universal) compact quantum group Proof. In view of Proposition 5.5, we just have to discuss the 4 half-liberated cases. The idea here will be that for the spheres SN −1 , ¯SN −1 R,∗ , SN −1 , ¯SN −1 , SN −1 similar to the one for the spheres SN −1 [uia, ukc] = uiaukc − ukcuia by quantities of type [uia, ujb, ukc] = uiaujbukc − ukcujbuia. coaction on ¯SN −1 We only discuss the twisted case, the proof in the untwisted case being similar. For a R,∗ , we have two sets of conditions to be verified, as follows: R C,∗∗ , ¯SN −1 R,∗ , ¯SN −1 C,∗∗ the proof will be , by replacing the commutators C R C – For i, j, k distinct, we must have ZiZjZk = −ZkZjZi. We have: ZiZjZk = Xa,b,c distinct uiaujbukc ⊗ zazbzc +Xa6=c uiaujaukc ⊗ zazazc uiaujbuka ⊗ zazbza +Xa6=b uiaujauka ⊗ zazaza +Xa6=b +Xa uiaujbukb ⊗ zazbzb Now by using Lemma 5.6, all three sums appearing at left must vanish, and the 2 sums on the right must add up to 0 too. From the vanishing of the first sum we conclude, by proceeding as in the proof of Proposition 5.5, that the coordinates uia satisfy the relations abc = cba, when their span is (3, 3). Similarly, from the vanishing of the other sums we obtain abc = −cba for a (3, 2) span, and abc = −cba for a (3, 1) span. LIBERATIONS AND TWISTS OF REAL AND COMPLEX SPHERES 31 – For i 6= k we must have ZiZiZk = ZkZiZi. We have: uiauibukc ⊗ zazbzc +Xa6=c ZiZiZk = Xa,b,c distinct uiauiaukc ⊗ zazazc uiauibukb ⊗ zazbzb uiauibuka ⊗ zazbza +Xa6=b uiauiauka ⊗ zazaza +Xa6=b +Xa From the first sum we get abc = −cba for a (3, 2) span, from the next three sums we get abc = cba for a (2, 2) span, and from the last sum we get abc = cba for a (2, 1) span. Since we have as well, trivially, abc = cba for a (1, 1) span, we have reached to the defining relations for the quantum group ¯O∗ Finally, the proof for the sphere ¯SN −1 and by using the last assertion in Lemma 5.6. N , and we are done. C,∗∗ is similar, by adding ∗ exponents in the middle, (cid:3) Observe that Theorem 5.7 above is a quantum isometry group computation, in the affine sense of [32]. More precisely, if we define the affine actions on the real/complex spheres to be the actions of closed subgroups G ⊂ O+ N given by coaction maps of type Φ(zi) = Pj uij ⊗ zj, then Theorem 5.7 computes the corresponding quantum isometry groups. We refer to [32] for full details regarding the affine action formalism. N /U + 6. Further results, conclusion We discuss in this section a number of further topics, including the construction and basic properties of the integration functional for our 10 spheres, the Riemannian aspects of these spheres, and a proposal for an extended formalism, comprising 18 spheres. In order to construct the integration, we use the associated quantum group: Definition 6.1. Given one of the spheres SN −1 group, and we let R× N ⊂ C(U × N ) be the subalgebra generated by u11, . . . , u1N . × , we denote by U × N the associated quantum By the universal property of C(SN −1 × ) → R× xi → u1i, and by composing with the restriction of the Haar functional I : C(U × we obtain a trace tr : C(SN −1 × ) → C. In order to prove that tr is ergodic, we use: × ) we have a morphism π : C(SN −1 N mapping N ) → C, Lemma 6.2. The following formula holds, over the sphere SN −1 × , δπ(j1, . . . , jl)zα1 j1 . . . zαl jl = 1 Xj1...jl for any exponent vector α = (α1, . . . , αk) ∈ {1,∗}k, and any pairing π ∈ P×(α). 32 TEODOR BANICA Proof. In the untwisted case this was proved in [5]. Let us discuss now the case of ¯SN −1 By switching as for putting π in standard form, π′ = ⊓ . . .⊓, we obtain: ε(ker j)zj1 . . . zjl R . Xj1...jl ¯δπ(j1, . . . , jl)zj1 . . . zjl = Xj:ker j≤π = Xj ′:ker j ′≤π′ = Xj ′ 3...j ′ 1j ′ l−1 zj ′ 1 . . . zj ′ l z2 j ′ 1 . . . z2 j ′ l−1 = 1 For the sphere ¯SN −1 Pj zjz∗ using abc → cba switches as in [5], and in the free cases the result is clear. the proof is similar, with the last equality coming this time from j zj = 1. Finally, in the half-liberated cases the proof is similar as well, by (cid:3) j =Pj z∗ C Now back to the trace constructed above, we have here: Proposition 6.3. Consider the trace tr : C(SN −1 canonical surjection onto the first row algebra of U × × ) → C obtained by composing the N with the Haar functional. (1) tr is invariant, (id ⊗ tr)Φ(x) = tr(x)1. (2) tr is ergodic, (I ⊗ id)Φ = tr(.)1. (3) tr is the unique positive unital invariant trace on C(SN −1 × ). Proof. We use a general method from [5], which was further developed in [7]. The idea is that the result will follow by using the Weingarten integration formula: (1) This is clear, by using the invariance of the Haar integral of C(U × (2) It is enough to check the equality on a product zα1 ik . The left term is: i1 . . . zαk N ). (I ⊗ id)Φ(zα1 i1 . . . zαk ik ) = Xj1...jk I(uα1 i1j1 . . . uαk ikjk)zα1 j1 . . . zαk jk = Xj1...jk Xπ,σ∈P×(α) = Xπ,σ∈P×(α) δπ(i) δσ(j)W α kN (π, σ)zα1 j1 . . . zαk jk δπ(i)W α kN (π, σ) Xj1...jk δσ(j)zα1 j1 . . . zαk jk By using Lemma 6.2 the sum on the right is 1, so we get: (I ⊗ id)Φ(zα1 i1 . . . zαk ik ) = Xπ,σ∈P×(α) δπ(i)W α kN (π, σ)1 LIBERATIONS AND TWISTS OF REAL AND COMPLEX SPHERES 33 On the other hand, another application of the Weingarten formula gives: tr(zα1 i1 . . . zαk ik )1 = I(uα1 1i1 . . . uαk 1ik )1 = Xπ,σ∈P×(α) = Xπ,σ∈P×(α) δπ(1) δσ(i)W α kN (π, σ)1 δσ(i)W α kN (π, σ)1 Since the Weingarten function is symmetric in π, σ, this finishes the proof. (3) Let τ : C(SN −1 × ) → C be a trace satisfying the invariance condition. We have: τ (I ⊗ id)Φ(x) = (I ⊗ τ )Φ(x) = I(id ⊗ τ )Φ(x) = I(τ (x)1) = τ (x) On the other hand, according to the formula in (2) above, we have as well: Thus we obtain τ = tr, which finishes the proof. τ (I ⊗ id)Φ(x) = τ (tr(x)1) = tr(x) (cid:3) We make now the following convention, for the reminder of this paper: Definition 6.4. We agree from now on to replace each algebra C(SN −1 completion with respect to the canonical trace, coming from U × N . × ) with its GNS As a first observation, the classical spheres SN −1 are left unchanged by this modification, because the trace comes from the usual uniform measure on them. The free spheres SN −1 C,+ are however “cut” by this construction, for instance because this happens at the quantum group level, since O+ N are not coamenable. See [33]. R,+ , SN −1 N , U + , SN −1 R C Regarding the various half-liberations and twists, here we do not know. The faithfulness question for the trace of SN −1 R,∗ , which was raised some time ago in [5], is still open. The point with the above convention is that we have: Proposition 6.5. The following algebras, with generators and traces, are isomorphic: (1) The algebra C(SN −1 (2) The row algebra R× × ), with generators z1, . . . , zN , and with the trace. N ⊂ C(U × N ) generated by u11, . . . , u1N , with the integration. Proof. Consider the canonical quotient map π : C(SN −1 N , used in the proof of Proposition 6.2. The invariance property of the integration functional I : C(U × N ) → C shows that tr′ = Iπ satisfies the invariance condition in Proposition 6.2, so we have tr = tr′. Together with the positivity of tr and with the basic properties of the GNS construction, this shows that π is indeed an isomorphism, and we are done. (cid:3) × ) → R× As in [5], we can now construct spectral triples for our spheres, in some weak sense. C,+ , and so we have surjective ), and we can construct the Laplacian filtration as The idea is that we have inclusions SN −1 maps C(SN −1 projection/pullback of the Laplacian filtration for SN −1 × ⊂ SN −1 C,+ / SN −1 × ) → C( SN −1 C,+ ) → C( SN −1 ⊂ SN −1 . R R R 34 TEODOR BANICA More precisely, we have the following construction: × i1 . . . zαr Definition 6.6. Associated to each sphere SN −1 A = C(SN −1 zi, z∗ is the spectral triple (A, H, D), where × ), the dense subalgebra A is the linear span of all finite words in the generators i , and the operator D acting on H = L2(A, tr) is defined as follows: (1) Set Hk = span(zα1 (2) Define Ek = Hk ∩ H ⊥ (3) Finally, set Dx = λkx, for any x ∈ Ek, where λk are distinct numbers. As pointed out in [5], it is quite unclear what the correct eigenvalues should be. In the various half-liberated cases the problem can be probably approached by using the geometry of the associated projective planes [26]. In the free cases the situation seems to require the use of advanced analytic techniques, like those in [20], [28]. k−1, so that we have H = ⊕∞ ir r ≤ k, α ∈ {1,∗}r). k=0Ek. This type of issue is in fact well-known in the quantum group context, for noncommu- tative manifolds constructed by using various liberation procedures. See [16]. Without precise eigenvalues, we are in fact in the orthogonal filtration framework of [6], [38]. As explained there, having such a filtration suffices for constructing a quantum isometry group. In our case, we can formulate the following result: Theorem 6.7. We have G+(SN −1 spectral triple sense of [31], for all the 5 real spheres. × ) = O× N , with the quantum isometry group taken in the Proof. This was proved in [5] in the untwisted case, and the proof in the twisted case is similar. Consider indeed the standard coaction Φ : C(SN −1 N ) ⊗ C(SN −1 × ). This extends to a unitary representation on the GNS space H × N , that we denote by U. We have Φ(Hk) ⊂ C(O× N ) ⊗ Hk, which reads U(Hk) ⊂ Hk. By unitarity we get as well U(H ⊥ k , so each Ek is U-invariant, and U, D must commute. Thus, Φ is isometric with respect to D. Finally, the universality of O× (cid:3) × ) → C(O× N follows from Theorem 5.7. k ) ⊂ H ⊥ In the complex case the situation is more delicate, and would require a good understand- ing of the notion of complex affine action, in the noncommutative Riemannian geometry setting. For an exposition of some of the technical difficulties here, see [32]. There are of course many other questions regarding our 10 spheres, and their geometry. Besides the two fundamental questions raised above, regarding the faithfulness of the trace on the full algebra, and the construction of the eigenvalues, further interesting questions regard orientability issues, and the existence of a Dirac operator, cf. [14], [23]. To summarize, regarding the geometric structure of our spheres, we have so far more questions than answers. We intend to clarify the situation in a future paper. We would like to discuss now a possible extension of our formalism, from 10 to 18 spheres. The idea is that such an extension should come in three steps, as follows: I. First, the projective planes for the 10 spheres can be computed by using methods from C , P N R,+ [5], [9], by using Schur-Weyl duality. These are as follows, where P N C , ¯P N R , ¯P N R , P N LIBERATIONS AND TWISTS OF REAL AND COMPLEX SPHERES 35 are by definition the projective planes for SN −1 R , SN −1 C , ¯SN −1 R , ¯SN −1 C , SN −1 R,+ : P N C / P N C / P N R,+ ¯P N C ¯P N C P N R P N C P N R,+ ¯P N C ¯P N R II. We recall from [1] that the free complexification operation amounts in multiplying the standard coordinates by a unitary which is free from them. The free complexifications of the 10 spheres can be computed by using the projective planes and techniques from [1], [36], the conclusion being that the diagram is as follows, with SN −1 C,∗ obtained via relations of type ab∗c = ±cb∗a, and SN −1 C,# obtained via relations of type ab∗ = ±ba∗, a∗b = ±b∗a: / SN −1 C,∗ / SN −1 C,+ ¯SN −1 C,∗ ¯SN −1 C,∗ SN −1 C,∗ SN −1 C,# SN −1 C,∗ SN −1 C,+ ¯SN −1 C,∗ ¯SN −1 C,# III. The problem now is that, when adding these 4 new spheres, we will lose the fact that our set of spheres is stable by intersection. More precisely, in order for this to hold, C,◦ ∩ SN −1 we must add 4 more spheres, namely SN −1 . The diagram of inclusions between the 18 spheres is then as follows: C,# ∩ SN −1 C,∗∗ and SN −1 C,◦ = SN −1 C,− = SN −1 C SN −1 C,# SN −1 C,∗∗ $■■■■ :✉✉✉✉ :✉✉✉✉ $■■■■ :✉✉✉✉ SN −1 C,◦ SN −1 C SN −1 C,− :✉✉✉✉ $■■■■ ¯SN −1 C,# SN −1 C,∗ / SN −1 C,+ ¯SN −1 C,∗ ¯SN −1 C,◦ ¯SN −1 C,∗∗ ¯SN −1 C,− ¯SN −1 C SN −1 R / SN −1 R,∗ SN −1 R,+ ¯SN −1 R,∗ ¯SN −1 R By functoriality, the set of 18 spheres follows to be stable by free complexification. Regarding the extension of our various results, the first, and main problem, concerns the axiomatization. The complexification formula zi = uxi suggests to use diagrams with each leg labelled either ◦× or ו, with the simplification rules ◦• → ∅ and •◦ → ∅. We / / o o o o / / O O / / O O O O O O o o O O o o / / o o o o / / O O / / O O O O O O o o O O o o $ z z $ : / o o d d z z : $ : d d d d z z : d d O O / / / O O O O o o O O o o O O 36 TEODOR BANICA believe that an axiomatization is possible along these lines, and that this should lead to an extension of the other results as well, but we do not have any precise result here. Regarding some further extensions of our 10 + 8 formalism, interesting here, as a tech- nical ingredient, would be to have classification results for the easy quantum groups UN ⊂ G ⊂ U + In principle, the needed ingredients for dealing with such questions are available from [8], [30], [37]. In practice, however, it is not clear what the “19-th sphere” should be. N , or more generally for the easy quantum groups ON ⊂ G ⊂ U + N . As a general conclusion, in the undeformed world we have 10 + 8 main geometries. For the simplest such geometry, the one of RN , the group ON appears twice, first as a quantum isometry group, ON = G+(SN −1 . The situation is similar in the complex case, and for the remaining 8 + 8 geometries as well. With this perspective in mind, several results concerning the subgroups G ⊂ ON , taken either as groups, or as manifolds, should have extensions to other geometries. ), and second as a manifold, ON ⊂ SN 2−1 R R This adds to the various questions raised throughout the paper. References [1] T. Banica, A note on free quantum groups, Ann. Math. Blaise Pascal 15 (2008), 135–146. [2] T. Banica, J. Bichon and B. Collins, The hyperoctahedral quantum group, J. Ramanujan Math. Soc. 22 (2007), 345–384. [3] T. Banica, J. Bichon and S. Curran, Quantum automorphisms of twisted group algebras and free hypergeometric laws, Proc. Amer. Math. Soc. 139 (2011), 3961–3971. [4] T. Banica and B. Collins, Integration over compact quantum groups, Publ. Res. Inst. Math. Sci. 43 (2007), 277–302. [5] T. Banica and D. Goswami, Quantum isometries and noncommutative spheres, Comm. Math. Phys. 298 (2010), 343–356. [6] T. Banica and A. Skalski, Quantum symmetry groups of C∗-algebras equipped with orthogonal filtrations, Proc. Lond. Math. Soc. 106 (2013), 980–1004. [7] T. Banica, A. Skalski and P.M. So ltan, Noncommutative homogeneous spaces: the matrix case, J. Geom. Phys. 62 (2012), 1451–1466. [8] T. Banica and R. Speicher, Liberation of orthogonal Lie groups, Adv. Math. 222 (2009), 1461–1501. [9] T. Banica and R. Vergnioux, Invariants of the half-liberated orthogonal group, Ann. Inst. Fourier 60 (2010), 2137–2164. [10] H. Bercovici and V. Pata, Stable laws and domains of attraction in free probability theory, Ann. of Math. 149 (1999), 1023–1060. [11] J. Bhowmick, F. D’Andrea and L. Dabrowski, Quantum isometries of the finite noncommutative geometry of the standard model, Comm. Math. Phys. 307 (2011), 101–131. [12] J. Bhowmick, F. D’Andrea, B. Das and L. Dabrowski, Quantum gauge symmetries in noncommuta- tive geometry, J. Noncommut. Geom. 8 (2014), 433–471. [13] J. Bhowmick and D. Goswami, Quantum isometry groups: examples and computations, Comm. Math. Phys. 285 (2009), 421–444. [14] J. Bhowmick and D. Goswami, Quantum group of orientation preserving Riemannian isometries, J. Funct. Anal. 257 (2009), 2530–2572. [15] J. Bhowmick and D. Goswami, Quantum isometry groups of the Podle´s spheres, J. Funct. Anal. 258 (2010), 2937–2960. LIBERATIONS AND TWISTS OF REAL AND COMPLEX SPHERES 37 [16] J. Bhowmick, C. Voigt and J. Zacharias, Compact quantum metric spaces from quantum groups of rapid decay, arxiv:1406.0771. [17] J. Bichon and M. Dubois-Violette, Half-commutative orthogonal Hopf algebras, Pacific J. Math. 263 (2013), 13–28. [18] A.H. Chamseddine and A. Connes, The spectral action principle, Comm. Math. Phys. 186 (1997), 731–750. [19] A.H. Chamseddine and A. Connes, Why the standard model, J. Geom. Phys. 58 (2008), 38–47. [20] F. Cipriani, U. Franz and A. Kula, Symmetries of L´evy processes on compact quantum groups, their Markov semigroups and potential theory, J. Funct. Anal. 266 (2014), 2789–2844. [21] A. Connes, Noncommutative geometry, Academic Press (1994). [22] A. Connes, Gravity coupled with matter and foundation of noncommutative geometry, Comm. Math. Phys. 182 (1996), 155–176. [23] A. Connes, On the spectral characterization of manifolds, J. Noncommut. Geom. 7 (2013), 1–82. [24] A. Connes and M. Dubois-Violette, Moduli space and structure of noncommutative 3-spheres, Lett. Math. Phys. 66 (2003), 91–121. [25] A. Connes and G. Landi, Noncommutative manifolds, the instanton algebra and isospectral defor- mations, Comm. Math. Phys. 221 (2001), 141–160. [26] F. D’Andrea, L. Dabrowski and G. Landi, The noncommutative geometry of the quantum projective plane, Rev. Math. Phys. 20 (2008), 979–1006. [27] L. Dabrowski, F. D’Andrea, G. Landi and E. Wagner, Dirac operators on all Podle´s quantum spheres, J. Noncommut. Geom. 1 (2007), 213–239. [28] B. Das and D. Goswami, Quantum Brownian motion on noncommutative manifolds: construction, deformation and exit times, Comm. Math. Phys. 309 (2012), 193–228. [29] B. Das, D. Goswami and S. Joardar, Rigidity of action of compact quantum groups on compact, connected manifolds, arxiv:1309.1294. [30] A. Freslon, On the partition approach to Schur-Weyl duality and free quantum groups, arxiv: 1409.1346. [31] D. Goswami, Quantum group of isometries in classical and noncommutative geometry, Comm. Math. Phys. 285 (2009), 141–160. [32] D. Goswami, Existence and examples of quantum isometry groups for a class of compact metric spaces, arxiv:1205.6099. [33] S. Neshveyev and L. Tuset, Compact quantum groups and their representation categories, SMF (2013). [34] A. Nica and R. Speicher, Lectures on the combinatorics of free probability, Cambridge Univ. Press (2006). [35] P. Podle´s, Quantum spheres, Lett. Math. Phys. 14 (1987), 193–202. [36] S. Raum, Isomorphisms and fusion rules of orthogonal free quantum groups and their complexifica- tions, Proc. Amer. Math. Soc. 140 (2012), 3207–3218. [37] S. Raum and M. Weber, The full classification of orthogonal easy quantum groups, arxiv: 1312.3857. [38] M. Thibault de Chanvalon, Quantum symmetry groups of Hilbert modules equipped with orthogonal filtrations, J. Funct. Anal. 266 (2014), 3208–3235. [39] J.C. Varilly, Quantum symmetry groups of noncommutative spheres, Comm. Math. Phys. 221 (2001), 511–524. [40] D.V. Voiculescu, K.J. Dykema and A. Nica, Free random variables, AMS (1992). [41] S. Wang, Free products of compact quantum groups, Comm. Math. Phys. 167 (1995), 671–692. [42] S. Wang, Quantum symmetry groups of finite spaces, Comm. Math. Phys. 195 (1998), 195–211. 38 TEODOR BANICA [43] S.L. Woronowicz, Compact matrix pseudogroups, Comm. Math. Phys. 111 (1987), 613–665. [44] S.L. Woronowicz, Tannaka-Krein duality for compact matrix pseudogroups. Twisted SU(N) groups, Invent. Math. 93 (1988), 35–76. T.B.: Department of Mathematics, Cergy-Pontoise University, 95000 Cergy-Pontoise, France. [email protected]
1004.0614
1
1004
2010-04-05T11:31:12
Subalgebras of $C(\Omega,M_n)$ and their modules
[ "math.OA" ]
We give an operator space characterization of subalgebras of $C(\Omega,M_n)$. We also describe injective subspaces of $C(\Omega,M_n)$ and then give applications to sub-TROs of $C(\Omega,M_n)$. Finally, we prove an `$n$-minimal version' of the Christensen-Effros-Sinclair representation theorem.
math.OA
math
SUBALGEBRAS OF C(Ω, Mn) AND THEIR MODULES JEAN ROYDOR Abstract. We give an operator space characterization of sub- algebras of C(Ω, Mn). We also describe injective subspaces of C(Ω, Mn) and then give applications to sub-TROs of C(Ω, Mn). Finally, we prove an 'n-minimal version' of the Christensen-Effros- Sinclair representation theorem. 1. Introduction and preliminaries Let n ∈ N∗. An operator space X is called n-minimal if there exists a compact Hausdorf space Ω and a completely isometric map i : X → C(Ω, Mn). The readers are referred to [13] and [7] for details on operator space theory. Recall that the C∗-algebra C(Ω, Mn) can be identified ∗-isomorphically with C(Ω) ⊗min Mn or Mn(C(Ω)) (see [12, Proposition 12.5] for details). Obviously, in the case n = 1, we just deal with the well-known class of minimal operator spaces. Smith noticed that any linear map into Mn is completely bounded and its cb norm is achieved at the nth amplification i.e. (cid:107)u(cid:107)cb = (cid:107)idMn ⊗ u(cid:107) (see [12, Proposition 8.11]). Clearly, this property remains true for maps into C(Ω, Mn). In fact, Pisier showed that this property characterized n-minimal operator spaces. More precisely, if X is an operator space such that any linear map u into X is necessarily completely bounded and (cid:107)u(cid:107)cb = (cid:107)idMn ⊗ u(cid:107), then X is n-minimal (see [14, Theorem 18]). We now recall a few facts about injectivity (see [7], [12] or [2] for details). A Banach space X is injective if for any Banach spaces Y ⊂ Z, each contractive map u : Y → X has a contractive exten- sion u : Z → X. Since the 50's, it is known that a Banach space is injective if and only if it is isometric to a C(K)-space with K a Stonean space and dual injective Banach spaces are exactly L∞-spaces (see [6] for more details). More recently, injectivity has also been stud- ied in operator spaces category. Analogously, an operator space X is said to be injective if for any operator spaces Y ⊂ Z, each completely contractive map u : Y → X has a completely contractive extension u : Z → X. Note that a Banach space is injective if and only if it is 1991 Mathematics Subject Classification. 47L30,47L25. 1 2 JEAN ROYDOR injective as a minimal operator space. Let X be an operator space, (Y, i) is an injective envelope of X if Y is an injective operator space, i : X → Y is a complete isometry and for any injective operator space Z with i(X) ⊂ Z ⊂ Y , then Z = Y . Sometimes, we may forget the completely isometric embedding. In fact, any operator space admits a unique injective envelope (up to complete isometry) and we write I(X) the injective envelope of X. See [7, Chapter 6] for a proof of this construction. Obviously, an (cid:96)∞-direct sum of n-minimal operator spaces is again n-minimal. In the next proposition, we give some other easy properties of n-minimal operator spaces : Proposition 1.1. Let X be an n-minimal operator space. i) Then its bidual X∗∗ and its injective envelope I(X) are n-minimal ii) If moreover, X is a dual operator space, then there is a set I and too. a w∗-continuous complete isometry i : X → (cid:96)∞ I (Mn). Proof. The first assertion of i) follows from C(Ω, Mn)∗∗ = Mn(C(Ω))∗∗ = Mn(C(Ω)∗∗) ∗-isomorphically. For the second, suppose X ⊂ C(Ω, Mn) completely isometrically. From the description of injective Banach spaces, I(C(Ω)) = C(Ω(cid:48)) with Ω(cid:48) Stonean. Then X ⊂ C(Ω(cid:48), Mn) and this last C∗-algebra is injective, so I(X) ⊂ C(Ω(cid:48), Mn) completely isometrically. Suppose that W is an operator space predual of X. Then X = CB(W, C) and if I = ∪nBall(Mn(W )), we have a w∗-continuous com- plete isometry ψ : X −→ ⊕∞ w∈IMnw (where nw = m if w ∈ Mm(W )) defined by ψ(x) = ([x(wij)])w∈I. Let x ∈ Mk(X) = CB(W, Mk). As X is n-minimal, by [12, Proposition 8.11], (cid:107)x∗(cid:107)cb = (cid:107)idMn ⊗ x∗(cid:107), k → X denotes the adjoint map. However, for any l, where x∗ : M∗ (cid:107)idMl ⊗ x(cid:107) = (cid:107)idMl ⊗ x∗(cid:107). Hence, (cid:107)x(cid:107)cb = (cid:107)idMn ⊗ x(cid:107) and so, in the definition of ψ, we can majorize the nw's by n and obtain a complete isometry. We reviewed that an injective minimal operator space is a C∗-algebra, but this property is lost for n-minimal operator spaces (as soon as n ≥ 2). Generally, an injective operator space only admits a structure of ternary ring of operators. We recall that a closed subspace X of a C∗-algebra is a ternary ring of operators (TRO in short) if XX (cid:63)X ⊂ X, here X (cid:63) denotes the adjoint space of X. And a W ∗-TRO is w∗-closed subspace of a von Neumann algebra stable under the preceding 'triple product'. TROs and W ∗-TROs can be regarded as generalization of C∗-algebras and W ∗-algebras. For instance, The Kaplansky density SUBALGEBRAS OF C(Ω, Mn) AND THEIR MODULES 3 Theorem and the Sakai Theorem remain valid for TROs (see e.g. [6]). A triple morphism between TROs is a linear map which preserves their 'triple products'. This category enjoys some 'rigidity properties' like C∗-algebras category (see e.g. [6] or [2, Section 8.3] for details). So far we have seen that certain properties of the minimal case 'pass' to the n-minimal situation. Therefore, the basic idea of this paper is to extend valid results in the commutative case to the more general n-minimal case. A first commutative result that can be extended to the n-minimal case is a theorem on operator algebras due to Blecher. We recall that an operator algebra is a closed subalgebra of B(H), see [2] or [12] for some backgrounds and developments. And an operator algebra is said to be approximately unital if it possesses a contractive approximate identity. In [1], Blecher showed that an approximately unital operator algebra which is minimal is in fact a uniform algebra (i.e a subalgebra of a commutative C∗-algebra). So here, let A be an approximately unital operator algebra and assume that A is n-minimal. Then we can obtain a completely isometric homomorphism from A into a certain C(Ω, Mn) (see Corollary 2.3). Of course, we can ask this type of question in various categories of operator spaces. More precisely, let C denote a certain subcategory of the category of operator spaces with completely contractive maps. Let X be an object of C which is n-minimal (as an operator space), can we obtain a completely isometric morphism of C from X into a C∗-algebra of the form C(Ω, Mn) ? For example in Proposition 1.1, we answered this question in the category of dual operator spaces and w∗-continuous completely contractive maps. We will also give a positive answer in the category of : - C∗-algebras and ∗-homomorphisms (see Theorem 2.2) ; - von Neumann algebras and w∗-continuous ∗-homomorphisms (see Re- mark 2.4) ; - approximately unital operator algebras and completely contractive homomorphisms (see Corollary 2.3) ; - operator systems and completely positive unital maps (see Corollary 3.3) ; - TRO and triple morphisms (see Proposition 4.1) ; - W ∗-TRO and w∗-continuous triple morphisms (see Corollary 4.5). It means that, in any of the previous categories, the n-minimal opera- tor space structure encodes the additional structure. Since the injective envelope of an n-minimal operator space is n-minimal too (see Propo- sition 1.1), passing to the injective envelope will be a useful technique to answer these preceding questions. In any case, the description of 4 JEAN ROYDOR n-minimal injective operator spaces (established in Theorem 3.5) will be of major importance. The Christensen-Effros-Sinclair theorem (CES-theorem in short) is a second example of theorem that could be treated in the n-minimal case. Let A be an operator algebra (or more generally a Banach algebra endowed with an operator space structure) and let X be an operator space which is a left A-module. Then following [2, Chapter 3], we say that X is a left h-module over A if the action of A on X induces a completely contractive map from A ⊗h X in X (where ⊗h denotes the Haagerup tensor product). The CES-theorem states that if X is a non-degenerate h-module over an approximately unital operator al- gebra A (i.e. AX is dense in X), then there exists a C∗-algebra C, a complete isometry i : X → C and a completely contractive homomor- phism π : A → C such that i(a · x) = π(a)i(x) for any a ∈ A, x ∈ X. We will prove that if X is n-minimal, we can choose C to be n-minimal too. This leads to an 'n-minimal version' of the CES-theorem. The case n = 1 has been treated (see [3]) in a Banach space framework ; here we will use an operator space approach based on the multiplier algebra of an operator space. 2. Subalgebras of C(Ω, Mn) Recall that a C∗-algebra is subhomogeneous of degree ≤ n if it is con- tained ∗-isomorphically in a C∗-algebra of the form C(Ω, Mn), where Ω is compact Hausdorf space. Hence n-minimality could be seen as an operator space analog of subhomogeneity of degree ≤ n. We also recall the well-known characterization of subhomogeneous C∗-algebras Indeed, a C∗-algebra A is subhomoge- in terms of representations. neous of degree ≤ n if and only if every irreducible representation of A has dimension no greater than n. The 'if part' is easily obtained taking a separating family of irreducible representations. Conversely, if A is contained ∗-isomorphically in C(Ω, Mn), then every irreducible representation of A extends to one on C(Ω, Mn) (because irreducible representations correspond to pure states). And as any irreducible representation of C(Ω, Mn) has dimension no greater than n, we can conclude (the author thanks Roger Smith for these explanations). Lemma 2.1. Let k ∈ N∗, Ω a compact Hausdorf space and tk the transpose mapping tk : C(Ω, Mk) → C(Ω, Mk), [fij] (cid:55)→ [fji] SUBALGEBRAS OF C(Ω, Mn) AND THEIR MODULES 5 Then for any l ∈ N∗, (cid:107)idMl ⊗ tk(cid:107) = inf (k, l). Thus tk is completely bounded and (cid:107)idMk ⊗ tk(cid:107) = (cid:107)tk(cid:107)cb = k. Proof. The equality (cid:107)tk(cid:107)cb = k is obtained in adapting the proof of [7, Proposition 2.2.7]. Hence in the case k ≤ l, by [12, Proposition 8.11]) we obtain (cid:107)idMl ⊗ tk(cid:107) = inf (k, l). Next we prove (cid:107)idMl ⊗ tk(cid:107) ≤ l. let π be the cyclical permutation matrix  ∈ Ml(C(Ω, Mk)).  0 0 Ik 0 ... 0 0 π = ··· ··· . . . ··· 0 0 Ik Ik 0 ... 0 Let Dl : Ml(C(Ω, Mk)) → Ml(C(Ω, Mk)) be the diagonal truncation of Ml i.e. Dl(ij ⊗ y) = δijij ⊗ y where ij (i, j ≤ l) denotes the matrix units of Ml and y ∈ C(Ω, Mk). Let x = [xij]i,j≤l ∈ Ml(C(Ω, Mk)) and for simplicity of notation, we wrote t(x) = idMl⊗tk(x) ∈ Ml(C(Ω, Mk)). i=0 (cid:107)Dl(t(x)πi)(cid:107) (because π is unitary). To conclude it suffices to majorize each terms of the previous sum by the norm of x. However, for any i, Dl(t(x)πi) i=0 Dl(t(x)πi)π−i, and so (cid:107)t(x)(cid:107) ≤ (cid:80)l−1 jj⊗tk(xpj qj )(cid:107)2 = (cid:107) l(cid:88) j=1 jj ⊗ tk(xpj qj ) and we can majorize its norm, )(cid:107) = maxj{(cid:107)tk(xpj qj x∗ jj⊗tk(xpj qj x∗ Then t(x) = (cid:80)l−1 is of the form(cid:80)l (cid:107) l(cid:88) pj qj )(cid:107)} pj qj j=1 j=1 pj qj but xpj qj x∗ is a selfadjoint element of C(Ω, Mk), so its norm is un- )(cid:107) = (cid:107)xpj qj(cid:107)2 ≤ (cid:107)x(cid:107)2. Finally, for any changed by tk and (cid:107)tk(xpj qj x∗ i, (cid:107)Dl(t(x)πi)(cid:107) ≤ (cid:107)x(cid:107) which enable us to conclude. Moreover in adapting [7, Proposition 2.2.7], we have easily (cid:107)idMl ⊗ tk(cid:107) = l, if l ≤ k. pj qj In the next theorem, we denote by Aop the opposite structure of a C∗-algebra A (see e.g. [13, Paragraph 2.10] or [2, Paragraph 1.2.25] for details). More generally, if X is an operator space, X op is the same vector space but with the new matrix norms defined by (cid:107)[xij](cid:107)Mn(X op) = (cid:107)[xji](cid:107)Mn(X) for any [xij] ∈ Mn(X). Hence the assumption (iii) in the next theorem is equivalent to (cid:107)idA ⊗ tk(cid:107) ≤ n for any k ∈ N∗, where tk denotes the transpose mapping from Mk to Mk discussed above. Theorem 2.2. Let A be a C∗-algebra. Then the following are equiva- lent : 6 JEAN ROYDOR (i) A is subhomogeneous of degree ≤ n. (ii) A is n-minimal. (iii) (cid:107)id : A → Aop(cid:107)cb ≤ n. Proof. (i) ⇒ (ii) is obvious and (ii) ⇒ (iii) follows from the first equality in the previous lemma. Suppose (iii). Let π : A → B(H) be an irreducible representation and k ∈ N∗ such that Mk ⊂ B(H) ; from the first paragraph of this section, we must prove that k ≤ n. Using the previous lemma (with a singleton as Ω), there is x ∈ Mk(Mk) ⊂ Mk(B(H)) satisfying k = (cid:107)idMk ⊗ tk(x)(cid:107) and (cid:107)x(cid:107) ≤ 1. The representation πk = idMk ⊗ π is also irreducible so the commutant πk(Mk(A))(cid:48) = CIH k, thus by the von Neumann's double commutant theorem Then by the Kaplansky density theorem, there exists a net (xλ)λ∈Λ ⊂ Mk(π(A)) converging to x in the σ-strong operator topology and such that (cid:107)xλ(cid:107) ≤ 1. Therefore idB(H) ⊗ tk(xλ) tends to idMk ⊗ tk(x) in the w∗-topology and by the semicontinuity of the norm in the w∗-topology, we have = Mk(B(H)). Mk(π(A)) so k = (cid:107)idMk ⊗ tk(x)(cid:107) ≤ lim sup (cid:107)idB(H) ⊗ tk(xλ)(cid:107) Let  > 0. For any λ, there exists yλ ∈ Mk(A) such that xλ = πk(yλ) and (cid:107)yλ(cid:107) ≤ 1 + . By assumption, λ (cid:107)idA ⊗ tk(cid:107) ≤ n Moreover (idB(H)⊗tk)◦πk = πk◦(idA⊗tk). Combining these arguments we finally obtain k = (cid:107)idMk ⊗ tk(x)(cid:107) ≤ lim supλ (cid:107)idB(H) ⊗ tk(πk(yλ))(cid:107) ≤ lim supλ (cid:107)πk(idA ⊗ tk(yλ))(cid:107) ≤ (cid:107)idA ⊗ tk(cid:107)(1 + ) ≤ n(1 + ). Hence k ≤ n. Now we extend (i) ⇔ (ii) of the previous theorem, which concerns C∗-algebras, to the larger category of operator algebras and completely contractive homomorphisms. Corollary 2.3. Let A be an approximately unital operator algebra. Then the following are equivalent : (i) There exists a compact Hausdorf space Ω and a completely iso- metric homomorphism π : A → C(Ω, Mn). SUBALGEBRAS OF C(Ω, Mn) AND THEIR MODULES 7 (ii) A is n-minimal. Proof. (i) ⇒ (ii) is obvious. Suppose (ii). We know that the injec- tive envelope I(A) is a C∗-algebra and there is a completely isometric homomorphism from A into I(A) (see [2, Corollary 4.2.8]). Since A is n-minimal, I(A) is n-minimal too, by Proposition 1.1. Applying Theorem 2.2 to I(A), we can conclude. Remark 2.4. Using the well-known description of subhomogeneous W ∗-algebras, we easily obtained that, if M is a W ∗-algebra and M is n-minimal, then via a normal ∗-isomorphism. Here Ωi is a measure space and ni ≤ n, for any i ∈ I. This result will be extended to the category of W ∗-TROs (see Corollary 4.5). i∈IL∞(Ωi, Mni) M = ⊕∞ 3. Injective n-minimal operator spaces Before describing injective n-minimal operator spaces, we can treat the more 'rigid' case of injective n-minimal C∗-algebras as an easy consequence of [16]. Proposition 3.1. Let A be an n-minimal C∗-algebra. Then the fol- lowing are equivalent : (i) A is injective. (ii) There exists a finite family of Stonean compact Hausdorf spaces i∈IC(Ωi, Mni) ∗-isomorphically with ni ≤ (Ωi)i∈I such that A = ⊕∞ n, for any i ∈ I. Proof. As A is injective, A is monotone complete (see [7, Theorem 6.1.3]). Thus A is an AW ∗-algebra. Moreover, by [16, Proposition 6.6], A either contains M∞ = ⊕∞ k Mk or A is of the desired form. The first alternative is impossible because A is n-minimal, which ends the 'only if' part. The converse is clear, since each Ωi is Stonean. Remark 3.2. This theorem enables us to give a short proof of (ii) ⇒ (i) in Theorem 2.2. If A is an n-minimal C∗-algebra, its injective enve- lope I(A) is n-minimal too (by Proposition 1.1). I(A) is a C∗-algebra and contains A ∗-isomorphically (see [7, Theorem 6.2.4]). Applying the previous proposition to I(A), we obtain that with ni ≤ n, for any i ∈ I. And now it is not difficult to construct a ∗-isomorphism from A into C(Ω, Mn) where Ω denotes the (finite) disjoint union of the Ωi's. I(A) = ⊕∞ i∈IC(Ωi, Mmi) ∗-isomorphically 8 JEAN ROYDOR We recall that an operator space X is unital if there exists e ∈ X and a complete isometry from X into a certain B(H) which sends e on IH. From the result below, an n-minimal operator system can embed into a C∗-algebra of the form C(Ω, Mn) via a unital complete order isomorphism. Corollary 3.3. Let X be a unital operator space. Then the following are equivalent : (i) There exists a compact Hausdorf space Ω and a completely iso- metric unital map π : X → C(Ω, Mn). (ii) X is n-minimal. Proof. (i) ⇒ (ii) is obvious. Suppose (ii). We know that the injective envelope I(X) is a C∗-algebra and there is a unital complete isometry from X into I(X) (see [2, Corollary 4.2.8]). As X is n-minimal, I(X) is n-minimal too (by Proposition 1.1). By the previous theorem I(X) = ⊕∞ i∈IC(Ωi, Mni) ∗-isomorphically. Next we show that for any i there exists a unital complete isometry ϕi : Mni → Mn. By iteration, we only need to prove that for any k ∈ N∗, there exists a unital complete isometry from Mk into Mk+1. The map ik : Mk → Mk+1 (cid:55)→ x ⊕ trk(x) x (where trk denotes the normalized trace on Mk) is a unital complete order isomorphism and thus a unital complete isometry. We can define a unital complete isometry ψ : ⊕∞ i∈IC(Ωi, Mni) → C(Ω, Mn) (fi ⊗ xi)i fi ⊗ ϕi(xi) (cid:55)→ (cid:80) i where Ω denotes the disjoint union of Ωi's and fi the continuous ex- tension by 0 of fi on Ω. Finally, we have X ⊂ I(X) ⊂ C(Ω, Mn) via unital complete isometries. Remark 3.4. This last corollary cannot be extended to the category of operator algebras and completely contractive homomorphisms. In fact, if π : Mp → C(Ω, Mq) is a unital completely contractive homomorphism then π is positive so it is a ∗-homomorphism. Therefore (composing by an evaluation) we can obtain a unital ∗-homomorphism from Mp in Mq and thus p divides q (see [12, Exercise 4.11]). SUBALGEBRAS OF C(Ω, Mn) AND THEIR MODULES 9 We must recall a crucial construction of the injective envelope of an operator space X which will be useful in this paper (see [2, Paragraph 4.4.2] for more details on this construction). Assume that X ⊂ B(H), we can consider its Paulsen system (cid:19) (cid:18) C X X (cid:63) C S(X) = ⊂ M2(B(H)) where X (cid:63) denotes the adjoint space of X. The injective envelope of S(X) is the range of a completely contractive projection ϕ : M2(B(H)) → M2(B(H)) which leaves S(X) invariant. By [7, Theorem 6.1.3], I(S(X)) admits a C∗-algebraic structure but it is not necessarily a sub-C∗- algebra of M2(B(H)). However (cid:19) (cid:18) 1 0 0 0 p = and q = (cid:19) (cid:18) 0 0 0 1 = 1 − p (which are invariant by ϕ) are still orthogonal projections (i.e. selfad- joint idempotents) of the new C∗-algebra I(S(X)). Since they satisfy p + q = 1 and pq = 0, we can decompose I(S(X)) in 2 × 2 matrices, as follow : (cid:18) I11(X) I12(X) (cid:19) I(S(X)) = I21(X) I22(X) where I11(X) = pI(S(X))p and I22(X) = qI(S(X))q are injective C∗- algebras, I12(X) = pI(S(X))q is in fact the injective envelope of X and I21(X) = qI(S(X))p coincides with I12(X)(cid:63). Therefore, we obtain the Hamana-Ruan Theorem i.e. an injective operator space is an 'off- diagonal' corner of an injective C∗-algebra (see [7, Theorem 6.1.6]). It links the study of injective operator spaces to injective C∗-algebras (and, by the way, it proves that an injective operator space is a TRO). Theorem 3.5. Let X be an n-minimal operator space. Then the fol- lowing are equivalent : (i) X is injective. (ii) There exists a finite family of Stonean compact Hausdorf spaces i∈IC(Ωi, Mri,ki) completely isometrically (Ωi)i∈I such that X = ⊕∞ with ri, ki ≤ n, for any i ∈ I. Proof. (ii) ⇒ (i) is obvious. Let X be an injective n-minimal operator space. By the discussion above, we know that there exists an injective C∗-algebra A and a projection p ∈ A such that X = pA(1 − p) completely isometrically In fact A is the injective envelope of S(X) the Paulsen system of X (see above). As X is n-minimal, S(X) is 2n-minimal, so is A (by 10 JEAN ROYDOR Proposition 1.1). From Proposition 3.1, A = ⊕∞ i∈IC(Ωi, Mmi) ∗-isomorphically where mi ≤ 2n. For simplicity of notation, we will assume momentarily that the cardinal of I is equal to 1 and so X = pC(Ω, Mm)(1 − p) completely isometrically, for some projection p ∈ C(Ω, Mm). Using [5, Corollary 3.3] or [8, Theorem 3.2], there is a unitary u of C(Ω, Mm) such that for any ω ∈ Ω, upu∗(ω) is of the form diag(1, . . . , 1, 0, . . . , 0). So we may assume that for any ω ∈ Ω, p(ω) is a diagonal matrix of the form given above. For any k ≤ m, we define Ωk = {ω ∈ Ω : rg(p(ω)) = k} 1≤k≤m−1C(Ωk, Mk,m−k). k≤mC(Ωk, Mk,m−k) = ⊕∞ which is a closed subset of Ω (because the rank and the trace of a projection coincide) and the family (Ωk)k≤m forms a partition of Ω. Hence, any Ωk is open (and closed) in Ω, so Ωk is still Stonean. We have the completely isometric identifications X = pC(Ω, Mm)(1−p) = ⊕∞ Moreover, for any 1 ≤ k ≤ m − 1, we have the completely isometric embeddings and as X is n-minimal, it forces k ≤ n and m − k ≤ n ; if not, at least the row Hilbert space Rn+1 or the column Hilbert space Cn+1 would be n-minimal. Thus X has the announced form. In general, I is a finite set and Mk,m−k ⊂ C(Ωk, Mk,m−k) ⊂ X X = p ⊕∞ i∈I C(Ωi, Mmi)(1 − p) = ⊕∞ i∈IpiC(Ωi, Mmi)(1 − pi) where pi is a projection in C(Ωi, Mmi) and p = ⊕ipi. Applying the preceding argument to each terms piC(Ωi, Mmi)(1 − pi), we can con- clude. Corollary 3.6. Let X be an n-minimal dual operator space. Then the following are equivalent : (i) X is injective. (ii) There exists a finite family of measure spaces (Ωi)i∈I such that i∈IL∞(Ωi, Mri,ki) via a completely isometric w∗-homeomorphism X = ⊕∞ with ri, ki ≤ n, for any i ∈ I. Proof. From the previous theorem, X = ⊕∞ i C(Ki, Mri,ki) completely isometrically, where Ki is Stonean. Since X is a dual operator space, it forces C(Ki) to be a dual commutative C∗-algebra i.e. C(Ki) = L∞(Ωi) (via a normal ∗-isomorphism) for some measure space Ωi. SUBALGEBRAS OF C(Ω, Mn) AND THEIR MODULES 11 4. Application to n-minimal TROs In this section, we will use the description of injective n-minimal operator spaces to obtain results on n-minimal TROs. First, we will see that the n-minimal operator structure of a TRO determines its whole triple structure. See e.g. [6] or [2, Section 8.3] for details on TROs. Proposition 4.1. Let X be a TRO. The following are equivalent : (i) There exists a compact Hausdorf space Ω and an injective triple morphism π : X → C(Ω, Mn). (ii) X is n-minimal. Proof. (i) ⇒ (ii) follows from the fact that an injective triple mor- phism is necessarily completely isometric (see e.g. [6, Proposition 2.2] or [2, Lemma 8.3.2]). Suppose (ii). By [2, Remark 4.4.5 (1)], the injective envelope of X admits a TRO structure and X can be viewed as a sub-TRO of I(X). From Theorem 3.5, we can describe this injective envelope as a direct sum, I(X) = ⊕∞ i∈IC(Ωi, Mri,ki) completely isometrically. But the right hand side of the equality admits a canonical TRO struc- ture and it is known (see e.g. [2, Corollary 4.4.6]) that a surjective complete isometry between TROs is automatically a triple morphism. In addition, for any i, the embedding ϕi : Mri,ki → Mn into the 'up- left' corner of Mn is an injective triple morphism. As in the end of the proof of Corollary 3.3, we finally obtain X ⊂ I(X) = ⊕∞ i∈IC(Ωi, Mri,ki) ⊂ C(Ω, Mn) as TROs. For details on C∗-modules theory, the readers are referred to [11] or [2, Chapter 8] for an operator space approach. We must recall the construction of the linking C∗-algebra of a C∗-module. If X is left C∗-module over a C∗-algebra A then its conjugate vector space X is a right C∗-module over A with the action x · a = a∗x and inner product (cid:104)x, y(cid:105) = (cid:104)x, y(cid:105), for any a ∈ A, x, y ∈ X. We denote by AK(X) the C∗-algebra of 'compact' adjointable maps of X and then (cid:18) A L(X) = (cid:19) X X AK(X) 12 JEAN ROYDOR is a C∗-algebra too which is called the linking C∗-algebra of X. If X is an equivalence bimodule (see [2, Paragraph 8.1.2]) over two C∗- algebras A and B, we define (cid:19) (cid:18) A1 X X B1 (cid:19) (cid:18) A X X B L(X) = and L1(X) = (where A1 and B1 denote the unitizations of A and B) which are also C∗-algebras (see [2, Paragraph 8.1.17] for details on linking C∗- algebra). We can notice that X is an 'off-diagonal' corner of a C∗- algebra i.e. X = pL1(X)(1 − p) for some projection p ∈ L1(X). Hence a C∗-module admits a TRO structure. The converse will be seen later on, which will make the correspondence between C∗-modules, equiva- lence bimodules and TROs (see [2, Paragraph 8.1.19, 8.3.1]). Thus the next corollary is a reformulation of the previous proposition in the C∗- modules language. However, this corollary on representation of module action can be compared with Theorem 5.4. Corollary 4.2. Let X be a full left C∗-module over a C∗-algebra A. Then the following are equivalent : (i) There exists a compact Hausdorf space Ω, a complete isometry i : X → C(Ω, Mn) and a ∗-isomorphism σ : A → C(Ω, Mn) such that for any a ∈ A, x, y ∈ X i(a · x) = σ(a)i(x) σ((cid:104)x, y(cid:105)) = i(x)i(y)∗ (ii) X is n-minimal and A is subhomogeneous of degree ≤ n. (iii) X is n-minimal. Proof. Only (iii) ⇒ (i) needs a proof. Since X is a C∗-module, it's also a TRO (see above). From Proposition 4.1, there exists a com- pact Hausdorf space Ω and an injective triple morphism i : X → C(Ω, Mn). By [2, Corollary 8.3.5], we can construct a corner pre- serving ∗-isomorphism π : L(X) → M2(C(Ω, Mn)) such that i = π12. Choosing σ = π11, we obtain the desired relations. An equivalence bimodule version of the previous corollary could be stated. In the previous result we transfer n-minimality from X to A. We can treat the 'reverse' question ; let X be an equivalence bimodule over two n-minimal C∗-algebras, we will prove that X is n-minimal. But first, let us translate this proposition in the TROs language. Let X be a TRO contained in a C∗-algebra B via an injective triple mor- phism. As in the notation of the second section of [15], we define C(X) (resp. D(X)) the norm closure of span{xy∗, x, y ∈ X} (resp. SUBALGEBRAS OF C(Ω, Mn) AND THEIR MODULES 13 span{x∗y, x, y ∈ X}). As X is a sub-TRO of B, C(X) and D(X) are sub-C∗-algebras of B and (cid:18) C(X) A(X) = X X (cid:63) D(X) is a sub-C∗-algebras of M2(B). Hence a TRO can be regarded as an 'off- diagonal' corner of a C∗-algebra which prove totally the correspondence between C∗-modules, equivalence bimodules and TROs. And A(X) is also called the linking C∗-algebra of X. Analogously, in W ∗-TROs category, let X be a W ∗-TRO contained in a W ∗-algebra B via a w∗- continuous injective triple morphism. We define M (X) (resp. N (X)) the w∗-closure of span{xy∗, x, y ∈ X} (resp. span{x∗y, x, y ∈ X}). As X is a sub-W ∗-TRO of B, M (X) and N (X) are sub-W ∗-algebras of B and (cid:18) M (X) R(X) = X X (cid:63) N (X) is a sub-W ∗-algebras of M2(B). It is called the linking von Neumann algebra of X. In fact, the linking algebras do not depend on the em- bedding of X into a C∗-algebra. Obviously, if X is an equivalence bimodule over two C∗-algebras A and B, C(X) and D(X) play the roles of A and B in the correspon- dence between equivalence bimodules and TROs. Hence in the TROs language, we obtain (in the dual case) : Proposition 4.3. Let X be a W ∗-TRO such that M (X) and N (X) are n-minimal von Neumann algebras. Then X is n-minimal and (cid:19) (cid:19) X = ⊕∞ i L∞(Ωi)⊗Mri,ki where Ωi is a measure space, ri, ki ≤ n, for any i. Proof. We write R(X) the linking von Neumann of X. From [9, The- orem 6.5.2], there exist p1, p2 and p3 three central projections of R(X) such that R(X) = p1R(X) ⊕∞ p2R(X) ⊕∞ p3R(X) and for i = 1, 2, 3, piR(X) is a von Neumann algebra of type i or pi = 0. Since M (X) is n-minimal, M (X) is of type I. However, M (X) = pR(X)p for some projection p in R(X) and for any i, piM (X) = ppipM (X)ppip As the type is unchanged by compression (see [9, Exercise 6.9.16]), piM (X) is of type I or piM (X) = 0. On the other hand, for any i, piM (X) = pipR(X) = ppiR(X)pip 14 JEAN ROYDOR so piM (X) has the same type as piR(X) or piM (X) = 0. Thus piM (X) = 0 for i = 2, 3 i.e. pip = 0 for i = 2, 3. Symmetrically, using our assumption on N (X), we have pi(1 − p) = 0 for i = 2, 3. Hence pi = 0 for i = 2, 3 i.e. R(X) is of type I. Using [15, Theorem 4.1], X = ⊕∞ k L∞(Ωk)⊗MIk,Jk where Ωk is a measure space, Ik, Jk are sets and MIk,Jk = B((cid:96)2 ). Ik Since M (X) (resp. N (X)) is n-minimal, it forces the cardinal of Ik (resp. Jk) to be no greater than n, for any k. So X is n-minimal and has the desired form. , (cid:96)2 Jk Remark 4.4. In the next two results, we will use that the multiplier algebra of an n-minimal C∗-algebra is n-minimal too. It is due to Proposition 1.1. The next corollary on W ∗-TROs extends Remark 2.4. Corollary 4.5. Let X be a W ∗-TRO. The following are equivalent : (i) X is n-minimal. (ii) There exists a measure space Ω and a w∗-continuous injective (iii) There exists a finite family of measure spaces (Ωi)i∈I such that triple morphism π : X → L∞(Ω, Mn). X = ⊕∞ i∈IL∞(Ωi, Mri,ki) with ri, ki ≤ n, for any i ∈ I. Proof. Only (i) ⇒ (iii) needs a proof. Suppose (i). From Proposition 4.1, we can see X as a sub-TRO of C(Ω, Mn), hence by construction C(X) and D(X) are n-minimal C∗-algebras. By [10], M (X) (resp. N (X)) is the multiplier algebra of C(X) (resp. D(X)), so M (X) and N (X) are n-minimal W ∗-algebras (by Remark 4.4). The result follows from the previous proposition. Finally, we can generalize (ii) ⇔ (iv) ⇔ (v) of [2, Proposition 8.6.5] on minimal TROs to the n-minimal case. Theorem 4.6. Let X be a TRO, the following are equivalent : (i) X is n-minimal. (ii) X∗∗ is an injective n-minimal operator space (see Corollary 3.6). (iii) C(X) and D(X) are n-minimal C∗-algebras. Proof. (ii) ⇒ (i) and (i) ⇒ (iii) are obvious. Suppose (iii). From [10, Proposition 2.4], we know that the multiplier algebra of C(X∗∗) is C(X)∗∗ and this C∗-algebra is n-minimal by our assumption on C(X) and Remark 4.4. Moreover by [15], M (X∗∗) is also the multiplier algebra of C(X∗∗), so M (X∗∗) is n-minimal too. The same argument works for N (X∗∗) and we can apply Proposition 4.3 to X∗∗. SUBALGEBRAS OF C(Ω, Mn) AND THEIR MODULES 15 5. An n-minimal version of the CES-theorem To prove the 'n-minimal' version the CES-Theorem we need the no- tion of left multiplier algebra of an operator space X. A left multiplier of an operator space X is a map u : X → X such that there exist a C∗-algebra A containing X via a complete isometry i and a ∈ A satis- fying i(u(x)) = ai(x) for any x ∈ X. Let Ml(X) denote the set of left multipliers of X. And the multiplier norm of u is the infimum of (cid:107)a(cid:107) over all possible A, i, a as above. In fact Blecher-Paulsen proved that any left multiplier can be represented in the embedding of X into the C∗-algebra (discussed in section 3) (cid:18) I11(X) I(X)(cid:63) (cid:19) I(X) I22(X) I(S(X)) = More precisely, for any left multiplier u of norm no greater than 1, there exists a unique a ∈ I11(X) of norm no greater than 1 such that u(x) = ax for any x ∈ X (see [2, Theorem 4.5.2]). This result enables us to consider Ml(X) as an operator subalgebra of I11(X) (see the proof of [2, Proposition 4.5.5] and [2, Paragraph 4.5.3] for more details) and Ml(X) = {a ∈ I11(X), aX ⊂ X} as operator algebras. The product used in the preceding centered for- mula is the one on the C∗-algebra I(S(X)). And the operator alge- bra Ml(X) is called the multiplier algebra of X. We let Al(X) = ∆(Ml(X)) denote the diagonal (see [2, Paragraph 2.1.2]) of Ml(X), this C∗-algebra is called the left adjointable multiplier algebra of X and Al(X) = {a ∈ I11(X), aX ⊂ X and a∗X ⊂ X} ∗-isomorphically. In fact, if X happens to be originally a C∗-algebra, Al(X) is just its multiplier algebra, and we recover Remark 4.4. Symmetrically, the right multiplier algebra of X is given by Mr(X) = {b ∈ I22, Xb ⊂ X} and its diagonal Ar(X) = {b ∈ I22, Xb ⊂ X and Xb∗ ⊂ X} is the right adjointable multiplier algebra of X. Lemma 5.1. Let X be an operator space and I(X) its injective enve- lope. Then there exists a completely contractive unital homomorphism θ : Ml(X) → Ml(I(X)) such that θ(u)X = u, for any u ∈ Ml(X). And thus, θAl(X) : Al(X) → Al(I(X)) is a ∗-isomorphism. Moreover, the same results hold for right multipliers. Proof. Let u ∈ Ml(X), then u can be represented by an element a in {a ∈ I11(X), aX ⊂ X}. And using the multiplication inside I(S(X)), 16 JEAN ROYDOR aI(X) ⊂ I(X), so a can be seen as an element of Ml(I(X)) which will be written θ(u). Therefore, θ is an injective unital completely contractive homomorphism. The rest of the proof follows from [2, Paragraph 2.1.2]. In the next lemma, we use the C∗-envelope of a unital operator space, see [2, Theorem 4.3.1] for details. And we write Rn (resp. Cn) the row (resp. column) Hilbert space of dimension n. If X is an operator space, we let Cn(X) be the minimal tensor product of Cn and X or equivalently (cid:110) x1 0 ··· ··· ... xn 0 ··· ... 0 ... 0  , xi ∈ X (cid:111) ⊂ Mn(X). Cn(X) = The definition of Rn(X) is similar using a row instead of a column. Adapting the proof of the first example of the third section of [17], we can obtain : Lemma 5.2. Let A be an injective C∗-algebra and k ∈ N∗. Then (1) Ml(Rk(A)) = A ∗-isomorphically and the action is given by : a · (x1, . . . , xk) = (ax1, . . . , axk), for any a, xi ∈ A (2) Mr(Ck(A)) = A ∗-isomorphically and the action is given by : for any a, xi ∈ A  x1 (cid:110)(cid:18) α1A ... xk  · a =  ,  x1a (cid:19) ... xka x y∗ βIn ⊗ 1A , α, β ∈ C, x, y ∈ Rn(A) Proof. We only prove (1), the proof of (2) is similar. Since Rn = n, C), the Paulsen system S of Rn(A) is B((cid:96)2 S = Clearly the C∗-algebra C∗(S) generated by S (inside Mn+1(A)) coin- e (S) of S cides with Mn+1(A). Next we show that the C∗-envelope C∗ e (S), there is a surjective is Mn+1(A). By the universal property of C∗ ∗-homomorphism π : C∗(S) (cid:16) C∗ e (S) such that the following commu- tative diagram holds (cid:111) ⊂ Mn+1(A). C∗(S) π $IIIIIIIII / C∗ e (S) S? $ $ $  O O   / We let (cid:19) (cid:18) 1A 0 0 0 (cid:18) 0 0 0 In ⊗ 1A (cid:19) 17 ). SUBALGEBRAS OF C(Ω, Mn) AND THEIR MODULES p = π( ) and q = π( e (S) satisfying p + q = 1 and pq = 0. Then p and q are projections of C∗ e (S) in '2 × 2' matrix corners. Hence π is Thus we can decompose C∗ corner preserving and there exist π1, π2, π3, π4 such that for any a ∈ A, (cid:19) b ∈ Mn(A), x, y ∈ Rn(A), (cid:18) π1(a) π2(x) (cid:18) a x (cid:19) π( y∗ b ) = π3(y)∗ π4(b) . The (1,2) corners of S and of C∗(S) coincide so π2 is injective (because π extends to C∗(S) the inclusion S ⊂ C∗ e (S)). Similarly π3 is injective. On the other hand, for any a ∈ A, x ∈ Rn(A), π2(ax) = π1(a)π2(x). Thus choosing 'good x', it shows that π1 is injective too. Analogously, using π2(xb) = π2(x)π4(b), for any b ∈ Mn(A), x ∈ Rn(A), the previous argument works to prove the injectivity of π4. Finally, π is injective and so C∗ Mn+1(A) is an injective C∗-algebra. Therefore I(S) = Mn+1(A) e (S) = Mn+1(A). By assumption on A, (cid:19) ∗-isomorphically (cid:18) 1A 0 (cid:19) (cid:18) 1A 0 I(S) 0 0 = A. 0 0 and I11(Rn(A)) = This proves (1). Remark 5.3. We acknowledge that after the paper was submitted, D. Blecher pointed out to the author a more general result : let X be an operator space, then for any p, q ∈ N∗, Ml(Mp,q(X)) = Mp(Ml(X)). We outline the proof. As in [2, Paragraph 4.4.11], we can define the C∗-algebra C(X) = I(X)I(X)∗. Using [2, Corollary 4.6.12], we note that C(Mp,q(X)) = Mp(C(X)). Moreover, from [4], the multiplier algebra of C(X) coincides with I11(X) i.e. M(C(X)) = I11(X). 18 JEAN ROYDOR Hence, using the two previous facts, we can compute Ml(Mp,q(X)) = {a ∈ I11(Mp,q(X)), aMp,q(X) ⊂ Mp,q(X)} = {a ∈ M(C(Mp,q(X))), aMp,q(X) ⊂ Mp,q(X)} = {a ∈ M(Mp(C(X))), aMp,q(X) ⊂ Mp,q(X)} = {a ∈ Mp(M(C(X))), aijX ⊂ X, ∀ i, j} = {a ∈ Mp(I11(X)), aijX ⊂ X, ∀ i, j} = Mp(Ml(X)). The next theorem enables to represent completely contractively a module action on an n-minimal operator space into a C∗-algebra of the form C(Ω, Mn). It constitutes the main result of this section and generalizes (i) ⇔ (iii) of [3, Theorem 2.2]. Theorem 5.4. Let A be a Banach algebra endowed with an operator space structure (resp. a C∗-algebra). Let X be an n-minimal operator space which is also a left Banach A-module. Assume that there is a net (et)t ⊂ Ball(A) satisfying et · x → x, for any x ∈ X. The following are equivalent : (i) X is a left h-module over A. (ii) There exists a compact Hausdorf space Ω, a complete isometry i : X → C(Ω, Mn) and a completely contractive homomorphism (resp. ∗-homomorphism) π : A → C(Ω, Mn) such that i(a · x) = π(a)i(x), for any a ∈ A, x ∈ X Proof. Suppose (i). We first treat the Banach algebra case. By Blecher's oplication Theorem (see [2, Theorem 4.6.2]), we know that there is a completely contractive homomorphism η : A → Ml(X) such that η(a)(x) = a · x, for any a ∈ A, x ∈ X. Using θ ob- tained in Lemma 5.1, we have a completely contractive homomorphism σ = θ ◦ η : A → Ml(I(X)) satisfying σ(a)(x) = a · x, for any a ∈ A, x ∈ X. Moreover, I(X) is an injective n-minimal operator space, so I(X) = ⊕∞ i∈IC(Ωi, Mri,ki) completely isometrically where the Ωi's are Stonean and ri, ki ≤ n, for any i ∈ I. We have the completely isometric unital isomorphisms Ml(I(X)) = ⊕∞ = ⊕∞ = ⊕∞ = ⊕∞ i Ml(C(Ωi, Mri,ki)) i Ml(Cri ⊗min Rki ⊗min C(Ωi)) i Mri(Ml(Rki ⊗min C(Ωi))) i Mri(C(Ωi)) (by Lemma 5.2) SUBALGEBRAS OF C(Ω, Mn) AND THEIR MODULES 19 and via these last identifications, the action of Ml(I(X)) on I(X) is the one inherited from the obvious left action of Mri on Mri,ki. More precisely for any u = (fi ⊗ yi)i ∈ Ml(I(X)) and x = (gi ⊗ xi)i ∈ I(X), u(x) = (figi ⊗ yixi)i. : Mri → Mn (resp. ϕi : Mri,ki → Mn) be the For each i, let ϕi embedding of Mri (resp. Mri,ki) in the 'up-left corner' of Mn. Hence, as in the end of the proof of Corollary 3.3, we have now a ∗-isomorphism ψ : Ml(I(X)) → C(Ω, Mn) (fi ⊗ yi)i fi ⊗ ϕi(yi) i (cid:55)→ (cid:80) (cid:55)→ (cid:80) and a complete isometry j : which verify I(X) → C(Ω, Mn) (gi ⊗ xi)i i gi ⊗ ϕi(xi) j(u(x)) = ψ(u)j(x) for any u ∈ Ml(I(X)), x ∈ I(X) Finally Ω, i = jX and π = ψ ◦ σ satisfy the desired relations. If A is a C∗-algebra, we conclude using the fact that a contractive homomor- phism between C∗-algebras is necessarily a ∗-homomorphism. Remark 5.5. (1) From the previous result, a C∗-algebra which acts 'suitably' on an n-minimal operator space is necessarily an exten- sion of a subhomogeneous C∗-algebra of degree ≤ n. (2) Suppose that A is unital and its action too (i.e. 1 · x = x for any x in X). In the previous result, we cannot expect to obtain a unital completely contractive homomorphism π. Because when A is an operator algebra and A = X, the assumption (i) is verified (see the BRS theorem [2, Theorem 2.3.2]). Hence this particular case leads back to the Remark 3.4. The theorem below could be considered as an 'n-minimal version' of the CES-theorem (see [2, Theorem 3.3.1]). It is the bimodule version of Theorem 5.4 and its proof is 'symmetrically' the same using the two lemmas above. Theorem 5.6. Let A and B be two Banach algebras endowed with an operator space structure (resp. two C∗-algebras). Let X be an n- minimal operator space which is also a Banach A-B-bimodule. Assume that there is a net (et)t ⊂ Ball(A) (resp. (fs)s ⊂ Ball(B)) satisfying et · x → x (resp. x · fs → x), for any x ∈ X. The following are equivalent : (i) X is an h-bimodule over A and B. 20 JEAN ROYDOR (ii) There exists a compact Hausdorf space Ω, a complete isometry i : X → C(Ω, Mn) and two completely contractive homomor- phisms (resp. ∗-homomorphisms) π : A → C(Ω, Mn) and θ : B → C(Ω, Mn) such that i(a · x · b) = π(a)i(x)θ(b), for any a ∈ A, b ∈ B, x ∈ X. The next result states that if A and B are originally n-minimal op- erator algebras, then π and θ can be chosen completely isometric. This corollary generalizes [3, Corollary 2.10]. Corollary 5.7. Let A, B and X be three n-minimal operator spaces such that A and B are approximately unital operator algebras and X is a Banach A-B-bimodule. Assume that there is a net (et)t ⊂ Ball(A) (resp. (fs)s ⊂ Ball(B)) satisfying et · x → x (resp. x· fs → x), for any x ∈ X. The following are equivalent : (i) X is a left h-module over A. (ii) There exists a compact Hausdorf space Ω, a complete isometry i : X → C(Ω, Mn) and completely isometric homomorphisms π : A → C(Ω, Mn) and θ : B → C(Ω, Mn) such that i(a · x · b) = π(a)i(x)θ(b), for any a ∈ A, b ∈ B, x ∈ X. Proof. From Theorem 5.6, there exists a compact Hausdorf space K0, a complete isometry j : X → C(K0, Mn) and completely contractive homomorphisms π0 : A → C(K0, Mn) and θ0 : B → C(K0, Mn) satis- fying j(a · x · b) = π0(a)i(x)θ0(b), for any a ∈ A, b ∈ B, x ∈ X. Moreover by Corollary 2.3, there exists a compact Hausdorf space KA (resp. KB) and a completely isometric homomorphism πA : A → C(KA, Mn) (resp. θB : B → C(KB, Mn)). Let C = C(KA, Mn) ⊕∞ C(K0, Mn) ⊕∞ C(KB, Mn) = C(Ω, Mn) where Ω is the disjoint union of KA, KB and K0. Let i : X → C(Ω, Mn) defined by i(x) = 0 ⊕ j(x) ⊕ 0, for any x ∈ X so i is a complete isometry. Let π : A → C(Ω, Mn) (resp. θ : B → C(Ω, Mn)) defined by π(a) = πA(a)⊕π0(a)⊕0, for any a ∈ A (resp. θ(b) = 0⊕θ0(b)⊕θB(b), for any b ∈ B ). Hence, π and θ are completely isometric homomorphisms. Finally, Ω, π, θ and i satisfy the desired relation. References [1] D.P. Blecher, Commutativity in operator algebras. Proc. Amer. Math. Soc. 109(1990), 709-715. SUBALGEBRAS OF C(Ω, Mn) AND THEIR MODULES 21 [2] D.P. Blecher and C. Le Merdy, Operator algebras and their modules - an op- erator space approach. Oxford University Press, 2004. [3] D.P. Blecher and C. Le Merdy, On function and operator modules. Proc. Amer. Math. Soc. 129(2001), 833-844. [4] D.P. Blecher and V.I. Paulsen, Multiplier of operator spaces, and the injective envelope. Pacific J. Math. 200(2001), 1-17. [5] D. Deckard and C. Pearcy, On matrices over the ring of continuous functions on a Stonian space. Proc. Amer. Math. Soc. 14(1963), 322-328. [6] E. Effros, N. Ozawa and Z.J. Ruan, On injectivity and nuclearity for operator spaces. Duke Math. J. 110(2001), 489-521. [7] E. Effros and Z.J. Ruan, Operator spaces. London Mathematical Society Mono- graphs 23, Oxford University Press, New-York, 2000. [8] K. Grove and G.K. Pedersen, Diagonalizing matrices over C(X). J. Funct. Anal. 59(1984), 65-89. [9] R. Kadison and J. Ringrose, Fundamentals of the theory of operator algebras, Vol. II. Academic Press, New-York, 1986. linking C∗-algebras. J. Funct. Anal. 195(2002), 262-305. [10] M. Kaur and Z.J. Ruan, Local properties of ternary rings of operators and their [11] E.C. Lance, Hilbert C∗-modules - A toolkit for operator algebraist. London Math. Soc. Lecture Notes, 210, Cambridge University Press, Cambridge, 1995. [12] V. Paulsen, Completely bounded maps and operator algebras. Cambridge Stud- ies in Advanced Mathematics 78, Cambridge University Press, Cambridge, 2002. [13] G. Pisier, An introduction to the theory of operator spaces. London Mathemat- ical Society Lecture Note Series 294, Cambridge University Press, Cambridge, 2003. [14] G. Pisier, Exact operator spaces. Colloque sur les alg`ebres d'op´erateurs. in "Recent advances in operator algebras" (Orl´eans 1992) Ast´erisque (Soc. Math. France) 232(1995), 159-186. [15] Z.J. Ruan, Type decomposition and the rectangular AFD property for W ∗- [16] R.R. Smith and D.P. Williams, The decomposition property for C∗-algebras. J. TRO's. Canad. J. Math 56(2004), 843-870. Operator Theory 16(1986), 5-74. [17] C. Zhang, Representation of operator spaces. J. Operator Theory 33(1995), 327-351. Departement de Math´ematiques, Universit´e de Franche-Comt´e, 25030 Besanc¸on cedex, France E-mail address: [email protected]
1705.07665
4
1705
2019-03-03T21:15:26
On the $C^*$-algebra generated by the Koopman representation of a topological full group
[ "math.OA", "math.GR" ]
Let $(X,T,\mu)$ be a Cantor minimal sytem and $[[T]]$ the associated topological full group. We analyze $C^*_\pi([[T]])$, where $\pi$ is the Koopman representation attached to the action of $[[T]]$ on $(X,\mu)$. Specifically, we show that $C^*_\pi([[T]])=C^*_\pi([[T]]')$ and that the kernel of the character $\tau$ on $C^*_\pi([[T]])$ coming from weak containment of the trivial representation is a hereditary $C^*$-subalgebra of $C(X)\rtimes\mathbb{Z}$. Consequently, $\ker\tau$ is stably isomorphic to $C(X)\rtimes\mathbb{Z}$, and $C^*_\pi([[T]]')$ is not AF. We also prove that if $G$ is a finitely generated, elementary amenable group and $C^ *(G)$ has real rank zero, then $G$ is finite.
math.OA
math
ON THE C ∗-ALGEBRA GENERATED BY THE KOOPMAN REPRESENTATION OF A TOPOLOGICAL FULL GROUP EDUARDO SCARPARO Abstract. Let (X, T, µ) be a Cantor minimal sytem and [[T ]] the associated topological full group. We analyze C ∗ π ([[T ]]), where π is the Koopman repre- sentation attached to the action of [[T ]] on (X, µ). Specifically, we show that C ∗ π ([[T ]]′) and that the kernel of the character τ on C ∗ π([[T ]]) coming from containment of the trivial representation is a hereditary C ∗-subalgebra of C(X) ⋊ Z. Consequently, ker τ is stably isomorphic to C(X) ⋊ Z, and C ∗ π([[T ]]′) is not AF. π([[T ]]) = C ∗ We also prove that if G is a finitely generated, elementary amenable group and C ∗(G) has real rank zero, then G is finite. 1. Introduction In this work, we study the real rank zero and AF properties for certain classes of group C ∗-algebras. The motivations are the classical equivalence between amenabil- ity of a group and nuclearity of its C ∗-algebra, and the equivalence between local finiteness of a group and finiteness of its uniform Roe algebra worked out in [21], [11] and [20]. For a compact metric space X, both C(X) being AF and having real rank zero are equivalent to total disconnectedness of X. If a group G is countable and locally finite, then C ∗(G) is clearly AF (hence it has real rank zero). It is an open problem whether there exists a non-locally finite group G such that C ∗(G) is AF. In [10, Theorem 2], Kaniuth proved that if G is a nilpotent group and C ∗(G) has real rank zero, then G is locally finite. In section 2, we show that if G is a finitely generated, elementary amenable group, and C ∗(G) has real rank zero, then G is finite (Theorem 2.3). Our proof relies on the fact that infinite, finitely generated, elementary amenable groups virtually map onto Z [7, Chapter I, Lemma 1]. Let (X, T, µ) be a Cantor minimal system and π the Koopman representation associated to the action of the topological full group [[T ]] on (X, µ). Notice that C ∗([[T ]]) does not have real rank zero, since [[T ]] maps onto Z (by [18, Theorem 1.1(i)], or [5, Proposition 5.5]). On the other hand, by results of Matui, the commutator subgroup [[T ]]′ is simple ([13]) and non-locally finite (this follows from much sharper results from [14]). Hence, commutators of topological full groups form a class which is not covered by Theorem 2.3. 2010 Mathematics Subject Classification. 22D25. Key words and phrases. group C ∗-algebra, real rank zero, topological full group. This work was supported by CNPq, National Council for Scientific and Technological Devel- opment - Brazil. 1 2 EDUARDO SCARPARO π([[T ]]) = C ∗ Futhermore, it was proven by Juschenko and Monod ([8]) that [[T ]] is amenable. In Section 3, we prove that C ∗([[T ]]′) is not AF. This is done by showing that C ∗ π([[T ]]) coming from containment of the trivial representation is a hereditary C ∗-subalgebra of C(X) ⋊ Z. Consequently, ker τ is stably isomorphic to C(X) ⋊ Z, and C ∗ π([[T ]]′) is not AF and has real rank zero. π([[T ]]′), and that the kernel of the character τ on C ∗ In Section 4, we discuss examples coming from odometers. 2. Elementary amenable groups and real rank zero Recall that the class of elementary amenable groups is the smallest class of groups containing all abelian and all finite groups, and closed under taking subgroups, quotients, extensions and inductive limits. We will use the following fact about elementary amenable groups, due to Hillman ([7, Chapter I, Lemma 1]). See also [19, Lemma 1] for a slightly different proof. Lemma 2.1 ([7]). If G is an infinite, finitely generated, elementary amenable group, then there is a subgroup of finite index of G which admits a homomorphism onto Z. A C ∗-algebra A is said to have real rank zero if every hereditary C ∗-subalgebra of A has an approximate unit of projections (not necessarily increasing). We refer the reader to, for example, [3, Section V.7] for other equivalent definitions of real rank zero. Lemma 2.2. If A is an infinite-dimensional, real rank zero C ∗-algebra, then it contains a sequence of non-zero, orthogonal projections. Proof. Since A is infinite-dimensional, there is a sequence (an)n∈N ⊂ A of non-zero, positive elements such that ajak = 0 when j 6= k (see, for example, [9, Exercise 4.6.13] or [15]). For each n ∈ N, take a non-zero projection pn in the hereditary (hence real rank (cid:3) zero) C ∗-subalgebra anAan. By construction, pjpk = 0 when j 6= k. Theorem 2.3. If G is a finitely generated, elementary amenable group and C ∗(G) has real rank zero, then G is finite. Proof. Suppose G is infinite. By Lemma 2.1, there is a subgroup H of G with finite index n, and Φ : H → Z a surjective homomorphism. Let ϕ : C ∗(H) → C ∗(Z) be the ∗-homomorphism induced by Φ, and ϕn : Mn(C ∗(H)) → Mn(C ∗(Z)) the inflation of ϕ. There is an injective ∗-homomorphism ψ : C ∗(G) → Mn(C ∗(H)) such that the image of ϕn ◦ ψ is infinite-dimensional. For the convenience of the reader, we sketch the construction of ψ, which is standard. Let x1, . . . , xn ∈ G be such that x1 = e and G = ⊔n i=1xiH. Consider the following unitary defined on canonical basis vectors: U : n Mi=1 ℓ2(H) → ℓ2(G) δi,h 7→ δxih. Let S : B(ℓ2(G)) → Mn(B(ℓ2(H)) be the isomorphism induced by U . KOOPMAN REPRESENTATION OF A TOPOLOGICAL FULL GROUP 3 By using the left regular representations λG and λH , we see C ∗(G) as contained in B(ℓ2(G)) and analogously for C ∗(H). It is easy to check that S(λG(g)) ∈ Mn(C ∗(H)) for every g ∈ G. Hence, S(C ∗(G)) ⊂ Mn(C ∗(H)). Furthermore, for h ∈ H, we have that S(λG(h))1,1 = λH (h). Let ψ := SC ∗(G). Then ϕn(ψ(C ∗(G))) is infinite-dimensional. Hence, by Lemma 2.2, Mn(C ∗(Z)) ≃ Mn(C(T)) contains a sequence of non-zero, orthogonal projections. Since T is connected, we get a contradiction. Hence, G is finite. (cid:3) Remark 2.4. Recall that a C ∗-algebra A is said to have property (SP) if every non- zero hereditary C ∗-subalgebra of A contains a non-zero projection. Furthermore, A is said to have residual property (SP) if every quotient of A has property (SP) (see [16, Section 7] for more details about these properties). In the proof of Theorem 2.3, the only aspects of real rank zero that were used are that it implies property (SP) and that having real rank zero is closed under taking quotients. In particular, Theorem 2.3 remains true if one replaces "real rank zero" by "residual property (SP)". 3. Koopman representation of a topological full group Given a unitary representation π of a group G, we denote by C ∗ π(G) the C ∗- algebra generated by the image of π. We will denote the Cantor set by X. Let α be an action of a group G on X by homeomorphisms. The topological full group associated to α, denoted by [[α]], is the group of all homeomorphisms γ on X for which there exists a finite partition of X into clopen sets {Ai}n i=1 and g1, . . . , gn ∈ G such that γAi = αgi Ai for 1 ≤ i ≤ n. That is, [[α]] consists of the homeomorphisms on X which are locally given by the action α. Fix T a minimal homeomorphim on X. We denote by [[T ]] the topological full group associated to the Z-action induced by T . Let µ be a T -invariant probability measure on X. Note that µ is also invari- ant under the action of [[T ]] on X. Let π : [[T ]] → B(L2(X, µ)) be given by π(g)(f ) := f ◦ g−1, for g ∈ [[T ]] and f ∈ L2(X, µ). This π is the so called Koopman representation associated to the action of [[T ]] on (X, µ). We will use the faithful representation of C(X) ⋊ Z in B(L2(X, µ)), with C(X) acting by multiplication operators, and, for n∈ Z, δn := π(T n), so that C(X)⋊ Z := span{f δn : f ∈ C(X), n ∈ Z}. Given g ∈ [[T ]] and {Ai}n i=1 a partition of X into clopen sets such that gAi = T niAi for 1 ≤ i ≤ n, notice that (1) π(g) =X 1T ni (Ai)δni. In particular, C ∗ π([[T ]]) ⊂ C(X) ⋊ Z. Definition 3.1. Given n ∈ N, we say that a subset A ⊂ X is n-disjoint if are pairwise disjoint. A, T (A), . . . , T n−1(A) 4 EDUARDO SCARPARO Suppose A ⊂ X is a clopen and n-disjoint set. Consider the symmetric group Sn acting on {0, . . . , n − 1}. For σ ∈ Sn, let σA ∈ [[T ]] be given by (2) σA(x) =(T σ(i)−i(x), x, if 0 ≤ i < n and x ∈ T i(A) if x /∈ ⊔n−1 i=0 T i(A), x ∈ X. Note that, for 0 ≤ i < n, σA(T i(A)) = T σ(i)(A). Lemma 3.2. Let n ≥ 4 and A ⊂ X be a clopen and n-disjoint set. For every σ ∈ Sn, it holds that π(σA) ∈ C ∗ π([[T ]]′). Proof. Notice first that {1T i(A)δi−j}0≤i,j<n forms a system of matrix units in C(X) ⋊ Z of type Mn(C) (we see Mn(C) as matrices indexed by the set {0, . . . , n − 1}). Let B := (⊔n−1 i=0 T i(A))c and ϕ : C⊕ Mn(C) → C(X) ⋊ Z be the ∗-homomorphism given by ϕ(α, eij ) := α1B + 1T i(A)δi−j , for α ∈ C and 0 ≤ i, j ≤ n − 1. Let ρ : Sn → C ⊕ Mn(C) be the direct sum of the trivial representation and the permutation representation. ϕ(ρ(σ)). Given σ ∈ Sn, by (1) and (2), it holds that π(σA) = 1B +P 1T σ(i)(A)δσ(i)−i = Furthermore, since n ≥ 4, the permutation representations of S ′ n and Sn decom- pose into the direct sum of a trivial representation and an irreducible representation of degree n − 1. Therefore, we have that C ∗ π([[T ]]′) for any σ ∈ Sn. ρ ((Sn)′) = C ∗ ρ (Sn). Hence, π(σA) ∈ C ∗ (cid:3) Given A ⊂ X clopen, consider the continuous function tA : A → N x 7→ min{k ≥ 1 : T k(x) ∈ A}. This is the so called function of first return to A. Notice that, for j ∈ Z, it holds that (3) tT j (A) ◦ T jA = tA. Let TA ∈ [[T ]] be defined by (4) TA(x) =(T tA(x)(x), x, if x ∈ A otherwise. , x ∈ X. If B ⊂ X is a clopen set disjoint from A, then TA and TB commute. In order to prove Lemma 3.4, we will have to analyze the spectrum of C ∗-algebras generated by certain commuting unitaries, and the next lemma will be useful for this. We consider the circle T as a pointed space with basepoint 1. Lemma 3.3. The universal C ∗-algebra generated by commuting unitaries z1, . . . , zn T), with subject to the relations {(zi − 1)(zj − 1) = 0 : 1 ≤ i 6= j ≤ n} is C(Wn k=1 KOOPMAN REPRESENTATION OF A TOPOLOGICAL FULL GROUP 5 zk being given by n zk : T → C _i=1 (x, i) 7→(x, 1, if i = k if i 6= k. T → Tn which takes x in the i-th copy of T and sends it into (F (x)i)1≤i≤n ∈ Tn such that F (x)i := x and F (x)j := 1 if T) be given by F ′(f ) := f ◦ F , for f ∈ C(Tn). For 1 ≤ i ≤ n, let wi ∈ C(Tn) be given by wi(y) := yi, for y ∈ Tn. Then Proof. Consider the embedding F : Wn j 6= i. Also let F ′ : C(Tn) → C(Wn i=1 i=1 F ′(wi) = zi. Assume n > 1. Let A := C ∗({(wi − 1)k(wj − 1)l : i 6= j and k, l ∈ N}). We claim that ker F ′ = A. Clearly, A ⊂ ker F ′. Let Y := Tn \ Im(F ). Notice that ker F ′ = {f ∈ C(Tn) : f Im(F ) = 0} ≃ C0(Y ). By the Stone-Weierstrass Theorem, in order to show that A = C0(Y ), it is sufficient to show that, for every y ∈ Y , there is f ∈ A such that f (y) 6= 0, and that A separates the points of Y . The proof of the former condition is trivial, so we only show that A separates the points of Y . Take (x1, . . . , xn), (y1, . . . yn) ∈ Y distinct points. There is i such that xi 6= yi. Without loss of generality, assume xi 6= 1. Take j 6= i such that xj 6= 1. Then, by choosing k ∈ N appropriately, we get (xi − 1)k(xj − 1) 6= (yi − 1)k(yj − 1). Since C(Tn) is the universal C ∗-algebra generated by n commuting unitaries and T) is generated by {z1, . . . , zn}, the result follows. (cid:3) i=1 C(Wn Lemma 3.4. Let A ⊂ X be a clopen and 3-disjoint set. Then π(TA) ∈ C ∗ Proof. Given σ ∈ S3, x ∈ A and 0 ≤ i, j < 3, we have that σATT i(A)σ−1 T j(x) if j 6= σ(i) and π([[T ]]′). A (T j(x)) = σATT i(A)σ−1 A (T σ(i)(x)) = σATT i(A)(T i(x)) = T σ(i)−iT tT i(A)(T i(x))(T i(x)) (∗) = T σ(i)−iT t (T σ(i)(x))(T i(x)) T σ(i) (A) (T σ(i)(x))(T σ(i)(x)) T σ(i) (A) = T t = TT σ(i)(A)(T σ(i)(x)), where the equality in (*) is due to (3). Hence, σATT i(A)σ−1 A = TT σ(i)(A). In particular, for 0 ≤ i, j < 3, we have that TT i(A)(TT j (A))−1 ∈ [[T ]]′. If 0 ≤ i 6= j < 3, then T i(A) and T j(A) are disjoint, hence (π(TT i(A)) − 1)(π(TT j (A)) − 1) = 0. Then, by Lemma 3.3, there is a ∗-homomorphism ϕ : C 3 _i=1 T! → C ∗({π(TT i(A)) : 0 ≤ i < 3} zi 7→ π(TT i−1(A)). 6 EDUARDO SCARPARO Furthermore, by the Stone-Weierstrass theorem, C(W3 j : 1 ≤ i, j ≤ 3}. {ziz∗ i=1 T) is generated by (cid:3) Hence, π(TA) ∈ C ∗ π([[T ]]′). Theorem 3.5. Let (X, T, µ) be a Cantor minimal system and π the Koopman rep- resentation associated to the action of [[T ]] on (X, µ). Then C ∗ π([[T ]]′). π([[T ]]) = C ∗ Proof. By [6, Theorem 4.7], given m ∈ N, [[T]] is generated by {TA, σA : σ ∈ Sn, A ⊂ X is clopen and n-disjoint}. [n≥m By Lemmas 3.2 and 3.4, the result follows. (cid:3) Notice that 1X ∈ L2(X, µ) is invariant under π([[T ]]). Therefore, π contains the trivial representation. Lemma 3.6. Let ρ : G → B(H) be a unitary representation which weakly contains the trvial representation, and τ the associated character on C ∗ ρ (G). Then ker τ = span{1 − ρ(g) : g ∈ G}. Proof. Given d ∈ ker τ and ǫ > 0, take d′ ∈ span ρ(G) such that kd − d′k < ǫ 2 . Then kd − (d′ − τ (d′))k = k(d − d′) + τ (d′ − d)k < ǫ. Furthermore, d′ − τ (d′) ∈ ker τ ∩ span ρ(G). Since ker τ ∩ span ρ(G) = span{1 − ρ(g) : g ∈ G}, the result follows. (cid:3) Theorem 3.7. Let τ be the character on C ∗ the trivial representation. Then ker τ is a hereditary C ∗-subalgebra of C(X) ⋊ Z. π([[T ]]) coming from containment of Proof. We are going to show that, for a ∈ C(X) ⋊ Z and b, c ∈ ker τ , it holds that bac ∈ ker τ . Given A ⊂ X clopen and 2-disjoint, notice that (δ0 − δ1)1A(δ0 − δ−1) = δ0 − (1(A∪T (A))cδ0 + 1T (A)δ1 + 1Aδ−1) ∈ C ∗ π([[T ]]). By using telescoping sums, it follows that, for n, m ∈ Z and A ⊂ X 2-disjoint and clopen, (δ0 − δn)1A(δ0 − δm) ∈ C ∗ π([[T ]]). Given g, h ∈ [[T ]], take a basis B of 2-disjoint, clopen sets for the topology of X. Moreover, assume that, for each A ∈ B, there is n(A), m(A) ∈ Z such that gA = T n(A)A and hh−1(A) = T m(A)h−1(A). Then (δ0 − π(g))1A(δ0 − π(h)) = 1A − δn(A)1A − 1Aδm(A) + δn(A)1Aδm(A) π([[T ]]). = (δ0 − δn(A))1A(δ0 − δm(A)) ∈ C ∗ Since C(X) = span{1A : A ∈ B}, we conclude that, for g, h ∈ [[T ]] and f ∈ C(X), (δ0 − π(g))f (δ0 − π(h)) ∈ C ∗ π([[T ]]). By Lemma 3.6 and the fact that C(X) ⋊ Z = span{f δn : f ∈ C(X), n ∈ Z}, we conclude that, for b, c ∈ ker τ and a ∈ C(X) ⋊ Z, bac ∈ C ∗ π([[T ]]). Since τ is a character, the result follows. (cid:3) KOOPMAN REPRESENTATION OF A TOPOLOGICAL FULL GROUP 7 Corollary 3.8. Let τ be the character on C ∗ π([[T ]]) coming from containment of the trivial representation. Then ker τ is stably isomorphic to C(X) ⋊ Z. In particular, C ∗ π([[T ]]′) has real rank zero and C ∗([[T ]]′) is not AF. Proof. By Theorem 3.7 and the fact that C(X) ⋊ Z is simple, it follows that ker τ is a full, hereditary C ∗-subalgebra of C(X) ⋊ Z. Therefore, [2, Theorem 2.8] implies that ker τ is stably isomorphic to C(X) ⋊ Z. Furthermore, by Theorem 3.5, C ∗ π([[T ]]′). Since C(X) ⋊ Z has real rank zero (see, for instance, [17] for a proof of this fact), and K1(C(X) ⋊ Z) ≃ Z, and K1(A) = 0 for any AF-algebra A, the conclusion follows. (cid:3) π([[T ]]) = C ∗ 4. Odometers We start this section by giving a description of C ∗ π([[T ]]) when T is an odometer map. Given m ∈ N, let Zm := Z/mZ. Example 4.1. Let (nk) be a strictly increasing sequence of natural numbers such that, for every k, nknk+1. Let ρk : Znk+1 → Znk be the surjective homomorphism such that ρk(1) = 1, and define X := {(xk) ∈ Yk∈N Znk : ρk(xk+1) = xk, ∀k ∈ N}. Consider T : X → X (xk) 7→ (xk + 1). Then (X, T ) is a Cantor minimal system, the so called odometer of type (nk). For k ∈ N and l ∈ Znk , let U (k, l) := {(xm) ∈ X : xk = l}. Using the notation from (2) and (4), let, for k ∈ N, Γk := h{TU(k,l), σU(k,0) ∈ [[T ]] : l ∈ Znk , σ ∈ Snk }i. As proven by Matui in [14, Proposition 2.1], Γk ⊂ Γk+1, For k ∈ N, let Ak := span{1U(k,l)δm : l ∈ Znk , m ∈ Z}. Then Ak ⊂ Ak+1, and Γk ≃ Znk ⋊ Snk , and Sk Γk = [[T ]]. C(X) ⋊ Z =Sk Ak. ϕk(1U(k,l)) = el,l, for l ∈ Znk , and, for z ∈ T, Fix k ∈ N and consider the isomorphism ϕk : Ak → C(T, MZnk (C)), such that (ϕk(δ1)(z))i,j :=  if 0 < i ≤ nk − 1 and j = i − 1 if i = 0 and j = nk − 1 1, z, 0, otherwise. Let π : [[T ]] → U (C(X) ⋊ Z) be the homomorphism coming from the Koopman representation and Bk := {b ∈ MZnk (C) : ∀i, j ∈ Znk ,Pr bi,r =Ps bs,j}. Then, for σ ∈ Snk , we have that ϕk(π(σU(k,0))) =P eσ(i),i and Furthermore, ϕk(C ∗(π({TU(k,l) : l ∈ Znk }))) ≃ C(Wl∈Znk C ∗({ϕk(π(σU(k,0))) : σ ∈ Snk }) ≃ Bk. (C)) : f (1) ∈ Bk}. {f ∈ C(T, MZnk T) and ϕk(C ∗ π(Γk)) = In [4], Dykema and Rørdam gave examples of non-locally finite groups G such red(G) has real rank zero. As far as we are aware, there is no known example that C ∗ of non-locally finite group G such that C ∗(G) has real rank zero. 8 EDUARDO SCARPARO Question 4.2. Let (X, T ) be an odometer as in Example 4.1. Does C ∗([[T ]]′) have real rank zero? Example 4.3. Let (X, T ) be an odometer of type (nk) as in Example 4.1. Consider J : X → X (xk) 7→ (−xk). Then J is an involutive homeomorphism on X such that JT J = T −1. Hence, T and J induce an action α of the infinite dihedral group Z ⋊ Z2 on X. We will use Matui's technique ([14, Proposition 2.1]) in order to compute [[α]]. For every γ ∈ (Z ⋊ Z2) \ {e}, it holds that {x ∈ X : αγ(x) = x} has empty interior (it consists of at most two elements). Hence, given g ∈ [[α]], there exists a unique continuous function cg : X → Z ⋊ Z2 such that, for x ∈ X, g(x) = αc(g)(x). For k ∈ N and l ∈ Znk , let U (k, l) be as in Example 4.1 and Γk := {g ∈ [[α]] : cg is constant on U (k, l) for l ∈ Znk }. Define Jk,l ∈ [[α]] by Jk,l(x) =(T 2lJ(x), x, if x ∈ U (k, l) otherwise, x ∈ X. Then Γk = h{TU(k,l), Jk,l, σU(k,0) : l ∈ Znk , σ ∈ Snk }i and (5) Γk ≃ (Z ⋊ Z2)nk ⋊ Snk , Γk ⊂ Γk+1, and [k Γk = [[α]]. Notice that the constant sequence (0) ∈ X is a fixed point for J. Hence, [1, Theorem 3.5] implies that C(X) ⋊ (Z ⋊ Z2) is AF (see also [12]). Moreover, it follows from (5) that the abelianization of [[α]] is locally finite. Therefore, the two obstructions that were used for ruling out the possibility of C ∗([[T ]]) and C ∗([[T ]]′) being AF do not hold for C ∗([[α]]). Question 4.4. Let α be as in Example 4.3. Is C ∗([[α]]) AF?. Acknowledgements Part of this work was carried out while the author was attending the research program Classification of operator algebras: complexity, rigidity, and dynamics at the Mittag-Leffler Institute. The author thanks the organizers of the program and the staff of the institute for the excellent work conditions provided. The author also thanks J. Carri´on, T. Giordano, K. Li and J. Rout for helpful conversations related to topics of this work. References [1] Bratteli, O., Evans, D. E., and Kishimoto, A. Crossed products of totally disconnected spaces by Z2 ∗ Z2. Ergodic Theory Dynam. Systems 13, 3 (1993), 445 -- 484. [2] Brown, L. G. Stable isomorphism of hereditary subalgebras of C ∗-algebras. Pacific J. Math. 71, 2 (1977), 335 -- 348. [3] Davidson, K. R. C ∗-algebras by example, vol. 6 of Fields Institute Monographs. American Mathematical Society, Providence, RI, 1996. [4] Dykema, K. J., and Rørdam, M. Projections in free product C ∗-algebras. II. Math. Z. 234, 1 (2000), 103 -- 113. KOOPMAN REPRESENTATION OF A TOPOLOGICAL FULL GROUP 9 [5] Giordano, T., Putnam, I. F., and Skau, C. F. Full groups of Cantor minimal systems. Israel J. Math. 111 (1999), 285 -- 320. [6] Grigorchuk, R. I., and Medinets, K. S. On the algebraic properties of topological full groups. Mat. Sb. 205, 6 (2014), 87 -- 108. [7] Hillman, J. A. The algebraic characterization of geometric 4-manifolds, vol. 198 of London Mathematical Society Lecture Note Series. Cambridge University Press, Cambridge, 1994. [8] Juschenko, K., and Monod, N. Cantor systems, piecewise translations and simple amenable groups. Ann. of Math. (2) 178, 2 (2013), 775 -- 787. [9] Kadison, R. V., and Ringrose, J. R. Fundamentals of the theory of operator algebras. Vol. I, vol. 15 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 1997. Elementary theory, Reprint of the 1983 original. [10] Kaniuth, E. Group C ∗-algebras of real rank zero or one. Proc. Amer. Math. Soc. 119, 4 (1993), 1347 -- 1354. [11] Kellerhals, J., Monod, N., and Rørdam, M. Non-supramenable groups acting on locally compact spaces. Doc. Math. 18 (2013), 1597 -- 1626. [12] Kumjian, A. An involutive automorphism of the Bunce-Deddens algebra. C. R. Math. Rep. Acad. Sci. Canada 10, 5 (1988), 217 -- 218. [13] Matui, H. Some remarks on topological full groups of Cantor minimal systems. Internat. J. Math. 17, 2 (2006), 231 -- 251. [14] Matui, H. Some remarks on topological full groups of Cantor minimal systems II. Ergodic Theory Dynam. Systems 33, 5 (2013), 1542 -- 1549. [15] Ogasawara, T. Finite-dimensionality of certain Banach algebras. J. Sci. Hiroshima Univ. Ser. A. 17 (1954), 359 -- 364. [16] Pasnicu, C., and Phillips, N. C. Crossed products by spectrally free actions. J. Funct. Anal. 269, 4 (2015), 915 -- 967. [17] Phillips, N. C. Crossed products of the Cantor set by free minimal actions of Zd. Comm. Math. Phys. 256, 1 (2005), 1 -- 42. [18] Putnam, I. F. The C ∗-algebras associated with minimal homeomorphisms of the Cantor set. Pacific J. Math. 136, 2 (1989), 329 -- 353. [19] Scarparo, E. Dynamical characterization of paradoxicality for groups. Oberwolfach Rep. 13, 3 (2016), 2326 -- 2328. [20] Scarparo, E. Characterizations of locally finite actions of groups on sets. Glasgow Mathe- matical Journal (2017), 1 -- 4. [21] Wei, S. On the quasidiagonality of Roe algebras. Sci. China Math. 54, 5 (2011), 1011 -- 1018. Departamento de Matem´atica, Universidade Federal de Santa Catarina, Campus Universit´ario Trindade, 88040-900, Florian´opolis - SC, Brazil E-mail address: [email protected]
1805.09321
1
1805
2018-05-23T09:18:00
Characterization of numerical radius parallelism in $C^*$-algebras
[ "math.OA", "math.FA" ]
Let $v(x)$ be the numerical radius of an element $x$ in a $C^*$-algebra $\mathfrak{A}$. First, we prove several numerical radius inequalities in $\mathfrak{A}$. Particularly, we present a refinement of the triangle inequality for the numerical radius in $C^*$-algebras. In addition, we show that if $x\in\mathfrak{A}$, then $v(x) = \frac{1}{2}\|x\|$ if and only if $\|x\| = \|\mbox{Re}(e^{i\theta}x)\| + \|\mbox{Im}(e^{i\theta}x)\|$ for all $\theta \in \mathbb{R}$. Among other things, we introduce a new type of parallelism in $C^*$-algebras based on numerical radius. More precisely, we consider elements $x$ and $y$ of $\mathfrak{A}$ which satisfy $v(x + \lambda x) = v(x) + v(y)$ for some complex unit $\lambda$. We show that this relation can be characterized in terms of pure states acting on $\mathfrak{A}$.
math.OA
math
CHARACTERIZATION OF NUMERICAL RADIUS PARALLELISM IN C∗-ALGEBRAS ALI ZAMANI Abstract. Let v(x) be the numerical radius of an element x in a C ∗-algebra A. First, we prove several numerical radius inequalities in A. Particularly, we present a refinement of the triangle inequality for the numerical radius in C ∗- algebras. In addition, we show that if x ∈ A, then v(x) = 1 2kxk if and only if kxk = kRe(eiθx)k+kIm(eiθx)k for all θ ∈ R. Among other things, we introduce a new type of parallelism in C ∗-algebras based on numerical radius. More precisely, we consider elements x and y of A which satisfy v(x+λx) = v(x)+v(y) for some complex unit λ. We show that this relation can be characterized in terms of pure states acting on A. 1. Introduction and preliminaries Let A be a unital C∗-algebra with unit denoted by e. We denote by U(A) and Z(A) the group of all unitary elements in A and the centre of A, respectively. For an element x of A, we denote by Re(x) = 1 2i (x − x∗) the real and the imaginary part of x. Let A′ denote the dual space of A, and define the set of normalized states of A by 2(x + x∗) and Im(x) = 1 S(A) = {ϕ ∈ A′ : ϕ(e) = kϕk = 1}. A linear functional ϕ ∈ A′ is said to be positive, and write ϕ ≥ 0, if ϕ(x∗x) ≥ 0 for all x ∈ A . Note that the set of normalized states S(A) is nothing but S(A) = {ϕ ∈ A′ : ϕ ≥ 0 and ϕ(e) = 1}. Recall that a positive linear functional ϕ on A is said to be pure if for every positive functional ψ on A satisfying ψ(x∗x) ≤ ϕ(x∗x) for all x ∈ A, there is a scalar 0 ≤ µ ≤ 1 such that ψ = µϕ. The set of pure states on A is denoted by P(A). The numerical range of an element x ∈ A is V (x) = {ϕ(x) : ϕ ∈ S(A)}. It is a nonempty compact and convex set of the complex plane C, and its maximum modulus is the numerical radius v(x) of x; i.e. v(x) = sup{z : z ∈ V (x)}. It is well-known that v(·) define a norm on A, which is equivalent to the C∗-norm k · k. In fact, the following inequalities are well-known: (x ∈ A). 1 2kxk ≤ v(x) ≤ kxk (1.1) It is a basic fact that the norm v(·) is self-adjoint (i.e., v(x∗) = v(x) for every x ∈ A) and also, if x is normal, then v(x) = kxk. Also, since P(A) coincides with 2010 Mathematics Subject Classification. Primary 46L05; Secondary 47A12, 47A30, 46B20. Key words and phrases. C ∗-algebra, numerical radius, inequality, parallelism. 1 2 A. ZAMANI the set of all extremal points of S(A), thus for every x ∈ A we have v(x) = sup ϕ∈S(A)ϕ(x) = sup ϕ∈P(A)ϕ(x). When A = B(H ) is the C∗-algebra of all bounded linear operators on a com- plex Hilbert space (cid:0)H ,h·,·i(cid:1) and T ∈ B(H ), it well known that V (T ) is the closure of W (T ), the spatial numerical range of T defined by W (T ) = (cid:8)hT x, xi : x ∈ H ,kxk = 1(cid:9). It is known as well that W (T ) is a nonempty bounded con- vex subset of C (not necessarily closed), and its supremum modulus, denoted by ω(x) = sup{z : z ∈ W (T )}, is called the spatial numerical radius of T and coincides with v(T ). For more material about the numerical radius and other information on the basic theory of algebraic numerical range, we refer the reader to [5] and [11]. Some other related topics can be found in [3, 8, 9, 12, 18, 20, 22, 26]. Now, let (X ,k· k) be a normed space. An element x ∈ X is said to be norm -- parallel to another element y ∈ X (see [21, 28]), in short x k y, if kx + λyk = kxk +kyk for some λ ∈ T. Here, as usual, T is the unit cycle of the complex plane In the context of continuous functions, the well-known Daugavet equation C. kT + Idk = kTk + 1 is a particular case of parallelism. Here Id denotes the identity function. This property of a function, apart from being interesting in its own right, arises naturally in problems dealing with best approximations in function spaces; see [23] and the references therein. In the framework of inner product spaces, the norm -- parallel relation is exactly the usual vectorial parallel relation, that is, x k y if and only if x and y are linearly dependent. In the setting of normed linear spaces, two linearly dependent vectors are norm -- parallel, but the converse is false in general. Some characterizations of the norm -- parallelism for operators on various Banach spaces and elements of an arbitrary Hilbert C∗-module were given in [6, 10, 17, 24, 27, 28, 29]. Now, let us introduce a new type of parallelism in C∗-algebras based on nu- merical radius. Definition 1.1. An element x ∈ A is called the numerical radius parallel to another element y ∈ A, denoted by xkv y, if v(x + λx) = v(x) + v(y) for some λ ∈ T. It is easy to see that the numerical radius parallelism is reflexive (xkv x), sym- metric (xkv y if and only if y kv x) and R-homogenous (xkv y ⇒ αxkv βy for all α, β ∈ R)). Notice that two linearly dependent elements are numerical radius parallel. The converse is however not true, in general. The organization of this paper will be as follows. Inspired by the numerical radius inequalities of bounded linear operators in [1], [2], [13], [14], [15], [16], [25] and by using some ideas of them, we firstly state a useful characterization of the numerical radius for elements of a C∗-algebra, as follows: v(x) = sup θ∈RkRe(eiθx)k. CHARACTERIZATION OF NUMERICAL RADIUS PARALLELISM 3 We then apply it to prove that several numerical radius inequalities in C∗- algebras. Moreover, we give new improvements of the inequalities (1.1). We also give an expression of v(x) in terms of the real and imaginary parts of x ∈ A, as follows: v(x) = sup α2+β2=1(cid:13)(cid:13)αRe(x) + βIm(x)(cid:13)(cid:13). Particularly, then we show that if x ∈ A, then v(x) = 1 2kxk if and only if kxk = kRe(eiθx)k + kIm(eiθx)k for all θ ∈ R. Our results generalize recent numerical radius inequalities of bounded linear operators due to Kittaneh et al. [14, 16, 25]. In addition, we present a refinement of the triangle inequality for the numerical radius in C∗-algebras. We then apply it to give a necessary condition for the numerical radius parallelism. Furthermore, for two elements x and y in A we show that xkv y if and only if there exists a pure state ϕ on A such that ϕ(x)ϕ(y) = v(x)v(y). Finally, we prove that if c ∈ Z(A) ∩ U(A), then xkv y holds exactly when cxkv cy. We start our work with the following lemma. 2. Main results Lemma 2.1. Let A be a C∗-algebra and let ϕ be a state over A. For x ∈ A the following statements hold. (i) sup (ii) sup θ∈R(cid:12)(cid:12)Re(cid:0)eiθϕ(x)(cid:1)(cid:12)(cid:12) = ϕ(x). θ∈R(cid:12)(cid:12)Im(cid:0)eiθϕ(x)(cid:1)(cid:12)(cid:12) = ϕ(x). Proof. We may assume that ϕ(x) 6= 0 otherwise (i) and (ii) trivially hold. (i) Put eiθ0 = ϕ(x) ϕ(x) . Then we have ϕ(x) = (cid:12)(cid:12)Re(cid:0)eiθ0ϕ(x)(cid:1)(cid:12)(cid:12) ≤ sup θ∈R (cid:12)(cid:12)Re(cid:0)eiθϕ(x)(cid:1)(cid:12)(cid:12) ≤ sup θ∈R (cid:12)(cid:12)eiθϕ(x)(cid:12)(cid:12) = ϕ(x), and hence ϕ(x) = supθ∈R(cid:12)(cid:12)Re(cid:0)eiθϕ(x)(cid:1)(cid:12)(cid:12). (ii) By replacing x in (i) by ix, we obtain sup θ∈R (cid:12)(cid:12)Im(cid:0)eiθϕ(x)(cid:1)(cid:12)(cid:12) = sup θ∈R (cid:12)(cid:12)Re(cid:0)eiθϕ(ix)(cid:1)(cid:12)(cid:12) = ϕ(ix) = ϕ(x). (cid:3) Now, we are in a position to state two useful characterizations of the numerical radius for elements of a C∗-algebra. Theorem 2.2. Let A be a C∗-algebra. For x ∈ A the following statements hold. (i) sup (ii) sup θ∈RkRe(eiθx)k = v(x). θ∈RkIm(eiθx)k = v(x). Proof. (i) Since Re(eiθx) is self adjoint for any θ ∈ R, we have kRe(eiθx)k = v(Re(eiθx)). 4 A. ZAMANI Therefore, we get sup θ∈R kRe(eiθx)k = sup v(Re(eiθx)) sup sup ϕ∈S(A)ϕ(cid:0)Re(eiθx)(cid:1) ϕ∈S(A)Re(eiθϕ(x)) θ∈R Re(eiθϕ(x)) θ∈R = sup θ∈R = sup θ∈R = sup ϕ∈S(A) ϕ∈S(A)ϕ(x) = sup sup = v(x). (cid:0)by Lemma 2.1 (i)(cid:1) Thus sup θ∈RkRe(eiθx)k = v(x). (ii) By replacing x in (i) by ix, we reach that sup θ∈R kIm(eiθx)k = sup θ∈R kRe(eiθ(ix))k = v(ix) = v(x). (cid:3) Recall that the Crawford number of T ∈ B(H ) is defined by c(T ) = inf{hT x, xi : x ∈ H ,kxk = 1}. This concept is useful in studying linear operators (e.g., see [1, 26], and their references). The Crawford number of z ∈ A can be defined by c(z) = inf{ϕ(z) : ϕ ∈ S(A)}. In the following theorem, we give new improvement of the inequalities (1.1). Theorem 2.3. Let A be a C∗-algebra. For x ∈ A the following statements hold. (i) 1 2kxk ≤ 1 (ii) v(x) ≤ 1 2q(cid:13)(cid:13)x2 + x∗2(cid:13)(cid:13) + 2c(x2) ≤ v(x). 2q(cid:13)(cid:13)x2 + x∗2(cid:13)(cid:13) + 2v(x2) ≤ 1 2(cid:0)kxk + kx2k 1 2(cid:1) ≤ kxk. Proof. (i) Let x ∈ A. By [19, Theorem 3.3.6] there is a state ϕ over A such that ϕ(cid:0)x2 + x∗2(cid:1) = (cid:13)(cid:13)x2 + x∗2(cid:13)(cid:13). (2.1) CHARACTERIZATION OF NUMERICAL RADIUS PARALLELISM 5 Let θ0 be a real number such that ϕ(x2) = e2iθ0 ϕ(x2). Then, by Theorem 2.2 (i), we have v(x) ≥ kRe(eiθ0x)k = = = ≥ = = ≥ 1 1 1 2keiθ0x + e−iθ0x∗k 1 2q(cid:13)(cid:13)(cid:0)eiθ0x + e−iθ0x∗(cid:1)(cid:0)eiθ0x + e−iθ0x∗(cid:1)∗(cid:13)(cid:13) 2q(cid:13)(cid:13)x2 + x∗2 + 2Re(e2iθ0x2)(cid:13)(cid:13) 2q(cid:12)(cid:12)ϕ(cid:0)x2 + x∗2 + 2Re(e2iθ0 x2)(cid:1)(cid:12)(cid:12) 2q(cid:12)(cid:12)ϕ(cid:0)x2 + x∗2(cid:1) + 2Re(cid:0)e2iθ0ϕ(x2)(cid:1)(cid:12)(cid:12) 2q(cid:13)(cid:13)x2 + x∗2(cid:13)(cid:13) + 2ϕ(x2) 2q(cid:13)(cid:13)x2 + x∗2(cid:13)(cid:13) + 2c(x2) ≥ 1 2kxk, 1 1 1 (cid:0)by (2.1)(cid:1) which proves the inequalities in (i). (ii) By Theorem 2.2 (i), as in the proof of (i) we get v(x) = sup 1 sup sup 1 2 1 2 1 θ∈R kRe(eiθx)k θ∈R q(cid:13)(cid:13)x2 + x∗2 + 2Re(e2iθx2)(cid:13)(cid:13) θ∈R q(cid:13)(cid:13)x2 + x∗2k + 2kRe(e2iθx2)(cid:13)(cid:13) 2r(cid:13)(cid:13)x2 + x∗2k + 2 sup θ∈R kRe(e2iθx2)(cid:13)(cid:13) 2q(cid:13)(cid:13)x2 + x∗2(cid:13)(cid:13) + 2v(x2) 2pkxk2 + kx2k + 2v(x2) 2pkxk2 + 3kx2k 2qkxk2 + 2kxkkx2k 2(cid:0)kxk + kx2k 2(cid:1) ≤ kxk, (cid:0)by (1.1)(cid:1) 2 + kx2k 1 1 1 1 1 1 = ≤ ≤ = ≤ ≤ ≤ = (cid:0)since kx2k = kx2k 1 2kx2k 1 2 ≤ kxkkx2k 1 2(cid:1) (cid:3) which proves the inequalities in (ii). As a consequence of Theorem 2.3, we have the following result. 2kxk. Corollary 2.4. Let A be a C∗-algebra. v(x) = 1 Proof. Since x2 = 0, by Theorem 2.3 (ii), we obtain v(x) ≤ 1 2kxk. We also have that 1 2kxk ≤ v(x) for every x ∈ A. Thus v(x) = 1 2(cid:0)kxk + kx2k 2kxk. If x ∈ A is such that x2 = 0, then 2(cid:1) = (cid:3) 1 1 6 A. ZAMANI The following result is another consequence of Theorem 2.3. Corollary 2.5. Let A be a C∗-algebra. If x ∈ A is such that v(x) = kxk, then kx2k = kxk2. Proof. It follows from Theorem 2.2 (ii) that v(x) = kxk implies kxk ≤ 1 kx2k 2 , or equivalently kx2k = kxk2. 2(cid:0)kxk + 2(cid:1) ≤ kxk. Thus kxk = kx2k In the following theorem we give an expression of v(x) in terms of the real and (cid:3) 1 1 imaginary parts of x ∈ A. Theorem 2.6. Let A be a C∗-algebra and let x ∈ A. Then for α, β ∈ R, the following statements hold. (i) sup α2+β2=1(cid:13)(cid:13)αRe(x) + βIm(x)(cid:13)(cid:13) = v(x). (ii) max(cid:8)kRe(x)k,kIm(x)k(cid:9) ≤ v(x). Proof. (i) Let θ ∈ R. Put α = cos θ and β = − sin θ. We have Re(eiθx) = eiθx + e−iθx∗ 2 (cos θ + i sin θ)x + (cos θ − i sin θ)x∗ = 2 = (cos θ) x + x∗ 2 − (sin θ) x − x∗ 2i = αRe(x) + βIm(x). Therefore sup θ∈R kRe(eiθx)k = sup α2+β2=1(cid:13)(cid:13)αRe(x) + βIm(x)(cid:13)(cid:13), and hence by Theorem 2.2 (i) we obtain v(x) = sup α2+β2=1(cid:13)(cid:13)αRe(x) + βIm(x)(cid:13)(cid:13). In the next result, we obtain a necessary and sufficient condition for v(x) = (ii) By setting (α, β) = (1, 0) and (α, β) = (0, 1) in (i), we get kRe(x)k ≤ v(x) and kIm(x)k ≤ v(x). Thus max(cid:8)kRe(x)k,kIm(x)k(cid:9) ≤ v(x). 2kxk to hold. We will need the following lemma. Lemma 2.7. [28, Corollary 4.4] Let A be a C∗-algebra and let x, y ∈ A. Then the following statements are equivalent: (cid:3) 1 (i) x k y. (ii) There exists a state ϕ over A such that ϕ(x∗y) = kxkkyk. Theorem 2.8. Let A be a C∗-algebra and let x ∈ A. Then the following state- ments are equivalent: 2kxk. (i) v(x) = 1 (ii) kxk = kRe(eiθx)k + kIm(eiθx)k for all θ ∈ R. CHARACTERIZATION OF NUMERICAL RADIUS PARALLELISM 7 Proof. (i)⇒(ii) Suppose that v(x) = 1 2kxk. Then for any θ ∈ R, we have kxk = keiθxk = kRe(eiθx) + iIm(eiθx)k ≤ kRe(eiθx)k + kIm(eiθx)k ≤ 2 max(cid:8)kRe(eiθx)k,kIm(eiθx)k(cid:9) ≤ 2v(eiθx) = 2v(x) = kxk, and hence kxk = kRe(eiθx)k + kIm(eiθx)k. (cid:0)by Theorem 2.6 (ii)(cid:1) (ii)⇒(i) Suppose (ii) holds. Thus for all θ ∈ R, so Re(eiθx) k Im(eiθx). By Lemma 2.7, there exists a state ϕ over A such that and hence From this it follows that v(cid:0)Re(eiθx)Im(eiθx)(cid:1) = kRe(eiθx)k kIm(eiθx)k, so by Theorem 2.2 (ii) we reach that (2.2) o (2.3) (2.4) (2.5) On the other hand, Im(cid:0)Re(eiθx)Im(eiθx)(cid:1) = Im(cid:16)( 2 kRe(eiθx) + iIm(eiθx)k = kRe(eiθx)k + kIm(eiθx)k, (cid:12)(cid:12)(cid:12) ϕ(cid:16)(cid:0)Re(eiθx)(cid:1)∗Im(eiθx)(cid:17)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)ϕ(cid:0)Re(eiθx)Im(eiθx)(cid:1)(cid:12)(cid:12) = kRe(eiθx)k kIm(eiθx)k. = kRe(eiθx)k kIm(eiθx)k, 1 = )( 4i 4i 2i eiθx + e−iθx∗ kRe(eiθx)k kIm(eiθx)k = (cid:13)(cid:13)Im(cid:0)(Re(eiθx)Im(eiθx)(cid:1)(cid:13)(cid:13). eiθx − e−iθx∗ )(cid:17) = Im(cid:16) e2iθx2 − e−2iθx∗2 − xx∗ + x∗x (cid:17) 2in e2iθx2 − e−2iθx∗2 − xx∗ + x∗x − −4i = Im(cid:0)Re(x)Im(x)(cid:1), kRe(eiθx)k kIm(eiθx)k = (cid:13)(cid:13)Im(cid:0)Re(x)Im(x)(cid:1)(cid:13)(cid:13). kRe(eiθx)k = kxk +qkxk2 − 4(cid:13)(cid:13)Im(cid:0)Re(x)Im(x)(cid:1)(cid:13)(cid:13) kIm(eiθx)k = kxk −qkxk2 − 4(cid:13)(cid:13)Im(cid:0)Re(x)Im(x)(cid:1)(cid:13)(cid:13) xx∗ − x∗x 2 2 = 4 . and by (2.2) we get Thus for all θ ∈ R, by (ii) and (2.3) we obtain and e−2iθx∗2 − e2iθx2 − xx∗ + x∗x 8 Since A. ZAMANI Re(eiθx) = (cos θ + i sin θ)x + (cos θ − i sin θ)x∗ 2 = cos θRe(x) − sin θIm(x) and Im(eiθx) = (cos θ + i sin θ)x − (cos θ − i sin θ)x∗ 2i = sin θRe(x) + cos θIm(x) So, from relations (2.4) and (2.5), we conclude that the functions kRe(eiθx)k,kIm(eiθx)k are continuous on θ ∈ R and therefore they must be constant, i.e., kRe(eiθx)k = kIm(eiθx)k = 1 2kxk (θ ∈ R). Thus sup θ∈RkRe(eiθx)k = 1 2kxk. 1 2kxk. Now, by Theorem 2.2 (i) we conclude that v(x) = (cid:3) Another, we present new improvement of the inequalities (1.1). Theorem 2.9. Let A be a C∗-algebra. For x ∈ A the following statements hold. (i) 1 2kxk ≤ 1 2pkx∗x + xx∗k ≤ v(x). (ii) v(x) ≤ 1√2pkx∗x + xx∗k ≤ kxk. 2kxk ≤ 1 Proof. (i) Let x ∈ A. Clearly, 1 tions, 2pkx∗x + xx∗k. But, by simple computa- x∗x + xx∗ = 2Re2(x) + 2Im2(x). Consequently, by Theorem 2.6 (ii) we get 1 2pkx∗x + xx∗k = ≤ ≤ 2pkx∗x + xx∗k ≤ v(x). 1 1 2q(cid:13)(cid:13)2Re2(x) + 2Im2(x)(cid:13)(cid:13) 2p2kRe(x)k2 + 2kIm(x)k2 2p2v2(x) + 2v2(x) = v(x) 1 Therefore 1 (ii) Obviously, 1√2pkx∗x + xx∗k ≤ kxk. Now, let π : A → B(H ) be a non- degenerate faithful representation of A on some Hilbert space H (see [7, Theorem 2.6.1]). Let α, β ∈ R satisfy α2 + β2 = 1. Then for any unit vector ξ ∈ H , we CHARACTERIZATION OF NUMERICAL RADIUS PARALLELISM 9 have 0 0 0 0 0 0 Im(π(x)) (cid:13)(cid:13)π(cid:0)αRe(x) + βIm(x)(cid:1)ξ(cid:13)(cid:13) = (cid:13)(cid:13)(cid:13)(cid:13) βξ(cid:21)(cid:13)(cid:13)(cid:13)(cid:13) (cid:20)π(Re(x)) π(Im(x)) (cid:21)(cid:20)αξ ≤ (cid:13)(cid:13)(cid:13)(cid:13) (cid:21)(cid:13)(cid:13)(cid:13)(cid:13) (cid:20)π(Re(x)) π(Im(x)) = (cid:13)(cid:13)(cid:13)(cid:13) (cid:21)(cid:13)(cid:13)(cid:13)(cid:13) (cid:20)Re(π(x)) = (cid:13)(cid:13)(cid:13)(cid:13) Im(π(x)) 0(cid:21)(cid:13)(cid:13)(cid:13)(cid:13) (cid:20)Re(π(x)) (cid:21)(cid:20)Re(π(x)) 0 = (cid:13)(cid:13)Re2(π(x)) + Im2(π(x))(cid:13)(cid:13) √2(cid:13)(cid:13)π(x)∗π(x) + π(x)π(x)∗(cid:13)(cid:13) √2(cid:13)(cid:13)π(x∗x + xx∗)(cid:13)(cid:13) 1 √2kx∗x + xx∗k Im(π(x)) = = = 1 1 2 . 1 0 0 1 2 1 2 1 2 1 2 (since π is isometric) (since π is representation) Hence we have (cid:13)(cid:13)π(αRe(x) + βIm(x))ξ(cid:13)(cid:13) ≤ 1√2pkx∗x + xx∗k and so by taking the supremum over all ξ ∈ H we obtain kπ(αRe(x) + βIm(x))(cid:13)(cid:13) ≤ 1√2pkx∗x + xx∗k. From this it follows that (cid:13)(cid:13)αRe(x) + βIm(x)(cid:13)(cid:13) ≤ 1√2pkx∗x + xx∗k and hence sup α2+β2=1(cid:13)(cid:13)αRe(x) + βIm(x)(cid:13)(cid:13) ≤ 1 √2pkx∗x + xx∗k. Now, by Theorem 2.6 (i) we conclude that v(x) ≤ 1√2pkx∗x + xx∗k. (cid:3) In what follows, r(x) stands for the spectral radius of an arbitrary element x in a C∗-algebra A. It is well known that for every x ∈ A, we have r(x) ≤ kxk and that equality holds in this inequality if x is normal. In the following lemma we obtain a spectral radius inequality for sums of elements in C∗-algebras. Lemma 2.10. Let A be a C∗-algebra and let z, w ∈ A. Then r(z + w) ≤ 1 2(cid:16)kzk + kwk +p(kzk − kwk)2 + 4 min{kzwk,kwzk}(cid:17). Proof. We first recall that [13, Corollary 1] tells us that r(T + S) ≤ 1 2(cid:16)kTk + kSk +p(kTk − kSk)2 + 4 min{kT Sk,kSTk}(cid:17), (2.6) for all bounded linear operators T, S that acting on a Hilbert space. Now, let π : A → B(H ) be a non-degenerate faithful representation of A on some Hilbert space H (see [7, Theorem 2.6.1]). Since π is isometric, by letting T = π(z) and 10 A. ZAMANI S = π(w) in (2.6), we obtain r(z + w) = r(cid:0)π(z) + π(w(cid:1)) 1 ≤ 2(cid:16)kπ(z)k + kπ(w)k +p(kπ(z)k − kπ(w)k)2 + 4 min{kπ(z)π(w)k,kπ(w)π(z)k}(cid:17) 1 = 2(cid:16)kzk + kwk +p(kzk − kwk)2 + 4 min{kzwk,kwzk}(cid:17), and the statement is proved. (cid:3) Now, we present a refinement of the triangle inequality for the numerical radius in C∗-algebras. Theorem 2.11. Let A be a C∗-algebra. For x, y ∈ A the following statements hold. (i) (ii) v(x + y) ≤ 1 2(cid:0)v(x) + v(y)(cid:1) + 1 2r(cid:0)v(x) − v(y)(cid:1)2 + 4 sup θ∈R (cid:13)(cid:13)Re(eiθx)Re(eiθy)(cid:13)(cid:13) ≤ v(x) + v(y). v(x + y) ≤ 1 2(cid:0)v(x) + v(y)(cid:1) + 1 2r(cid:0)v(x) − v(y)(cid:1)2 ≤ v(x) + v(y). + 4 sup θ∈R (cid:13)(cid:13)Im(eiθx)Im(eiθy)(cid:13)(cid:13) Proof. (i) Since Re(eiθ(x + y)) is self adjoint for any θ ∈ R, we have kRe(eiθ(x + y))k = r(cid:0)Re(eiθ(x + y))(cid:1). CHARACTERIZATION OF NUMERICAL RADIUS PARALLELISM 11 So, by letting z = Re(eiθx) and w = Re(eiθy) in Lemma 2.10, we obtain kRe(eiθ(x + y))k = r(cid:0)Re(eiθ(x + y))(cid:1) 1 = r(cid:0)Re(eiθx) + Re(eiθy)(cid:1) 2(cid:16)kRe(eiθx)k + kRe(eiθy)k ≤ +p(kRe(eiθx)k − kRe(eiθy)k)2 + 4kRe(eiθx)Re(eiθy)k(cid:17) = (cid:13)(cid:13)(cid:13)(cid:13) (cid:20) kRe(eiθx)k pkRe(eiθx)Re(eiθy)k θ∈RkRe(eiθx)k ≤ (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)   = (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) pkRe(eiθx)Re(eiθy)k (cid:21)(cid:13)(cid:13)(cid:13)(cid:13) kRe(eiθy)k θ∈RpkRe(eiθx)Re(eiθy)k  sup  θ∈RkRe(eiθy)k θ∈RpkRe(eiθx)Re(eiθy)k   θ∈RpkRe(eiθx)Re(eiθy)k θ∈RpkRe(eiθx)Re(eiθy)k (by Theorem 2.2 (i)) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)   v(x) v(y) sup sup sup sup sup 1 = 2(cid:0)v(x) + v(y)(cid:1) 1 + 2r(v(x) − v(y))2 + 4sup θ∈RkRe(eiθx)Re(eiθy)k (by the norm monotonicity of matrices with nonnegative entries) Therefore, for every θ ∈ R we have kRe(eiθ(x + y))k ≤ 1 2(cid:0)v(x) + v(y)(cid:1) 1 + 2r(v(x) − v(y))2 + 4sup θ∈RkRe(eiθx)Re(eiθy)k, and hence sup θ∈RkRe(eiθ(x + y))k ≤ 1 2(cid:0)v(x) + v(y)(cid:1) 1 + 2r(v(x) − v(y))2 + 4sup θ∈RkRe(eiθx)Re(eiθy)k. Now, by Lemma 2.2 (i) and the above inequality we get 1 2(cid:16)v(x) + v(y) +r(v(x) − v(y))2 + 4sup v(x + y) ≤ Furthermore, by Lemma 2.2 (i) we have θ∈RkRe(eiθx)Re(eiθy)k(cid:17). (2.7) sup θ∈RkRe(eiθx)Re(eiθy)k ≤ sup θ∈RkRe(eiθx)ksup θ∈RkRe(eiθy)k = v(x)v(y). Thus the inequalities (i) follow from (2.7) and the above inequality. (ii) It is enough to replace x and y in (i) by ix and iy, respectively. (cid:3) 12 A. ZAMANI Corollary 2.12. Let A be a C∗-algebra and let x, y ∈ A. If xkv y then there exists θ0 ∈ R such that the following statements hold. (i) sup (ii) sup θ∈R(cid:13)(cid:13)Re(eiθx)Re(ei(θ+θ0)y)(cid:13)(cid:13) = sup θ∈R(cid:13)(cid:13)Im(eiθx)Im(ei(θ+θ0)y)(cid:13)(cid:13) = sup θ∈RkRe(eiθx)kkRe(eiθy)k. θ∈RkIm(eiθx)kkIm(eiθy)k. Proof. Since xkv y, so there exists θ0 ∈ R such that v(x + eiθ0y) = v(x) + v(y) = v(x) + v(eiθ0y). Hence by Theorem 2.11 it follows that and sup θ∈R (cid:13)(cid:13)Re(eiθx)Re(ei(θ+θ0)y)(cid:13)(cid:13) = v(x)v(y) sup θ∈R (cid:13)(cid:13)Im(eiθx)Im(ei(θ+θ0)y)(cid:13)(cid:13) = v(x)v(y). These, together with Theorem 2.2, imply that (i) and (ii). (cid:3) In the following result we characterize the numerical radius parallelism for elements of a C∗-algebra. Theorem 2.13. Let A be a C∗-algebra and let x, y ∈ A. Then the following statements are equivalent: (i) xkv y. (ii) There exists a pure state ϕ on A such that ϕ(x)ϕ(y) = v(x)v(y). Proof. (i)⇒(ii) Let xkv y. Thus v(x+λy) = v(x)+v(y) for some λ ∈ T. Therefore, there exists a pure state ϕ on A such that ϕ(x + λy) = v(x + λy). From this it follows that v(x) + v(y) = v(x + λy) = ϕ(x + λy) ≤ ϕ(x) + ϕ(y) ≤ v(x) + ϕ(y) ≤ v(x) + v(y), and hence ϕ(x) = v(x) and ϕ(y) = v(y). Thus ϕ(x)ϕ(y) = v(x)v(y). trivially holds. Put λ = ϕ(x)ϕ(y) ϕ(x)ϕ(y) ϕ(y). Since (ii)⇒(i) Suppose (ii) holds. We may assume that ϕ(x)ϕ(y) 6= 0 otherwise (i) . Here, ϕ(y) denotes the complex conjugate of v(x)v(y) = ϕ(x)ϕ(y) ≤ ϕ(x)v(y) ≤ v(x)v(y), CHARACTERIZATION OF NUMERICAL RADIUS PARALLELISM 13 we have v(x) = ϕ(x) and so v(y) = ϕ(y). Therefore, v(x) + v(y) = ϕ(x) + ϕ(y) ϕ(y) ϕ(y) ϕ(x)ϕ(y) ϕ(x)ϕ(y) ϕ(x) + ϕ(x) + = (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) = (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ϕ(y)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ϕ(y)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) = ϕ(x + λy) ≤ v(x + λy) ≤ v(x) + v(y). This implies that v(x + λy) = v(x) + v(y) and hence xkv y. (cid:3) As a consequence of the preceding theorem, we have the following result. Corollary 2.14. Let A be a C∗-algebra with identity e. Then for every x ∈ A, xkv e. Proof. Let x ∈ A. Thus there exists a pure state ϕ on A such that ϕ(x) = v(x) and so ϕ(x)ϕ(e) = ϕ(x) = v(x) = v(x)v(e). Therefore, Theorem 2.13 tells us that xkv e. (cid:3) As an immediate consequence of Theorem 2.13, Lemma 2.1 and Theorem 2.2, we have the following result. Corollary 2.15. Let A be a C∗-algebra and let x, y ∈ A. following statements hold. (i) There exists a pure state ϕ on A such that If xkv y then the (ii) There exists a pure state ϕ on A such that sup θ∈R(cid:12)(cid:12)Re(cid:0)eiθϕ(x)(cid:1)(cid:12)(cid:12)(cid:12)(cid:12)Re(cid:0)eiθϕ(y)(cid:1)(cid:12)(cid:12) = sup θ∈R(cid:12)(cid:12)Im(cid:0)eiθϕ(x)(cid:1)(cid:12)(cid:12)(cid:12)(cid:12)Im(cid:0)eiθϕ(y)(cid:1)(cid:12)(cid:12) = sup θ∈R(cid:13)(cid:13)Re(eiθx)(cid:13)(cid:13)(cid:13)(cid:13)Re(eiθy)(cid:13)(cid:13). θ∈R(cid:13)(cid:13)Im(eiθx)(cid:13)(cid:13)(cid:13)(cid:13)Im(eiθy)(cid:13)(cid:13). sup We closed this paper with the following equivalence theorem. In fact, our next result is a characterization of left or right homogenous for the numerical radius parallelism in unital C∗-algebras. Theorem 2.16. Let A be a C∗-algebra with identity e and let c ∈ Z(A) ∩ U(A). Then for every x, y ∈ A the following statements are equivalent: (i) xkv y. (ii) cxkv cy. (iii) xc kv yc. Proof. Firstly, we show that v(cz) = v(z) = v(zc) for all z ∈ A. Let ϕ be a pure state on A. By [7, Proposition 2.4.4] there exist a Hilbert space H , an irreducible representation π : A → B(H ) and a unit vector ξ ∈ H such that for any z ∈ A 14 A. ZAMANI we have ϕ(z) = hπ(z)ξ, ξi. Since c ∈ Z(A), by [4, Proposition II.6.4.13], there exists α ∈ C \ {0} such that π(c) = αe. Now from c ∈ U(A) it follows that kξk=1phπ(c∗c)ξ, ξi = 1. kξk=1(cid:13)(cid:13)π(c)ξ(cid:13)(cid:13) = sup α = kπ(c)k = sup Therefore for any z ∈ A we obtain ϕ(cz) = (cid:12)(cid:12)hπ(cz)ξ, ξi(cid:12)(cid:12) = (cid:12)(cid:12)hπ(c)π(z)ξ, ξi(cid:12)(cid:12) = (cid:12)(cid:12)hαπ(z)ξ, ξi(cid:12)(cid:12) = (cid:12)(cid:12)hπ(z)ξ, ξi(cid:12)(cid:12) = (cid:12)(cid:12)ϕ(z)(cid:12)(cid:12). kξk=1phπ(c)ξ, π(c)ξi = sup From this it follows that v(cz) = sup ϕ∈P(A)(cid:12)(cid:12)ϕ(cz)(cid:12)(cid:12) = sup ϕ∈P(A)(cid:12)(cid:12)ϕ(z)(cid:12)(cid:12) = v(z), and hence v(cz) = v(z). By using a similar argument we conclude that v(z) = v(zc). Now, let x, y ∈ A. Hence xkv y if and only if v(x + λy) = v(x) + v(y) for some λ ∈ T, or equivalently, if and only if v(cid:0)c(x + λy)(cid:1) = v(cx) + v(cy). This holds if and only if v(cx + λcy) = v(cx) + v(cy) for some λ ∈ T, or equivalently, if and only if cxkv cy. Therefore, (i)⇔(ii). The proof of the equivalence (i)⇔(iii) is similar, so we omit it. (cid:3) References 1. A. Abu-Omar and F. Kittaneh, Upper and lower bounds for the numerical radius with an application to involution operators, Rocky Mountain J. Math. 45 (2015), no. 4, 1055 -- 1064. 2. A. Abu-Omar and F. Kittaneh, Notes on some spectral radius and numerical radius in- equalities, Studia Math. 227 (2015), no. 2, 97 -- 109. 3. M. Bakherad and K. Shebrawi, Upper bounds for numerical radius inequalities involving off-diagonal operator matrices, Ann. Funct. Anal. (to appear). 4. B. Blackadar, Operator algebras: theory of C ∗ -- algebras and von Neumann algebras. In: Operator algebras and non-commutative geometry, III. Vol. 122, Encyclopaedia of mathe- matical sciences. Berlin: Springer-Verlag; 2006. 5. FF. Bonsall and J. Duncan, Numerical ranges of operators on normed spaces and elements of normed algebras, Vol. 2, Londonmathematical society lecture note series. London: Cam- bridge University Press, 1971. 6. T. Bottazzi, C. Conde, M.S. Moslehian, P. W´ojcik and A. Zamani, Orthogonality and parallelism of operators on various Banach spaces, J. Aust. Math. Soc. (to appear). 7. J. Dixmier, C ∗-Algebras, North-Holland, Amsterdam, 1981. 8. S. S. Dragomir, A survey of some recent inequalities for the norm and numerical radius of operators in Hilbert spaces, Banach J. Math. Anal. 1 (2007), no. 2, 154 -- 175. 9. M. El-Haddad and F. Kittaneh, Numerical radius inequalities for Hilbert space operators, II. Studia Math. 182 (2007), no. 2, 133 -- 140. 10. P. Grover, Orthogonality of matrices in the Ky Fan k-norms, Linear Multilinear Algebra, 65 (2017), no. 3, 496 -- 509. 11. K. E. Gustafson and D. K. M. Rao, Numerical range. The field of values of linear operators and matrices, Universitext. Springer-Verlag, New York, 1997. 12. O. Hirzallah, F. Kittaneh and K. Shebrawi, Numerical radius inequalities for certain 2 × 2 13. F. Kittaneh, A numerical radius inequality and an estimate for the numerical radius of the operator matrices, Integral Equations Operator Theory 71 (2011), no. 1, 129 -- 147. Frobenius companion matrix, Studia Math. 158 (2003), no. 1, 11 -- 17. 14. F. Kittaneh, Numerical radius inequalities for Hilbert space operators, Studia Math. 168 (2005), no. 1, 73 -- 80. CHARACTERIZATION OF NUMERICAL RADIUS PARALLELISM 15 15. F. Kittaneh, Spectral radius inequalities for Hilbert space operators, Proc. Amer. Math. Soc. 134 (2006), 385 -- 390. 16. F. Kittaneh, M.S. Moslehian and T. Yamazaki, Cartesian decomposition and numerical radius inequalities, Linear Algebra Appl. 471 (2015), 46 -- 53. 17. A. Mal, D. Sain and K. Paul, On some geometric properties of operator spaces, arXiv:1802.06227v1 [math.FA] 17 Feb 2018. block matrices, J. Math. Phys. 57 (2016), no. 1, 015201, 15pp. 18. M.S. Moslehian and M. Sattari, Inequalities for operator space numerical radius of 2 × 2 19. G. J. Murphy, C ∗-Algebras and Operator Theory, Academic Press, New York, 1990. 20. M. Sababheh, Numerical radius inequalities via convexity, Linear Algebra Appl. 549 (2018), 67 -- 78. 21. A. Seddik, Rank one operators and norm of elementary operators, Linear Algebra Appl. 424 (2007), 177 -- 183. Algebra Appl. 523 (2017), 1 -- 12. 22. K. Shebrawi, Numerical radius inequalities for certain 2 × 2 operator matrices II, Linear 23. D. Werner, An elementary approach to the Daugavet equation, Interaction between func- tional analysis, harmonic analysis, and probability (Columbia, MO, 1994), 449 -- 454, Lecture Notes in Pure and Appl. Math., 175, Dekker, New York, 1996. 24. P. W´ojcik, Norm-parallelism in classical M -ideals, Indag. Math. (N.S.) 28 (2017), no. 2, 287 -- 293. 25. T. Yamazaki, On upper and lower bounds of the numerical radius and an equality condition, Studia Math. 178 (2007), no. 1, 83 -- 89. 26. A. Zamani, Some lower bounds for the numerical radius of Hilbert space operators, Adv. Oper. Theory 2 (2017), 98 -- 107. 27. A. Zamani, The operator-valued parallelism, Linear Algebra Appl. 505 (2016), 282 -- 295. 28. A. Zamani and M. S. Moslehian, Exact and approximate operator parallelism, Canad. Math. Bull. 58(1) (2015), 207 -- 224. 29. A. Zamani and M. S. Moslehian, Norm-parallelism in the geometry of Hilbert C ∗-modules, Indag. Math. (N.S.) 27 (2016), no. 1, 266 -- 281. Department of Mathematics, Farhangian University, Iran E-mail address: [email protected]
1802.06917
1
1802
2018-02-20T00:09:13
Relative Weak Injectivity for Operator Systems
[ "math.OA" ]
We investigate the notion of relative weak injectivity and its nuclearity related properties in the category of operator systems. We obtain several characterizations of the weak expectation property. We show that (c,max)-nuclearity characterizes Kirchberg and Wasserman's C*-systems. Namioka and Phelps' test systems, which detects nuclear C*-algebras, is shown to characterize nuclear C*-systems. We study quasi-nuclearity in the operator system setting and prove that quasi-nuclearity and nuclearity are equivalent, in other words, (er,max)-nuclearity and (min,max)-nuclearity are equivalent.
math.OA
math
RELATIVE WEAK INJECTIVITY FOR OPERATOR SYSTEMS ALI S. KAVRUK Abstract. We investigate the notion of relative weak injectivity and its nuclearity related properties in the category of operator systems. We obtain several characterizations of the weak expectation property. We show that (c,max)-nuclearity characterizes Kirchberg and Wasserman's C*-systems. Namioka and Phelps' test systems, which detects nuclear C*- algebras, is shown to characterize nuclear C*-systems. We study quasi-nuclearity in the operator system setting and prove that quasi-nuclearity and nuclearity are equivalent, in other words, (er,max)-nuclearity and (min,max)-nuclearity are equivalent. Weak expectation property (WEP) is a fundamental nuclearity related property introduced by C. Lance [21], [20]. For von Neumann algebras WEP coincides with injectivity, and can be interpreted as a weak form of classical completely positive factorization property. If a C*-algebra A has WEP then A∗∗ is injective relative to A, therefore such objects are also known as weakly injective C*-algebras [18], [25]. WEP is studied in various categories in operator algebras [27], [10], [3], [15], [22]. A more general framework of weak injectivity introduced by E. Kirchberg [17]: for C*-algebras A ⊆ B, A is said to be relatively weakly injective in B, or weakly relatively injective, w.r.i. in short, if the canonical inclusion of A into A∗∗ admits a conditional expectation on B. This is a nuclearity related property in the sense that A is w.r.i. in B if and only if A ⊗max C ⊆ B ⊗max C for every C*-algebra C, in other words, the projective tensor product behaves injectively. A recent work [12] exposes that relative weak injectivity coincides with tight Riesz interpolation property and Riesz-Arveson extension property for C*-algebras. Our main purpose in this paper is to study the notion of relative weak injectivity in the operator system category. As in weak expectation property, weak relative injectivity has two natural extensions to operator systems, namely w.r.i. and r.d.c.i., relative double commutant injectivity. We define w.r.i. for a pair of operator system S1 ⊆ S2 analogously: S1 is said to be w.r.i . in S2 if the canonical inclusion S1 into S ∗∗ 1 extends to a ucp map on S2. A weaker condition is defined as follows: S1 has r.d.c.i . in S2 if for all representation S1 ⊆ B(H), the inclusion i of S1 into B(H) extends to a ucp map i on S2 such a way that i(S2) ⊆ i(S1)′′, the bi-commutant of the image of i. It is worth mentioning that in the definition of r.d.c.i., embedding of S1 into every B(H) is essential, in fact, one can always find an inclusion S1 ⊆ B(H) such that the requirement in the definition holds. W.r.i. and r.d.c.i. are different in operator system category, however, coincide for a pair of C*-algebras. Note: The notion of relative weak injectivity is also studied previously in [2]. We remark that the w.r.i. definition proposed in [2] coincides with r.d.c.i. definition in this article. We first observe that w.r.i. is a nuclearity related property: for S1 ⊆ S2, S1 is w.r.i. in S2 if and only if S1 ⊗max T ⊆ S2 ⊗max T for every operator system T . One of our characterization for w.r.i. involves in extension property into matrix systems: S1 is w.r.i. in S2 if and only if for every matrix system R (i.e. an operator subsystem of a matrix algebra Mn for some n) every ucp map ϕ : S1 → R, has a ucp extension ϕ : S2 → R. This allows 2010 Mathematics Subject Classification. Primary 46L06, 46L07; Secondary 46L05, 47L25, 47L90. Key words. operator system, tensor product, nuclear C*-algebras, injectivity. 1 2 ALI S. KAVRUK us to obtain a tensorial characterization via matrix quotients: S1 is w.r.i. in S2 if and only if S1 ⊗max (Mn/J) ⊆ S2 ⊗max (Mn/J) for every n and null-subspace J ⊂ Mn. As a corollary we obtain that S has the weak expectation property if and only if the minimal and the maximal tensor product coincide on S ⊗ (Mn/J) for every n and null-subspace J ⊂ Mn. As the quotient has the lifting property, we recover the equivalence of (el,max)-nuclearity and weak expectation property. Arveson's extension theorem is one of the most fundamental tool in operator algebras [1]. Combining with a result of Choi and Effros, it states that for operator systems S1 ⊆ S2 and Hilbert space H, every cp map ϕ : S1 → B(H) extends to cp map ϕ : S2 → B(H). There is also a partial order on the self-adjoint completely bounded maps given by ψ ≤ φ if φ − ψ is a cp map. Relative weak injectivity is the key property for Arveson's extension theory which also obeys the order of cp maps: S1 ⊆ S2 is w.r.i. if and only if every state ϕ on S1 has a state extension ϕ on S2 such that every positive linear functional ψ on S1 with ψ ≤ ϕ has a positive extension ψ on S2 with ψ ≤ ϕ. Moreover, this can be achieved such a way the E : [ϕ] → [ ϕ] given by ψ 7→ ψ is a ucp map for which the restriction map is the inverse ucp map. (Here [ϕ] denotes the Effros system associated with ϕ as explained in Section 1.) An operator system S for which the bidual operator system S∗∗ has structure of a C*- algebra is called a C*-system [19]. Nuclear operator systems are such examples, in fact, if S is nuclear then the bidual operator system S ∗∗ is injective, thus has a structure of an injective von Neumann algebra [17]. We prove that being a C*-system is a nuclearity related property and is equivalent to (c,max)-nuclearity in the sense of [15]. In fact, S is a C*-system if and only if S is w.r.i. in its universal C*-algebra C ∗ u(S). In tensor theory of compact convex sets the square is a test object to verify semi-simplexity [23]. Therefore, from the nuclearity viewpoint, Namioka and Phelps' test systems identify the nuclear objects in function systems. Recall that these test systems are defined by W2n = {(ai) ∈ ℓ∞ 2n : a1 + a2 + · · · + an = an+1 + an+2 + · · · a2n} ⊆ ℓ∞ 2n. In [13] we streamline their results in non-commutative setting by showing that W6 detects nuclear C*-algebras. We extend this result to C*-systems: a C*-system S is nuclear if and only if we have a canonical complete order isomorphism S ⊗min W6 = S ⊗max W6. Unfortunately such a property for general operator systems remains open. However we prove that an operator system S is nuclear if and only if S ⊗min R = S ⊗max R for every matrix system R. This, in particular, implies that (er,max)-nuclearity is equivalent to nuclearity, extending C. Lance's notion of quasi-nuclearity to general operator systems. In Section 1 we review basics aspects of operator systems that are required for our work herein. In Section 2 we obtain several equivalent formulations of relative weak injectivity and deduce that C*-systems and (c,max)-nuclear objects coincide. Section 3 includes non- commutative analogue of Namioka and Phelps' theory extended on C*-system and quasi- nuclearity for operator systems. 1. Preliminaries By an operator system S we mean a unital ∗-closed subspace of B(H) together with the induced matricial order structure, where B(H) denotes the von Neumann algebra of bounded linear operators on a Hilbert space H. We refer the reader to [24] for an excellent source of operator systems and their abstract characterizations given by Choi and Effros [4]. A map between operator systems ϕ : S1 → S2 is called completely positive, cp in short, if the nth-amplification idn ⊗ ϕ : Mn ⊗ S1 → Mn ⊗ S2 is positive for all n. If ϕ is also unital, i.e. ϕ(e) = e, we will say that ϕ is a ucp map. We assume familiarity with the universal and RELATIVE WEAK INJECTIVITY FOR OPERATOR SYSTEMS 3 enveloping C*-algebra of an operator system S, denoted by C ∗ map ϕ : S → B(H) u(S) and C ∗ e (S), resp. For a cp [ϕ] = span{ψ : S → B(H) : ψ is cp with ψ ≤ ϕ} ⊆ CB(S, B(H)) is an operator system with unit ϕ, which we call the Effros system associated with ϕ. 1.1. Duality. The topological dual S ∗ of an operator system S can be endowed with a matricial order structure via matricial cp maps. More precisely, after defining the self-adjoint idempotent ∗ via f ∗(s) = f (s∗), we declare (fij) ∈ Mn(S ∗) positive if S ∋ s 7−→ (fij(s)) ∈ Mn is cp. The collection of the cones of the positive elements {Mn(S ∗)+}∞ n=1 forms a strict, compatible matricial order structure on S ∗. In general an Archimedean matrix order may fail to exist for this matricially ordered space. If dim(S) < ∞ then a faithful state w on S can be assigned as an Archimedean order unit for S ∗ [4]. In general, for any state f on S [f ] = span{g : g ≤ f } ⊆ S∗ declares an operator system with unit f , which we call the Effros system as above. 1.2. Quotients. A subspace J ⊂ S is called a kernel if J is kernel of a ucp map defined from S (equivalently kernel of a cp map). A kernel is typically a non-unital ∗-closed subspace but these properties, in general, do not characterize a kernel. A matricial order structure on the algebraic quotient S/J can be defined by Qn = {(sij + J) : (sij) ∈ Mn(S)+)}. The Archimedeanization process, i.e, completion of the cones {Qn} relative to order topology induced by (e + J) ⊗ In (see [26], [16]), yields the operator system quotient S/J. The universal property of the quotient ensures that if ϕ : S → T is a ucp map then the induced map ϕ : S/ker(ϕ) → T is again a ucp map [8]. ϕ is called a quotient (resp., complete quotient) map if ϕ is an order (resp. a complete order) inclusion. In particular a surjective ucp map ϕ is completely quotient if and only if the adjoint ϕ† : T ∗ → S ∗ is a complete order inclusion. A finite dimensional ∗-closed subspace J of an operator system S which does not include any positive elements other that 0 is called a null-subspace. Any null-subspace is a kernel [11]. If R ⊆ Mn is an operator system then R∗ ∼= Mn/J for some null-subspace J ⊂ Mn. To n → R∗ of the inclusion of R into Mn, which is see this one can simply take the adjoint i† : M ∗ ∼= Mn [8] and observing that a quotient map. Now using Farenick and Paulsen's identity M ∗ n the kernel of i† is a null-subspace we get R∗ ∼= Mn/J. Conversely (Mn/J)∗ can be embedded into Mn as an operator subsystem. To see this one can take the adjoint of the quotient map from Mn into Mn/J. We leave the details to the reader. 1.3. Minimal tensor product. For operator systems S and T we define C min n = {[xij ] ∈ Mn(S ⊗ T ) : [(φ ⊗ ψ)(xij )] ≥ 0 for all p, q ∈ N, for all ucp φ : S → Mp and ψ : T → Mq}. n }∞ The collection of cones {C min n=1 forms a strict compatible matricial ordering for the alge- braic tensor S ⊗ T . Moreover, 1S ⊗ 1T is a Archimedean order unit. Therefore the triplet (S ⊗ T , {C min n=1, 1S ⊗ 1T ) forms an operator system which we call the minimal tensor product of S and T and denote by S ⊗min T . We refer the reader to [16] for details. The min- imal tensor product is spatial, injective and functorial. By the representation of the minimal tensor we mean the identification CP(S, T ) ∼= (S ∗ ⊗min T )+ whenever dim(S) < ∞ [11]. n }∞ 4 ALI S. KAVRUK 1.4. Maximal tensor product. Let S and T be two operator systems. We define Dmax n = {X ∗(S ⊗ T )X : S ∈ Mp(S)+, T ∈ Mq(T )+, X is pq × n matrix, p, q ∈ N}. The collection of the cones {Dmax order unit for the matrix ordered space (S ⊗ T , {Dmax Archimedean, which can be resolved by Archimedeanization process. We define }∞ n=1 are strict and compatible. Moreover, 1⊗1 is a matricial }). Nonetheless 1 ⊗ 1 may fail to be n n C max n = {X ∈ Mn(S ⊗ T ) : X + ǫ(1 ⊗ 1)n ∈ Dmax n for all ǫ > 0}. n The collection {C max } forms a strict, compatible matrix ordering on S ⊗ T for which 1 ⊗ 1 is an Archimedean matrix order unit. We let S ⊗maxT denote the resulting tensor product. max is functorial and projective [16], [9]. By the representation of the maximal tensor product we mean the canonical identification CP(S ⊗max T , C) ∼=CP(S, T ∗). 1.5. Commuting tensor product. We construct the commuting tensor product S ⊗c T of two operator systems S and T via ucp maps with commuting ranges, that is, the collection of matricial positive cones are defined by C com n = {X ∈ Mn(S ⊗ T ) : for all Hilbert spaces H and for all ucp maps ϕ : S → B(H), ψ : T → B(H) with commuting ranges we have (ϕ ⊗ ψ)n(X) ≥ 0}. We refer [16] for basic properties of this tensor product and recall that if one of the tensorant has a structure of a C*-algebra then c and max coincide, that is, for an operator system S and a C*-algebra A we have a canonical complete order isomorphism S ⊗c A = S ⊗max A. 1.6. Some asymmetric tensor products. For operator systems S and T we define the left injective and right injective tensor products by the inclusions S ⊗el T :⊆ I(S) ⊗max T and S ⊗er T :⊆ S ⊗max I(T ). The tensor product el is a left injective in the sense that for any operator systems S1 ⊆ S2 and T , we have a canonical complete order embedding S1 ⊗el T ⊆ S2 ⊗el T . Besides, el is maximal left injective tensor product. Analogous properties for for the right injective tensor products hold. We refer [16] for details. 1.7. Nuclearity. For operator system tensor products α and β we write α ≤ β if for every operator systems S and T the canonical map S ⊗β T → S ⊗α T is ucp. For example, the tensor products we discuss above exhibit min ≤ el, er ≤ c ≤ max. An operator system S is said to be (α, β)-nuclear if S ⊗α T = S ⊗β T for every operator system T . Local liftability, weak expectaion, completely positive factorization, exactness are examples of intrinsic char- acterizations of nuclearity related properties. We refer the reader [15] for details and remark that, in this article, we work on (c,max)-nuclearity and (er,max)-nuclearity. An operator system S is said to be nuclear if it is (min,max)-nuclear, i.e., S ⊗min T = S ⊗max T for all T . CP F P exactness DCEP = quasi−nuc. min ≤ el LLP , er ≤ c ≤ max C ∗−syst. C ∗−nuclearity W EP RELATIVE WEAK INJECTIVITY FOR OPERATOR SYSTEMS 5 2. Relative Weak Injectivity As we defined in the introduction, an operator subsystem S1 of an operator system S2 is in S2 if the canonical inclusion of S1 into its bidual operator system S ∗∗ 1 said to be w.r.i. extends to ucp map on S2. i / S ∗∗ 1 8♣ ♣ ♣ ♣ ♣ ♣ ♣ ♣ ♣ ♣ ucp i ♣ ♣ ♣ S1 j S2 An operator system S is said to have the weak expectation property (WEP) if the canonical inclusion of S ֒→ S ∗∗ extends to a ucp map on I(S). Note that this is equivalent to S having w.r.i. in I(S). We start with the following: Theorem 2.1. The following are equivalent for S1 ⊆ S2 : (1) S1 is w.r.i. in S2; (2) for all matrix systems R, every ucp map ϕ : S1 → R has a ucp extension ϕ : S2 → R; (3) for every operator systems T , every ucp map ϕ : S1 → T ∗∗ extends to a ucp map ϕ : S2 → T ∗∗; (4) for every n and null-subspace J ⊂ Mn we have unital order embedding (5) for every operator system T we have a complete order embedding S1 ⊗max (Mn/J) ⊆ S2 ⊗max (Mn/J); S1 ⊗max T ⊆ S2 ⊗max T ; (6) every state ϕ on S1 has a state extension ϕ on S2 such that if ψ is positive linear functional on S1 with ψ ≤ ϕ, then ψ has positive extension ψ on S2 with ψ ≤ ϕ. Moreover, this can be achieved such a way that ψ 7→ ψ is a cp map from [ϕ] to [ ϕ]; (7) S ∗∗ 1 is w.r.i. in S ∗∗ 2 has a ucp inverse. Proof. We shall follow the pattern (1) ⇒ (5) ⇒ (6) ⇒ (4) ⇒ (2) ⇒ (3) ⇒ (1) ⇔ (7). 2 , moreover the inclusion of S ∗∗ 1 into S ∗∗ (1) ⇒ (5): By Lemma 6.5 of [15], for every operator system S and T we have a complete order inclusion S ⊗max T ⊂ S ∗∗ ⊗max T . Now let j be the inclusion of S1 in S2 and let i : S2 → S ∗∗ 1 . Functoriality of the maximal tensor product ensures that 1 be the ucp map extending the canonical inclusion i : S1 ֒→ S ∗∗ S1 ⊗max T j⊗id −−−→ S2 ⊗max T i⊗id−−−→ S ∗∗ 1 ⊗max T are ucp maps. Since the composition is a complete order embedding the first map, j ⊗ id, must have the same property. (5) ⇒ (6): Let ϕ be state on S1 and let Q : S ∗ 1 be the canonical quotient map. By the assumption we have S1 ⊗max [ϕ] ⊆ S2 ⊗max [ϕ]. So we must have that every positive linear functional on S1 ⊗max [ϕ] extends to positive linear functional on S2 ⊗max [ϕ]. By employing the representation of the maximal tensor product, the later condition can be rephrased as follows: every cp map γ : [ϕ] → S ∗ 2 so that Q ◦ γ = γ. In particular if we let γ be the canonical inclusion [ϕ] ⊂ S ∗ 1 the result follows. In fact, for each positive linear functional ψ ≤ ϕ, ψ = γ(ψ) yields the desired map in (6). 1 lifts to a cp map γ : [ϕ] → S ∗ 2 → S ∗ (6) ⇒ (4): Let n be a positive integer and J be a null subspace in Mn. As a first step we will prove that S1 ⊗max (Mn/J) ⊂ S2 ⊗max (Mn/J) order isomorphically. To do so, we need to prove that every positive linear functional f on S1 ⊗max (Mn/J) extends to a positive linear functional on S2 ⊗max (Mn/J). By the representation of the maximal tensor product, _      / 8 6 ALI S. KAVRUK let γf : Mn/J → S ∗ 1 be the cp map associated with f . By rescaling f , if necessary, we may suppose that γf (In + J) is a state in S ∗ 1 , say ϕ. Let ϕ be a state extension of ϕ on S2 as in (6) and E : [ϕ] → [ ϕ] be the cp map. Define γ : Mn/J → S ∗ 2 by γ = E ◦ γf . Let g be the corresponding positive linear functional on S2 ⊗max (Mn/J). It is not difficult to show that g extends f . This proves the desired order inclusion. Now for any operator system S we have Mk(S ⊗max (Mn/J)) ∼= S ⊗max Mk(Mn/J) ∼= S ⊗max (Mk ⊗ Mn)/(Mk ⊗ J). Since Mk ⊗J is a null-subspace of Mk ⊗Mn, it follows that S1 ⊗max (Mn/J) ⊂ S2 ⊗max (Mn/J) holds completely order isomorphically. This finishes the proof. ∼= M ∗ n via eij n/J for some null-subspace J ⊂ M ∗ (4) ⇒ (2): Let R ⊂ Mn be a matrix system and let ϕ : S1 → R be a ucp map. Note that R∗ ∼= M ∗ n. By using the completely positive identification 7→ δij/n, where {eij} is the canonical matrix units in Mn and {δij} Mn is the corresponding dual basis, we may suppose that R∗ is a matrix quotient by a null- subspace. Let fϕ : S1 ⊗max R∗ → C be the associated positive linear functional. Since S1 ⊗max R∗ ⊂ S2 ⊗max R∗ order isomorphically, fϕ extends to a positive linear functional fϕ : S1 ⊗max R∗ → C. Now it is not difficult to show that the corresponding cp map ψ : S2 → R extends ϕ. wot (2) ⇒ (3): Let T be an operator system and ϕ : S1 → T ∗∗ be a ucp map. We can concretely represent T ∗∗ ⊆ B(H) such a way that T ∗∗ ∼= T weak*-wot homeomorphically. Let {Hα} be the net of finite dimensional Hilbert subspaces of H directed under inclusion. We let Pα be the projection onto Hα. Consider the "matrix system" given by Rα = PαT ∗∗Pα ⊂ B(H). We let ϕα : S1 → Rα by s 7→ Pαϕ(s)Pα. By our assumption, ϕα extends to ucp map ϕα : S2 → Rα. (Note that ϕ and ϕ are contractive cp maps when the image is extended to B(H).) Let ϕ be a point-weak cluster point of the net { ϕα}. We let { ϕβ} be the subnet converging ϕ (in point-ultraweak topology). First note that ϕ(s) = ϕ(s) for any s ∈ S1. In fact, the sequence { ϕβ(s)} = {Pβϕ(s)Pβ} converges ϕ(s) in wot and hence in ultraweak topology as it's bounded. Secondly we shall prove that ϕ(S2) ⊂ T . Let us fix x ∈ S2 and set ϕ(x) = y. By definition ϕβ(x) ∈ PβT Pβ, so let tβ ∈ T such that ϕβ(x) = PβtβPβ. Note that the sequence { ϕβ(x)} also converges to y in wot. We claim that tβ converges y in wot. Given h1, h2 ∈ H we fix β0 such that h1, h2 ∈ Hβ0, implying that for any β ≥ β0 Pβhi = hi for i = 1, 2. Now for any such β wot htβh1, h2i − hyh1, h2i = hPβtβPβh1, h2i − hyh1, h2i = h( ϕβ (x) − y)h1, h2i → 0. Since {tβ} ⊂ T wot , y must belong to T wot . This finishes the proof. (3) ⇒ (1): Follows from definition. (1) ⇒ (7): We simply take the second adjoint of the maps appear in the definition of w.r.i.: / S ∗∗ 1 ⇒ S ∗∗ 1 i∗∗ / (S∗∗ 1 )∗∗ P / S ∗∗ 1 i 8♣ ♣ ♣ ♣ ♣ ♣ ♣ ♣ ♣ i ♣ ♣ ♣ ♣ S1 j S2 7♥ ♥ ♥ ♥ j ∗∗ ♥ ♥ ♥ ♥ ♥ i∗∗ ♥ ♥ ♥ ♥ S ∗∗ 2 The canonical inclusion of S ∗∗ 1 in S ∗∗ 1 → S ∗∗ S ∗∗∗∗ (7) ⇒ (1): Let E : S ∗∗ 2 . Consider the canonical inclusion S ∗ in (S ∗∗ 1 )∗∗ coincides with the second adjoint i∗∗, so S ∗∗ 1 is w.r.i. 1 )∗∗. Let P be the adjoint of this inclusion, 1 ֒→ (S∗ 1 . The composition P ◦ i∗∗ gives the desired map from S ∗∗ 2 ֒→ S ∗∗ 2 . Clearly Ej(S1) yields (cid:3) 1 be the ucp inverse of j∗∗ : S ∗∗ 1 2 → S ∗∗ to S ∗∗ 1 . the desired extension in the definition of w.r.i. _      / 8 _      / / / / 7 RELATIVE WEAK INJECTIVITY FOR OPERATOR SYSTEMS 7 The following is a restatement of the above theorem with the pair S ⊆ I(S). The impli- cation (1) ⇒ (3) was already shown in [15] and (3) ⇒ (1) is shown in [9]. Theorem 2.2. The following properties of an operator system S are equivalent: (1) S has the weak expectation property; (2) for every n and null-subspace J ⊂ Mn we have a complete order isomorphism S ⊗min (Mn/J) = S ⊗max (Mn/J); (3) for every operator system T we have a complete order inclusion S ⊗max T ⊆ I(S) ⊗max T , in other words, S is (el,max)-nuclear; (4) S is approximately injective for the matrix systems in the sense that for every n, matrix system R ⊆ Mn, cp map ϕ : R → S and ǫ > 0 there is a cp map ϕ : Mn → S such that k ϕR − ϕkcb ≤ ǫ. Proof. The equivalence of (1) and (3) follows from the previous theorem. Before we proceed we remark that (2) can be rephrased as follows: (2') for any n and null-subspace J ⊂ Mn we have S ⊗max (Mn/J) ⊆ I(S) ⊗max (Mn/J). In fact, the quotient Mn/J has the lifting property so the min and the max tensor products coincide on I(S) ⊗ (Mn/J). Simply replacing max by min, and using the injectivity of min, we get the equivalence of (2) and (2'). Clearly (3) implies (2'). Now we will show that (2') implies (1). Let R ⊆ Mn be a matrix system and let ϕ : S → R be a ucp map. By the representation of the maximal tensor product let fϕ : S ⊗max R∗ → C be the associated positive linear functional. Since R∗ ∼= Mn/J (see Subsection 1.2) fϕ extends to a positive linear functional fϕ : I(S) ⊗max R∗ → C. Now using the representation of max again, we obtain a cp map ϕ : I(S) → R, which extends ϕ. By using the previous theorem we deduce that S is w.r.i. in I(S), so (1) follows. The statement in (4) can be rewritten in the tensor form: (4') S ⊗ Mn → S ⊗min (Mn/J) is a quotient map for any n and null-subspace J ⊂ Mn. To see this let us assume (4'). Let R ⊆ Mn be an operator system and let ϕ : R → S be a cp map. By the representation of the minimal tensor product, let u be the associated positive element in S ⊗min R∗. We declare a faithful state f ∈ R∗ as unit. Now, by the assumption, id⊗Q −−−→ S ⊗min R∗ is a quotient map, so for any ǫ > 0, there exists an element Uǫ ≥ 0 S ⊗ M ∗ n in S ⊗ M ∗ n such that idS ⊗ Q(Uǫ) = u + ǫ(1S ⊗ f ). Uǫ corresponds to cp map ϕǫ : Mn → S. One can verify that ϕǫR − ϕ = ǫf (·)1S . Since cb-norm of a state is 1 we obtain (4). Simply reversing the steps, the reader can verify that (4) implies (4'), hence they are equivalent. We complete our proof by showing (2) and (4') are equivalent. The projectivity of the maximal tensor product [9] ensures that the operator system structure on S ⊗ (Mn/J) arising from the quotient map S ⊗ Mn coincides with the maximal tensor product. Therefore (2) and (4') are equivalent. This finishes our proof. (cid:3) Question: Let J = span{(1, 1, 1, −1, −1, −1)} ⊆ ℓ∞ coincides with the dual operator system W ∗ A is w.r.i. in B if and only if 6 /J 6 [7]. We show in [13] that, for C*-algebras A ⊆ B, 6 . The quotient operator system ℓ∞ A ⊗max (ℓ∞ 6 /J) ⊆ B ⊗max (ℓ∞ 6 /J) . We don't know if such a tensorial inclusion implies w.r.i. for operator systems. In particular, a unital C*-algebra A has WEP if and only if A ⊗min (ℓ∞ 6 /J) [14]. We don't know if a similar property characterizes WEP for general operator system. 6 /J) = A ⊗max (ℓ∞ 8 ALI S. KAVRUK The following is a restatement of Theorem 2.1 for the pair S ⊆ C ∗ u(S). Technically they u(S). The equivalence of (1) and (3) was known to are all equivalent to S being w.r.i. in C ∗ K. H. Han. Theorem 2.3. The following are equivalent for an operator system S: (1) S is (c,max) nuclear, that is, for any operator system T we have S ⊗c T = S ⊗max T ; (2) for every n and null-subspace J ⊂ Mn, we have S ⊗c (Mn/J) = S ⊗max (Mn/J); (3) S is a C*-system, that is, S ∗∗ has structure of a C*-algebra. Proof. The operator system structure on S ⊗ T arising from the inclusion C ∗ u(S) ⊗max T coincides with the commuting tensor product. Therefore the statements (1) and (2) are both equivalent to S being w.r.i. in C ∗ u(S), one can easily verify that every C*-system S is w.r.i. in C ∗ u(S). Therefore (3) implies (1) and (2). u(S)∗∗, with a ucp inverse. Now by a Finally, if S is w.r.i. in C ∗ fundamental result of Choi and Effros (see Theorem 15.2 in [24]) the ucp idempotent from C ∗ (cid:3) u(S)∗∗ onto S ∗∗ declares a C*-algebra structure on S ∗∗, hence S is a C*-system. u(S). By using the universal property of C ∗ u(S) then S ∗∗ is w.r.i. in C ∗ We end this section with a brief summary of relative double commutant injectivity in the operator system category. This notion is introduced in [2] (where the author prefers to use w.r.i. rather than r.d.c.i.). For the completeness of this work we overview the nuclearity related aspects of this property. As we defined at the introduction, an operator subsystem S1 of an operator system S2 is said to have r.d.c.i . in S2 if every representation i : S1 ֒→ B(H) extends to a ucp map i : S2 → B(H) such that i(S2) ⊆ i(S1)′′. We remark that an operator system S has r.d.c.i. in its injective envelope I(S) if and only if it has the double commutant injectivity property (see Section 7 of [15] for this nuclearity related property.) In the following C ∗(F∞) denotes the full group C*-algebra of the free group on countably infinite number of generators. Theorem 2.4 ([2]). The following are equivalent for S1 ⊆ S2 : (1) S1 has r.d.c.i. in S2; (2) for any operator system T we have S1 ⊗c T ⊆ S2 ⊗c T ; (3) for any unital C*-algebra A we have S1 ⊗max A ⊆ S2 ⊗max A; (4) we have S1 ⊗max C ∗(F∞) ⊆ S2 ⊗max C ∗(F∞); (5) C ∗ u(S1) is w.r.i. in C ∗ u(S2). 3. C*-systems and Quasi-nuclearity An operator system S is said to be nuclear if it is (min,max)-nuclear, that is, for every operator system T the minimal and the maximal tensor product on S ⊗ T coincide. A unital C*-algebra A is classically defined as nuclear if A ⊗min B = A ⊗max B for every C*-algebra B. It is elementary to verify that A is a nuclear C*-algebra if and only if A is nuclear as an operator system. In a recent study [12] we extend a classical result of Namioka and Phelps [23] to non-commutative setting by proving that Namioka and Phelps' test system W6 = {(ai)6 i=1 : a1 + a2 + a3 = a4 + a5 + a6} ⊆ ℓ∞ 6 detects nuclear C*-algebras. More precisely, a unital C*-algebra A is nuclear if and only if the minimal and the maximal tensor product on A ⊗ W6 coincide. Unfortunately such a property for general operator systems remains open. Here is an extension to C*-systems: Theorem 3.1. A C*-system S is nuclear if and only if we have a canonical complete order isomorphism S ⊗min W6 = S ⊗max W6. RELATIVE WEAK INJECTIVITY FOR OPERATOR SYSTEMS 9 Proof. We will only prove the non-trivial direction (⇐). Let i be a von Neumann algebra embedding of S ∗∗ into a B(H), and let i0 = iS . Let [i0] ⊆ CB(S, B(H)) be the Effros system. We will first prove that [i0] ∼= i0(S)′ ⊆ B(H) then show that [i0] has WEP. First observe that we have a canonical embedding of i0(S)′ into [i0] given by A 7→ Ai0 (here Ai0 is given by s 7→ Ai0(s)). We also have a natural surjective map R : [i] → [i0] given by the restriction: for a cp map ψ : S∗∗ → B(H) with ψ ≤ i we define R(ψ) = ψS . Surjectivity follows from the fact that whenever ψ ≤ i0 then the weak extension ψ∗∗ : S ∗∗ → B(H) satisfies ψ∗∗ ≤ i and R(ψ∗∗) = ψ. This shows that R is a quotient map. In this particular case, by Radon- Nikodym theory of Arveson [1], [i] coincides with i(S ∗∗)′ = i0(S)′. As R is both injective and projective it must be a bijective complete order isomorphism. This proves that [i0] ∼= i0(S)′, in particular, [i0] has structure of a von Neumann algebra. Next we will show that [i0] has WEP, equivalently is injective. Our assumption is equivalent to the statement that S ⊗max W6 ⊆ S ⊗ ℓ∞ 6 completely order isomorphically. Let γ : W6 → [i0] be a cp map. By the representation of the maximal tensor product we obtain a cp map Γ : S ⊗max W6 → B(H). Γ extends to a cp map Γ from S ⊗ ℓ∞ 6 → [i0]. By [14] 6 it follows that [i0] ∼= i0(S)′ has WEP. A fundamental result of Effros and Lance [6, Prop.3.7], then, implies that i0(S)′′ = i(S ∗∗)′′ = i(S ∗∗) ∼= S ∗∗ is an injective von Neumann algebra. Finally, by Kirchberg's characterization of nuclearity [17], we conclude that S is nuclear. (cid:3) into B(H). Going backward we obtain a cp map γ : ℓ∞ Question: If S ⊗min W6 = S ⊗max W6 can we conclude that S is nuclear? Remark: In the previous theorem we can replace W6 by the 4-dimensional operator system W2,3 = {(a1, ..., a5) : 3(a1 + a2) = 2(a3 + a4 + a5)} ⊆ ℓ∞ 5 . In fact W ∗ 2,3 detects WEP for C*-algebras [14]. A C*-algebra A is nuclear if and only if A ⊗max B1 ⊆ A ⊗max B2 for every C*-algebras B1 ⊆ B2, where that later condition is called quasi-nuclearity (or semi-nuclearity) [20]. In fact if A is nuclear we can replace max by min, so the result simply follows from the injectivity of the minimal tensor product. Conversely, by fixing a representation A ⊆ B(H) with A∗∗ ∼= A , a canonical C*-algebra inclusion A ⊗max A′ ⊆ A ⊗max B(H), if holds, is solely sufficient to conclude that A is nuclear. In fact the canonical ucp map A ⊗max A′ ∋ a ⊗ a′ 7→ aa′ ∈ B(H) extends to a ucp map on A ⊗max B(H), say ϕ. Employing Choi's theory of multiplicative domain, ϕB(H) must have an image sitting in A′. We conclude that A′ is injective, therefore A′′ ∼= A∗∗ is injective, which is sufficient to conclude that A is nuclear [5]. The following result is an operator system analogue of quasi-nuclearity. ′′ Theorem 3.2. The following are equivalent for an operator system S: (1) S is nuclear; (2) S is (er,max)-nuclear; (3) S is quasi-nuclear, that is, for every operator systems T1 ⊆ T2 we have (4) for any matrix system R we have S ⊗min R = S ⊗max R. S ⊗max T1 ⊆ S ⊗max T2; Proof. We first observe that (2) and (3) are equivalent. In fact, given T1 ⊆ T2, the right injectivity of er guarantees that S ⊗er T1 ⊆ S ⊗er T2. If S is (er,max)-nuclear then we can replace er by max, so (3) follows. Conversely, assuming (3), given an operator system T , we have the inclusion S ⊗max T ⊆ S ⊗max I(T ). As the operator system structure on S ⊗ T arising from the inclusion S ⊗max I(T ) is er, we conclude that S is (er,max)-nuclear. 10 ALI S. KAVRUK Clearly (1) implies equivalent conditions (2) and (3). Here we will prove that (3) implies (1): as a first step we show that S is a C*-system. In fact, for an operator system T , our assumption guarantees that we have an inclusion S ⊗max T ⊆ S ⊗max C ∗ u(T ). As the operator system structure on S ⊗T arisen from the inclusion S ⊗max C ∗ u(T ) coincides with the commuting tensor product, we obtain that S ⊗c T = S ⊗max T . Thus, S is (c,max)-nuclear, equivalently, by Theorem 2.3, S is a C*-system. Now, our assumption also yields that we have a complete order inclusion S ⊗max W6 ⊆ S ⊗max ℓ∞ 6 is nuclear, the later inclusion yields that S ⊗min W6 = S ⊗max W6. Hence, by the above theorem, S is nuclear. 6 . As ℓ∞ (1) clearly implies (4). Therefore it suffices to prove that (4) implies (3). Given a matrix system R ⊆ Mn, the condition in (4) is equivalent to the canonical complete order embedding S⊗maxR ⊆ S⊗Mn. Let f be a state on S and let [f ] ⊆ S ∗ be the corresponding Effros system. As a first step we wish to prove that [f ] has structure of an injective von Neumann algebra. Let ϕ : R → [f ](⊆ S ∗) be a cp map. By the representation of the maximal tensor product we obtain a positive linear functional Γϕ : S ⊗max R → C. By our assumption, Γϕ extends to a positive linear functional Γϕ : S ⊗ Mn → C. Employing the representation of the max again, we obtain a cp map ϕ : Mn → S ∗, which extends ϕ. As ϕ(1) = ϕ(1) and [f ] includes all the positive linear functionals dominated by f , we conclude that image of ϕ sits in [f ]. This shows [f ] is approximately injective (actually injective) for the matrix systems, hence, by Theorem 2.2 it has WEP. This means that we must have the canonical complete order embedding [f ] ⊗max S ⊆ I([f ]) ⊗max S. Now consider the identity map i : [f ] → [f ](⊆ S∗), which corresponds to a positive linear functional γ on [f ] ⊗max S. By the previous inclusion, γ extends to a positive linear functional γ on I([f ]) ⊗max S. By using the representation of the max again, we obtain a cp map i : I([f ]) → S ∗, which extends i. Clearly the image of i sits in [f ]. Therefore we obtain a ucp extension i : I([f ]) → [f ] of i. A fundamental result of Choi and Effros ([24, Thm. 15.2]) implies that [f ] must have a structure of a C*-algebra. Finally, we observe that every increasing bounded net of positive elements in [f ] must have a supremum. We leave the details to the reader and conclude that [f ] must have a structure of a von Neumann algebra. As WEP and injectivity coincides for von Neumann algebras we obtain the first part of our proof. Now, let T1 ⊆ T2 be given. Let f : S ⊗max T1 → C be a positive linear functional. It suffices to show that f extends to a positive linear functional f : S ⊗max T2 → C. Let ϕf : T1 → S ∗ be the cp map corresponding f . By rescaling we may suppose that ϕf (1) is a state. Let g denotes this state. Clearly the image of ϕf sits in [g] ⊆ S ∗. As [g] is an injective object, ϕf extends to a cp map ϕf : T2 → [g](⊆ S ∗). Now let f : S ⊗max T2 → C be the associated positive linear functional. It is elementary to verify that f extends f . This shows that we have an order inclusion S ⊗max T1 ⊆ S ⊗max T2. Finally tensoring with Mn one can show that the inclusion holds completely order isomorphically. This finishes our proof. (cid:3) Question: An operator system S is said to be C*-quasi-nuclear if for every A ⊆ B, where A is a C*-subalgebra of the C*-algebra B, we have S ⊗max A ⊆ S ⊗max B. Clearly, C*-nuclearity implies C*-quasi-nuclearity. We don't know if the inverse is true. C*-nuclearity is well known to be equivalent to (min,c)-nuclearity [15]. It can be verified that C*-quasi-nuclearity is equivalent to (er,c)-nuclearity. We end this section with the following trio for an operator system S [17], [15], [9]: S is (min,max)-nuclear ⇐⇒ S ∗∗ is an injective von Neumann algebra; S is (el,max)-nuclear ⇐⇒ S ∗∗ is a von Neumann algebra which is injective relative to S S is (c,max)-nuclear ⇐⇒ S ∗∗ is a von Neumann algebra. (i.e. S has WEP); RELATIVE WEAK INJECTIVITY FOR OPERATOR SYSTEMS 11 References [1] W. B. Arveson, Subalgebras of C*-algebras I, Acta Math. 123 (1969) 141-224. [2] A. Bhattacharya, Relative weak injectivity of operator system pairs, J. Math. Anal. Appl., Vol. 420, Issue 1, (2014) [3] D. P. Blecher, B. L. Duncan, Nuclearity-related properties for nonselfadjoint algebras, J. Op. Theory Vol. 65, Issue 1 (2011) [4] M. Choi, E. Effros, Injectivity and operator spaces, J. Functional Analysis 24 (1977) 156-209. [5] E. Effros, Aspects of non-commutative order, in C*-algebras and applications to physics (Proc. 2nd Japan-USA Seminar, Los Angeles, 1977), eds. [6] E. Effros, C. Lance, Tensor products of operator algebras, Adv. in Math., 25 (1977), pp. 1-34 [7] D. Farenick, A. S. Kavruk, V. I. Paulsen and I. G. Todorov, Operator systems from discrete groups, Comm. Math. Phys., Volume 329, Issue 1 (2014), Page 207-238) [8] D. Farenick, V. I. Paulsen, Operator system quotients of matrix algebras and their tensor products, Mathematica Scandinavica, Vol. 111 (2012) [9] K. H. Han, On maximal tensor products and quotient maps of operator systems, J. Math. Anal. Appl. 384(2):12 (2011) [10] U. Haagerup, Self-polar forms, conditional expectations and the weak expectation property for C-algebras. Unpublished manuscript (1995). [11] A. S. Kavruk, Nuclearity related properties in operator systems. J. Op. Theory, V 71 Issue 1 (2014) [12] A. S. Kavruk, On a non-commutative analogue of a classical result of Namioka and Phelps, J. Funct. Anal. Vol 269, Issue 10, (2015), Pages 3282-3303 [13] A. S. Kavruk, Relative weak injectivity for C*-algebras. preprint (arXiv:1610.08599 ), 2016. [14] A. S. Kavruk, The weak expectation property and Riesz interpolation, preprint (arXiv:1201.5414), 2012. [15] A. S. Kavruk, V. I. Paulsen, I. G. Todorov, and M. Tomforde, Quotients, exactness and nuclearity in the operator system category, Adv. Math. Volume 235, 1 March 2013, Pages 321-360 [16] A. S. Kavruk, V. I. Paulsen, I. G. Todorov, and M. Tomforde, Tensor Products of Operator Systems, J. Funct. Anal., Vol 261, Issue 2, (2011) [17] E. Kirchberg, On non-semisplit extensions, tensor products and exactness of group C*-algebras, Invent. Math. 112 (1993) 449-489. [18] E. Kirchberg, Some properties of QWEP C-algebras, presentation at University of Copenhagen, 2012 [19] E. Kirchberg, S. Wassermann, C*-Algebras Generated by Operator Systems, J. Funct. Anal. V 155, Issue 2, (1998) [20] C. Lance, Tensor products and nuclear C*-algebras, Proceedings of Symposia in Pure Mathematics, Vol. 38 (1982) 379-399. [21] C. Lance, On nuclear C*-algebras, J. Functional Analysis 12(1973), 157-176. [22] J. Liang, Operator-Valued Kirchberg Theory and Its Connection to Tensor Norms and Correspondence, Doctoral Dissertation, University of Illinois Urbana-Champaign (2015) [23] I. Namioka and R. R. Phelps, Tensor products of compact convex sets, Pacific J. Math. Volume 31, Number 2 (1969), 469-480 [24] V. I. Paulsen, Completely bounded maps and operator algebras, Cambridge Studies in Advanced Mathe- matics 78, Cambridge University Press, 2002. [25] V. I. Paulsen, Weak Expectations and the Injective Envelope, Trans. Amer. Math. Soc. V. 363 (2011) [26] V. I. Paulsen and M. Tomforde, Vector spaces with an order unit, Indiana Univ. Math. J., 58 (3) (2007) [27] G. Pisier, Introduction to Operator Space Theory, Cambridge University Press, 2003 Department of Mathematics & Applied Mathematics, Virginia Commonwealth University Richmond, VA 23220, U.S.A. E-mail address: [email protected]
1410.1451
2
1410
2015-02-08T21:33:10
Ergodic theorems in fully symmetric spaces of $\tau-$measurable operators
[ "math.OA", "math.FA" ]
In [11], employing the technique of noncommutative interpolation, a maximal ergodic theorem in noncommutative Lp-spaces, 1 < p < infinity, was established and, among other things, corresponding maximal ergodic inequalities and individual ergodic theorems were derived. In this article, we derive maximal ergodic inequalities in noncommutative Lp-spaces directly from [25] and apply them to prove corresponding individual and Besicovitch weighted ergodic theorems. Then we extend these results to noncommutative fully symmetric Banach spaces with Fatou property and non-trivial Boyd indices, in particular, to noncommutative Lorentz spaces Lpq. Norm convergence of ergodic averages in noncommutative fully symmetric Banach spaces is also studied.
math.OA
math
ERGODIC THEOREMS IN FULLY SYMMETRIC SPACES OF τ−MEASURABLE OPERATORS VLADIMIR CHILIN AND SEMYON LITVINOV Abstract. In [11], employing the technique of noncommutative interpolation, a maximal ergodic theorem in noncommutative Lp −spaces, 1 < p < ∞, was established and, among other things, corresponding maximal ergodic inequal- ities and individual ergodic theorems were derived. In this article, we derive maximal ergodic inequalities in noncommutative Lp −spaces directly from [25] and apply them to prove corresponding individual and Besicovitch weighted ergodic theorems. Then we extend these results to noncommutative fully sym- metric Banach spaces with Fatou property and non-trivial Boyd indices, in particular, to noncommutative Lorentz spaces Lp,q. Norm convergence of ergodic averages in noncommutative fully symmetric Banach spaces is also studied. 1. Preliminaries and introduction Let H be a Hilbert space over C, B(H) the algebra of all bounded linear operators in H, k · k∞ the uniform norm in B(H), I the identity in B(H). If M ⊂ B(H) is a von Neumann algebra, denote by P(M) = {e ∈ M : e = e2 = e∗} the complete lattice of all projections in M. For every e ∈ P(M) we write e⊥ = I − e. If ei(H) is denoted by Vi∈I {ei}i∈I ⊂ P(M), the projection on the subspace Ti∈I ei. A linear operator x : Dx → H, where the domain Dx of x is a linear subspace of H, is said to be affiliated with the algebra M if yx ⊆ xy for every y from the commutant of M. Assume now that M is a semifinite von Neumann algebra equipped with a faith- ful normal semifinite trace τ . A densely-defined closed linear operator x affiliated with M is called τ -measurable if for each ǫ > 0 there exists such e ∈ P(M) with τ (e⊥) ≤ ǫ that e(H) ⊂ Dx. Let us denote by L0(M, τ ) the set of all τ -measurable operators. It is well-known [21] that if x, y ∈ L0(M, τ ), then the operators x + y and xy are densely-defined and preclosed. Moreover, the closures x + y (the strong sum) and xy (the strong product) and x∗ are also τ -measurable and, equipped with these operations, L0(M, τ ) is a unital ∗-algebra over C. For every subset If X ⊂ L0(M, τ ), the set of all self-adjoint operators in X is denoted by X h, whereas the set of all positive operators in X is denoted by X +. The partial order ≤ in Lh 0 (M, τ ) is defined by the cone L+ 0 (M, τ ). Date: February 8, 2015. 2010 Mathematics Subject Classification. 47A35(primary), 46L52(secondary). Key words and phrases. Semifinite von Neumann algebra, maximal ergodic inequality, non- commutative ergodic theorem, bounded Besicovitch sequence, noncommutative fully symmetric space, Boyd indices. 1 2 VLADIMIR CHILIN AND SEMYON LITVINOV The topology defined in L0(M, τ ) by the family V (ǫ, δ) = {x ∈ L0(M, τ ) : kxek∞ ≤ δ for some e ∈ P(M) with τ (e⊥) ≤ ǫ} (cid:0)W (ǫ, δ) = {x ∈ L0(M, τ ) : kexek∞ ≤ δ for some e ∈ P(M) with τ (e⊥) ≤ ǫ}(cid:1) , ǫ > 0, δ > 0, of (closed) neighborhoods of zero is called the measure topology (resp., the bilaterally measure topology). It is said that a sequence {xn} ⊂ L0(M, τ ) converges to x ∈ L0(M, τ ) in measure (bilaterally in measure) if this sequence converges to x in measure topology (resp., in bilaterally measure topology). It is known [3, Theorem 2.2] that xn → x in measure if and only if xn → x bilaterally in measure. For basic properties of the measure topology in L0(M, τ ), see [18]. A sequence {xn} ⊂ L0(M, τ ) is said to converge to x ∈ L0(M, τ ) almost uni- formly (a.u.) (bilaterally almost uniformly (b.a.u.)) if for every ǫ > 0 there exists such e ∈ P(M) that τ (e⊥) ≤ ǫ and k(x − xn)ek∞ → 0 (resp., ke(x − xn)ek∞ → 0). It is clear that every a.u. convergent (b.a.u. convergent) to x sequence in L0(M, τ ) converges to x in measure (resp., bilaterally in measure, hence in measure). For a positive self-adjoint operator x =R ∞ 0 λdeλ affiliated with M one can define τ (x) = sup n τ(cid:18)Z n 0 λdeλ(cid:19) =Z ∞ 0 λdτ (eλ). If 1 ≤ p < ∞, then the noncommutative Lp−space associated with (M, τ ) is defined as Lp = (Lp(M, τ ),k · kp) = {x ∈ L0(M, τ ) : kxkp = (τ (xp))1/p < ∞}, where x = (x∗x)1/2, the absolute value of x (see [24]). Naturally, L∞ = (M,k·k∞). If xn, x ∈ Lp and kx − xnkp → 0, then xn → x in measure [12, Theorem 3.7]. Besides, utilizing the spectral decomposition of x ∈ L+ p , it is possible to find a sequence {xn} ⊂ L+ in particular, kxnkp ≤ kxkp for all n and kx − xnkp → 0. [25]: Let T : L1 ∩ M → L1 ∩ M be a positive linear map that satisfies conditions of p ∩ M such that 0 ≤ xn ≤ x for each n and xn ↑ x; (Y ) T (x) ≤ I and τ (T (x)) ≤ τ (x) ∀ x ∈ L1 ∩ M with 0 ≤ x ≤ I. It is known [25, Proposition 1] that such a T admits a unique positive ultraweakly continuous linear extension T : M → M. In fact, T contracts M: Proposition 1.1. Let T be the extension to M of a positive linear map T : L1 ∩ M → L1 ∩ M satisfying condition (Y). Then kT (x)k∞ ≤ kxk∞ for every x ∈ M. Proof. Since the trace τ is semifinite, there exists a net {pα}α∈Λ ⊂ P(M), where Λ is a base of neighborhoods of zero of the ultraweak topology ordered by inclusion, such that 0 < τ (pα) < ∞ for erery α and pα → I ultraweakly. Then T (xα) → T (I) ultraweakly. Since kT (pα)k∞ ≤ 1, and the unit ball of M is closed in ultraweak topology, we conclude that kT (I)k∞ ≤ 1. Therefore, by [19, Corollary 2.9], kTkM→M = kT (I)k∞ ≤ 1. (cid:3) ERGODIC THEOREMS IN FULLY SYMMETRIC SPACES OF τ −MEASURABLE OPERATORS3 In [11, Theorem 4.1], a maximal ergodic theorem in noncommutative Lp−spaces, 1 < p < ∞, was established for the class of positive linear maps T : M → M satisfying the condition (JX) kT (x)k∞ ≤ kxk∞ ∀ x ∈ M and τ (T (x)) ≤ τ (x) ∀ x ∈ L1 ∩ M+. Remark 1.1. Due to Proposition 1.1, (JX) ⇔ (Y). Besides, by [11, Lemma 1.1], a positive linear map T : M → M that satisfies (JX) uniquely extends to a positive linear contraction T in Lp, 1 < p < ∞. In the sequel, we shall write T ∈ DS+ = DS+(M, τ ) to indicate that the map T : L1 + M → L1 + M is the unique positive linear extension of a positive linear map T : M → M satisfying condition (JX). Such T is often called positive Dunford-Schwartz transformation (see, for example, [26]). Assume that T ∈ DS+ and form its ergodic averages: (1) Mn = Mn(T ) = 1 n + 1 T k, n = 1, 2, ... . nXk=0 The following fundamental result provides a maximal ergodic inequality in L1 for the averages (1). Theorem 1.1. [25] If T ∈ DS+, then for every x ∈ L+ e ∈ P(M) that τ (e⊥) ≤ kxk1 ǫ and sup n keMn(x)ek∞ ≤ ǫ. 1 and ǫ > 0, there is such Here is a corollary of Theorem 1.1, a noncommutative individual ergodic theorem The next result, an extension of Theorem 1.2, was established in [11]. of Yeadon: Theorem 1.2. [25] If T ∈ DS+, then for every x ∈ L1 the averages Mn(x) converge b.a.u. to some bx ∈ L1. Theorem 1.3 ([11], Corollary 6.4). Let T ∈ DS+, 1 < p < ∞, and x ∈ Lp. Then the averages Mn(x) converge b.a.u. to some bx ∈ Lp. If p ≥ 2, these averages converge also a.u. The proof of Theorem 1.3 in [11] is based on an application of a weak type (p, p) maximal inequality for the averages (1), an Lp−version of Theorem 1.1. Note that the proof of this inequality itself relies on Theorem 1.1 and essentialy involves an intricate technique of noncommutative interpolation. Below (Theorem 2.1) we provide a simple, based only on Theorem 1.1, proof of such a maximal inequality. As an application of Theorem 2.1, we prove Besicovitch weighted noncommuative ergodic theorem in Lp, 1 < p < ∞, (Theorem 3.1), which contains Theorem 1.3 as a particular case. Theorem 3.1 is an extension of the corresponding result for L1 in [3]. Note that, in [17], Theorem 1.3 was derived from Theorem 1.1 by utilizing the notion of uniform equicontinuity at zero of a family of additive maps into L0(M, τ ). Lp−spaces with 1 ≤ p < ∞, allows us to establish its validity for a wide class of noncommutative fully symmetric spaces with Fatou property. As a consequence, we obtain an individual ergodic theorem in noncommutative Lorentz spaces Lp,q. Having available Besicovitch weighted ergodic theorem for noncommutative Then we can write (4) ǫ Z ∞ x =Z ǫ λdeλ ≤ ǫ1−pZ ∞ λdeλ +Z ∞ ǫ 0 ǫ λpdeλ ≤ ǫ1−pxp. λdeλ ≤ xǫ + ǫ1−pxp, 4 VLADIMIR CHILIN AND SEMYON LITVINOV The last section of the article is devoted to a study of the mean ergodic ergodic theorem in noncommutative fully symmetric spaces in the case where T ∈ DS(M, τ ). 2. Maximal ergodic inequalities in noncommutative Lp−spaces Everywhere in this section T ∈ DS+. Assume that a sequence of complex numbers {βk}∞ k=0 is such that βk ≤ C for every k. Let us denote (2) Mβ,n = Mβ,n(T ) = 1 n + 1 nXk=0 βkT k. Theorem 2.1. If 1 ≤ p < ∞, then for every x ∈ Lp and ǫ > 0 there is e ∈ P(M) such that (3) τ (e⊥) ≤ 4(cid:18)kxkp ǫ (cid:19)p and sup n keMβ,n(x)ek∞ ≤ 48Cǫ. Proof. Let first βk ≡ 1. In this case, Mβ,n = Mn. Fix ǫ > 0. Assume that x ∈ L+ p , and let x =R ∞ 0 λdeλ be its spectral decomposition. Since λ ≥ ǫ implies λ ≤ ǫ1−pλp, we have where xǫ =R ǫ As xp ∈ L1, Theorem 1.1 entails that there exists e ∈ P(M) satisfying 0 λdeλ. τ (e⊥) ≤ kxpk1 ǫ (cid:19)p ǫp =(cid:18)kxkp and sup n keMn(xp)ek∞ ≤ ǫp. It follows from (4) that 0 ≤ Mn(x) ≤ Mn(xǫ) + ǫ1−pMn(xp) and 0 ≤ eMn(x)e ≤ eMn(xǫ)e + ǫ1−peMn(xp)e for every n. Since xǫ ∈ M, the inequality holds, and we conclude that kT (xǫ)k∞ ≤ kxǫk∞ ≤ ǫ sup n keMn(x)ek∞ ≤ ǫ + ǫ = 2ǫ. If x ∈ Lp, then x = (x1 − x2) + i(x3 − x4), where xj ∈ L+ every j = 1, . . . , 4. As we have shown, there exists ej ∈ P(M) such that (5) n kejMn(xj )ejk∞ ≤ 2ǫ, sup j ) ≤(cid:18)kxjkp ǫ (cid:19)p ≤(cid:18)kxkp ǫ (cid:19)p τ (e⊥ , p and kxjkp ≤ kxkp for j = 1, . . . , 4. ERGODIC THEOREMS IN FULLY SYMMETRIC SPACES OF τ −MEASURABLE OPERATORS5 Now, let {βk}∞ 0 ≤ Imβk + C ≤ 2C, it follows from the decomposition (6) Mβ,n = (Reβk + C)T k + 1 k=0 ⊂ C satisfy βk ≤ C for every k. As 0 ≤ Reβk + C ≤ 2C and nXk=0 nXk=0 (Imβk + C)T k − C(1 + i)Mn n + 1 i n + 1 and (5) that sup n kejMβ,n(xj )ejk∞ ≤ 6C sup n kejMn(xj )ejk∞ ≤ 12Cǫ, j = 1, . . . , 4. Finally, letting e = 4Vj=1 ej, we arrive at (3). (cid:3) Remark 2.1. Note that (5) provides the following extension of the maximal ergodic inequality given in Theorem 1.1 for p = 1: for every x ∈ L+ p and ǫ > 0 there exists e ∈ P(M) such that τ (e⊥) ≤(cid:18)kxkp ǫ (cid:19)p and sup n keMn(x)ek∞ ≤ 2ǫ. To refine Theorem 2.1 when p ≥ 2 we turn to the fundamental result of Kadison [13]: Theorem 2.2 (Kadison's inequality). Let S : M → M be a positive linear map such that S(I) ≤ I. Then S(x)2 ≤ S(x2) for every x ∈ Mh. We will need the following technical lemma; see the proof of [3, Theorem 2.7] or m,n=1 ⊂ L0(M, τ ) be such that for any n the sequence m=1 converges in measure to some an ∈ L0(M, τ ). Then there exists {amkn}∞ k,n=1 [17, Theorem 3.1]. Lemma 2.1. Let {amn}∞ {amn}∞ such that for any n we have amkn → an a.u. as k → ∞. Proposition 2.1 (cf. then for every x ∈ Lh Proof. Let x =R ∞ xm = R m kx2 − x2 [11], proof of Remark 6.5). If 2 ≤ p < ∞ and T ∈ DS+, p and ǫ > 0, there exists e ∈ P(M) such that τ (e⊥) ≤ ǫ and keMn(x)2ek∞ ≤ keMn(x2)ek∞, n = 1, 2, . . . −∞ λdeλ be the spectral decomposition of x ∈ Lh p, and let −m λdeλ. Then, since x ∈ Lp, we clearly have kx − xmkp → 0. Besides, mkp/2 → 0, so kMn(x2) − Mn(x2 m)kp/2 → 0 for every n, which implies that Mn(x2 m) → Mn(x2) in measure, n = 1, 2, . . . Also kMn(x) − Mn(xm)kp → 0 for every n, hence Mn(xm) → Mn(x) in measure and Mn(xm)2 → Mn(x)2 in measure, n = 1, 2, . . . In view of Lemma 2.1, it is possible to find a subsequence {xmk} ⊂ {xm} such that Mn(x2 mk ) → Mn(x2) and Mn(xmk )2 → Mn(x)2 a.u., n = 1, 2, . . . Then one can construct such e ∈ P(M) that τ (e⊥) ≤ ǫ and mk )ek∞ → keMn(x2)ek∞ and keMn(xmk )2ek∞ → keMn(x)2ek∞ keMn(x2 for every n. Since, by Kadison's inequality, we have keMn(xmk )2ek∞ ≤ keMn(x2 mk )ek∞, k, n = 1, 2, . . . , 6 VLADIMIR CHILIN AND SEMYON LITVINOV the result follows. (cid:3) Theorem 2.3. If 2 ≤ p < ∞, then for every x ∈ Lp and ǫ > 0 there is such e ∈ P(M) that (7) n kMβ,n(x)ek∞ ≤ 4√C(2 + √C)ǫ. τ (e⊥) ≤ 6(cid:18)kxkp ǫ (cid:19)p and sup Proof. Pick x ∈ Lh such that (8) τ (e⊥ p . Since x2 ∈ L+ p/2, referring to (5), we can present e1 ∈ P(M) ǫ (cid:19)p ǫ2 (cid:19)p/2 =(cid:18)kxkp n ke1Mn(x2)e1k∞ ≤ 2ǫ2. and sup By Proposition 2.1, there is e2 ∈ P(M) such that 1 ) ≤(cid:18)kx2kp/2 2 ) ≤(cid:18)kxkp ǫ (cid:19)p τ (e⊥ and sup n ke2Mn(x)2e2k∞ ≤ sup n ke2Mn(x2)e2k∞. sup and Then, letting e = e1 ∧ e2, we obtain τ (e⊥) ≤ 2(cid:16) kxkp ǫ (cid:17)p n kMn(x)ek∞ =(cid:18)sup n kMn(x)ek2 n keMn(x2)ek∞(cid:19)1/2 ≤(cid:18)sup n keMn(x)2ek∞(cid:19)1/2 =(cid:18)sup k=0 ⊂ C, βk ≤ C, in accordance with the decomposition (6), we denote β,n = (Reβk + C)T k, M (I) If {βk}∞ M (R) ∞(cid:19)1/2 (Imβk + C)T k. √2ǫ. β,n = ≤ = 1 1 nXk=0 n = 1 nXk=0 n + 1 Let x = x1 + ix2 ∈ Lp, where xj ∈ Lh it follows from (8) that there is f1 ∈ P(M) such that ǫ (cid:19)p 1 ) ≤(cid:18)kx1kp n kf1Mn(x2 and sup τ (f ⊥ 1)f1k∞ ≤ 2ǫ2. p and kxjkp ≤ kxkp, j = 1, 2. Since x2 1 ∈ L+ p/2, Therefore we have sup β,n (x2 n kf1M (R) 1)f1k∞ ≤ 4Cǫ2 β,n : M → M and M (I) Since M (R) ing (2C)−1M (R) inequality, we obtain and sup n kf1M (I) β,n(x2 1)f1k∞ ≤ 4Cǫ2. β,n (I) ≤ I and (2C)−1M (I) β,n : M → M are positive linear maps satisfy- β,n(I) ≤ I for each n, applying Kadison's and This in turn entails β,n (x2 1) β,n (x1)2 ≤ M (R) M (R) β,n(x1)2g12 ≤ M (I) M (I) β,n(x2 1). and sup n kf1M (R) β,n (x1)2f1k∞ ≤ sup n kf1M (R) β,n (x2 1)f1k∞ sup n kf1M (I) β,n(x1)2f1k∞ ≤ sup n kf1M (I) β,n(x2 1)f1k∞. ERGODIC THEOREMS IN FULLY SYMMETRIC SPACES OF τ −MEASURABLE OPERATORS7 Therefore sup n kM (R) β,n (x1)f1k2 ∞ = sup n kf1M (R) β,n (x1)2f1k∞ ≤ sup n kf1M (R) β,n (x2 1)f1k∞ ≤ 4Cǫ2, and similarly sup n kM (I) β,n(x1)f1k2 ∞ ≤ 4Cǫ2. 1 ) ≤ 3(cid:16) kxkp ǫ (cid:17)p and Then, letting g1 = e ∧ f1, we derive τ (g⊥ n kMβ,n(x1)g1k∞ ≤ 2√C(2 + √C)ǫ. sup Similarly, one can find g2 ∈ P(M) with τ (g⊥ 2 ) ≤ 3(kx2kp/ǫ)p such that sup n kMβ,n(x2)g2k∞ ≤ 2√C(2 + √C)ǫ. Finally, we conclude that e = g1 ∧ g2 ∈ P(M) satisfies (7). Remark 2.2. Beginning of the proof of Theorem 2.3 contains the following max- if 2 ≤ p < ∞, given x ∈ Lh imal ergodic inequality for the ergodic averages (1): and ǫ > 0, there exists e ∈ P(M) such that (cid:3) p τ (e⊥) ≤ 2(cid:18)kxkp ǫ (cid:19)p and sup n keMn(x)ek∞ ≤ √2ǫ. 3. Besicovitch weighted ergodic theorem in noncommutative Lp−spaces In this section, using maximal ergodic inequalities given in Theorems 2.1 and 2.3, we prove Besicovitch weighted ergodic theorem in noncommutative Lp−spaces, 1 < p < ∞. As was already mentioned, this extends the corresponding result for p = 1 from [3]. Everywhere in this section T ∈ DS+. We will need the following technical lemma. Lemma 3.1 (see [2], Lemma 1.6). Let X be a linear space, and let Sn : X → L0(M, τ ) be a sequence of additive maps. Assume that x ∈ X is such that for every ǫ > 0 there exists a sequence {xk} ⊂ X and a projection e ∈ P(M) satisfying the following conditions: (i) the sequence {Sn(x + xk)} converges a.u. (b.a.u.) as n → ∞ for each k; (ii) τ (e⊥) ≤ ǫ; (iii) supn kSn(xk)ek∞ → 0 (resp., supn keSn(xk)ek∞ → 0) as k → ∞. Then the sequence {Sn(x)} also converges a.u. (resp., b.a.u.) Using Theorems 2.1 and 2.3, we obtain a corollary: Corollary 3.1. Let 1 ≤ p < ∞ (2 ≤ p < ∞). Then the set {x ∈ Lp : {Mβ,n(x)} converges b.a.u.} (resp., {x ∈ Lp : {Mβ,n(x)} converges a.u.}) is closed in Lp. 8 VLADIMIR CHILIN AND SEMYON LITVINOV Proof. Denote C = {x ∈ Lp : {Mβ,n(x)} converges b.a.u.}. Fix ǫ > 0. Theorem 2.1 implies that for every given k ∈ N there is such γk > 0 that for every x ∈ Lp with kxkp < γk it is possible to find ek,x ∈ P(M) for which τ (e⊥ k,x) ≤ ǫ 2k and sup n kek,xMβ,n(x)ek,xk∞ ≤ 1 k . Pick x ∈ C, the closure of C in Lp. Given k, let yk ∈ C satisfy kyk − xkp < γk. Denoting yk − x = xk, choose a sequence {ek} ⊂ P(M) to be such that , k = 1, 2, ... . τ (e⊥ k ) ≤ ǫ 2k and sup n kekMβ,n(xk)ekk∞ ≤ 1 k Then we have x + xk = yk ∈ C for every k. Also, letting e = Vk≥1 ek, we have τ (e⊥) ≤ ǫ and sup Consequently, Lemma 3.1 yields x ∈ C. maining part of the statement. Analogously, applying Theorem 2.3 instead of Theorem 2.1, we obtain the re- (cid:3) n keMβ,n(xk)ek∞ ≤ 1 k . Corollary 3.1, in the particular case where βk ≡ 1, allows us to present a new, direct proof of Theorem 1.3. Proof. Assume first that p ≥ 2. Since the map T generates a contraction in the real Hilbert space (Lh 2 , (·,·)τ ) [25, Proposition 1], where (x, y)τ = τ (xy), x, y ∈ Lh 2 , it is easy to verify that the set H0 = {x ∈ Lh 2 : T (x) = x} + {x − T (x) : x ∈ Lh 2} is dense in (Lh set Lh 2 ,k · k2) (see, for example [10, Ch.VIII, §5]). Therefore, because the 2 ∩ M is dense in Lh H1 = {x ∈ Lh p and T contracts Lh 2 : T (x) = x} + {x − T (x) : x ∈ L2 p, we conclude that the set h ∩ M} 2 ,k·k2). Besides, if y = x− T (x), x ∈ Lh is also dense in (Lh 2 ∩M, then the sequence Mn(y) = (n + 1)−1(x− T n+1(x)) converges to zero with respect to the norm k·k∞, hence a.u. Therefore H1 + iH1 is a dense in L2 subset on which the averages Mn converge a.u. This, by Corollary 3.1, implies that {Mn(x)} converges a.u. for all x ∈ L2. Further, since the set Lp ∩ L2 is dense in Lp, Corollary 3.1 implies that the sequence {Mn(x)} converges a.u. for each x ∈ Lp (to some bx ∈ L0(M, τ )). Then {Mn(x)} converges to bx in measure. Since Mn(x) ∈ Lp and kMn(x)kp ≤ 1, n = 1, 2, . . . , by Theorem 1.2 in [3], bx ∈ Lp. Let now 1 < p < ∞. By the first part of the proof, the sequence {Mn(x)} converges b.a.u. for all x ∈ L2. But Lp ∩ L2 is dense in Lp, and Corollary 3.1 entails b.a.u. convergence of the averages Mn(x) for all x ∈ Lp. Remembering that b.a.u. convergence implies convergence in measure (see Section 1), we conclude, as before, that Mn(x) → bx ∈ Lp b.a.u.. (cid:3) Let C1 = {z ∈ C : z = 1} be the unit circle in C. A function P : Z → C is said to be a trigonometric polynomial if P (k) = Ps j=1 zjλk j , k ∈ Z, for some 1 ⊂ C1. A sequence {βk}∞ s ∈ N, {zj}s 1 ⊂ C, and {λj}s k=0 ⊂ C is called a bounded Besicovitch sequence if (i) βk ≤ C < ∞ for all k; ERGODIC THEOREMS IN FULLY SYMMETRIC SPACES OF τ −MEASURABLE OPERATORS9 (ii) for every ǫ > 0 there exists a trigonometric polynomial P such that lim sup n 1 n + 1 nXk=0 βk − P (k) < ǫ. that our argument essentially relies on [20, Theorem 1.22.13]. Assume now that M has a separable predual. The reason for this assumption is Since L1 ∩ M ⊂ L2, using Theorem 1.3 for p = 2 (or [5, Theorem 3.1]) and repeating steps of the proof of [3, Lemma 4.2], we arrive at the following. Proposition 3.1. For any trigonometric polynomial P and x ∈ L1 ∩ M, the averages converge a.u. 1 n + 1 nXk=0 P (k)T k(x) Next, it is easy to verify the following (see the proof of [3, Theorem 4.4]). Proposition 3.2. If {βk} is a bounded Besicovitch sequence, then the averages (2) converge a.u. for every x ∈ L1 ∩ M. Here is an extension of [3, Theorem 4.6] to Lp−spaces, 1 < p < ∞. Theorem 3.1. Assume that M has a separable predual. Let 1 < p < ∞, and let {βk} be a bounded Besicovitch sequence. Then for every x ∈ Lp the averages (2) converge b.a.u. to some bx ∈ Lp. If p ≥ 2, these averages converge a.u. set L1 ∩M is dense in Lp. The inclusion bx ∈ Lp follows as in the proof of Theorem Proof. In view of Proposition 3.2 and Corollary 3.1, we only need to recall that the 1.3. (cid:3) 4. Individual ergodic theorems in noncommutative fully symmetric spaces Let x ∈ L0(M, τ ), and let {eλ}λ≥0 be the spectral family of projections for the absolute value x of x. If t > 0, then the t-th generalized singular number of x (see [12]) is defined as A Banach space (E,k·kE) ⊂ L0(M, τ ) is called fully symmetric if the conditions µt(x) = inf{λ > 0 : τ (e⊥ λ ) ≤ t}. sZ µt(y)dt ≤ sZ 0 0 x ∈ E, y ∈ L0(M, τ ), µt(x)dt for all s > 0 imply that y ∈ E and kykE ≤ kxkE. It is known [6] that if (E,k · kE) is a fully symmetric space, xn, x ∈ E, and kx − xnkE → 0, then xn → x in measure. A fully symmetric space (E,k · kE) is said to possess Fatou property if the conditions xα ∈ E+, xα ≤ xβ for α ≤ β, and sup xα ∈ E and kxkE = sup α kxαkE < ∞ imply that there exists x = sup α kxαkE. The space (E,k·kE) α is said to have order continuous norm if kxαkE ↓ 0 whenever xα ∈ E and xα ↓ 0. Let L0(0,∞) be the linear space of all (equivalence classes of) almost every- where finite complex-valued Lebesgue measurable functions on the interval (0,∞). 10 VLADIMIR CHILIN AND SEMYON LITVINOV We identify L∞(0,∞) with the commutative von Neumann algebra acting on the Hilbert space L2(0,∞) via multiplication by the elements from L∞(0,∞) with the trace given by the integration with respect to Lebesgue measure. A Banach space E ⊂ L0(0,∞) is called fully symmetric Banach space on (0,∞) if the condition above holds with respect to the von Neumann algebra L∞(0,∞). Let E = (E(0,∞),k · kE) be a fully symmetric function space. For each s > 0 let Ds : E(0,∞) → E(0,∞) be the bounded linear operator given by Ds(f )(t) = f (t/s), t > 0. The Boyd indices pE and qE are defined as pE = lim s→∞ , qE = lim s→+0 log s log kDskE log s . log kDskE It is known that 1 ≤ pE ≤ qE ≤ ∞ [16, II, Ch.2, Proposition 2.b.2]. A fully symmetric function space is said to have non-trivial Boyd indices if 1 < pE and qE < ∞. For example, the spaces Lp(0,∞), 1 < p < ∞, have non-trivial Boyd indices: pLp(0,∞) = qLp(0,∞) = p [1, Ch.4, §4, Theorem 4.3]. If E(0,∞) is a fully symmetric function space, define E(M) = E(M, τ ) = {x ∈ L0(M, τ ) : µt(x) ∈ E} and set kxkE(M) = kµt(x)kE, x ∈ E(M). It is shown in [6] that (E(M),k · kE(M)) is a fully symmetric space. If 1 ≤ p < ∞ and E = Lp(0,∞), the space (E(M),k·kE(M)) coincides with the noncommutative Lp−space (Lp(M, τ ),k · kp) because kxkp = ∞Z  t (x)dt  = kxkE(M) µp 1/p 0 [24, Proposition 2.4]. It was shown in [4, Proposition 2.2] that if M is non-atomic, then every noncom- mutative fully symmetric (E,k · kE) ⊂ L0(M, τ ) is of the form (E(M),k · kE(M)) for a suitable fully symmetric function space E(0,∞). Let Lp,q(0,∞), 1 ≤ p, q < ∞, be the classical function Lorentz space, that is, the space of all such functions f ∈ L0(0,∞) that ∞Z (t1/pµt(f ))q dt t < ∞. 1/q kfkp,q =  0   It is known that for q ≤ p the space (Lp,q(0,∞),k·kp,q) is a fully symmetric function space with Fatou property and order continuous norm. In addition, Lp,p = Lp. In the case 1 < p < q, the function k · kp,q is a quasi-norm on Lp,q(0,∞), but there exists a norm · (p,q) on Lp,q(0,∞) that is equivalent to the norm k · kp,q and such that (Lp,q(0,∞), · (p,q)) is a fully symmetric function space with Fatou property and order continuous norm [1, Ch.4, §4]. In addition, if 1 ≤ q ≤ p < ∞ (1 < p < ∞, 1 ≤ q < ∞), then p(Lp,q(0,∞),k·kp,q) = q(Lp,q(0,∞),k·kp,q) = p ERGODIC THEOREMS IN FULLY SYMMETRIC SPACES OF τ −MEASURABLE OPERATORS11 [1, Ch.4, §4, Theorem 4.3] (resp., p(Lp,q(0,∞),k·k(p,q)) = q(Lp,q(0,∞),k·k(p,q)) = p [1, Ch.4, §4, Theorem 4.5]). Using function Lorentz space (Lp,q(0,∞),k · kp,q) ((Lp,q(0,∞),k · k(p,q))), one can define noncommutative Lorentz space Lp,q(M, τ ) =  x ∈ L0(M, τ ) : kxkp,q =  ∞Z (t1/pµt(x))q dt t 0 1/q    < ∞  that is fully symmetric with respect to the norm k · kp,q for 1 ≤ q ≤ p (resp., with respect to the norm · (p,q) for q > p > 1). In addition, the norm k · kp,q (resp., · (p,q)) is order continuous [7, Proposition 3.6] and satisfies Fatou property [8, Theorem 4.1]. These spaces were first introduced in the paper [14]. Following [15], a Banach couple (X, Y ) is a pair of Banach spaces, (X,k · kX ) and (Y,k · kY ), which are algebraically and topologically embedded in a Housdorff topological space. With any Banach couple (X, Y ) the following Banach spaces are associated: (i) the space X ∩ Y equipped with the norm kxkX∩Y = max{kxkX,kxkY }, x ∈ X ∩ Y ; (ii) the space X + Y equipped with the norm kxkX+Y = inf{kykX + kzkY : x = y + z, y ∈ X, z ∈ Y }, x ∈ X + Y. Let (X, Y ) be a Banach couple. A linear map T : X + Y → X + Y is called a bounded operator for the couple (X, Y ) if both T : X → X and T : Y → Y are bounded. Denote by B(X, Y ) the linear space of all bounded linear operators for the couple (X, Y ). Equipped with the norm kTkB(X,Y ) = max{kTkX→X,kTkY →Y }, this space is a Banach space. A Banach space Z is said to be intermediate for a Banach couple (X, Y ) if X ∩ Y ⊂ Z ⊂ X + Y with continuous inclusions. If Z is intermediate for a Banach couple (X, Y ), then it is called an interpolation space for (X, Y ) if every bounded linear operator for the couple (X, Y ) acts boundedly from Z to Z. If Z is an interpolation space for a Banach couple (X, Y ), then there exists a constant C > 0 such that kTkZ→Z ≤ CkTkB(X,Y ) for all T ∈ B(X, Y ). An interpolation space Z for a Banach couple (X, Y ) is called an exact interpolation space if kTkZ→Z ≤ kTkB(X,Y ) for all T ∈ B(X, Y ). space for the Banach couple (L1(0,∞), L∞(0,∞)) [15, Ch.II, §4, Theorem 4.3]. Every fully symmetric function space E = E(0,∞) is an exact interpolation We need the following noncommutative interpolation result for the spaces E(M). Theorem 4.1. [6, Theorem 3.4] Let E, E1, E2 be fully symmetric function spaces on (0,∞). Let M be a von Neumann algebra with a faithful semifinite normal trace. If (E1, E2) is a Banach couple and E is an exact interpolation space for (E1, E2), then E(M) is an exact interpolation space for the Banach couple (E1(M), E2(M)). 12 VLADIMIR CHILIN AND SEMYON LITVINOV It follows now from [15, Ch.II, Theorem 4.3] and Theorem 4.1 that every non- commutative fully symmetric space E(M), where E = E(0,∞) is a fully symmetric function space, is an exact interpolation space for the Banach couple (L1(M),M). Let T ∈ DS+(M, τ ). Let E(0,∞) be a fully symmetric function space. Since the noncommutative fully symmetric space E(M) is an exact interpolation space for the Banach couple (L1(M, τ ),M), we conclude that T (E(M)) ⊂ E(M) and T is a positive linear contraction on (E(M),k · kE(M)). Thus T k(x) ∈ E(M) Mn(x) = n + 1 1 nXk=0 for each x ∈ E(M) and all n. Besides, the inequalities kT (x)k1 ≤ kxk1, x ∈ L1, kT (x)k∞ ≤ kxk∞, x ∈ M imply that n≥1kMnkL1→L1 ≤ 1 and sup sup n≥1kMnkM→M ≤ 1. Since the noncommutative fully symmetric space E(M) is an exact interpolation space for the Banach couple (L1(M, τ ),M), we have (9) n≥1kMnkE(M)→E(M) ≤ 1. sup Now, let {βk}∞ k=0 ⊂ C satisfy βk ≤ C, k = 1, 2, . . . . As 0 ≤ Reβk + C ≤ 2C and 0 ≤ Imβk + C ≤ 2C, it follows from (6) that n≥1kMβ,nkL1→L1 ≤ 6C and sup sup n≥1kMβ,nkM→M ≤ 6C. Since the noncommutative fully symmetric space E(M) is an exact interpolation space for the Banach couple (L1(M, τ ),M), we obtain (10) n≥1kMβ,nkE(M)→E(M) ≤ 6C. sup The following theorem is a version of Theorem 1.3 for noncommutative fully symmetric Banach spaces with non-trivial Boyd indices. Theorem 4.2. Let E(0,∞) be a fully symmetric function space with Fatou property and non-trivial Boyd indices. If T ∈ DS+(M, τ ), then for any given x ∈ E(M, τ ) the averages Mn(x) converge b.a.u. to some bx ∈ E(M, τ ). If pE(0,∞) > 2, these averages converge a.u. Proof. Since E(0,∞) has non-trivial Boyd indices, according to [16, II, Ch.2, Propo- sition 2.b.3], there exist such 1 < p, q < ∞ that the space E(0,∞) is intermediate for the Banach couple (Lp(0,∞), Lq(0,∞)). Since (see [6, Proposition 3.1]), we have (Lp + Lq)(M, τ ) = Lp(M, τ ) + Lq(M, τ ) E(M, τ ) ⊂ Lp(M, τ ) + Lq(M, τ ). Then x = x1 + x2, where x1 ∈ Lp(M, τ ), x2 ∈ Lq(M, τ ), and, by Theorem 1.3, there exist such bx1 ∈ Lp(M, τ ) and bx1 ∈ Lq(M, τ ) that Mn(xj) converge b.a.u. to bxj, j = 1, 2. Therefore Mn(x) → bx = bx1 +bx2 ∈ Lp(M, τ ) + Lq(M, τ ) ⊂ L0(M, τ ) ERGODIC THEOREMS IN FULLY SYMMETRIC SPACES OF τ −MEASURABLE OPERATORS13 Following the proof of Theorem 4.2, we obtain its extended version: b.a.u., hence Mn(x) → bx in measure. Since E(M) satisfies Fatou property, the unit ball of E(M) is closed in the measure topology [8, Theorem 4.1], and (9) implies that bx ∈ E(M). If pE(0,∞) > 2, then the numbers p and q can be chosen such that 2 < p, q < ∞. Utilizing Theorem 1.3 and repeating the argument above, we conclude that the averages Mn(x) converge to bx a.u. (cid:3) Theorem 4.3. Let E(0,∞) be a fully symmetric function space with Fatou prop- If T ∈ DS+ and x ∈ E(M, τ ) is such that x = x1 + ··· + xn(x), where erty. xi ∈ Lpj (x)(M, τ ), pj(x) ≥ 1, j = 1, . . . , n(x), then the averages Mn(x) converge b.a.u. to some bx ∈ E(M, τ ). If pj(x) ≥ 2 for all j = 1, . . . , n(x), these averages converge a.u. Since any function Lorentz space E = Lp,q(0,∞) with 1 < p < ∞ and 1 ≤ q < ∞ has non-trivial Boyd indices pE = qE = p, we have the following corollary of Theorem 4.2. Theorem 4.4. Let 1 < p < ∞ and 1 ≤ q < ∞. Then, given x ∈ Lp,q(M, τ ), the averages Mn(x) converge b.a.u. to some bx ∈ Lp,q(M, τ ). If p > 2, these averages converge a.u. Remark 4.1. If 1 ≤ q ≤ p, then Lp,q(M, τ ) ⊂ Lp,p(M, τ ) = Lp(M, τ ) (see [14] and [23, Lemma 1.6]). Then it follows directly from Theorem 1.3 along with the ending of the proof ofthe first part of Theorem 4.2 that for every x ∈ Lp,q(M, τ ) the averages Mn(x) converge to some bx ∈ Lp,q(M, τ ) b.a.u. (a.u. for p ≥ 2). noncommutative fully symmetric space E(M, τ ). Theorem 4.5. Assume that M has a separable predual. Let E(0,∞) be a fully symmetric function space with Fatou property and non-trivial Boyd indices. Let {βk} be a bounded Besicovitch sequence. If T ∈ DS+(M, τ ), then for any given If x ∈ E(M, τ ) the averages Mβ,n(x) converge b.a.u. pE(0,∞) > 2, these averages converge a.u. Proof of Theorem 4.5 uses Theorem 3.1 and the inequality (10) and is analogous to the proof of Theorem 4.2. The next theorem is a version of Besicovitch weighted ergodic theorem for a to some bx ∈ E(M, τ ). Immediately From Theorem 4.5 we obtain the following individual ergodic the- orem for Lorentz spaces Lp,q(M, τ ) (cf. Theorem 4.4). Theorem 4.6. Let M have a separable predual. If 1 < p < ∞ and 1 ≤ q < ∞, then for any x ∈ Lp,q(M, τ ) the averages Mβ,n converge b.a.u. to some bx ∈ Lp,q(M, τ ). If p > 2, these averages converge a.u. Remark 4.2. If 1 ≤ q ≤ p, then Lp,q(M, τ ) ⊂ Lp(M, τ ), and it follows directly from Theorem 3.1 along with the ending of the proof of the first part of Theorem 4.2 that for every x ∈ Lp,q(M, τ ) the averages Mβ,n converge to some bx ∈ Lp,q(M, τ ) b.a.u. (a.u. for p ≥ 2). 5. Mean ergodic theorems in noncommutative fully symmetric spaces Let M be a von Neumann algebra with a faithful normal semifinite trace τ . In [24] the following mean ergodic theorem for noncommutative fully symmetric spaces was proven. 14 VLADIMIR CHILIN AND SEMYON LITVINOV Theorem 5.1. Let E(M) be a noncommutative fully symmetric space such that (i) L1 ∩ M is dense in E(M); (ii) kenkE(M) → 0 for any sequence of projections {en} ⊂ L1 ∩ M with en ↓ 0; (iii) kenkE(M)/τ (en) → 0 for any increasing sequence of projections {en} ⊂ L1 ∩ M with τ (en) → ∞. Then, given x ∈ E(M) and T ∈ DS+(M, τ ), there exists bx ∈ E(M) such that kbx − Mn(x)kE(M) → 0. It is clear that any noncommutative fully symmetric space (E(M),k·kE(M)) with order continuous norm satisfies conditions (i) and (ii) of Theorem 5.1. Besides, in the case of noncommutative Lorentz space Lp,q(M, τ ), the inequality p > 1 together with kekp,q =(cid:18) p q(cid:19)1/q τ (e)1/p, e ∈ L1 ∩ P(M) imply that condition (iii) is also satisfied. Therefore Theorem 5.1 entails the fol- lowing. Corollary 5.1. Let 1 < p < ∞, 1 ≤ q < ∞, T ∈ DS+, and x ∈ Lp,q(M, τ ). Then there exists bx ∈ Lp,q(M, τ ) such that kbx − Mn(x)kp,q → 0. The next theorem asserts convergence in the norm k · kE(M) of the averages Mn(x) for any noncommutative fully symmetric space (E(M),k·kE(M)) with order continuous norm, under the assumption that τ (I) < ∞. Theorem 5.2. Let τ be finite, and let E(M, τ ) be a noncommutative fully sym- metric space with order continuous norm. Then for any x ∈ E(M) and T ∈ DS+ there exists bx ∈ E(M) such that kbx − Mn(x)kE(M) → 0. Proof. Since the trace τ is finite, we have M ⊂ E(M, τ ). As the norm k · kE(M) is order continuous, applying spectral theorem for selfadjoint operators in E(M, τ ), we conclude that M is dense in (E(M, τ ),k · kE(M)). Therefore M+ is a funda- mental subset of (E(M, τ ),k · kE(M)), that is, the linear span of M+ is dense in (E(M, τ ),k · kE(M)). Show that the sequence {Mn(x)} is relatively weakly sequentially compact for every x ∈ M+. Without loss of generality, assume that 0 ≤ x ≤ I. Since T ∈ DS+, we have 0 ≤ Mn(x) ≤ Mn(I) ≤ I for any n. By [9, Proposition 4.3], given y ∈ E+(M, τ ), the set {a ∈ E(M, τ ) : 0 ≤ a ≤ y} is weakly compact in (E(M, τ ),k·kE(M)), which implies that the sequence {Mn(x)} is relatively weakly sequentially compact in (E(M, τ ),k · kE(M)). Since sup n≥1kMnkE(M)→E(M) ≤ 1 (see (9)) and 0 ≤(cid:13)(cid:13)(cid:13)(cid:13) T n(x) n (cid:13)(cid:13)(cid:13)(cid:13)E(M) ≤ kxkE(M) n → 0 whenever x ∈ M+, the result follows by Corollary 3 in [10, Ch.VIII, §5]. (cid:3) Remark 5.1. In the commutative case, Theorem 5.2 was established in [22]. It was also shown that if M = L∞(0, 1), then for every fully symmetric Banach function space E(0, 1) with the norm that is not order continuous there exists such T ∈ DS+ and x ∈ E(M) that the averages Mn(x) do not converge in (E(M),k · kE(M)). ERGODIC THEOREMS IN FULLY SYMMETRIC SPACES OF τ −MEASURABLE OPERATORS15 The following proposition is a version of Theorem 5.1 for noncommutative fully symmetric space with order continuous norm with condition (iii) being replaced by non-triviality of the Boyd indices of E(0,∞). Note that we do not require T to be positive. Proposition 5.1. Let E(0,∞) be a fully symmetric function space with non-trivial Boyd indices and order continuous norm. Then for any x ∈ E(M, τ ) and T ∈ DS(M, τ ) there exists such bx ∈ E(M, τ ) that kbx − Mn(x)kE(M) → 0. Proof. By [16, Theorem 2.b.3], it is possible to find such 1 < p, q < ∞ that Lp(0,∞) ∩ Lq(0,∞) ⊂ E(0,∞) ⊂ Lp(0,∞) + Lq(0,∞) with continuous inclusion maps. In particular, kfkE(0,∞) ≤ CkfkLp(0,∞)∩Lq(0,∞) for all f ∈ Lp(0,∞) ∩ Lq(0,∞) and some C > 0. Hence kxkE(M,τ ) ≤ CkxkLp(M,τ )∩Lq(M,τ ) for all x ∈ L := Lp(M, τ ) ∩ Lq(M, τ ). Therefore the space L is continuously embedded in E(M, τ ). Besides, it follows as in Theorem 5.2 that L is a fundamental subset of (E(M, τ ),k · kE(M)). Show that for everty x ∈ L the sequence {Mn(x)} is relatively weakly sequen- tially compact in (E(M, τ ),k · kE(M)). Since p, q > 1, the spaces Lp(M, τ ) and Lq(M, τ ) are reflexive. As T ∈ DS and x ∈ Lp(M, τ ) ∩ Lq(M, τ ), we conclude that the averages {Mn(x)} converge in (Lp(M, τ ),k · kp) and in (Lq(M, τ ),k · kq) to bx1 ∈ Lp(M, τ ) and to bx2 ∈ Lq(M, τ ), respectively [10, Ch.VIII, §5, Corollary 4]. This implies that the sequence {Mn(x)} converges to bx1 and to bx2 in measure, hence bx1 = bx2 := bx. Since L is continuously embedded in E(M, τ ), the sequence {Mn(x)} converges to bx with respect to the norm k · kE(M), thus, it is relatively weakly sequentially compact in (E(M, τ ),k · kE(M)). Now we can proceed as in the ending of the proof of Theorem 5.2. (cid:3) References [1] C. Bennett, R. Sharpley, Interpolation of Operators, Academic Press Inc. (London) LTD, 1988. [2] V. Chilin, S. Litvinov, Uniform equicontinuity for sequences of homomorphisms into the ring of measurable operators, Methods of Funct. Anal. Top., 12 (2)(2006), 124-130. [3] V. Chilin, S. Litvinov, and A. Skalski, A few remarks in non-commutative ergodic theory, J. Operator Theory, 53 (2)(2005), 331-350. [4] V. I. Chilin, F. A. Sukochev, Weak convergence in non-commutative symmetric spaces, J. Operator Theory, 31 (1994), 35-65. [5] D. C¸ omez, S. Litvinov, Ergodic averages with vector-valued Besicovitch weights, Positivity, 17 (2013), 27-46. [6] P. G. Dodds, T. K. Dodds, and B. Pagter, Fully symmetric operator spaces, Integr. Equat. Oper. Theory, 15 (1992), 942-972. [7] P. G. Dodds, T. K. Dodds, and B. Pagter, Noncommutative Kothe duality, Trans. Amer. Math. Soc., 339(2) (1993), 717-750. [8] P. G. Dodds, T. K. Dodds, F. A. Sukochev, and O. Ye. Tikhonov, A Non-commutative Yoshida-Hewitt theorem and convex sets of measurable operators closed locally in measure, Positivity, 9 (2005), 457-484. [9] P. G. Dodds, B. Pagter and F. A. Sukochev, Sets of uniformly absolutely continuous norm in symmetric spaces of measurable operators, Trans. Amer. Math. Soc., (2014) (to appear). [10] N. Dunford and J. T. Schwartz, Linear Operators, Part I: General Theory, John Willey and Sons, 1988. [11] M. Junge, Q. Xu, Noncommutative maximal ergodic theorems, J. Amer. Math. Soc., 20 (2)(2007), 385-439. 16 VLADIMIR CHILIN AND SEMYON LITVINOV [12] T. Fack, H. Kosaki, Generalized s-numbers of τ -mesaurable operators, Pacific J. Math., 123(1986), 269-300. [13] R. V. Kadison, A generalized Schwarz inequality and algebraic invariants for operator alge- bras, Ann. of Math. (2), 56(1952), 494-503. [14] H. Kosaki, Non-commutative Lorentz spaces associated with a simi-finite von Neumann al- gebra and applications, Proc. Japan DSad., Ser A, 57 (1981), 303-306. [15] S. G. Krein, Ju. I. Petunin, and E. M. Semenov, Interpolation of Linear Operators, Translations of Mathematical Monographs, Amer. Math. Soc., 54, 1982. [16] J. Lindenstraus, L. Tsafriri, Classical Banach spaces I-II, Springer-Verlag, Berlin Heidel- berg New York. 1977. [17] S. Litvinov, Uniform equicontinuity of sequences of measurable operators and non- commutative ergodic theorems, Proc. of Amer. Math. Soc., 140 (2012), 2401-2409. [18] E. Nelson, Notes on non-commutative integration, J. Funct. Anal., 15 (1974), 103-116. [19] V. Paulsen, Completely Bounded Maps and Operator Algebras, Cambridge University Press, 2002. [20] S. Sakai, C ∗ 1971. −algebras and W ∗ −algebras, Springer-Verlag, Berlin Heidelberg New York, [21] I. E. Segal, A non-commutative extension of abstract integration, Ann. of Math., 57 (1953), 401-457. [22] A. Veksler, An ergodic theorem in symmetric spaces, Subirsk. Mat. Zh, 24 (1985), 189-191 (in Russian). [23] H. Yanhou, T. N. Bekjan, The dual on noncommutative Lorentz spaces, Acta Math. Sci., 31 B(5)(2011), 2067-2080. [24] F. J. Yeadon, Non-commutative Lp −spaces, Math. Proc. Camb. Phil. Soc., 77(1975), 91-102. [25] F. J. Yeadon, Ergodic theorems for semifinite von Neumann algebras-I, J. London Math. Soc., 16 (2)(1977), 326-332. [26] F. J. Yeadon, Ergodic theorems for semifinite von Neumann algebras: II, Math. Proc. Camb. Phil. Soc., 88 (1980), 135-147. National University of Uzbekistan, Tashkent, 700174, Uzbekistan E-mail address: [email protected] Pennsylvania State University, Hazleton, PA 18202, USA E-mail address: [email protected]
1905.00978
2
1905
2019-06-19T01:04:18
Radial operators on polyanalytic Bargmann-Segal-Fock spaces
[ "math.OA", "math.FA" ]
The paper considers bounded linear radial operators on the polyanalytic Fock spaces $\mathcal{F}_n$ and on the true-polyanalytic Fock spaces $\mathcal{F}_{(n)}$. The orthonormal basis of normalized complex Hermite polynomials plays a crucial role in this study; it can be obtained by the orthogonalization of monomials in $z$ and $\overline{z}$. First, using this basis, we decompose the von Neumann algebra of radial operators, acting in $\mathcal{F}_n$, into the direct sum of some matrix algebras, i.e. radial operators are represented as matrix sequences. Secondly, we prove that the radial operators, acting in $\mathcal{F}_{(n)}$, are diagonal with respect to the basis of the complex Hermite polynomials belonging to $\mathcal{F}_{(n)}$. We also provide direct proofs of the fundamental properties of $\mathcal{F}_n$ and an explicit description of the C*-algebra generated by Toeplitz operators in $\mathcal{F}_{(n)}$, whose generating symbols are radial, bounded, and have finite limits at infinity.
math.OA
math
Radial operators on polyanalytic Bargmann -- Segal -- Fock spaces Egor A. Maximenko, Ana Mar´ıa Teller´ıa-Romero June 20, 2019 Dedicated to Nikolai L. Vasilevski, our guide in this area of mathematics, on the occasion of his 70th birthday Abstract The paper considers bounded linear radial operators on the polyanalytic Fock spaces Fn and on the true-polyanalytic Fock spaces F(n). The orthonormal basis of normalized complex Hermite polynomials plays a crucial role in this study; it can be obtained by the orthogonalization of monomials in z and z. First, using this basis, we decompose the von Neumann algebra of radial operators, acting in Fn, into the direct sum of some matrix algebras, i.e. radial operators are represented as matrix sequences. Secondly, we prove that the radial operators, acting in F(n), are diagonal with respect to the basis of the complex Hermite polynomials belonging to F(n). We also provide direct proofs of the fundamental properties of Fn and an explicit description of the C*-algebra generated by Toeplitz operators in F(n), whose generating symbols are radial, bounded, and have finite limits at infinity. AMS Subject Classification (2010): Primary 22D25; Secondary 30H20, 47B35. Keywords: radial operator, polyanalytic function, Bargmann -- Segal -- Fock space, von Neumann algebra. 1 Introduction and main results The theory of bounded linear operators in spaces of analytic functions has been intensively developed since the 1980s. In particular, the general theory of operators on the Bargmann- Segal-Fock space (for the sake of brevity, we will say just "Fock space") is explained in the book of Zhu [36]. Nevertheless, the complete understanding of the spectral properties is achieved only for some special classes of operators, in particular, for Toeplitz operators with generating symbols invariant under some group actions, see Vasilevski [34], Grudsky, Quiroga-Barranco, and Vasilevski [11], Dawson, ´Olafsson, and Quiroga-Barranco [8]. The simplest class of this type consists of Toeplitz operators with bounded radial generating symbols. Various properties of these operators (boundedness, compactness, and eigenval- ues) have been studied by many authors, see [13, 20, 24, 37]. The C*-algebra generated by such operators was explicitly described in [12, 32] for the nonweighted Bergman space, 1 in [6, 15] for the weighted Bergman space, and in [10] for the Fock space. Loaiza and Lozano [21, 22] studied radial Toeplitz operators in harmonic Bergman spaces. The spaces of polyanalytic functions, related with Landau levels, have been used in mathematical physics since 1950s; let us just mention a couple of recent papers: [3, 14]. A connection of these spaces with wavelet spaces and signal processing is shown by Abreu [1] and Hutn´ık [16, 17]. Various mathematicians contributed to the rigorous mathematical theory of square-integrable polyanalytic functions. Our research is based on results and ideas from [2, 4, 5, 30, 33]. Hutn´ık, Hutn´ıkov´a, Ram´ırez Ortega, S´anchez-Nungaray, Loaiza, and other authors [18, 19, 23, 26, 29] studied vertical and angular Toeplitz operators in polyanalytic and true- polyanalytic spaces, Bergman and Fock. In particular, vertical Toeplitz operators in the n-analytic Bergman space over the upper half-plane are represented in [26] as n×n matrices whose entries are continuous functions on (0, +∞), with some additional properties at 0 and +∞. Recently, Rozenblum and Vasilevski [27] investigated Toeplitz operators with distri- butional symbols and showed that Toeplitz operators in true-polyanalytic Fock spaces are equivalent to some Toeplitz operators with distributional symbols in the analytic Fock space. In this paper, we analyze radial operators in Fock spaces of polyanalytic or true- polyanalytic functions. We denote by µ the Lebesque measure on the complex plane and by γ the Gaussian measure on the complex plane: dγ(z) = 1 π e−z2 dµ(z). In what follows, we work with the space L2(C, γ) and its subspaces, and denote its norm by k · k. A very useful orthonormal basis in L2(C, γ) is formed by complex Hermite polynomials bj,k, j, k ∈ N0 := {0, 1, 2, . . .}; see Section 2. Given n in N := {1, 2, . . .}, let Fn be the subspace of L2(C, γ) consisting of all n- analytic functions belonging to L2(C, γ). It is known that Fn is a closed subspace of L2(C, γ); moreover, it is a RKHS (reproducing kernel Hilbert space). We denote by F(n) the orthogonal complement of Fn−1 in Fn. z = 1}, let Rn,τ be the rotation operator acting in Fn by the rule (Rn,τ f )(z) := f (τ−1z). For every τ in T := {z ∈ C : The family (Rn,τ )τ∈T is a unitary representation of the group T in the space Fn. We denote by Rn the commutant of {Rn,τ : τ ∈ T} in B(Fn), i.e. the von Neumann algebra that consists of all bounded linear operators acting in Fn that commute with Rn,τ for every τ in T. In other words, the elements of Rn are the operators intertwining the representation (Rn,τ )τ∈T of the group T. The elements of Rn are called radial operators in Fn. In a similar manner, we denote by R(n),τ the rotation operators acting in F(n) and by R(n) the von Neumann algebra of radial operators in F(n). The principal tool in the study of Rn is the following orthogonal decomposition of Fn: (1) Fn = Md=−n+1 Dd,min{n,n+d}. 2 Here the "truncated diagonal subspaces" Dd,m are defined as the linear spans of bj,k with j − k = d and 0 ≤ j, k < m. Another description of Dd,m is given in Proposition 3.7. The main results of this paper are explicit decompositions of the von Neumann algebras Rn and R(n) into direct sums of factors. The symbol ∼= means that the algebras are isometrically isomorphic. Theorem 1.1. Let n ∈ N. Then Rn consists of all operators belonging to B(Fn) that act invariantly on the subspaces Dd,min{n,n+d}, for d ≥ −n + 1. Furthermore, Rn ∼= ∞Md=−n+1 B(Dd,min{n,n+d}) ∼= ∞Md=−n+1 Mmin{n,n+d}. Theorem 1.2. Let n ∈ N. Then R(n) consists of all operators belonging to B(F(n)) that are diagonal with respect to the orthonormal basis (bp,n−1)∞p=0. Furthermore, R(n) ∼= ℓ∞(N0). In particular, Theorems 1.1 and 1.2 imply that the algebra Rn is noncommutative for n ≥ 2, whereas R(n) is commutative for every n in N. In Section 2 we recall the main properties of the complex Hermite polynomials bp,q. In Section 3 we give direct proofs of the principal properties of the spaces Fn and F(n). Section 4 contains some general remarks about unitary representations in RKHS, given by changes of variables. Section 5 deals with radial operators, describes the von Neumann algebra of radial operators in L2(C, γ), and proves Theorems 1.1 and 1.2. Finally, in Sec- tion 6 we make some simple observations about Toeplitz operators generated by bounded radial functions and acting in the spaces Fn and F(n). Another natural method to prove (1) and Theorems 1.1, 1.2 is to represent L2(C, γ) as a tensor product L2(T, dµT )⊗ L2([0, +∞), e−r2 2r dr) and to apply the Fourier transform of the group T. We prefer to work with the canonical basis because this method seems more elementary. Comparing our Theorem 1.1 with the main results of [23,26,29], we would like to point out three differences. 1. We study the von Neumann algebra Rn of all radial operators, instead of C*-algebras generated by Toeplitz operators with radial symbols (such C*-algebras can be objects of study in a future). 2. The dual group of T is the discrete group Z, therefore matrix sequences appear instead of matrix functions. 3. In [23, 26, 29], all matrices have the same order n, whereas in our Theorem 1.1 the matrices have orders 1, 2, . . . , n − 1, n, n, . . .. 2 Complex Hermite polynomials Most results of Sections 2 and 3 are well known to experts [2, 5, 33]. Nevertheless, our proofs are more direct than the ideas found in the bibliography. 3 Given a function f : C → C, continuously differentiable in the R2-sense, we define A†f and A†f by A†f =(cid:18)z − A†f =(cid:18)z − ∂ ∂z(cid:19) f = − ez z ∂ ∂z(cid:19) f = − ez z ∂ ∂ ∂z(cid:0)e−z z f(cid:1) , ∂z(cid:0)e−z z f(cid:1) . The operators A† and A† are known as (nonnormalized) creation operators with respect to z and z, respectively. For every p, q in N0, denote by mp,q the monomial function mp,q(z) := zp zq. Following Shigekawa [30, Section 7] we define the normalized complex Hermite polynomials as bp,q := 1 √p! q! (A†)q(A†)pm0,0 (p, q ∈ N0). (2) Notice that [30] defines complex Hermite polynomials without the factor polynomials appear also in Balk [5, Section 6.3]. Let us show explicitly some of them: 1√p! q! . These b0,0(z) = 1, b0,1(z) = z, b0,2(z) = 1√2 z2, b1,0(z) = z, b2,0(z) = 1√2 z2, b1,1(z) = z2 − 1, b2,1(z) = 1√2 z(z2 − 2), b1,2(z) = 1√2 z(z2 − 2), b2,2(z) = 1 2 (z2 − 4z2 + 2). For every p, α in N0, we denote by L(α) the associated Laguerre polynomial. Recall p the Rodrigues formula, the explicit expression, and the orthogonality relation for these polynomials: L(α) n (x) = L(α) n (x) = n! dn x−α ex dxn(cid:0)e−x xn+α(cid:1), (−1)k(cid:18)n + α n − k(cid:19)xk nXk=0 k! , L(α) n (x)L(α) m (x) xα e−x dx = (n + α)! n! Z +∞ 0 δm,n. (3) (4) (5) (6) Lemma 2.1. Let n, α ∈ N. Then exy ∂n Proof. Apply Rodrigues formula (3) and the chain rule: n (xy). ∂xn(cid:0)e−xy xn+α(cid:1) = n! xαL(α) ∂xn(cid:0)e−xy(xy)n+α(cid:1) = n! e−xy(xy)αL(α) ∂n n (xy) yn. Canceling the factor yn+α in both sides yields (6). 4 Proposition 2.2. For every p, q in N0, In other words, bp,q(z) = bp,q =s min{p, q}! q p q q! p! (−1)q zp−qL(p−q) q p! q! (−1)p zq−pL(q−p) min{p,q}Xs=0 (cid:18)max{p, q} s max{p, q}! (cid:19) ∂zz2 = ∂ Proof. Let p, q ∈ N0, p ≥ q. Notice that ∂ ∂p ∂zp e−zz ezz ∂q ∂zq bp,q(z) = (z2), (z2), if p ≥ q; if p ≤ q. (7) (−1)s (min{p, q} − s)! mp−s,q−s. (8) ∂z (z z) = z. By (2) and (6), (−1)p+q √p! q! (−1)q √p! q! = ezz ∂q ∂zq(cid:0)e−zz zp(cid:1) =s q! p! (−1)qzp−qL(p−q) q (z z). In the case when p ≤ q, we first notice that the operators A† and A† commute on the space of polynomial functions. Reasoning as above, but swapping the roles of z and z, we arrive at the second case of (7). Finally, with the help of (4), we pass from (7) to (8). Formula (8) can also be derived directly from (2), by applying mathematical induction and working with binomial coefficients. Denote by ℓ(α) m the normalized Laguerre function: m (t) :=s m! ℓ(α) (m + α)! tα/2 e−t/2 L(α) m (t) (m, α ∈ N0). Corollary 2.3. For every p, q in N0, bp,q(rτ ) = (−1)min{p,q}τ p−q er2/2 ℓ(p−q) min{p,q} (r2) (r ≥ 0, τ ∈ T). (9) (10) It is convenient to treat the family (mp,q)p,q∈N0 as an infinite table, and to think in terms of its columns or diagonals (parallel to the main diagonal). Given d in Z and n in N0, let Dd,n be the subspace of L2(C, γ) generated by the first n monomials in the diagonal with index d: Dd,n := span{mp,q : p, q ∈ N0, min{p, q} < n, p − q = d}. Proposition 2.4. The family (bp,q)p,q∈N0 is an orthonormal basis of L2(C, γ). This family can be obtained from (mp,q)∞p,q=0 by applying the Gram -- Schmidt orthogonalization. 5 Proof. 1. The orthonormality is easy to verify by passing to polar coordinates and using (7) with the orthogonality relation (5). 2. The formula (8) tells us that the functions bp,q are linear combinations of mp−s,q−s with 0 ≤ s ≤ min{p, q}. Inverting these formulas, mp,q results a linear combination of bp−s,q−s with 0 ≤ s ≤ min{p, q}. So, for every d in Z and every n in N0, Dd,n = span{bp,q : p, q ∈ N0, min{p, q} < n, p − q = d}. (11) Jointly with the orthonormality of (bp,q)∞p,q=0, this means that the family (bp,q)∞p,q=0 is obtained from (mp,q)∞p,q=0 by applying the orthogonalization in each diagonal. 3. Due to 2, it is sufficient to prove that the polynomials in z and z form a dense subset of L2(C, γ). Notice that the set of polynomial functions in z and z coincides with the set of polynomial functions in Re(z) and Im(z). Suppose that f ∈ L2(C, γ) and f is orthogonal to the polynomials Re(z)j Im(z)k for all j, k in N0. Denote by g the function g(x, y) = f (x + i y) e−x2−y2 and consider its Fourier transform: bg(u, v) =ZR2 ∞Xj=0 = e−2π i(xu+yv) f (x + i y) e−x2−y2 dx dy (−2π i u)j(−2π i v)k j! k! ZR2 ∞Xk=0 xjykf (x + i y) e−x2−y2 dx dy = 0. By the injective property of the Fourier transform, we conclude that g vanishes a.e. As a consequence, f also vanishes a.e. Remark 2.5. The second part of the proof of Proposition 2.4 implies that for every d in Z, every q ≥ max{0,−d} every k in Z with max{0,−d} ≤ k ≤ q, hmd+k,k, bd+q,qi =(pq! (d + q)!, k = q; k < q. 0, (12) Formula (11) means that the first n elements in the diagonal d of the table (bp,q)p,q∈N0 generate the same subspace as the first n elements in the diagonal d of the table (mp,q)p,q∈N0. For example, D−1,3 = span{m0,1, m1,2, m2,3} = span{b0,1, b1,2, b2,3}, D2,2 = span{m2,0, m3,1} = span{b2,0, b3,1}. In the following tables we show generators of D2,2 (green) and D−1,3 (blue). b0,4 m0,0 m0,1 m0,2 m0,3 m0,4 b0,1 b0,0 b0,3 b0,2 m1,0 m1,1 m1,2 m1,3 m1,4 m2,0 m2,1 m2,2 m2,3 m2,4 m3,0 m3,1 m3,2 m3,3 m3,4 m4,0 m4,1 m4,2 m4,3 m4,4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . b1,0 b2,0 b3,0 b4,0 . . . b1,1 b2,1 b3,1 b4,1 . . . b1,2 b2,2 b3,2 b4,2 . . . b1,3 b2,3 b3,3 b4,3 . . . b1,4 b2,4 b3,4 b4,4 . . . 6 . . . . . . . . . . . . . . . . . . Given d in Z, we denote by Dd the closure of the subspace of L2(C, γ) generated by the monomials mp,q, where p − q = d: Dd := clos(cid:0)span{mp,q : p, q ∈ N0, p − q = d}(cid:1). Proposition 2.4 implies the following properties of the "diagonal subspaces" Dd, d ∈ Z. Corollary 2.6. The sequence (bq+d,q)∞q=max{0,−d} Corollary 2.7. The space Dd consists of all functions of the form is an orthonormal basis of Dd. f (rτ ) = τ dh(r2) (r ≥ 0, τ ∈ T), where h ∈ L2([0, +∞), e−x dx). (13) Moreover, kfk = khkL2([0,+∞),e−x dx). Corollary 2.8. The space L2(C, γ) is the orthogonal sum of the subspaces Dd: L2(C, γ) =Md∈Z Dd. (14) Here we show the generators of D1 (green) and D−2 (blue): m0,0 m0,1 m0,2 m0,3 m1,0 m1,1 m1,2 m1,3 m2,0 m2,1 m2,2 m2,3 m3,0 m3,1 m3,2 m3,3 . . . . . . . . . . . . . . . . . . . . . . . . . . . b0,0 b1,0 b2,0 b3,0 . . . b0,1 b1,1 b2,1 b3,1 . . . b0,2 b1,2 b2,2 b3,2 . . . b0,3 b1,3 b2,3 b3,3 . . . . . . . . . . . . . . . . . . 3 Bargmann -- Segal -- Fock spaces of polyanalytic functions Fix n in N. Let Fn be the space of n-polyanalytic functions belonging to L2(C, γ), and F(n) be the true-n-polyanalytic Fock space defined in [33] by F(n) := {f ∈ Fn : f ⊥ Fn−1}. Proposition 3.1. Let R > 0. Then there exists a number Cn,R > 0 such that for every f in Fn and every z in C with z ≤ R, f (z) ≤ Cn,Rkfk. Proof. Let Pn be the polynomial in one variable of degree ≤ n − 1 such that Z 1 0 Pn(x)xj dx = δj,0 (j ∈ {0, . . . , n − 1}). 7 (15) (16) The existence and uniqueness of such a polynomial follows from the invertibility of the Hilbert matrix(cid:2)1/(j + k + 1)(cid:3)n−1 Cn,R :=(cid:18) max j,k=0. Put x∈[0,1]Pn(x)(cid:19) 1 πZ(R+1)D ew2 dµ(w)!1/2 . Let f ∈ Fn and z ∈ C, with z ≤ R. It is known [5, Section 1.1] that f can be expanded into a uniformly convergent series of the form f (w) = ∞Xj=0 n−1Xk=0 αj,k(w − z)j(w − z)k, where αj,k are some complex numbers. Using the change of variables w = z + r ei ϑ and the property (16), we obtain the following version of the mean value property of polyanalytic functions: f (z) = 1 πZz+D f (w)Pn(w − z2) dµ(w). (17) After that, estimating Pn by its maximum value, multiplying and dividing by ew2/2, and applying the Schwarz inequality, we arrive at (15). Remark 3.2. The constant Cn,R, found in the proof of Proposition 3.1, is not optimal. The exact upper bound for the evaluation functionals in Fn is given in Corollary 3.16. Proposition 3.3. Fn is a RKHS. Proof. Let (gn)n∈N be a Cauchy sequence in Fn. By Proposition 3.1, this sequence con- verges pointwise on C and uniformly on compacts to a function f . By [5, Corollary 1.8], the function f is n-analytic. On the other hand, let h be the limit of the sequence (gn)n∈N in L2(C, γ). Then for every compact K in C, the sequence of the restrictions gnK con- verges in the L2(K, γ)-norm simultaneously to fK and to hK . Therefore h coincides with f a.e. and f ∈ L2(C, γ), i.e. f ∈ Fn. So, Fn is a Hilbert space. The boundedness of the evaluation functionals is established in Proposition 3.1. Proposition 3.4. The family (bp,q)p∈N0,q<n is an orthonormal basis of Fn. Proof. We already know that this family is contained in Fn and is orthonormal. Let us verify the total property. Our reasoning uses ideas of Ramazanov [25, proof of Theorem 2]. Suppose that f ∈ Fn and hf, bp,qi = 0 for every p ∈ N0, q < n. We have to show that f = 0. By the decomposition of polyanalytic functions [5, Section 1.1], there exists a family of numbers (αj,k)j∈N0,k<n such that f (z) = n−1Xk=0 ∞Xj=0 αj,kmj,k(z), where each of the inner series converges pointwise on C and uniformly on compacts. For j=0 αj,kmj,k. Given r > 0, every ν in N0, we denote by Sν the partial sum Sν :=Pn−1 k=0Pν 8 the sequence (Sν)ν∈N0 converges to f uniformly on rD. For every p, q in N0 with q < n, using the orthogonality on rD between bp,q and mj,k with j − k 6= p − q, we obtain ZrD f bp,q dγ = lim ν→∞ZrD Sν bp,q dγ = n−1Xk=0 αk+p−q,kZrD mk+p−q,kbp,q dγ. The functions f bp,q and mk+p−q,k bp,q are integrable on C with respect to the measure γ. Therefore their integrals over C are the limits of the corresponding integrals over rD, as r tends to infinity. Since hf, bp,qi = 0, the coefficients αj,k must satisfy the following infinite system of homogeneous linear equations: n−1Xk=0 hmk+p−q,k, bp,qiαk+p−q,k = 0 (p ∈ N0, 0 ≤ q < n). (18) Now we fix d > −n and restrict ourselves to the equations (18) with p − q = d, which yields an s × s system represented by the matrix Md, where s = min{n, n + d}, and Md := [hmd+k,k, bd+q,qi]n−1 q,k=max{0,−d} . By (12), Md is an upper triangular matrix with nonzero diagonal entries, hence Md is invertible. So, all coefficients αj,k are zero. Corollary 3.5. F(n) is a RKHS, and the sequence (bp,n−1)p∈N0 is an orthonormal basis of F(n). We denote by Pn and P(n) the orthogonal projections acting in L2(C, γ), whose images are Fn and F(n), respectively. They can be explicitly defined in terms of the corresponding reproducing kernels: (Pnf )(z) = hf, Kn,zi, (P(n)f )(z) = hf, K(n),zi. Corollary 3.6. If f ∈ Fn, then f = ∞Xj=0 n−1Xk=0 hf, bj,kibj,k, where the series converges in the L2(C, γ)-norm and uniformly on the compacts. In par- ticular, if f ∈ F(n), then f = hf, bj,n−1ibj,n−1. ∞Xj=0 (19) For example, (bp,2)p∈N0 is an orthonormal basis of F(3), and (bp,q)p∈N0,q<3 is an or- thonormal basis of F3: b0,1 b1,1 b2,1 b3,1 ... b0,0 b1,0 b2,0 b3,0 ... b0,2 b1,2 b2,2 b3,2 ... b0,3 b1,3 b2,3 b3,3 ... . . . . . . . . . . . . . . . b0,0 b1,0 b2,0 b3,0 ... b0,1 b1,1 b2,1 b3,1 ... b0,2 b1,2 b2,2 b3,2 ... b0,3 b1,3 b2,3 b3,3 ... . . . . . . . . . . . . . . . 9 Using Proposition 3.4, Corollary 2.6, and formula (11) gives Dd ∩ Fn =(Dd,min{n,n+d}, d ≥ −n + 1; d < −n + 1. {0}, (20) Here is a description of the subspaces Dd,m in terms of the polar coordinates. Proposition 3.7. For every m in N0 and every d in Z with d ≥ −m + 1, the space Dd,m consists of all functions of the form f (rτ ) = τ drdQ(r2) (r ≥ 0, τ ∈ T), where Q is a polynomial of degree ≤ m − 1. Moreover, kfk = kQkL2([0,+∞),xd e−x dx). Proof. Apply formula (11) and the orthonormality of the polynomials L(d) L2([0, +∞), xd e−x dx). k in the space The decomposition of Fn into a direct sum of "truncated diagonals" shown below follows from Proposition 3.4 and plays a crucial role in the study of radial operators. Proposition 3.8. Fn = ∞Md=−n+1 Dd,min{n,n+d}. (21) Let us illustrate Proposition 3.8 for n = 3 with a table (we have marked in different shades of blue the basic functions that generate each truncated diagonal): b0,0 b1,0 b2,0 b3,0 ... b0,1 b1,1 b2,1 b3,1 ... b0,2 b1,2 b2,2 b3,2 ... b0,3 b1,3 b2,3 b3,3 ... . . . . . . . . . . . . . . . The upcoming fact was proved by Vasilevski [33]. We obtain it as a corollary from Propo- sition 2.4 and Corollary 3.5. Corollary 3.9. The space L2(C, γ) is the orthogonal sum of the subspaces F(m), m ∈ N: L2(C, γ) = Mm∈N F(m). For every f in F(n), define A†nf by (A†nf )(z) = 1 √n (A†f )(z) = 10 1 √n(cid:18)z − ∂ ∂z(cid:19) f (z). Definition (2) of the family (bp,q)p,q∈N0 implies that A†nbp,n−1 = bp,n. The next picture shows the action of A†2 on basic elements: (22) b0,0 b1,0 b2,0 b3,0 ... b0,1 7→ b0,2 b1,1 7→ b1,2 b2,1 7→ b2,2 b3,1 7→ b3,2 ... ... b0,3 b1,3 b2,3 b3,3 ... . . . . . . . . . . . . . . . Proposition 3.10. A†n is an isometric isomorphism from F(n) onto F(n+1). Proof. Vasilevski [33] proved this fact by using the Fourier transform. Here we give another proof. Write f as in (19). It is known [5, Corollary 1.9] that the derivative ∂ ∂z can be applied to the each term of the series. Therefore A†nf = ∞Xj=0 hf, bj,n−1i A†nbj,n−1 = ∞Xj=0 hf, bj,n−1i bj,n, and kA†nfk = kfk. Also, using the decomposition into series, we see that A†n is surjective. Now we are going to prove explicit formulas (26) and (27) for the reproducing kernels of F(n) and Fn, respectively. These formulas were published by Balk [5, Section 6.3], without using the terminology of Laguerre polynomials, and by Askour, Intissar, and Mouayn [4], though they defined the space F(n) in a different (but equivalent) way. Our proof uses the operators A†n and thereby continues the work of Vasilevski [33]. Lemma 3.11. Let H be a RKHS and (ej)∞j=0 be an ortonormal sequence in H. Then the seriesP∞j=0 ej(z)2 converges. Proof. Denote by KH,z the reproducing kernel of H. From the reproducing property and Bessel's inequality, ∞Xj=0 ej(z)2 = ∞Xj=0 hKH,z, eji2 ≤ kKH,zk2 = KH,z(z) < ∞. Lemma 3.12. For every n in N0 and every z, w in C, K(n+1),z(w) = 1 n(cid:18)z − ∂ ∂z(cid:19)(cid:18)w − ∂ ∂w(cid:19) K(n),z(w). (23) 11 Proof. It is well known that the reproducing kernel of a RKHS H with an orthonormal basis (ej)j∈N0 can be derived from the series KH,z(w) = ∞Xj=0 ej(z)ej(w). (24) In our case, we use the orthonormal basis (bp,n)p∈N0 of the space F(n+1). For a fixed z in C, put αp = bp,n(z). So, K(n+1),z = bp,n(z)bp,n = αpbp,n = ∞Xp=0 ∞Xp=0 ∞Xp=0 αpA†n−1bp,n−1. From Lemma 3.11 we know that (αp)∞p=0 ∈ ℓ2, thus the series P∞p=0 αpbp,n−1 converges in F(n). Since A†n−1 is a bounded operator in F(n), we can interchange it with the sum operator. Therefore K(n+1),z(z) = 1 √n(cid:18)w − ∂ ∂w(cid:19) ∞Xp=0 bp,n(z)bp,n−1(w). +∞. Following the same ideas as above, but swapping the roles of z and w, we factorize Now we fix w in C, write bp,n as A†n−1bp,n−1, and use the fact that P∞p=0 bp,n−1(w)2 < (cid:0)w − ∂ ∂w(cid:1) from the series: K(n+1),z(w) = 1 n(cid:18)z − ∂ ∂z(cid:19)(cid:18)w − ∂ ∂w(cid:19) ∞Xp=0 bp,n−1(z)bp,n−1(w). The last sum equals K(n),z(w), which yields (23). Corollary 3.13. For every n in N0 and every z, w in C, Proposition 3.14. The reproducing kernel of F(n) is given by K(n),z(w) = 1 ∂ ∂z(cid:19)n−1(cid:18)w − (n − 1)!(cid:18)z − K(n),z(w) = ezwLn−1(w − z2). ∂ ∂w(cid:19)n−1 ezw. (25) (26) Proof. Using the definition of creation operators, formula (25) and identity (6) for Laguerre polynomials we have K(n),z(w) = = ezz (n − 1)! ezz (n − 1)! ∂n−1 ∂wn−1 (e−wwewz)(cid:19) ∂ zn−1(cid:18)e−zzeww ∂n−1 ∂ zn−1(cid:16)e−z(z−w)(z − w)n−1(cid:17) ∂n−1 = ezw e(z−w)(z−w) (n − 1)! = ezwLn−1(z − w2). ∂n−1 ∂(z − w)n−1(cid:16)e−(z−w)(z−w)(z − w)n−1(cid:17) 12 Corollary 3.15. The reproducing kernel of Fn is Kn,z(w) = ezwL(1) Proof. Use (26) and the formula L(1) n−1(w − z2). k=0 Lk(x). Corollary 3.16. For every f in Fn and every z in C, 2 kfk. m (x) =Pm−1 f (z) ≤ √n e z2 (27) (28) The equality is achieved when f = Kn,z. Proof. Indeed, kKn,zk2 = Kn,z(z) = ez2 L(1) n−1(0) = n ez2 . We finish this section with a couple of simple results about the Berezin transform and Toeplitz operators in Fn. Given a RKHS H over a domain Ω with a reproducing kernel (Kz)z∈Ω, the corresponding Berezin transform BerH acts from B(H) to the space B(Ω) of bounded functions by the rule BerH(S)(z) = hSKz, KziH hKz, KziH = (SKz)(z) Kz(z) . Stroethoff proved [31] that BerH is injective for various RKHS of analytic functions, in particular, for H = F1. Englis noticed [9, Section 2] that BerH is not injective for various RKHS of harmonic functions. The reasoning of Englis can be applied without any changes to n-analytic functions with n ≥ 2. Proposition 3.17. Let n ≥ 2. Then BerFn is not injective. Proof. Let u and v be some linearly independent elements of Fn such that f , g ∈ Fn. For example, u(z) = b0,0(z) = 1 and v(z) = b1,0(z) = z. Following [9, Section 2], consider S ∈ B(Fn) given by (29) With the help of the reproducing property we easily see that the function BerFn(S) is the zero constant, although the operator S is not zero. Sf := hf, uiv − hf, viu. Given a measure space Ω and a function g in L∞(Ω), we denote by Mg the multipli- cation operator defined on L2(Ω) by Mgf := gf . If H is a closed subspace of L2(Ω), then the Toeplitz operator TH,g is defined on H by TH,g(f ) := PH (gf ) = PHMgf. For H = Fn and H = F(n), we write just Tn,g and T(n),g, respectively. Proposition 3.18. Let g ∈ L∞(C) and Tn,g = 0. Then g = 0 a.e. Proof. For n = 1, this result was proven in [7, Theorem 4]. Let us recall that proof which also works for n ≥ 2. The condition Tn,g = 0 implies that for all j, k in N0 hg, mj,ki =ZC g(z) zjzk dγ(z) = hgmk,0, mj,0i = hTn,gmk,0, mj,0i = 0. Since {mj,k : j, k ∈ N0} is a dense subset of L2(C, γ), g = 0 a.e. 13 4 Unitary representations defined by changes of variables This section states some simple general facts about unitary group representations in RKHS, defined by changes of variables. Suppose that (Ω, ν) is a measure space, H is a RKHS over Ω, with the inner product inherited from L2(Ω), (Kz)z∈Ω is the reproducing kernel of H, and PH ∈ B(L2(Ω)) is the orthogonal projection whose image is H: (PH f )(z) = hf, KziL2(Ω). Furthermore, let G be a locally compact group, and α be a group action in Ω. So, for every τ in G we have a "change of variables" α(τ ) : Ω → Ω, which satisfies α(τ1τ2) = α(τ1)◦α(τ2). Suppose that the function ρ, defined by the following rule, is a strongly continuous unitary representation of the group G in the space L2(Ω): ρ(τ )f := f ◦ α(τ−1) (f ∈ L2(Ω), τ ∈ G). In other words, we suppose that ρ(τ )f ∈ L2(Ω), kρ(τ )fkL2(Ω) = kfkL2(Ω), and ρ(τ )f depends continuously on τ . Proposition 4.1. The following conditions are equivalent. (a) ρ(τ )(H) ⊆ H for every τ in G. (b) ρ(τ )PH = PH ρ(τ ) for every τ in G. (c) The reproducing kernel is invariant under simultaneous changes of variables in both arguments: Kα(τ )(z)(α(τ )(w)) = Kz(w) (τ ∈ G, z, w ∈ Ω). (d) ρ(τ )Kz = Kα(τ )(z) for every z in Ω and every τ in G. Proof. Obviously, (a) is equivalent to (b). Suppose (a) and prove (c): Kα(τ )(z)(α(τ )(w)) = (ρ(τ−1)Kα(τ )(z))(w) = hρ(τ−1)Kα(τ )(z), KwiL2(Ω) = hKα(τ )(z), ρ(τ )KwiL2(Ω) = (ρ(τ )Kw)(α(τ )(z)) = Kw(z) = Kz(w). Suppose (c) and prove (d): (ρ(τ )Kz)(w) = Kz(α(τ−1)(w)) = Kα(τ )(z)(α(τ )(α(τ−1)(w))) = Kα(τ )(z)(w). Suppose (d) and prove (a). Let f ∈ H. Then (ρ(τ )f )(z) = f (α(τ−1)(z)) = hf, Kα(τ −1)(z)iL2(Ω) = hρ(τ )f, ρ(τ )Kα(τ −1)(z)iL2(Ω) = hρ(τ )f, KziL2(Ω). Suppose that the conditions (a) -- (d) of Proposition 4.1 are fulfilled. For every τ in G we denote by ρH(τ ) the compression of the operator ρ(τ ) to the invariant subspace H. Then ρH is a unitary representation of G in H. Let us relate this unitary representation with the Berezin transform of operators. 14 Proposition 4.2. Let S ∈ B(H) and τ ∈ G. Then BerH(ρH (τ−1)SρH (τ ))(z) = BerH(S)(α(τ )(z)) (z ∈ Ω). (30) Proof. BerH(ρH (τ−1)SρH(τ ))(z) = (ρH (τ−1)SρH(τ )Kz)(z) Kz(z) = (SKα(τ )(z)(α(τ )(z)) Kα(τ )(z)(α(τ )(z)) = BerH (S)(α(τ )(z)). Corollary 4.3. Let S ∈ B(H) such that Sρ(τ ) = ρ(τ )S for every τ in G. Then the function BerH(S) is invariant under α, i.e. BerH(S) ◦ α(τ ) = BerH(S) for every τ in G. If BerH is injective, then the inverse of the Corollary 4.3 is also true. The rest of this section does not assume that H has a reproducing kernel; it can be just a closed subspace of L2(Ω). We are going to state some elementary results about the interaction of ρH with Toeplitz operators. These results are well known for many particular cases; see [8, Lemma 3.2 and Corollary 3.3] for the case when H is a Bergman space of analytic functions. Lemma 4.4. Let g ∈ L∞(Ω) and τ ∈ G. Then Mgρ(τ ) = ρ(τ )Mg◦α(τ ). Proof. Put u := g ◦ α(τ ). Given f in L2(Ω), Mgρ(τ )f = (u ◦ α(τ−1)) (f ◦ α(τ−1)) = (uf ) ◦ α(τ−1) = ρ(τ )Muf. Proposition 4.5. Let g ∈ L∞(Ω) and τ ∈ G. Then TgρH(τ ) = ρH(τ )Tg◦α(τ ). (31) Proof. Use Lemma 4.4 and the assumption PH ρ(τ ) = ρ(τ )PH : TgρH(τ )f = PH Mgρ(τ )f = PH ρ(τ )Mg◦α(τ )f = ρ(τ )PH Mg◦α(τ )f = ρH (τ )Tg◦α(τ )f. Corollary 4.6. Let g ∈ L∞(Ω) such that g◦α(τ ) = g for every τ in G. Then Tg commutes with ρH(τ ) for every τ in G. Corollary 4.7. Suppose that the mapping L∞(Ω) → B(H) defined by a 7→ Ta is injective. Let g ∈ L∞(X) such that Tg commutes with ρH(τ ) for every τ in G. Then for every τ in G the functions g ◦ α(τ ) and g coincide a.e. 15 5 Von Neumann algebras of radial operators The methods of this section are similar to ideas from [12,24,37]. We start with two simple general schemes, stated in the context of von Neumann algebras, and then apply them to radial operators in L2(Ω, γ), in Fn, and in F(n). Proposition 5.2 uses the concept of the (bounded) direct sum of von Neumann algebras [28, Definition 1.1.5]. Definition 5.1. Let H be a Hilbert space, U be a self-adjoint subset of B(H), and (Wj)j∈J We say that this family diagonalizes U if the following two conditions are satisfied. be a finite or countable family of nonzero closed subspaces of H such that H =Lj∈J Wj. 1. For each j in J and each U in U , there exists λU,j in C such that Wj ⊆ ker(λU,jI−U ), i.e. U (v) = λU,jv for every v in Wj. 2. For every j, k in J with j 6= k, there exists U in U such that λU,j 6= λU,k. Proposition 5.2. Let H, U , and (Wj)j∈J be like in Definition 5.1. Denote by A the commutant of U . Then A = {S ∈ B(H) : ∀j ∈ J S(Wj) ⊆ Wj}, (32) and A is isometrically isomorphic toLj∈J B(Wj). Proof. 1. Since U is a self-adjoint subset of B(H), its commutant A is a von Neumann algebra [35, Proposition 18.1]. 2. Notice that if U ∈ U and j ∈ J, then λU ∗,j = λU,j. Indeed, for every v in Wj \ {0} λU,jkvk2 H = hλU,jv, viH = hU v, viH = hv, U∗viH = hv, λU ∗,jviH = λU ∗,j kvk2 H . 3. Let S ∈ A, j ∈ J, f ∈ Wj. We are going to prove that Sf ∈ Wj. If k ∈ J \ {j} and g ∈ Wk, then there exists U in U with λU,j 6= λU,k, and λU,jhSf, giH = hSU f, giH = hU Sf, giH = hSf, U∗giH = λU,khSf, giH . which implies that hSf, giH = 0. Since H =Lk∈H Wj, the vector Sf expands into the series of the form Sf =Pq∈J hq with hk ∈ Wk. For every k in J \ {j}, hhq, hkiH = khkk2 H . 0 = hSf, hkiH = hhk, hkiH + Xq∈J\{k} Thus, Sf = hj ∈ Wj. U , j in J, and g in Wj, 4. Now suppose that S ∈ B(H) and S(Wj) ⊆ Wj for every j ∈ J. Then for every U in U Sg = U (Sg) = λU,jSg = S(λU,jg) = SU g. In general, if f in H, then f =Pj∈J gj with some gj in Wj, and SU gj = SU f. U Sf =Xj∈J U Sgj =Xj∈J 16 Given S in A, for every j in J we denote by Aj the compression of S onto the invariant 5. Using (32) we are going to prove that A is isometrically isomorphic to Lj∈J B(Wj). subspace Wj. Then the family (Aj)j∈J belongs toLd∈J B(Wd), and kSk = supj∈J kAjk. Conversely, given a bounded sequence (Aj)j∈J with Aj in B(Wj), we put SXj∈J gj =Xj∈J isomorphisms between A andLj∈J B(Wj). Lj∈J CIWj . Then S(Wj) ⊆ Wj for every j in J, thus S ∈ A. Thereby we have constructed isometrical Proposition 5.2 implies that the von Neumann algebra generated by U consists of all operators that act as scalar operators on each Wj, and can be naturally identified with Ajgj (gj ∈ Wj). Proposition 5.3. Let H, U , and (Wj)j∈J be like in Definition 5.1, and H1 be a closed subspace of H invariant under U . For every U in U , denote by U1 the compression of U onto the invariant subspace H1, and put U1 := {U1 : U ∈ U}, J1 := {j ∈ J : Wj ∩ H1 6= {0}}. Then H1 =Mj∈J1 (Wj ∩ H1), (33) and the family (Wj ∩ H1)j∈J diagonalizes U1. Proof. Denote by P1 the orthogonal projection that acts in H and has image H1. The condition that H1 is invariant under U means that P1 ∈ A. By (32), for every j in J the subspace P1(Wj) is contained in Wj and therefore coincides with Wj ∩ H1. This easily implies (33). If U ∈ U and j ∈ J, then Wj ∩ H1 ⊆ ker(λU,jIH1 − U1). So, the eigenvalues λU1,j If j, k ∈ J1 and j 6= k, then there exists U in U such that λU,j 6= λU,k, which means coincide with λU,j for every j in J1. that λU1,j 6= λU1,k. Radial operators in L2(C, γ) For each τ in T, denote by Rτ the rotation operator acting in L2(C, γ): (34) The family (Rτ )τ∈T is a unitary representation of the group T in L2(C, γ). Notice that we are in the situation of Section 4, with Ω = C, ν = γ, G = T, α(τ )(z) = τ z, ρ(τ ) = Rτ . (Rτ f )(z) = f (τ−1z). Denote by R the set of all radial operators acting in L2(C, γ): R = {S ∈ B(L2(C, γ)) : ∀τ ∈ T Rτ S = SRτ}. Since the set {Rτ : τ ∈ T} is an autoadjoint subset of B(L2(C, γ)), its commutant R is a von Neumann algebra. 17 Lemma 5.4. The family (Dd)d∈Z diagonalizes the collection {Rτ : τ ∈ T} in the sense of Definition 5.1. Proof. If τ ∈ T and d ∈ Z, then Dd ⊆ ker(τ−dI − Rτ ). (35) Indeed, for every p, q ∈ Z with p − q = d the basic function bp,q is an eigenfunction of Rτ associated to the eigenvalue τ−d: Rτ bp,q = τ q−pbp,q = τ−dbp,q, (36) and by Corollary 2.6 the functions bp,q with p − q = d form an orthonormal basis of Dd. Another way to prove (35) is to use Corollary 2.7. If d1, d2 ∈ Z and d1 6= d2, then τ−d1 6= τ−d2 for many values of τ , for example, for d1−d2 or for τ = ei ϑ with any irrational ϑ. τ = e i π Proposition 5.5. The von Neumann algebra R consists of all operators that act invari- antly on Dd for every d in Z, and is isometrically isomorphic toLd∈Z B(Dd). Proof. This is a consequence of Proposition 5.2 and Lemma 5.4. Now we will describe all radial operators of finite rank. Remark 5.6. It is well known that every linear operator of a finite rank m, acting in a Hilbert space H, can be written in the form Sf = mXk=1 ξkhf, ukiH vk, (37) where ξ1, . . . , ξm ∈ C\{0}, u1, . . . , um and v1, . . . , vm are some orthonormal lists of vectors in H. Corollary 5.7. Let m ∈ N and S ∈ B(L2(C, γ)) such that the rank of S is m. Then S is radial if and only if there exist d1, . . . , dm in Z such that S has the form (37), where uj, vj, ξj are like in Remark 5.6, and additionally uj, vj ∈ Ddj for every j in {1, . . . , m}. Proof. This is a simple consequence of Proposition 5.5. Suppose that S is radial. For every d in Z let Ad be the compression of S to Dd. There is only a finite set of d such that Ad 6= 0. Apply Remark 5.6 to each of the nonzero operators Ad and join the obtained decompositions. Following Zorboska [37], we will describe radial operators in term of the "radialization" Rad : B(L2(C, γ)) → B(L2(C, γ)) defined by Rad(S) :=ZT Rτ SRτ −1 dµT(τ ), 18 where µT is the normalized Haar measure on T. The integral is understood in the weak sense, i.e. the operator Rad(S) is actually defined by the equality of the corresponding sesquilinear forms: hRad(S)f, gi =ZThRτ SRτ −1f, gi dµT(τ ). Making an appropriate change of variables in the integral and using the invariance of the measure µT, we see that Rad(S) ∈ R. This immediately implies the following criterion of radial operators in terms of the radialization. Proposition 5.8. Let S ∈ B(L2(C, γ)). Then S ∈ R if and only if Rad(S) = S. Radial operators in Fn Let n ∈ N. Obviously, the reproducing kernel of Fn, given by (27), is invariant under simultaneous rotations in both arguments: Kn,τ z(τ w) = Kn,z(w) (z, w ∈ C, τ ∈ T). (38) Therefore, by Proposition 4.1, Fn is invariant under rotations, and Pn ∈ R. For every τ in T, we denote by Rn,τ the compression of Rτ onto the space Fn. In other words, the operator Rn,τ acts in Fn and is defined by (34). The family (Rn,τ )τ∈T is a unitary representation of T in Fn. Let Rn be the von Neumann algebra of all bounded linear radial operators acting in Fn. Denote by Mn the following direct sum of matrix algebras: Mn := ∞Md=−n+1 Mmin{n,n+d} = −1Md=−n+1 Mn+d! ⊕ ∞Md=0 Mn! . The elements of Mn are matrix sequences of the form A = (Ad)∞d=−n+1, where Ad ∈ Mn+d if d < 0, Ad ∈ Mn if d ≥ 0, and Now we are ready to prove Theorem 1.1. sup d≥−n+1kAdk < +∞. Proof of Theorem 1.1. By Propositions 5.2, 5.3 and formula (20), Rn is isometrically iso- morphic to the direct sum of B(Dd,min{n,n+d}), with d ≥ −n + 1. Using the orthonormal basis (bd+k,k)n−1 of the space Dd,min{n,n+d}, we represent linear operators on this space as matrices. Define Φn : Rn → Mn by k=max{0,−d} Φn(S) =(cid:16)[hSbd+k,k, bd+j,ji]n−1 j,k=max{0,−d}(cid:17)∞ . (39) d=−n+1 Then Φn is an isometrical isomorphism. 19 Similarly to Corollary 5.7, there is a simple description of radial operators of finite rank acting in Fn. Of course, now d1, . . . , dm ≥ −n + 1. By Corollary 4.3, if S ∈ Rn, then BerFn(S) is a radial function. For n = 1, the Berezin transform BerF1 is injective. So, if S ∈ B(F1) and the function BerF1(S) is radial, then S ∈ R1. For n ≥ 2, there are nonradial operators S with radial Berezin transforms. Example 5.9. Let n ≥ 2. Define u, v, and S like in the proof of Proposition 3.17. Then Ber(S) is the zero constant. In particular, Ber(S) is a radial function. On the other hand, Sb0,0 = b1,0, the subspace D0 is not invariant under S, and thus S is not radial. Radial operators in F(n) Let n ∈ N. By Proposition 4.1 and formula (26), the subspace F(n) is invariant under the rotations Rτ for all τ in T. Denote the corresponding compression of Rτ by R(n),τ . Let R(n) be the von Neumann algebra of all radial operators in F(n). Proof of Theorem 1.2. Corollaries 2.6 and 3.5 give Dd ∩ F(n) =(Cbd+n−1,n−1, d ≥ −n + 1, d < −n + 1. {0}, (40) By Propositions 5.2, 5.3 and formula (40), R(n) consists of the operators that act invari- antly on Cbd+n−1,n−1, d ≥ −n + 1, i.e. are diagonal with respect to the basis (bp,n−1)∞p=0. Therefore the function Φ(n) : R(n) → ℓ∞(N0), defined by Φ(n)(S) =(cid:0)hSbp,n−1, bp,n−1i(cid:1)∞ is an isometric isomorphism. p=0, (41) Similarly to Corollary 5.7, there is a simple description of radial operators of finite rank acting in F(n). 6 Radial Toeplitz operators in polyanalytic spaces A measurable function g : C → C is called radial if for every τ in T the equality g(τ z) = g(z) is true for a.e. z in C. If g ∈ L2(C, γ), then this condition means that Rτ g = g for every τ in T. It is easy to see that a function g in L∞(C) is radial if and only if there exists a in Given a function a in L∞([0, +∞)), letea be its extension defined on C as L∞([0, +∞)) such that g =ea. compute the matrix of this operator with respect to the basis (bp,q)p,q∈N0. Put ea(z) := a(z) (z ∈ C). By Lemma 4.4, the multiplication operator Mea, acting in L2(C, γ), is radial. Let us βa,d,j,k := heabj+d,j, bk+d,ki (d ∈ Z, j, k, j + d, k + d ∈ N0). 20 (42) Passing to the polar coordinates and using (10) we get Proposition 6.1. Let a ∈ L∞([0, +∞)). Then Mea ∈ R, and 0 (t)ℓ(d) min{j,j+d} a(√t) ℓ(d) βa,d,j,k =Z +∞ hMeabp,q, bj,ki = heabp,q, bj,ki = δp−q,j−kβa,p−q,q,k. min{k,k+d} (t) dt. Proof. Use the fact that Mea is radial and the orthogonality of the "diagonal subspaces". Then apply the definition of βa,d,j,k. Proposition 6.2. Let g ∈ L∞(C). Then the opeator Tn,g is radial if and only if the function g is radial. Proof. Apply Proposition 3.18 and Corollaries 4.6, 4.7. Proposition 6.3. Let a ∈ L∞([0, +∞)). Then T(n),ea ∈ R(n), the operator T(n),ea is diagonal with respect to the orthonormal basis (bp,n−1)∞p=0, and the sequence λa,n of the corresponding eigenvalues can be computed by λa,n(p) = βa,p−n+1,n−1,n−1 =Z +∞ 0 a(√t)(cid:0)ℓ(p−n+1) min{p,n−1} (t)(cid:1)2 dt (p ∈ N0). (43) Proof. From Corollary 4.6 we get T(n),ea ∈ R(n). Due to Proposition 6.1 and Theorem 1.2, λa,n(p) = (Φ(n)(T(n),ea))p = hT(n),eabp,n−1, bp,n−1i = βa,p−n+1,n−1,n−1. Given a class G ⊆ L∞(C) of generating symbols, we denote by T(n)(G) the C*- subalgebra of B(F(n)) generated by the set {T(n),g : g ∈ G}. Let RB be the space of all radial bounded functions on C, and RBC be the space of all radial bounded functions on C having a finite limit at infinity. We are going to describe the algebra T(n)(RBC). Lemma 6.4. Let m ∈ N0 and x > 0. Then sup lim d→∞ 0≤t≤xℓ(d) m (t) = 0. Proof. For each t < x, we write ℓ(d) bounds: m (t) explicitly by (9) and (4), then apply simple upper m (t) =s m! ℓ(d) ≤s m! (m + d)! t d 2 e− t 2 L(d) m (t) ≤s m! (m + d)! e−t/2 (m + d)! (d + j)! tj+ d 2 ≤ (m + 1)√m! 2 j! mXj=0(cid:18)m + d m − j(cid:19) tj+ d p(m + d)! (m + d)m (1 + t)m+ d 2 . √m! (m + 1) (m + d)m (1 + x)m+ d 2 , p(m + d)! 21 Then, (m + d)! mXj=0 0≤t≤xℓ(d) sup m (t) ≤ and the last expression tends to 0 as d tends to ∞. The following lemma and proposition are similar to [34, Lemma 7.2.3 and Theo- rem 7.2.4]. Lemma 6.5. Let a ∈ L∞([0, +∞)), v ∈ C, and lim r→+∞ a(r) = v. Then In particular, βa,d,j,k = δj,kv lim d→+∞ (j, k ∈ N0). λa,n(p) = v lim p→∞ (n ∈ N). (44) (45) Proof. 1. First, suppose that v = 0 and j = k. For every x > 0 and d ≥ 0, βa,d,j,j ≤Z x j (t)(cid:1)2 dt +Z +∞ 0 a(√t)(cid:0)ℓ(d) j (t)(cid:19)2 ≤ xkak∞(cid:18) sup 0≤t≤xℓ(d) + sup a(√t)(cid:0)ℓ(d) t>x a(√t). x j (t)(cid:1)2 dt Let ε > 0. Using the assumption that a(r) → 0 as r → +∞, we choose x such that the second summand is less than ε/2. After that, applying Lemma 6.4 with this fixed x, we make the first summand less than ε/2. inequality and the result of the first part of this proof. 2. If v = 0, j, k ∈ N0, then we obtain limd→+∞ βa,d,j,k = 0 by applying the Schwarz 3. For general v in C, we rewrite a in the form (a − v1(0,+∞)) + v1(0,+∞). Since β1(0,+∞),d,j,k =Z +∞ 0 ℓ(d) j (t)ℓ(d) k (t) dt = δj,k, the limit relation (44) follows from the result of the second part of this proof. Proposition 6.6. The C*-algebra T(n)(RBC) is isometrically isomorphic to c(N0). Proof. Recall that Φ(n) is an isometrical isomorphism R(n) → ℓ∞(N0) defined by (41). By Proposition 6.3, Φ(n)({Tb : b ∈ RBC}) = L, where L := {λa,n : a ∈ L∞([0, +∞)), ∃v ∈ C lim r→+∞ a(r) = v}. So, T(n)(RBC) is isometrically isomorphic to the C*-subalgebra of ℓ∞(N0) generated by the set L. By Lemma 6.5, L ⊆ c(N0). Our objective is to show that the C*-subalgebra of c(N0) generated by L coincides with c(N0). The space c(N0) may be viewed as the C*-algebra of the continuous functions on the compact N0 ∪{+∞}. The set L is a vector subspace of c(N0) which contains the constants and is closed under the pointwise conjugation. In order to apply the Stone -- Weierstrass theorem, we have to prove that the set L separates the points of N0 ∪ {+∞}. For every u in (0, +∞], define au to be the characteristic function 1(0,u). Then λau,n(p) =Z u2 0 (cid:0)ℓ(p−n+1) min{p,n−1} (t)(cid:1)2 dt. 22 Let p, q ∈ N0, p 6= q. If λau,n(p) = λau,n(q) for all u > 0, then for all t > 0 (cid:0)ℓ(p−n+1) min{p,n−1} (t)(cid:1)2 =(cid:0)ℓ(q−n+1) min{q,n−1} (t)(cid:1)2, which is not true. So, the set L separates p and q. Now let p ∈ N0 and q = +∞. Put u = 1. Then λa1,n(p) > 0, but λa1,n(+∞) = limr→+∞ a1(r) = 0. So, the set L separates p and +∞. Recall that Φn : Rn → Mn is defined by (39). Proposition 6.7. Let a ∈ L∞([0, +∞)). Then Tn,ea ∈ Rn, and the d-th component of the sequence Ψ(Tn,ea) is the matrix Proof. Apply Corollary 4.6 and Proposition 6.1. Ψ(Tn,ea)d =(cid:2)βa,d,j,k(cid:3)n−1 j,k=max{0,−d} . Let Cn be the C*-subalgebra of Mn that consists of all matrix sequences that have scalar limits: Cn := {A ∈ Mn : Proposition 6.8. Φn(Tn(RBC)) ⊆ Cn. Proof. Follows from Lemma 6.5. ∃v ∈ C lim d→+∞ Ad = v In}. We finish this section with a couple of conjectures. Conjecture 6.9. The C*-algebra T(n)(RB) is isometrically isomorphic to the C*-algebra of bounded square-root-oscillating sequences. The concept of square-root-oscillating sequences and a proof of Conjecture 6.9 for n = 1 can be found in [10]. Conjecture 6.10. Φn(Tn(RBC)) = Cn. Various results, similar to Conjecture 6.10, but for Toeplitz operators in other spaces of functions or with generating symbols invariant under other group actions, were proved by Loaiza, Lozano, Ram´ırez Ortega, S´anchez Nungaray, Gonz´alez-Flores, L´opez-Mart´ınez, and Arroyo-Neri [22, 23, 26, 29]. Acknowledgements The authors are grateful to the CONACYT (Mexico) scholarships and to IPN-SIP projects (Instituto Polit´ecnico Nacional, Mexico) for the financial support. This research is inspired by many works of Nikolai Vasilevski. We also thank Jorge Iv´an Correo Rosas for discus- sions of the proof of Proposition 3.4. 23 References [1] L.D. Abreu, On the structure of Gabor and super Gabor spaces, Monatsh. Math. 161 (2010), 237 -- 253, doi:10.1007/s00605-009-0177-0. [2] L.D. Abreu and H.G. Feichtinger, Function spaces of polyanalytic functions, Harmonic and Complex Analysis and its Applications, Birkhauser, 2014, 1 -- 38, doi:10.1007/978-3-319-01806-5 1. [3] A.T. Ali, F. Bagarello, and J.P. Gazeau, D-pseudo-bosons, complex Hermite polyno- mials, and integral quantization, Symmetry Integr. Geom. 11 (2015), 078, 23 pages, doi:10.3842/SIGMA.2015.078. [4] N. Askour, A. Intissar, and Z. Mouayn, Explicit formulas for reproducing kernels of generalized Bargmann spaces, C. R. Acad. Sci. Paris, Ser. I 325 (1997), 707 -- 712, doi:10.1016/S0764-4442(97)80045-6. [5] M.B. Balk, Polyanalytic Functions, Akad.-Verl., 1991. [6] W. Bauer, C. Herrera Yanez, and N. Vasilevski, Eigenvalue characterization of radial operators on weighted Bergman spaces over the unit ball, Integr. Equat. Oper. Th. 78 (2014), 1 -- 30, doi:10.1007/s00020-013-2101-1. [7] C.A. Berger, L.A. Coburn, Toeplitz operators and quantum mechanics, J. Funct. Anal. 68 (1986), 273 -- 299, doi:10.1016/0022-1236(86)90099-6. [8] M. Dawson, G. ´Olafsson, and R. Quiroga-Barranco, Commuting Toeplitz operators on bounded symmetric domains and multiplicity-free restrictions of holomorphic discrete series, J. Funct. Anal. 268 (2015), 1711 -- 1732, doi:10.1016/j.jfa.2014.12.002. [9] M. Englis, Berezin and Berezin-Toeplitz quantizations for general function spaces, Rev. Mat. Complut. 19 (2006), 385 -- 430, http://eudml.org/doc/41908. [10] K. Esmeral and E. Maximenko, Radial Toeplitz operators on the Fock space and square-root-slowly oscillating sequences, Complex Anal. Oper. Th. 10 (2016), 1655 -- 1677, doi:10.1007/s11785-016-0557-0. [11] S. Grudsky, R. Quiroga-Barranco, and N. Vasilevski, Commutative C*-algebras of Toeplitz operators and quantization on the unit disk, J. Funct. Anal. 234 (2006), 1 -- 44, doi:10.1016/j.jfa.2005.11.015. [12] S.M. Grudsky, E.A. Maximenko, and N.L. Vasilevski, Radial Toeplitz operators on the unit ball and slowly oscillating sequences, Commun. Math. Anal. 14:2 (2013), 77 -- 94, https://projecteuclid.org/euclid.cma/1356039033. [13] S. Grudsky and N. Vasilevski, Toeplitz operators on the Fock space: Radial compo- nent effects, Integr. Equat. Oper. Th. 44 (2002), 10 -- 37, doi:10.1007/BF01197858. 24 [14] A. Haimi and H. Hedenmalm, The polyanalytic Ginibre ensembles, J. Stat. Phys. 153 (2013), 10 -- 47, doi:10.1007/s10955-013-0813-x. [15] C. Herrera Yanez, N. Vasilevski, and E.A. Maximenko, Radial Toeplitz operators revisited: Discretization of the vertical case Integr. Equat. Oper. Th. 83 (2015), 49 -- 60, doi:10.1007/s00020-014-2213-2. [16] O. Hutn´ık, On the structure of the space of wavelet transforms, C. R. Acad. Sci. Paris, Ser. I 346 (2008), 649 -- 652, doi:10.1016/j.crma.2008.04.013. [17] O. Hutn´ık, A note on wavelet subspaces, Monatsh. Math. 160 (2010), 59 -- 72, doi:10.1007/s00605-008-0084-9. [18] O. Hutn´ık, E. Maximenko, and A. Miskov´a, Toeplitz localization operators: functions density, Complex Anal. Oper. Th. 10 (2016), 1757 -- 1774, spectral doi:10.1007/s11785-016-0564-1. [19] O. Hutn´ık and M. Hutn´ıkov´a, Toeplitz operators on poly-analytic spaces via time- scale analysis, Oper. Matrices 8 (2015), 1107 -- 1129, doi:10.7153/oam-08-62. [20] B. Korenblum and K. Zhu, An application of Tauberian theorems to Toeplitz opera- tors, J. Oper. Th. 33 (1995), 353 -- 361, https://www.jstor.org/stable/24714916. [21] M. Loaiza and C. Lozano, On C*-algebras of Toeplitz operators on the 105 -- 130, Integr. Equat. Oper. Th. 76 (2013), harmonic Bergman space, doi:10.1007/s00020-013-2046-4. [22] M. Loaiza and C. Lozano, On Toeplitz operators on the weighted harmonic Bergman space on the upper half-plane, Complex Anal. Oper. Th. 9 (2014), 139 -- 165, doi:10.1007/s11785-014-0388-9. [23] M. Loaiza and J. Ram´ırez-Ortega, Toeplitz operators with homogeneous symbols acting on the poly-Bergman spaces of the upper half-plane, Integr. Equat. Oper. Th. 87 (2017), 391 -- 410, doi:10.1007/s00020-017-2350-5. [24] R. Quiroga-Barranco, Separately radial and radial Toeplitz operators on the unit ball and representation theory, Bol. Soc. Mat. Mex. 22 (2016), 605 -- 623 doi:10.1007/s40590-016-0111-0 [25] A.K. Ramazanov, Representation of the space of polyanalytic functions as a direct sum of orthogonal subspaces. Application to rational approximations, Math. Notes 66 (1999), 613 -- 627, doi:10.1007/BF02674203. [26] J. Ram´ırez Ortega and A. S´anchez-Nungaray, Toeplitz operators with vertical symbols acting on the poly-Bergman spaces of the upper half-plane, Complex Anal. Oper. Th. 9 (2015), 1801 -- 1817, doi:10.1007/s11785-015-0469-4. 25 [27] G. Rozenblum and N.L. Vasilevski, Toeplitz sesquilinear forms, Operator Theory: Adv. and Appl. 262, Birkhauser, 2018, 287 -- 304. doi:10.1007/978-3-319-62527-0 9. operators via [28] S. Sakai, C∗-algebras and W ∗-algebras, Springer-Verlag, 1971. [29] A. S´anchez-Nungaray, C. Gonz´alez-Flores, R.R. L´opez-Mart´ınez, and J.L. Arroyo- Neri, Toeplitz operators with horizontal symbols acting on the poly-Fock spaces, J. Funct. Spaces 2018 (2018), Article ID 8031259, 8 pages, doi:10.1155/2018/8031259. [30] I. Shigekawa, Eigenvalue problems for the Schrodinger operator with the mag- netic field on a compact Riemannian manifold, J. Funct. Anal. 75 (1987), 92 -- 127, doi:10.1016/0022-1236(87)90108-X. [31] K. Stroethoff, The Berezin transform and operators on spaces of analytic functions, Banach Center Publ. 38 (1997), 361 -- 380, doi:10.4064/-38-1-361-380. [32] D. Su´arez, The eigenvalues of limits of radial Toeplitz operators, Bull. Lond. Math. Soc. 40 (2008), 631 -- 641, doi:10.1112/blms/bdn042. [33] N.L. Vasilevski, Poly-Fock spaces, Operator Theory: Adv. and Appl. 117, Birkhauser, 2000, 371 -- 386, doi:10.1007/978-3-0348-8403-7 28. [34] N.L. Vasilevski, Commutative Algebras of Toeplitz Operators on the Bergman Space, Birkhauser, 2008, doi:10.1007/978-3-7643-8726-6. [35] K. Zhu, An Introduction to Operator Algebras, CRC Press, 1993. [36] K. Zhu, Analysis on Fock Spaces, Springer, 2012, doi:10.1007/978-1-4419-8801-0. [37] N. Zorboska, The Berezin transform and radial operators, Proc. Amer. Math. Soc. 131 (2003), 793 -- 800, https://www.jstor.org/stable/1194482. Egor A. Maximenko Instituto Polit´ecnico Nacional Escuela Superior de F´ısica y Matem´aticas Ciudad de M´exico Mexico e-mail: [email protected] https://orcid.org/0000-0002-1497-4338 Ana Mar´ıa Teller´ıa-Romero Instituto Polit´ecnico Nacional Escuela Superior de F´ısica y Matem´aticas Ciudad de M´exico Mexico e-mail: [email protected] https://orcid.org/0000-0002-9821-9398 26
1803.09212
2
1803
2018-07-18T09:04:10
Shape, Scale, and Minimality of Matrix Ranges
[ "math.OA", "math.FA" ]
We study containment and uniqueness problems concerning matrix convex sets. First, to what extent is a matrix convex set determined by its first level? Our results in this direction quantify the disparity between two product operations, namely the product of the smallest matrix convex sets over $K_i \subseteq \mathbb{C}^d$, and the smallest matrix convex set over the product of $K_i$. Second, if a matrix convex set is given as the matrix range of an operator tuple $T$, when is $T$ determined uniquely? We provide counterexamples to results in the literature, showing that a compact tuple meeting a minimality condition need not be determined uniquely, even if its matrix range is a particularly friendly set. Finally, our results may be used to improve dilation scales, such as the norm bound on the dilation of (non self-adjoint) contractions to commuting normal operators, both concretely and abstractly.
math.OA
math
SHAPE, SCALE, AND MINIMALITY OF MATRIX RANGES BENJAMIN PASSER† Abstract. We study containment and uniqueness problems concerning matrix convex sets. First, to what extent is a matrix convex set determined by its first level? Our results in this direction quantify the disparity between two product operations, namely the product of the smallest matrix convex sets over Ki ⊆ Cd, and the smallest matrix convex set over the product of Ki. Second, if a matrix convex set is given as the matrix range of an operator tuple T , when is T determined uniquely? We provide counterexamples to results in the literature, showing that a compact tuple meeting a minimality condition need not be determined uniquely, even if its matrix range is a particularly friendly set. Finally, our results may be used to improve dilation scales, such as the norm bound on the dilation of (non self-adjoint) contractions to commuting normal operators, both concretely and abstractly. 8 1 0 2 l u J 8 1 ] . A O h t a m [ 2 v 2 1 2 9 0 . 3 0 8 1 : v i X r a 1. Introduction The noncommutative generalization of a function system is called an operator system, and many crucial objects in the study of function systems also generalize to the noncommutative setting [4, 15]. Definition 1.1. An operator system is a self-adjoint, unital subspace S of a unital C ∗-algebra A. If A = C(X) is commutative, then S is also called a function system on X. Any unital C ∗-algebra is spanned by its positive (or more precisely, positive semidefinite) elements, which by definition are self-adjoint elements a ∈ A whose spectra are contained in the nonnegative real line. We write a ≥ 0 when a is positive, noting that positivity of a is equivalent to the claim that a factorization a = bb∗ exists for some b ∈ A. While an operator system S ⊆ A might not have a multiplicative structure of its own, by considering the given C ∗-algebra in which S lives, one may point out the set of positive elements in S. In particular, S is also spanned by its positive elements, which make up a crucial part of the operator system structure, as in the abstract definition found in [6]. The above discussion of positivity applies equally well to the set of n × n matrices over S, as Mn(S) embeds into the unital C ∗-algebra Mn(A). Further, to any map φ : S → T between operator systems, one may also produce maps φ(n) : Mn(S) → Mn(T ) which apply φ entrywise. The relevant notion of morphism between operator systems is a map which respects all of the above structure, on every matrix level, as in the following definition. Definition 1.2. Let S ⊆ A and T ⊆ B be operator systems. Then a linear map φ : S → T is a unital completely positive map, or UCP map, if φ(1) = 1 and each φ(n) is positive -- for any matrix s ∈ Mn(S) such that s ≥ 0, it follows that φ(n)(s) ≥ 0. Date: July 8, 2021. 2010 Mathematics Subject Classification. 47A20, 47A13, 46L07, 47L25. Key words and phrases. matrix convex set; dilation; operator system; matrix range. † Partially supported by a Zuckerman Fellowship at the Technion. 1 Any unital C ∗-algebra A is by default an operator system, and Arveson's extension theo- rem [1, Theorem 1.2.3] implies that for an operator system S ⊆ A, any UCP map S → B(H) extends to a UCP map A → B(H). Therefore, when studying the interpolation problem for UCP maps into B(H), the choice of domain is generally not important. (In contrast, it is often of great interest if an extension of a UCP map given by Arveson's extension theorem is unique, and this problem is certainly domain-sensitive. See, for example, the role of unique extensions in Arveson's Hyperrigidity Conjecture [4, Conjecture 4.3].) The interpolation problem for UCP maps reduces to consideration of the matrix range, defined below. Definition 1.3. The matrix range of A = (A1, . . . , Ad) ∈ B(H)d, denoted W(A) = is a subset of the matrix universe Md = n defined on each n × n level by M d ∞Sn=1Wn(A), Wn(A) := {(Q1, . . . , Qd) ∈ M d n : ∃ UCP map ψ : B(H) → Mn with ψ(Ai) = Qi}. From a slight reworking of [2, Theorem 2.4.2] in [8, Theorem 5.1], if A ∈ B(H)d and B ∈ B(K)d, then a UCP map φ : B(H) → B(K) mapping φ(Ai) = Bi exists precisely if W(B) ⊆ W(A). Thus, the interpolation problem for UCP maps reduces to the consideration of (all) UCP maps whose codomains are finite-dimensional. For any A, the matrix range W(A) is a closed and bounded matrix convex set [8, Proposition 2.5]. More precisely, each Wn(A) is closed as a subset of M d n with the product norm topology, and there is a uniform bound (independent of n) on the norm of any member of any tuple belonging to Wn(A). Matrix convexity is defined as follows. Definition 1.4. A set S = of direct sums and UCP maps. That is, S meets the following conditions. ∞Sn=1Sn ⊆ Md is matrix convex if it is closed under the application • If X ∈ Sn and Y ∈ Sm, then X ⊕ Y ∈ Sn+m. • If X ∈ Sn and φ : Mn → Mm is a UCP map, then (φ(X1), . . . , φ(Xd)) ∈ Sm. In fact, operator systems and matrix convex sets are dual to each other [9]. From Choi's theorem [5], which characterizes completely positive maps between matrix algebras, an equiv- alent definition of matrix convexity follows. Definition 1.5. Let S be a subset of Md. For each i ∈ {1, . . . , N}, let Y i ∈ Sni and let Vi : Cn → Cni be a linear map, such that V ∗ i Vi = In. Then ∞Sn=1 NPi=1 NXi=1 2 (1.6) X := V ∗ i Y iVi is called a matrix convex combination of Y 1, . . . , Y N ∈ S. Definition 1.7. A set S ⊆ Md is matrix convex if whenever X is a matrix convex combi- nation of Y 1, . . . , Y N ∈ S, it follows that X ∈ S. Note in particular that if X ∈ Sn and Y ∈ Sm, then X ⊕ Y = V ∗ 2 Y V2 for a natural choice of coisometries V1 and V2, so X ⊕ Y is a matrix convex combination of X and Y . Further, any unitary conjugation U ∗XU is a matrix convex combination of X that uses only one summand. 1 XV1 + V ∗ It is of great interest what the proper notion of extreme point should be in the matrix con- vex setting, analogous to the Krein-Milman theorem (as well as Milman's converse) and the Minkowski/Steinitz theorem in the compact convex setting [19, 22]. Two major candidates are matrix extreme points and absolute extreme points, and both have been characterized in dilation-theoretic terms [11, Theorem 1.1]. However, both candidates have limitations. If S is a closed and bounded matrix convex set, then S is generated by its matrix extreme points, in the sense that the smallest closed matrix convex set containing these points is equal to S. However, it is possible for a matrix extreme point to be a nontrivial matrix convex combination of other matrix extreme points, and it is not known if there is a smaller generating set. While the definition of an absolute extreme point forces its representation as a matrix convex combination to be essentially unique, it is possible for S to have no absolute extreme points at all [10, Corollary 1.1]. Given a compact convex set K ⊆ Cd, there might be many matrix convex sets S such that S1 = K, but there is always a smallest and largest choice of S [8, Definition 4.1 and Proposition 4.3]. They may be presented in multiple equivalent ways: (1.8) W min(K) = {X ∈ Md : there is a normal dilation N of X with σ(N) ⊆ K} = {X ∈ Md : there is a normal matrix dilation N of X with σ(N) ⊆ K} ajXj! ≤ bI) ajxj! ≤ b for all x ∈ K, then Re dXj=1 if Re dXj=1 and (1.9) W max(K) =(X ∈ Md : = {X ∈ Md : W1(X) ⊆ K}. We remind the reader that when a tuple N = (N1, . . . , Nd) is called normal, this means that the operators N1, . . . , Nd are normal and commute with each other. Further, Y ∈ B(K)d is a dilation of X ∈ B(H)d if there exists an isometry V : H → K such that Xi = V ∗YiV for each i. Equivalently, X is a compression of Y . There is a considerable amount of information buried in the previous definitions. For example, as W min(K) is by definition the smallest matrix convex set spanned by the scalar set K, this spanning property does not include a closure operation. However, the second formulation of W min(K) (with added dimension bounds) shows it is actually closed. More detail from [8] and [21] is given below, and the following proposition may be seen as a manipulation of the Stinespring dilation procedure [23]. Proposition 1.10. ([8, Corollary 2.8 and Proposition 4.4]) If N ∈ B(H)d is a normal tuple, then W(N) = W min(K), where K is the convex hull of σ(N). Since every compact convex set K ⊆ Cd may be written as the convex hull of σ(N) for a diagonal operator N, it follows that any W min(K) may be written as a matrix range W(N), and it is therefore also closed and bounded. Alternatively, a compactness argument can be used, as if T ∈ W min (K), then there is a normal tuple of matrices with a fixed dimension bound (depending on n) which dilates T and has spectrum in K. Proposition 1.11. (reformulation of [8, Theorem 7.1] and [21, Proposition 2.3]) If T ∈ B(H)d has W(T ) ⊆ W min(K), then there is a normal dilation N of T with σ(N) ⊆ ext(K). n 3 If, in addition, T acts on a finite-dimensional space of dimension n, then we may choose N to act on a space of dimension 2n3(d + 1) + 1 or lower. Note that since T above need not act on a finite-dimensional space, Proposition 1.11 shows that the existence of normal dilations for a family of matrices may be used to produce a normal dilation of an infinite-dimensional operator. In the background of this claim lies the fact that UCP maps (and hence Stinespring dilations) behave very well with respect to limits in pointwise topologies. For most of the problems we pursue, there is no harm in considering a tuple (T1, . . . , Td) of d operators as a tuple (X1, Y1, . . . , Xd, Yd) of 2d self-adjoint operators instead. In particular, this notational change does not affect the definitions of W(T ), W min(K), and W max(K). Therefore, whenever it is possible, we will restrict proofs to the self-adjoint setting B(H)d sa and assume that any scalar tuples we consider belong to Rd. We also note that it is possible for W max(K) and W min(K) to be equal, and this occurs if and only if K is a simplex. More specifically, for a compact convex set K ⊆ Rd, W max(K) = W min(K) ⇐⇒ W max (1.12) 2d−1(K) ⇐⇒ K is a simplex 2d−1(K) = W min holds from [21, Theorem 4.1]. See also [13, Theorem 4.7] for the equivalence of the first and third items when K is a polyhedron, phrased in the language of operator systems. If T ∈ B(H)d sa and A is an invertible affine transformation on Rd, then W(A(T )) = A(W(T )), (1.13) where A is applied to operator tuples in the natural way, as in [21, §3]. This also implies that for any compact convex set K ⊆ Rd, W min(A(K)) = A(W min(K)) (1.14) Similarly, if H ⊆ Rd is an affine subspace with orthogonal projection PH : Rd → H, then (1.15) W max(A(K)) = A(W max(K)). W max(K) ⊆ W min(L) =⇒ W max(K ∩ H) ⊆ W min(PH(L)) and by [21, Lemma 3.2]. When our computations take place entirely in a proper affine subspace of Rd, we may then use (1.13), (1.14), and (1.15) to reduce the ambient space to Rn for n < d. This allows us to prove results for all compact convex sets by focusing only on convex bodies (compact convex sets with nonempty interior). Roughly speaking, this corresponds to throwing out useless 0 operators in a tuple (T1, . . . , Td, 0, . . . , 0) to focus on (T1, . . . , Td) alone. If one considers the graded product NYj=1 S j := ∞[n=1 NYj=1 S j n of matrix convex sets, then it is evident from (1.9) that W max NQj=1 Ki! = However, such a factorization generally does not exist for W min. The root of the problem is that if (A1, . . . , Ad) and (B1, . . . , Bd) are normal tuples, then (A1, . . . , Ad, B1, . . . , Bd) might fail to be normal, as the various Ai and Bj might not commute. In section 2 we consider NQj=1W max(Ki). 4 containment problems of the form W min(Ki) ⊆ W min NYj=1 nYj=1 ci · Ki! . For symmetric Ki, it is possible to derive such containments from estimates concerning products of simplices and products of diamonds, as in Theorem 2.9. Since the matrix convex set S consisting of all d-tuples of matrix contractions may be written as can then obtain a dilation scale result in Corollary 2.18 for tuples of contractions (see also Theorem 4.4). In contrast, for Ki which are not necessarily symmetric, we show in Corollary 2.33 that estimates derived from a very slight modification of dilations in [8] cannot be improved. dQj=1W min(D), we In section 3, we first consider the interaction between dilation theorems and compactness of operators. While we cannot guarantee that compactness is preserved in a dilation that comes from Proposition 1.11, it does hold from Proposition 3.1 that if T is compact and W(T ) ⊆ W min(K), then there is a compact normal dilation N of T with spectrum in a neighborhood of K. Moreover, if a compact tuple T has W(T ) = W min(K), then K has the shape one would expect if it were assumed that T is also normal, as in Theorem 3.12. However, unlike in the finite-dimensional setting, the assumption that T is minimal for its matrix range does not characterize T up to unitary equivalence, and a minimal compact T with W(T ) = W min(K) need not be normal. In particular, for many K there exist uncountably many inequivalent, compact, minimal tuples with matrix range equal to W min(K), as in Corollary 3.15. Similarly, if T is compact, a minimal summand S of T with the same matrix range might not exist, from Example 3.22. These results indicated the need for additional assumptions in the theorems of [8, §6], which have now been corrected in response (see [7]). In the restricted setting of W min sets and compact operators, we consider some alternative relaxations of the problem in Propositions 3.21 and 3.24, which we believe could be useful starting points for further study. The remainder of section 3 concerns operator tuples which are not necessarily compact. A simple spectral theorem argument in Theorem 3.26 shows that there is a minimal normal T for matrix range W(T ) = W min(K) if and only if K satisfies a simple geometric condition: the isolated extreme points of K are dense in the set of all extreme points of K. In this case, T must be diagonal with eigenvalues at the isolated extreme points of K. However, if K has at least three extreme points, then Corollary 3.27 shows that the same condition on isolated extreme points implies the existence of uncountably many non-normal minimal tuples for matrix range W min(K). Finally, for any compact convex set K with at least three extreme points, there is a tuple T with W(T ) = W min(K) such that T has no summand which is minimal for the same matrix range, and such that T has no normal summands at all, as in Theorem 3.29. Finally, in section 4, we consider two matrix convex set containments that are demon- strated by explicit dilation procedures. First, we dilate tuples of contractions to normal tuples in Theorem 4.4, with a new norm bound (see also Corollary 2.18). Second, we give a lower bound for the matrix range of a universal tuple of anticommuting self-adjoint unitaries, using an explicit dilation procedure developed in Theorem 4.13 and Corollary 4.16. 5 As in [21], for nonempty compact convex sets K and L in Euclidean space, define 2. Products of Minimal Sets and θ(K) := inf{C > 0 : W max(K) ⊆ C · W min(K)} θ(K, L) := inf{C > 0 : W max(K) ⊆ C · W min(L)}. Two disparate estimates (2.1) θ([−1, 1]d) = √d < d = θ([0, 1]d) were computed in [21, Theorems 6.4 and 6.7], along with a non-uniform version of the first equality, (2.2) W max([−1, 1]d) ⊆ W min dYj=1 [−aj, aj]! ⇐⇒ dXj=1 a−2 j ≤ 1. Equation (2.1) is equivalent to the following dilation results. (1) If (X1, . . . , Xd) ∈ B(H)d sa is a tuple of self-adjoint contractions, then there exists a dilation tuple (M1, . . . , Md) of commuting self-adjoint operators with norm Mi ≤ √d, and √d is the optimal constant. (2) If (P1, . . . , Pd) ∈ B(H)d sa is a tuple of positive contractions, then there exists a dilation tuple (N1, . . . , Nd) of commuting positive operators with norm Ni ≤ d, and d is the optimal constant. Both computations are paired with explicit dilation procedures. The disparity between the two constants emphasizes the fact that, when dilating self-adjoint contractions to self- adjoint operators which commute, the preservation of another relation among the contrac- tions (namely, positivity) significantly alters the norm bound one can achieve. Consistent with this idea, the dilation procedure in [21, Theorem 6.7] which demonstrates (1) begins by replacing each Xi with a self-adjoint unitary. Therefore, the dilation procedure is generally not able to preserve additional properties of the tuple. Namely, it cannot • preserve compactness of each Xi, or • preserve the satisfaction of other linear or non-linear inequalities by the Xi (which • successfully use the fact that some of the Xi might already commute to lower the are independent from −1 ≤ Xi ≤ 1), or norm of the dilation in this special case without modification. In contrast, an earlier explicit dilation procedure that demonstrates θ([−1, 1]d) ≤ d in [8, §7], while far from achieving the optimal constant, does preserve compactness information, and it simultaneously demonstrates multiple containments of the form W max(K) ⊆ W min(L). In this section, we expand upon the third bullet point, in pursuit of the following problem. Problem 2.3. Compute when dQj=1W min(Ki) ⊆ W min (L). Namely, if a large tuple M consists ni ), where each M [i] is normal with joint spectrum of smaller subtuples M [i] = (M [i] σ(M [i]) ⊆ Ki, does M admit a normal dilation N with joint spectrum in L? 1 , . . . , M [i] 6 In particular, we show that when the Ki are symmetric and L is the product of (perhaps distinct) multiples of Ki, containment theorems follow from seemingly unrelated dilation constants. These constants are defined in reference to products of simplices and products of diamonds. Definition 2.4. The standard simplex ∆n in Rn refers to the convex hull of 0 and the standard basis vectors e1, . . . , en. The corresponding standard diamond ⋄n is the ℓ1 unit ball in Rn, i.e., the smallest symmetric convex set containing ∆n. Definition 2.5. Call a tuple (a1, . . . , ad) of positive numbers an SD-tuple if it holds that for any n ∈ Z+, (2.6) Similarly, let aj ·⋄n! . n(cid:1) ⊆ W min dYj=1 W max(cid:0)∆d n,⋄d θ(∆d n) U(d) := sup n∈Z+ = inf{C > 0 : (C, . . . , C) is an SD-tuple of length d} be called the uniform SD-constant for d-tuples. Dilation techniques from [8] imply that SD-tuples exist and that U(d) is finite, and such results will be clarified later in this section. The most common use of SD-tuples will come in the following form. Proposition 2.7. Fix n1, . . . , nd ∈ Z+ and let P be a tuple of self-adjoint operators P [i] j , j ≥ 0, and P [i] j ≤ I.) If (a1, . . . , ad) is an SD-tuple, then there exists a dilation Q ≻ P 1 ≤ i ≤ d, 1 ≤ j ≤ ni, on B(H) such that W1(P ) ⊆ (cid:18) dQi=1 ∆ni(cid:19). (That is, P [i] for each i, consisting of self-adjoint operators with the following properties. niPj=1 j has σ(Q[i] (1) Each Q[i] (2) The operators Q[i] (3) If i is fixed, then distinct members of the tuple Q[i] = (Q[i] (4) If H is finite-dimensional, then Q acts on a finite-dimensional space as well. j ) ⊆ {−ai, 0, ai}. j all commute with each other. 1 , . . . , Q[i] ni) are orthogonal. Proof. By introducing the 0 operator into tuples if necessary, we may assume that n1, . . . , nd are equal to a fixed n. Since W max(cid:0)∆d n(cid:1) ⊆ W min dQj=1 aj ·⋄n! holds, Proposition 1.11 shows aj·⋄n, where Q acts on a finite-dimensional space if H is finite-dimensional. The extreme points of there is a normal dilation Q of P whose joint spectrum lies in the extreme points of dQj=1 aj ·⋄n are positioned exactly so that the remaining properties (1)-(3) also hold. 7 dQj=1 j are positive, or that all products of distinct Q[i] Remark 2.8. It is important to note that the properties listed do not imply that the operators Q[i] j are zero. We abuse notation somewhat by letting Q = (Q[1], . . . , Q[d]) denote the "conjoined tuple" consisting of all the matrices Q[i] Theorem 2.9. Let (a1, . . . , ad) be an SD-tuple, and for each 1 ≤ i ≤ d, let Ki ⊆ Rni be a compact convex set with Ki = −Ki. Then (2.10) j , ordered by i first and j second. holds. Consequently, it also holds that dYi=1 W max dYi=1 Ki! ≤ inf W min(Ki) ⊆ W min dYi=1 Ki! ⊆ W min dYi=1 a is SD(cid:20) max aiKi! θ(Ki)aiKi! aiθ(Ki)(cid:21) ≤ U(d) · max 1≤i≤d 1≤i≤d θ(Ki). θ dYi=1 (2.11) and (2.12) Proof. Fix m ∈ Z+. For each i ∈ {1, . . . , d}, let N [i] = (N [i] ni ) be a normal tuple of self-adjoint m × m matrices with joint spectrum satisfying σ(N [i]) ⊆ Ki. The normal tuple N [i] admits a joint diagonalization, so we may specify a tuple P [i] = (P [i] m ) of 1 , . . . , N [i] 1 , . . . , P [i] P [i] k = Im and a collection of eigenvalues λ[i] j,k such mutually orthogonal projections with that mPk=1 N [i] j = λ[i] j,kP [i] k mXk=1 mXk=1 holds. By assumption, for each i and k, the tuple (λ[i] ni,k) belongs to Ki. Consider the conjoined tuple P = (P [1], . . . , P [d]), which has W1(P ) ⊆ ∆d m, so that we may form a dilation Q = (Q[1], . . . , Q[d]) guaranteed by Proposition 2.7. All of the matrices Q[i] k Q[i] k commute with each other, and if i is fixed but k 6= l, it follows that Q[i] l = 0. Finally, each matrix Q[i] k ) ⊆ {−ai, 0, ai}. Therefore, for each fixed i, the matrices k has σ(Q[i] 1,k, . . . , λ[i] (2.13) M [i] j := λ[i] j,kQ[i] k are essentially given in jointly diagonalized form, and the tuple M [i] = (M [i] ni ) is normal with joint spectrum contained in −aiKi ∪ {0} ∪ aiKi = aiKi. Further, for any choices of i and j, the self-adjoint matrices M [i] j commute with each other, so the conjoined tuple M = (M [1], . . . , M [d]) is also normal. Finally, the joint spectrum of M is contained 1 , . . . , M [i] in σ(M [i]), which is contained in dQi=1 follow from the equality W max(cid:18) dQi=1 U(d). dQi=1 Ki(cid:19) = aiKi, so (2.10) holds. The remaining identities then dQi=1W max(Ki) and the definitions of θ(Ki) and 8 A consequence of this result is that SD-tuples may be defined with reference only to diamonds, as opposed to both simplices and diamonds. Corollary 2.14. A tuple (a1, . . . , ad) of positive numbers is an SD-tuple if and only if for each n ∈ Z+, Proof. The forward direction is given by Theorem 2.9 for the choice of symmetric set Ki = (2.15) dYj=1 W min (⋄n) ⊆ W min dYj=1 aj ·⋄n! . ⋄n. For the converse, note that (2.15) directly implies (2.6), as W max(cid:0)∆d n(cid:1) = W max(∆n) = W min(∆n) ⊆ dYj=1 dYj=1 W min(⋄n) ⊆ W min dYj=1 dYj=1 aj ·⋄n! . We may also dilate tuples of contractions using the previous results. Let D be the closed unit disk and fix any Y ∈ W min(D), where we may choose to view Y as a single matrix (instead of two self-adjoint matrices). Since Y has a normal dilation N with "joint" spectrum in S1 ⊆ D, it follows that N ≤ 1 and consequently Y ≤ 1. On the other hand, if X is any matrix with X ≤ 1, then the Halmos dilation procedure √I − XX ∗ −X ∗ (cid:19) is a unitary dilation (2.16) X is a contraction =⇒ U :=(cid:18) √I − X ∗X X of [14] shows that X ∈ W min(D). It therefore follows that W min(D) is precisely the collection of (not necessarily self-adjoint) matrix contractions. Applying the graded product, we obtain the following simple fact: (2.17) dYj=1 W min(D) = {T ∈ Md : for each i ∈ {1, . . . , d},Ti ≤ 1}. Corollary 2.18. Let H be a Hilbert space of any dimension, and let T ∈ B(H)d be a tuple of (not necessarily self-adjoint) contractions. Then for any SD-tuple (a1, . . . , ad), it holds that • W(T ) ⊆ W min(cid:18) dQi=1 aiD(cid:19), and • there exists a normal tuple N which dilates T and has Ni ≤ ai for each i. In particular, we may choose ai = U(d), and if H is finite-dimensional, we may choose N which acts on a finite-dimensional space. Proof. Since T is a tuple of contractions, any tuple A ∈ W(T ) consists of matrix contrac- tions and is therefore contained in aiD(cid:19). Since we then have W(T ) ⊆ W min(cid:18) dQi=1 dQj=1W min(D) by (2.17). By Theorem 2.9, it follows that aiD(cid:19), applying Proposition A ∈ W min(cid:18) dQi=1 1.11 finishes the proof. 9 W max([−1, 1]d) = dYj=1 Finally, from (2.2) we conclude that W max([−1, 1]) = dPj=1 a−2 j ≤ 1. W min([−1, 1]) ⊆ W min dYj=1 dYj=1 [−aj, aj]! . See Theorem 4.4 for an explicit dilation procedure that begins with a tuple T ∈ B(H)d of contractions and ends with a normal tuple N satisfying Ni ≤ √2d for each i, where the constant √2d is not necessarily optimal. The optimal dilation scale of self-adjoint contrac- tions is known from (2.2), allowing us to place bounds on the collection of SD-tuples. Corollary 2.19. If (a1, . . . , ad) is an SD-tuple, then j ≤ 1. Consequently, U(d) ≥ √d. Proof. If (a1, . . . , ad) is an SD-tuple, then by Theorem 2.9, it holds that a−2 dPj=1 W min([−1, 1]) ⊆ W min dYj=1 [−aj, aj]! . dYj=1 However, the minimal and maximal matrix convex set over an interval are identical, so we have dPi=1 dPi=1 Following the orthogonal case of [8, Theorem 7.7], given a1, . . . , ad > 0 with a−1 i = 1, we dPi=1 have that for any orthonormal system Pi = viv∗ i of rank one projections such that Pi = Id (that is, v1, . . . , vd form a basis of Cd), the unit vector w := a−1/2 vd has haiPiw, wi = 1. Applying a change of basis, we find that for any tuple (a1, . . . , ad) of positive numbers with a−1 i = 1, there exist d × d matrices Q1, . . . , Qd with v1 + . . . + a−1/2 d 1 (2.20) Qi = Q∗ i , Rank(Qi) = 1, σ(Qi) = {0, ai}, QiQj = 0 for i 6= j, Qi =(cid:18)1 ∗ ∗ ∗(cid:19) . It follows that in this circumstance, the dilation technique of [8, Theorem 7.7] replaces a self-adjoint operator Ti with Ti ⊗ Qi, so that (T1 ⊗ Q1, . . . , Td ⊗ Qd) is a normal tuple. We will use such a dilation, where we replace each Ti with a tuple containing multiple operators. Theorem 2.21. For each 1 ≤ i ≤ d, let Ki ⊆ Rni be a compact convex set with 0 ∈ Ki for each i. For positive numbers t1, . . . , td, let L[t1, . . . , td] ⊆ tiKi be the convex hull of all the sets {0} × tiKi × {0}, 1 ≤ i ≤ d, where 0 denotes a tuple of zeroes of the appropriate size. Then for any positive numbers a1, . . . , ad with dQt=1 a−1 i ≤ 1, it follows that (2.22) nYi=1 W min (Ki) ⊆ W min(L[a1, . . . , ad]) ⊆ W min nYi=1 10 aiKi! dPi=1 holds. Consequently, it also holds that (2.23) and (2.24) W max nYi=1 θ nYi=1 Ki! ⊆ W min(L[θ(K1)a1, . . . , θ(Kd)ad]) ⊆ W min nYi=1 Ki! ≤ θ(Ki)ai(cid:21) ≤ d · max i ≤1(cid:20) max ai>0, P a−1 1≤i≤d 1≤i≤d inf θ(Ki). θ(Ki)aiKi! dPi=1 a−1 Proof. Since 0 ∈ Ki, we may shrink the ai so that i = 1. From (2.20), there are mutually orthogonal projections P1, . . . , Pd ∈ Md such that Qi := aiPi has entry 1 in the top-left corner. For each i, let M [i] ∈ B(H)ni sa be a normal tuple with σ(M [i]) ⊆ Ki. Dilate each operator M [i] j by choosing N [i] j := M [i] j ⊗ Qi. Then the conjoined tuple N = (N [1], . . . , N [d]) is a normal dilation of M = (M [1], . . . , M [d]). The spectrum of each N [i] is in aiKi, and moreover N [i] l = 0 if i 6= k. It follows that the joint spectrum of N is contained in L([a1, . . . , ad]), and we may conclude that (2.22) holds. Next, (2.23) follows from (2.22), as any tuple of operators with numerical range in Ki admits a normal dilation with joint spectrum in W min(θ(Ki)Ki), by definition. Finally, (2.24) follows immediately from (2.23). j N [k] Corollary 2.25. If (a1, . . . , ad) is a tuple of positive numbers with (a1, . . . , ad) is an SD-tuple. Consequently, U(d) ≤ d. Proof. Theorem 2.21 shows that for any n ∈ Z+, W max(∆d n) ⊆ W min dYj=1 θ(∆n)aj∆n! = W min dYj=1 dPi=1 aj∆n! ⊆ W min dYj=1 so by definition (a1, . . . , ad) is an SD-tuple. aj⋄n! , a−1 i ≤ 1, then All together, we have that for positive tuples (a1, . . . , ad), (2.26) 1 a1 + . . . + 1 ad ≤ 1 =⇒ (a1, . . . , ad) is an SD-tuple =⇒ 1 a2 1 + . . . + 1 d ≤ 1, a2 and in particular (2.27) √d ≤ U(d) ≤ d. alence, then SD-tuples may be characterized by the identity Analogous to the computation (2.1), we suspect that if one of the implications is an equiv- a−2 j ≤ 1. However, the estimates of Theorem 2.21 are optimal for certain positive sets, such as [0, 1]d, once again emphasizing that dilation problems concerning symmetric sets are (or have the potential to be) considerably more flexible than those concerning positive sets. Recall that the difficulties of positive-to-positive dilation were abstracted in [21] to simplex-pointed sets. dPj=1 11 Definition 2.28. A convex body K ⊆ Rd is simplex-pointed at x ∈ K if there exists a basis v1, . . . , vd of Rd such that the convex hull of x, x + v1, . . . , x + vd is contained in K and is a neighborhood of x in K, where K is equipped with the relative topology from Rd. Equivalently, there is an invertible affine transformation A on Rd such that A(K) ⊆ [0,∞)d, A(x) = 0, and A(K) includes a neighborhood of 0 in [0,∞)d. The following theorem expands upon the techniques in [21] and gives a partial answer to [21, Problem 8.6]: if W max(K) ⊆ W min(L), under what circumstances can we conclude that there is a simplex Π with K ⊆ Π ⊆ L? Theorem 2.29. Let K ⊆ Rd be a convex body which is simplex-pointed at x ∈ K with v1, . . . , vd as in Definition 2.28. Let L be another convex body which is simplex-pointed at the same point x with the same vector data v1, . . . , vd, and suppose that W max (L). If (K) ⊆ W min 2 2 (2.30) then the convex hull of x, x + t1v1, . . . , x + tdvd is a simplex Π with K ⊆ Π ⊆ L. Consequently, if K is simplex-pointed at x = 0 with vector data v1, . . . , vd, and ti := max{t ≥ 1 : x + tvi ∈ L}, then S := {conv(0, s1v1, . . . , sdvd) : s1, . . . , sd ≥ 1}, θ(K) = min{C ≥ 1 : there exists Π ∈ S with K ⊆ Π ⊆ C · K}. ti := max{t ≥ 1 : tei ∈ L}, Fix any interior point c of K, and let q = (d − 1) · max Proof. After an invertible affine transformation (linear if x = 0), we may suppose that K and L are contained in [0,∞)d, x = 0, and the standard simplex ∆d is contained in K and L. Letting e1, . . . , ed denote the standard basis vectors in Rd, we wish to prove that if (2.31) then the convex hull Π of 0, t1e1, . . . , tded has K ⊆ Π (as the containment Π ⊆ L is trivial). 1≤i≤d{ci}. Let ε > 0 be small enough that the qε-neighborhood Nqε of the line segment conv(0, c) has Nqε∩ [0,∞)d ⊆ K. Next, let P = (P1, . . . , Pd) be a tuple of 2 × 2 rank 1 projections such that Ran(Pi) ∩ Ran(Pj) = {0} if i 6= j, but Pi − Pj < ε. Setting X = (c1P1, . . . , cdPd) and Q = (c1P1, c2P1, . . . , cdP1), we have that W1(X) ⊆ [0,∞)d is within the qε-neighborhood of W1(Q) = conv(0, c). This implies two key facts. First, we have that W1(X) ⊆ K. Second, applying a vector state corresponding to a unit vector in Ran(P1) shows that (2.32) ∃v ∈ W1(X) with v − c < qε. 2 Since W1(X) ⊆ K and X is a tuple of 2 × 2 matrices, it follows that X ∈ W max (K) ⊆ W min (L), and X admits a normal dilation with joint spectrum in L ⊆ [0,∞)d. By design, the tuple X has the property that if i 6= j and T is a matrix with 0 ≤ T ≤ Xi and 0 ≤ T ≤ Xj, then T = 0. Therefore, by [21, Lemma 6.3], there is a possibly distinct normal dilation Z of X such that σ(Z) ⊆ L ∪ {0} = L and ZiZj = 0 for i 6= j. That is, σ(Z) ⊆ L consists of points which have at most one nonzero coordinate. From (2.31) and the definition of Π immediately thereafter, we have that σ(Z) ⊆ Π and X ∈ W min(Π). We then conclude from (2.32) that there is a point v ∈ W1(X) ⊆ W min (Π) = Π such that v − c < qε. Since we may repeat the procedure for arbitrarily small ε, and Π (which does not depend on ε) is closed, we have that c ∈ Π. Next, c was an arbitrary interior point of K, so the interior of 2 1 12 K is contained in Π. Finally, K is a convex body, so K is the closure of its interior, and it follows that K ⊆ Π. Theorem 2.29 directly generalizes the computations of dilation scale in [21, Theorem 6.4]. Moreover, it also implies [21, Theorem 8.8], as the technical condition given therein shows that K is a simplex-pointed set, ∆ is a simplex containing K that emanates from x in the same direction as the given vector data v1, . . . , vd, and ∆ is minimal among simplices which contain K. Thus, we find that the particularly precise perturbation method used in the proof of [21, Theorem 8.8] is ultimately not necessary. Applied to products of simplices, Theorem 2.29 implies that the estimates of Theorem 2.21 are optimal when each Ki is a simplex. Corollary 2.33. For each 1 ≤ i ≤ d, let ni ≥ 1 and let Ki be a simplex of dimension ni with 0 ∈ Ki. Then for positive scalars a1, . . . , ad, W max(cid:18) dQi=1 W min(cid:18) dQi=1 Ki(cid:19) = i ≤ 1. In particular, θ(cid:18) dQi=1 aiKi(cid:19) if and only if a−1 dQi=1W min (Ki) is contained in Ki(cid:19) = d. Proof. The forward direction is proved in Theorem 2.21. For the converse, we may assume Ki is the standard simplex by applying a linear transformation and restriction to a proper dPi=1 dQi=1 subspace, if necessary. Since Ki is then simplex-pointed at x = 0 with vector data given by the standard basis e1, . . . , en1+...+nd, Theorem 2.29 shows that if W max(cid:18) dQi=1 W min(cid:18) dQi=1 Ki(cid:19) ⊆ aiKi(cid:19), then the simplex L spanned by 0, a1e1, . . . , a1en1, a2en1+1, . . . , a2en1+n2, . . . , aden1+...+nd−1+1, . . . , aden1+...+nd must have Ki ⊆ L. This implies a−1 i ≤ 1 by consideration of the point (1, 0n1−1, 1, 0n2−1, . . . , 1, 0nd−1). dQi=1 dPi=1 3. Compactness and Minimality In this section, we primarily consider problems related to the matrix ranges of compact operator tuples. First, recall that Proposition 1.11 demonstrates that if T is an operator tuple with W(T ) ⊆ W min(K), then T has a normal dilation N with joint spectrum in K. Moreover, if T is a matrix tuple, then there exists a choice of N which is also a matrix tuple. However, no claim is made about the preservation of compactness if T acts on an infinite-dimensional space. Below we prove an approximate result in this direction. Proposition 3.1. Let T ∈ K(H)d be a tuple of compact operators acting on an infinite- dimensional space, with W(T ) ⊆ W min(K). Then for any ε > 0, there exists a compact normal dilation N of T such that N decomposes as a direct sum A ⊕ B with Ai < ε and σ(B) ⊆ (1 + ε)K. Proof. We may assume that T consists of compact self-adjoint operators. Since T acts on an infinite-dimensional space, 0 belongs to W1(T ) ⊆ K. Given δ > 0, let F1, . . . , Fd be self-adjoint finite rank operators with Ti − Fi < δ/2. Then if P is the finite-dimensional projection onto the sum of the kernels and cokernels of the Fi, it follows from conjugation by P that P FiP − P TiP = Fi − P TiP < δ/2. Therefore, Ti − P TiP < δ. 13 First, consider the tuple X = (T1 − P T1P, . . . , Td − P TdP ). The operators in the tuple are compact, but since P need not be a reducing subspace for T , W1(X) might not be contained in K. However, since Ti − P TiP < δ, the dilation technique of the second half of the proof of [8, Theorem 7.4], which preserves compactness of self-adjoint operators, produces a compact normal dilation R of X with Ri < dδ. Next, we dilate Y = (P T1P, . . . , P TdP ), which is unitarily equivalent to (M1⊕0, . . . , Md⊕ 0) for some matrix tuple M with M ∈ W(T ) ⊆ W min(K). Since M is a matrix tuple, it admits a normal matrix dilation with spectrum in K by Proposition 1.11, so Y admits a finite rank (hence compact) normal dilation S with σ(S) ⊆ K ∪ {0} = K. Finally, we combine the two dilations. The tuples R and S might act on different spaces, but this is easily remedied with the addition of zero summands, if necessary. Now, R + S is a dilation of T , but it might not be a normal tuple, as the Ri and Sj might not commute. Given a, b > 0 with 1 b = 1, let Q1 and Q2 be positive 2 × 2 matrices with entry 1 in the top left corner such that σ(Q1) = {0, a}, σ(Q2) = {0, b}, and Q1Q2 = Q2Q1 = 0, following (2.20). Let Ai = Ri ⊗ Q1 and Bj = Sj ⊗ Q2, so that AiBj = 0 = BjAi, and the orthogonal sum N = A + B is a normal dilation of T . We have that Ai < adδ for each i, and σ(B) ⊆ bK. Given ε > 0, we complete the proof by noting that we could have chosen b = 1 + ε, a = 1 a + 1 1−1/b , and δ = ε ad. Combining Proposition 3.1 with (2.2) (from [21, Theorem 6.7]), we find that the dilation of compact self-adjoint contractions may be bounded in the following sense. Corollary 3.2. Let (a1, . . . , ad) be a tuple of positive numbers such that given any tuple T ∈ K(H)d dilation N of compact self-adjoint operators with Ni ≤ ai. sa of compact self-adjoint contractions, there exists a normal Similarly, there are bounds on the dilation of compact contractions which are not neces- a−2 i < 1. Then dPi=1 sarily self-adjoint. Corollary 3.3. Let (a1, . . . , ad) be a tuple of positive numbers such that for some ε > 0, (1+ε)·(a1, . . . , ad) is an SD-tuple. Then given any tuple T ∈ K(H)d of compact contractions, there exists a normal dilation N of compact operators with Ni ≤ ai. It would be interesting to know if compact-to-compact dilation could be achieved without the approximation of spectrum used in Proposition 3.1. This would be useful even in the particular case of Corollary 3.2, as compact-to-compact dilation without perturbation of bounds would demonstrate that the use of Halmos dilation (3.4) X is a (self-adjoint) contraction =⇒ √I − X ∗X √I − XX ∗ −X ∗ (cid:19) is a (self-adjoint) unitary dilation U :=(cid:18) in [21, Theorem 6.7] is not optimal, as it immediately removes compactness. On the other hand, if W(T ) ⊆ W min(K), then for the diagonal operator tuple N with eigenvalues at all points of K, there is a UCP map sending Ni to Ti. If a compact normal dilation of T with joint spectrum in K exists, then we may find such a UCP map which also maps ∗-polynomials in N1, . . . , Nd to compact operators. Because the proof of Arveson's extension theorem relies on limits in a pointwise topology, not the norm topology, this may be too much to ask. X 14 M d ∞Sn=1 Below we consider a different sense of matrix approximation. Equip the matrix universe Md = n with a norm · that is decreasing under UCP maps. More precisely, equip each matrix level with a norm ·n, such that if φ : M d m is a UCP map and A ∈ M d n, then φ(A)m ≤ An. For example, we may take (A1, . . . , Ad) to be the sum of the operator norms of each Ai. With the norm · fixed, we may consider the Hausdorff n → M d n=1 Sn, which we abbreviate as follows. topology on subsets S =S∞ n n, so that S ≈ε T if and only if Sn ≈ε Tn for each n. n have W min(K) ≈ε W(A). Then there is a polyhedron L ⊆ K Definition 3.5. We write S ≈ε T if for every S ∈ S there exists T ∈ T with S − T < ε, and for every T ′ ∈ T , there exists S′ ∈ S with T ′ − S′ < ε. We apply the same notation for subsets of a fixed matrix level M d Proposition 3.6. Let A ∈ M d with at most 2n3(d + 1) + 1 vertices such that K ≈2ε L. Proof. By definition, we have that for any m ∈ Z+ and any tuple T ∈ W min m (K), there exists a UCP map φ : M d m such that φ(A) − T < ε. Similarly, since A ∈ Wn(A), there exists an n× n matrix tuple B ∈ W min (K) such that A− B < ε, which implies that φ(B)− T < 2ε. That is, any tuple in W min(K) may be approximated within 2ε by a tuple in W(B). Since we also have that W(B) ⊆ W min(K), it holds that W(B) ≈2ε W min(K). Let N be a normal dilation of B with joint spectrum in K, where we may suppose the mem- bers of N are matrices of dimension at most 2n3(d + 1) + 1 by Proposition 1.11. Then since W(B) ≈2ε W min(K) and W(B) ⊆ W(N) ⊆ W min(K), it holds that W(N) ≈2ε W min(K). Restricting to the first level yields W1(N) ⊆ K and W1(N) ≈2ε K. By normality of N, L := W1(N) is a polyhedron with at most 2n3(d + 1) + 1 vertices. n → M d When approximation is replaced by equality, the vertex count 2n3(d + 1) + 1 may be replaced by the more pleasant n. This may be deduced from [3], but we will present a proof which arises in pursuit of the following problem: if T ∈ K(H)d is a tuple of compact operators with W(T ) = W min(K), to what extent is the shape of K restricted by the fact that T is compact? In addition, if K is fixed, to what extent is T determined by K? Such results appeared to be within the scope of [8, §6], but one of our contributions here is the presentation of counterexamples to the claims therein. In response, the authors uploaded an arxiv correction [7] which addresses these examples with additional assumptions. In both versions, the arguments center around a minimality condition for operator tuples, as in [8, Definition 6.1]. We repeat that definition here. Definition 3.7. A tuple T = (T1, . . . , Td) ∈ B(H)d is said to be minimal, or minimal for its matrix range, if the restriction of T to any proper reducing subspace has strictly smaller matrix range. If one intends to determine T uniquely from its matrx range W(T ), the presence of some minimality condition (though not necessarily the one above) is natural. We show that even in the compact case, this particular condition is not sufficient to determine T from W(T ) up to unitary equivalence. Similarly, given a compact tuple T which is not minimal, there might not be a summand which is minimal for the same matrix range. Finally, a minimal compact tuple need not have the property that the unital C ∗-algebra it generates is isomorphic to the C ∗-envelope of the operator system it generates. All three results were claimed positively in [8], extending the uniqueness results of [3, 16, 24] for matrix ranges or free spectrahedra to 15 the compact setting. Therefore, our counterexamples show that Definition 3.7 is insufficient for consideration of compact tuples acting on infinite-dimensional spaces. The reader is invited to read [7, §6] and see how the definition of "nonsingularity" therein covers the non-pathological cases presented in this section (which motivated said definition). Recall that for a matrix convex set S, there are multiple relevant notions of extreme point. Given a matrix convex combination (3.8) X = V ∗ i Y iVi, mXi=1 Y i ∈ Sni, Vi : Cn → Cni, V ∗ i Vi = 1, mXi=1 one calls the point X ∈ Sn • an absolute extreme point of S if whenever each Vi : Cn → Cni is nonzero, it follows that each Y i contains a summand (possibly equal to Y i) which is unitarily equivalent to X. that each Y i is unitarily equivalent to X. • a matrix extreme point of S if whenever each Vi : Cn → Cni is surjective, it follows • a Euclidean extreme point of S if X is an extreme point of the convex set Sn in the usual sense. These definitions may be characterized in dilation-theoretic terms, as in [11, Theorem 1.1]. Note that at the scalar level n = 1, there is no difference between matrix extreme points and Euclidean extreme points, as surjectivity of the Vi forces ni = 1, in which case (3.8) reduces to a traditional convex combination. We will need the following lemma, which concerns the application of vector states to normal tuples. Lemma 3.9. Let K ⊂ Cd be a compact convex set and suppose λ is an extreme point of K. If N ∈ B(H)d is a normal tuple with σ(N) ⊆ K, and some v ∈ H has v = 1 and hNjv, vi = λj for each j, then Njv = λjv for each j. Proof. The joint reducing subspace of N generated by v is separable, so we may assume H is separable. Therefore H is a finite or countable direct sum of L2(µi) for regular Borel measures µi on σ(N), which we may extend to measures on K in the usual way. We may write Nj = ⊕Mπj for πj the jth coordinate function on σ(N), and v = (f1, f2, . . .) for fi ∈ L2(µi) with PR fi2 dµi = 1. We need to prove that πjfi = λjfi a.e. [µi]. Equivalently, if we let dνi = fi2 dµi and ν(E) = P νi(E), so ν is a probability measure, we need to prove that ZK i dµi =XhMπj fi, fii = hNjf, fi = λj = πj(λ). By definition, it holds that for each j ∈ {1, . . . , d}, πj dνi =XZK πj dν =XZK πj = λj a.e. [ν]. πj · f 2 Therefore, for the affine function system S = span{1, π1, . . . , πd} ⊂ C(K), ν is a representing measure for the point λ. Since λ is an extreme point of K, it is a Choquet boundary point of S, and the representing measure is unique: ν = δλ. Combined with the above equality R πj dν = λj, this says that πj = λj a.e. [ν]. It follows that πjfi = λjfi a.e. [µi] for each i and j, and the proof is complete. 16 In Lemma 3.9, the tuple N may act on an infinite-dimensional space. Even so, some manipulation of matrix ranges shows that Lemma 3.9 is actually equivalent to the claim λ is an extreme point of K =⇒ λ is an absolute extreme point of W min(K). (3.10) We leave the details of the equivalence between Lemma 3.9 and (3.10) to the reader, and we remark that (3.10) is certainly well-known. In particular, any minimal set W min(K) is spanned by its absolute extreme points. See also the general result [20, Corollary 6.12], which shows that if S is a matrix convex set spanned by its matrix extreme points from a fixed level Sn, then S is also spanned by its absolute extreme points. That result pairs quite nicely with [10, Corollary 1.1], which concerns precisely the opposite scenario -- there exists a family of matrix convex sets which have no absolute extreme points at all. The construction of this family relies heavily on the use of compact operators acting on infinite-dimensional spaces. We will use Lemma 3.9 in tandem with the following facts, which may be demonstrated by matrix computations. • If T is a d-tuple of matrices and λ is an extreme point of W1(T ), then there is a unit • If T is a d-tuple of compact operators and λ is a nonzero extreme point of W1(T ), In particular, we note that detection of the point 0 ∈ W1(T ) for T a compact tuple on an infinite-dimensional space might not be achieved using a vector state. For example, consider vector v with hTiv, vi = λi. then there is a unit vector v with hTiv, vi = λi. the single diagonal operator S = Ln∈Z+ 1 n . If T ∈ K(H1)d has W(T ) = W min(K), then T dilates to a normal tuple N ∈ B(H2)d with the same matrix range by Proposition 1.11, though compactness of T might be lost in the dilation. If v is a unit vector in H1, then by definition, the corresponding vector state gives the same result whether it is applied to T or to N. Further, if v is a joint eigenvector for N, then because v belongs to the smaller Hilbert space H1, it is also a joint eigenvector for T . We may use these facts to characterize when W(T ) = W min(K) for T a compact or matrix tuple. While we find that the shape of K is what one would expect if T were actually normal, T itself might be minimal and non-normal. n be a tuple of matrices with W(T ) = W min(K). Then every Theorem 3.11. Let T ∈ M d extreme point of K is a joint eigenvalue of T , and in particular K must be a polyhedron with at most n vertices. Proof. Since K = W1(T ) and T is finite-dimensional, if λ is an extreme point of K, there is a vector state v such that hTiv, vi = λi. Writing a normal dilation N of T , conv(σ(N)) = K, we have that similarly hNiv, vi = λi. By Lemma 3.9, since λ is extreme and detected by the vector state v, we conclude that Niv = λiv. Now, as v belongs to the Hilbert space on which T acts, we also have Tiv = λiv, i.e. v is a joint eigenvector for T . Finally, as T consists of n × n matrices, there are only up to n possible extreme points of K. Theorem 3.11 and its proof are listed for completeness, as whenever T is a matrix tuple, we may assume T is of the smallest dimension possible and apply any one of the various uniqueness results for free spectrahedra or matrix ranges of matrices. See, for example, [3, §1], [16, Theorem 3.12 and Proposition 3.17], [24, Theorem 1.2], and [7, Definition 6.3 and Theorem 6.9]. We now consider a version for compact operator tuples. 17 Theorem 3.12. Let H be a Hilbert space (of any finite or infinite dimension). If there exists a tuple T ∈ K(H)d of compact operators with W(T ) = W min(K), then every nonzero extreme point of K is a joint eigenvalue for T , and ext(K) is either a finite set or a sequence tending to zero. Proof. We begin by following the logic of the previous proof, noting that since T is a compact tuple, we may detect nonzero extreme points λ of K through vector states, λi = hTiv, vi. We similarly conclude that each such λ is a joint eigenvector of T using a normal dilation, Lemma 3.9, and a restriction. Since ext(K) \ {0} is contained in the set of eigenvalues of a compact operator tuple, it follows that ext(K) is either a finite set or a sequence tending to zero. Theorems 3.11 and 3.12 show that if a matrix tuple, or a compact tuple acting on an infinite-dimensional space, has W(T ) = W min(K), then the shape of K is "precisely what one would expect" from examination of normal tuples with the same properties. In particular, the theorems exhibit a decomposition T ∼= N ⊕ M where N is a normal tuple. However, we note that the qualification of nonzero extreme point in Theorem 3.12 is problematic, in that if K is a polyhedron with 0 as a vertex, we might have that W(N) is a proper subset of W min(K). In particular, we might not be able to find a minimal normal summand for the same matrix range. There are two distinct questions one can consider regarding minimality in this context, as in [8]. First, if a compact tuple T has matrix range W min(K), and T is minimal for its matrix range, is T determined up to unitary equivalence? Second, if T is not minimal, does it have a summand which is minimal for the same matrix range? Theorems 3.11 and 3.12 answer both questions affirmatively in most cases, but there are pathological examples concerning the point 0 ∈ W1(T ) when T acts on an infinite-dimensional space. First, we consider the question of uniqueness. Corollary 3.13. Let K ⊆ Cd be a compact convex set, and let H be a Hilbert space (of any finite or infinite dimension). If T ∈ K(H)d is minimal for its matrix range W(T ) = W min(K), then the following hold. (1) If K has infinitely many extreme points, then H is separable and infinite-dimensional, and T is diagonal with eigenvalues at the nonzero extreme points of K, which are isolated in the extreme points and form a sequence tending to zero. (2) If T acts on an n-dimensional space H, then K is a polyhedron with exactly n vertices, and T is diagonal with eigenvalues at the vertices of K. (3) If K is a polyhedron with n vertices, none of which are 0, then H must be n- dimensional, and T is diagonal with eigenvalues at the vertices of K. Proof. In cases (1) and (3), application of Theorem 3.12 shows that T admits eigenvectors for joint eigenvalues at each nonzero extreme point of K. The assumptions of either case show that the resulting normal summand N of T has W(N) = W min(K). By minimality of T , we have that N = T , so T is diagonal with the prescribed eigenvalues and the dimension of H is determined by the number of nonzero extreme points of K, which is either finite or a sequence tending to zero from Theorem 3.12. In case (2), we instead apply Theorem 3.11, so T has joint eigenvectors for each vertex of K (possibly including zero). By minimality, T must be equal to the resulting normal summand, so T is diagonal and the dimension of H is determined by the number of vertices n of K. 18 S1 = ∞Xn=1 1 3n Pen S2 = 1 3n Pvn ∞Xn=1 T1 = 1 0 S1  0 T2 = 1 S2  We note that if H is assumed finite-dimensional, then more general uniqueness results were proved in [3, 16, 24], either in terms of matrix ranges or free spectrahedra. Therefore, Corollary 3.13 is primarily of use to determine the shape of K based on T (or vice-versa) when H is infinite-dimensional, or to determine when the dimension of H must be finite or infinite based on other assumptions. In all cases of the "non-pathological" Corollary 3.13, T is normal, and in particular, T is diagonal with eigenvalues at the isolated extreme points of K. Later, in Theorem 3.26, we will demonstrate that regardless of compactness, all minimal normal tuples will take this form. In particular, if a normal tuple N is minimal for matrix range W(N) = W min(K), then the isolated extreme points of K must be dense in ext(K). However, not all minimal tuples are normal, even in the compact case. Indeed, the following example shows why item (3) of Corollary 3.13 must be only a partial converse to item (2). Example 3.14. Let K = span{(0, 0), (1, 0), (0, 1)} be the standard simplex in R2. Given an orthonormal basis {e1, e2, . . .} of an infinite-dimensional separable Hilbert space, let v1 = 1 √2 e3 + . . ., and extend v1 to an orthonormal basis {v1, v2, . . .} for the same space. Define the operators Si and Ti as follows. 1 √8 1 √4 e2 + e1 + Now, T = (T1, T2) ∈ K(H)2 sa is a compact tuple of positive operators whose numerical range includes (0, 0), (1, 0), and (0, 1), where we note that (0, 0) is detected by a limit of vector states. Moreover, we have 0 ≤ T1 + T2 ≤ I, so it follows that W1(T ) is precisely equal to K. Since K is a simplex, there is only one matrix convex set with K as its scalar level, and W(T ) = W min(K). By Theorem 3.12, R has joint eigenvectors for the eigenvalues (1, 0) and (0, 1). Since S1+S2 ≤ 2 3I, the only possible eigenvectors are those exhibited in the direct sum decomposition of T . Suppose eH is a reducing subspace for T such that the restriction R has W(R) = W min(K). Therefore, to show eH = H, we need only show that eH includes the entire domain of the summand S = (S1, S2). The intersection of eH to the domain of S is a reducing subspace for S, which we denote by L. The subspace L is nontrivial, as we must have (0, 0) ∈ W1(R), so L includes a vector x with hx, eni 6= 0 for some fixed n. Since S1 has distinct nonzero eigenvalues at the ei, manipulation of the functional calculus shows that this particular en belongs to L. However, hen, v1i 6= 0, so applying the same trick to the eigenvector basis of S2 shows that v1 ∈ L. Finally, for all m ∈ Z+, hv1, emi 6= 0, so examining S1 again shows and T is minimal for its matrix range. that em ∈ L for all m, and L is the entire domain of S. That is, S is irreducible, eH = H, Finally, note that by replacing 1 3np , p ∈ [1,∞), one can construct uncountably many examples of T , no two of which are unitarily equivalent. Corollary 3.15. Let K ⊆ Cd be a polyhedron with at least 3 vertices, one of which is 0, and fix a separable infinite-dimensional Hilbert space H. Then there are uncountably many unitarily inequivalent tuples T ∈ K(H)d such that W(T ) = W min(K) and T is minimal for its matrix range. 3n with 1 19 Proof. Let T have joint eigenvalues at all of the nonzero vertices, and produce an additional summand S of T which is determined from a selection of three vertices 0, v1, v2 and the technique of Example 3.14, after an invertible linear transformation. The above results may also be formulated in the language of operator systems, after we recall some additional notation and definitions. Definition 3.16. If T = (T1, . . . , Td) ∈ B(H)d, then we let ST denote the operator system (i.e., self-adjoint unital subspace of B(H)) generated by T1, . . . , Td. Given an operator system S, we let C ∗(S) denote the C ∗-algebra generated by S, which is necessarily unital. If S is an operator system inside a unital C ∗-algebra A, the particular structure of A is not generally relevant unless S generates A. It is usually easier to ignore A and write "S ⊆ C ∗(S) is an operator system," where it is important to note that the operator system structure of S alone might not determine the C ∗-algebra S generates. That is, it is possible for two operator systems S1 and S2 to be completely isometrically isomorphic while C ∗(S1) and C ∗(S2) are not isomorphic. However, given any concrete representation S ⊆ C ∗(S), there is a quotient of C ∗(S) that produces the "smallest" C ∗-algebra into which S embeds. Definition 3.17. Let S be an operator system. The Shilov ideal of S inside C ∗(S) is the largest ideal I such that the quotient C ∗(S) → C ∗(S)/I is completely isometric on S. The C ∗-envelope of S, denoted C ∗ e (S), is the quotient of C ∗(S) by the Shilov ideal. The existence of the Shilov ideal is a very deep result, and the structure of C ∗ e (S) does not depend on the choice of the initial concrete representation of S inside a C ∗-algebra (see [18] for additional background). We begin with a natural and well-known example. Example 3.18. Let T ∈ B(H)d be a tuple with W(T ) = W min(K). Then C ∗ e (ST ) is iso- morphic to the commutative C ∗-algebra C(ext(K)) of continuous complex-valued functions on ext(K). To see this, let Z = (Z1, . . . , Zd) denote the tuple of coordinate functions on ext(K). Since the convex hull of ext(K) is K, we have that W(Z) = W min(K) = W(T ), and there is a unital completely isometric map ST → SZ mapping Ti 7→ Zi by [8, Theorem 5.1]. e (SZ) is a quotient of C ∗(SZ) = C(ext(K)). If I is a nontrivial It follows that C ∗ ideal in C ∗(SZ), then C ∗(SZ)/I may be written as C(X) for a proper compact subset K of e (ST ) ∼= C ∗ ext(K). If eZ := (Z1 + I, . . . , Zd + I), then it follows that W(eZ) = W min(conv(X)), where conv(X) is a proper subset of K. In particular, W(eZ) 6= W(Z), so the quotient map by I is not completely isometric as a map SZ → S eZ by [8, Theorem 5.1]. We conclude that the Shilov ideal of SZ in C ∗(SZ) is trivial, and C ∗ e (ST ) ∼= C ∗ e (SZ) ∼= C(ext(K)). Because the C ∗-envelope computed above is certainly commutative, we find that each example found using the construction in Corollary 3.15 must have nontrivial Shilov ideal. Corollary 3.19. For any d ≥ 2, there exist uncontably many unitarily inequivalent tuples T ∈ K(H)d sa of compact self-adjoint operators such that T is minimal for its matrix range, but C ∗(ST ) 6∼= C ∗ Proof. In Corollary 3.15, the constructed tuple T has C ∗ e (ST ) isomorphic to the commutative C ∗-algebra of functions on the finite set ext(K). However, the operators Ti do not commute, so C ∗(ST ) cannot be isomorphic to C ∗ e (ST ). In particular, the Shilov ideal of ST in C ∗(ST ) is nontrivial. e (ST ), and the Shilov ideal must be nontrivial. 20 Since there are now numerous examples of minimal compact tuples which are not uniquely determined by the matrix ranges, even when the matrix ranges considered are of the form W min(K), we can consider relaxing the problem somewhat. One option is to replace mini- mality with a stronger condition, as in the following definition. Definition 3.20. A tuple T = (T1, . . . , Td) ∈ B(H)d is said to be fully compressed if the compression of T to any proper subspace of H has strictly smaller matrix range. Any tuple which is fully compressed is also minimal in the sense of Definition 3.7, but the reverse implication certainly does not hold. In fact, it is easy to see that if a compact tuple has matrix range W(T ) = W min(K), then the assumption that T is fully compressed allows us to uniquely determine T . That is, the counterexamples we consider in this section are no longer counterexamples in the new framework. Proposition 3.21. Suppose that S ∈ K(H1)d and T ∈ K(H2)d are fully compressed compact tuples with matrix range W min(K). Then S and T are unitarily equivalent. Proof. Since any fully compressed tuple is automatically minimal, Corollary 3.13 shows that S and T are unitarily equivalent if ext(K) is an infinite set, or if K is a polyhedron which does not have 0 as a vertex. If K is a polyhedron with 0 as a vertex, then any fully compressed compact tuple with matrix range W min(K) is of the form λ1 ⊕ · · · ⊕ λn ⊕ Q, where λ1, . . . , λn are the nonzero vertices of K and 0 ∈ W1(Q). If Q is anything other than (0, . . . , 0) ∈ Cd, there is a proper compression Q′ of Q with 0 ∈ W1(Q), and hence a proper compression of the original tuple with the same matrix range W min(K). It is conceivable that in the general setting, if T is a fully compressed tuple, then W(T ) determines T up to unitary equivalence. However, one should expect the closed and bounded matrix convex sets S which may be obtained as the matrix ranges of fully compressed tuples to be fairly restricted. We now consider a separate, but very much related problem which was examined in [8]. If T is compact with W(T ) = W min(K), does there exist a summand of T which is minimal for the same matrix range? For this problem, there is again an issue with the point 0 and detection by non-vector states, but the obstruction is far more elementary. Example 3.22. Let K 6= {0} be a polyhedron with vertices 0, v1, . . . , vn. Then the diagonal tuple N with joint eigenvalues v1, . . . , vn, 1 3vn, . . . is compact and normal with matrix range W(N) = W min(K). However, it is not minimal for its matrix range, and it admits no minimal summand with the same matrix range, as any summand with the same matrix range has infinitely many eigenvalues converging to zero. 2vn, 1 However, if one carefully avoids zero, then minimal summands are easy to pick out, as in the following corollary. Corollary 3.23. Let T ∈ K(H)d have matrix range W(T ) = W min(K). Then in any of the following circumstances, there exists a decomposition T ∼= N ⊕ M where N is minimal for the same matrix range. (1) H has finite dimension (so K is a polyhedron with at most dim(H) vertices), or (2) ext(K) is infinite (so dim(H) is infinite, and ext(K) is a sequence tending to 0), or (3) K is a polyhedron and 0 is not a vertex of K, or 21 (4) T has 0 as a joint eigenvalue. We may choose N diagonal with eigenvalues at isolated extreme points of K. For cases (1), (2), and (3), this is the only choice of N. For case (4), the choice might not be unique. Proof. For case (1), application of Theorem 3.11 shows that K is a polyhedron with at most dim(H) vertices and T ∼= N ⊕ M, where N is diagonal with eigenvalues for each vertex. For case (2), Theorem 3.12 shows that ext(K) is a sequence tending to 0, and we may write T ∼= N ⊕ M where N is diagonal with a single eigenvalue for each nonzero extreme point of K. Note in particular that in this case, every point of ext(K) \ {0} is isolated, and 0 is either not an extreme point, or it is an extreme point which is not isolated. In either case, 0 is not used as an eigenvalue of N, as it is not needed. N the diagonal operator with those eigenvalues. For case (3), Theorem 3.12 produces eigenvectors for each vertex of K, so T ∼= N ⊕ M for For case (4), we need only consider a polyhedron K with 0 as a vertex, as otherwise we may apply case (2) or (3). Since T by assumption has a joint eigenvector for 0, and Theorem 3.12 produces joint eigenvectors for the other vertices of K, we find that T ∼= N ⊕ M where N is a finite-dimensional diagonal tuple with an eigenvalue corresponding to each vertex of K. Finally, we consider uniqueness. In cases (1), (2), and (3), if M is another minimal summand of T with the same matrix range, then Theorems 3.11 and 3.12 produce the diagonal operator N as a summand of M, a contradiction. In case (4), however, it is possible that K is a polyhedron with vertices 0, v1, v2, . . . , vn, and that T is the direct sum of 0 and an operator S of the form in Corollary 3.15, which has each vi as a joint eigenvalue as well as an additional irreducible summand. We may choose a minimal summand N which is diagonal with eigenvalues at 0, v1, . . . , vn, or we may choose a minimal summand M = S. As in the comments after Theorem 3.11, we note that there is some overlap with Corollary 3.23 and results in which uniqueness results for matrix ranges or their polar duals apply. We also note that the pathology involving 0 disappears if one accepts summands of a tuple that is approximately unitarily equivalent to T . Corollary 3.24. Suppose T ∈ K(H)d is a tuple of compact operators with matrix range such that T ′ is approximately unitarily equivalent to T , T ′ decomposes as T ′ = S ⊕ X, and W(T ) = W min(K). Then there is a decomposition H ∼= eH ⊕ X and a tuple T ′ ∈ K(H)d S ∈ B(eH)d is minimal for the matrix range W(S) = W min(K). Proof. We may assume that H is infinite-dimensional. From Corollary 3.23, a minimal summand of T exists if K is a polyhedron with 0 6∈ ext(K), or if ext(K) is an infinite sequence tending to 0. The only remaining case is that K is a polyhedron which has 0 as a vertex. In this case, T decomposes as λ1 ⊕ · · · λn ⊕ Q, where the λi are the nonzero vertices of K and Q is some infinite-dimensional compact tuple (which may or may not have 0 as a joint eigenvalue). Since Q is approximately unitarily equivalent to 0 ⊕ Q, an operator T ′ approximately unitarily equivalent to T admits a finite-dimensional normal summand S with joint eigenvectors at every vertex of K. This summand S is therefore minimal. We now consider operator tuples which are not necessarily compact. A simple spectral theorem argument will show that minimal normal tuples for matrix range W min(K) exist if and only if K satisfies a geometric condition on its extreme points: the isolated extreme 22 points of K are dense in ext(K). The condition also implies uniqueness of minimal normal tuples with W(T ) = W min(K). However, the same exact condition also guarantees there are many non-normal minimal tuples with the same matrix range. Given a compact convex set K ⊆ Cd, let IK = {x ∈ ext(K) : ∃V ⊂ Cd open such that V ∩ ext(K) = {x}} be the set of isolated extreme points, and note that we may replace either instance of ext(K) (or both instances) with ext(K) without changing the result. Further, if r > 0 is fixed, then there are only finitely many x ∈ ext(K) which satisfy x − y ≥ r for all y ∈ ext(K) \ {x}, so Ik is a finite or countably infinite set. Moreover, a point x ∈ K is in IK if and only if the closed convex hull of ext(K) \ {x} is a proper subset of K. Lemma 3.25. Let N ∈ B(H)d be a normal tuple which is minimal for its matrix range W(N) = W min(K). Then σ(N) = ext(K). Proof. Since W(N) = W min(K), we have that conv(σ(N)) = K by Proposition 1.10. Note that a closure of the convex hull is not needed. First, suppose that ext(K) 6⊆ σ(N). Since σ(N) is closed, it follows that there is an extreme point x of K which is missing from σ(N). From conv(σ(N)) = K we find that x is nontrivial convex combination of points from σ(N) ⊆ K, a contradiction. Next, assume that ext(K) is properly contained in σ(N). Let V be an open set in Cd such that V ∩ ext(K) = ∅ and there exists y ∈ σ(N) ∩ V . Letting (π1, . . . , πd) denote the coordinate functions on σ(N), there is a spectral measure E defined on Borel subsets of σ(N) with Ni =Zσ(N ) πi dE, Li =Zσ(N ) πi · Iσ(N )\V dE, and the support of E is exactly σ(N). Since V is open and intersects σ(N), it follows that E(V ∩σ(N)) is a nonzero projection, and the set of Borel functions which vanish on V ∩σ(N) is a proper, nonzero reducing subspace of each Ni. The restriction to the complement, i.e. It follows that has joint spectrum which includes all of ext(K), since V ∩ ext(K) = ∅. conv(σ(L)) = K and W(L) = W min(K), a contradiction of the minimality of N. Theorem 3.26. Let K be a nonempty compact convex subset of Cd. Then there is a normal, minimal tuple N with W(N) = W min(K) if and only if IK = ext(K). Moreover, in this case, the only such N is diagonal with eigenvalues at each λ ∈ IK. Proof. Step I. Assume IK = ext(K). Let D be diagonal with joint eigenvalues from the countable set IK, so σ(D) = IK = ext(K) and W(D) = W min(K). Any projection that reducing subspace for D, then the restriction R misses an eigenvalue x ∈ IK. Since x is isolated, the convex hull of σ(R) does not equal K, and therefore W(R) 6= W min(K). We conclude that D is minimal, and in particular a normal minimal tuple for matrix range W min(K) exists. If N is any other minimal normal tuple for the same matrix range, then by Lemma 3.25, σ(N) = ext(K), so every point x ∈ IK is an atom of the spectral measure. commutes with D corresponds to a direct sum of eigenspaces for D, so if eH is a proper 23 *E Oc ∪ Vn! z, z+ =*E ∞[n=1 ∞[n=1 Vn! z, z+ ≤ ∞Xn=1 hE(Vn)z, zi ≤ 1 2n+1 = 1 2 . ∞Xn=1 It follows that N has eigenvalues at each x ∈ IK, and D is a summand of N with the same matrix range. By minimality, N must equal D. Step II. Suppose there exists a minimal normal N, but IK is a proper subset of ext(K). We know from Lemma 3.25 that σ(N) = ext(K). Let O be an open set in Cd such that O ∩ ext(K) 6= ∅ and O ∩ IK = ∅, and let E be the spectral measure that represents N. Case (a). Suppose there is a point x ∈ O ∩ ext(K) with E({x}) 6= 0. Since x is not an isolated extreme point, it follows that E(ext(K)\{x}) is a projection onto a proper reducing subspace of N whose restriction has the same spectrum, a contradiction of minimality. Case (b). Suppose that for every x ∈ O ∩ ext(K), E({x}) = 0. Since the support of E is σ(N) = ext(K), and O is an open set which includes some extreme points, it follows that E(O) must be a nontrivial projection. Let z be a unit vector in its range, so hE(O)z, zi = 1. Also let xn ∈ O ∩ σ(N) form a sequence which is dense in O ∩ σ(N). Because E({xn}) = 0, there is an open neighborhood xn ∈ Vn ⊂ O such that hE(Vn)z, zi ≤ 1 2n+1 holds, and Therefore, P := E(cid:18)Oc ∪ Vn(cid:19) is a proper nonzero projection. The spectrum of NRan(P ) contains at least {x1, x2, . . .} ∪ (σ(N) \ O) by design, and the closure of the union is σ(N). This contradicts the minimality of N. ∞Sn=1 The condition IK = ext(K) also allows us to expand Corollary 3.15 to the non-compact setting (without the need to assume K is a polyhedron). That is, the condition which characterizes existence and uniqueness of minimal normal tuples also guarantees the existence of a plethora of non-normal minimal tuples for the same matrix range W min(K). Corollary 3.27. Let K be a compact convex set with at least three extreme points, such that IK = ext(K). Then there are uncountably many unitarily inequivalent tuples T such that T is minimal for matrix range W(T ) = W min(K). For such T , the Shilov ideal of ST in C ∗(ST ) is trivial if and only if T is normal. Proof. Let IK = {v1, v2, v3, . . .}. Following an affine transformation of Corollary 3.15, choose T to be the direct sum of v2, v3, . . . and an irreducible tuple S. The summand S is chosen such that W1(S) is contained in the simplex conv[v1, v2, v3] and v1 ∈ W1(S), but W1(S) does not include any extreme points of K besides v1. The uncountably many options for T are all minimal for matrix range W(T ) = W min(K). Next, let T be any minimal tuple for matrix range W(T ) = W min(K), and let N be a diagonal operator with eigenvalues at IK, which is dense in ext(K). The C ∗-envelope of ST is isomorphic to C(ext(K)), and in particular is commutative. Therefore, if T is not normal, the Shilov ideal of ST in C ∗(ST ) is nontrivial. However, if T is normal, then by Theorem 3.26, T is unitarily equivalent to N. Finally, we verify that the Shilov ideal of N in SN is trivial, which follows from the fact that C ∗(SN ) ∼= C(ext(K)). In particular, any quotient of C ∗(SN ) by a nontrivial ideal I gives rise to a normal tuple (M1, . . . , Md) = (N1 + I, . . . , Nd + I) whose joint spectrum σ(M) is strictly smaller than σ(N) = ext(K). It follows that W1(M) = conv(σ(M)) is a proper subset of W1(N) = K, and the unital map 24 sending Ni 7→ Mi is certainly not completely isometric. That is, the nontrivial ideal I cannot be the Shilov ideal. We also produce another pathological example using the simplex. Consider the universal C ∗-algebra (3.28) Ad := C ∗(x1, . . . , xd xi = x∗ i , x2 i = 1, xixj = −xjxi for i 6= j), which has played a major role, under various guises, in previous problems concerning matrix convex sets and free spectrahedra [17, 8, 21]. The generators may be realized as a tuple F [d] = (F [d] 1 , . . . F [d] 1 = 1 and the following recursive identities for d ≥ 2: d ) of 2d−1 × 2d−1 matrices defined by F [1] F [d] j := F [d−1] j 0 ⊗(cid:18)1 0 −1(cid:19) , 1 ≤ j ≤ d − 1, := I2d−2 ⊗(cid:18)0 1 1 0(cid:19) . F [d] d i Anticommutation of the F [d] 1 + . . . + λdF [d] implies that for any real d-tuple λ with λℓ2 = 1, it follows that λ1F [d] d ≤ 1. In particular, W1(F [d]) is contained in the closed ℓ2 unit ball Bd, so F [d] ∈ W max(Bd). Elementary computations show that the unit sphere Sd−1 is contained in W1(F [d]), which then implies that W1(F [d]) = Bd. Theorem 3.29. Let K be a compact convex set with at least 3 extreme points. Then there is a tuple A ∈ B(H)d such that W(A) = W min(K), A has no nontrivial normal summands, and any summand B of A such that W(B) = W(A) has the property that B is not minimal for its matrix range. Proof. We may assume that K is a convex body in Rd, d ≥ 2, and construct a tuple of self-adjoint operators. For k ∈ Z+, choose xk ∈ Rd and ck ∈ (0,∞) such that the collection {x1 + c1Bd, x2 + c2Bd, . . .} of ℓ2-balls has the following properties. For each k, there exists a simplex Sk with xk + ckBd ⊂ Sk ⊆ int(K), and for i 6= j, the balls xi + ciBd and xj + cj · Bd do not intersect. Finally, every λ ∈ ext(K) is a limit point of the union of all the balls. d ), so that W1(A(k)) = xk + ckBd, and let A(k). Then W(A) is a closed matrix convex set whose first level has ext(K) ⊆ W1(A), A = which shows that K ⊆ W1(A) and W min(K) ⊆ W(A). On the other hand, because Sk is a simplex, we have that each A(k) admits a normal dilation with spectrum inside Sk. Therefore, A admits a normal dilation with spectrum inside K, and W(A) ⊆ W min(K) holds. Finally, W(A) = W min(K). C2d−1 which is reducing for A. Let P denote the projection onto K and write P in block form [Pij] with Pij ∈ M2d−1. Fix any i 6= j, so that W1(A(i)) and W1(A(j)) are disjoint compact convex sets, meaning there is a hyperplane which separates them. We may therefore fix constants b1, . . . , bd such that the self-adjoint operators Let A(k) be the tuple xk + ck · (F [d] ∞Lk=1 Suppose K is a nontrivial subspace of 1 , . . . , F [d] ∞Li=1 T (k) := bmA(k) m dXm=1 25 have W1(T (i)) and W1(T (j)) disjoint, which gives that T (i) and T (j) have disjoint spectrum. Letting T = bmAm, dXm=1 we have that σ(T ) ⊂ R is a compact set which includes the spectrum of any summand T (k). Fix a continuous function f on σ(T ), which we may apply using the functional calculus, such that f (T (i)) = 0 and f (T (j)) is invertible. Since P commutes with T , we have that P f (T ) = f (T )P , which in block form shows that Pijf (T (j)) = f (T (i))Pij. By the choice of f , we have that Pij = 0. Therefore, P is a direct sum of projections Pii ∈ M2d−1 corresponding to reducing subspaces of A(i). Since every reducing subspace K of A is a direct sum of reducing subspaces for the A(i), it follows that if A has a nontrivial normal summand, then there is some A(i) which has a nontrivial normal summand as well. Since (F [d] d ) has no nontrivial normal summand, this is impossible. Similarly, if B is a summand of A with the same matrix range W min(K), then B is a direct sum of A(i)Ki for reducing subspaces Ki of A(i). Since each set W1(A(i)) is contained in the interior of K, it follows that B must have infinitely many summands, with detection of the extreme points of K in W1(B) unaffected by the removal of one summand. Finally, B is not minimal for its matrix range W(B) = W min(K). 1 , . . . , F [d] 4. Scaled Containments In this section, we consider two matrix convex set containments which may be demon- strated by explicit dilation procedures. We first consider the problem of dilating tuples T ∈ B(H)d of (not necessarily self-adjoint) contractions to normal tuples N ∈ B(K)d such that Ni ≤ C for each i. Recall from (2.17) that dYj=1 (4.1) W min(D) = {T ∈ Md : for each i ∈ {1, . . . , d},Ti ≤ 1} and that an abstract dilation result for contractions can be found in Corollary 2.18, in the language of SD-tuples. Below we show that the constant C = √2d can be achieved explicitly. Lemma 4.2. Suppose S ∈ B(H)d satisfies 1 + . . . + SdS∗ Then there is a normal dilation M of S with Mi ≤ √2d and MiMj = 0 for i 6= j. 1S1 + . . . + S∗ d = I = S∗ S1S∗ dSd. Proof. Write Sj = Xj + iYj, so that the two identities given imply that X 2 1 + Y 2 1 + . . . + X 2 d + Y 2 d = I and [X1, Y1] + [X2, Y2] + . . . + [Xd, Yd] = 0. From [21, Theorem 6.6] and Proposition 1.11, there is a normal dilation (Q1, R1, . . . , Qd, Rd) of (X1, Y1, . . . , Xd, Yd) with joint spectrum in the extreme points of √2d ·⋄2d. Therefore, the self-adjoint operators Qi and Rj have norm √2d and satisfy QiRj = 0 for all i, j and QiQj = 0 = RiRj if i 6= j. It follows that (S1, . . . , Sd) has a normal dilation (M1, . . . , Md) = (Q1 + iR1, . . . , Qd + iRd) such that Mi = √2d and MiMj = 0 for i 6= j. 26 Remark 4.3. For d = 2, this estimate cannot be improved. Consider the elementary 2 × 2 matrices S1 = E12 and S2 = E21 = S∗ If a normal 1 , which meet the identities required. dilation (M1, M2) has Mj ≤ r and M1M2 = 0, then it follows that M1 + M ∗ 2 ≤ r. Since 2 = 2S1, which has norm 2, this implies that r ≥ 2 = √2d. M1 + M ∗ Theorem 4.4. Suppose T ∈ B(H)d is a tuple of (not necessarily self-adjoint) contractions. Then there is a normal dilation N of T with Ni ≤ √2d for each i. It follows that 2 is a dilation of S1 + S∗ (4.5) W min(D) ⊆ The constant √2d is not necessarily optimal. Proof. For convenience, we label the operators in T as T0, . . . , Td−1. Use Halmos dilation (if necessary) to obtain a dilation tuple U = (U0, . . . , Ud−1) where each Ui is unitary. Let ω be a primitive dth root of unity, and define the averages √2d · W min(cid:16)D d(cid:17) . dYj=1 (4.6) Sj := 1 d ωjkUk. d−1Xk=0 A simple computation using the identity 1 + ωr + . . . + ωr(d−1) = 0 for ωr 6= 1 shows that d−1Xj=0 SjS∗ j = I = S∗ j Sj. d−1Xj=0 ω−jnMn By Lemma 4.2, there is a normal dilation M of S with Mj ≤ √2d and MiMj = 0 for i 6= j. Moreover, it follows from (4.6) that Nj := d−1Xn=0 is a dilation of Uj (and hence also of Tj). Since MiMj = 0 for i 6= j, we have that Nj ≤ √2d, and N = (N0, . . . , Nd−1) is a normal tuple. Corollary 4.7. Suppose T ∈ K(H)d is a tuple of (not necessarily self-adjoint) compact contractions. Then for any ε > 0, there is a normal tuple N consisting of compact operators such that N is a dilation of T and Ni ≤ √2d + ε for each i. d(cid:17), which follows from Proof. Apply Proposition 3.1 to the claim W(T ) ⊆ W min(cid:16)√2d · D The constant √2d in Theorem 4.4 strictly improves the constant min{d, 2√d} from [21, Corollary 6.11] when d ≥ 3, but for d = 2, the two constants are equal. While Lemma 4.2 is optimal when d = 2, the proof of Theorem 4.4 does not use the full strength of the lemma. Namely, the final step of the proof (when d = 2) seeks to show that N2 := M1 − M2 N1 := M1 + M2 Theorem 4.4. are normal operators which commute and have some norm bound Ni ≤ C. Lemma 4.2 in the case d = 2 shows that C = √2 · 2 = 2 can be obtained, with the stronger condition that the building blocks M1 and M2 are normal operators with M1M2 = 0 = M2M1. Thus, and 27 we cannot necessarily conclude that the constant in Theorem 4.4 is optimal. We remind the reader that due to the estimate in Corollary 2.18, knowledge of the optimal constant in Theorem 4.4 also produces a bound on U(d), which might improve the bound (2.27). As noted in (4.1), the set S of all d-tuples of matrix contractions is equal to dQj=1W min(D), d(cid:17). In fact, the containment (4.5) holds and in particular, it is strictly smaller than W max(cid:16)D even though the larger scale 2√d in W max(cid:16)D d(cid:17) ⊆ 2√d · W min(cid:16)D d(cid:17) λ ∈ Cd, λ1 + . . . + λd ≤ 1 is optimal by [21, Corollary 6.11]. However, while S is not a maximal matrix convex set, S is trivially equal to the set of matrix tuples T ∈ Md such that (4.8) The scalar tuple λ is selected from the complex ℓ1 ball, which is dual to the complex ℓ∞ ball . Therefore, (4.8) may be considered a C-linear analogue of the real inequalities (1.9) that D characterize sets of the form W max(K). Further, since each Sn is closed under multiplication by n × n unitary matrices, S is a free circular matrix convex set in the sense of [12, §1]. We now pursue another matrix convex set containment through explicit dilation. As in (3.28), consider the universal C ∗-algebra λ1T1 + . . . + λdTd ≤ 1. =⇒ d i = 1, xixj = −xjxi for i 6= j) i , x2 Ad := C ∗(x1, . . . , xd xi = x∗ (4.9) and the concrete realization of the generators in the tuple F [d] = (F [d] j are 2d−1 × 2d−1 matrices defined by F [1] F [i] d ≥ 2: ⊗(cid:18)1 0 −1(cid:19) , 1 ≤ j ≤ d − 1, := I2d−2 ⊗(cid:18)0 1 1 0(cid:19) . := F [d−1] F [d] d F [d] j 0 j d ), where the 1 = 1 and the following recursive identities for 1 , . . . F [d] If Bd denotes the closed ℓ2 unit ball in Rd, then [17, Proposition 14.14] (adjusted to the self-adjoint complex setting) shows that the free spectrahedron 1 + X2 ⊗ F [2] DF [2] :=nX ∈ M2 : X1 ⊗ F [2] DF [2] = W min(B2). 2 ≤ Io satisfies (4.10) Two proofs are given in [17, Proposition 14.14], one of which uses an explicit dilation proce- dure. The polar dual (see [8, §3] for details) of (4.10) is the equivalent expression (4.11) W(F [2]) = W max(B2). It is not clear how to extend the techniques of [17, Proposition 14.14] in order to prove the extension of (4.10) or its dual (4.11) to d > 2. We will focus on providing explicit dilation evidence for the dual formulation (4.12) W(F [d]) ? = W max(Bd), 28 noting that the containment ⊆ is trivial. Moreover, while the polar dual may allow one to switch between two equivalent problems, we note that explicit dilation information does not generally survive applying the polar dual. Since the members of F [d] are generators of the universal C ∗-algebra Ad, application of Stinespring factorization shows that a tuple T ∈ Md is in W(F [d]) if and only if there is a dilation A of T consisting of self-adjoint unitiaries Ai such that AiAj = −AjAi for i 6= j. Thus, a proof that (4.12) holds would imply that the existence of such a dilation A is characterized by the satisfaction of linear inequalities by T . We will provide some evidence for (4.12) by using more restrictive linear inequalities, which place W1(T ) in a rectangular prism inside the ball. Theorem 4.13. Let X = (X1, . . . , Xd) ∈ B(H)d sa be a d-tuple of self-adjoint contractions, and define AC(X) ⊆ B(H)sa as the collection of all self-adjoint operators which anticommute a−2 with each Xj. j ≤ 1, then there exists a dilation tuple A ∈ B(H ⊗ C4d−1)d sa of X with the following properties. If a1, . . . , ad > 0 satisfy dPj=1 • For each j, Aj ≤ aj. • For j 6= k, Aj and Ak anticommute. 0 • For each W ∈ AC(X), W ⊗(cid:18)1 0 −1(cid:19)⊗2(d−1) anticommutes with every Aj. Gj ≤r12 + 1 r2 = aj bj , 1 ≤ j ≤ d − 1, 29 √12 + r2 = ad. E ≤ i i 0 with a−2 0 −W(cid:19) = W ⊗(cid:18)1 • The block entries of each Aj are in the real unital C ∗-algebra generated by X1, . . . , Xd. Proof. The case d = 1 is trivial, as no dilation is necessary. We proceed by induction: suppose the theorem holds for d − 1. Given a tuple X ∈ B(H)d sa of self-adjoint contractions, p1 − X 2 consider first the Halmos dilations Yi = (cid:18) Xi −Xi (cid:19) ∈ B(H ⊗ C2)sa, which p1 − X 2 anticommute with(cid:18)W 0 0 −1(cid:19) for each W ∈ AC(X). Given a1, . . . , ad > 0 j ≤ 1, let ad = √1 + r2, and define bj = aj√1+1/r2 for 1 ≤ j ≤ d − 1, so that b−2 j ≤ 1. Make intermediate dilations as follows: dPj=1 Gj := commutes with the off-diagonal term, so applying the C ∗-norm identity A =pAA∗ to  , 1 ≤ j ≤ d − 1, so that E anticommutes with G1, . . . , Gd−1. Within each operator, the diagonal term anti- E :=(cid:18) Yd −rI −rI −Yd(cid:19) , these self-adjoint operators shows that YjYd + YdYj YjYd + YdYj Yj 2r 2r −Yj d−1Pj=1 Moreover, since Y1, . . . , Yd anticommute with W ⊗(cid:18)1 G1, . . . , Gd−1 anticommute with W ⊗(cid:18)1 0 −1(cid:19)⊗2 {E} ∪(W ⊗(cid:18)1 0 −1(cid:19)⊗2 (4.14) 0 0 0 0 −1(cid:19) for W ∈ AC(X), it follows that . That is, : W ∈ AC(X)) ⊆ AC(G). bj 0 Apply the inductive assumption (scaled by aj bj ) to the tuple G ∈ B(H ⊗ C4)d−1 sa , the collection AC(G), and the scalars b1, . . . , bd−1. It follows that there exist pairwise anticom- muting dilations A1, . . . , Ad−1 ∈ B(H ⊗ C4 ⊗ C4d−2)sa = B(H ⊗ C4d−1)sa of G1, . . . , Gd−1 with norm Aj ≤ aj · bj = aj and with block entries in the real unital C ∗-algebra generated by G1, . . . , Gd−1. Moreover, for each V ∈ AC(G), A1, . . . , Ad−1 anticommute with V ⊗(cid:18)1 0 −1(cid:19)⊗2(d−2) . By (4.14), for each W ∈ AC(X), A1, . . . , Ad−1 anticommute with W ⊗(cid:18)1 0 −1(cid:19)⊗2 ⊗(cid:18)1 = W ⊗(cid:18)1 E ⊗(cid:18)1 0 −1(cid:19)⊗2(d−2) A1, . . . , Ad−1. Finally, since E anticommutes with W ⊗(cid:18)1 it follows that Ad anticommutes with W ⊗(cid:18)1 0 −1(cid:19)⊗2(d−1) 0 −1(cid:19)⊗2(d−1) 0 −1(cid:19)⊗2 , then Ad = E ≤ aj, and (4.14) shows that Ad anticommutes with as an examination of the intermediate dilations shows that the block entries of Ad belong to the real unital C ∗-algebra generated by the Xi. 0 −1(cid:19)⊗2(d−2) for each W ∈ AC(X), . The inductive step is complete, . Similarly, if Ad := 0 0 0 0 0 0 j = a2 sa be a tuple of pairwise anticommuting self-adjoint op- sa consisting of pairwise j I. Moreover, we may choose Mj such We note that the operators Aj constructed in Theorem 4.13 have Aj = aj, but they do not necessarily satisfy A2 j I. This may be remedied by a "step-by-step" modification of the Halmos dilation procedure, designed to preserve pairwise anticommutation. We include the details for completeness. Proposition 4.15. Let A ∈ B(H)d erators with Aj ≤ aj. Then there is a dilation M ∈ B(H ⊗ C2d)d anticommuting self-adjoint operators Mj with M 2 2dLi=1 pa2 A1 1I − A2 as if X anticommutes with A1, then X ⊕ X anticommutes with M1. Therefore, we may induct, so we assume the result holds for d − 1. sa is a tuple of pairwise anticommuting self-adjoint operators with Aj ≤ aj, and we let B = (A1, . . . , Ad−1), then B admits a dilation L ∈ B(H ⊗ C2d−1)sa with L2 j I and LjLk = −LkLj for j 6= k. Proof. The case d = 1 follows from the Halmos dilation M1 =(cid:18) that for every X ∈ B(H)sa that anticommutes with A1, . . . , Ad, M1, . . . , Md. If A ∈ B(H)d 1I − A2 −A1 X anticommutes with pa2 j = a2 (cid:19), j = a2 1 1 30 Moreover, we may choose Lj such that if X anticommutes with A1, . . . , Ad−1, then anticommutes with L1, . . . , Ld−1. In particular, this applies to Ad, and we may define X 2d−1Li=1 Mj :=(cid:18)Lj 0 −Lj(cid:19) , 1 ≤ j ≤ d − 1, Md+1 := 0  2d−1Li=1 sa2 dI − Ad 2d−1Li=1 sa2 A2 d dI − 2d−1Li=1 2d−1Li=1 Ad .  A2 d − It follows that M 2 j = a2 j I for all j, MjMk = −MkMj if j 6= k, and if X anticommutes with A1, . . . , Ad, then X anticommutes with M1, . . . , Md. 2dLj=1 Finally, we may place a lower bound on the matrix range of the tuple F [d], which consists of (universal) pairwise anticommuting, self-adjoint unitaries. it follows that B := ( 1 c1 j ≤ 1, then W max dQj=1 [−cj, cj]! ⊆ W(F [d]). Corollary 4.16. If c1, . . . , cd ≥ 0 satisfyP c2 Proof. We may assume cj > 0, as W(F [d]) is closed. Given a tuple X ∈ W max dQj=1 [−cj, cj]!, dPj=1(cid:16) 1 cj(cid:17)−2 dPj=1 c2 j ≤ 1, Theorem 4.13 and Proposition 4.15 shows that B admits a dilation M consisting of pairwise anticommuting self-adjoint operators with M 2 I for each j. Rescaling by cj shows that X admits a dilation L consisting of pairwise anticommuting self-adjoint uni- taries. This implies that L1, . . . , Ld satisfy the relations of (4.9), and hence there is a unital ∗-homomorphism F [d] 7→ Lj. Composition with a compression shows that there is a UCP map F [d] j Xd) is a tuple of self-adjoint contractions. Since X1, . . . , 1 cd j = 1 c2 j 7→ Xj, and finally X ∈ W(F [d]). j = With (4.11) and Corollary 4.16, it is within the realm of possibility that (4.12) holds in full generality. We conclude by noting that Ad := C ∗(x1, . . . , xd xi = x∗ i , x2 i = 1, xixj = −xjxi for i 6= j) is but one of many universal C ∗-algebras that produces potential scaled containments 1 C · W max(K) ⊆ W(x1, . . . , xd) that may be posed as (noncommutative) dilation problems through Stinespring factorization. It would be of great interest to the author if, in addition to a resolution of (4.12), there were a general method by which one could compute or bound the optimal scale C, depending on the universal C ∗-algebra and the relationship between K and W1(x1, . . . , xd). 31 Acknowledgments I am grateful to Orr Shalit and Adam Dor-On for their comments, and to the referee for sig- nificant improvements (especially regarding section 3 and alternative notions of minimality). References [1] William Arveson. Subalgebras of C ∗-algebras. Acta Math., 123:141 -- 224, 1969. 2 [2] William Arveson. Subalgebras of C ∗-algebras. II. Acta Math., 128(3-4):271 -- 308, 1972. 2 [3] William Arveson. The noncommutative Choquet boundary III: operator systems in matrix algebras. Math. Scand., 106(2):196 -- 210, 2010. 15, 17, 19 [4] William Arveson. The noncommutative Choquet boundary II: hyperrigidity. Israel J. Math., 184:349 -- 385, 2011. 1, 2 [5] Man Duen Choi. Completely positive linear maps on complex matrices. Linear Algebra and Appl., 10:285 -- 290, 1975. 2 [6] Man Duen Choi and Edward G. Effros. Injectivity and operator spaces. J. Functional Analysis, 24(2):156 -- 209, 1977. 1 [7] Kenneth R. Davidson, Adam Dor-On, Orr Moshe Shalit, and Baruch Solel. Dilations, inclusions of matrix convex sets, and completely positive maps. Corrected version in arxiv:1601.07993v3. 5, 15, 16, 17 [8] Kenneth R. Davidson, Adam Dor-On, Orr Moshe Shalit, and Baruch Solel. Dilations, inclusions of matrix convex sets, and completely positive maps. Int. Math. Res. Not. IMRN, (13):4069 -- 4130, 2017. 2, 3, 5, 6, 7, 10, 14, 15, 18, 20, 21, 25, 28 [9] Edward G. Effros and Soren Winkler. Matrix convexity: operator analogues of the bipolar and Hahn- Banach theorems. J. Funct. Anal., 144(1):117 -- 152, 1997. 2 [10] Eric Evert. Matrix convex sets without absolute extreme points. Linear Algebra Appl., 537:287 -- 301, 2018. 3, 17 [11] Eric Evert, J. William Helton, Igor Klep, and Scott McCullough. Extreme points of matrix convex sets, free spectrahedra and dilation theory. arxiv:1612.00025. 3, 16 [12] Eric Evert, J. William Helton, Igor Klep, and Scott McCullough. Circular free spectrahedra. J. Math. Anal. Appl., 445(1):1047 -- 1070, 2017. 28 [13] Tobias Fritz, Tim Netzer, and Andreas Thom. Spectrahedral containment and operator systems with finite-dimensional realization. SIAM J. Appl. Algebra Geom., 1(1):556 -- 574, 2017. 4 [14] Paul R. Halmos. Normal dilations and extensions of operators. Summa Brasil. Math., 2:125 -- 134, 1950. 9 [15] Masamichi Hamana. Injective envelopes of C ∗-algebras. J. Math. Soc. Japan, 31(1):181 -- 197, 1979. 1 [16] J. William Helton, Igor Klep, and Scott McCullough. The matricial relaxation of a linear matrix in- equality. Math. Program., 138(1-2, Ser. A):401 -- 445, 2013. 15, 17, 19 [17] J. William Helton, Igor Klep, Scott McCullough, and Markus Schweighofer. Dilations, linear matrix inequalities, the matrix cube problem and beta distributions. arxiv:1412.1481. 25, 28 [18] Evgenios T.A. Kakariadis. Notes on the C ∗-envelope and the Silov ideal. 20 [19] M. Krein and D. Milman. On extreme points of regular convex sets. Studia Math., 9:133 -- 138, 1940. 3 [20] Tom-Lukas Kriel. An introduction to matrix convex sets and free spectrahedra. arxiv:1611.03103. 17 [21] Benjamin Passer, Orr Moshe Shalit, and Baruch Solel. Minimal and maximal matrix convex sets. J. Funct. Anal., 274(11):3197 -- 3253, 2018. 3, 4, 6, 11, 12, 13, 14, 25, 26, 27, 28 [22] E. Steinitz. Bedingt konvergente Reihen und konvexe Systeme. J. Reine Angew. Math., 146:1 -- 52, 1916. 3 [23] W. Forrest Stinespring. Positive functions on C ∗-algebras. Proc. Amer. Math. Soc., 6:211 -- 216, 1955. 3 [24] Aljaz Zalar. Operator Positivstellensatze for noncommutative polynomials positive on matrix convex sets. J. Math. Anal. Appl., 445(1):32 -- 80, 2017. 15, 17, 19 Faculty of Mathematics, Technion-Israel Institute of Technology, Haifa, Israel E-mail address: [email protected] 32
1601.01482
2
1601
2016-01-08T08:33:29
Representations of the canonical commutation relations--algebra and the operators of stochastic calculus
[ "math.OA", "math-ph", "math.FA", "math-ph" ]
We study a family of representations of the canonical commutation relations (CCR)-algebra (an infinite number of degrees of freedom), which we call admissible. The family of admissible representations includes the Fock-vacuum representation. We show that, to every admissible representation, there is an associated Gaussian stochastic calculus, and we point out that the case of the Fock-vacuum CCR-representation in a natural way yields the operators of Malliavin calculus. And we thus get the operators of Malliavin's calculus of variation from a more algebraic approach than is common. And we obtain explicit and natural formulas, and rules, for the operators of stochastic calculus. Our approach makes use of a notion of symmetric (closable) pairs of operators. The Fock-vacuum representation yields a maximal symmetric pair. This duality viewpoint has the further advantage that issues with unbounded operators and dense domains can be resolved much easier than what is possible with alternative tools. With the use of CCR representation theory, we also obtain, as a byproduct, a number of new results in multi-variable operator theory which we feel are of independent interest.
math.OA
math
REPRESENTATIONS OF THE CANONICAL COMMUTATION RELATIONS–ALGEBRA AND THE OPERATORS OF STOCHASTIC CALCULUS PALLE JORGENSEN AND FENG TIAN Abstract. We study a family of representations of the canonical commuta- tion relations (CCR)-algebra (an infinite number of degrees of freedom), which we call admissible. The family of admissible representations includes the Fock- vacuum representation. We show that, to every admissible representation, there is an associated Gaussian stochastic calculus, and we point out that the case of the Fock-vacuum CCR-representation in a natural way yields the op- erators of Malliavin calculus. And we thus get the operators of Malliavin’s calculus of variation from a more algebraic approach than is common. And we obtain explicit and natural formulas, and rules, for the operators of stochastic calculus. Our approach makes use of a notion of symmetric (closable) pairs of operators. The Fock-vacuum representation yields a maximal symmetric pair. This duality viewpoint has the further advantage that issues with unbounded operators and dense domains can be resolved much easier than what is possible with alternative tools. With the use of CCR representation theory, we also obtain, as a byproduct, a number of new results in multi-variable operator theory which we feel are of independent interest. Contents Introduction 1. 2. Unbounded operators and the CCR-algebra 2.1. Unbounded operators between different Hilbert spaces 2.2. The CCR-algebra, and the Fock representations 2.3. An infinite-dimensional Lie algebra 2.4. Gaussian Hilbert space 3. The Malliavin derivatives 3.1. A derivation on the algebra D Infinite-dimensional ∆ and ∇Φ 3.2. 3.3. Realization of the operators 3.4. The unitary group 4. The Fock-state, and representation of CCR, realized as Malliavin calculus 5. Conclusions: the general case References 2 3 3 4 5 6 9 15 18 20 21 22 25 25 2000 Mathematics Subject Classification. Primary 81S20, 81S40, 60H07, 47L60, 46N30, 65R10, 58J65, 81S25. Key words and phrases. Canonical commutation relations, representations, unbounded opera- tors, closable operator, unbounded derivations, spectral theory, duality, Gaussian fields, probabil- ity space, stochastic processes, discrete time, path-space measure, stochastic calculus. 1 1. Introduction 2 Both the study of quantum fields, and of quantum statistical mechanics, entails families of representations of the canonical commutation relations (CCRs). In the case of an infinite number of degrees of freedom, it is known that we have existence of many inequivalent representations of the CCRs. Among the representations, some describe such things as a nonrelativistic infinite free Bose gas of uniform density. But the representations of the CCRs play an equally important role in the kind of infinite-dimensional analysis currently used in a calculus of variation approach to Gaussian fields, It¯o integrals, including the Malliavin calculus. In the literature, the infinite-dimensional stochastic operators of derivatives and stochastic integrals are usually taken as the starting point, and the representations of the CCRs are an afterthought. Here we turn the tables. As a consequence of this, we are able to obtain a number of explicit results in an associated multi-variable spectral theory. Some of the issues involved are subtle because the operators in the representations under consideration are unbounded (by necessity), and, as a result, one must deal with delicate issues of domains of families of operators and their extensions. The representations we study result from the Gelfand-Naimark-Segal construc- tion (GNS) applied to certain states on the CCR-algebra. Our conclusions and main results regarding this family of CCR representations (details below, especially sects 4 and 5) hold in the general setting of Gaussian fields. But for the benefit of read- ers, we have also included an illustration dealing with the simplest case, that of the standard Brownian/Wiener process. Many arguments in the special case carry over to general Gaussian fields mutatis mutandis. In the Brownian case, our initial Hilbert space will be L = L2 (0,∞). From the initial Hilbert space L , we build the ∗-algebra CCR (L ) as in Section 2.2. We will show that the Fock state on CCR (L ) corresponds to the Wiener measure P. Moreover the corresponding representation π of CCR (L ) will be acting on the Hilbert space L2 (Ω, P) in such a way that for every k in L , the operator π(a(k)) is the Malliavin derivative in the direction of k. We caution that the representations of the ∗-algebra CCR (L ) are by unbounded operators, but the operators in the range of the representations will be defined on a single common dense domain. Example: There are two ways to think of systems of generators for the CCR- algebra over a fixed infinite-dimensional Hilbert space (“CCR” is short for canonical commutation relations.): (i) an infinite-dimensional Lie algebra, or (ii) an associative ∗-algebra. With this in mind, (ii) will simply be the universal enveloping algebra of (i); see [Dix77]. While there is also an infinite-dimensional “Lie” group corresponding to (i), so far, we have not found it as useful as the Lie algebra itself. All this, and related ideas, supply us with tools for an infinite-dimensional sto- chastic calculus. It fits in with what is called Malliavin calculus, but our present approach is different, and more natural from our point of view; and as corollaries, we obtain new and explicit results in multi-variable spectral theory which we feel are of independent interest. 3 There is one particular representation of the CCR version of (i) and (ii) which is especially useful for stochastic calculus. In the present paper, we call this rep- resentation the Fock vacuum-state representation. One way of realizing the repre- sentations is abstract: Begin with the Fock vacuum state (or any other state), and then pass to the corresponding GNS representation. The other way is to realize the representation with the use of a choice of a Wiener L2-space. We prove that these two realizations are unitarily equivalent. By stochastic calculus we mean stochastic derivatives (e.g., Malliavin deriva- tives), and integrals (e.g., It¯o-integrals). The paper begins with the task of realizing a certain stochastic derivative operator as a closable operator acting between two Hilbert spaces. 2. Unbounded operators and the CCR-algebra 2.1. Unbounded operators between different Hilbert spaces. While the the- ory of unbounded operators has been focused on spectral theory where it is then natural to consider the setting of linear endomorphisms with dense domain in a fixed Hilbert space; many applications entail operators between distinct Hilbert spaces, say H1 and H2. Typically the facts given about the two differ greatly from one Hilbert space to the next. Let Hi, i = 1, 2, be two complex Hilbert spaces. The respective inner products will be written (cid:104)·,·(cid:105)i, with the subscript to identify the Hilbert space in question. Definition 2.1. A linear operator T from H1 to H2 is a pair D ⊂ H1, T , where D is a linear subspace in H1, and T ϕ ∈ H2 is well-defined for all ϕ ∈ D. We say that D = dom (T ) is the domain of T , and ⊂ ; ϕ ∈ D G (T ) = (cid:26)(cid:18) ϕ (cid:19) T ϕ (cid:27) (cid:33) (cid:32)H1⊕ H2 is the graph. By closure, we shall refer to closure in the norm of H1 ⊕ H2, i.e., If the closure G (T ) is the graph of a linear operator, we say that T is closable. (2.1) (2.2) T−−→ H2, then T is closable if hi ∈ Hi. If dom (T ) is dense in H1, we say that T is densely defined. = (cid:107)h1(cid:107)2 h2 1 + (cid:107)h2(cid:107)2 2 , (cid:13)(cid:13)(cid:13)(cid:13)(cid:18)h1 (cid:19)(cid:13)(cid:13)(cid:13)(cid:13)2 (cid:110) h2 ∈ H2 ; ∃C = Ch2 < ∞ s.t. (cid:104)T ϕ, h2(cid:105)2 ≤ C (cid:107)ϕ(cid:107)1 , ∀ϕ ∈ dom (T ) dom (T ∗) = Definition 2.2. Let H1 subspace dom (T ∗) ⊂ H2 defined as follows: T−−→ H2 be a densely defined operator, and consider the (cid:111) (2.3) (2.4) Then, by Riesz’ theorem, there is a unique h1 ∈ H1 s.t. (cid:104)T ϕ, h2(cid:105)2 = (cid:104)ϕ, h1(cid:105)1 , and we set T ∗h2 = h1. Lemma 2.3. Given a densely defined operator H1 and only if dom (T ∗) is dense in H2. Proof. See [DS88]. Remark 2.4 (Notation and Facts). (1) The abbreviated notation H1 T T ∗ 4 (cid:3) H2 will be used when the domains of G(cid:0)T(cid:1) = G (T ) T and T ∗ are understood from the context. (2) Let T be an operator H1 T−−→ H2 and Hi, i = 1, 2, two given Hilbert spaces. Assume D := dom (T ) is dense in H1, and that T is closable. Then there is a unique closed operator, denoted T such that (2.5) where “—” on the RHS in (2.5) refers to norm closure in H1⊕H2, see (2.2). (3) It may happen that dom (T ∗) = 0. See Example 2.5 below. Example 2.5. An operator T : H1 −→ H2 with dense domain s.t. dom (T ∗) = 0, i.e., “extremely” non-closable. Set Hi = L2 (µi), i = 1, 2, where µ1 and µ2 are two mutually singular measures on a fixed locally compact measurable space, say X. The space D := Cc (X) is dense in both H1 and in H2 with respect to the two L2-norms. Then, the identity mapping T ϕ = ϕ, ∀ϕ ∈ D, becomes a Hilbert space operator H1 Using Definition 2.2, we see that h2 ∈ L2 (µ2) is in dom (T ∗) iff ∃h1 ∈ L2 (µ1) T−−→ H2. such that Since D is dense in both L2-spaces, we get ϕ h1 dµ1 = ϕ h2 dµ2, ∀ϕ ∈ D. (2.6) (2.7) h1 dµ1 = h2 dµ2, E E where E = supp (µ2). ´ ´ A h1 dµ1 = ´ A h2 dµ2 > 0. But ´ Now suppose h2 (cid:54)= 0 in L2 (µ2), then there is a subset A ⊂ E s.t. h2 > 0 on A, µ2 (A) > 0, and A h1 dµ1 = 0 since µ1 (A) = 0. This contradiction proves that dom (T ∗) = 0; and in particular T is unbounded and non-closable. T−−→ H2 be a densely defined operator, and assume that Theorem 2.6. Let H1 dom (T ∗) is dense in H2, i.e., T is closable, then both of the operators T ∗T and T T ∗ are densely defined, and both are selfadjoint. Moreover, there is a partial isometry U : H1 −→ H2 with initial space in H1 A h2 dµ2, and and final space in H2 such that T = U(cid:0)T ∗T(cid:1) 1 2 =(cid:0)T T ∗(cid:1) 1 2 U. (2.8) (Eq. (2.8) is called the polar decomposition of T .) (cid:3) Proof. See, e.g., [DS88]. 2.2. The CCR-algebra, and the Fock representations. There are two ∗- algebras built functorially from a fixed (single) Hilbert space L ; often called the one-particle Hilbert space (in physics). The dimension dim L is called the number of degrees of freedom. The case of interest here is when dim L = ℵ0 (countably infinite). The two ∗-algebras are called the CAR, and the CCR-algebras, and they are extensively studied; see e.g., [BR81]. Of the two, only CAR(L ) is a C∗-algebra. * * j j 5 The operators arising from representations of CCR(L ) will be unbounded, but still having a common dense domain in the respective representation Hilbert spaces. In both cases, we have a Fock representation. For CCR(L ), it is realized in the sym- metric Fock space Γsym (L ). There are many other representations, inequivalent to the respective Fock representations. a∗ (h), h ∈ L , subject to Let L be as above. The CCR(L ) is generated axiomatically by a system, a (h), [a (h) , a (k)] = 0, ∀h, k ∈ L , and [a (h) , a∗ (k)] = (cid:104)h, k(cid:105)L 1. (2.9) Notation. In (2.9), [·,·] denotes the commutator. More specifically, if A, B are elements in a ∗-algebra, set [A, B] := AB − BA. The Fock States ωF ock on the CCR-algebra are specified as follows: ωF ock (a (h) a∗ (k)) = (cid:104)h, k(cid:105)L with the vacuum property ωF ock (a∗ (h) a (h)) = 0, ∀h ∈ L ; For the corresponding Fock representations π we have: [π (h) , π∗ (k)] = (cid:104)h, k(cid:105)L IΓsym(L ), (2.10) (2.11) (2.12) where IΓsym(L ) on the RHS of (2.12) refers to the identity operator. Some relevant papers regarding the CCR-algebra and its representations are [AW63, Arv76a, Arv76b, PS72a, PS72b, AW73, GJ87, JP91]. 2.3. An infinite-dimensional Lie algebra. Let L be a separable Hilbert space, i.e., dim L = ℵ0, and let CCR (L ) be the corresponding CCR-algebra. As above, its generators are denoted a (k) and a∗ (l), for k, l ∈ L . We shall need the following: Proposition 2.7. (1) The “quadratic” elements in CCR (L ) of the form a (k) a∗ (l), k, l ∈ L , span a Lie algebra g (L ) under the commutator bracket. (2) We have [a (h) a∗ (k) , a (l) a∗ (m)] = (cid:104)h, m(cid:105)L a (l) a∗ (k) − (cid:104)k, l(cid:105)L a (h) a∗ (m) , (3) If {εi}i∈N is an ONB in L , then the non-zero commutators are as follows: for all h, k, l, m ∈ L . Set γi,j := a (εi) a∗ (εj), then, for i (cid:54)= j, we have (2.13) (2.14) (2.15) All other commutators vanish; in particular, {γi,i i ∈ N} spans an abelian sub-Lie algebra in g (L ). [γi,i, γj,i] = γj,i; [γi,i, γi,j] = −γi,j; and [γj,i, γi,j] = γi,i − γj,j. Note further that, when i (cid:54)= j, then the three elements γi,i − γj,j, γi,j, and γj,i (2.16) span (over R) an isomorphic copy of the Lie algebra sl2 (R). (4) The Lie algebra generated by the first-order elements a (h) and a∗ (k) for h, k ∈ L , is called the Heisenberg Lie algebra h (L ). It is normalized by g (L ); indeed we have: [a (l) a∗ (m) , a (h)] = −(cid:104)m, h(cid:105)L a (l) , and [a (l) a∗ (m) , a∗ (k)] = (cid:104)l, k(cid:105)L a∗ (m) , ∀l, m, h, k ∈ L . 6 Proof. The verification of each of the four assertions (1)-(4) uses only the fixed axioms for the CCR, i.e., [a (k) , a (l)] = 0, [a∗ (k) , a∗ (l)] = 0, and [a (k) , a∗ (l)] = (cid:104)k, l(cid:105)L 1, k, l ∈ L ; (2.17) (cid:3) where 1 denotes the unit-element in CCR (L ). Corollary 2.8. Let CCR (L ) be the CCR-algebra, generators a (k), a∗ (l), k, l ∈ L , and let [·,·] denote the commutator Lie bracket; then, for all k, h1,··· , hn ∈ L , and all p ∈ R [x1,··· , xn] (= the n-variable polynomials over R), we have [a (k) , p (a∗ (h1) ,··· , a∗ (hn))] = (a∗ (h1) ,··· , a∗ (hn))(cid:104)k, hi(cid:105)L . ∂p ∂xi (2.18) n(cid:88) i=1 Proof. The verification of (2.18) uses only the axioms for the CCR, i.e., the com- (cid:3) mutation relations (2.17) above, plus a little combinatorics. We shall now return to a stochastic variation of formula (2.18), the so called Malliavin derivative in the direction k. In this, the system (a∗ (h1) ,··· , a∗ (hn)) in (2.18) instead takes the form of a multivariate Gaussian random variable. 2.4. Gaussian Hilbert space. The literature on Gaussian Hilbert space, white noise analysis, and its relevance to Malliavin calculus is vast; and we limit ourselves here to citing [BØSW04, AJL11, AJ12, VFHN13, AJS14, AJ15, AØ15], and the papers cited there. Setting and Notation. L : a fixed real Hilbert space (Ω,F, P): a fixed probability space L2 (Ω, P): the Hilbert space L2 (Ω,F, P), also denoted by L2 (P) E: the mean or expectation functional, where E (··· ) = ´ Ω (··· ) dP Definition 2.9. Fix a real Hilbert space L and a given probability space (Ω,F, P). We say the pair (L , (Ω,F, P)) is a Gaussian Hilbert space. A Gaussian field is a linear mapping Φ : L −→ L2 (Ω, P), such that {Φ (h) h ∈ L } is a Gaussian process indexed by L satisfying: (1) E (Φ (h)) = 0, ∀h ∈ L ; (2) ∀n ∈ N, ∀l1,··· , ln ⊂ L , the random variable (Φ (l1) ,··· , Φ (ln)) is jointly Gaussian, with i.e., ((cid:104)li, lj(cid:105))n fields, see the discussion below.) E (Φ (li) Φ (lj)) = (cid:104)li, lj(cid:105) , (2.19) i=1 = the covariance matrix. (For the existence of Gaussian 7 Figure 2.1. The multivariate Gaussian (Φ (h1) ,··· Φ (hn)) and its distribution. The Gaussian with Gramian matrix (Gram ma- trix) Gn, n = 2. Remark 2.10. For all finite systems {li} ⊂ L , set Gn = ((cid:104)li, lj(cid:105))n i,j=1, called the Gramian. Assume Gn non-singular for convenience, so that det Gn (cid:54)= 0. Then there is an associated Gaussian density g(Gn) on Rn, −n/2 (det Gn) −1/2 exp g(Gn) (x) = (2π) (2.20) The condition in (2.19) assumes that for all continuous functions f : Rn −→ R (e.g., polynomials), we have Rn − 1 2 (cid:18) n x(cid:11) (cid:10)x, G−1 (cid:19) (cid:124) (cid:123)(cid:122) real valued (cid:125)) E(f (Φ (l1) ,··· , Φ (ln)) = Rn f (x) g(Gn) (x) dx; (2.21) where x = (x1,··· , xn) ∈ Rn, and dx = dx1 ··· dxn = Lebesgue measure on Rn. See Figure 2.1 for an illustration. In particular, for n = 2, (cid:104)l1, l2(cid:105) = (cid:104)k, l(cid:105), and f (x1, x2) = x1x2, we then get E (Φ (k) Φ (l)) = (cid:104)k, l(cid:105), i.e., the inner product in L . For our applications, we need the following facts about Gaussian fields. Fix a Hilbert space L over R with inner product (cid:104)·,·(cid:105)L . Then (see [Hid80, AØ15, Gro70]) there is a probability space (Ω,F, P), depending on L , and a real linear mapping Φ : L −→ L2 (Ω,F, P), i.e., a Gaussian field as specified in Definition 2.9, satisfying (2.22) It follows from the literature (see also [JT14]) that Φ (k) may be thought of as a generalized It¯o-integral. One approach to this is to select a nuclear Fréchet space S with dual S(cid:48) such that , 2(cid:107)k(cid:107)2 ∀k ∈ L . E(cid:0)eiΦ(k)(cid:1) = e− 1 (2.23) forms a Gelfand triple. In this case we may take Ω = S(cid:48), and Φ (k), k ∈ L , to be the extension of the mapping S (cid:44)→ L (cid:44)→ S(cid:48) (2.24) defined initially only for ϕ ∈ S, but, with the use of (2.24), now extended, via (2.22), from S to L . See also Example 2.12 below. S(cid:48) (cid:51) ω −→ ω (ϕ) = (cid:104)ϕ, ω(cid:105) 8 Example 2.11. Fix a measure space (X,B, µ). Let Φ : L2 (µ) −→ L2 (Ω, P) be a Gaussian field such that E (ΦAΦB) = µ (A ∩ B) , ´ ∀A, B ∈ B where ΦE := Φ (χE), ∀E ∈ B; and χE denotes the characteristic function. In this case, L = L2 (X, µ). Then we have Φ (k) = X k (x) dΦ, i.e., the It¯o-integral, and the following holds: (2.25) E (Φ (k) Φ (l)) = (cid:104)k, l(cid:105) = k (x) l (x) dµ (x) for all k, l ∈ L = L2 (X, µ). Eq. (2.25) is known as the It¯o-isometry. Example 2.12 (The special case of Brownian motion). There are many ways of realizing a Gaussian probability space (Ω,F, P). Two candidates for the sample space: Case 1. Standard Brownian motion process: Ω = C (R), F = σ-algebra gener- ated by cylinder sets, P = Wiener measure. Set Bt (ω) = ω (t), ∀ω ∈ Ω; and Φ (k) = R k (t) dBt, ∀k ∈ L2 (Ω, P). ´ X Case 2. The Gelfand triples: S (cid:44)→ L2 (R) (cid:44)→ S(cid:48), where S = the Schwartz space of test functions; S(cid:48) = the space of tempered distributions. Set Ω = S(cid:48), F = σ-algebra generated by cylinder sets of S(cid:48), and define Φ (k) :=(cid:98)k (ω) = (cid:104)k, ω(cid:105) , k ∈ L2 (R) , ω ∈ S(cid:48). E(cid:0)ei(cid:104)k,·(cid:105)(cid:1) = Note Φ is defined by extending the duality S ←→ S(cid:48) to L2 (R). The probability measure P is defined from ei(cid:98)k(ω)dP (ω) = e 2(cid:107)k(cid:107)2 L2(R), − 1 S(cid:48) by Minlos’ theorem [Hid80, AØ15]. Definition 2.13. Let D ⊂ L2 (Ω,F, P) be the dense subspace spanned by functions F , where F ∈ D iff ∃n ∈ N, ∃h1,··· , hn ∈ L , and p ∈ R [x1,··· , xn] = the polynomial ring, such that F = p (Φ (h1) ,··· , Φ (hn)) : Ω −→ R. (See the diagram below.) The case of n = 0 corresponds to the constant function 1 on Ω. Note that Φ (hi) ∈ L2 (Ω, P). (Φ(h1),··· ,Φ(hn)) Ω Rn F p / R Lemma 2.14. The polynomial fields D in Def. 2.13 form a dense subspace in L2 (Ω, P). Proof. The easiest argument below takes advantage of the isometric isomorphism of L2 (Ω, P) with the symmetric Fock space Γsym (L ) = H0(cid:124)(cid:123)(cid:122)(cid:125) 1 dim ⊕ ∞(cid:88) n=1 (cid:124) (cid:123)(cid:122) (L ⊗ ··· ⊗ L ) . n-fold symmetric (cid:125) ' ' 7 7 / For ki ∈ L , i = 1, 2, there is a unique vector eki ∈ Γsym (L ) such that (cid:10)ek1, ek2(cid:11) Γsym(L ) = (cid:104)k1, k2(cid:105)n n! = e(cid:104)k1,k2(cid:105)L . ∞(cid:88) n=0 Moreover, Γsym (L ) (cid:51) ek W0−−−→ eΦ(k)− 1 2(cid:107)k(cid:107)2 L ∈ L2 (Ω, P) mapping onto L2 (Ω, P). Hence D is dense in L2 (Ω, P), as span(cid:8)ek k ∈ L(cid:9) is extends by linearity and closure to a unitary isomorphism Γsym (L ) W−−→ L2 (Ω, P), (cid:3) dense in Γsym (L ). Lemma 2.15. Let L be a real Hilbert space, and let (Ω,F, P, Φ) be an associated Gaussian field. For n ∈ N, let {h1,··· , hn} be a system of linearly independent vec- tors in L . Then, for polynomials p ∈ R [x1,··· , xn], the following two conditions are equivalent: p (Φ (h1) ,··· , Φ (hn)) = 0 a.e. on Ω w.r.t P; and ∀ (x1,··· , xn) ∈ Rn. (2.26) (2.27) Proof. Let Gn = ((cid:104)hi, hj(cid:105))n i,j=1 be the Gramian matrix. We have det Gn (cid:54)= 0. Let g(Gn) (x1,··· , xn) be the corresponding Gaussian density; see (2.20), and Figure 2.1. Then the following are equivalent: p (x1,··· , xn) ≡ 0, 9 (cid:3) (1) Eq. (2.26) holds; (2) p (Φ (h1) ,··· , Φ (hn)) = 0 in L2 (Ω,F, P); (3) E(cid:16)p (Φ (h1) ,··· , Φ (hn))2(cid:17) ´ = Rn p (x)2 g(Gn) (x) dx = 0; (4) p (x) = 0 a.e. x w.r.t. the Lebesgue measure in Rn ; (5) p (x) = 0, ∀x ∈ Rn; i.e., (2.27) holds. 3. The Malliavin derivatives Below we give an application of the closability criterion for linear operators T between different Hilbert spaces H1 and H2, but having dense domain in the first Hilbert space. In this application, we shall take for T to be the so called Malliavin derivative. The setting for it is that of the Wiener process. For the Hilbert space H1 we shall take the L2-space, L2 (Ω, P) where P is generalized Wiener measure. Below we shall outline the basics of the Malliavin derivative, and we shall specify the two Hilbert spaces corresponding to the setting of Theorem 2.6. We also stress that the literature on Malliavin calculus and its applications is vast, see e.g., [BØSW04, AØ15]. Settings. It will be convenient for us to work with the real Hilbert spaces. Let (Ω,F, P, Φ) be as specified in Definition 2.9, i.e., we consider the Gaussian field Φ. Fix a real Hilbert space L with dim L = ℵ0. Set H1 = L2 (Ω, P), and H2 = L2 (Ω → L , P) = L2 (Ω, P) ⊗ L , i.e., vector valued random variables. For H1, the inner product (cid:104)·,·(cid:105)H1 is where E (··· ) = (cid:104)F, G(cid:105)H1 = ´ Ω (··· ) dP is the mean or expectation functional. Ω F G dP = E (F G) ; (3.1) On H2, we have the tensor product inner product: If Fi ∈ H1, ki ∈ L , i = 1, 2, then 10 Equivalently, if ψi : Ω −→ L , i = 1, 2, are measurable functions on Ω, we set (cid:104)F1 ⊗ k1, F2 ⊗ k2(cid:105)H2 = (cid:104)F1, F2(cid:105)H1 (cid:104)k1, k2(cid:105)L = E (F1F2)(cid:104)k1, k2(cid:105)L . (cid:104)ψ1, ψ2(cid:105)H2 = (cid:104)ψ1 (ω) , ψ2 (ω)(cid:105)L dP (ω) ; Ω (3.2) (3.3) (3.7) (3.8) where it is assumed that Ω L dP (ω) < ∞, (cid:107)ψi (ω)(cid:107)2 ´ ∞ Remark 3.1. In the special case of standard Brownian motion, we have L = L2 (0,∞), and set Φ (h) = 0 h (t) dΦt (= the It¯o-integral), for all h ∈ L . Recall E(cid:16)Φ (h)2(cid:17) we then have h (t)2 dt, ∞ i = 1, 2. (3.5) (3.4) = or equivalently (the It¯o-isometry), 0 (cid:107)Φ (h)(cid:107)L2(Ω,P) = (cid:107)h(cid:107)L , ∀h ∈ L . (3.6) The consideration above also works in the context of general Gaussian fields; see Section 2.4. Definition 3.2. Let D be the dense subspace in H1 = L2 (Ω, P) as in Definition 2.13. The operator T : H1 −→ H2 (= Malliavin derivative) with dom (T ) = D is specified as follows: For F ∈ D, i.e., ∃n ∈ N, p (x1,··· , xn) a polynomial in n real variables, and h1, h2,··· , hn ∈ L , where F = p (Φ (h1) ,··· , Φ (hn)) ∈ L2 (Ω, P) . Set T (F ) = (cid:19) (cid:18) ∂ n(cid:88) ∂xj j=1 (Φ (h1) ,··· , Φ (hn)) ⊗ hj ∈ H2. p In the following two remarks we outline the argument for why the expression for T (F ) in (3.8) is independent of the chosen representation (3.7) for the particular F . Recall that F is in the domain D of T . Without some careful justification, it is not even clear that T , as given, defines a linear operator on its dense domain D. The key steps in the argument to follow will be the result (3.12) in Theorem 3.8 below, and the discussion to follow. There is an alternative argument, based instead on Corollary 2.8; see also Section 5 below. Remark 3.3. It is non-trivial that the formula in (3.8) defines a linear operator. Reason: On the LHS in (3.8), the representation of F from (3.7) is not unique. So we must show that p (Φ (h1) ,··· , Φ (hn)) = 0 =⇒ RHS(3.8) = 0 as well. (The dual pair analysis below (see Def. 3.6) is good for this purpose.) m(cid:88) i=1 n(cid:88) i=1 11 Suppose F ∈ D has two representations corresponding to systems of vectors h1,··· , hn ∈ L , and k1,··· , km ∈ L , with polynomials p ∈ R [x1,··· , xn], and q ∈ R [x1,··· , xm], where F = p (Φ (h1) ,··· , Φ (hn)) = q (Φ (k1) ,··· , Φ (km)) . (3.9) We must then verify the identity: (Φ (h1) ,··· , Φ (hn)) ⊗ hi = ∂p ∂xi (Φ (k1) ,··· , Φ (km)) ⊗ ki. (3.10) ∂q ∂xi The significance of the next result is the implication (3.9) =⇒ (3.10), valid for all choices of representations of the same F ∈ D. The conclusion from (3.12) in Theorem 3.8 is that the following holds for all l ∈ L : E(cid:0)(cid:10)LHS(3.10), l(cid:11)(cid:1) = E(cid:0)(cid:10)RHS(3.10), l(cid:11)(cid:1) = E (F Φ (l)) . Moreover, with a refinement of the argument, we arrive at the identity (cid:10)LHS(3.10) − RHS(3.10), G ⊗ l(cid:11) valid for all G ∈ D, and all l ∈ L . But span{G ⊗ l G ∈ D, l ∈ L } is dense in H2 H2 = 0, (cid:0)= L2 (P) ⊗ L(cid:1) w.r.t. the tensor- Hilbert norm in H2 (see (3.2)); and we get the desired identity (3.10) for any two representations of F . Remark 3.4. An easy case where (3.9) =⇒ (3.10) can be verified “by hand”: Let F = Φ (h)2 with h ∈ L \{0} fixed. We can then pick the two systems {h} and {h, h} with p (x) = x2, and q (x1, x2) = x1x2. A direct calculus argument shows that LHS(3.10) = RHS(3.10) = 2Φ (h) ⊗ h ∈ H2. We now resume the argument for the general case. Definition 3.5 (symmetric pair). For i = 1, 2, let Hi be two Hilbert spaces, and suppose Di ⊂ Hi are given dense subspaces. We say that a pair of operators (S, T ) forms a symmetric pair if dom (T ) = D1, and dom (S) = D2; and moreover, (cid:104)T u, v(cid:105)H2 = (cid:104)u, Sv(cid:105)H1 (3.11) holds for ∀u ∈ D1, ∀v ∈ D2. graphs: It is immediate that (3.11) may be rewritten in the form of containment of T ⊂ S∗, S ⊂ T ∗. In that case, both S and T are closable. We say that a symmetric pair is maximal if T = S∗ and S = T ∗. H1 T S H2 We will establish the following two assertions: (1) Indeed T from Definition 3.2 is a well-defined linear operator from H1 to H2 . (2) Moreover, (S, T ) is a maximal symmetric pair (see Definitions 3.5, 3.6). & & f f 12 T−−→ H2 be the Malliavin derivative with D1 = dom (T ), Definition 3.6. Let H1 see Definition 3.2. Set D2 = D1 ⊗ L = algebraic tensor product, and on dom (S) = D2, set S (F ⊗ k) = −(cid:104)T (F ) , k(cid:105) + MΦ(k)F, ∀F ⊗ k ∈ D2, where MΦ(k) = the operator of multiplication by Φ (k). Note that both operators S and T are linear and well defined on their respective dense domains, Di ⊂ Hi, i = 1, 2. For density, see Lemma 2.14. It is a “modern version” of ideas in the literature on analysis of Gaussian pro- cesses; but we are adding to it, giving it a twist in the direction of multi-variable operator theory, representation theory, and especially to representations of infinite- dimensional algebras on generators and relations. Moreover our results apply to more general Gaussian processes than covered so far. Lemma 3.7. Let (S, T ) be the pair of operators specified above in Definition 3.6. Then it is a symmetric pair, i.e., (cid:104)T u, v(cid:105)H2 = (cid:104)u, Sv(cid:105)H1 , ∀u ∈ D1, ∀v ∈ D2. Equivalently, (cid:104)T (F ) , G ⊗ k(cid:105)H2 = (cid:104)F, S (G ⊗ k)(cid:105)H1 ∀F, G ∈ D, ∀k ∈ L . , In particular, we have S ⊂ T ∗, and T ⊂ S∗(containment of graphs.) Moreover, the two operators S∗S and T ∗T are selfadjoint. (For the last conclusion in the lemma, see Theorem 2.6.) Theorem 3.8. Let T : H1 −→ H2 be the Malliavin derivative, i.e., T is an unbounded closable operator with dense domain D consisting of the span of all the functions F from (3.7). Then, for all F ∈ dom (T ), and k ∈ L , we have E ((cid:104)T (F ) , k(cid:105)L ) = E (F Φ (k)) . (3.12) Proof. We shall prove (3.12) in several steps. Once (3.12) is established, then there is a recursive argument which yields a dense subspace in H2, contained in dom (T ∗); and so T is closable. Moreover, formula (3.12) yields directly the evaluation of T ∗ : H2 −→ H1 as follows: If k ∈ L , set 1 ⊗ k ∈ H2 where 1 denotes the constant function “one” on Ω. We get ∞ T ∗ (1 ⊗ k) = Φ (k) = k (t) dΦt (= the It¯o-integral.) (3.13) 0 The same argument works for any Gaussian field; see Definition 2.9. We refer to the literature [BØSW04, AØ15] for details. The proof of (3.12) works for any Gaussian process L (cid:51) k −→ Φ (k) indexed by an arbitrary Hilbert space L with the inner product (cid:104)k, l(cid:105)L as the covariance kernel. Formula (3.12) will be established as follows: Let F and T (F ) be as in (3.7)- Step 1. For every n ∈ N, the polynomial ring R [x1, x2,··· , xn] is invariant (3.8). under matrix substitution y = M x, where M is an n × n matrix over R. Step 2. Hence, in considering (3.12) for {hi}n onalize the n × n Gram matrix ((cid:104)hi, hj(cid:105))n may assume that the system {hi}n i=1 ⊂ L , h1 = k, we may diag- i,j=1; thus without loss of generality, we i=1 is orthogonal and normalized, i.e., that (cid:104)hi, hj(cid:105) = δij, ∀i, j ∈ {1,··· , n} , 13 (3.14) and we may take k = h1 in L . that the joint distribution of {Φ (hi)}n Rn, i.e., Step 3. With this simplification, we now compute the LHS in (3.12). We note i=1 is thus the standard Gaussian kernel in −n/2 e− 1 (cid:80)n i=1 x2 i , (3.15) 2 gn (x) = (2π) with x = (x1,··· , xn) ∈ Rn. We have x1gn (x) = − ∂ ∂x1 gn (x) (3.16) by calculus. Step 4. A direct computation yields LHS(3.12) = = by (3.14) = by (3.15) = int. by parts = by (3.16) = by (3.14) = (cid:19) (cid:18) ∂p E ((cid:104)T (F ) , h1(cid:105)L ) E ∂x1 ∂p ∂x1 p (x1,··· , xn) (Φ (h1) ,··· Φ (hn)) (x1,··· , xn) gn (x1,··· , xn) dx1 ··· dxn (x1,··· , xn) dx1 ··· dxn Rn x1p (x1,··· , xn) gn (x1,··· , xn) dx1 ··· dxn ∂gn ∂x1 Rn − Rn E (Φ (h1) p (Φ (h1) ,··· , Φ (hn))) E (Φ (h1) F ) = RHS(3.12), which is the desired conclusion (3.12). (cid:3) T−−→ H2 be as in Theorem 3.8, i.e., T is the Corollary 3.9. Let H1, H2, and H1 Malliavin derivative. Then, for all h, k ∈ L = L2 (0,∞), we have for the closure T of T the following: E(cid:0)(cid:104)T (eΦ(h)), k(cid:105)L(cid:1) = e T (eΦ(h)) = eΦ(h) ⊗ h, 2(cid:107)h(cid:107)2 1 and L (cid:104)h, k(cid:105)L . Here T denotes the graph-closure of T . Moreover, T ∗T (eΦ(k)) = Φ (k) − (cid:107)k(cid:107)2 L (cid:17) eΦ(k). (3.17) (3.18) (3.19) (cid:16) n(cid:88) Proof. Eqs. (3.17)-(3.18) follow immediately from (3.12) and a polynomial approx- imation to see (3.7). In particular, eΦ(h) ∈ dom(cid:0)T(cid:1), and T(cid:0)eΦ(h)(cid:1) is well defined. ex = lim n→∞ 0 , x ∈ R; xj j! For (3.19), we use the facts for the Gaussians: E(eΦ(k)) = e 1 2(cid:107)k(cid:107)2 , and E(Φ (k) eΦ(k)) = (cid:107)k(cid:107)2 e 1 2(cid:107)k(cid:107)2 . 14 (cid:3) Example 3.10. Let F = Φ (k)k, (cid:107)k(cid:107) = 1. We have T Φ (k)n = nΦ (k)n−1 ⊗ k T ∗T Φ (k)n = −n (n − 1) Φ (k)n−2 + nΦ (k)n and similarly, T eΦ(k) = eΦ(k) ⊗ k T ∗T eΦ(k) = eΦ(k) (Φ (k) − 1) . Let (S, T ) be the symmetric pair, we then have the inclusion T ⊂ S∗, i.e., containment of the operator graphs, G(cid:0)T(cid:1) ⊂ G (S∗). In fact, we have Proof. We will show that G (S∗) (cid:9) G(cid:0)T(cid:1) = 0, where (cid:9) stands for the orthogo- nal complement in the direct sum-inner product of H1 ⊕ H2. Recall that H1 = L2 (Ω, P), and H2 = H1 ⊗ L . Corollary 3.11. T = S∗. Using (3.17), we will prove that if F ∈ dom (S∗), and (cid:19)(cid:29) (cid:19) (cid:18) F , S∗F (cid:28)(cid:18) eΦ(k) E(cid:16) eΦ(k) ⊗ k (cid:17) = 0, ∀k ∈ L =⇒ F = 0, which is equivalent to But it is know that for the Gaussian filed, span(cid:8)eΦ(k) k ∈ L(cid:9) is dense in H1, eΦ(k) (F + (cid:104)S∗F, k(cid:105)) = 0, ∀k ∈ L . (3.20) and so (3.20) implies that F = 0, which is the desired conclusion. We can finish the proof of the corollary with an application of Girsanov’s theo- rem, see e.g., [BØSW04] and [Pri10]. By this result, we have a measurable action τ of L on (Ω,F, P), i.e., L τ−→ Aut (Ω,F) a.e. on Ω, ∀k, l ∈ L (see also sect 5 below) s.t. τk (F) = F for all k ∈ L , and τk ◦ τl = τk+l P ◦ τ−1 k (cid:28) P (3.21) with dP ◦ τ−1 k dP = e− 1 2(cid:107)k(cid:107)2 L eΦ(k), a.e. on Ω. (3.22) Returning to (3.20). An application of (3.22) to (3.20) yields: F (· + k) + (cid:104)S∗ (F ) (· + k) , k(cid:105)L = 0 a.e. on Ω; (3.23) where we have used “· + k” for the action in (3.21). Since τ in (3.21) is an action by measure-automorphisms, (3.23) implies F (·) + (cid:104)S∗ (F ) (·) , k(cid:105)L = 0; (3.24) 15 If F (cid:54)= 0 in L2 (Ω,F, P), then the second term in again with k ∈ L arbitrary. (3.24) would be independent of k which is impossible with S∗ (F ) (·) (cid:54)= 0. But if S∗ (F ) = 0, then F (·) = 0 (in L2 (Ω,F, P)) by (3.24); and so the proof is (cid:3) completed. Remark 3.12. We recall the definition of the domain of the closure T . The following is a necessary and sufficient condition for an F ∈ L2 (Ω,F, P) to be in the domain of T : F ∈ dom(cid:0)T(cid:1) ⇐⇒ ∃ a sequence {Fn} ⊂ D s.t. E(cid:16)Fn − Fm2 + (cid:107)T (Fn) − T (Fm)(cid:107)2 L (cid:17) = 0. (3.25) lim n,m→∞ When (3.25) holds, we have: T (F ) = lim (3.26) where the limit on the RHS in (3.26) is in the Hilbert norm of L2 (Ω,F, P) ⊗ L . Corollary 3.13. Let (L , Ω,F, P, Φ) be as above, and let T and S be the two operators from Corollary 3.11. Then, for the domain of T , we have the following: For random variables F in L2 (Ω,F, P), the following two conditions are equiv- n→∞ T (Fn) alent: (1) F ∈ dom(cid:0)T(cid:1); (2) ∃C = CF < ∞ s.t. E (F S (ψ))2 ≤ C E(cid:16)(cid:107)ψ (·)(cid:107)2 L (cid:17) holds for ∀ψ ∈ span{G ⊗ k G ∈ D, k ∈ L }. Recall equivalently, S (· ⊗ k) = MΦ(k) · −(cid:104)T (·) , k(cid:105)L ; S (G ⊗ k) = Φ (k) G − (cid:104)T (G) , k(cid:105)L for all G ∈ D, and all k ∈ L . Proof. Immediate from the previous corollary. (cid:3) 3.1. A derivation on the algebra D. The study of unbounded derivations has many applications in mathematical physics; in particular in making precise the time dependence of quantum observables, i.e., the dynamics in the Schrödinger picture; — in more detail, in the problem of constructing dynamics in statistical mechanics. An early application of unbounded derivations (in the commutative case) can be found in the work of Silov [Šil47]; and the later study of unbounded derivations in non-commutative C∗-algebras is outlined in [BR81]. There is a rich in variety unbounded derivations, because of the role they play in applications to dynamical systems in quantum physics. But previously the theory of unbounded derivations has not yet been applied systematically to stochastic analysis in the sense of Malliavin. In the present section, we turn to this. We begin with the following: Lemma 3.14 (Leibniz-Malliavin). Let H1 (3.7)-(3.8). Then, T−−→ H2 be the Malliavin derivative from 16 (1) dom (T ) =: D, given by (3.7), is an algebra of functions on Ω under point- (2) H2 is a module over D where H2 = L2 (Ω, P) ⊗ L (= vector valued L2- wise product, i.e., F G ∈ D, ∀F, G ∈ D. random variables.) (3) Moreover, i.e., T is a module-derivation. T (F G) = T (F ) G + F T (G) , ∀F, G ∈ D, (3.27) Notation. The eq. (3.27) is called the Leibniz-rule. By the Leibniz, we refer to the traditional rule of Leibniz for the derivative of a product. And the Malliavin derivative is thus an infinite-dimensional extension of Leibniz calculus. Proof. To show that D ⊂ H1 = L2 (Ω, P) is an algebra under pointwise multiplica- tion, the following trick is useful. It follows from finite-dimensional Hilbert space geometry. i=1 ⊂ L , such Let F, G be as in Definition 2.13. Then ∃p, q ∈ R [x1,··· , xn], {li}n F = p (Φ (l1) ,··· , Φ (ln)) , and G = q (Φ (l1) ,··· , Φ (ln)) . That is, the same system l1,··· , ln may be chosen for the two functions F and G. that For the pointwise product, we have i.e., the product in R [x1,··· , xn] with substitution of the random variable F G = (pq) (Φ (l1) ,··· , Φ (ln)) , Eq. (3.27) ⇐⇒ ∂(pq) ∂xi polynomials. Note that n(cid:88) i=1 T (F G) = (Φ (l1) ,··· , Φ (ln)) : Ω −→ Rn. = ∂p ∂xi q + p ∂q ∂xi , which is the usual Leibniz rule applied to (pq) (Φ (l1) ,··· , Φ (ln)) ⊗ li. ∂ ∂xi (cid:3) Remark 3.15. There is an extensive literature on the theory of densely defined unbounded derivations in C∗-algebras. This includes both the cases of abelian and non-abelian ∗-algebras. And moreover, this study includes both derivations in these algebras, as well as the parallel study of module derivations. So the case of the Malliavin derivative is in fact a special case of this study. Readers interested in details are referred to [Sak98], [BJKR84], [BR79], and [BR81]. Definition 3.16. Let (L , Ω,F, P, Φ) be a Gaussian field, and T be the Malliavin derivative with dom (T ) = D. For all k ∈ L , set Tk (F ) := (cid:104)T (F ) , k(cid:105) , F ∈ D. (3.28) In particular, let F = p (Φ (l1) ,··· , Φ (l1)) be as in (3.7), then (Φ (l1) ,··· , Φ (l1))(cid:104)li, k(cid:105) . Tk (F ) = n(cid:88) ∂p ∂xi i=1 Corollary 3.17. Tk is a derivative on D, i.e., Tk (F G) = (TkF ) G + F (TkG) , ∀F, G ∈ D, ∀k ∈ L . (3.29) (cid:3) Proof. Follows from (3.27). Corollary 3.18. Let (L , Ω,F, P, Φ) be a Gaussian field. Fix k ∈ L , and let Tk be the Malliavin derivative in the k direction. Then on D we have Tk + T ∗ [Tk, T ∗ k = MΦ(k), and l ] = (cid:104)k, l(cid:105)L IL2(Ω,P). Proof. For all F, G ∈ D, we have E (Tk (F ) G) + E (F Tk (G)) = by (3.29) = by (3.12) E (Tk (F G)) E (Φ (k) F G) which yields the assertion in (3.30). Eq. (3.31) now follows from (3.30) and the (cid:3) fact that [Tk, Tl] = 0. Definition 3.19. Let (L , Ω,F, P, Φ) be a Gaussian field. For all k ∈ L , let Tk be Malliavin derivative in the k-direction (eq. (3.28)). Assume L is separable, i.e., dim L = ℵ0. For every ONB {ei}∞ (cid:88) i=1 in L , let T ∗ N := Tei . ei i 17 (3.30) (3.31) (3.32) (3.33) (3.34) (N is the CCR number operator. See Section 4 below.) 1 = 0, ∀i. Similarly, Example 3.20. N 1 = 0, since Tei N Φ (k) = Φ (k) N Φ (k)2 = −2(cid:107)k(cid:107)2 1 + 2Φ (k)2 , ∀k ∈ L . To see this, note that(cid:88) T ∗ ei TeiΦ (k) = = i T ∗ (cid:104)ei, k(cid:105) 1 Φ (ei)(cid:104)ei, k(cid:105) (cid:33) = Φ (k) , = Φ (cid:104)ei, k(cid:105) ei ei (cid:88) (cid:88) (cid:32)(cid:88) i i i (cid:88) which is (3.33). The verification of (3.34) is similar. Theorem 3.21. Let {ei} be an ONB in L , then Proof. Note the span of (cid:8)eΦ(k) k ∈ L(cid:9) is dense in L2 (Ω, P), and both sides of Tei = N. (3.35) ei i T ∗T = T ∗ (3.35) agree on eΦ(k), k ∈ L . Indeed, by (3.32), T ∗T eΦ(k) = N eΦ(k) = (cid:16) Φ (k) − (cid:107)k(cid:107)2(cid:17) eΦ(k). (cid:3) Corollary 3.22. Let D := T ∗T . Specialize to the case of n = 1, and consider F = f (Φ (k)), k ∈ L , f ∈ C∞ (R); then D (F ) = −(cid:107)k(cid:107)2 L f(cid:48)(cid:48) (Φ (k)) + Φ (k) f(cid:48) (Φ (k)) . (3.36) (cid:3) Proof. A direct application of the formulas of T and T ∗. Remark 3.23. If (cid:107)k(cid:107)L = 1 in (3.36), then the RHS in (3.36) is obtained by a substitution of the real valued random variable Φ (k) into the deterministic function δ (f ) := − f + x f. (3.37) (cid:18) d (cid:19)2 dx (cid:18) d (cid:19) dx 18 Then eq. (3.36) may be rewritten as D (f (Φ (k))) = δ (f ) ◦ Φ (k) , Corollary 3.24. If {Hn}n∈N0 on R, then we get for ∀k ∈ L , (cid:107)k(cid:107)L = 1, the following eigenvalues (3.38) , N0 = {0, 1, 2,···}, denotes the Hermite polynomials f ∈ C∞ (R) . D (Hn (Φ (k))) = n Hn (Φ (k)) . (3.39) Proof. It is well-known that the Hermite polynomials Hn satisfies (3.40) (cid:3) and so (3.39) follows from a substitution of (3.40) into (3.38). Theorem 3.25. The spectrum of T ∗T , as an operator in L2 (Ω,F, P), is as follows: δ (Hn) = n Hn, ∀n ∈ N0, (cid:0)T ∗T(cid:1) = N0 = {0, 1, 2,···} . specL2(P) Proof. We saw that the L2 (P)-representation is unitarily equivalent to the Fock (cid:3) vacuum representation, and π (Fock-number operator) = T ∗T . 3.2. Infinite-dimensional ∆ and ∇Φ. Corollary 3.26. Let (L , Ω,F, P, Φ) be a Gaussian field, and let T be the Malliavin derivative, L2 (Ω, P) T−−→ L2 (Ω, P)⊗ L . Then, for all F = p (Φ (h1) ,··· , Φ (hn)) ∈ D (see Definition 3.2), we have T ∗T (F ) = − n(cid:88) (cid:124) i=1 ∂2p ∂xi (cid:123)(cid:122) ∆F n(cid:88) (cid:124) i=1 (cid:125) (Φ (h1) ,··· , Φ (hn)) + Φ (hi) ∂p ∂xi (Φ (h1) ,··· , Φ (hn)) , (cid:123)(cid:122) ∇ΦF (cid:125) which is abbreviated T ∗T = −∆ + ∇Φ. (3.41) (For the general theory of infinite-dimensional Laplacians, see e.g., [Hid03].) Proof. (Sketch) We may assume the system {hi}n (cid:104)hi, hj(cid:105) = δij. Hence, for F F = p (Φ (h1) ,··· , Φ (hn)) ∈ D, we have i=1 ⊂ L is orthonormal, i.e., n(cid:88) ∂p ∂xi T F = i=1 T ∗T (F ) = − n(cid:88) n(cid:88) i=1 + i=1 (Φ (h1) ,··· , Φ (hn)) ⊗ hi, and (Φ (h1) ,··· , Φ (hn)) ∂2p ∂x2 i Φ (hi) ∂p ∂xi (Φ (h1) ,··· , Φ (hn)) which is the assertion. For details, see the proof of Theorem 3.8. (cid:3) 19 Definition 3.27. Let (L , Ω,F, P, Φ) be a Gaussian field. On the dense domain D ⊂ L2 (Ω, P), we define the Φ-gradient by ∂p ∂xi ∇ΦF = (Φ (h1) ,··· , Φ (hn)) , i=1 Φ (hi) (3.42) for all F = p (Φ (h1) ,··· , Φ (hn)) ∈ D. (Note that ∇Φ is an unbounded operator in L2 (Ω, P), and dom (∇Φ) = D.) Lemma 3.28. Let ∇Φ be the Φ-gradient from Definition 3.27. The adjoint operator ∇∗ Φ, i.e., the Φ-divergence, is given as follows: ∇∗ Φ (G) = Φ (hi)2 − n G − ∇Φ (G) , ∀G ∈ D. (3.43) n(cid:88) (cid:32) n(cid:88) i=1 Proof. Fix F, G ∈ D as in Definition 3.2. Then ∃n ∈ N, p, q ∈ R [x1,··· , xn], and {hi}n i=1 ⊂ L , such that (cid:33) F = p (Φ (h1) ,··· , Φ (hn)) G = q (Φ (h1) ,··· , Φ (hn)) . Further assume that (cid:104)hi, hj(cid:105) = δij. In the calculation below, we use the following notation: x = (x1,··· , xn) ∈ Rn, dx = dx1 ··· dxn = Lebesgue measure, and gn = gGn = standard Gaussian distribution in Rn, see (3.15). Then, we have E ((∇ΦF ) G) Φ (hi) ∂p ∂xi (Φ (h1) ,··· , Φ (hn)) q (Φ (h1) ,··· , Φ (hn)) (cid:19) which is the desired conclusion in (3.43). Remark 3.29. Note T ∗ k is not a derivation. In fact, we have T ∗ k (F G) = T ∗ for all F, G ∈ D, and all k ∈ L . k (F ) G + F T ∗ k (G) − Φ (k) F G, However, the divergence operator ∇Φ does satisfy the Leibniz rule, i.e., ∇Φ (F G) = (∇ΦF ) G + F (∇ΦG) , ∀F, G ∈ D. (cid:3) (cid:18) = = i=1 i=1 Rn n(cid:88) E n(cid:88) = − n(cid:88) = − n(cid:88) E(cid:16) n(cid:88) = E(cid:16) F G i=1 i=1 i=1 = xi ∂p ∂xi (x) q (x) gn (x) dx ∂ ∂xi Rn Rn p (x) p (x) (cid:18) q (x) + xi (xiq (x) gn (x)) dx (cid:19) F GΦ (hi)2(cid:17) − nE (F G) − E (F∇ΦG) (cid:17)(cid:17) − E (F∇ΦG) , (cid:16)(cid:88)n (x) − q (x) x2 Φ (hi)2 − n ∂q ∂xi i i=1 (cid:18) ∂gn ∂xi (cid:19) = −xign gn (x) dx 20 3.3. Realization of the operators. Theorem 3.30. Let ωF ock be the Fock state on CCR (L ), see (2.10)-(2.11), and let πF denote the corresponding (Fock space) representation, acting on Γsym (L ), see Lemma 2.14. Let W : Γsym (L ) −→ L2 (Ω, P) be the isomorphism given by (3.44) Here L2 (Ω, P) denotes the Gaussian Hilbert space corresponding to L ; see Defini- tion 2.9. For vectors k ∈ L , let Tk denote the Malliavin derivative in the direction k; see Definition 3.2. L , 2(cid:107)k(cid:107)2 k ∈ L . W(cid:0)ek(cid:1) := eΦ(k)− 1 We then have the following realizations: Tk = W πF (a (k)) W ∗, and MΦ(k) − Tk = W πF (a∗ (k)) W ∗; (3.45) (3.46) valid for all k ∈ L , where the two identities (3.45)-(3.46) hold on the dense domain D from Lemma 2.14. Remark 3.31. The two formulas (3.45)-(3.46) take the following form, see Figs 3.1-3.2. In the proof of the theorem, we make use of the following: Lemma 3.32. Let L , CCR (L ), and ωF (= the Fock vacuum state) be as above. Then, for all n, m ∈ N, and all h1,··· , hn, k1,··· , km ∈ L , we have the following identity: ωF (a (h1)··· , a (hn) a∗ (km)··· a (k1)) (cid:10)h1, ks(1) (cid:11) L (cid:10)h2, ks(2) (cid:11) L ···(cid:10)hn, ks(n) (cid:11) = δn,m (cid:88) s∈Sn L (3.47) where the summation on the RHS in (3.47) is over the symmetric group Sn of all permutations of {1, 2,··· , n}. (In the case of the CARs, the analogous expression on the RHS will instead be a determinant.) Proof. We leave the proof of the lemma to the reader; it is also contained in [BR81]. (cid:3) Remark 3.33. In physics-lingo, we say that the vacuum-state ωF is determined by its two-point functions ωF (a (h) a∗ (k)) = (cid:104)h, k(cid:105)L , and ∀h, k ∈ L . ωF (a∗ (k) a (h)) = 0, Proof of Theorem 3.30. We shall only give the details for formula (3.45). The mod- ifications needed for (3.46) will be left to the reader. Since W in (3.44) is an isomorphic isomorphism, i.e., a unitary operator from Γsym (L ) onto L2 (Ω, P), we may show instead that TkW = W πF (a (k)) (3.48) holds on the dense subspace of all finite symmetric tensor polynomials in Γsym (L ); or equivalently on the dense subspace in Γsym (L ) spanned by ∈ Γsym (L ) , l ∈ L ; Γ (l) := el := ∞(cid:88) l⊗n√ (3.49) n=0 n! 21 (cid:3) see also Lemma 2.14. We now compute (3.48) on the vectors el in (3.49): TkW(cid:0)el(cid:1) = Tk (cid:16) 2(cid:107)k(cid:107)2 L (cid:17) (cid:0)eΦ(k)(cid:1) = W πF (a (k))(cid:0)el(cid:1) , eΦ(k)− 1 2(cid:107)k(cid:107)2 2(cid:107)k(cid:107)2 = e− 1 = e− 1 L Tk L (cid:104)k, l(cid:105)L eΦ(l) (by Lemma 2.14) (by Remark 3.3) valid for all k, l ∈ L . Γsym (L ) πF (a(k)) Γsym (L ) W W L2 (Ω, P) Tk / L2 (Ω, P) Figure 3.1. The first operator. Γsym (L ) πF (a∗(k)) Γsym (L ) W W L2 (Ω, P) MΦ(k)−Tk / L2 (Ω, P) Figure 3.2. The second operator. 3.4. The unitary group. For a given Gaussian field (L , Ω,F, P, Φ), we studied the CCR (L )-algebra, and the operators associated with its Fock-vacuum repre- sentation. From the determination of Φ by E(cid:0)eiΦ(k)(cid:1) = e− 1 (3.50) we deduce that (Ω,F, P, Φ) satisfies the following covariance with respect to the group Uni (L ) := G (L ) of all unitary operators U : L −→ L . 2(cid:107)k(cid:107)2 L , k ∈ L ; We shall need the following: Definition 3.34. We say that α ∈ Aut (Ω,F, P) iff the following three conditions hold: (2) F = α (F); more precisely, F =(cid:8)α−1 (B) B ∈ F(cid:9) where (1) α : Ω −→ Ω is defined P a.e. on Ω, and P (α (Ω)) = 1. α−1 (B) = {ω ∈ Ω α (ω) ∈ B} . (3.51) (3) P = P ◦ α−1, i.e., α is a measure preserving automorphism. / /     / / /     / Note that when (1)-(3) hold for α, then we have the unitary operators Uα in UαF = F ◦ α, (3.52) 22 (UαF ) (ω) = F (α (ω)) , a.e. ω ∈ Ω, L2 (Ω,F, P), or more precisely, valid for all F ∈ L2 (Ω,F, P). Theorem 3.35. Aut (Ω,F, P) s.t. (1) For every U ∈ G (L ) (= the unitary group of L ), there is a unique α ∈ Φ (U k) = Φ (k) ◦ α, (3.53) or equivalently (see (3.52)) (3.54) (2) If T : L2 (Ω, P) −→ L2 (Ω, P)⊗L is the Malliavin derivative from Definition Φ (U k) = Uα (Φ (k)) , ∀k ∈ L . 3.2, then we have: T Uα = (Uα ⊗ U ) T. (3.55) Proof. The first conclusion in the theorem is immediate from the above discussion, and we now turn to the covariance formula (3.55). Note that (3.55) involves unbounded operators, and it holds on the dense sub- space D in L2 (Ω, P) from Lemma 2.14. Hence it is enough to verify (3.55) on L , k ∈ L . Using Lemma 2.14, we then vectors in L2 (Ω, P) of the form eΦ(k)− 1 get: 2(cid:107)k(cid:107)2 (cid:0)eΦ(k)− 1 2(cid:107)k(cid:107)2 LHS(3.55) L T(cid:0)eΦ(U k)(cid:1) L(cid:1) = e− 1 = (Uα ⊗ U )(cid:0)eΦ(k)− 1 2(cid:107)k(cid:107)2 2(cid:107)U k(cid:107)2 L eΦ(U k) ⊗ (U k) 2(cid:107)k(cid:107)2 L(cid:1) = e− 1 = RHS(3.55) (by (3.53)) (by Remark 3.3) (cid:3) 4. The Fock-state, and representation of CCR, realized as Malliavin calculus We now resume our analysis of the representation of the canonical commutation relations (CCR)-algebra induced by the canonical Fock state (see (2.9)). In our analysis below, we shall make use of the following details: Brownian motion, It¯o- integrals, and the Malliavin derivative. The general setting. Let L be a fixed Hilbert space, and let CCR (L ) be the ∗-algebra on the generators a (k), a∗ (l), k, l ∈ L , and subject to the relations for the CCR-algebra, see Section 2.2: and [a (k) , a (l)] = 0, [a (k) , a∗ (l)] = (cid:104)k, l(cid:105)L 1 (4.1) (4.2) where [·,·] is the commutator bracket. A representation π of CCR (L ) consists of a fixed Hilbert space H = Hπ (the representation space), a dense subspace Dπ ⊂ Hπ, and a ∗-homomorphism π : CCR (L ) −→ End (Dπ) such that 23 (4.3) The representation axiom entails the commutator properties resulting from (4.1)- (4.2); in particular π satisfies Dπ ⊂ dom (π (A)) , ∀A ∈ CCR. [π (a (k)) , π (a (l))] F = 0, (cid:2)π (a (k)) , π (a (l)) ∗(cid:3) F = (cid:104)k, l(cid:105)L F, and (4.4) (4.5) ∀k, l ∈ L , ∀F ∈ Dπ; where π (a∗ (l)) = π (a (l)) (Ω,FΩ, P) is the standard Wiener probability space, and In the application below, we take L = L2 (0,∞), and Hπ = L2 (Ω,FΩ, P) where ∗ . Φt (ω) = ω (t) , ∀ω ∈ Ω, t ∈ [0,∞). (4.6) For k ∈ L , we set ∞ Φ (k) = k (t) dΦt (=the It¯o-integral.) The dense subspace Dπ ⊂ Hπ is generated by the polynomial fields: For n ∈ N, h1,··· , hn ∈ L = L2R (0,∞), p ∈ Rn −→ R a polynomial in n real 0 variables, set F = p (Φ (h1) ,··· , Φ (hn)) , and (cid:19) (cid:18) ∂ n(cid:88) ∂xj j=1 π (a (k)) F = (Φ (h1) ,··· , Φ (hn))(cid:104)hj, k(cid:105) . p (4.7) (4.8) It follows from Lemma 3.14 that Dπ is an algebra under pointwise product and that π (a (k)) (F G) = (π (a (k)) F ) G + F (π (a (k)) G) , (4.9) ∀k ∈ L , ∀F, G ∈ Dπ. Equivalently, Tk := π (a (k)) is a derivation in the algebra Dπ (relative to pointwise product.) Theorem 4.1. With the operators π (a (k)), k ∈ L , we get a ∗-representation π : CCR (L ) −→ End (Dπ), i.e., π (a (k)) = the Malliavin derivative in the direction k, π (a (k)) F = (cid:104)T (F ) , k(cid:105)L , (4.10) (cid:3) Proof. We begin with the following Lemma 4.2. Let π, CCR (L ), and Hπ = L2 (Ω,FΩ, P) be as above. For k ∈ L , we shall identify Φ (k) with the unbounded multiplication operator in Hπ: ∀F ∈ Dπ, ∀k ∈ L . For F ∈ Dπ, we have π (a (k)) ∗ ∗ π (a (k)) valid on the dense domain Dπ ⊂ Hπ. Dπ (cid:51) F (cid:55)−→ Φ (k) F ∈ Hπ. (4.11) F = −π (a (k)) F + Φ (k) F ; or in abbreviated form: (4.12) = −π (a (k)) + Φ (k) Proof. This follows from the following computation for F, G ∈ Dπ, k ∈ L . Setting Tk := π (a (k)), we have 24 E (Tk (F ) G) + E (F Tk (G)) = E (Tk (F G)) = E (Φ (k) F G) . Hence Dπ ⊂ dom (T ∗ conclusion (4.12). k ), and T ∗ k (F ) = −Tk (F ) + Φ (k) F , which is the desired (cid:3) Proof of Theorem 4.1 continued. It is clear that the operators Tk = π (a (k)) form a commuting family. Hence on Dπ, we have for k, l ∈ L , F ∈ Dπ: [Tk, T ∗ l ] (F ) = [Tk, Φ (l)] (F ) = Tk (Φ (l) F ) − Φ (l) (Tk (F )) = Tk (Φ (l)) F = (cid:104)k, l(cid:105)L F by (4.12) by (4.9) by (4.8) which is the desired commutation relation (4.2). The remaining check on the statements in the theorem are now immediate. (cid:3) Corollary 4.3. The state on CCR (L ) which is induced by π and the constant function 1 in L2 (Ω, P) is the Fock-vacuum-state, ωF ock. Proof. The assertion will follow once we verify the following two conditions: Ω Ω T ∗ k Tk (1) dP = 0 TkT ∗ l (1) dP = (cid:104)k, l(cid:105)L (4.13) (4.14) and for all k, l ∈ L . l ) (1) = (cid:104)k, l(cid:105)L 1. See (3.13). This in turn is a consequence of our discussion of eqs (2.10)-(2.11) above: The Fock state ωF ock is determined by these two conditions. The assertions (4.13)-(4.14) (cid:3) follow from Tk (1) = 0, and (TkT ∗ Corollary 4.4. For k ∈ L2R (0,∞) we get a family of selfadjoint multiplication operators Tk + T ∗ k = MΦ(k) on Dπ where Tk = π (a (k)). Moreover, the von Neu- mann algebra generated by these operators is L∞ (Ω, P), i.e., the maximal abelian L∞-algebra of all multiplication operators in Hπ = L2 (Ω, P). Remark 4.5. In our considerations of representations π of CCR (L ) in a Hilbert space Hπ, we require the following five axioms satisfied: (4) (cid:2)π (a (k)) , π (a (l)) (1) a dense subspace Dπ ⊂ Hπ; (2) π : CCR (L ) −→ End (Dπ), i.e., Dπ ⊂ ∩A∈CCR(L )dom (π (A)); (3) [π (a (k)) , π (a (l))] = 0, ∀k, l ∈ L ; (5) π (a∗ (k)) ⊂ π (a (k)) Note that in our assignment for the operators π (a (k)), and π (a∗ (k)) in Lemma 4.2, we have all the conditions (1)-(5) satisfied. We say that π is a selfadjoint representation. ∗(cid:3) = (cid:104)k, l(cid:105)L IHπ, ∀k, l ∈ L ; and ∗, ∀k ∈ L . If alternatively, we define ρ : CCR (L ) −→ End (Dπ) (4.15) with the following modification:(cid:40) ρ (a (k)) = Tk, k ∈ L , and ρ (a∗ (k)) = Φ (k) 25 (4.16) then this ρ will satisfy (1)-(3), and [ρ (a (k)) , ρ (a∗ (l))] = (cid:104)k, l(cid:105)L IHπ ; but then ρ (a (k)) (cid:38) ρ (a (k)) ∗; i.e., non-containment of the respective graphs. One generally says that the representation π is (formally) selfadjoint, while the second representation ρ is not. 5. Conclusions: the general case Definition 5.1. A representation π of CCR (L ) is said to be admissible iff (Def.) ∃ (Ω,F, P) as above such that Hπ = L2 (Ω,F, P), and there exists a linear mapping Φ : L −→ L2 (Ω,F, P) subject to the condition: For every n ∈ N, and every k, h1,··· , hn ∈ L , the following holds on its natural dense domain in Hπ: For every p ∈ R [x1,··· , xn], we have (cid:104)k, hi(cid:105)L M ∂p π ([a (k) , p (a∗ (h1) ,··· , a∗ (hn))]) = n(cid:88) (5.1) (Φ(h1),··· ,Φ(hn)), ∂xi with the M on the RHS denoting “multiplication.” Corollary 5.2. i=1 (1) Every admissible representation π of CCR (L ) yields an associated Malli- avin derivative as in (5.1). (2) The Fock-vacuum representation πF is admissible. Proof. (1) follows from the definition combined with Corollary 2.8. (2) is a direct (cid:3) consequence of Lemma 3.7 and Theorem 3.8; see also Corollary 4.3. Acknowledgement. The co-authors thank the following colleagues for helpful and en- lightening discussions: Professors Sergii Bezuglyi, Ilwoo Cho, Paul Muhly, Myung- Sin Song, Wayne Polyzou, and members in the Math Physics seminar at The Uni- versity of Iowa. References [AJ12] [AJ15] [AJL11] [AJS14] [AØ15] Daniel Alpay and Palle E. T. Jorgensen, Stochastic processes induced by singular operators, Numer. Funct. Anal. Optim. 33 (2012), no. 7-9, 708–735. MR 2966130 Daniel Alpay and Palle Jorgensen, Spectral theory for Gaussian processes: reproducing kernels, boundaries, and L2-wavelet generators with fractional scales, Numer. Funct. Anal. Optim. 36 (2015), no. 10, 1239–1285. MR 3402823 Daniel Alpay, Palle Jorgensen, and David Levanony, A class of Gaussian pro- cesses with fractional spectral measures, J. Funct. Anal. 261 (2011), no. 2, 507–541. MR 2793121 (2012e:60101) Daniel Alpay, Palle Jorgensen, and Guy Salomon, On free stochastic processes and their derivatives, Stochastic Process. Appl. 124 (2014), no. 10, 3392–3411. MR 3231624 Nacira Agram and Bernt Øksendal, Malliavin Calculus and Optimal Control of Sto- chastic Volterra Equations, J. Optim. Theory Appl. 167 (2015), no. 3, 1070–1094. MR 3424704 [Arv76a] William Arveson, Aspectral theorem for nonlinear operators, Bull. Amer. Math. Soc. 82 (1976), no. 3, 511–513. MR 0417882 (54 #5930) 26 [Arv76b] [AW63] , Spectral theory for nonlinear random processes, Symposia Mathematica, Vol. XX (Convegno sulle Algebre Csp∗ e loro Applicazioni in Fisica Teorica, Convegno sulla Teoria degli Operatori Indice e Teoria K, INDAM, Rome, 1975), Academic Press, London, 1976, pp. 531–537. MR 0474479 (57 #14118) H. Araki and E. J. Woods, Representations of the canonical commutation relations describing a nonrelativistic infinite free Bose gas, J. Mathematical Phys. 4 (1963), 637–662. MR 0152295 (27 #2275) Rep. Mathematical Phys. 4 (1973), 227–254. MR 0330350 (48 #8687) [AW73] , Topologies induced by representations of the canonical commutation relations, [BJKR84] O. Bratteli, P. E. T. Jorgensen, A. Kishimoto, and D. W. Robinson, A C∗- algebraic Schoenberg theorem, Ann. Inst. Fourier (Grenoble) 34 (1984), no. 3, 155–187. MR 762697 (86b:46105) [BØSW04] Francesca Biagini, Bernt Øksendal, Agnès Sulem, and Naomi Wallner, An introduction to white-noise theory and Malliavin calculus for fractional Brownian motion, Proc. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci. 460 (2004), no. 2041, 347–372, Stochastic analysis with applications to mathematical finance. MR 2052267 (2005a:60107) Ola Bratteli and Derek W. Robinson, Operator algebras and quantum statistical me- chanics. Vol. 1, Springer-Verlag, New York-Heidelberg, 1979, C∗- and W ∗-algebras, algebras, symmetry groups, decomposition of states, Texts and Monographs in Physics. MR 545651 (81a:46070) [BR79] [BR81] [Dix77] [DS88] [GJ87] [Gro70] [Hid80] [Hid03] [JP91] [JT14] [Pri10] [PS72a] [PS72b] [Sak98] , Operator algebras and quantum-statistical mechanics. II, Springer-Verlag, New York-Berlin, 1981, Equilibrium states. Models in quantum-statistical mechanics, Texts and Monographs in Physics. MR 611508 (82k:82013) Jacques Dixmier, Enveloping algebras, North-Holland Publishing Co., Amsterdam- New York-Oxford, 1977, North-Holland Mathematical Library, Vol. 14, Translated from the French. MR 0498740 (58 #16803b) Nelson Dunford and Jacob T. Schwartz, Linear operators. Part II, Wiley Classics Li- brary, John Wiley & Sons Inc., New York, 1988, Spectral theory. Selfadjoint operators in Hilbert space, With the assistance of William G. Bade and Robert G. Bartle, Reprint of the 1963 original, A Wiley-Interscience Publication. MR 1009163 (90g:47001b) James Glimm and Arthur Jaffe, Quantum physics, second ed., Springer-Verlag, New York, 1987, A functional integral point of view. MR 887102 (89k:81001) Leonard Gross, Abstract Wiener measure and infinite dimensional potential theory, Lectures in Modern Analysis and Applications, II, Lecture Notes in Mathematics, Vol. 140. Springer, Berlin, 1970, pp. 84–116. MR 0265548 (42 #457) Takeyuki Hida, Brownian motion, Applications of Mathematics, vol. 11, Springer- Verlag, New York, 1980, Translated from the Japanese by the author and T. P. Speed. MR 562914 (81a:60089) , Laplacians in white noise analysis, Finite and infinite dimensional analysis in honor of Leonard Gross (New Orleans, LA, 2001), Contemp. Math., vol. 317, Amer. Math. Soc., Providence, RI, 2003, pp. 137–142. MR 1966892 (2005b:60172) Palle E. T. Jorgensen and Robert T. Powers, Positive elements in the algebra of the quantum moment problem, Probab. Theory Related Fields 89 (1991), no. 2, 131–139. MR 1110533 (92k:47090) P. Jorgensen and F. Tian, Noncommutative analysis, Multivariable spectral theory for operators in Hilbert space, Probability, and Unitary Representations, ArXiv e-prints (2014). Nicolas Privault, Random Hermite polynomials and Girsanov identities on the Wiener space, Infin. Dimens. Anal. Quantum Probab. Relat. Top. 13 (2010), no. 4, 663–675. MR 2754322 (2012a:60160) K. R. Parthasarathy and K. Schmidt, Factorisable representations of current groups and the Araki-Woods imbedding theorem, Acta Math. 128 (1972), no. 1-2, 53–71. MR 0390124 (52 #10950) , Positive definite kernels, continuous tensor products, and central limit theo- rems of probability theory, Lecture Notes in Mathematics, Vol. 272, Springer-Verlag, Berlin-New York, 1972. MR 0622034 (58 #29849) Shôichirô Sakai, C∗-algebras and W ∗-algebras, Classics in Mathematics, Springer- Verlag, Berlin, 1998, Reprint of the 1971 edition. MR 1490835 (98k:46085) 27 [Šil47] G. E. Šilov, On a property of rings of functions, Doklady Akad. Nauk SSSR (N. S.) 58 (1947), 985–988. MR 0027131 (10,258b) [VFHN13] Frederi Viens, Jin Feng, Yaozhong Hu, and Eulalia Nualart (eds.), Malliavin calculus and stochastic analysis, Springer Proceedings in Mathematics & Statistics, vol. 34, Springer, New York, 2013, A Festschrift in honor of David Nualart. MR 3155263 (Palle E.T. Jorgensen) Department of Mathematic s, The University of Iowa, Iowa City, IA 52242-1419, U.S.A. E-mail address: [email protected] URL: http://www.math.uiowa.edu/~jorgen/ (Feng Tian) Department of Mathematics, Hampton University, Hampton, VA 23668, U.S.A. E-mail address: [email protected]
1703.00546
1
1703
2017-03-01T23:31:33
Noncommutative versions of the arithmetic-geometric mean inequality
[ "math.OA" ]
Recht and R\'{e} introduced the noncommutative arithmetic geometric mean inequality (NC-AGM) for matrices with a constant depending on the degree $d$ and the dimension $m$. In this paper we prove AGM inequalities with a dimension-free constant for general operators. We also prove an order version of the AGM inequality under additional hypothesis. Moreover, we show that our AGM inequality almost holds for many examples of random matrices .
math.OA
math
NONCOMMUTATIVE VERSIONS OF THE ARITHMETIC-GEOMETRIC MEAN INEQUALITY WAFAA ALBAR, MARIUS JUNGE, AND MINGYU ZHAO Recht and R´e in [21] introduced the noncommutative arithmetic geometric mean inequality (NC-AGM) for matrices with a constant depending on the degree d and the dimension m. In this paper we prove AGM inequalities with a dimension-free constant for general operators. We also prove an order version of the AGM in- equality under additional hypothesis. Moreover, we show that our AGM inequality almost holds for many examples of random matrices . 1. Introduction Variations of the arithmetic-geometric mean (AGM) inequality have many applications in analysis and geometry. As pointed out by R´e and Recht in [21], noncommutative versions of the AGM inequalities are relevant to machine learning. In particular, their proof, which employed the classical MacLaurin inequalities, led to improved convergence rate of the of the algorithms in machine learning. Let us recall the famous MacLaurin inequalities for positive real numbers x1, ..., xn and the normalized d-th symmetric sums as Sd =(cid:18)n d(cid:19)−1 xi . τ=kYi∈τ Xτ∈[n] where 1 ≤ d ≤ n. According to the MacLaurin inequalities, we have In particular, S1 ≥ n√Sn is the standard AGM inequality. For more details about S1 ≥ 2√S2 ≥ 3√S3 ≥ ... ≥ n√Sn . the classical AGM inequality see [5]. In this paper, we will discuss noncommutative versions of MacLaurin's inequalities. Indeed, we will consider a generalized AGM inequality for the norm and the order. It may come as a surprise to the operator algebra community that these inequalities are motivated by problems in machine learning, stochastic gradient method (see Buttou [2] and the reference there is in [21]), and randomized coordinates descent (see Nesterov [15]). This interesting connection and an overview of known results on this topic can be found in [20] and [21]. In fact, these methods contain an iteration procedure which can be performed with or without replacement samples. Recht and R´e, in [20], study the performance of both. They show that the expected convergence rate without replacement is faster than that with replacement. They proved this result by using a particular AGM inequality. Partially supported by DMS 1501103 and BigData 1447879. 1 2 WAFAA ALBAR, MARIUS JUNGE, AND MINGYU ZHAO In the effort to generalize the classical AGM inequality to the noncommutative setting, a standard but naive procedure in noncommutative analysis is to replace scalars by operators. Famous examples of this strategy are Cauchy-Schwarz type in- equalities for C∗-modules, Khintchine, and martingale inequalities.(See e.g. LP[12], LPP[13], PXu[18], Narcisse[19], J[6] , JXu1[9], JXu2[9]. For a general survey see [18].) Proving these noncommutative extensions often employs a combination of functional analytic and combinatorial methods. In fact, the key results of this pa- per heavily rely on Pisier's interpretation of Rota's Mobius formulae for partitions. A NC-AGM inequality would ask whether (1.1) A1 ··· An ? ≤ ( 1 n Aj)n n Xj=1 holds for positive operators A1, ..., An on a Hilbert space. (In this context we shall interpret x ≤ y as requiring that y − x is positive semi-definite.) However, for positive operators A and B, the product AB may not be positive or even self- adjoint, so the inequality (1.1) may not make sense. Inspired by Recht and R´e, we modify (1.1) by replacing the left hand side with the average of all the products of the operators Ai, which turns out to be self-adjoint. Following the MacLaurin approach, we may now ask whether the AGM inequality holds on average, i.e. (1.2) 1 n! Xσ∈Sn Aσ(1) ··· Aσ(n) ? ≤ ( 1 n n Xj=1 Aj)n. Unfortunately, we can not prove (1.2) in general. A milder version of (1.2) is to ask for (1.3) k 1 n! Xσ∈Sn Aσ(1) ··· Aσ(n)k ? ≤ k( 1 n n Xj=1 Aj )nk , where kxk = kxkB(H) refers to the standard operator norm of bounded operators on a Hilbert space H. The inequality (1.3) is a particular case of the noncommutative MacLaurin inequalities discussed in [20]. Indeed, for fixed d we may consider the following average product of noncommutative operators of length d: Pd(A1, ..., An) = 1 n··· (n − d + 1) X 1≤j1,...,jd≤n all different Aj1 ··· Ajd . We refer to the example in [21] for the fact that the symmetrization for the operators in the AGM inequality is required. In [20], R´e and Recht posed the following question: Is it true that for positive bounded operators A1, ..., An on a Hilbert space one has (1.4) kPd(A1, ..., An)k1/d ≤ kP1(A1, ..., An)k ? They proved that (1.4) holds when A1, ..., An are matrices that mutually commute. Moreover, they observed that for operators A1, ..., An on an m-dimensional Hilbert space one has (1.5) kPd(A1, ..., An)k1/d B(ℓm 2 ) ≤ m kP1(A1, ..., An)k . NONCOMMUTATIVE VERSIONS OF THE AGM INEQUALITY 3 We will prove the AGM inequality for the norm with a constant independent of the dimension m. Theorem 1.1. For operators A1, ..., An ≥ 0 on a Hilbert space H, kPd(A1, ..., An)k1/d ≤ d kP1(A1, ..., An)k. Let us now consider the "order version" of the AGM inequality. Here we add generally, it is good to start with d = 3. the additional assumption P Ai = n. In order to illustrate the technique we use Theorem 1.2. Let n ≥ 6. If A1, ..., An are self-adjoint operators such thatPi Ai = n. Then P3(A1, ..., An)1/3 ≤ 1. For the proof we consider the mean-zero operators ai := Ai − 1. Observe the operators ai are self-adjoint and ai = 0. It follows easily that n Pi=1 P1(a1, ..., an) = (n−1)! 1(cid:19)P1(a1, ..., an) +(cid:18)3 P3(A1, ..., An) = 1 +(cid:18)3 Straightforward computations using P ai = 0 reveal that n! Pi n! (cid:16)(P ai)2 −P a2 n! Pi6=j (n − 3)! P2(a1, ..., an) = (n−2)! aiaj = (n−2)! ai = 0 = n! n! a2 = 2 aiajak) (n − 3)! (n − 3)! P3(a1, ..., an) = i )(Xk ak) − (Xi ( Xi6=j6=k (cid:16)( Xi6=j6=k ai)3 − (Xi=j i(cid:17). (cid:16)Xi This leads to the form P3(A1, ..., An) = 1 − 3 gether with P a3 i ≤ kaikP a2 i ≤ nP a2 i , this yields n(n − 1)X a2 P3(A1, ..., An) ≤ 1 − (1.6) i + a3 n! 3 2(cid:19)P2(a1, ..., an) +(cid:18)3 3(cid:19)P3(a1, ..., an). i(cid:17) = − (n−2)! n! P a2 i ai)(Xj=k a2 j ) −Xj Xi=k6=j aiajai + 2( Xi=j=k a3 i )(cid:17) i . To- 2 i + n(n−1)(n−2)P a3 n(n−1)P a2 n(n − 1)(n − 2)X a2 2n i . 2n n(n−1)(n−2) ≤ 3 n(n−1) holds for all n ≥ 6, the right side of (1.6) is at most 1 Since and we are done. A far-reaching generalization of this idea leads to the following result. Theorem 1.3. Fix n and d. Suppose A1, ..., An and ai are defined as above, Ai = n and Pi i) P1(A1, ..., An) = Pi Ai n = 1, j ) ii) k(P a2 1 2k ≤ n 3d . Then the AGM inequality holds in the order sense: Pd(A1, ..., An) ≤ P1(A1, ..., An)d = 1. 4 WAFAA ALBAR, MARIUS JUNGE, AND MINGYU ZHAO Note that these techniques work efficiently when d is very large. This paper is organized as follows. In section 2, we review the analytic and com- binatorial tools needed to prove Theorem 1.2, especially Pisier's interpretation of Rota's results on Mobius transforms for partitions. In section 3, we combine the results from section 2 with Pisier's group construction for partitions in [17] to ob- tain our key norm and order estimate. In section 4 and 5, a combination of Pisier's partition method and probabilistic results allow "almost AGM" inequalities hold in many different scenarios. We confirm the AGM inequality up to ε for many ran- dom matrices, in particular Wishart random matrices, more general vector-valued moments of convex bodies, and freely independent operators. We should point out that in contrast to results on averages of random matrices in R´e and Recht in [21], our estimates hold with high probabilities. 2. Partition and Mobius Formula We need some definitions from the combinatorial theory of partitions. Let Pd be the lattice of all the partitions of {1, ..., d}. For two partitions σ and π, we write σ ≤ π if every block of the partition σ is contained in some block of π (i.e., any block of the partition of π can be written as a union of blocks of σ). In other words, π is a refinement of σ. There are two trivial partitions, 0 and 1, where 0 is the partition into n singletons and 1 is the partition of a single block. For a partition π, ν(π) is the number of the blocks of the partition π and ri(π) is the number of i=1 ri(π) = ν(π). blocks of π with cardinality i such that Pd For more information on partitions, see [1] and [22]. i=1 iri(π) = d; and Pd Let us recall some main results on the Mobius function µ in [17] which are crucial for our paper. Proposition 2.1. (Pisier's Mobius inversion formula) For any d ∈ N there exists a function µ : Pd × Pd −→ Z such that for every vector space V and functions φ : Pd −→ V and ψ : Pd −→ V , we have the following properties: i) If ψ(σ) = Pπ≤σ ii) If ψ(σ) = Pπ≥σ iii) Moreover, ∀σ 6= 0, P0≤π≤σ φ(π), then φ(σ) = Pπ≤σ φ(π), then φ(σ) = Pπ≥σ µ(π, σ) = 0. µ(π, σ)ψ(π); µ(σ, π)ψ(π); The next result provides precise formulas for the Mobius function µ in special cases. Theorem 2.2. The Mobius function satisfies the following properties: i=1[(−1)i−1(i − 1)!]ri(π), and consequently, i) µ( 0, 1) = (−1)d−1(d − 1)!. ii) µ( 0, π) =Qd iii) Pπ∈Pd µ( 0, π) = d!. If σ is a partition of {1, ..., d}, then there exists a coordinate function f : {1, ..., d} → {1, ..., ν(σ)} such that f−1(t) = At where each At represents a block in our par- tition. Note that this coordinate function isn't unique. For every partition σ we can fix an enumeration of the blocks f : {1, 2, ..., d} −→ {1, 2, ...,σ} where σ:=hj1, j2, ..., jdi. This means jr = js if and only if r, s ∈ Ar,s where Ar,s is a block NONCOMMUTATIVE VERSIONS OF THE AGM INEQUALITY 5 in σ = hj1, j2, ..., jdi. Using this notation we define the restricted and full partition for elements from an algebra. Definition 2.3. Let A be an algebra and xi defined by: hσi = Xhj1,j2,...,jdi=σ The full partition with elements xji is given by: [σ] = Xπ≥σ hπi. ji ∈ A . The restricted partition is j1 ...xd x1 jd . The restricted and full partitions, which are denoted as hσi and [σ], respectively, give expressions for the elements in the given B(H) according to the algebraic com- binatorial partition σ. In order to understand the difference between the definition of restricted partition and full partition, consider the following example. Example 2.4. Let both the numbers of total samples and chosen samples be 3 (n = d = 3). Then for the full partition [1 2, 3] , with the assumption that xi j = xj we have [1 2, 3] = (X x2 i )(X xi) = h1 2, 3i + h1 2 3i i1 xi3 + Xi1=i2=i3 = Xi1=i26=i3 x2 x3 i1 . Whereas the restricted partition h1 2, 3i is defined as h1 2, 3i =Pi1=i26=i3 We reformulate Pisier's Mobius inversion formula in our context. x2 i1 xi3 . Proposition 2.5. Let xk (2.1) (2.2) Moreover, we have (2.3) hπi = Xν≥π hπi = Xν≤π j ∈ A as above. Then we have µ(π, ν)[ν], where [π] = Xν≥π µ(π, ν)[ν], where [π] = Xν≤π h 0i = [ 0] + X0(cid:8)ν≤ 1 µ( 0, ν)[ν]. hνi, hνi. In [17], in order to separate different partition blocks into disjoint subspaces, Pisier uses a trick to embed operators xik ∈ B(H) into B(K ⊗ H) (for another Hilbert space K). Our first goal is to modify Pisier's trick by using matrix units. Consider first the trivial partition that has only one block [1, 2,··· , d]. We can write 1 = [1, 2,··· , d] = X x1 i1 id i2 ··· xd x2 = (X e1i1 ⊗ x1 × (X eid−1id−1 ⊗ xd−1 i1 ) × (X ei2i2 ⊗ x2 i2 ) × ··· ) × (X eid1 ⊗ xd id ). id−1 6 WAFAA ALBAR, MARIUS JUNGE, AND MINGYU ZHAO Now if we have 6 elements and our partition σ has two crossing blocks, one con- taining {1, 3, 4, 6} and the other containing {2, 5}, then the full partition of σ will be of the form: [σ] = Xi1=i3=i4=i6 i2=i5 xi1 xi2 xi3 xi4 xi5 xi6 . We rewrite these elements into a tensor form, as follows: Zi1 = e1i1 ⊗ 1 ⊗ xi1 , Zi3 = ei3i3 ⊗ 1 ⊗ xi3 , Zi5 = 1 ⊗ ei51 ⊗ xi5 , Zi2 = 1 ⊗ e1i2 ⊗ xi2 Zi4 = ei4i4 ⊗ 1 ⊗ xi4 Zi6 = ei61 ⊗ 1 ⊗ xi6 . With this new notation, we get xi1 xi2 xi3 xi4 xi5 xi6 Zi1 Zi2 Zi3 Zi4 Zi5 Zi6 = i2=i5 [σ] = Xi1=i3=i4=i6 Xi1,i2,i3,i4,i5,i6 (Xij Yj=1 = 6 Zij ) = Zj, 6 Yj=1 (Zj :=Xij Zij ). In a more general setting, assume σ has more than one block. Denote A1,. . . ,Aσ as the blocks of the partition σ with cardinality larger than one. Then we define (2.4) as follows: where jk ∈ B(H)⊗σ Z k ⊗ B(H) ∀k ∈ A1, Z k ∀k ∈ A2, Z k ∀k ∈ Aσ, Z k jk = tA1(jk) ⊗ 1 ⊗ ··· ⊗ xk jk = 1 ⊗ tA2(jk) ⊗ 1 ··· ⊗ xk jk = 1 ⊗ ··· ⊗ tAσ (jk) ⊗ xk jk jk jk , jk = min Am otherwise jk = max Am. e1jk ejkjk ejk1 tAm(jk) =  Z k jk = 1 ⊗ ··· ⊗ 1 ⊗ xk . jk Here, min Am means the smallest index number and max Am means the largest index number in the partition Am. Finally, if k belongs to singleton block of the partition σ, then we set To sum up, the method places each element into larger spaces, which will allow us to interchange the summation and multiplication as in the above example and the following lemma. Lemma 2.6. For an arbitrary partition σ for d elements, we have [σ] = Xi1,...,id i1 ...Z d Z 1 id . NONCOMMUTATIVE VERSIONS OF THE AGM INEQUALITY 7 Indeed, this immediately follows from ij · Z k Z i ik = 0, if ij 6= ik. Follow Pisier's result in [17]; we deduce the following norm estimate. Theorem 2.7. For an arbitrary partition σ for d elements, we have d Z k k[σ]kB(H) ≤ Yk=1(cid:16)kXjk jkk · 1σs(k) + kXjk Moreover, k[σ]kB(H) ≤Qk∈σs kPjk Zjkk ×Qk∈σns k(Zjk )k, where k(Zjk )k = max{kP Zjk1 2 , kP Z∗jkp Z∗jk1k Zjkpk Z k 1 1 jkk · 1σns(k)(cid:17). 2 , supjk kZjkk}. Here σs means the set of singletons in the partition σ, and σns means the set of non-singleton elements in the partition σ. The functions 1σns(k), 1σs(k) represent the characteristic functions, i.e. 1σns(k) =(cid:26) 1 k ∈ σns 0 otherwise , 1σs(k) =(cid:26) 1 k ∈ σs 0 otherwise Proof. Taking the norm for the full partition, we have j1 ··· xd x1 jdk The equality (2.5) comes from Lemma 2.6. The equality (2.6) follows from the definition of Z k jk , which means it allows us to perform summation first and then multiplication. Next, Zjkk kXjk Zjkk · (1min Am + 1max Am + 1mid Am) kXjk Zjkk · Yk∈σns kXjk Zjkk × Yk∈Am⊂σns Zjkk× k[σ]kB(H) ≤ Yk∈σs ≤ Yk∈σs ≤ Yk∈σs Yk∈Am⊂σns(cid:16)kXjk kXjk kXjk Zjkk · 1min Am + kXjk Zjkk · 1max Am + kXjk Zjkk · 1mid Am(cid:17) (2.5) (2.6) k[σ]k = kXπ≥σ = k Xj1,j2,...,jd = k Yk∈σsXjk ≤ k Yk∈σsXjk kXjk ≤ Yk∈σs hπik = k Xhj1,··· ,jdi≥σ j1 ··· Z d Z 1 jdk jk · Yk∈σnsXjk Z k jkk jkk · k Yk∈σnsXjk kXjk jkk · Yk∈σns Z k Z k Z k Z k jkk Z k jkk. 8 WAFAA ALBAR, MARIUS JUNGE, AND MINGYU ZHAO 1 Zjkk× kXjk ≤ Yk∈σs Yk∈Am⊂σns(cid:16)kXjk 2 · 1min Am + kXjk ≤ Yk∈σs k(Zjk )k, where k(Zjk )k = max{kP Zjk1 The next corollary states the norm estimate in B(H) rather than in B(K ⊗ H). For simplicity we replace xk Zjk Z∗jkk Zjkk × Yk∈σns 2 , kP Z∗jkp 2 · 1max Am + sup 2 , supjk kZjkk}. jk kZjkk · 1mid Am(cid:17) kXjk Z∗jk1k Zjkpk Zjkk Z∗jk 1 1 1 ik by xik . Corollary 2.8. If σ is a partition and xjk is a self-adjoint operator for arbitrary k ∈ {1, ..., d}, then Proof. We need to discuss two cases: 1 2 kX x2 jkk kX xjkk · Yk∈σns k[σ]kB(H) ≤ Yk∈σs Zjkk = kP 1 ⊗ ··· ⊗ xjkk = k1 ⊗ ··· ⊗P xjkk = kP xjkk. 2 = kX[1 ⊗ ··· ⊗ e1jk1 ⊗ ··· ⊗ xjk1 ] · [1 ⊗ ··· ⊗ ejk1 1 ⊗ ··· ⊗ x∗jk1 = kX 1 ⊗ ··· ⊗ e11 ⊗ ··· ⊗ xjk1 x∗jk1k = k1 ⊗ ··· ⊗X xjk1 2 = kX xjk1 2 = kX[1 ⊗ ··· ⊗ e1jkp ⊗ ··· ⊗ x∗jkp = kX 1 ⊗ ··· ⊗ e11 ⊗ ··· ⊗ x∗jkp xjkpk = kX x∗jkp 2 = kX x2 1 jkpk 2 . ] · [1 ⊗ ··· ⊗ ejkp 1 ⊗ ··· ⊗ xjkp ]k 2 = k1 ⊗ ··· ⊗X x∗jkp xjkpk 2 = kX x2 jk1k x∗jk1k x∗jk1k xjkp k 1 2 . 1 2 1 1 1 1 1 1 2 ]k 1 2 1 2 (i) For k ∈ σs, kPj (ii) For Am ∈ σns, kX Zjk1 Z∗jk1k 1 (2.7) and kX Z∗jkp Zjkpk (2.8) For the middle term, we have sup sup jk kZjkk = sup k∈{k2,...,kp−1} (2.9) jk kZ∗jk sup k∈Am kX x2 jkk ≤ sup Combining (i) and (ii) finishes the proof. k∈{k2,...,kp−1} 1 2 . 1 2 = Zjkk sup k∈{k2,...,kp−1} jk kx∗jk sup xjkk 1 2 3. AGM inequality for the norm and for the order In this section we prove the AGM inequality for the norm and for the order. We need the following lemma which handles positive or self-adjoint operators {xik} in a C*-algebra A . Lemma 3.1. (i) If xjk ≥ 0, then kP x2 2 ≤ kP xjkk. jkk 2 = k(P x2 (ii) If xjk are self-adjoint, then kP x2 2k. jkk jk ) 1 1 1 NONCOMMUTATIVE VERSIONS OF THE AGM INEQUALITY 9 Proof. (i) Indeed, we have kX x2 jkk 1 1 2 jk 2 = kX x ≤ (kX xjkk = kX xjkk. 1 2 1 2 xjk x 1 jkk 2 · kX xjkk · kX xjkk 1 2 ) 1 2 (ii) Holds trivially using kx2k = kxk2, for x = (P x2 1 2 . jk ) 3.1. AGM inequality for the norm. Now we have done all the preparation to prove the NC-AGM inequality for the norm. Theorem 3.2. Suppose x1, . . . , xn are positive operators in B(H). Then (3.1) kPd(x1, ..., xn)k1/d B(H) ≤ d kP1(x1, ..., xn)kB(H). Proof. From Corollary 2.8 and Lemma 3.1, we deduce that for a given arbitrary partition σ and positive elements xjk = xj, we have Recall identity 2.3 from Proposition 2.5: k[σ]kB(H) ≤ kX xjkd. (3.2) Taking the norm of both sides of the equality (3.2) we get µ( 0, ν)[ν], whereXυ(cid:9) 0 h1,··· , di = [1,··· , d] +Xυ(cid:9) 0 kh1,··· , dikB(H) = k[1,··· , d] +Xυ(cid:9) 0 µ( 0, ν)[ν]kB(H) µ( 0, ν) = d! − 1. µ( 0, ν)k[ν]kB(H) ≤ k[1,··· , d]kB(H) +Xυ(cid:9) 0 ≤ kX xjkd ≤ d!kX xjkd = d!ndk = d!ndkP1(x1, ..., xn)kB(H). B(H) + (d! − 1)kX xjkd nX xjkd B(H) B(H) 1 B(H) Thus, kPd(x1, ..., xn)kB(H) ≤ d!nd(n − d)! n! Denote C(n, d) := d!nd(n−d)! n! , and for fixed d define f (n) := kP1(x1, ..., xn)kB(H). log n d−1 Pi=0 d!nd n−i . Then C(n, d) = = d!nd(n − d)! n n! n n · n − 1 · = d! · = d! · exp(f (n)). n(n − 1)(n − 2)··· (n − d + 1) n n − 2 ··· n − d + 1 n 10 WAFAA ALBAR, MARIUS JUNGE, AND MINGYU ZHAO Since f (n) is a decreasing function in n, C(n, d) is also a decreasing function with respect to the variable n. From the definition of d, we know n ≥ d, so max C(n, d) = n≥d C(d, d) = dd. 3.2. AGM inequality for the order. Recall that the average product is defined by: Pd(x1, x2, ..., xn) = (n − d)! n! Xhσi= 0 xi1 ...xid . Lemma 3.3. Let {xi} be a finite family of positive operators in B(H) which satisfy the condition n xi = n. If ai := xi − 1 then Pi=1 (3.3) Pd(x1, x2, ..., xn) = 1 + d Xk=1(cid:18)d k(cid:19)Pk(a1, a2, ..., an). Proof. This lemma can be proved by two methods. The first method is by induction which is left to the reader. For the convenience of the reader we give the second proof, using the binomial identity. Then we have Pd(x1, ..., xn) = = xi1 ...xid (n − d)! n! Xhσi= 0 (n − d)! n! Xhσi= 0 (ai1 + 1)(ai2 + 1)...(aid + 1) = 1 + λkPk(a1, ..., an). d Xk=1 Let x1 = x2 = .... = xn = t, where t = a + 1. Then Pd(x1, ..., xn) = td = (1 + a)d = 1 + d Xk=1(cid:18)d k=1(cid:0)d k(cid:19)ak, k(cid:1)Pk(a1, ..., an). In Theorem 1.2 for d=3, we deduce that each term in P3(x1, ..., xn) has an upper i . For d > 3, we need the following lemma. k(cid:1), so Pd(x1, ..., xn) = 1 +Pd which implies that λk =(cid:0)d bound of some scalar multiple of P a2 Lemma 3.4. If {xi},{ai} are defined as above, then 2 ≤ kXi 1 1 2 . x2 i k In particular, kaikk ≤ nkk 1 Proof. Since we have a2 i , k 2 . max i kaik ≤ kX a2 ik n2 Pi x2 i k j ≤P a2 kaik = ka2 ik i =X a2 X x2 1 1 2 . 2 ≤ kX a2 ik i + n ≥X a2 i Moreover, for each ai, we have xi = ai + 1. Thus This finishes the proof. NONCOMMUTATIVE VERSIONS OF THE AGM INEQUALITY 11 Note that for a partition with d = 3, the proof of the AGM inequality in the order sense was easily done in the introduction. However, the proof is much more complicated for d ≥ 4. The complication comes from crossing partitions, so we need the following useful known lemma [16]. Lemma 3.5. Assume a, b ∈ B(H) and t ≥ 0. Then (1) −(a∗a + b∗b) ≤ a∗b + b∗a ≤ a∗a + b∗b (2) ab + b∗a∗ ≤ t2aa∗ + t−2b∗b To prove (1), we start by observing (a + b)∗(a + b), (a− b)∗(a− b) ≥ 0. This directly gives −(a∗a + b∗b) ≤ a∗b + b∗a and a∗b + b∗a ≤ a∗a + b∗b. It is clear that (2) is a special case of (1), using the assumptions that a = ta∗ and b = t−1b for the upper bound of (1). The two previous lemmas will help in establishing our result for general case of the Zi is defined as at the beginning of Section 3.1. We now provide upper and lower bounds for Pd(ai1 , ..., ain ). AGM inequality for the order. For convenience, we will write Ai := Pi Zi where Lemma 3.6. If {ai} and {xi} are defined as above, then for S = kP x2 ik1/2 d! Sd−2X a2 i . d! Sd−2 X a2 i ≤ Pd(a1, a2,··· , an) ≤ (n − d)! (n − d)! Proof. From Proposition 2.5, we know − n! n! n! (n − d)! Pd(a1, a2,··· , an) = h 0id = [ 0]d + X0(cid:8)ν≤ 1 µ( 0, ν)[ν]d. We will prove first the case when µ( 0, ν) ≥ 0. We will obtain an upper bound for the sum [ν]d by introducing [¯ν]d as the following: ai1 ai2 ··· aid , [ν]d = Xh [¯ν]d := Xh Here the ¯ν can be viewed as the transposition of the partition ν. By Theorem 2.2, i=1[(−1)i−1(i − 1)!]ri(π). So µ( 0, ν) = µ( 0, ¯ν). Thus, we can aid aid−1 ··· ai1 . i1,i2,··· ,idi≥ν i1,i2,··· ,idi≥ν sum these two items together. we have µ( 0, π) = Qd Claim: For every partition ν and S = kP x2 (3.4) ik1/2 we have − 2 Sd−2X a2 i ≤ [ν]d + [¯νd] ≤ 2 Sd−2X a2 i . The idea here is to use our modification of Pisier's trick for these two partitions. Recall that Zi1 = e1i1⊗ai1 is for the first component in the partition, Zij = ejj⊗aij is for the elements in the middle of the partition, and Zid = eid1 ⊗ aid is for the 12 WAFAA ALBAR, MARIUS JUNGE, AND MINGYU ZHAO last element in the partition. Then we have i1,i2,··· ,idi≥ν [ν]d + [¯ν]d = Xh Zi1 ...Xid By applying Lemma 3.5 with S = kP x2 ai1 ai2 ··· aid + aid aid−1 ··· ai1 Zid +Xid = A1...Ad + A∗d...A∗1. i k1/2, [ν]d + [¯ν]d = A1 ··· Ad + A∗d ··· A∗1 = Xi1 ...Xi1 (3.5) Z∗id Z∗i1 ≤ t2A1A∗1 + t−2A∗d ··· A∗2A2 ··· Ad kA∗j AjkA∗dAd ≤ t2A1A∗1 + t−2 d−1 d−1 ≤ t2A1A∗1 + t−2 Yj=2 Yj=2 Yj=2 ≤ t2A1A∗1 + t−2 ≤ t2A1A∗1 + t−2kX a2 ≤ kX a2 d−1 kAjk2A∗dAd kX a2 ikd−2A∗dAd jkA∗dAd (3.6) (3.7) ikd/2−1(A1A∗1 + A∗dAd) ≤ 2X a2 i Sd−2. Indeed, if our partition contains the singleton then [ν]d + [¯ν]d is already zero. Hence we may assume there are no singletons in our partition as it also can be noticed in inequality (3.7). Indeed, if the index is a singleton in partition ν, then it is controlled by the summation norm kP aik which is zero by our construction. On the other hand, if the index is in a non-singleton block, then by Theorem 2.8, it is controlled by the square norm kP a2 ik. Therefore, in both cases, kAik is controlled by the square norm of ai. To get inequality (3.6), we may apply the norm equality as in equality (2.7) from section 2. For the inequality (3.7), we use Lemma (3.5) by choosing t2 = Sd/2−1. Then we have µ( 0, ν)[ν]d = X0(cid:8)ν≤ 1 µ( 0, ν)[ν]d n! 1 = µ( 0, ν)[ν]d Pd(a1, a2,··· , an) = [ 0]d + X0(cid:8)ν≤ 1 (n − d)! = Xµ( 0,ν)≥0 µ( 0, ν)[ν]d + Xµ( 0,ν)≤0 2(cid:16) Xµ( 0,ν)≥0 µ( 0, ν)[ν]d + Xµ( 0,¯ν)≥0 2(cid:16) Xµ( 0,ν)≤0 µ( 0, ν)[ν]d + Xµ( 0,¯ν)≤0 Pd(a1, a2,··· , an) ≤ Xµ( 0,ν)≥0 µ( 0, ¯ν)[¯ν]d(cid:17) µ( 0, ¯ν)[¯ν]d(cid:17) µ( 0, ν)Sd−2X a2 =Xµ( 0, ν) Sd−2(X a2 + 1 n! (n − d)! µ( 0, ν)Sd−2X a2 i i − Xµ( 0,ν)≤0 i ) = d! Sd−2X a2 i . NONCOMMUTATIVE VERSIONS OF THE AGM INEQUALITY 13 For the lower bound, the proof is similar to the one above replacing A1 by −A1. Theorem 3.7. (AGM inequality for the order) Fix n and d. Let x1, ..., xn be self-adjoint operators such that Pi xi = n and ai = xi − 1 as above. Assume the following conditions hold: 3d . Using this upper bound for the average of noncommutative operators ai with the identity (3.3) where S = = 1. i) P1(x1, ..., xn) = Pi xi n = 1, i ) ii) k(P x2 1 2k ≤ n 3d . Then the AGM inequality holds in the order sense: 3d , we have Pd(x1, x2,··· , xn) ≤(cid:16)Pi xi n (cid:17)d Proof. According to Lemma 3.4, we have kP a2 ik1/2 ≤ n kP x2 i k1/2 ≤ ∆n and let ∆ := 1 Xk=1(cid:18)d k(cid:19)Pk(a1, a2, ..., an) 2(cid:19) (n − 2)! = 1 −(cid:18)d ≤ 1 −(cid:18)d 2(cid:19) (n − 2)! Pd(x1, x2, ..., xn) = 1 + Xk=3(cid:18)d Xk=3(cid:18)d (X a2 (X a2 i ) + i ) + n! n! d d d Now we need the following condition: (3.8) (cid:18)d 2(cid:19) (n − 2)! n! ? ≥ Simplifying the right hand side gives d Xk=3(cid:18)d k(cid:19) (n − k)! n! k!∆k−2nk−2 = k(cid:19)Pk(a1, a2, ..., an) k(cid:19) (n − k)! n! k!∆k−2nk−2(X a2 i ). k!∆k−2nk−2. d! (d − k)!k! (n − k)!k! n! ∆k−2nk−2 (n − k)! n! ∆k−2nk−2 k(cid:19) (n − k)! n! d d Xk=3(cid:18)d Xk=3 Xk=3 n(n − 1) d! d d (d − k)! 1 Xk=3 = = d! nk−2 (d − k)! (n − 2)··· (n − k + 1) ∆k−2. (3.9) Fix k, and denote f (n) := g(n) := log f (n) = Pk−1 (3.10) nk−2 (n−2)···(n−k+1) . Then, by taking the logarithm, we have n−i . Observe that g(n) is a decreasing function and i=2 log n thus f (n) is a decreasing function as well. Therefore, we get the inequality: nk−2 (n − 2)··· (n − k + 1) ≤ dk−2 (d − 2)··· (d − k + 1) . 14 WAFAA ALBAR, MARIUS JUNGE, AND MINGYU ZHAO We continue the calculation in (3.9) with the help of inequality (3.10), we have d Xk=3(cid:18)d k(cid:19) (n − k)! n! k!∆k−2nk−2 = ≤ ≤ (n − 2)··· (n − k + 1) (d − 2)··· (d − k + 1) ∆k−2 ∆k−2 nk−2 dk−2 d d! d d! (d − k)! (d − k)! Xk=3 Xk=3 Xk=3 d∆(1 − (d∆)d−2) d d(d − 1)dk−2∆k−2 1 n(n − 1) 1 n(n − 1) 1 n(n − 1) d(d − 1) n(n − 1) = 1 − d∆ 3d we deduce indeed d(d−1) n(n−1) ≤ 1−d∆ ≤ (cid:0)d d∆ d(d − 1) n(n − 1) 2(cid:1) (n−2)! n! d∆ . 1 − d∆ and this With our choice of ∆ = 1 completes the proof. 4. AGM inequality for random matrices In this section, we prove a version of the NC-AGM inequality for random matrices. We start with a deviation inequality. Let us use the norm Xp = (EkXkp B(H))1/p defined for a random variable X : Ω → B(H). Proposition 4.1. Let {ai} be a family of self-adjoint random operators. Let ε > 0, p ≥ 2, pd = p d and xi = ai + 1. Define (i) εp :=(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 1 (ii) δp := 1 nP ai(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)p , nP ai − E 1 i )1/2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)p , n(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(P a2 (iii) γp := max(εp, δp). Eai = 0 and γp ≤ 1 Assume Pi 3d and ε = 3dγp Then, Pd(x1, ..., xn) − EPd(x1, ..., xn)pd ≤ ε. Proof. From the assumption above, we get that Fix a partition ν. According to Theorem (2.7) and by using Holder's inequality we have that = εp. 1 (cid:12)(cid:12)(cid:12)(cid:12)p (cid:12)(cid:12)(cid:12)(cid:12) nX ai)(cid:12)(cid:12)(cid:12)(cid:12) ( (cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) ∞ ≤ E k(X a2 = E k(X a2 Ek[ν]kpd i )1/2k(d−νs)pd ∞ i )1/2k (d−νs)pd d ∞ d ∞ ! k(X ai)kνspd · k(X ai)k νspd d ∞ d ! NONCOMMUTATIVE VERSIONS OF THE AGM INEQUALITY 15 pd d ∞ ! pd νs ∞! pd νs p p i )1/2kpdd ≤ E(k(X a2 = E(k(X a2 = (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)X a2 ∞ ! pd(d−νs ) ∞! pd (d−νs ) pdd Ek(X ai)kpdd Ek(X ai)kp i )1/2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)p!pd(d−νs) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)X ai(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)p!pdνs i )1/2kp = (δp · n)pd(d−νs).(εp · n)pdνs = δpd(d−νs) np = δpd(d−νs) εpdνs p npdd. εpdνs p p p Since γp = max(δp, εp), (4.1) [ν]pd = (Ek[ν]kpd ∞) 1 pd ≤ γd pdd · nd By using our definition of γp and the upper bound for inequality (4.1) we obtain ∞)1/pd (EkPk(a1, ..., an) − EPk(a1, ..., an)kpd ≤ ≤ ≤ 2 n! Xµ(0, ν)(cid:16)E(k[ν] − E[ν]k∞)pd(cid:17)1/pd (n − k)! (n − k)! n! Xµ(0, ν) · 2(Ek[ν]kpd (n − k)! ∞)1/pd pdknk. k!γk n! (4.2) From the above we will have Pd(x1, ..., xn) − EPd(x1, ..., xn)pd d d d = (Ek (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)pd (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) k(cid:19)(Pk(a1, ..., an) − EPk(a1, ..., an))(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Xk=1(cid:18)d Xk=1(cid:18)d k(cid:19)(Pk(a1, ..., an) − EPk(a1, ..., an))kpd Xk=1(cid:18)d k(cid:19)(Ek(Pk(a1, ..., an) − EPk(a1, ..., an))kpd Xk=1(cid:18)d k(cid:19) (n − k)! d!(n − k)!nk Xk=1 (d − k)!n! pdk · nk = 2 k!(d − k)! Xk=1 ≤ 2 ≤ 2 k!γk γk p . ≤ n! d! d d d (4.3) ∞)1/pd ∞)1/pd (n − k)! n! k!γk pdk · nk Recall the definition γpdk = max(δpdk, εpdk). Each δpdk, εpdk is increasing since Lpdk is defined as probability space which is norm increasing in probability measure. Thus γpdk ≤ γpdd = γp,∀ 1 ≤ k ≤ d, which justifies the last inequality (4.3). Let f (n) = d!(n−k)!nk . This function is a decreasing function in n, so f (d) = max f (n) = (d−k)!n! 16 WAFAA ALBAR, MARIUS JUNGE, AND MINGYU ZHAO dk. Then we have Pd(x1, ..., xn) − EPd(x1, ..., xn)pd d ≤ 2 Xk=1 (d · γp)k = 2 · d · γp (1 − (d · γp)d) 1 − d · γp ≤ 2 · 1−ε/2 . d · γp 1 − d · γp ≤ ε. The last inequality follows from d · γp ≤ ε We now present conditions for positive random operators {xi} where ai = xi − 1. Note that for A :=Pn ai n , we have )kp. Therefore, whenever we control the xi's, we control the ai's. Lemma 4.2. Let {xi} be a family of self-adjoint random operators. Then EkA − EAkp = Ek(P xj ) − E(P xj i=1 n n (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(X(xi − 1)2)1/2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)p ≤ 6(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(X x2 i )1/2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)p 1 n n 1 n 1 n 1 1 yi)k = Xi=1 1 √nk Indeed n e1,i ⊗ yik. ej,1 ⊗ φ(yi)jk = k By triangle inequality, we can get yjk ≤ kXi e1,i ⊗ xik. Also, i )1/2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)p = P xi ⊗ ei,1p is given by the column norm. Proof. Observe that(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(P x2 Define operators φ : Cn(B(H)) → Cn(B(H)) and Φ : Cn → Cn such that Φ(αi) = nPi ( 1 αi)j where φ = Φ ⊗ Id. Then it is easy to check that kΦkcb = kφkcb ≤ 1. Xj=1 Xj=1 k Denote zi := xi − Exi, so kPi (Id + φ)(zi) = xi − Exi + Xj=1 ej,1 ⊗ ( e1,i ⊗ (Id + Φ)(zi)k ≤ 2kPi nX xi − (xi − 1) = (Id + φ)(zi) + Exi − nX xi, nX Exi = xi − 1 − Exi + nX xi. (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)X(xi − 1) ⊗ ei,1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)X(Id + φ)(zi) ⊗ ei,1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) +(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)X Exi ⊗ ei,1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) +(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) nX xi) ⊗ ej,1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Xj (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) nX xi) ⊗ ej,1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ 2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)X zi ⊗ ei,1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) +(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)X Exi ⊗ ei,1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) +(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Xj (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ 2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)X(xi − Exi) ⊗ ei,1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) +(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)X Exi ⊗ ei,1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) +(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) nX xi) ⊗ ej,1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Xj nX xi(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) ≤ 2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)X xi ⊗ ei,1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) + 3(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)X Exi ⊗ ei,1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) +(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) ≤ 2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(X x2 2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12). 2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ 6(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(X x2 2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) + 3(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)X xi ⊗ ei,1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) +(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(X x2 1 i ) 1 i ) 1 i ) 1 ( 1 ( 1 ( 1 NONCOMMUTATIVE VERSIONS OF THE AGM INEQUALITY 17 The second-to-last inequality P Exi ⊗ ei,1 ≤ P xi ⊗ ei,1 follows from the fact that conditional expectation from E : L∞(Ω, B(H)) → B(H) is a complete contraction. The inequality is true by the Cauchy-Schwarz inequality. (cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) 1 (cid:12)(cid:12)(cid:12)(cid:12) ≤(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(X x2 (cid:12)(cid:12)(cid:12)(cid:12) nX xi(cid:12)(cid:12)(cid:12)(cid:12) i ) 1 2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Thanks to Theorem 4.1 and Lemma 4.2 we obtain the following deviation result. Theorem 4.3. Let p ≥ 2, pd := p d , and {xi} be a random family of positive operators such that Exi = 1. Define (i) εp :=(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 1 (ii) δp := 1 , nP xi(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)p nP xi − E 1 i )1/2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)p , n(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(P x2 (iii) γp := max(εp, 4δp). If 3d · γp ≤ 1 then Pd(x1, ..., xn) − EPd(x1, ..., xn)pd ≤ 3d · γp. Corollary 4.4. If in addition {xi} are matrix-valued i.i.d. Then Pd(x1, ..., xn)pd ≤ 1 + 3d · γp. Proof. Since xi's are matrix-valued i.i.d, then E(Pd(x1, ..., xn)) = Pd(Ex1, ..., Exn). Moreover, for ε := 3d · γp by (ii) in the above Theorem 4.3, we have k(X E(xi)2)1/2k = kX E(xi) ⊗ ei,1kCn⊗B(H) ≤ kX xi ⊗ ei,1kCn⊗B(H) ≤ δp · n. Then we can use Theorem 3.7 for E(xi)'s and the classical AGM inequality (here δp ≤ 1 4 γp ≤ 1 4d < 1 3d ). E(Pd(x1, ..., xn)) ≤ P1(Ex1, ..., Exn) =X Exi n = 1. Using the upper bound above and Theorem 4.3, we have the required inequality. Pd(x1, ..., xn)pd ≤ P1(Ex1, ..., Exn)pd + ǫ ≤ 1 + ǫ. 4.1. Application to Log concave measures. In this section we want to study random AGM inequalities for log-concave measures. Definition 4.5. A Borel measure µ on n-dimensional Euclidean space Rn is called logarithmically concave (or log-concave) if for any compact subsets A and B of Rn and 0 ≤ λ ≤ 1 we have µ(cid:0)λA + (1 − λ)B(cid:1) ≥ µ(cid:0)A(cid:1)λ µ(cid:0)B(cid:1)(1−λ) . 18 WAFAA ALBAR, MARIUS JUNGE, AND MINGYU ZHAO Let us recall the isotropic measure µ in Rn. Definition 4.6. The isotropic measure µ is the measure which satisfies ZRn hθ, xi2dµ(x) = Lµkθk2, for all θ ∈ Rn where Lµ is denoted as isotropic constant. Also let us recall Rosenthal's inequality, which will be used frequently in this section. n n n k Theorem 4.7 (Rosenthal inequality [10]). Let Ai be a fully independent sub-algebra over N where N ⊂ M and M is a von Neumann algebra, and 1 ≤ p < ∞. Let xi ∈ Lp(Ai) with EN (xi) = 0. Then Xi=1 Xi=1 EN (x∗i xi + xix∗i )1/2kp, p( xikp ≤ C max{√pk p)1/p}. Xi=1 kxikp Theorem 4.8. Let n, d ∈ N, p ≥ 2. Let (Rd, µ) be log-concave Borel measure µ in isotropic position on Rd with constant L. Define random variable y : Rd → Rd by y(ω) = ω√L where ω ∈ Rd. Let yi be independent copies of y. Then xi(ω) := yi(ω)ihyi(ω) is a d × d random matrix satisfying (i) ∀i, Exi = 1, (ii) P(xi − Exi)p ≤ γp · n, (iii) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)P x2 i(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) p ≤ γp · n, where γp = p1/2q d 2C√ln dδ1/2  2C(ln n)3δ p ≥ ln n or d ≤ n d ≥ n d , ε ≥ 0, and pk := p (iv) Moreover, assume γp ≤ (1− 2 k . Then the following n + p5/2 d 2+ε ) 1 ln n5 ln n5 1/2 n hold. • Pk(x1, ..., xn) − EPk(x1, ..., xn)pk ≤ ε. • The AGM inequality holds Pk(x1, ..., xn)pk ≤ (1 + 2ε). Proof. We apply Rosenthal's inequality for q ≥ p to xi − 1 instead of xi. Let us introduce the norm in the space Lq(Sq) where Sq is the Schatten class, xq := (Ekxikq Sq ) So, we have 1 q = (Z kx(ω)kq qdµ) 1 q . (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)X(xi − Exi)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)q ≤ c max{√qX E((xi − 1)∗(xi − 1) + (xi − 1)(xi − 1)∗) ≤ c√qX E((xi − 1)∗(xi − 1) + (xi − 1)(xi − 1)∗) ≤ 2c√qX E(xi − 1)2 ≤ 2c√qX Ex2 + cq(Xxi − Exiq + 2cqn1/qx1q i q) 1 2 q 2 1 2 q 2 1 2 q 2 1 q 1 2 q 2 , q(Xxi − Exiq + cq(Xxi − Exiq q) 1 q q) 1 q } NONCOMMUTATIVE VERSIONS OF THE AGM INEQUALITY 19 By Rosenthal's inequality, we need to separately estimate the two terms of the right side. We denote (4.4) I = (X Ex2 i )1/2)q and II = x1q. We claim that (ii) holds for γq and Ex2 (see [14] where k.k is seminorm), we have i ≤ dExi ≤ cd · 1. Using Borel inequality Recall that Ekyk2 = 1 d q ≤ CqEkykX ≤ Cq(Ekyk2 X ) (Ekykq X) Pi=1 Ehei, ω√Li2 = d. So, we have for xi := x1 = yihy 1θi = Ehθ, yihy, yihθ, yi = Ekyk2hθ, yi2 hθ, Ex2 1 2 . 1 ≤ (Ekyk4) ≤ C4 2 (Ehθ, yi4) i ≤ dExi ≤ cd · 1. This implies k(P Ex2 4 Ekyk2E(hθ, yi2) = C4 i.e. Ex2 our claim for (I). For (II), note that the q-norm is defined to be xq = (Etrxq) Let's first take q = m be an integer. We have i )1/2)kq ≤ C · d1/2+1/q which proves 1 q . 4 · d kθk2. 1 2 xm i =yiihyim = yiihyi, yii···hyi, yiihyi =yiikyik2(m−1)hyi. Then, by using the Borel inequality (see [14]), we have Etr(xm i ) =Etr(yiikyik2(m−1)hyi) = E(kyik2m 2 ) ≤ (C · 2m)2m((Ekyk2 ≤ (C · 2m)2mdm. 2)1/2)2m So we get the inequality xm ≤ (C·2m)2d for arbitrary integer m. Then for any real number q, we can find an integer m, such that m ≤ q ≤ m + 1, and by interpolation between m and m + 1, we get (4.5) xq ≤ (C · 2q)2 d. Thanks to (4.5), we can now prove condition (iii). (4.6) Combining (I) and (II) we obtain )1/2 q 2 1/2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(X x2 i(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) i )1/2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)q ≤(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)X x2 2 ≤ (X(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)x2 i(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) q ≤ (Xxiq)1/2 = √nx1q ≤ √nx1q ≤ √nd (C · 2q)2. (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)X(xi − 1)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)q ≤c(qnd)1/2d1/q + cqCn1/qq2d q + C′n1/qq3d. = C(qn)1/2d 2 + 1 1 20 WAFAA ALBAR, MARIUS JUNGE, AND MINGYU ZHAO And then divide each term by n, we have P(xi − Exi)q n q n ≤ C(q, d, n) := ( 2 + 1 = ( q + )1/2 d 1 q n )1/2d 1 2 + 1 q + n 1 q −1q3d q2n 1 q qd n d n = d1/q(cid:16)q1/2( = d1/q(cid:16)(q)1/2( )1/2 + q3 d n d n )1/2 + q3( If we denote d n = δ, then n1/q d1/q(cid:17) )1−1/q(cid:17). d n P(xi − Exi)q n 1 1 1 2 δ q (q 2 + q3δ1− 1 q ) ≤ d = d1/qq1/2δ1/2(1 + q5/2δ1/2−1/q). Now our goal is to find γq = inf q≥q0 over q where q0 ≥ 2. Define f (q, δ) := q5/2δ1/2−1/q and consider d1/qq1/2δ1/2(1 + q5/2δ1/2−1/q) by optimization g := ln f (q, δ) = ln q + ( 5 2 1 2 − 1 q ) ln δ, with derivative g′ = 5 δ . Since 2 f (q, δ) is a convex function then it has no more than one minimum point which is q(δ). Then we have to consider the following cases for the choices of q, q2 ln δ. The critical point for f (q, δ) is q(δ) = 2 5 ln 1 q + 1 1 (1) q0 ≤ ln d ≤ q1 where q1 = ( 1 (2) ln d ≤ q0 ≤ ln n (3) ln d ≤ ln n ≤ q0 δ )1/5. This can be done by using optimization over q for the term d1/qq1/2δ1/2. For the first case, we choose q = ln d and C(q, δ) = 2C√ln dδ1/2 where f (q, δ) ≤ 1. We also calculate q1 which represents the upper bound for our choice of q from q5/2δ1/2 = 1. For the second case, if ( n d , then we simply choose q = ln n. This leads to d n ln n5 . We can summarize the cases in the following ⋍ 1 ln n4 ≤ 1 d )1/5 ≥ ln n q1/2q d 2C√ln dδ1/2 2C(ln n)3δ γq =  n + q5/2 d n q ≥ ln n or d ≤ n d ≥ n ln n5 ln n5 We apply the estimate for q ≥ p and appeal to Theorem 4.3 and Corollary 4.4 to deduce the AGM inequality. 4.2. Wishart random variable matrices. Let us recall the definition of Wishart random matrices. Let [gi r,s] is a family of d×m Gaussian random matrices such that i ∈ [1, n], r ∈ [1, d] and s ∈ [1, m]. Define Gi = 1√m [gi rs] and xi = GiG∗i . We call the matrices xi d × d Wishart random matrices. Then we have Exi = EGiG∗i = 1 which implies that Pn i=1 Exi = n. In this section we assume that m ≥ n. Let us list some useful lemmas which will be used in the main theorem. Each of these lemmas proves one of the conditions of Theorem 4.3 separately. NONCOMMUTATIVE VERSIONS OF THE AGM INEQUALITY 21 q√n . Those d× d Wishart random matrices {xi} from above satisfy (4.7) Lemma 4.9. Let εq,m,n,d =(cid:16)√d+√m√m (cid:17)2 n(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(X x2 Proof. Denote A = 1√mPr,s 1 i )1/2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)q ≤ εq,m,n,d. gr,ser,s. Then for all h ∈ H, and x = AA∗ E(h, x2h) =E(h,AA∗2h) = E(h, AA∗AA∗h) = E(AA∗h, AA∗h) =EkAA∗hk2 ≤ E(kAk2 op · kA∗hk2) ≤ EkAk2 op · EkA∗hk2. d m Note that EkA∗hk2 = E(h, A∗Ah) = khk2. Using Chevet's inequality [4], Xr=1 (4.8) EkAk = Ek where X = lm Kahane's inequality we have that gr,sesk) + E(k 2 and Y = ld Xr=1 m d Xs=1 √2(cid:16)√d + √m Xs=1 gr,ser⊗eskX ⊗Y ≤ E(k 2. We deduce that if A = 1√mP gi √m (cid:17) =: C(d, m) op)1/2 ≤ op)1/2 ≤ C(d, m). xi2 = (Ekxik2 (4.9) Therefore (EkAk2 gr,serk), rser ⊗ es then by using For q ≥ 2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(X x2 i )1/2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)q (4.10) 1/2 =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)X x2 i(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) q/2 ≤ (X(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)x2 i(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)q/2)1/2 q)1/2 ≤ √nxiq = √n[(EkAk2q)1/2q]2 ≤ (Xxi2 ≤ √n(√q)2[(EkAk2)1/2]2 = 2q√n( √d + √m √m )2 The last inequality comes from Kahane's inequality (see proposition 3.3.1 and proposition 3.4.1 in [11]) and inequality (4.9). Thus, taking εq,m,n,d =(cid:0) √d+√m√m (cid:1)2 2q√n , we have The following lemma is used to prove the first condition in Theorem 4.3. Lemma 4.10. For d × d Wishart random variables xi, the following is satisfied (4.11) 1 ′ 1 i )1/2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)q ≤ εq,m,n,d. n(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(X x2 n(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)X(xi − Exi)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)q ≤ γ q, q ≤ ln d ≤ n q ≥ ln d . where γ ′ q =(C′ ln d q ln d q q max{p q C′d n 1 n , q n} 22 WAFAA ALBAR, MARIUS JUNGE, AND MINGYU ZHAO Proof. By Rosenthal's inequality, we have 1 q Ex2 (xi − Exi)kq q) i )1/2q i )1/2q (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)X(xi − Exi)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)q ≤ (EkXi ≤c √q(E(Xi ≤c √q(E(Xi ≤ √qd ≤ √qnd ≤ d q (E(X x2 q q[1 +r d q [1 +r d Ex2 i ) m m 1 1 1 1 q)1/q q)1/q + q(Xxi − Exiq (Exiq q) √d + √m q)1/q + qn q · max i 1 1 1 q ]2 2 q ∞)1/q + qn 1 q d q q[ √m ]2 + qn 1 q d 1 q q[1 +r d m ]2 ]2(√qnq + q2n 1 q ). The second-to-last inequality uses Kahane's inequality [11] and inequality (4.10). Dividing the inequality by n, we obtain (4.12) P(xi − Exi)q n 1 q [1 +r d m ≤ d ]2q(r q n √q n + √qn 1 q ) . Let 2 ≤ q0 ≤ q. We have two cases to estimate the upper bound: (1) q0 ≤ ln d ≤ n (2) ln d ≤ q0 ≤ q 1 We follow the optimization for q from the proof of Theorem 4.8. Define f (q) = √qn q2 = 0 at q = 2 ln n. Then q , and consider g(q) = ln f (q) = 1 q ln n, then g′(q) = 1 2q − ln n 2 ln q + 1 C q n Moreover, by (4.12), when d ≤ m, we obtain d 1 q [1 +r d m ]2q(r q n n √q n f (q) ≤(Cp q +r q √qn q − 1 n 1 2 ≤ q<n q ≥ n q [1 +r d q q max{r q m 1 1 2 ) ≤ 2Cd ≤ 8Cd , q n} ]2q max{r q q n}. n , n Denote F (d, n) = 8Cd F (d, n) = C′ ln d q ln d we get F (d, n) = C′d 1 1 n , q n , q n}. We choose q = ln d and we get that q q max{p q n if we have q0 ≤ ln d ≤ n. Otherwise we choose q ≥ q0, and q q max{p q γq =(C′ ln d q ln d q q max{p q n} q ≥ ln d . n}. Moreover, q ≤ ln d ≤ n n , q C′d n 1 We apply the estimate for q ≥ p and appeal to Theorem 4.3 and Corollary 4.4. NONCOMMUTATIVE VERSIONS OF THE AGM INEQUALITY 23 Now, we can prove the AGM inequality for random matrices which holds up to (1 + ε). Theorem 4.11. Let {xi} be a family of self-adjoint family of d×d Wishart random matrices. For 2 ≤ p ≤ ln d ≤ n, we have n (ii) 1 n E(xi) = 1; (cid:12)(cid:12)(cid:12)(cid:12)p ≤ γpn; (cid:12)(cid:12)(cid:12)(cid:12)Pi (cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) (xi − E(xi))(cid:12)(cid:12)(cid:12)(cid:12) (i) (cid:12)(cid:12)(cid:12)(cid:12) Pi=1 (cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) (iii) (cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)p ≤ γpn, where γp = C′ ln d q ln d (cid:12)(cid:12)(cid:12)(cid:12) 2(cid:12)(cid:12)(cid:12)(cid:12) (Pi • Pk(x1, ..., xn) − EPk(x1, ..., xn)pk ≤ ε. • The random AGM inequality holds, (iv) Moreover, for ε ≥ 0 if γp ≤ ε 3k , pk := p 1 x2 i ) Pk(x1, ..., xn)pk ≤ (1 + 2ε). n , p0 ≤ ln d ≤ n; k then the following hold. Proof. Condition (ii) comes from definition of the Wishart random matrices. For condition (i) we directly use Lemma 4.10 for the case when pk ≤ ln d ≤ n. For condition (iii), we use Lemma 4.9. This implies that all the conditions of Theorem 4.4 are satisfied, since pk ≤ ln d ≤ n. Thus, we get the random AGM inequality. 5. application on Pisier's construction and freely independent Let (M, τ ) be a von Neumann algebra where τ is a faithful normal and normalized trace. An example of a finite von Neumann algebra is given by the group von Neu- mann algebra L(G) associated to the left regular representation λ(G) of a discrete group G. It is defined as the strong operator closure of the linear span of λ(G). Recall that Lp(M, τ ) where 1 ≤ p < ∞ is defined as the completion of M with respect to the norm kxkp = (τ (xp))1/p (see Pisier [18] for more details). Note that L(G) = L∞(L(G)) and L(G) ⊂ Lp(L(G)). We want to prove a version of the AGM inequality with respect to the norm k.kp. For this version of the AGM inequality, we need the following key lemma. Lemma 5.1. Let M be a von Neumann algebra. Let ν be a partition. Then there exists a group G and bi(j) ∈ L(G) such that for xi(j) ∈ Lp(M ), the elements Xi(j) = bi(j) ⊗ xi(j) ∈ Lp(L(G) ⊗ M ) satisfy (5.1) Xi1 (1)Xi2 (2)...Xid (d). [ν] = EM Xi1,i2,i3,...,id Moreover, (5.2) Xi(j)kp ≤ kXi (C maxnk(P xi(j)∗xi(j))1/2kp,k(P xi(j)xi(j)∗)1/2kpo kP xi(j) kp j ∈ An.s ∈ σn.s {j} ∈ σs, where C is a universal constant. Note that bi(j) = 1 if {i} ∈ σs. Remark 5.2. The norm inequality (5.2) was proved by Pisier for even integers p ≥ 2 in [18]. The general case follows from [8]. 24 WAFAA ALBAR, MARIUS JUNGE, AND MINGYU ZHAO Now we can state the AGM inequality for Lp(M ) where p ≥ d. Theorem 5.3. Let M be a von Neumann algebra and xi ∈ Lp(M, τ )s.a satisfy the following condition for some δ ≥ 0, k(X x2 d ≤(cid:16)1 + (δC)(d! − 1)(cid:17) nd(n − d)! i )1/2kp ≤ δkX xikp. kPd(x1, ..., xn)k p Then we have xikd p. (5.3) 1 n n! k n X1 We will only give the sketch of the proof of this theorem since it is similar to the proof of Theorem 3.2 for pd = p d ≥ 1. Proof. By using Lemma 5.1, Holder's inequality and the contraction of conditional expectation we have that khσikpd ≤ kX xikd ≤ kX xikd p +Xυ(cid:9) 0 p +Xυ(cid:9) 0 i )1/2kvn.s p kX xikvsp µ( 0, ν)Cvn.sk(X x2 µ( 0, ν)(δC)vn.skX xikd nd(n − d)! p. k 1 nX xikd p. Thus for δC ≤ 1 (5.4) kPd(x1, ..., xn)kpd ≤ (1 + (δC)(d! − 1)) n! Remark 5.4. If δ ≤ 1, we get the AGM inequality with a constant C(d, n) = Cddd. As a matter of completeness, we want to include the limit case of the Wishart random matrices as an application for the AGM inequality. Let's first give the definition of freely independent von Neumann algebra (for more details see [23]). Definition 5.5. Let {Ai} be a family of unital von Neumann subalgebras of A. Then {Ai} is called a freely independent algebra (with respect to a unital linear functional φ ) if φ(x1...xn) = 0 whenever φ(xj ) = 0 for all xj ∈ Aij and i1 6= i2, i2 6= i3, ... We say that operators xi ∈ Ai are freely independent if their algebra {Ai} are freely independent. In the following theorem we prove the deviation inequality up to ε and apply this to the AGM inequality. Theorem 5.6. If {xi} are freely independent in von Neumann algebra M such that (1) EM (xi) = 1 (2) sup i kxik ≤ C and 2 + (4√n)C ≤ εn 3d , then (1) kPd(x1, ..., xn) − EPd(x1, ..., xn)k∞ ≤ ε (2) kPd(x1, ..., xn)k∞ ≤ 1 + ε. NONCOMMUTATIVE VERSIONS OF THE AGM INEQUALITY 25 Proof. Let ai = xi−1. By assumption we have EM (ai) = 0. By simple modification of Voiculescu's inequality [7] , we get that k(X a2 i )1/2k = kX ei1 ⊗ aik ≤ supkaik + 2k(X EM (a2 i ))1/2k ≤ 2(1 + C) + 2√nC ≤ 2 + (4√nC) ≤ εn 3d . (5.5) Indeed, kaik = kxi − 1k ≤ 1 + kxik ≤ 1 + C and EM (a2 i ) = EM (xi − EM (xi))2 = EM (x2 i ) − EM (xi)2 xix1/2 ) ≤ EM (x2 ≤ kxikEM (xi) = kxik ≤ C. i ) = EM (x1/2 i i Again, using Voiculescu's inequality we have, (5.6) kX aik ≤ supkaik + 2k(X EM (a2 Following the proof of Proposition 4.1, we get i ))1/2k ≤ 2(1 + C) + 2√nC ≤ εn 3d . (5.7) Applying the techniques of Proposition 4.1 to the case p = ∞, we have (5.8) k[ν]k ≤ k(X a2 kPk(a1, ..., an) − EPk(a1, ..., an)k ≤ 2 i )1/2kd−vskX aikvs ≤ ( (n − k)! εn 3d )k. )d. k!( n! εn 3d Then we have kPd(x1, ..., xn) − EPd(x1, ..., xn)k = k (5.9) ≤ 2 ≤ 2 k(cid:19)(Pk(a1, ..., an) − EPk(a1, ..., an))k d!(n − k)! (d − k)!n! {z )k ≤ ε. ε 3d ε 3d } nk )k ( f (n) is a decreasing function (d)k( d d Xk=1(cid:18)d Xk=1 Xk=1 d Then we have to apply Theorem 3.7 for yi = Exi instead of xi, where P yi n = 1. Note that by free independence, we have EPd(x1, ..., xn) = Pd(Ex1, ..., Exn) using the fact that {xn} in Pd(x1, ..., xn) has no repetition. kPd(x1, ..., xn)k ≤ kP1(Ex1, ..., Exn)k + ǫ ≤ 1 + ǫ. Remark 5.7. The norm version of the AGM inequality also holds for the family of freely independent {xi}. Indeed, we have that (5.10) kPd(x1, ..., xn)k ≤ (1 + ε)k 1 In this case we use again the Voiculescu inequality and deduce that nX xikd. k 1 nX xi − 1 nX Exik ≤ ε 3d . 26 WAFAA ALBAR, MARIUS JUNGE, AND MINGYU ZHAO This implies k 1 nP xik ≥ 1 − ε 3d . Hence, Note that (1+ε)1/d 3d ) ≈ 1+ε (1− ε kPd(x1, ..., xn)k1/d ≤ (1 + ε)1/d 3d ) k (1 − ε 1−ε . This is true since we have (1 − x)n ≥ (1 − nx) 1 nX xik. for x ∈ [0, 1] and n ≥ 1. Applying this inequality for x = ε ). d) = (1 − Thus, AGM inequality is true up to the constant 1+ε 1− ε )d ≥ (1 − (1 − ε 3d ε 3d ε 3 3 ≈ 1 + ε. 3d , we have Another interesting application for freely independent copies {xi} is given as follows: Corollary 5.8. Let {xi} be self-adjoint freely independent copies over an algebra B such that EB(x1) = 1B and kx1k ≤ C. Then the AGM inequality holds up to (1 + ε). Proof. Using the free independence for the {xi}'s, where d ≤ p ≤ ∞ we get (1) EB(xi) = 1B (2) kxikp = kx1kp ≤ C (3) k(P x2 i )1/2kp ≤ ckxikpn1/p + √nk(EM x2 i )1/2kp Indeed, for the property (3) we just apply a version of Voiculescu's inequality for free variables [8], n n n Note that 1)1/2kp xikp ≥ k EM (xi)kp − k kX xi ⊗ ei1kp ≤ c(Xkxikp)1/p + k(X EM (x∗i xi))1/2kp ≤ Cn1/p + √nk(EM x2 X1 X1 k ≥ nkEM (x1)kp −(cid:16) Cn1/p + √nkEM (x2 Now, if A ≤ n 2kEM (x1)k, then we have X1 x2 i )1/2kp ≤ 2 k( i )1/2kp(cid:17) } X1 k xikp = n−1/2 2 Cn1/p−1/2 + 2kEM (x2 1)1/2kp kEM (x1)kp X1 (cid:16)xi − EM (xi)(cid:17)kp {z 1)1/2kp Cn1/p + √nkEM (x2 nkEM (x1)kp k A n n n X1 xikp. ! } Then we get X1 k( n Cn {z X1 i )1/2kp ≤ δnk x2 n xikp, NONCOMMUTATIVE VERSIONS OF THE AGM INEQUALITY 27 where δn = Cn√n . Then for √n ≫ d! we have δn → 0. This implies that when n is large enough we get the AGM inequality as follows: n (5.11) kPd(x1, ..., xn)k p d ≤ (1 + ε)k xikd p. X1 References [1] George E Andrews. The theory of partitions, volume 2. Cambridge university press, 1998. [2] L´eon Bottou. Online learning and stochastic approximations. On-line learning in neural net- works, 17(9):25, 1998. [3] Xixuan Feng, Arun Kumar, Benjamin Recht, and Christopher R´e. Towards a unified ar- chitecture for in-rdbms analytics. In Proceedings of the 2012 ACM SIGMOD International Conference on Management of Data, pages 325 -- 336. ACM, 2012. [4] Yehoram Gordon. Some inequalities for Gaussian processes and applications. Israel Journal of Mathematics, 50(4):265 -- 289, 1985. [5] Godfrey Harold Hardy, John Edensor Littlewood, and George P´olya. Inequalities. Cambridge university press, 1952. [6] Marius Junge. Doob's inequality for non-commutative martingales. Journal fur die Reine und Angewandte Mathematik, pages 149 -- 190, 2002. [7] Marius Junge. Embedding of the operator space oh and the logarithmic little Grothendieck inequality. Inventiones mathematicae, 161(2):225 -- 286, 2005. [8] Marius Junge, Javier Parcet, Quanhua Xu, et al. Rosenthal type inequalities for free chaos. The Annals of Probability, 35(4):1374 -- 1437, 2007. [9] Marius Junge and Quanhua Xu. Noncommutative Burkholder/Rosenthal inequalities. Annals of probability, pages 948 -- 995, 2003. [10] Marius Junge, Qiang Zeng, et al. Noncommutative Bennett and Rosenthal inequalities. The Annals of Probability, 41(6):4287 -- 4316, 2013. [11] Stanislaw Kwapien and Wojbor A Woyczynski. Random series and stochastic integrals: single and multiple. Boston, 1992. [12] Fran¸coise Lust-Piquard. In´egalit´es de khintchine dans cp (1¡ p¡). CR Acad. Sci. Paris, 303:289 -- 292, 1986. [13] Fran¸coise Lust-Piquard and Gilles Pisier. Non commutative Khintchine and Paley inequali- ties. Arkiv for Matematik, 29(1):241 -- 260, 1991. [14] Vitali D Milman and Gideon Schechtman. Asymptotic theory of finite dimensional normed spaces. 1986. [15] Yu Nesterov. Efficiency of coordinate descent methods on huge-scale optimization problems. SIAM Journal on Optimization, 22(2):341 -- 362, 2012. [16] Vern Paulsen. Completely bounded maps and operator algebras, volume 78. Cambridge Uni- versity Press, 2002. [17] Gilles Pisier et al. An inequality for p-orthogonal sums in non-commutative l {p}. Illinois Journal of Mathematics, 44(4):901 -- 923, 2000. [18] Gilles Pisier and Quanhua Xu. Non-commutative lp-spaces. Handbook of the geometry of Banach spaces, 2:1459 -- 1517, 2003. [19] Narcisse Randrianantoanina. Non-commutative martingale transforms. Journal of Functional Analysis, 194(1):181 -- 212, 2002. [20] Benjamin Recht and Christopher Re. Parallel stochastic gradient algorithms for large-scale matrix completion. submitted for publication. Preprint available at http://pages. cs. wisc. edu/ brecht/publications. html, 2011. [21] Benjamin Recht and Christopher R´e. Beneath the valley of the noncommutative arithmetic- conjectures, case-studies, and consequences. arXiv preprint geometric mean inequality: arXiv:1202.4184, 2012. [22] Gian-Carlo Rota. On the foundations of combinatorial theory i. theory of mobius functions. Probability theory and related fields, 2(4):340 -- 368, 1964. [23] Dan V Voiculescu, Ken J Dykema, and Alexandru Nica. Free random variables. Number 1. American Mathematical Soc., 1992. 28 WAFAA ALBAR, MARIUS JUNGE, AND MINGYU ZHAO UNVERSITY OF ILLINOIS AT URBANA-CHAMPAIGN E-mail address: [email protected] UNVERSITY OF ILLINOIS AT URBANA-CHAMPAIGN E-mail address: [email protected] UNVERSITY OF ILLINOIS AT URBANA-CHAMPAIGN E-mail address: [email protected]
1506.07572
2
1506
2016-08-29T15:21:22
The equivariant Cuntz semigroup
[ "math.OA", "math.DS", "math.FA" ]
We introduce an equivariant version of the Cuntz semigroup, that takes an action of a compact group into account. The equivariant Cuntz semigroup is naturally a semimodule over the representation semiring of the given group. Moreover, this semimodule satisfies a number of additional structural properties. We show that the equivariant Cuntz semigroup, as a functor, is continuous and stable. Moreover, cocycle conjugate actions have isomorphic associated equivariant Cuntz semigroups. One of our main results is an analog of Julg's theorem: the equivariant Cuntz semigroup is canonically isomorphic to the Cuntz semigroup of the crossed product. We compute the induced semimodule structure on the crossed product, which in the abelian case is given by the dual action. As an application of our results, we show that freeness of a compact Lie group action on a compact Hausdorff space can be characterized in terms of a canonically defined map into the equivariant Cuntz semigroup, extending results of Atiyah and Segal for equivariant $K$-theory. Finally, we use the equivariant Cuntz semigroup to classify locally representable actions on direct limits of one-dimensional NCCW-complexes, generalizing work of Handelman and Rossmann.
math.OA
math
THE EQUIVARIANT CUNTZ SEMIGROUP EUSEBIO GARDELLA AND LUIS SANTIAGO Abstract. We introduce an equivariant version of the Cuntz semigroup, that takes an action of a compact group into account. The equivariant Cuntz semi- group is naturally a semimodule over the representation semiring of the given group. Moreover, this semimodule satisfies a number of additional structural properties. We show that the equivariant Cuntz semigroup, as a functor, is continuous and stable. Moreover, cocycle conjugate actions have isomorphic associated equivariant Cuntz semigroups. One of our main results is an analog of Julg's theorem: the equivariant Cuntz semigroup is canonically isomorphic to the Cuntz semigroup of the crossed product. We compute the induced semi- module structure on the crossed product, which in the abelian case is given by the dual action. As an application of our results, we show that freeness of a compact Lie group action on a compact Hausdorff space can be characterized in terms of a canonically defined map into the equivariant Cuntz semigroup, extending results of Atiyah and Segal for equivariant K-theory. Finally, we use the equivariant Cuntz semigroup to classify locally representable actions on direct limits of one-dimensional NCCW-complexes, generalizing work of Handelman and Rossmann. Contents Introduction 1. 2. The Equivariant Cuntz semigroup 3. The semiring Cu(G) and the category CuG 4. A Hilbert module picture of CuG(A, α) 5. Julg's theorem and the Cu(G)-semimodule structure on Cu(A ⋊α G) 6. Examples and computations 7. A characterization of freeness using the equivariant Cuntz semigroup 8. Classification of actions using CuG References 1 4 13 20 27 42 48 51 55 1. Introduction Equivariant K-theory for compact groups acting on topological spaces was in- troduced by Atiyah (the paper [Seg68], by Segal, contains a basic treatment of the theory). One of the first applications of this theory was a striking characterization of freeness of a compact Lie group action ([AS69]; see also Theorem 1.1.1 in [Phi87]). Equivariant K-theory was later defined and studied for actions of compact groups Date: 5 September 2018. 2010 Mathematics Subject Classification. Primary 46L55; Secondary 46L35, 46L80. Key words and phrases. C ∗-algebras, group actions, Cuntz semigroup, crossed product, cocycle equivalence. 1 2 EUSEBIO GARDELLA AND LUIS SANTIAGO on noncommutative C∗-algebras. A fundamental result in this area is Julg's iden- tification ([Jul81]) of the equivariant K-theory of a given action, with the ordinary K-theory of its associated crossed product. In a different direction, each of the dif- ferent statements in Atiyah-Segal's characterization of freeness, interpreted in the context of C∗-algebras, can be taken as possible definitions of "noncommutative freeness". This is the approach taken by Phillips in [Phi87]. Equivariant K-theory has also been used as an invariant for compact group actions ([HR85], [Gar14b]), and its definition has been extended to actions of more general objects, such as quantum groups ([ZZ08]). Equivariant K-theory also plays a role in the so called Baum-Connes conjecture with coefficients (where one considers the equivariant K- theory of a C∗-algebra other than C.) On the other hand, the Cuntz semigroup Cu(A) of a C∗-algebra A, first consid- ered by Cuntz in the 70's ([Cun78]), has been intensively studied in the last decade since Toms successfully used it ([Tom08]) to distinguish two non-isomorphic C∗- algebras with identical Elliott invariant (as well as identical real and stable ranks). Coward, Elliott and Ivanescu ([CEI08]) suggested that the Cuntz semigroup could be used as an invariant for C∗-algebras (in many cases, finer than K0), and this semigroup has since then been used to obtain positive classification results of not necessarily simple C∗-algebras. The most general result in this direction is due to Robert ([Rob12]): he showed that the Cuntz semigroup is a complete invariant for (not necessarily simple) unital direct limits of 1-dimensional noncommutative CW-complexes with trivial K1. (This class contains all AI-algebras.) We refer the reader to [BP08] and [APT14] for thorough developments of the theory of the Cuntz semigroup. In this paper, we study an equivariant version of the Cuntz semigroup for com- pact group actions on C∗-algebras. For an action α : G → Aut(A) of a com- pact group G on a C∗-algebra A, we denote its equivariant Cuntz semigroup by CuG(A, α). This is a partially ordered semigroup, with a natural semimodule struc- ture over the representation semiring Cu(G) of G. We explore some basic proper- ties of the functor (A, α) 7→ CuG(A, α), such as continuity, stability, passage to full hereditary subalgebras, cocycle equivalence invariance, etc. One of the main results of this work (Theorem 5.3) is an analog of Julg's theorem for the Cuntz semigroup: CuG(A, α) is naturally isomorphic to Cu(A ⋊α G). The induced Cu(G)-semimodule structure on Cu(A ⋊α G) is computed in Theorem 5.14 (see Proposition 5.16 for a simpler description when G is abelian). We prove an analog of Atiyah-Segal's char- acterization of freeness using the Cuntz semigroup; see Theorem 7.6. Applications to classification of actions are given in the last section; see Theorem 8.4. There are a number of reasons to be interested in an equivariant version of the Cuntz semigroup. Its use for classification of C∗-algebras suggests that this should also be a useful invariant for group actions. As a semigroup, CuG(A, α) is just the Cuntz semigroup of the crossed product, but the additional Cu(G)-semimodule structure makes this a finer invariant. The fact that cocycle equivalent actions have isomorphic equivariant Cuntz semigroups is a stronger statement than the fact that their crossed products have isomorphic Cuntz semigroups. In a different direction, our study is a first step towards a bivariant, equivariant version of the Cuntz semigroup, which is currently being developed by Tornetta; see [Tor16]. (A bivariant version of the Cuntz semigroup has been introduced in [BTZ16].) 3 We have organized this paper as follows. In Section 2, and after reviewing the definition of the Cuntz semigroup and the Cuntz category Cu, we introduce the equivariant Cuntz semigroup using positive invariant elements in suitable stabiliza- tions of the algebra (Definition 2.7). The main result of this section, Corollary 2.14, asserts that the equivariant Cuntz semigroup is a functor from the category of G- C∗-algebras (that is, C∗-algebras with an action of G), to the category Cu. In Section 3, we introduce the representation semiring Cu(G) of G (Definition 3.1), which corresponds to the equivariant Cuntz semigroup of G acting on C (Theorem 3.4), and define a canonical Cu(G)-action on CuG(A, α); see Definition 3.10. With this semimodule structure, the equivariant Cuntz semigroup becomes a functor from G- C∗-algebras to a distinguished category of Cu(G)-semimodules (see Definition 3.7 and Theorem 3.11). We finish this section by showing that the functor CuG is stable (Proposition 3.12) and preserves countable inductive limits (Proposition 3.13). Section 4 is devoted to giving two pictures of the equivariant Cuntz semigroup using equivariant Hilbert modules. In one of these pictures, we identify CuG(A, α) with the ordinary Cuntz semigroup of K(HA)G, where HA is the universal equivari- ant Hilbert module for (A, α) introduced by Kasparov in [Kas80] (see Definition 4.5). In the second picture, we identify CuG(A, α) with the Cu(G)-semimodule obtained by taking a suitable equivalence relation (Definition 4.4) on the class of equivariant Hilbert modules; see Corollary 4.15. Section 5 contains some of our main results. In Theorem 5.3, we construct a natural Cu-isomorphism between CuG(A, α) and Cu(A ⋊α G). (This is a Cuntz semigroup analog of Julg's theorem for K-theory; see [Jul81].) The induced Cu(G)- semimodule structure on Cu(A ⋊α G) is computed in Theorem 5.14, with an easier description available when G is abelian; see Proposition 5.16. As an application, we show that invariant full hereditary subalgebras have canonically isomorphic equivariant Cuntz semigroups (Proposition 5.17). Section 6 contains several computations of the equivariant Cuntz semigroups of a number of dynamical systems. Most of our computations use either the Rokhlin property (which are handled with results from [GS16]) or are pullbacks of dynamical systems (which are handled with Theorem 6.7). In Section 7, we apply the theory developed in the previous sections to prove a characterization of freeness of a compact Lie group action on a compact Hausdorff space, in terms of a certain canonical map into the equivariant Cuntz semigroup; see Theorem 7.6. This characterization resembles (and depends on) Atiyah-Segal's characterization of freeness using equivariant K-theory ([AS69]). Finally, in Section 8, we use the equivariant Cuntz semigroup to classify cer- tain direct limit actions on a class of stably finite C∗-algebras containing all AI- algebras; see Theorem 8.4. Our results extend the work of Handelman and Ross- mann ([HR85]) on locally representable actions on AF-algebras. Throughout the paper, topological groups are always assumed to be Hausdorff. classes of irreducible representations of G. When G is abelian, it is well-known that If G is a locally compact group, we will denote by bG the set of unitary equivalence every irreducible representation is one dimensional, and hence bG has a natural group abelian, then its dual group bG is discrete. A unitary representation µ : G → U(Hµ) structure. When G is compact (but not necessarily abelian), then every irreducible representation is automatically finite dimensional. Finally, if G is compact and 4 EUSEBIO GARDELLA AND LUIS SANTIAGO of G on a Hilbert space Hµ will usually be abbreviated to (Hµ, µ). We say that (Hµ, µ) is separable, or finite dimensional, if Hµ is. The unitary equivalence class of (Hµ, µ) is denoted by [µ]. For a locally compact group G, we denote by Cu(G) the class of unitary equiv- alence classes of unitary representations of G on separable Hilbert spaces. It is easy to see, fixing a separable Hilbert space and restricting to representations on it, that Cu(G) is in fact a set. This set has important additional structure that will not be discussed until it is needed in Section 3. The set Cu(G) will play mostly a notational role in the first few sections. We sometimes make a slight abuse of notation and do not distinguish between elements in bG (or Cu(G)) and irreducible (separable) unitary representations of G. If A is a C∗-algebra, we denote by idA : A → A its identity map. For a unitary u in its multiplier algebra M (A), we write Ad(u) for the automorphism of A given by Ad(u)(a) = uau∗ for a ∈ A. Writing U(M (A)) for the unitary group of M (A), observe that the map Ad : U(M (A)) → Aut(A) is a group homomorphism. For Hilbert spaces H1, H2, we write B(H1, H2) for the Banach space of bounded, linear operators from H1 to H2. We write N for {1, 2, . . .}, Z≥0 for {0, 1, 2, . . .}, and Z≥0 for {0, 1, 2, . . . , ∞}. For a locally compact group G and a C∗-algebra A, an action of G on A will always mean a group homomorphism α : G → Aut(A) satisfying the following con- tinuity condition: for a ∈ A, the map G → A, given by g 7→ αg(a) for g ∈ G, is continuous. (This is usually referred to as "strong continuity".) We write AG, or Aα if we need to stress the action α, for the C∗-subalgebra AG = {a ∈ A : αg(a) = a for all g ∈ G}. Acknowledgements. The authors are grateful to Chris Phillips for valuable conversations on equivariant K-theory. The first named author would like to thank Tron Omland for helpful email correspondence regarding coactions, and also Gabriele Tornetta for numerous discussions on the topic of this work, as well as for pointing out that our previous version of Theorem 8.1 had an unnecessary assump- tion. Finally, we thank the anonymous referee for his comments and suggestions, which greatly improved the present paper, and particularly made it more accessible to the reader. Most of this work was done while the first named author was visiting the Westfalische Wilhelms-Universitat Munster in June-September of 2013, and while both authors were participating in the Thematic Program on Abstract Harmonic Analysis, Banach and Operator Algebras, at the Fields Institute for Research in Mathematical Sciences at the University of Toronto, in January-June 2014. The authors wish to thank both Mathematics departments for their hospitality and for providing a stimulating research environment. 2. The Equivariant Cuntz semigroup In this section, for a continuous action α : G → Aut(A) of a compact group G on a C∗-algebra A, we define its equivariant Cuntz semigroup CuG(A, α) (Definition 2.7), and explore some basic properties. The main result of this section, Corollary 2.14, asserts that the equivariant Cuntz semigroup is a functor from the category of G-C∗-algebras to the Cuntz category Cu. 5 2.1. The Cuntz semigroup and the category Cu. This subsection is devoted to reviewing the construction of the ordinary (that is, non equivariant) Cuntz semi- group, as well as the definition of the Cuntz category Cu. Let A be a C∗-algebra and let a, b ∈ A be positive elements. We say that a is Cuntz subequivalent to b, written a - b, if there exists a sequence (dn)n∈N in A, n − ak = 0. We say that a is Cuntz equivalent to b, written such that lim n→∞ kdnbd∗ a ∼ b, if a - b and b - a. It is clear that - is a preorder on the set of positive elements of A, and thus ∼ is and equivalence relation. We denote by [a] the Cuntz equivalence class of the element a ∈ A+. Definition 2.1. The Cuntz semigroup of A, denoted by Cu(A), is defined as the set of equivalence classes of positive elements of A ⊗ K. Addition in Cu(A) is given by [a] + [b] =(cid:20)(cid:18)a 0 b(cid:19)(cid:21) , 0 where the positive element inside the brackets in the right hand side is identified with its image in A ⊗ K under any ∗-isomorphism of M2(A ⊗ K) with A ⊗ K induced by an ∗-isomorphism of M2(K) and K. The set Cu(A) becomes a partially ordered semigroup when equipped with the partial order [a] ≤ [b] if a - b. It is easy to see that any ∗-homomorphism φ : A → B induces an order preserving map Cu(φ) : Cu(A) → Cu(B) defined by Cu(φ)([a]) = [(φ ⊗ idK)(a)] for a ∈ (A ⊗ K)+. It was shown in [CEI08, Theorem 1] that Cu is a functor from the category of C∗-algebras to a subcategory of the category of ordered abelian semigroups. We now proceed to define this category, which we denote by Cu. Let S be an ordered semigroup and let s, t ∈ S. We say that s is compactly contained in t, and denote this by s ≪ t, if whenever (tn)n∈N is an increasing sequence in S such that t ≤ sup tn, there exists k ∈ N such that s ≤ tk. A sequence n∈N (sn)n∈N in S is said to be rapidly increasing, if sn ≪ sn+1 for all n ∈ N. Definition 2.2. An ordered abelian semigroup S is an object in the category Cu if it has a zero element and it satisfies the following properties: (O1) Every increasing sequence in S has a supremum; (O2) For every s ∈ S, there exists a rapidly increasing sequence (sn)n∈N in S such that s = sup n∈N sn. (O3) If (sn)n∈N and (tn)n∈N are increasing sequences in S, then sup n∈N sn + sup n∈N tn = sup n∈N (sn + tn); (O4) If s1, s2, t1, t2 ∈ S satisfy s1 ≪ t1 and s2 ≪ t2, then s1 + s2 ≪ t1 + t2. Let S and T be semigroups in the category Cu. An order preserving semigroup map ϕ : S → T is a morphism in the category Cu if it preserves the zero element and it satisfies the following properties: (M1) If (sn)n∈N is an increasing sequence in S, then ϕ(cid:18)sup n∈N sn(cid:19) = sup n∈N ϕ(sn); 6 EUSEBIO GARDELLA AND LUIS SANTIAGO (M2) If s, t ∈ S satisfy s ≪ t, then ϕ(s) ≪ ϕ(t). The following observation will be used repeatedly. Remark 2.3. Let M be a partially ordered semigroup with zero element, and let S be a semigroup in Cu. Suppose that there exists a semigroup morphism ϕ : M → S (or ϕ : S → M ) preserving the zero element and such that: (1) ϕ preserves the order, that is, x ≤ y in M implies ϕ(x) ≤ ϕ(y) in S; (2) ϕ is an order embedding, that is, ϕ(x) ≤ ϕ(y) in S implies x ≤ y in M (this implies that ϕ is injective); and (3) ϕ is surjective. Then M belongs to Cu, and ϕ is a Cu-isomorphism. In particular, ϕ automatically preserves suprema of increasing sequences and the compact containment relation. Let A be a C∗-algebra and let a ∈ A. Then it can be checked that [(a−ε)+] ≪ [a] [(a − ε)+]. In particular, it follows that Cu(A) for all ε > 0, and that [a] = sup ε>0 satisfies Axiom (O2). 2.2. The equivariant Cuntz semigroup. For the rest of this section, we fix a compact group G, a C∗-algebra A, and an action α : G → Aut(A). Let (Hµ, µ) and (Hν , ν) be separable unitary representations of G. We endow the Banach space B(Hµ, Hν) with the G-action given by g · T = νg ◦ T ◦ µg−1 , for g ∈ G and T ∈ B(Hµ, Hν). It is clear that K(Hµ, Hν) is an invariant, closed subspace, which we will endow with the restricted G-action. With these actions, a G-invariant linear map is precisely a map Hµ → Hν that is equivariant with respect to µ and ν. Let H1 and H2 be Hilbert spaces, and let A be a C∗-algebra. We will denote by K(H1, H2) ⊗ A the Banach subspace of the C∗-algebra K(H1 ⊕ H2) ⊗ A of operators of the form (cid:18) 0 0 ∗ 0 (cid:19) . With this convention, it is easy to check that for x ∈ K(H1, H2) ⊗ A and y ∈ K(H2, H3) ⊗ A, then the product (composition) xy belongs to K(H1, H3) ⊗ A. (We point out that what we denote by K(H1, H2) ⊗ A can be canonically identified with K(H1 ⊗ A, H2 ⊗ A), where H1 ⊗ A and H2 ⊗ A denote the exterior tensor products of the Hilbert modules. See, for example, [Lan95].) Remark 2.4. Let G be a locally compact group. If µ : G → U(Hµ) and ν : G → U(Hν ) are unitary representations, then δ = µ ⊕ ν is a unitary representation on H = Hµ ⊕ Hν, and conjugation by δ defines an action Ad(δ) : G → Aut(K(H)). If moreover α : G → Aut(A) is an action, there is a natural action γ of G on the C∗-algebra K(H) ⊗ A given by γg = Ad(δg) ⊗ αg for all g ∈ G. It is easy to check that K(Hµ, Hν) ⊗ A is invariant with respect to γ. We will therefore write (K(Hµ, Hν ) ⊗ A)G for those elements in K(Hµ, Hν ) ⊗ A that are fixed under γ. Definition 2.5. Let (Hµ, µ) and (Hν , ν) be separable unitary representations of G, and let a ∈ (K(Hµ) ⊗ A)G and b ∈ (K(Hν ) ⊗ A)G be positive elements. We say that a is G-Cuntz subequivalent to b, and denote this by a -G b, if there exists a sequence (dn)n∈N in (K(Hµ, Hν) ⊗ A)G such that lim n − ak = 0. We say n→∞ kdnbd∗ 7 that a is G-Cuntz equivalent to b, and denote this by a ∼G b, if a -G b and b -G a. The G-Cuntz equivalence class of a positive element a ∈ (K(Hµ) ⊗ A)G will be denoted by [a]G. We claim that the relation -G is transitive. To see this, let (Hµ, µ), (Hν , ν) and (Hλ, λ) be separable unitary representations of G, and let a ∈ (K(Hµ) ⊗ A)G, b ∈ (K(Hν ) ⊗ A)G, and c ∈ (K(Hλ) ⊗ A)G satisfy a -G b and b -G c. Fix ε > 0, and find x ∈ (K(Hµ, Hν ) ⊗ A)G such that ka − xbx∗k < ε 2 . Also, since b -G c there exists y ∈ (K(Hν , Hλ) ⊗ A)G such that kb − ycy∗k < ε 2kxk . The element z = xy belongs to (K(Hµ, Hλ) ⊗ A)G, and is easily seen to satisfy ka − zcz∗k < ε. Since ε > 0 is arbitrary, this implies that a -G c. In particular, it follows that ∼G is an equivalence relation. The following lemma is a simple corollary of [KR02, Lemma 2.4] and the defini- tion of G-Cuntz subequivalence. Proposition 2.6. Let (Hµ, µ) and (Hν, ν) be separable unitary representations of G, and let a ∈ (K(Hµ) ⊗ A)G and b ∈ (K(Hν ) ⊗ A)G be positive elements. The following are equivalent: (1) a -G b. (2) For every ε > 0, there exists d ∈ (K(Hµ, Hν ) ⊗ A)G such that (a − ε)+ = dbd∗. (3) For every ε > 0, there exist δ > 0 and d ∈ (K(Hµ, Hν) ⊗ A)G, such that (a − ε)+ = d(b − δ)+d∗. Definition 2.7. Let G be a compact group, let A be a C∗-algebra, and let α : G → Aut(A) be a continuous action. Define the equivariant Cuntz semigroup CuG(A, α), of the dynamical system (G, A, α), to be the set of G-Cuntz equivalence classes of positive elements in all of the algebras of the form (K(Hµ) ⊗ A)G, where (Hµ, µ) is a separable unitary representation of G. We define addition on CuG(A, α) as follows. Let (Hµ, µ) and (Hν , ν) be separable unitary representations of G, and let a ∈ (K(Hµ) ⊗ A)G and b ∈ (K(Hν ) ⊗ A)G be positive elements. Denote by a ⊕ b the positive element a ⊕ b =(cid:18)a 0 b(cid:19) 0 in (K(Hµ ⊕ Hν ) ⊗ A)G, and set [a]G + [b]G = [a ⊕ b]G. (One must check that the definition is independent of the representatives, but this is routine.) Finally, we endow CuG(A, α) with the partial order given by [a]G ≤ [b]G if a -G b. (One has to again check that the order is well defined; we omit the proof.) It is clear that if β : G → Aut(B) is another continuous action of G on a C∗- algebra B, and if ψ : A → B is an equivariant ∗-homomorphism, then ψ induces an ordered semigroup homomorphism CuG(ψ) : CuG(A, α) → CuG(B, β), given by CuG(ψ)([a]G) = [(idK(Hµ) ⊗ ψ)(a)]G for a ∈ (K(Hµ) ⊗ A)G. The rest of this section is devoted to proving that the equivariant Cuntz semi- group is a functor from the category of G-C∗-algebras to the category Cu (Definition 2.2). This will be accomplished in Corollary 2.14. 8 EUSEBIO GARDELLA AND LUIS SANTIAGO We point out that in Section 3, we will show that the equivariant Cuntz semi- group has additional structure, and that CuG(A, α) belongs to a certain category of semimodules; see Definition 3.7 and Theorem 3.11. Lemma 2.8. Let (Hµ, µ) and (Hν, ν) be separable unitary representations of G, and let a ∈ (K(Hµ) ⊗ A)G be a positive element. Suppose that there exists V ∈ (B(Hµ, Hν) ⊗ A)G satisfying V ∗V = idHµ ⊗ idA. Then V aV ∗ is a positive element in (K(Hν ) ⊗ A)G, and a ∼G V aV ∗. Moreover, if W ∈ (B(Hµ, Hν) ⊗ A)G is another element satisfying W ∗W = idHµ ⊗ idA, then W aW ∗ ∼G V aV ∗. Proof. It is clear that V aV ∗ is a G-invariant element in K(Hν ) ⊗ A. Likewise, for n ∈ N, we have a n V ∗ ∈ (K(Hν ) ⊗ A)G. Now, 1 so a -G V aV ∗. Similarly, V a lim n→∞(cid:13)(cid:13)(cid:13)(cid:16)a1/nV ∗(cid:17) (V aV ∗)(cid:16)a1/nV ∗(cid:17)∗ n→∞(cid:13)(cid:13)(cid:13)(cid:16)V a1/n(cid:17) a(cid:16)V a1/n(cid:17)∗ − a(cid:13)(cid:13)(cid:13) = lim n→∞(cid:13)(cid:13)(cid:13)a1/naa1/n − a(cid:13)(cid:13)(cid:13) = 0, − V aV ∗(cid:13)(cid:13)(cid:13) = 0. n ∈ (K(Hµ) ⊗ A)G, and one shows that lim 1 We conclude that a ∼G V aV ∗, as desired. The last part of the statement is immediate. (cid:3) Let (Hµ, µ) and (Hν , ν) be separable unitary representations of G, such that (Hµ, µ) is unitarily equivalent to a subrepresentation of (Hν, ν). Then there exists Wµ,ν ∈ B(Hµ, Hν)G satisfying W ∗ µ,ν Wµ,ν = idHµ . Set V A µ,ν = Wµ,ν ⊗ idA ∈ B(Hµ, Hν) ⊗ A. It is clear that V A µ,ν is G-invariant and that (V A µ,ν )∗V A µ,ν = idHµ ⊗ idA. Set µ,ν = Ad(V A ιA µ,ν ) : (K(Hµ) ⊗ A)G → (K(Hν ) ⊗ A)G. Then ιA µ,ν is a ∗-homomorphism, since V A µ,ν is an isometry. Let jA µ,ν = Cu(Ad(V A µ,ν )) : Cu((K(Hµ) ⊗ A)G) → Cu((K(Hν ) ⊗ A)G) be the Cu-morphism given by jA the context, we will write Vµ,ν for V A µ,ν = Cu(ιA µ,ν ). Whenever A and α are clear from µ,ν , and similarly for ιµ,ν and jµ,ν . Lemma 2.9. Adopt the notation from the discussion above. (1) The map jµ,ν is independent of the choice of Vµ,ν. (2) If (Hν , ν) is unitarily equivalent to a subrepresentation of the separable representation (Hλ, λ), then jν,λ ◦ jµ,ν = jµ,λ. Proof. The first part is an immediate consequence of Lemma 2.8. The second one is straightforward. (cid:3) Next, we show that CuG(A, α) is an object in Cu. Proposition 2.10. Let α : G → Aut(A) be an action of a compact group G on a C∗-algebra A. Then CuG(A, α) is a semigroup in Cu. 9 Proof. We will check axioms (O1) through (O4) from Definition 2.2. We divide the proof into four claims. Claim 1: CuG(A, α) satisfies (O1). Let (sn)n∈N be an increasing sequence in CuG(A, α). Choose separable representations (Hµn , µn) of G and positive elements an ∈ (K(Hµn ) ⊗ A)G with [an]G = sn for n ∈ N. (Hµn , µn), which is a separable representation. Since µn is a Set (Hµ, µ) = Ln∈N subrepresentation of µ, there exists a natural map jµn,µ : Cu((K(Hµn ) ⊗ A)G) → Cu((K(Hµ) ⊗ A)G) in the category Cu. Since Cu((K(Hµ) ⊗ A)G) is an object in Cu, the supremum jµn,µ(sn) exists in Cu((K(Hµ) ⊗ A)G). Choose a positive element a ∈ s = sup n∈N (K(Hµ) ⊗ A)G with [a] = s. We claim that [a]G is the supremum of (sn)n∈N in CuG(A, α). Let t ∈ CuG(A, α) satisfy sn ≤ t for all n ∈ N. Choose a separable representation (Hν, ν) and a positive element b ∈ (K(Hν ) ⊗ A)G representing t. Then jµn,µ⊕ν(sn) ≤ jµn,µ⊕ν(t) for all n ∈ N. It is clear that a - b in (K(Hµ) ⊗ A)G, so [a]G ≤ [b]G in CuG(A, α), as desired. Claim 2: CuG(A, α) satisfies (O2). Given s ∈ CuG(A, α) choose µ ∈ Cu(G) and a positive element a ∈ (K(Hµ) ⊗ A)G with [a]G = s. Then ([(a − 1 n )+])n∈N is a rapidly increasing sequence in Cu((K(Hµ) ⊗ A)G), and its supremum is [a]. Using the description of suprema from the above paragraph, it follows that the same is true for the elements [(a − 1 n )+]G in CuG(A, α), as desired. Claim 3: CuG(A, α) satisfies (O3). Let (sn)n∈N and (tn)n∈N be increasing se- quences in Cu(G, α). For n ∈ N, choose separable representations (Hµn , µn) and (Hνn , νn) of G, and positive elements an ∈ (K(Hµn )⊗ A)G and bn ∈ (K(Hνn )⊗ A)G satisfying [an]G = sn and [bn]G = tn. Set (Hµ, µ) =Mn∈N (Hµn , µn) and (Hν , ν) =Mn∈N (Hνn , νn), which are separable representations of G. Similarly to what was one in Claim 1, find positive elements a ∈ (K(Hµ) ⊗ A)G and b ∈ (K(Hν ) ⊗ A)G satisfyin [a] = sup n∈N jµn,µ([an]) ∈ Cu((K(Hµ)⊗A)G) and [b] = sup n∈N jνn,ν([bn]) ∈ Cu((K(Hν )⊗A)G). Then [a]G = sup n∈N sn and [b]G = sup n∈N tn, by the proof of Claim 1. Using that Cu((K(Hµ⊕ν ) ⊗ A)G) is an object in Cu at the second step, we get jµ,µ⊕ν ([a]) + jν,µ⊕ν ([b]) = sup n∈N [jµn,µ⊕ν([an]) + jνn,µ⊕ν([bn])] . jµn,µ⊕ν([an]) + sup n∈N jνn,µ⊕ν([bn]) sup n∈N It is then clear that [a]G + [b]G is the supremum of (sn + tn)n∈N in CuG(A, α), and the claim is proved. Claim 4: CuG(A, α) satisfies (O4). Let s1, s2, t1, t2 ∈ CuG(A, α) satisfy sj ≪ tj for j = 1, 2. In order to check that s1 + s2 ≪ t1 + t2, let (rn)n∈N be an increasing 10 EUSEBIO GARDELLA AND LUIS SANTIAGO sequence in CuG(A, α) satisfying t1 + t2 ≤ sup n∈N rn. Find a large enough separable representation (Hµ, µ) of G, and positive elements a1, a2, b1, b2, cn ∈ (K(Hµ) ⊗ A)G, satisfying [aj]G = sj, [bj]G = tj and [cn]G = rn for j = 1, 2 and for n ∈ N. [cn] in Cu((K(Hµ) ⊗ A)G). Since Then [aj] ≪ [bj], for j = 1, 2, and [b1] + [b2] ≤ sup n∈N Cu((K(Hµ)⊗A)G) is an object in Cu, there exists m ∈ N such that [a1]+[a2] ≤ [cm]. It follows that s1 + s2 = [a1]G + [a2]G ≤ [cm]G = rm, as desired. This finishes the proof of the claim and the proposition. (cid:3) Remark 2.11. It follows from the above proof that the supremum of an increasing sequence of elements in CuG(A, α) can be computed in a single space of the form (K(Hµ) ⊗ A)G. Likewise, compact containment can also be verified in a single separable representation. Let us consider the action on K(ℓ2(N) ⊗ Hµ ⊗ A) induced by tensor product of the trivial action of G on ℓ2(N) and the given action of G on Hµ ⊗ A. Then K(ℓ2(N) ⊗ Hµ ⊗ A)G is naturally isomorphic to K(ℓ2(N)) ⊗ ((K(Hµ) ⊗ A)G). Thus, the Cuntz semigroup of (K(Hµ) ⊗ A)G can be naturally identified with the set of (ordinary) Cuntz equivalences classes of positive elements in (K(ℓ2(N) ⊗ Hµ) ⊗ A)G. Fix [µ] ∈ Cu(G). Then the inclusion (K(ℓ2(N) ⊗ Hµ) ⊗ A)G ֒→ G[ν]∈Cu(G) (K(Hν ) ⊗ A)G induces a semigroup homomorphism (1) iµ : Cu((K(Hµ) ⊗ A)G) → CuG(A, α), which is given by iµ([a]) = [a]G for a positive element a ∈ (K(ℓ2(N) ⊗ Hµ) ⊗ A)G. By Lemma 2.8, the map iµ satisfies iµ ◦ jν,µ = iν whenever ν is equivalent to a subrepresentation of µ. It is also clear that iµ preserves suprema of increasing sequences, the compact containment relation, and that it is an order embedding. Define a preorder ≤ on Cu(G) by setting [µ] ≤ [ν] if µ is equivalent to a subrep- resentation of ν. It is clear that (Cu(G), ≤) is a directed set. For use in the next theorem, we recall that by Corollary 3.1.11 in [APT14], the category Cu is closed under direct limits indexed over an arbitrary directed set. Theorem 2.12. The direct limit of the direct system (cid:0)(Cu((K(Hµ) ⊗ A)G))[µ]∈Cu(G), (jµ,ν )[µ],[ν]∈Cu(G),[µ]≤[ν](cid:1) , in the category Cu, is naturally isomorphic to the pair (CuG(A, α), (iµ)µ∈Cu(G)). For ease of notation, we will denote elements of Cu(G) and representatives with the same symbols. Proof. We will show that (CuG(A, α), (iµ)µ∈Cu(G)) satisfies the universal property of the direct limit in Cu. Let (S, (γµ)µ∈Cu(G)) be a pair consisting of a semigroup S in the category Cu and Cu-morphisms γµ : Cu((K(Hµ) ⊗ A)G) → S, 11 for µ ∈ Cu(G), satisfying γν ◦ jµ,ν = γµ for all ν, µ ∈ Cu(G) with µ ≤ ν. Define a map γ : CuG(A, α) = [µ∈Cu(G) iµ(Cu((K(Hµ) ⊗ A)G)) → S by γ(iµ(s)) = γµ(s) for s ∈ Cu((K(Hµ) ⊗ A)G). The proof will be finished once we prove that γ is a well-defined morphism in Cu. We divide the proof into a number of claims. Claim: γ is a well defined order preserving map. For this, it is enough to show the following. Given µ, ν ∈ Cu(G) and given s ∈ Cu((K(Hµ) ⊗ A)G) and t ∈ Cu((K(Hν ) ⊗ A)G), if iµ(s) ≤ iν(t), then Let µ, ν, s and t be as above. Then γ(iµ(s)) ≤ γ(iν(t)). iµ⊕ν(jµ,µ⊕ν (s)) = iµ(s) ≤ iν(s) = iµ⊕ν (jν,µ⊕ν (t)). Since iµ⊕ν is an order embedding, we deduce that jµ,µ⊕ν (s) ≤ jν,µ⊕ν (t). Hence, γ(iµ(s)) = γµ(s) = γµ⊕ν(jµ,µ⊕ν (s)) ≤ γµ⊕ν(jν,µ⊕ν (t)) = γν (t) = γ(iν(t)). The claim is proved. Claim: γ is a semigroup homomorphism. Given µ, ν ∈ Cu(G), given s ∈ Cu((K(Hµ) ⊗ A)G), and given t ∈ Cu((K(Hν ) ⊗ A)G), we have γ(iµ(s) + iν(t)) = γµ⊕ν (jµ,µ⊕ν (s) + jν,µ⊕ν (t)) = γµ⊕ν (jµ,µ⊕ν (s)) + γµ⊕ν (jν,µ⊕ν (t)) = γ(iµ(s)) + γ(iν(t)), so the claim follows. Claim: γ preserves suprema of increasing sequences (condition (M1) in Definition 2.2). Let (xn)n∈N be an increasing sequence in CuG(A, α), and let x ∈ CuG(A, α) be its supremum. For each n ∈ N, choose [µn] ∈ Cu(G) and an element sn ∈ Cu((K(Hµn ) ⊗ A)G) such that iµn (sn) = xn. Likewise, choose [µ] ∈ Cu(G) and an element s ∈ Cu((K(Hµ) ⊗ A)G) such that iµ(s) = x. Set ν = µ ⊕ µn. Then ∞Ln=1 iν(jµn,ν(sn)) = iµn (sn) ≤ iµn+1(sn+1) = iν(jµn+1,ν(sn+1)) for all n ∈ N. It follows that jµn,ν(sn) ≤ jµn+1,ν(sn+1) for all n ∈ N, since iν is an order embedding. In other words, (jµn,ν(sn))n∈N is an increasing sequence in 12 EUSEBIO GARDELLA AND LUIS SANTIAGO Cu((K(Hν ) ⊗ A)G). Since suprema of increasing sequences exist in Cu((K(Hν ) ⊗ A)G) and iν and γν are maps in Cu we get γ(x) =γ(iµ(s)) = γ(cid:18)sup = γ(iν(cid:18)sup n∈N = sup n∈N γ(iµn (sn)) = sup n∈N γ(xn). n∈N iµn (sn)(cid:19) = γ(cid:18)sup jµn,ν(sn)(cid:19)) = γν(sup n∈N n∈N iν(jµn,ν(sn))(cid:19) jµn,ν(sn)) = sup n∈N γν(jµn,ν(sn)) Hence γ(x) is the supremum of (γ(xn))n∈N, proving the claim. Claim: γ preserves the compact containment relation (condition (M2) in Definition 2.2). Given µ, ν ∈ Cu(G), and given s ∈ Cu((K(Hµ) ⊗ A)G) and t ∈ Cu((K(Hν ) ⊗ A)G), suppose that iµ(s) ≪ iν(t). Then iµ⊕ν(jµ,µ⊕ν (s)) ≪ iν(jν,µ⊕ν (t)). Since iµ⊕ν is a morphism in the category Cu and it is an order embedding, we deduce that jµ,µ⊕ν (s) ≪ jν,µ⊕ν (t). Hence, γ(iµ(s)) = γµ⊕ν (jµ,µ⊕ν (s)) ≪ γν⊕ν(jν,µ⊕ν (t)) = γ(iν(t)). We conclude that γ is a morphism in Cu, so the proof is complete. (cid:3) We can now show that semigroup homomorphisms between the equivariant Cuntz semigroups induced by equivariant ∗-homomorphisms are morphisms in Cu. Proposition 2.13. Let β : G → Aut(B) be an action of G on a C∗-algebra B, and let φ : A → B be an equivariant ∗-homomorphism. Then the induced map CuG(φ) : CuG(A, α) → CuG(B, β) is a morphism in the category Cu. Proof. For [µ] ∈ Cu(G), set φµ = idK(Hµ) ⊗ φ : (K(Hµ) ⊗ A, Ad(µ) ⊗ α) → (K(Hµ) ⊗ B, Ad(µ) ⊗ β). Then φµ is equivariant. Its induced map Cu(φµ) : Cu((K(Hµ) ⊗ A)G) → Cu((K(Hµ) ⊗ B)G), between the Cuntz semigroups of the corresponding fixed point algebras, is a mor- phism in Cu. For [µ] ≤ [ν], we have Consequently, the maps jB µ,ν ◦ Cu(φµ) = Cu(φν ) ◦ jA µ,ν . µ ◦ Cu(φµ) : Cu((K(Hµ) ⊗ A)G) → CuG(B, β) iB satisfy µ ◦ Cu(φµ) = (iB iB ν ◦ Cu(φν )) ◦ jA µ,ν for [µ] ≤ [ν]. The universal property of the direct limit provides a Cu-morphism κ : CuG(A, α) → CuG(B, β) satisfying κ ◦ iA have µ = iB µ ◦ Cu(φµ) for all [µ] ∈ Cu(G). For s ∈ Cu((K(Hµ) ⊗ A)G), we κ(iA µ (s)) = iB µ (Cu(idK(Hµ) ⊗ φ)([a])) = CuG(φ)([a]). We conclude that κ = CuG(φ), and hence CuG(φ) is a morphism in Cu. µ (Cu(φµ)(s)) = iB (cid:3) 13 Since CuG obviously preserves composition of maps, we get the following. Corollary 2.14. The equivariant Cuntz semigroup CuG is a functor from the category of G-C∗-algebras to the category Cu. 3. The semiring Cu(G) and the category CuG 3.1. The semiring Cu(G). Let G be a compact group. Denote by V (G) the semigroup of equivalence classes of finite dimensional representations of G, the operation being given by direct sum. Recall that the representation ring R(G) of G is the Grothendieck group of V (G). The product structure on R(G) is induced by the tensor product of representations. The construction of R(G) resembles that of K-theory, while the object we define below is its Cuntz semigroup analog. Recall that a semiring is a set R with two binary operations + and · on R, which satisfy all axioms of a unital ring except for the axiom demanding the existence of additive inverses. Definition 3.1. The representation semiring of G, denoted by Cu(G), is the set of all equivalence classes of unitary representations of G on separable Hilbert spaces. Addition in Cu(G) is given by the direct sum of representations, while product in Cu(G) is given by the tensor product. We endow Cu(G) with the order: [µ] ≤ [ν] if µ is unitarily equivalent to a subrepresentation of ν. Since the tensor product of representations is associative, it is clear that Cu(G) is indeed a semiring. Lemma 3.2. Let A be a C∗-algebra, let G be a compact group, and let α : G → Aut(A) be an action. Let (Hµ, µ) be a separable unitary representation of G, and let a ∈ K(Hµ)G be a positive element. Set H = a(Hµ), and let a′ be the restriction of a to H. Then a′ is a G-invariant strictly positive element in K(H), and there exists a sequence (dn)n∈N in K(Hµ, H)G such that lim n→∞ kd∗ na′dn − ak = 0 and lim n→∞ kdnad∗ n − a′k = 0. Proof. Denote by BHµ and BH the unit balls Hµ and H, respectively. Since a′(BH) ⊆ a(BHµ ), it is clear that a′ is compact. 1 n (ξ) = ξ for all ξ ∈ H, we conclude that a′ is strictly positive. For Since lim n→∞ a n ∈ N, let dn : Hµ → H be the operator defined by restricting the codomain of a to H. Then dn ∈ K(Hµ, H)G, and d∗ ξ ∈ H. It is now clear that n : H → Hµ is given by d∗ 1 n(ξ) = a n (ξ) for all 1 n lim n→∞ kd∗ na′dn − ak = 0 and lim n→∞ kdnad∗ n − a′k = 0, so the proof is complete. (cid:3) The following observation will be used throughout without particular reference. Recall that a positive element x in a C∗-algebra A is said to be strictly positive if τ (x) > 0 for every positive linear map τ : A → C. Remark 3.3. Let G be a compact group, let A be a C∗-algebra, and let α : G → Aut(A) be an action. Denote by µ the normalized Haar measure on G. If x ∈ A is αg(x) dµ(g) is strictly positive in A. Indeed, a strictly positive element, then y =RG let τ : A → C be a positive linear map. For g ∈ G, the map τ ◦ αg : A → C is also 14 EUSEBIO GARDELLA AND LUIS SANTIAGO linear and positive, and so τ (αg(x)) > 0. Since g 7→ τ (αg(x)) is continuous, we deduce that τ (y) = τZG αg(x) dµ(g) =ZG so y is strictly positive, as desired. τ (αg(x)) dµ(g) > 0, Let (Hµ, µ) be a separable unitary representation of G. Since Hµ is separable, we get an invariant strictly positive element sµ of K(Hµ). K(Hµ) has a strictly positive element esµ. Moreover, by integrating g ·esµ over G, Theorem 3.4. Adopt the notation from the comments above. Then the map s : Cu(G) → CuG(C) given by s([µ]) = [sµ], for [µ] ∈ Cu(G), is well defined. Moreover, it is an isomorphism of ordered semigroups. Proof. We begin by showing that s is well defined. Let (Hµ, µ) and (Hν , ν) be separable unitary representations of G, with [µ] ≤ [ν]. Then there exists V ∈ B(Hµ, Hν)G such that V ∗V = idHµ. By Lemma 2.8, we have sµ ∼G V sµV ∗. Since sν is strictly positive, we also have V sµV ∗ -G sν. Thus, sµ -G sν. It follows that s is well defined and order preserving. We now show that s is an order embedding. Let sµ ∈ K(Hµ)G and sν ∈ K(Hν )G be strictly positive elements such that sµ -G sν. By [CES11, Proposition 2.5] applied to K(Hµ ⊕Hν) (with the convention from before Definition 2.5), there exists x ∈ K(Hµ, Hν) such that sµ = x∗x and xx∗ ∈ K(Hν )G. Also, by simply inspecting the proof of that proposition, one sees that x can be taken in K(Hµ, Hν)G. Let 2 be the polar decomposition of x. Then v belongs to B(Hµ, Hν)G, and x = v(x∗x) v∗v = idHµ . This implies that [µ] ≤ [ν]. In particular, s is injective. 1 To finish the proof, we show that s is surjective. Let (Hµ, µ) be a separable unitary representation of G, and let a be a strictly positive element in K(Hµ)G. Set Hν = a(Hµ), let ν be the restriction of µ to Hν, and let a′ : Hν → Hν be the restriction of a. By Lemma 3.2, a′ is a positive element in K(Hν )G, and a′ ∼G a. It follows that s([ν]) = [a], and the proof is complete. (cid:3) In particular, the above theorem shows that Cu(G) is a Cu-semiring in the sense of Definition 7.1.1 in [APT14]. Corollary 3.5. The semigroup Cu(G) is an object in Cu. In addition, (1) If ([µn])n∈N is an increasing sequence in Cu(G), then [µ] is the supremum of ([µn])n∈N if and only if [sµ] = sup n∈N [sµn ]; (2) [µ] ≪ [ν] if and only if [sµ] ≪ [sν ]. Recall that when G is compact, every unitary representation of G is equivalent to a direct sum of finite dimensional representations. Corollary 3.6. Let (Hµ, µ) be a separable unitary representation of G. Let (Hνk , νk)k∈N be a family of non-zero finite dimensional representations of G such that (Hµ, µ) ∼=Mk∈N (Hνk , νk). For n ∈ N, set µn = nLk=1 νk. Then [µn] ≪ [µ] for all n ∈ N, and [µ] = sup n∈N [µn]. 15 Proof. Let sµ be a strictly positive element of K(Hµ)G. For each n ∈ N, let pn be Hνk . Also, [pn] ≪ [sµ] for the unit of K(Hµn ). Then [sµ] = sup n∈N all n ∈ N, because pn is a projection. The result then follows from Corollary 3.5. (cid:3) [pn] since Hµ ∼= Lk∈N 3.2. The Cu(G)-semimodule structure on CuG(A, α). Throughout the rest of this section, we fix a compact group G, a C∗-algebra A, and a continuous action α : G → Aut(A). Recall that a (left) semimodule over a semiring R, or an R-semimodule, is a commutative monoid S together with a function · : R × S → S satisfying all the axioms of a module over a ring, except for the axiom demanding the existence of additive inverses. In this subsection, we show that CuG(A, α) has a natural Cu(G)-semimodule structure, which moreover satisfies a number of additional regularity properties. It follows that the equivariant Cuntz semigroups belong to a distinguished class of partially ordered semirings over Cu(G). We begin by defining this category, and then show that CuG(A, α) belongs to it; see Theorem 3.11. Definition 3.7. Denote by CuG the category defined as follows. The objects in CuG are partially ordered Cu(G)-semimodules (S, +, ·) such that: (O1) S is an object in Cu; (O2) if x, y ∈ S and r, s ∈ Cu(G) satisfy x ≤ y and r ≤ s, then r · x ≤ s · y; (O3) if x, y ∈ S and r, s ∈ Cu(G) satisfy x ≪ y and r ≪ s, then r · x ≪ s · y; (O4) if (xn)n∈N is an increasing sequence in S, and (rn)n∈N is an increasing sequence in Cu(G), then (rn · xn) =(cid:18)sup n∈N rn(cid:19) ·(cid:18)sup n∈N xn(cid:19) . sup n∈N The morphisms in CuG between two Cu(G)-semimodules S and T are all Cu(G)- semimodule homomorphisms ϕ : S → T in the category Cu. In particular, a Cu(G)-semimodule is a Cu-semimodule in the sense of Defini- tion 7.1.3 of [APT14]. Lemma 3.8. Axiom (O4) in Definition 3.7 is equivalent to the following. If (xn)n∈N is an increasing sequence in S, and (rn)n∈N is an increasing sequence in Cu(G), then (rn · x) =(cid:18)sup n∈N sup n∈N rn(cid:19) · x and sup n∈N (r · xn) = r ·(cid:18)sup n∈N xn(cid:19) for all r ∈ Cu(G) and for all x ∈ S. Proof. That Axiom (O4) implies the condition in the statement is immediate. Con- versely, suppose that S satisfies Axioms (O1), (O2) and (O3), and the condition in the statement. Let (xn)n∈N be an increasing sequence in S, and let (rn)n∈N be an xn by increasing sequence in Cu(G). For m ∈ N, we have rm · xm ≤ sup n∈N rn · sup n∈N 16 EUSEBIO GARDELLA AND LUIS SANTIAGO Axiom (O2), so sup m∈N (rm · xm) ≤(cid:18)sup For the opposite inequality, given m ∈ N we have n∈N rn(cid:19) ·(cid:18)sup (rn · xm) =(cid:18)sup xn(cid:19) . rn(cid:19) xm. n∈N n∈N sup n∈N (rn · xn) ≥ sup n∈N By taking sup m∈N , we conclude that Axiom (O4) is also satisfied. (cid:3) We will need to know the following: Theorem 3.9. The category CuG is closed under countable direct limits. Proof. Let (Sn, ϕn)n∈N be a direct system in the category CuG, with CuG-mor- phisms ϕn : Sn → Sn+1. For m ≥ n, we write ϕm,n : Sn → Sm+1 for the composi- tion ϕm,n = ϕm ◦ · · · ◦ ϕn. By Theorem 2 in [CEI08], the limit of this direct system exists in the category Cu, and we denote it by (S, (ψn)n∈N), where ψn : Sn → S is a Cu-morphism satisfying ψn+1 ◦ ϕn = ψn for all n ∈ N. We will use the description of the direct limit given in the proof of Theorem 2 in [CEI08], in the form given in Proposition 2.2 in [GS16]. We define a Cu(G)-semimodule structure on S as follows. Let x ∈ S, and choose ψn(xn) = x. Given elements xn ∈ Sn, for n ∈ N, such that ϕn(xn) ≪ xn+1 and sup n∈N r ∈ Cu(G), set r · x = sup n∈N ψn(r · xn). Claim: the Cu(G)-semimodule structure is well-defined and satisfies Axiom (O2). It is clearly enough to check Axiom (O2). Let x, y ∈ S with x ≤ y, and let r, s ∈ Cu(G) with r ≤ s. Choose elements xn, yn ∈ Sn, for n ∈ N, satisfying ψn(yn) = y. ϕn(xn) ≪ xn+1 and sup n∈N ψn(xn) = x, as well as ϕn(yn) ≪ yn+1 and sup n∈N Given n ∈ N, we have ψn(xn) ≪ x ≤ y = supm∈N ψm(ym), so there exists m0 ∈ N such that ψn(xn) ≤ ψm(ym) for all m ≥ m0. Without loss of generality, we may assume m0 ≥ n. Since ϕn(xn) ≪ xn+1, part (ii) of Proposition 2.2 in [GS16] implies that there exists n0 ∈ N with n0 ≥ m such that ϕn0,n(xn) ≤ ϕn0,m(ym). It follows that ϕn0,n(r · xn) ≤ ϕn0,m(s · ym) for all k ≥ n0. Composing with ψn0, we deduce that ψn(r · xn) ≤ ψm(s · ym) for all m ≥ m0. Taking first the supremum over m, and then the supremum over n, we conclude that sup n∈N ψn(r · xn) ≤ sup m∈N ψm(s · ym). The claim is proved. Claim: S satisfies Axiom (O3). Let x, y ∈ S with x ≪ y, and let r, s ∈ Cu(G) with r ≪ s. Then there exist n ∈ N, and x′, y′ ∈ Sn such that x′ ≪ y′ and x ≪ ψn(x′) ≪ ψn(y′) ≪ y. Then r · x′ ≪ s · y′, and hence r · x ≤ ψn(r · x′) ≪ ψn(s · y′) ≤ s · y, as desired. Claim: S satisfies Axiom (O4). It suffices to check the conditions in the state- ment of Lemma 3.8. Let (rn)n∈N be an increasing sequence in Cu(G), and let 17 x ∈ S. Choose elements xm ∈ Sm, for m ∈ N, such that ϕm(xm) ≪ xm+1 and sup m∈N ψm(xm) = x. Then rn ·(cid:18) sup m∈N ψm(xm)(cid:19) =(cid:18)sup n∈N rn(cid:19) · x, sup n∈N (rn · x) = sup n∈N sup m∈N ψm(rn · xm) = sup n∈N as desired. The other property in Lemma 3.8 is checked identically. This concludes the proof. (cid:3) We now define a Cu(G)-semimodule structure on CuG(A, α). Definition 3.10. Let (Hµ, µ) and (Hν , ν) be separable unitary representations of G, and let a ∈ (K(Hµ) ⊗ A)G be a positive element. In this definition, and to stress the role played by µ, we write [(Hµ, µ, a)]G for the G-Cuntz equivalence class of a. Use separability of Hν to choose a G-invariant strictly positive element sν ∈ K(Hν )G. We set [ν] · [(Hµ, µ, a)]G = [(Hν ⊗ Hµ, ν ⊗ µ, sν ⊗ a)]G . The following is one of the main results in this section. For use in its proof, we recall that any tensor product of C∗-algebras respects Cuntz subequivalence. Theorem 3.11. The Cu(G)-semimodule structure from Definition 3.10 is well de- fined. Moreover, with this structure, the semigroup CuG(A, α) becomes an object in CuG, and the equivariant Cuntz semigroup is a functor from the category of G-C∗-algebras to the category CuG. Proof. We will prove that the Cu(G)-semimodule structure is well defined together with condition O2 in Definition 3.7. So let [µ], [ν] ∈ Cu(G) and [a]G, [b]G ∈ CuG(A, α) satisfy [µ] ≤ [ν] and [a]G ≤ [b]G. By Theorem 3.4, we have sµ -G sν. Since we also have a -G b, we get sµ ⊗ a -G sν ⊗ b. Hence, [µ] · [a]G = [sµ ⊗ a]G ≤ [sν ⊗ b]G = [ν] · [b]G, as desired. It is immediate that [µ] · ([a]G + [b]G) = [µ] · [a]G + [µ] · [b]G and [µ] · ([ν] · [a]G) = ([µ] · [ν]) · [a]G, for all [µ], [ν] ∈ Cu(G) and for all [a]G, [b]G ∈ CuG(A, α). We conclude that CuG(A, α) is a Cu(G)-semimodule, and that it satisfies condition O2 in Definition 3.7. We now proceed to show that CuG(A, α) is an object in CuG. We already showed in Theorem 2.12 that it is an object in Cu, so condition O1 in Definition 3.7 is satisfied. We check condition O3. Suppose that [µ], [ν] ∈ Cu(G) and [a]G, [b]G ∈ CuG(A, α) satisfy [µ] ≪ [ν] and [a]G ≪ [b]G. By Theorem 3.4, we get [sµ]G ≪ [sν ]G in CuG(C). Without loss of generality, we may assume that sν and b are contractions. Recall that [sν]G = sup ε>0 [(sν − ε)+]G and [b]G = sup ε>0 [(b − ε)+]G. Using the definition of the compact containment relation, find ε > 0 such that [sµ]G ≤ [(sν − ε)+]G and [a]G ≤ [(b − ε)+]G. 18 EUSEBIO GARDELLA AND LUIS SANTIAGO Use ksνk ≤ 1 and kbk ≤ 1 at the third step to get [µ] · [a]G = [sµ ⊗ a]G ≤ [(sν − ε)+ ⊗ (b − ε)+]G ≤ [(sν ⊗ b − ε2)+]G ≪ [sν ⊗ b]G = [ν] · [b]G, so condition O3 is satisfied. We now check condition O4. Let ([µn])n∈N and ([an]G)n∈N be increasing se- [µn] and [a]G = quences in Cu(G) and CuG(A, α), respectively, and set [µ] = sup n∈N [an]G. Without loss of generality, we may assume that a is a contraction. For sup n∈N n ∈ N, denote by sµn an invariant strictly positive element in K(Hµn ), and denote by sµ an invariant strictly positive element in K(Hµ). By part (1) of Corollary 3.5, [sµn ]G in CuG(C). As before, we may assume that sµ is a we have [sµ]G = sup n∈N contraction. Then the sequence ([sµn ⊗ an]G)n∈N in CuG(A, α) is increasing. Set [c]G = sup n∈N [sµn ⊗ an]G. We claim that [c]G = [sµ ⊗ a]G. It is clear that [c]G ≤ [sµ ⊗ a]G. To check the opposite inequality, let ε > 0. Then there exists n ∈ N such that [(sµ − ε)+] ≤ [sµn ]G and [(a − ε)+]G ≤ [an]G. It follows that [(sµ ⊗ a − ε)+]G ≤ [(sµ − ε)+ ⊗ (a − ε)+]G ≤ [sµn ⊗ an]G. Hence, [(sµ ⊗ a − ε)+] ≤ [c]. Since [sµ ⊗ a]G = sup ε>0 [(sµ ⊗ a − ε)+]G, we deduce that [sµ ⊗ a]G ≤ [c]G, as desired. We have checked condition O4. Since CuG(A, α) is an object in Cu by Theorem 2.12, we conclude that CuG(A, α) is an object in CuG. It remains to argue that CuG is a functor into CuG. Let β : G → Aut(B) be a continuous action of G on a C∗-algebra B, and let φ : A → B be an equivariant ∗-homomorphism. By Corollary 2.14, CuG(φ) is a morphism in Cu, so we only need to check that it is a morphism of Cu(G)-semimodules. This is immediate, so the proof is complete. (cid:3) We mention here that naturality of the isomorphism in Theorem 2.12 implies that said isomorphism becomes a CuG-isomorphism when lim −→(cid:0)(Cu((K(Hµ) ⊗ A)G))µ∈Cu(G), (jµ,ν)µ,ν∈Cu(G),µ≤ν(cid:1) is endowed with the following Cu(G)-action. For separable representations (Hµ, µ) and (Hν , ν) of G, and for x ∈ Cu((K(Hµ) ⊗ A)G), we set [ν] · x = jν⊗µ,µ(x). 3.3. Functorial properties of CuG. In this subsection, we prove two functoriality properties of the equivariant Cuntz semigroup. Proposition 3.12 asserts that this functor is stable when the compact operators are given the trivial G-action. We will generalize this result in Corollary 5.7, where we replace the trivial action on K with an arbitrary inner action. (Stability fails if the action on K is not inner; see Example 5.8.) Then, in Proposition 3.13, we show that the equivariant Cuntz 19 semigroup preserves equivariant inductive limits of sequences. Both propositions in this subsection will be needed in Section 5. Proposition 3.12. Let q be any rank one projection on ℓ2(N), and denote by ιq : A → A ⊗ K(ℓ2(N)) the inclusion obtained by identifying A with A ⊗ qK(ℓ2(N))q. ιq(a) = a ⊗ q for a ∈ A. In other words, Give ℓ2(N) the trivial G-representation. Then ιq induces a natural CuG-isomor- phism CuG(ιq) : CuG(A, α) → CuG(A ⊗ K(ℓ2(N)), α ⊗ idK(ℓ2(N))). Proof. We abbreviate K(ℓ2(N)) to K. By Theorem 3.11, CuG(ιq) is a morphism in CuG. It thus suffices to check that it is an isomorphism in Cu. Let (Hµ, µ) be a separable unitary representation of G. Denote by κq µ : Cu((K(Hµ) ⊗ A)G) → Cu((K(Hµ) ⊗ A)G ⊗ K) the Cu-morphism induced by the inclusion as the corner associated to q. Then κq µ is an isomorphism (see Appendix 6 in [CEI08]). With the notation from Theorem 3.11, it is clear that for all ν ∈ Cu(G) with µ ≤ ν. jA⊗K µ,ν ◦ κq ν = κq µ ◦ jA µ,ν By the universal property of the direct limit in Cu, applied to the object CuG(A⊗ µ, for µ ∈ Cu(G), it follows that there exists a ◦ κq K, α ⊗ idK) and the maps iA⊗K Cu-morphism µ κq : CuG(A, α) → CuG(A ⊗ K, α ⊗ idK) satisfying κq ◦iA Finally, since κq complete. ◦κq µ. Since κq is induced by ιq, we must have κq = CuG(ιq). µ = iA⊗K µ is an isomorphism for all µ, the same holds for κq, so the proof is (cid:3) µ Proposition 3.13. Let (An, ιn)n∈N be a direct system of C∗-algebras with con- necting maps ιn : An → An+1. For n ∈ N, let α(n) : G → Aut(An) be a continuous action, and suppose that α(n+1) ◦ ιn = ιn ◦ α(n) for all n ∈ N. Set A = lim (An, ιn), −→ and α = lim −→ α(n). Then there exists a natural CuG-isomorphism CuG(An, α(n)) ∼= CuG(A, α). lim −→ Proof. By functoriality of CuG (see Corollary 2.14), the equivariant inductive sys- tem (An, α(n), ιn)n∈N induces the inductive system (cid:16)CuG(An, α(n)), CuG(ιn)(cid:17)n∈N CuG(An, α(n)). in CuG. By Theorem 3.9, its inductive limit exists in CuG, and we will denote it by lim −→ For n ∈ N, denote by ι∞,n : An → A the equivariant map into the direct limit. Then CuG(ι∞,n) is a morphism in CuG, and the universal property of inductive limits, there exists a CuG-morphism ϕ : lim −→ CuG(An, α(n)) → CuG(A, α). We claim that ϕ is an isomorphism. For this, it is enough to check that it is an isomorphism in Cu. Fix a separable representation (Hµ, µ) of G, and for n ∈ N, denote by ψn,µ : Cu((K(Hµ) ⊗ An)G) → Cu((K(Hµ) ⊗ An+1)G) 20 EUSEBIO GARDELLA AND LUIS SANTIAGO the Cu-morphism induced by ιn. For ν ∈ Cu(G) with µ ≤ ν, we have jµ,ν ◦ ψn,ν = ψn,ν ◦ jµ,ν . Denote by ψn : lim −→ µ∈Cu(G) Cu((K(Hµ) ⊗ An)G) → lim −→ µ∈Cu(G) Cu((K(Hµ) ⊗ An+1)G) the resulting Cu-morphism, and regard it with a Cu-morphism ϕn : CuG(An, α(n)) → CuG(An+1, α(n+1)). It is clear that ϕn = CuG(ιn). Using Theorem 2 in [CEI08], the direct limit of Cu((K(Hµ)⊗An)G) is (naturally) isomorphic, in the category Cu, to Cu((K(Hµ) ⊗ A)G). This isomorphism can be identified with the map ϕ, and this shows that ϕ is an isomorphism in Cu. This finishes the proof. (cid:3) We point out that in the previous proposition, we may allow direct limits over arbitrary directed sets, using Corollary 3.1.11 in [APT14] instead of Theorem 2 in [CEI08]. 4. A Hilbert module picture of CuG(A, α) In analogy with the non-equivariant case, the equivariant Cuntz semigroup can be constructed in terms of equivariant Hilbert modules. The goal of this section is to present this construction and identify it with CuG(A, α) in a canonical way. The description of CuG(A, α) provided in this section will be needed in Section 5, where we will prove that CuG(A, α) can be naturally identified with Cu(A ⋊α G) (Theorem 5.3). 4.1. Equivariant Hilbert C*-modules. Throughout this section, we fix a C∗- algebra A, a compact group G, and an action α : G → Aut(A). All modules will be right modules and a Hilbert A-module will mean a Hilbert C*-module over A. The reader is referred to [Lan95] for the basics of Hilbert C*-modules. Given Hilbert A-modules E and F , we let L(E, F ) and K(E, F ) denote the spaces of adjointable operators and compact operators from E to F , respectively. We write U(E, F ) for the set of unitaries between E and F . When E = F , we write L(E), K(E), and U(E) for L(E, E), K(E, E), and U(E, E), respectively. Definition 4.1. A Hilbert (G, A, α)-module is a pair (E, ρ) consisting of (1) a Hilbert A-module E, and (2) a strongly continuous group homomorphism ρ : G → U(E), satisfying (a) ρg(x · a) = ρg(x) · αg(a) for all g ∈ G, all x ∈ E and all a ∈ A, and (b) hρg(x), ρg(y)iE = αg (hx, yiE) for all g ∈ G and all x, y ∈ E. (The continuity condition for ρ means that for x ∈ E, the map G → E given by g 7→ ρg(x) is continuous.) Definition 4.2. If (E, ρ) is a Hilbert (G, A, α)-module and F is a Hilbert submod- ule of E satisfying ρg(F ) ⊆ F for all g ∈ G, we will write ρF for the (co-restricted) group homomorphism ρF : G → U(F ) given by (ρF )g(z) = ρg(z) for all g ∈ G and all z ∈ F . The pair (F, ρF ) will be called a Hilbert (G, A, α)-submodule of (E, ρ). 21 We say that E is countably generated if there exists a countable subset {ξn}n∈N ⊆ E such that is dense in E. ( kXn=1 ξnan : an ∈ A, k ∈ N) We will sometimes call Hilbert (G, A, α)-modules G-Hilbert (A, α)-modules, or just G-Hilbert A-modules if the action α is understood. Example 4.3. It is easy to check that if β : G → Aut(B) is an action of G on a C∗-algebra B, then the pair (B, β) is a G-Hilbert B-module. Given G-Hilbert A-modules (E, ρ) and (F, η), we let L(E, F )G and K(E, F )G de- note the subsets of L(E, F ) and K(E, F ), respectively, consisting of the equivariant operators. That is, L(E, F )G = {T ∈ L(E, F ) : T ◦ ρg = ηg ◦ T for all g ∈ G}, K(E, F )G = {T ∈ K(E, F ) : T ◦ ρg = ηg ◦ T for all g ∈ G}. (Note that L(E, F )G is the set of fixed points of L(E, F ), where for an adjointable operator T : E → F and g ∈ G, we set g · T = ηg ◦ T ◦ ρg−1 .) As before, L(E)G and K(E)G denote L(E, E)G and K(E, E)G, respectively. Definition 4.4. Let (E, ρ) and (F, η) be G-Hilbert A-modules. We say that (E, ρ) is isomorphic to (F, η), in symbols (E, ρ) ∼= (F, η), if there exists a unitary in L(E, F )G. We say that (E, ρ) is subequivalent to (F, η), in symbols (E, ρ) (cid:22) (F, η), if (E, ρ) is isomorphic to a direct summand of (F, η). (That is, if there exists V ∈ L(E, F )G such that V ∗V = idE.) Let I be a set and let (Ej, ρj)j∈I be a family of Hilbert A-modules. Then the direct sum with respect to the norm defined by the scalar product Hilbert direct sum Lj∈I ρj! is the completion of the corresponding algebraic Ej,Lj∈I *Mj∈I ξj ,Mj∈I ζj+ =Xj∈I hξj, ζji. Let H be a Hilbert space. By convention, the scalar product on H is linear in the second argument and conjugate linear in the first one. Let H ⊗ A denote the exterior tensor product of H and A, where A is considered as a right A-module over itself ([Lan95, Chapter 4]). That is, H ⊗ A is the completion of the algebraic tensor product H ⊗alg A in the norm given by the A-valued product for ξ1, ξ2 ∈ H and a1, a2 ∈ A. hξ1 ⊗ a1, ξ2 ⊗ a2i = hξ1, ξ2ia∗ 1a2 Definition 4.5. For each element [π] ∈ bG, choose a representative π : G → U(Hπ). Denote by HC the Hilbert space direct sum HC = M[π]∈ bG ∞Mn=1 Hπ, 22 EUSEBIO GARDELLA AND LUIS SANTIAGO and let πC : G → U(HC) be the unitary representation given by πC = M[π]∈ bG ∞Mn=1 π. The unitary representation (HC, πC) is easily seen not to depend on the choices of representatives π : G → U(Hπ) up to unitary equivalence. We define the universal G-Hilbert (A, α)-module (HA, πA) to be HA = HC ⊗ A and πA = πC ⊗ α. Remark 4.6. It is a classical result of Kasparov that when G is second count- able, then every countably generated G-Hilbert A-module is isomorphic to a direct summand of (HA, πA); see [Kas80, Theorem 2]. Remark 4.7. It is easy to check, using the Peter-Weyl theorem, that (HC, πC) is (unitarily equivalent) to the representation (L2(G) ⊗ ℓ2(N), λ ⊗ idℓ2(N)). An equivalent presentation of HC (and therefore of HA), when G is second countable, is with πC = L[µ]∈Cu(G) HC = M[µ]∈Cu(G) Hµ, µ. Lemma 4.8. Suppose that G is second countable, and let (E, (πA)E) be a count- ably generated G-Hilbert A-submodule of (HA, πA). Then there exists a separable subrepresentation (Hµ, µ) of (HC, πC) such that E ⊆ Hµ⊗A and (πA)E = (µ⊗α)E. (Hν ⊗ A) for any ξ ∈ HA, there exists a countable set Proof. Since HA = L[ν]∈Cu(G) Xξ ⊆ Cu(G) such that ξ belongs to L[ν]∈Xξ Now let {ξn}n∈N be a countable generating subset of E. Then X = Sn∈N countable subset of Cu(G). Set (Hν ⊗ A). Xξn is a (Hµ, µ) = M[ν]∈X (Hν, ν) . Then Hµ is separable. It is immediate that E ⊆ Hµ ⊗ A and (πA)E = (µ ⊗ α)E, so the proof is complete. (cid:3) Let (E, ρ) be a G-Hilbert A-module. Then the action G on E induces an action of G on the C∗-algebra K(E) by conjugation. The fixed point algebra of this action will be denoted by K(E)G. When (E, ρ) is the G-Hilbert A-module (Hµ ⊗ A, µ⊗ α), for some separable unitary representation (Hµ, µ) of G, then the induced action on K(Hµ ⊗ A) will be denoted by Ad(µ ⊗ α). Let E be a Hilbert A-module, and let ξ, ζ ∈ E. We denote by Θξ,ζ : E → E the A-rank one operator given by Θξ,ζ(η) = ξ · hζ, ηi for η ∈ E. Lemma 4.9. Let a ∈ K(HA)G, and set E = span{a(HA) ∪ a∗(HA)}, endowed with the restricted G-representation ρ. Then (E, ρ) is a countably generated G-Hilbert A-module and aE ∈ K(E)G. Proof. It is clear that E is invariant under (πA)g for all g ∈ G; thus (E, ρ) is a G-Hilbert A-module. Let ε > 0. Since a ∈ K(HA)G, there exist k ∈ N, and ξ1, . . . , ξk, ζ1, . . . , ζk ∈ HA satisfying 23 Use that a ∈ (aa∗)K(HA) and a ∈ K(HA)(a∗a) to choose n ∈ N with such that, in addition, It follows that a − kXj=1 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Θ(aa∗)1/n(ξj ),(a∗a)1/n(ζj )(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) < a − 1 1 ε 8 . n < 2 kaa∗k kXj=1 n < 2 and ka∗ak Θξj,ζj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)a − (aa∗)1/na(a∗a)1/n(cid:13)(cid:13)(cid:13) < (aa∗)1/nΘξj ,ζj (a∗a)1/n(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) kXj=1 ≤(cid:13)(cid:13)(cid:13)a − (aa∗)1/na(a∗a)1/n(cid:13)(cid:13)(cid:13) Θξj ,ζj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) + kaa∗k1/n(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) kXj=1 ε 2 . < ε. a − a − ka∗ak1/n For j = 1, . . . , k, the map Θ leaves E invariant, so its re- (aa∗) 1 n (ξj ),(a∗a) 1 n (ζj ) striction to E is a rank one operator in K(E). It follows that aE is the limit of a sequence finite rank operators on E, and hence it is compact. In particular, E is countably generated, because the range of each finite rank operator is finitely generated. Finally, it is clear that a is invariant, so a ∈ K(E)G. (cid:3) Let (E, ρ) be a G-Hilbert A-module, and let (F, ρF ) be a G-Hilber submodule of E. Then there is a canonical inclusion equivariant ι : K(F ) → K(E), which is defined as follows. For ξ, η ∈ F , we denote by ΘF ξ,η ∈ K(F ) the corresponding rank- one operator, and likewise for ΘE ξ,η. One must check that ι extends to an injective homomorphism K(F ) → K(E), and we omit the straightforward verification. That ι is equivariant is also clear. ξ,η ∈ K(E). Then ι is given by ι(ΘF ξ,η) = ΘE In particular, in the above situation, ι restricts to a canonical embedding ι : K(F )G → K(E)G. This fact will be used in the proof of the following lemma. Lemma 4.10. Suppose that G is second countable. Let (E, ρ) be a countably generated G-Hilbert A-module, and let (F, ρF ) be a countably generated G-Hilbert submodule of E. Then there exists a ∈ K(E)G such that aF is strictly positive and belongs to K(F ), and F = a(F ) = a(E). Proof. Use [Lan95, Proposition 6.7] to choose a strictly positive element c ∈ K(F ). Using the normalized Haar measure on G, it is easy to check that the element b =RG(g · c) dg is strictly positive and G-invariant. Moreover, one has c(F ) = b(F ). Using strict positivity of b, choose a sequence (bn)n∈N in K(F ) such that lim n→∞ kbbn − b 1 n k = 0. 24 EUSEBIO GARDELLA AND LUIS SANTIAGO By [Lan95, Equation 1.5], we have lim n→∞ bbnξ = ξ for all ξ ∈ F . It follows that b(F ) = F . Denote by ι : K(F )G → K(E)G the canonical inclusion described in the comments before this lemma. Set a = ι(c). Then a is strictly positive and invariant, and a(E) = a(F ) is dense in F . This finishes the proof. (cid:3) 4.2. The Hilbert module picture of CuG(A, α). We now define the relevant equivalence and subequivalence relations of G-Hilbert modules that will give rise to a different description of the equivariant Cuntz semigroup. Definition 4.11. Let (E, ρ) be a G-Hilbert A-module, and let (F, η) be a G-Hilbert A-submodule. We say that (F, η) is G-compactly contained in (E, ρ), if there exists b ∈ K(E) whose restriction to F is idF . We claim that the operator b in the definition above can be taken to be con- tractive and to belong to the fixed point algebra K(E)G. Contractivity is easy to arrange: simply divide by kbk. To choose b in the fixed point algebra, first note that if ξ ∈ F , then (g · b)(ξ) = ρg(b(ρg−1 (ξ))) = ξ. With dg denoting the normalized Haar measure on G, it follows that b′ =RG is invariant and its restriction to F is the identity. (g ·b) dg Definition 4.12. Let (E, ρ) and (F, η) be G-Hilbert A-modules. We say that (E, ρ) is G-Cuntz subequivalent to (F, η), and denote this by (E, ρ) -G (F, η), if every compactly contained G-Hilbert submodule of (E, ρ) is unitarily equivalent to a G-Hilbert submodule of (F, η). We say that (E, ρ) is G-Cuntz equivalent to (F, η), and denote this by (E, ρ) ∼G (F, η), if (E, ρ) -G (F, η) and (F, µ) -G (E, ν). The G-Cuntz equivalence class of the G-Hilbert A-module (E, ρ) is denoted by [(E, ρ)]. We denote by CuG H(A, α) the set of Cuntz equivalence classes of G-Hilbert A- modules. It is easy to check that the direct sum of G-Hilbert A-modules induces a well H(A, α) with the partial order given by H(A, α) defined operation on CuG [(E, ρ)] ≤ [(F, η)] if (E, ρ) -G (F, η). With this structure, it is clear that CuG is a partially ordered abelian semigroup. H(A, α). Endow CuG We now define a Cu(G)-semimodule structure on CuG H(A, α) and [µ] ∈ Cu(G), we set CuG H(A, α). For [(E, ρ)] ∈ [µ] · [(E, ρ)] = [(Hµ ⊗ E, µ ⊗ ρ)]. Similarly, Cu(K(HA)G) has a natural Cu(G)-semimodule structure (see Definition 4.5 for the definition of HA). Let a ∈ K(HA)G be a positive element. For a separable unitary representation (Hµ, µ) of G, let sµ ∈ K(Hµ)G be a strictly positive element. Identify Hµ ⊗ HA with a submodule of HA using the product in Cu(G), and set Let (Hµ, µ) be a separable unitary representation of G. Then (Hµ, µ) is unitarily equivalent to a subrepresentation of (HC, πC), and hence there exists an operator [µ] · [a] = [sµ ⊗ a]. satisfying V ∗ µ Vµ = idHµ . Vµ = Vµ,πA ∈ L(Hµ ⊗ A, HA)G 25 Define a map χ : CuG(A, α) → Cu(K(HA)G) as follows. Given a separable uni- tary representation (Hµ, µ) of G, and given a positive element a ∈ (K(Hµ) ⊗ A)G, set χ([a]G) = [VµaV ∗ µ ]. Proposition 4.13. The map χ : CuG(A, α) → Cu(K(HA)G), described above, is well defined. Moreover, it is an isomorphism in CuG. Proof. We divide the proof into a number of claims. Claim: χ is well defined, and it preserves the order. Let (Hµ, µ) and (Hν, ν) be separable unitary representations of G, and let a ∈ (K(Hµ)⊗A)G and b ∈ (K(Hν )⊗ A)G satisfy a -G b. Then there exists a sequence (dn)n∈N in (K(Hν , Hµ) ⊗ A)G, ν belongs to K(HA)G, and such that lim n→∞ n − ak = 0. It follows that VµdnV ∗ kdnbd∗ (VµdnV ∗ µ → VµaV ∗ µ , µ - Vν bV ∗ in the norm of (K(Hµ) ⊗ A)G, as n → ∞. This shows that VµaV ∗ the claim is proved. ν )∗ = VµdnbdnV ∗ ν )(VµdnV ∗ ν )(Vν bV ∗ ν , and µ and b′ = Vν bV ∗ Claim: χ is an order embedding. Let (Hµ, µ) and (Hν, ν) be separable unitary representations of G, and let a ∈ (K(Hµ) ⊗ A)G and b ∈ (K(Hν ) ⊗ A)G satisfy χ([a]G) ≤ χ([b]G). Set a′ = VµaV ∗ ν . Then there exists a sequence n)n∈N in K(HA)G such that lim (d′ n→∞ For each n ∈ N, set En = span (d′ n)∗(HA)). Then En is a countably generated G-Hilbert A-module by Lemma 4.9. Use Lemma 4.8 to choose a sepa- rable subrepresentation (Hn, (πC)Hn ) of (HC, πC), satisfying E ⊆ Hn ⊗ A ⊆ HA as G-Hilbert A-modules. Let Wµ,πA ∈ B(Hµ) and Wν,πA ∈ B(Hν) be the partial isometries implementing the isomorphisms of Hµ and Hν with Hilbert subspaces H′ kd′ n(HA) ∪ (d′ n)∗ − a′k = 0. ν of HC, respectively. Set µ and H′ nb′(d′ H = span H′ µ ∪ H′ Hn! ⊆ HC. ν ∪ [n∈N Then H is separable and the operators a′, b′, d′ to itself. Moreover, the restrictions a′′, b′′, d′′ belong to K(H ⊗ A)G. Moreover, we have nb′′(d′′ kd′′ lim n→∞ n)∗ − a′′k = 0, n, and (d′ n of a′, b′, d′ n)∗, for n ∈ N, map H ⊗ A n, for n ∈ N, to H ⊗ A, and thus a′′ - b′′ in K(H ⊗ A)G. Consequently, a′′ -G b′′. Lemma 2.8 implies that a ∼G a′′ and b ∼G b′′. We conclude that a -G b, as desired. Claim: χ is surjective. Let a ∈ K(HA)G. By Lemma 4.8, there exists a subrep- resentation (Hµ, µ) of (HC, πC) such that a(HA) ⊆ Hµ ⊗A as G-Hilbert A-modules. Let a′ be the restriction of a to Hµ ⊗ A. It is then clear that χ([a′]G) = [a], so the claim is proved. It follows that χ is a Cu-isomorphism. Claim: χ is a Cu(G)-semimodule morphism (and hence a CuG-isomorphism). It is enough to check that χ preserves the Cu(G)-action. This is immediate from the definitions. (cid:3) For a ∈ K(HA)G, we denote by HA,a the G-Hilbert A-module a(HA), and we let πA,a be the compression of πA to a(HA). 26 EUSEBIO GARDELLA AND LUIS SANTIAGO Theorem 4.14. Suppose that G is second countable. Then the map τ : Cu(K(HA)G) → CuG H(A, α), defined by τ ([a]) = [(HA,a, πA,a)] for a positive element a ∈ K ⊗ (K(HA)G), is a well defined natural isomorphism in CuG. Proof. We divide the proof into a number of claims. Claim: τ is well defined and it preserves the partial order. To show this, it suffices to prove that if a, b ∈ K(HA)G are positive elements with a -G b, then (HA,a, πA,a) -G (HA,b, πA,b). (See Definition 4.4.) Let (E, (πA)E) be a countably generated G-Hilbert A-module which is com- pactly contained in (HA,a, πA,a). By Lemma 4.10, there exists c ∈ K (HA,a)G ∼=(cid:16)a(K(HC) ⊗ A)a(cid:17)G such that cE is strictly positive and c(HA,a) = E. Then K(E) ∼= c(K(HC) ⊗ A)c. Use the definition of compact containment of G-Hilbert A-modules (Definition 4.11) to choose d ∈(cid:16)a(K(HC) ⊗ A)a(cid:17)G with kdk = 1 and dc = c. In particular, we have (d + c − 1)+ = c and d + c -G a. Apply Proposition 2.6 with ε = 1 to the elements d+c -G b to find f ∈ K (HA,b, HA,a)G such that (d+c−1)+ = f bf ∗. Set x = b 2 f ∗, which is an element in K (HA,a, HA,b)G. Since (d + c − 1)+ = c, we have 1 x∗x = c and xx∗ ∈(cid:16)b(K(HC) ⊗ A)b(cid:17)G . Set F = x(E), and let y : E → F be the operator obtained from x by restricting its domain to E and its codomain to F . Since x is invariant, we have y ∈ L(E, F )G. It is clear that y has dense range. Moreover, y∗ = (x∗)F ∈ L(F, E)G, and hence y∗(F ) = y∗(x(E)) = y∗(x(E)) = x∗x(E) = c(E) = E. It follows that both y and y∗ have dense range. By [Lan95, Proposition 3.8], it follows that E and F are unitarily equivalent. Moreover, it can be seen from the proof of that proposition that the unitary can be chosen in LG(E, F ). This shows that (E, (πA)E) is G-equivalent to a submodule of (HA,b, πA,b), as desired. This proves the claim. Claim: τ is an order embedding. Let a, b ∈ K(HA)G satisfy (HA,a, πA,a) -G (HA,b, πA,b) . Let ε > 0 and let f ∈ C0(0, kak] be a function that is linear on [0, ε] and constant equal to 1 on [ε, kak]. Then f (a) belongs to (cid:16)aK(HA)a(cid:17)G ε)+ = (a − ε)+. It follows that (cid:0)HA,(a−ε)+ , πA,(a−ε)+(cid:1) is compactly contained in U : (cid:0)HA,(a−ε)+ , πA,(a−ε)+(cid:1) → (HA,b, πA,b) . Set x = (a − ε)+U ∗, which is an element in K(HA,(a−ε)+ , b(HA)). Then (HA,a, πA,a), so there exists an equivariant unitary and satisfies f (a)(a − (a − ε)+ = xx∗ and x∗x = U (a − ε)+U ∗ ∈ K(bHA)G. 27 It follows that lim 1 n x∗xb 1 have (a − ε)+ = xx∗ ∼G x∗x, it follows that (a − ε)+ -G b. Since ε > 0 is arbitrary, we conclude that n − x∗x(cid:13)(cid:13)(cid:13) = 0. Therefore x∗x -G b. Since we also n→∞(cid:13)(cid:13)(cid:13)b [a] = sup ε>0 [(a − ε)+] ≤ [b], and the claim is proved. Claim: τ is surjective. Let (E, (πA)E) be a countably generated G-Hilbert A- module. Since G is assumed to be second countable, (E, (πA)E) is isomorphic to a G-Hilbert submodule of (HA, πA), by [Kas80, Theorem 2]. Use Lemma 4.10 to find a ∈ K(HA)G such that (E, (πA)E) ∼= (HA,a, πA,a). It follows that τ ([a]) = [(HA,a, πA,a)] = [(E, (πA)E)], and the claim follows. We deduce that τ is an isomorphism in Cu. Claim: τ is a Cu(G)-semimodule morphism (and hence an isomorphism in CuG). We only check that τ preserves the Cu(G)-action. Let (Hµ, µ) be a separa- ble unitary representation of G, and let sµ be a strictly positive element in K(Hµ)G. For a ∈ K(HA)G, we have τ ([µ] · [a]) = τ ([sµ ⊗ a]) = [(sµ ⊗ a)(Hµ ⊗ HA)] = [Hµ ⊗ HA,a] = [µ] · [HA,a]. This concludes the proof of the claim and of the theorem. (cid:3) The following is the main result of this section. Corollary 4.15. Suppose that G is second countable. Then there exists a natural CuG-isomorphism δ : CuG(A, α) → CuG H(A, α). Proof. By Theorem 4.14 and Proposition 4.13, the map δ = τ ◦ χ is the desired CuG-isomorphism. (cid:3) 5. Julg's theorem and the Cu(G)-semimodule structure on Cu(A ⋊α G) The goal of this section is to prove that if α : G → Aut(A) is a continuous action of a compact group G on a C∗-algebra A, then there exists a natural Cu- isomorphism between its equivariant Cuntz semigroup CuG(A, α) and Cu(A ⋊α G); see Theorem 5.3. This isomorphism allows us to endow Cu(A⋊αG) with a canonical Cu(G)-semimodule structure, and we compute it explicitly in Theorem 5.14. When G is abelian, this semimodule structure is particularly easy to describe: it is given by the dual action of α; see Proposition 5.16. We will prove these results using the equivariant Hilbert module picture of CuG(A, α) studied in the previous section. 5.1. Julg's Theorem. For the rest of this section, we fix a compact group G, a C∗- algebra A, and a continuous action α : G → Aut(A). The goal of this section is to prove the Cuntz semigroup analog of Julg's theorem; see Theorem 5.3. Most of the work has already been done in the previous section, and the only missing ingredients are Remark 5.1, which is essentially the Peter-Weyl theorem, and Proposition 5.2, which is noncommutative duality. 28 EUSEBIO GARDELLA AND LUIS SANTIAGO Let L2(G) denote the Hilbert space of square integrable functions on G with respect to its normalized Haar measure, and let λ : G → U(L2(G)) denote the left regular representation. Remark 5.1. By the Peter-Weyl Theorem ([Fol95, Theorem 5.12]), the G-Hilbert module (HC, πC) is unitarily equivalent (see Definition 4.4) to Therefore there exists an equivariant ∗-isomorphism (ℓ2(N) ⊗ L2(G), idℓ2(N) ⊗ λ). (K(HA), πA) → (K(ℓ2(N) ⊗ L2(G) ⊗ A), Ad(idℓ2(N) ⊗ λ ⊗ α)). It follows that there exists a natural ∗-isomorphism θ : K(HA)G → K(ℓ2(N) ⊗ L2(G) ⊗ A)G. The following result is standard, and it is a consequence of Landstad's du- ality. See, for example, Theorem 2.7 in [KOQ15], and specifically Example 2.9 in [KOQ15]. (The result can also be derived from Katayama's duality; see Theo- rem 8 in [Kat85].) Proposition 5.2. Let G be a locally compact group, let B be a C∗-algebra, and let δ : B → M (B ⊗ C∗(G)) be a normal coaction. Denote by B ⋊δ G the corresponding cocrossed product, and bybδ : G → Aut(B ⋊δ G) the dual action. Then there exists a canonical ∗-isomorphism ψ : (B ⋊δ G) bδ → B, which is moreover (δjG , δ)-equivariant (see Definition 2.8 in [KOQ15]). The following result is an analog of Julg's Theorem ([Jul81]; see also [Phi87, Theorem 2.6.1]) for the equivariant Cuntz semigroup. It asserts that the equivari- ant Cuntz semigroup is naturally isomorphic, in the category Cu, to the Cuntz semigroup of the crossed product. The isomorphism is obtained as a composition of the isomorphisms from Proposition 4.13, Remark 5.1, Appendix 6 in [CEI08] (this is stability of the functor Cu), and Proposition 5.2; see the commutative dia- gram at the end of the proof of Theorem 5.3. In some applications of this theorem, particularly in Theorem 5.14, the explicit construction of the isomorphism will be relevant. Theorem 5.3. Let G be a compact group, let A be a C∗-algebra, and let α : G → Aut(A) be a continuous action. Then there exists a natural Cu-isomorphism σ : CuG(A, α) → Cu(A ⋊α G). Proof. Endow K(L2(G)) with the action of conjugation by the left regular repre- sentation of G, endow K(ℓ2(N)) with the trivial G-action, and endow K(L2(G)) ⊗ A and K(ℓ2(N)) ⊗ K(L2(G)) ⊗ A with the corresponding tensor product actions. Then there exists a natural identification (K(ℓ2(N)) ⊗ K(L2(G)) ⊗ A)G = K(ℓ2(N)) ⊗ (K(L2(G)) ⊗ A)G. Since Cu is a stable functor (see Appendix 6 in [CEI08]), there exists a natural Cu-isomorphism κ : Cu(K(ℓ2(N) ⊗ L2(G) ⊗ A)G) → Cu(K(L2(G) ⊗ A)G). 29 By Remark 5.1, there exists a natural ∗-isomorphism θ : K(HA)G → K(ℓ2(N) ⊗ L2(G) ⊗ A)G. Denote by ψ : (K(L2(G)) ⊗ A)G → A ⋊α G the natural ∗-isomorphism obtained from Proposition 5.2 for B = A ⋊α G and δ = bα. (Recall that coactions of com- pact groups are automatically normal; see, for example, the end of the proof of Lemma 4.8 in [Gar15a].) With χ : CuG(A, α) → Cu(K(HA)G) denoting the natural Cu-isomorphism given by Proposition 4.13, define σ to be the following composi- tion: CuG(A, α) ❖ χ / Cu(K(HA)G) Cu(θ) / Cu((K(ℓ2(N)) ⊗ K(L2(G)) ⊗ A)G) ❖ ❖ σ ❖ '❖ ❖ Cu(A ⋊α G) κ Cu((K(L2(G)) ⊗ A)G). Cu(ψ) It is clear that σ is a natural isomorphism in the category Cu. (cid:3) 5.2. Semimodule structure on the crossed product. Theorem 5.3 provides an isomorphism CuG(A, α) ∼= Cu(A ⋊α G) as Cu-semigroups. We can give Cu(A ⋊α G) the unique Cu(G)-semimodule structure that makes this isomorphism into a CuG-isomorphism. To make this result useful, we must describe this semimodule structure internally. This takes some work, and we will need a series of intermediate results. This subsection is based, to some extent, on [Phi87]. We assume that G is a second countable group. The main technical difficulties are the absence of short exact sequences in the context of semigroups, and the fact that in the construction of the equivariant Cuntz semigroup, representations of the group G on infinite dimensional Hilbert spaces are allowed. Compactness of G is crucial in overcoming the latter. We need a standard definition. For a C∗-algebra A, we denote by M (A) its multiplier algebra. Definition 5.4. Let α, β : G → Aut(A) be continuous actions of a locally compact group G on a C∗-algebra A. We say that α and β are cocycle equivalent, if there exists a function ω : G → U(M (A)) satisfying: (1) ωgh = ωgβg(ωh) for all g, h ∈ G; (2) αg = Ad(ωg) ◦ βg for all g ∈ G; (3) For a ∈ A, the map G → A given by g 7→ ωga is continuous. Condition (1) in the definition above ensures that if α is defined by (2), then αg ◦ αh = αgh for all g, h ∈ G. Remark 5.5. It is well known that cocycle equivalent actions have isomorphic associated crossed products. Nevertheless, it does not follow from this that cocycle equivalent actions have isomorphic equivariant Cuntz semigroups, because we do not know how to compute the Cu(G)-semimodule structure of the crossed products. (That this is indeed the case is a consequence of Theorem 5.14.) In order to prove the result, however, we do need to know that some specific cocycle conjugate actions yield isomorphic equivariant Cuntz semigroups; see Proposition 5.6. For the rest of the subsection, we fix a continuous action α : G → Aut(A) of a compact group G on a C∗-algebra A. ' / /   o o 30 EUSEBIO GARDELLA AND LUIS SANTIAGO Proposition 5.6. Let β be an action of G on A which is cocycle equivalent to α. Suppose that A has an increasing countable approximate identity consisting of projections which are invariant for both α and β. Then there exists a natural CuG-isomorphism ε : CuG(A, α) → CuG(A, β). When A is unital, the map ε can be described as follows. Choose a cocycle ω : G → U(A) such that αg = Ad(ωg) ◦ βg for all g ∈ G. For a countably generated G-Hilbert (A, α)-module (E, ρ), the map ε is given by ε([(E, ρ)]) = ([E, ρω)], where ρω : G → U(E) is given by ρω g (x) = ρg(x)ωg for all g ∈ G and all x in E. Proof. Suppose first that A is unital. We must first argue why the map ε described in the statement is well-defined, and why it is an isomorphism in CuG. We fix a cocycle ω : G → U(A) such that αg = Ad(ωg) ◦ βg for all g ∈ G, and we fix a countably generated G-Hilbert (A, α)-module (E, ρ). Let ρω : G → U(E) be defined as in the statement. We claim that (E, ρω) is a countably generated G-Hilbert (A, β)-module. Since E was chosen to be countably generated to begin with, we shall only check that ρω is compatible with β and with the Hilbert module structure. Given a ∈ A, x ∈ E and g ∈ G, we have g (xa) = ρg(xa)ωg = ρg(x)αg(a)ωg = ρg(x)ωgβg(a) = ρω ρω g (x)βg(a), as desired. Moreover, for g ∈ G and x, y ∈ E, we have hρω g (x), ρω g (y)iE = hρg(x)ωg, ρg(y)ωgiE g hρg(x), ρg(y)iEωg = ω∗ = (Ad(ω∗ = βg(hx, yiE), g ) ◦ αg)(hx, yiE) thus proving the claim. The assignment (E, ρ) 7→ (E, ρω) is clearly surjective, and hence every G-Hilbert (A, β)-module has the form (E, ρω) for some G-Hilbert (A, α)-module (E, ρ). We claim that the assignment (E, ρ) 7→ (E, ρω) preserves G-Cuntz subequiva- lence. Let (E, ρ) and (E′, ρ′) be countably generated G-Hilbert (A, α)-modules, and suppose that (E, ρ) -G (E′, ρ′). If (F, ηω) is a G-Hilbert (A, β)-module that is compactly contained in (E, ρω), then it is straightforward to check that (F, η) If (F ′, η′) is a G-Hilbert (A, α)-module com- is compactly contained in (E, ρ). pactly contained in (E′, ρ′) such that (F ′, η′) ∼= (F, η), then one readily checks that (F ′, (η′)ω) ∼= (F, ηω). This shows that (E, ρω) -G (E′, (ρ′)ω), and proves the claim. Denote by ϕ : CuG(A, α) → CuG(A, β) the map given by [(E, ρ)] 7→ [(E, ρω)], g (x) = ρg(x)ωg for all g ∈ G and all x in E. We claim that where ρω is given by ρω ϕ is a CuG-morphism. We already showed that ϕ preserves the order. It is also easy to show that is preserves compact containment and suprema of increasing sequences. We show that it is a morphism of Cu(G)-semimodules. Given a separable unitary representation (Hµ, µ) of G, we must show that the diagram CuG(A, α) [µ]· CuG(A, α) ϕ ϕ CuG(A, β) [µ]· / CuG(A, β) / /     / 31 commutes. Given a countably generated G-Hilbert (A, α)-module (E, ρ), we have [(Hµ, µ)] · ϕ([(E, ρ)]) = [(E ⊗ Hµ, ρω ⊗ µ)] . On the other hand, the element ϕ ([(Hµ, µ)] · [(E, ρ)]) is represented by the G- Hilbert (A, β)-module (E ⊗ Hµ, (ρ ⊗ µ)ω), so it is enough to check that both actions on Hµ ⊗ E agree. Let g ∈ G, let x ∈ E and let ξ ∈ Hµ. Then ((ρ ⊗ µ)ω)g (x ⊗ ξ) = (ρg(x) ⊗ µg(ξ))ωg = (ρg(x)ωg) ⊗ µg(ξ) = ρω = (ρω ⊗ µ)g(x ⊗ ξ), g (x) ⊗ µg(ξ) thus showing that ϕ is a CuG-morphism. Since ϕ is clearly bijective, it follows that it is an isomorphism. Naturality is also clear. This proves the unital case. For the general case, let (en)n∈N be an increasing approximate identity in A consisting of projections that are invariant for both α and β. For n ∈ N, let α(n) : G → Aut(enAen) be the action given by α(n) (a) = αg(a) for all g ∈ G and a ∈ enAen, and similarly for β(n) : G → Aut(enAen). We claim that α(n) and β(n) are cocycle equivalent for all n ∈ N. g Choose a cocycle ω : G → U(M (A)) as in Definition 5.4. For g ∈ G and n ∈ N, one checks that ωgenω∗ g = (Ad(ωg) ◦ αg)(en) = βg(en) = en. Define ω(n) : G → U(enAen) by ω(n) that Ad(ω(n) also easy to verify, so the claim is proved. g = α(n) ) ◦ β(n) g g g = enωgen for g ∈ G. One readily checks for all n ∈ N and all g ∈ G. The cocycle condition is Note that there exist natural equivariant ∗-isomorphisms (A, α) = lim −→ (enAen, α(n)) and (A, β) = lim −→ (enAen, β(n)). By Proposition 3.13, there exist natural CuG-isomorphisms CuG(enAen, α(n)) and CuG(A, β) ∼= lim −→ CuG(A, α) ∼= lim −→ CuG(enAen, β(n)). For n ∈ N, denote by ϕ(n) : CuG(enAen, α(n)) → CuG(enAen, β(n)) the natural CuG-isomorphism provided by the unital case of this proposition. Naturality im- plies that the following diagram in CuG is commutative, where the top row is exact by Theorem 5 in [CRS10]: CuG(e1Ae1, α(1)) CuG(e2Ae2, α(2)) · · · / CuG(A, α) ϕ(1) ϕ(2) ϕ CuG(e1Ae1, β(1)) / CuG(e2Ae2, β(2)) / · · · / CuG(A, β). The universal property of the inductive limit in CuG shows that there exists a natural CuG-morphism ϕ : CuG(A, α) → CuG(A, β). This map is easily seen to be an isomorphism in CuG, so the proof is complete. (cid:3) The following result states that the equivariant Cuntz semigroup is a stable functor, as long as the algebra of compact operators is given an inner action. The / /   / /   /   ✤ ✤ ✤ / / / 32 EUSEBIO GARDELLA AND LUIS SANTIAGO case of the trivial action was already prove in Proposition 3.12. The result fails in general if the action on the compacts is not inner; see Example 5.8. Corollary 5.7. Let A be a unital C∗-algebra, and let (Hµ, µ) be a separable unitary representation of G. Let q be any rank one projection on Hµ, and consider the map ιq : A → A ⊗ K(Hµ), given by ιq(a) = a ⊗ q for a ∈ A. Let ε : CuG(A ⊗ K(Hµ), α ⊗ Ad(µ)) → CuG(A ⊗ K(Hµ), α ⊗ idK(Hµ)) be the natural CuG-isomorphism given by Proposition 5.6. Then ε−1 ◦ CuG(ιq) : CuG(A, α) → CuG(A ⊗ K(Hµ), α ⊗ Ad(µ)) is a natural CuG-isomorphism. Proof. We must first argue why the assumptions of Proposition 5.6 are met, so that the map ε really does exist. Firxt, since µ can be decomposed as a direct sum of finite dimensional representations by the Peter-Weyl Theorem, it follows that A ⊗ K(Hµ) has a countable approximate identity consisting of projections that are invariant under both α⊗Ad(µ) and α⊗idK(Hµ). And second, the actions α⊗Ad(µ) and α ⊗ idK(Hµ) are cocycle equivalent via the cocycle ω : G → U(A ⊗ K(Hµ)) given by ωg = 1A ⊗ µg for all g ∈ G. We conclude that Proposition 5.6 applies and that ε is a natural CuG-isomorphism. By Proposition 3.12, the map CuG(ιq) : CuG(A, α) → CuG(A ⊗ K(Hµ), α ⊗ idK(Hµ)) is natural CuG-isomorphism. It follows that ε−1 ◦CuG(ιq) is a natural isomorphism in CuG. (cid:3) As mentioned before, the result above fails if the action on K(H) is not inner, as the following example shows. We write Z2 = {−1, 1}. 0 Example 5.8. Define an action of G = Z2 × Z2 on A = M2 by letting (−1, 1) act by conjugation by (cid:18) 1 (cid:18) 0 0 −1 (cid:19) , and by letting (1, −1) act by conjugation by 0 (cid:19). Denote this action by α. (These unitaries commute up to a sign, so conjugation by them really does define an action of Z2 × Z2.) We claim that there is no isomorphism between CuG(M2, α) and CuG(M2, idM2 ). For this, it is enough to note that since there is an isomorphism 1 1 M2 ⋊idM2 (Z2 × Z2) ∼= M2 ⊕ M2 ⊕ M2 ⊕ M2, the semigroup CuG(M2, idM2 ) is isomorphic to Z≥0. On the other hand, one 4Lj=1 can compute CuG(M2, α) with elementary methods. For instance, one easily checks that M Z2×Z2 is simply the scalar matrices, and a similar computation shows that 2 (M2 ⊗ K(ℓ2(Z2 × Z2)))α⊗Ad(λ) ∼= M4. Thus, CuG(M2, α) = Z≥0, and the claim follows. We conclude that the analog of Corollary 5.7, where K(H) does not have an inner action, fails in general. 33 Denote by V (G) ⊆ Cu(G) the subsemigroup consisting of equivalence classes of finite dimensional unitary representations of G (so that the Grothendieck group of V (G) is the represenation ring R(G)). The following proposition describes the action of V (G) on Cu(A ⋊α G). Upon taking the supremum of an increasing sequence of finite dimensional representations, this result also leads to a method for computing the action of an arbitrary element in Cu(G). For use in its proof, we recall from the proof of Proposition 5.6 that for A unital and cocycle equivalent actions α and αω of G on A, the CuG-isomorphism ε : CuG(A, α) → CuG(A, αω) was constructed using G-Hilbert modules, and ε([(E, ρ)]) = [(E, ρω)], where the unitary representation ρω : G → U(E) is given by ρω g = ρgωg for g ∈ G. Proposition 5.9. Suppose that A is unital and let (Hµ, µ) be a separable unitary representation of G. Let p, q ∈ K(Hµ) be a µ-invariant projections, with q of rank one. Define equivariant ∗-homomorphisms ϕ : (A, α) → (A ⊗ K(Hµ), α ⊗ Ad(µ)) and ψ : (A, α) → (A ⊗ K(Hµ), α ⊗ idK(Hµ)) by ϕ(a) = a ⊗ p and ψ(a) = a ⊗ q for all a ∈ A. Let ε : CuG(A ⊗ K(Hµ), α ⊗ Ad(µ)) → CuG(A ⊗ K(Hµ), α ⊗ idK(Hµ)) be the cocycle equivalence isomorphism given by Proposition 5.6, associated to the cocycle equivalence g 7→ 1A ⊗ µg between α ⊗ Ad(µ) and α ⊗ idK(Hmu). Then the following CuG-diagram commutes: CuG(A, α) CuG(ϕ) [µpH]· CuG(A, α) CuG(ψ) CuG(A ⊗ K(Hµ), α ⊗ Ad(µ)) ε / CuG(A ⊗ K(Hµ), α ⊗ idK(Hµ)). Proof. Observe that CuG(ψ) is a natural isomorphism in CuG by Proposition 3.12. We only need to check that CuG(ψ)−1 ◦ ε ◦ CuG(ϕ) : CuG(A, α) → CuG(A, α) is multiplication by [µpHµ ]. Denote by (HA, πA) the canonical countably generated G-Hilbert (A, α)-module from Definition 4.5. Let (E, ρ) be a countably generated G-Hilbert (A, α)-module. Use Kasparov's absorption theorem (Theorem 2 in [Kas80]) to choose a πA-invariant projection r in L(HA)G such that (E, ρ) ∼= (rHA, πArHA). Then CuG(ϕ)([(E, ρ)]) =(cid:2)(cid:0)(r ⊗ p) (K(HA) ⊗ K(Hµ)) , πArHA ⊗ µpHµ(cid:1)(cid:3) = [E ⊗ pK(Hµ), ρ ⊗ µpHµ ]. g ∈ G and all x ∈ pK(Hµ). By the computation above, we have Denote the unitary representation Ad(µ)µ (see the comments before this propo- g)µg = µgx for all sition) of G on pK(Hµ) by eµ. In other words, eµg(x) = (µgxµ∗ (ε ◦ CuG(ϕ))([E, ρ]) = [(E ⊗ pK(Hµ), ρ ⊗eµ)]. We must compare the class of (E ⊗ pK(Hµ), ρ ⊗eµ) with the class of CuG(ψ)([E ⊗ pK(Hµ), ρ ⊗ Ad(µ)]), and show that they agree. / /     / 34 EUSEBIO GARDELLA AND LUIS SANTIAGO One checks that CuG(ψ)(cid:0)[(pHµ, µpHµ )] · [(E, ρ)](cid:1) is represented by (cid:0)E ⊗ pHµ ⊗ qK(Hµ), ρ ⊗ Ad(µ) ⊗ idK(Hµ)(cid:1) . Evaluating the maps in the diagram in the statement at (E, ρ), we get (E, ρ) ✤ ❴ (E ⊗ pHµ, ρ ⊗ µ) ❴ (E ⊗ pHµ ⊗ qK(Hµ), ρ ⊗ Ad(µ) ⊗ idK) (E ⊗ pK(Hµ), ρ ⊗ Ad(µ)) ✤ / (E ⊗ pK(Hµ), ρ ⊗eµ) . It is therefore enough to check that (cid:0)pHµ ⊗ qK(Hµ), µpHµ ⊗ idqK(Hµ)(cid:1) ∼= (pK(Hµ),eµ) as G-Hilbert K(Hµ)-modules. Fix a unit vector ξ(0) ∈ Hµ in the range of q and define σ : pHµ ⊗ qK(Hµ) → pK(Hµ) by σ(ξ ⊗b)(η) = hb∗(ξ(0)), ηiξ for all ξ ∈ pHµ, for all b ∈ qK(Hµ) and for all η ∈ Hµ, and extended linearly and continuously. Note that (p ◦ (σ(ξ ⊗ b))) (η) = hb∗(ξ(0)), ηip(ξ) = hb∗(ξ(0)), ηiξ, so the range of σ is really contained in pK(Hµ). Claim: σ is injective. Since p is a projection in K(Hµ), it has finite rank. For m = dim(pHµ), choose an orthonormal basis ξ1, . . . , ξm of pHµ. Given x ∈ mPj=1 pHµ ⊗ qK(Hµ), there exist b1, . . . , bm ∈ K(Hµ) such that x = ξj ⊗ bj. Assume that σ(x) = 0. By orthogonality of the vectors ξj, it follows that σ(ξj ⊗ bj) = 0 for j (ξ(0)) = 0. This all j = 1, . . . , m. Thus hη, b∗ j vanishes on span{ξ(0)} = qHµ, and so bj = 0 for j = 1, . . . , m. We shows that b∗ conclude that x = 0 and σ is injective. j (ξ(0))i = 0 for all η ∈ Hµ, and hence b∗ Claim: σ is surjective. Given a in K(Hµ), the map pa : Hµ → pHµ is map and has finite rank. For j = 1, . . . , m, denote by pj : Hµ → C × ξj the orthogonal projection. It follows from the Riesz Representation Theorem that there exist unit vectors ξ(0) in the range of pj and linear maps cj ∈ qjK(Hµ), for j = 1, . . . , m, such that j mXj=1 mXj=1 pa(η) = hc∗ j (ξ(0) j ), ηiξj for all η ∈ Hµ. (Recall our convention that inner products are linear on the second variable.) Since any two rank one projections are unitarily equivalent, it follows that there exist linear maps b1, . . . , bm ∈ qK(Hµ) such that pa(η) = hb∗ j (ξ(0)), ηiξj and thus pa = σ mPj=1 ξj ⊗ bj!, showing that σ is surjective. / /     / Claim: σ is a G-Hilbert K(Hµ)-homomorphism. Let ξ ∈ pHµ, b ∈ qK(Hµ), c ∈ K(Hµ), and η ∈ Hµ. Then 35 σ(ξ ⊗ b)(cη) = hc∗b∗(ξ(0)), ηiξ = h(bc)∗(ξ(0)), ηiξ = σ(ξ ⊗ bc)(η). Finally, for g ∈ G, b ∈ qK(Hµ), c ∈ K(Hµ), ξ ∈ pHµ, and η ∈ Hµ, one has σ(cid:0)(µpHµ ⊗ idqK(Hµ))g(ξ ⊗ b)(cid:1) = σ (µgξ ⊗ b) (η) = hb∗(ξ(0)), ηiµgξ = µghb∗(ξ(0)), ηiξ which shows that σ is equivariant. This finishes the proof. =eµg(σ(ξ ⊗ b))(η), (cid:3) The above result leads to a method for computing the Cu(G)-semimodule struc- ture on Cu(A ⋊α G). This description makes essential use of the cocycle equivalence isomorphism ε, and similarly to what happens with equivariant K-theory, it is in- convenient when trying to use it. To remedy this, we give an alternative description of the Cu(G)-action (Definition 5.10), which, even though it is not as transparent as the one in Proposition 5.9, has the advantage that all Cu-maps involved are induced by ∗-homomorphisms. Let (Hµ, µ) be a finite dimensional unitary representation of G, and denote by µ+ : G → U(Hµ⊕C) its direct sum with the trivial representation on C. Let pµ, qµ ∈ K(Hµ ⊕ C) be the projections onto Hµ and C, respectively. Define equivariant ∗- homomorphism ιpµ , ιqµ : A → A ⊗ K(Hµ ⊕ C) by ιpµ (a) = a ⊗ pµ and ιqµ (a) = a ⊗ pC for a ∈ A. Denote by cιpµ and cιqµ the corresponding maps on the crossed products by G. The Cu(ιqµ ) is invertible by Proposition 3.12, and it corresponds to multipli- cation by the class of the trivial representation by Proposition 5.9. By considering these maps at the level of the Cuntz semigroups, we have Cu(A ⋊α G) Cu(dιpµ ) / Cu(cid:0)(A ⊗ K(Hµ ⊕ C)) ⋊α⊗Ad(µ+) G(cid:1) Cu(dιqµ )−1 Cu(dιqµ ) Cu(A ⋊α G). Definition 5.10. Adopt the notation from the discussion above. We define a Cu(G)-semimodule structure on Cu(A ⋊α G) as follows. For a finite dimensional representation (Hµ, µ), and for s ∈ Cu(A ⋊α G), we set For an arbitrary separable unitary representation (Hν , ν), use compactness of G to choose irreducible representations (Hµn , µn) of G such that ν ∼= µn. For [µ] · s =(cid:0)Cu(cιqµ )−1 ◦ Cu(cιpµ )(cid:1) (s). m ∈ N, set νm = mLn=1 µn. For s ∈ Cu(A ⋊α G), we set [ν] · s = sup m∈N ([νm] · s) . ∞Ln=1 / / / o o 36 EUSEBIO GARDELLA AND LUIS SANTIAGO The following lemma shows that the above Cu(G)-semimodule structure is well- defined. Lemma 5.11. Let (Hν, ν) be a separable unitary representation of G, and find finite dimensional unitary representations (Hµn , µn) of G as in Definition 5.10. For m ∈ N, set νm = µn. Let s ∈ Cu(A ⋊α G). (1) The sequence ([νm] · s)n∈N is increasing in Cu(A ⋊α G). (2) The element [ν] · s = sup m∈N ([νm] · s) is independent of the decomposition mLn=1 ν ∼= Ln∈N µn. Proof. We only prove (1), since (2) follows the same idea. It suffices to show that if µ and ν are finite dimensional representations, then [µ ⊕ ν] · s ≤ [µ] · s for all s ∈ Cu(A ⋊α G). With the notation from the definition above, observe that the maps ϕµ, ϕν : A → A ⊗ K(Hµ ⊕ Hν ⊕ C) are orthogonal. Since orthogonal equivariant homomorphisms induce homomor- phisms between the respective crossed products which are also orthogonal, we de- duce that By evaluating the above identity at s, and using Definition 5.10, one deduces that [µ ⊕ ν] · s ≤ [µ] · s, as desired. (cid:3) Cu(bϕµ⊕ν ) = Cu(bϕµ) + Cu(bϕν). Lemma 5.12. The Cu(G)-semimodule structure on Cu(A ⋊α G) described above is compatible with taking suprema in Cu(G). Proof. Let (Hµn , µn)n∈N be a sequence of separable unitary representations of G such that ([µn])n∈N is increasing in Cu(G). Set [µ] = sup [µn]. Without loss of n∈N generality, we can assume that µn is a subrepresentation of µn+1 for all n ∈ N. In Hµn with Hµn ⊆ Hµn+1 for all n ∈ N. particular, we may assume that Hµ = Sn∈N It follows that µ+ = sup n∈N (µ+ n ) as representations of G on Hµ ⊕ C. Thus, for a ∈ A, we have (a ⊗ pHµn ) = sup ιpµ (a) = a ⊗ pHµ = sup n∈N n∈N(cid:0)ϕHµn (a)(cid:1) . Finally, since Cu(cιqµ ) is an isomorphism in Cu, we conclude that )(s)(cid:17) )(s)(cid:17)(cid:19) n∈N(cid:16)Cu(cιqµ )−1 ◦ Cu(ϕHµn = Cu(cιqµ )−1(cid:18)sup n∈N(cid:16)Cu(ϕHµn = Cu(cιqµ )−1 ◦ Cu(ϕHµ)(s) ([µn] · s) = sup = [µ] · s, sup n∈N for all s ∈ Cu(A ⋊α G), as desired. For later use, we record here the following fact. (cid:3) 37 Proposition 5.13. Let 0 / I ι / A π / B / 0 be an exact sequence of G-C∗-algebras, with actions γ : G → Aut(I) and β : G → Aut(B). Then CuG(I, γ) can be naturally identified with ker(CuG(π)), which by definition is ker(CuG(π)) = {s ∈ CuG(A, α) : CuG(π)(s) = 0} ⊆ CuG(A, α). Proof. Observe that ker(CuG(π)) only depends on the structure of CuG(A, α) as a semigroup, and is independent of the action of Cu(G). Denote by bπ : A ⋊α G → B ⋊β G the map induced by π. Using the natural isomorphisms from Theorem 5.3 for CuG(I, γ), CuG(A, α) and CuG(B, β), it is enough to show that Cu(I ⋊γ G) can be naturally identified with the subsemigroup of Cu(A ⋊α G). This follows immediately from Theorem 5 in [CRS10], so the proof is finished. (cid:3) {s ∈ Cu(A ⋊α G) : Cu(bπ)(s) = 0} We have now arrived at the main result of this section. Theorem 5.14. Let G be a compact group, let A be a C∗-algebra, and let α : G → Aut(A) be a continuous action. Then there exists a natural CuG-isomorphism CuG(A, α) ∼= Cu(A ⋊α G), where the Cu(G)-semimodule structure on Cu(A ⋊α G) is given by Definition 5.10. Proof. Assume first that A is unital. Let (Hµ, µ) be a finite dimensional unitary rality of the isomorphism in Theorem 5.3, there exists a commutative diagram representation of G, and let ιpµ , ιqµ ,cιpµ and cιqµ be as in Definition 5.10. By natu- CuG(ιpµ ) CuG(A, α) / CuG (A ⊗ K(Hµ ⊕ C), α ⊗ Ad(µ+)) CuG(A, α) CuG(ιqµ ) σA Cu(A ⋊α G) Cu(dιpµ ) / Cu(cid:0)(A ⊗ K(Hµ ⊕ C)) ⋊α⊗Ad(µ+) G(cid:1) σA Cu(A ⋊α G), Cu(dιqµ ) where all vertical arrows are the isomorphisms given by Theorem 5.3. By Proposition 5.9, the map CuG(ιqµ ) corresponds to multiplication by the class of the trivial repre- sentation in the Cu(G)-semimodule CuG (A ⊗ K(Hµ ⊕ C), α ⊗ Ad(µ+)). It follows that CuG(ιqµ ) is invertible. Thus Cu(cιqµ ) is also invertible, since the vertical ar- rows are invertible. By definition, Cu(cιqµ )−1 ◦ Cu(cιpµ ) is multiplication by [µ] on Cu(A⋊αG), and Proposition 5.9 shows that the composition CuG(ιqµ )−1◦CuG(ιpµ ) agrees with multiplication by [µ] on CuG(A, α). Commutativity of the diagram im- plies that the isomorphism σA : CuG(A, α) → Cu(A ⋊α G), which appears both in the left and right vertical arrows, commutes with multiplication by [µ]. Assume now that (Hν , ν) is a separable unitary representation of G. Since G is compact, it follows that K(Hν ) has an increasing approximate identity (en)n∈N consisting of G-invariant projections. For n ∈ N, denote by µn : G → U(enHν) the restriction of ν. It follows that [ν] = sup n∈N [µn] / / / /   /   o o   / o o 38 EUSEBIO GARDELLA AND LUIS SANTIAGO in Cu(G). Since the Cu(G)-semimodule structure on Cu(A ⋊α G) described above is compatible with taking suprema in Cu(G) by Lemma 5.12, it follows that the isomorphism σA : CuG(A, α) → Cu(A ⋊α G) commutes with multiplication by [ν], since it commutes with multiplication by [µn] for all n ∈ N by the above paragraph. This shows that σA is a Cu(G)-semimodule homomorphism. Now suppose that A is non-unital, and denote by eA its unitization. Define an extension eα : G → Aut(eA) of α to eA by setting eαg(a + λ1) = αg(a) + λ1 for all a ∈ A and all λ ∈ C. The short exact sequence of G-C∗-algebras induces the short exact sequence of crossed products 0 → A → eA → C → 0 (Recall that C ⋊id G ∼= C∗(G) for a locally compact group G.) 0 → A ⋊α G → eA ⋊ eα G → C∗(G) → 0. Denote by σA : CuG(A, α) → Cu(A⋊αG), by σ eA : CuG(eA,eα) → Cu(eA⋊ eαG), and by σC : CuG(C) → Cu(C∗(G)), the natural isomorphisms given by Theorem 5.3. Then the following diagram in Cu is commutative: Cu(A ⋊α G) σA CuG(A, α) CuG(bπ) Cu(C∗(G)) σC / CuG(C, idC). CuG(bπ) σ eA Cu(eA ⋊ eα G) / CuG(eA,eα) The maps σ eA and σC are Cu(G)-semimodule homomorphisms by the unital case. σ bA preserves the Cu(G)-action, so does σA. This finishes the proof. By Proposition 5.13, it follows that CuG(A, α) is the kernel of CuG(bπ). By com- mutativity of the diagram, σA is the restriction of σ bA to ker(CuG(bπ)), and since group, and let bG be its dual. If A is a C∗-algebra, then we write Cu(G) ⊗ Cu(A) We illustrate these methods by computing an easy example. Let G be a compact for the Cu(G)-semimodule Cu(G) ⊗ Cu(A) = {f : bG → Cu(A) : f has countable support}, with pointwise addition and partial order. The Cu(G)-action on Cu(G) ⊗ Cu(A) can be described as follows. Given [µ] ∈ Cu(G) and [π] ∈ bG, let mπ(µ) ∈ Z≥0 be the multiplicity of π in µ. Then (cid:3) [µ] = X[π]∈ bG mπ(µ) · [π]. For f ∈ Cu(G) ⊗ Cu(A), we set ([µ] · f )([π]) = mπ(µ)f ([π]) for π ∈ bG. that The tensor product notation is justified because of the following. One can check Cu(G) ∼= {f : bG → N : f has countable support}, / /   / /     / / with pointwise operations and partial order. Moreover, it is easy to check that Cu(G) ⊗ Cu(A) really is the tensor product in the category Cu of the semiring Cu(G) and the semigroup Cu(A), in the sense of Theorem 6.3.3 in [APT14]. 39 Proposition 5.15. Suppose that G acts trivially on A. Then CuG(A, idA) ∼= Cu(G) ⊗ Cu(A). Proof. Since G acts trivially on A, we have A ⋊α G ∼= A ⊗ C∗(G) canonically. For [π] ∈ bG, denote by dπ the dimension of π. Then C∗(G) ∼= L[π]∈ bG Mdπ , so Mdπ (A). For [τ ] ∈ bG, let A ⋊α G ∼= M[π]∈ bG ρτ : M[π]∈ bG Mdπ (A) → Mdτ (A) be the corresponding surjective ∗-homomorphism. We define a map ϕ : Cu(A ⋊α G) → Cu(G) ⊗ Cu(A) as follows. For a positive element A in K ⊗ (A ⋊α G) ∼= M[π]∈ bG K ⊗ Mdπ (A), of the representative of [a]. We must check that ϕ([a]) has countable support. Without loss of generality, assume that kak = 1. For 0 < ε < 1, there exists a finite K ⊗ Mdπ (A). Since set ϕ([a])([π]) = [ρπ(a)] ∈ Cu(A) for [π] ∈ bG. It is clear that ϕ([a]) is independent subset Xε of bG such that the element (a − ε)+ belongs to L[π]∈Xε ϕ([a]) = sup n∈N ψ "(cid:18)a − 1 n(cid:19)+#! , Cu(A) with finite support, so the support of ϕ([a]) is countable. Claim: ϕ preserves Cuntz subequivalence. Let a and b be positive elements in (K ⊗ Mdπ (A)) satisfying a - b. Without loss of generality, we may assume it follows that ϕ([a]) is a supremum of an increasing sequence of functions bG → L[π]∈ bG that kak = kbk = 1. Given n ∈ N, there exists m ∈ N such that (cid:0)a − 1 m(cid:1)+ - n(cid:1)+(cid:17) for all [π] ∈ bG. It follows that (cid:0)b − 1 n(cid:19)+#! ϕψ([b]). 1 n(cid:1)+. Hence, ρπ(cid:16)(cid:0)a − 1 ϕ "(cid:18)a − m(cid:1)+(cid:17) - ρπ(cid:16)(cid:0)b − 1 m(cid:19)+#! ≤ ϕ "(cid:18)b − 1 By taking supremum in m, we conclude that ϕ([a]) ≤ ϕ([b]), and the claim is proved. It is clear that the restriction of ϕ to the image in Cu(A ⋊α G) of the positive (K ⊗ Mdπ (A)) with finitely many nonzero coordinates preserves suprema of increasing sequences, preserves the compact containment relation, and is an order embedding. We want to show that ϕ is an isomorphism in the category elements in L[π]∈ bG 40 EUSEBIO GARDELLA AND LUIS SANTIAGO CuG. This will be a consequence of the next three claims, which will complete the proof of the proposition. Claim: ϕ is an order embedding. Let a and b be positive elements in M[π]∈ bG (K ⊗ Mdπ (A)) , ϕ, there exists n0 ∈ N such that increasing by the comments before this claim. In particular, for fixed n ∈ N, we we must have [a] ≤ [b]. For the general case, we can assume without loss of is rapidly n(cid:1)+i(cid:17)(cid:17)n∈N n(cid:1)+i(cid:17) by definition of and assume that ϕ([a]) ≤ ϕ([b]) in Cu(G) ⊗ Cu(A). If there exists a finite subset (K ⊗ Mdπ (A)), then it is clear that X ⊆ bG such that a and b belong to L[π]∈X generality that kak = kbk = 1. The sequence (cid:16)ϕ(cid:16)h(cid:0)a − 1 n(cid:1)+i(cid:17) ≪ ϕ([a]). Since ϕ([b]) = sup ϕ(cid:16)h(cid:0)b − 1 have ϕ(cid:16)h(cid:0)a − 1 n(cid:19)+#! ≪ ϕ "(cid:18)b − m(cid:19)+#! ϕ "(cid:18)a − "(cid:18)a − m(cid:19)+# , n(cid:19)+# ≤"(cid:18)b − n(cid:1)+ and(cid:0)b − 1 m(cid:1)+ have only finitely many nonzero coordinates. By because(cid:0)a − 1 Claim: ϕ is surjective. Let f : bG → Cu(A) be a function with countable support. Let (πn)n∈N be an enumeration of the support of f . For n ∈ N, let an ∈ K ⊗ A be a positive element with kank = 1 taking the supremum over m first, and then over n, we deduce that [a] ≤ [b], as desired. for all m ≥ n0. It follows that 1 1 n∈N 1 1 n satisfying [an] = f (πn) in Cu(A). Let a ∈ K ⊗ (A ⋊α G) ∼= M[π]∈ bG (K ⊗ Mdπ (A)) be the positive element determined by ρπn (a) = an for n ∈ N, and ρπ(a) = 0 for π /∈ supp(f ). It is then clear that ϕ([a]) = f . Claim: ϕ is a morphism in CuG. We need to check conditions (M1) and (M2) in Definition 2.2 and that ϕ is a Cu(G)-semimodule homomorphism. To check (M1), let (sn)n∈N be an increasing sequence in Cu(A ⋊α G), and set ϕ(sn). Let t ∈ Cu(G) ⊗ Cu(A) sn. We want to show that ϕ(s) = sup n∈N s = sup n∈N satisfy ϕ(sn) ≤ t for all n ∈ N. Since ϕ is surjective (see the previous claim), there exists r ∈ Cu(A ⋊α G) with ϕ(r) = t. Now, since ϕ is an order embedding (see claim above), we deduce that sn ≤ r for all n ∈ N. Hence s ≤ r by the definition of supremum. Thus ϕ(s) ≤ ϕ(r), and ϕ(s) = sup n∈N ϕ(sn). In order to check (M2), let s, t ∈ Cu(A⋊α G) satisfy s ≪ t. We want to show that ϕ(s) ≪ ϕ(t). To this end, let (rn)n∈N be an increasing sequence in Cu(G) ⊗ Cu(A) satisfying ϕ(t) ≤ supn∈N rn. For each n ∈ N, choose zn ∈ Cu(A ⋊α G) with ϕ(zn) = rn. Since ϕ is an order embedding, we have t ≤ supn∈N zn. Hence there exists m ∈ N with s ≤ zm, and so ϕ(s) ≤ ϕ(zm). This shows that ϕ(s) ≪ ϕ(t), as desired. For a positive element [µ ⊗ ρ] = X[π]∈ bG aτ ∈ K ⊗ Mdτ (A) ⊆ M[π]∈ bG we have [µ] · [aτ ] =h(mπ(µ ⊗ τ )aτ )[π]∈ bGi . (K ⊗ Mdπ (A)) ∼= K ⊗ (A ⋊α G), We will now check that ϕ is a Cu(G)-semimodule homomorphism. Observe first that A ⋊α G ∼= C∗(G) ⊗ A, and that the Cu(G)-module structure on Cu(A ⋊α G) ∼= Cu(C∗(G) ⊗ A) is the usual multiplication on Cu(C∗(G)) = Cu(G) and trivial on 41 Cu(A). Let [µ], [τ ] ∈ bG, and write [µ ⊗ ρ] as a linear combination mπ(µ ⊗ τ ) · [π]. Hence, for [π] ∈ bG, we have as desired. This completes the proof of the claim and the proposition. ϕ([µ] · [aτ ])([π]) = [mπ(µ ⊗ τ )aτ ] = ([µ] · ϕ([aτ ]))([bπ]), (cid:3) Similarly to what happens in equivariant K-theory, the Cu(G)-semimodule struc- ture on Cu(A ⋊α G) has a more concrete expression when G is abelian. We saw that Cu(G) consists of the suprema of all finite linear combinations of multiplication. In particular, it follows that a Cu(G)-semimodule structure on a partially ordered abelian semigroup that is compatible with suprema is necessarily elements of bG with coefficients in Z≥0, with coordinate-wise order, addition and completely determined by multiplication by the elements of bG. We denote by bα : bG → Aut(A ⋊α G) the dual action of α. proposition, we use the identification In the following HA =Mn∈NMπ∈ bG Hπ ⊗ A. Proposition 5.16. Let G be a compact abelian group, let A be a C∗-algebra, and let α : G → Aut(A) be a continuous action. Then for τ ∈ bG and s ∈ Cu(A ⋊α G), we have τ · s = Cu(bατ )(s). More precisely, the following diagram commutes: / CuG(A, α) CuG(A, α) τ · σ σ Cu(A ⋊α G) Cu( bατ ) / Cu(A ⋊α G), where σ : CuG(A, α) → Cu(A ⋊α G) is the natural Cu-isomorphism given by Theorem 5.3. Proof. Fix τ ∈ bG. By the construction of the Cu-isomorphism σ : CuG(A, α) → Cu(A ⋊α G) from Theorem 5.3, and adopting the notation in its proof, it is enough   /   / 42 EUSEBIO GARDELLA AND LUIS SANTIAGO to show that the following diagram commutes: Cu(K(HA)G) τ · / Cu(K(HA)G) Cu(θ) Cu(θ) Cu(K(ℓ2(N)) ⊗ K(L2(G) ⊗ A)G) Cu(K(ℓ2(N)) ⊗ K(L2(G) ⊗ A)G) κ κ Cu(K(L2(G) ⊗ A)G) Cu( bατ ) / Cu(K(L2(G) ⊗ A)G). The dual action bα : bG → Aut((A ⊗ K(L2(G)))G) has the following description. For χ ∈ bG, let uχ ∈ U(L2(G)) be the corresponding multiplication operator, which is given by uχ(ξ)(g) = χ(g)ξ(g) for all ξ ∈ L2(G) and for all g ∈ G. Define an automorphism γχ : A ⊗ K(L2(G)) → A ⊗ K(L2(G)) by γχ = idA ⊗ Ad(uχ). It is clear that γχ commutes with α ⊗ Ad(λg) restriction of γχ to (A ⊗ K(L2(G)))G. for all g ∈ G, and thus γχ leaves (A ⊗ K(L2(G)))G invariant. Then bαχ is the For χ ∈ bG, the action on K(HA)G can be described as follows. Write HA = ℓ2(N) ⊗ L2(G) ⊗ A, and let uχ be the unitary on L2(G) described above. Then wχ = idℓ2(N) ⊗ uχ ⊗ idA is a unitary on HA, and conjugation by wχ defines an automorphism of K(HA). This automorphism clearly commutes with the action of G on K(HA), so it defines, by restriction, an automorphism of its fixed point algebra K(HA)G. is commutative, and the result follows. Using these descriptions of the actions of bG, it is clear that the diagram above We close this section with an application to invariant hereditary subalgebras. (cid:3) The result is a Cuntz semigroup analog of Proposition 2.9.1 in [Phi87]. Proposition 5.17. Suppose that A is separable and G is second countable. Let B ⊆ A be an α-invariant hereditary subalgebra of A, and denote by β : G → Aut(B) the compression of α. If B is full, then the canonical inclusion induces a natural CuG-isomorphism CuG(B, β) → CuG(A, α). Proof. Under the canonical identification given by Theorem 5.3, the map in the statement becomes the map Cu(B ⋊β G) → Cu(A ⋊α G) induced by the inclusion. Now, Proposition 2.9.1 in [Phi87] shows that B⋊βG is a full hereditary subalgebra of A ⋊α G. Separability of the objects implies, by Brown's stability theorem, that they are stably isomorphic. It follows that the canonical map Cu(B⋊β G) → Cu(A⋊αG), which belongs to CuG by Theorem 3.11, is an isomorphism. (cid:3) 6. Examples and computations In this section, we compute the equivariant Cuntz semigroups of a number of dynamical systems. In most of our examples, the Rokhlin property for an action of a finite group comes up as an useful technical device that makes the computations possible. The Rokhlin property is also implicitly used in Theorem 8.1. We therefore recall its definition here.   /       / Definition 6.1. Let A be a σ-unital C∗-algebra, let G be a finite group, and let α : G → Aut(A) be an action. We say that α has the Rokhlin property if for every finite set F ⊆ A and every ε > 0, there exists orthogonal positive elements ag ∈ A, for g ∈ G, satisfying 43 (1) k(αg(ah) − agh)xk < ε for all g, h ∈ G; (2) kagx − xagk < ε for all g ∈ G and all x ∈ F ; < ε for all x ∈ F . ag! − x(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (3) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) x Pg∈G N =(cid:26)s ∈ S : ∃ (sn)n∈N in Sγ : sn ≪ sn+1 ∀ n ∈ N, s = sup Sγ n∈N sn(cid:27) . Let G be a finite group, let S be a semigroup in the category Cu, and let γ : G → Aut(S) be an action. Set Sγ = {s ∈ S : γg(s) = s for all g ∈ G}, and set For use in the following proposition, we observe that if α : G → Aut(A) is an action of a second countable compact group G, then its equivariant Cuntz semi- group CuG(A, α) can be canonically identified with Cu(cid:0)(A ⊗ K(L2(G)))α⊗Ad(λ)(cid:1), where the Cu(G)-semimodule structure is given by tensor product (where we iden- tify (L2(G) ⊗ Hµ, λ ⊗ µ) with (L2(G), λ) whenever (Hµ, µ) is a separable unitary representation of G). Proposition 6.2. Let A be a C∗-algebra, let G be a finite group, and let α : G → Aut(A) be an action with the Rokhlin property. Then there exists a natural CuG- isomorphism CuG(A, α) ∼= Cu(A)Cu(α) N , where the induced Cu(G)-semimodule structure on Cu(A)Cu(α) N is trivial. Proof. Denote by e ∈ K(ℓ2(G)) the projection onto the constant functions, and let ι : Aα ֒→ (A ⊗ K(L2(G)))α⊗Ad(λ) be the inclusion given by ι(a) = a ⊗ e for all a ∈ A. Since α has the Rokhlin property, ι induces a Cu-isomorphism between the Cuntz semigroups Cu(Aα) and Cu((A ⊗ K(L2(G)))α⊗Ad(λ)) by Proposition 2.6 in [Gar14c]. By the comments above this proposition, we deduce that CuG(A, α) is naturally isomorphic to Cu(Aα). Since Cu(Aα) can be canonically identified with Cu(A)Cu(α) by Theorem 4.1 in [GS16], we conclude that the inclusion of Aα into A induces a Cu-isomorphism of Cu(Aα) and Cu(A)Cu(α) . N N We give two concrete applications of the above computation. Recall that for is the identity map. We deduce from Theorem 5.14 that the Cu(G)-semimodule structure on CuG(A, α) is trivial. This finishes the proof. (cid:3) Finally, the dual action bα is approximately representable by part (i) of Proposi- tion 4.4 in [Naw12]. In particular,bατ is approximately inner for all τ ∈ bG, so Cu(bατ ) m ∈ N, there is a canonical identification Cu(Mm∞) ∼= Z≥0(cid:2) 1 m(cid:21) ⊔ R>0, Example 6.3. Let G be a finite group and set m = G. Let µG : G → Aut(Mm∞) be the action considered, for example, in [GS16, Example 2.1]. Then there is a CuG-isomorphism CuG(MG∞, µG) ∼= Z≥0(cid:20) 1 m(cid:3) ⊔ R>0 where the right-hand side carries the trivial Cu(G)-semimodule structure. 44 EUSEBIO GARDELLA AND LUIS SANTIAGO Proof. The Cu(G) -semimodule structure is trivial by Proposition 6.2, since µG has the Rokhlin property. Also, by Proposition 6.2, we have CuG(MG∞, µG) ∼= Cu(MG∞)Cu(µG) N . The computation follows since Cu(µG g ) = idCu(MG∞ ) for all g ∈ G. (cid:3) n ∈ T. Let γ : Zn → Aut(C(T)) be Example 6.4. Let n ∈ N and set ωn = e the action given by γk(f )(z) = f (ωk nz) for all z ∈ T and all k ∈ Zn. Let A be any unital Mn∞-absorbing C∗-algebra, and let α : Zn → Aut(A) be any action. Then there is a CuZn-semimodule isomorphism 2πi CuZn (C(T, A), γ ⊗ α) ∼= {f ∈ Lsc((T, Cu(A)) : f (ωnz) = f (z) for all z ∈ T} , where the Cu(Zn)-semimodule structure on the right-hand side is trivial. Proof. Note that γ has the unitary Rokhlin property from Definition 3.5 in [Gar14a] (take the unitary u in that definition to be u(z) = z for z ∈ T). One easily checks that γ ⊗ α also has the unitary Rokhlin property. Since A is assumed to absorb Mn∞, it follows from Theorem 3.19 in [Gar14a] that γ ⊗ α has the Rokhlin property. We deduce from Proposition 6.2 that the Cu(Zn)-semimodule structure on CuZn (C(T, A)) is trivial. Again by Proposition 6.2, we have CuZn (C(T, A), γ ⊗ α) ∼= Cu(C(T, A)Cu(γ⊗α) . N The result then follows from [APS11, Theorem 3.4]. (cid:3) Let W be the stably projectionless simple C∗-algebra studied in [Jac13]. This algebra has a unique tracial state and trivial K-groups. By [ERS11, Corollary 6.8], we have Cu(W) ∼= R≥0. Moreover, every automorphism of W is approximately inner by [Rob12, Theorem 1.0.1]. Consequently, if α : G → Aut(A) is an action of a group G on W, then Cu(αg) = idCu(W) for all g ∈ G. The following computations are similar to the previous ones, so we will omit them. Example 6.5. Let G be a finite group and let µ : G → Aut(W) be the unique (up to conjugacy) action with the Rokhlin property; see [Naw12]. Then CuG(W, µ) ∼= R≥0 and CuZn (C(T, W), γ ⊗ µ) ∼=(cid:8)Lsc(T, R≥0) : f (ωnz) = f (z) for all z ∈ T(cid:9) . Moreover, the corresponding Cu(G) and Cu(Zn)-semimodule structures are trivial. 6.1. Pullbacks of dynamical systems. Let (A, α) and (B, β) be G-C∗-algebras. Let I be an invariant closed two-sided ideal in A, and let φ : A → A/I denote the quotient map. Note that φ is equivariant when taking on A/I the action αA/I induced by α. Let ψ : (B, β) → (A/I, αA/I ) be an equivariant ∗-homomorphism. By [Ped99, Proposition 6.2], the following pullback exists in the category of G-C∗- algebras: (G, C, γ) πB πA (A, α) (B, β) ψ / (A/I, αA/I ), φ / /     / where we set C = A⊕A/I B, denote by πA : C → A and πB : C → B the correspond- ing quotient maps, and take γ = (α, β) to be the pullback action of G on C. By applying the functor CuG to the diagram above, we get the following commutative diagram in the category CuG: 45 CuG(C, γ) CuG(πB ) CuG(B, β) CuG(πA) CuG(ψ) CuG(A, α) CuG(φ) / CuG(A/I, αA/I ). Consider the pullback CuG(A, α) ⊕CuG(A/I,αA/I ) CuG(B, β) in the category of ordered semigroups with the Cu(G)-semimodule structure induced by the ones in CuG(A, α) and CuG(B, β). There is a natural morphism in the category CuG ρ : CuG(C, γ) → CuG(A, α) ⊕CuG(A/I,αA/I ) CuG(B, β) given by ρ([(a, b)]G) = ([a]G, [b]G) for [(a, b)]G ∈ CuG(C, γ). Proposition 6.6. Adopt the notation from the discussion above, and suppose that A/I ⋊αA/I G has stable rank one and that each of its closed two-sided ideals has trivial K1-group. Then ρ is an order embedding. Proof. Denote by bπA, bπB, bφ, and bψ the maps at the level of the crossed products induced by πA, πB, φ, and ψ, respectively. By Theorem 6.3 of [Ped99], the following diagram is a pullback C ⋊γ G bπA A ⋊α G bπB bφ B ⋊β G bψ / A/I ⋊αA/I G. In other words, there is a natural ∗-isomorphism C ⋊γ G ∼= A ⋊α G ⊕A/I⋊αA/I G B ⋊β G. Therefore, by Theorem 5.3, it is enough to show that the map Cu(A ⋊α G ⊕A/I⋊αA/I G B ⋊β G) → Cu(A ⋊α G) ⊕Cu(A/I⋊αA/I G) Cu(B ⋊β G), given by [(a, b)]G 7→ ([a]G, [b]G), is an order embedding. That this is the case is a consequence of [APS11, Theorem 3.1], by the assumptions on A/I ⋊αA/I G. (cid:3) We have arrived at the main result of this subsection. Theorem 6.7. Let X be a compact metric space and let G be a compact Lie group with dim(X) ≤ dim(G) + 1. Let γ : G → Aut(C(X)) be an action induced by a free action of G on X, and let Y be an invariant closed subset of X. Let (A, α) and (B, β) be G-C∗-algebras with A separable, of stable rank one, and such that the K1-groups of all its closed two-sided ideals are trivial. Let φ : (C(X, A), γ ⊗ α) → (C(Y, A), γC(Y ) ⊗ α) be the canonical equivariant quotient map, and let ψ : (B, β) → (C(Y, A), γC(Y ) ⊗ α) / /     / / /     / 46 EUSEBIO GARDELLA AND LUIS SANTIAGO be an equivariant *-homomorphism. Then the ordered semigroup (2) CuG(C(X, A), γ ⊗ α) ⊕CuG(C(Y,A),γC(Y )⊗α) CuG(B, β) belongs to the category CuG, and the map ρ from CuG(C(X, A) ⊕C(Y,A) B, (γ ⊗ α, β)) to the semigroup in Equation 2, given by ρ([(a, b)]G) = ([a]G, [b]G) for [(a, b)]G ∈ CuG(C(X, A) ⊕C(Y,A) B, (γ ⊗ α, β)), is an isomorphism in the category CuG. Proof. By Theorem 5.3, there are natural isomorphisms CuG(C(X, A), γ ⊗ α) ∼= Cu(C(X, A) ⋊γ⊗α G), CuG(C(Y, A), γC(Y ) ⊗ α) ∼= Cu(C(Y, A) ⋊γC(Y )⊗α G), CuG(B, β) ∼= Cu(B ⋊β G). Also, by [Ped99, Theorem 6.3], the following diagram is a pullback (C(X, A) ⊕C(Y,A) B) ⋊(γ⊗α,β) G eπB B ⋊β G bπC(X,A) bψ C(X, A) ⋊γ⊗α G / C(Y, A) ⋊γC(Y )⊗α G, bφ where bφ and bψ are the maps induced by φ and ψ, and the mapsbπC(X,A) andbπB are the maps induced by the canonical coordinate projections πC(X,A) : C(X, A)⊕C(Y,A) B → C(X, A) and πB : C(X, A) ⊕C(Y,A) B → B. Since the action of G on X is free, the crossed product C(X, A) ⋊γ⊗α G is a C(X/G)-algebra with fibers isomorphic to A⊗K(L2(G)) by Theorem 5.2 in [GHS16], and similarly for C(X, A) ⋊γC(Y )⊗α G. The assumptions on G and X imply that In addition, A ⊗ K(L2(G)) is separable, has sta- dim(Y /G) ≤ dim(X/G) ≤ 1. ble rank stable rank one, and K1(J) = 0 for every closed two-sided ideal J in A ⊗ K(L2(G)). Hence the conditions of [ABP13, Theorem 2.6] and [APS11, Theo- rem 3.3] are satisfied, and the statements in the theorem follow. (cid:3) We will now use the above theorem to compute the equivariant Cuntz semigroup of some C∗-dynamical systems. In the following example, for n ∈ N we identify Zn with the subgroup of T consisting of the n-th roots of unity. Example 6.8. Adopt the notation of Example 6.3. Let (B, β) be the Zn-dynamical system given by B = {f ∈ C(T, Mn∞ ) : f (ωk n) = f (1) ∈ C1Mn∞ for k ∈ Zn}, and β = (γ ⊗ µZn )B. Then CuZn (B, β) is isomorphic, in the category CuZn, to the semigroup n(cid:21) ⊔ R>0(cid:19) ⊕ C(cZn, Z≥0) :  g(η)  (f, g) ∈ Lsc(cid:18)T, Z≥0(cid:20) 1 and f (1) = Xη∈ cZn where for τ ∈ cZn, the corresponding semimodule structure is given by for all z ∈ T and for all η ∈ cZn. [τ · (f, g)] (z, η) = (f (z), g(τ + η)) f (ωnz) = f (z) for z ∈ T , / /     / 47 Proof. Set Y = Zn ⊆ T. Then (B, β) is given by the following pullback diagram: (B, β) (C, idC) ψ (C(T, Mn∞ ), γ ⊗ µZn) φ / (C(Y, Mn∞ ), γC(Y ) ⊗ µZn ), where φ is the restriction map to Y , and ψ is the map given by φ(z)(y) = z1Mn∞ for all z ∈ C and for all y ∈ Y . Recall that Mn∞ is simple and its K1-group is trivial. By Theorem 6.7, and using that the action of Zn on T induced by γ is free, we get the following pullback diagram in the category CuZn : CuZn(B, β) CuZn (C, idC) CuZn (ψ) CuZn (C(T, Mn∞ ), γ ⊗ µZn) CuZn (φ) / CuZn(C(Y, Mn∞ ), γC(Y ) ⊗ µZn). The isomorphism in the statement of the example now follows using the defi- nition of pullbacks, the computations given in Example 6.3 and Proposition 5.15 in [GHS16], and that the map CuZn (ψ) : C(cZn, Z≥0) → Z≥0(cid:20) 1 n(cid:21) ⊔ R>0 is given by CuZn (ψ)(f ) = Pj∈Zn f (k) for all f ∈ C(cZn, Z≥0). The semimodule structure described in the example can be computed using the following facts: (1) the semimodule structure on CuZn(C(T, Mn∞ ), γ ⊗ µZn ) is trivial since γ ⊗ µZn has the Rokhlin property; and (2) given τ ∈ cZn, the corresponding action on is given by (τ · g)(η) = g(τ + η) for all g ∈ C(cZn, Z≥0) and for all η ∈ cZn. CuZn(C, idC) ∼= C(cZn, Z≥0) The proof of the following example is similar to that of the previous example (cid:3) This finishes the proof. and so we will omit it. Example 6.9. Adopt the notation of Example 6.5. Let (B, β) be the Zn-dynamical system given by B = {f ∈ C(T, W) : f (k) = f (1) ∈ W α for k ∈ Zn ⊆ T}, and β = (γ ⊗ α)B. Then CuZn (B, β) is CuZn-isomorphic to the semigroup f (ωnz) = f (z) for z ∈ T (f, g) ∈ Lsc(cid:0)T, R≥0(cid:1) ⊕ C(cZn, R≥0) : g(η)   where for τ ∈ cZn, the corresponding semimodule structure is given by for all z ∈ T and for all η ∈ cZn. and f (1) = Xη∈ cZn [τ · (f, g)] (z, η) = (f (z), g(τ + η)) , / /     / / /     / 48 EUSEBIO GARDELLA AND LUIS SANTIAGO 7. A characterization of freeness using the equivariant Cuntz semigroup In this section, we give an application of the equivariant Cuntz semigroup in the context of free actions of locally compact spaces, which resembles Atiyah-Segal's characterization of freeness using equivariant K-theory; see [AS69]. Indeed, in Theorem 7.6, we characterize freeness of a compact Lie group action on a com- mutative C∗-algebra in terms of a certain canonical map to the equivariant Cuntz semigroup. We define this map, for arbitrary C∗-algebras, below. Definition 7.1. Let α : G → Aut(A) be a continuous action of a compact group G on a C∗-algebra A. We define a natural Cu-map φ : Cu(AG) → CuG(A, α) as follows. Given a positive element a ∈ K(ℓ2(N)) ⊗ AG, regard it as an element in (K(ℓ2(N)) ⊗ A)G by giving ℓ2(N) the trivial G-representation, and set φ([a]) = [a]G. Remark 7.2. Here is an alternative description of φ. Let ι : AG → A be the canonical inclusion. Since ι is equivariant, it induces a CuG-morphism Cu(ι) : CuG(AG) → CuG(A, α) between the equivariant Cuntz semigroups. Now, by Proposition 5.15, there exists a natural CuG-isomorphism CuG(AG) ∼= Cu(G) ⊗ Cu(AG). Then φ is the restriction of Cu(ι) to the second tensor factor. We need a proposition first, which is interesting in its own right. Let α : G → Aut(A) be a continuous action of a compact group G on a C∗-algebra A, and let a ∈ AG. We denote by ca : G → A the continuous function with constant value equal to a. Note that ca belongs to L1(G, A, α), and the assignment a 7→ ca defines a ∗-homomorphism c : AG → L1(G, A, α). (Recall that the product in L1(G, A, α) is given by twisted convolution.) Proposition 7.3. Let α : G → Aut(A) be a continuous action of a compact group G on a C∗-algebra A. Denote by σ : CuG(A, α) → Cu(A ⋊α G) the canonical Cu-isomorphism constructed in Theorem 5.3. Then there exists a commutative diagram Cu(AG) φ / CuG(A, α) w♦♦♦♦♦♦♦♦♦♦♦♦ σ Cu(c) Cu(A ⋊α G). Proof. Abbreviate K(ℓ2(N)) (with the trivial G-action) to K. By the construction of the map σ, we need to show that the following diagram is commutative: Cu(AG) φ / CuG(A, α) χ / Cu(K(HA)G) Cu(c) Cu(θ) Cu(A ⋊α G) Cu(ψ) Cu((K(L2(G)) ⊗ A)G) κ Cu((K ⊗ K(L2(G)) ⊗ A)G).   / w   / /   o o o o For an irreducible representation (Hπ, π) of G, set dπ = dim(Hπ). Write 49 HC = ℓ2(N) ⊗M[π]∈ bG  dπMj=1 Hπ . Let 1G : G → U(C) denote the trivial representation, and let W : ℓ2(N) ∼= ℓ2(N) ⊗ H1G ֒→ HC be the isometry corresponding to the canonical inclusion (H1G is just C). Write V : ℓ2(N) ⊗ A → HA for W ⊗ idA. Let a positive element a ∈ K ⊗ AG be given. Then (χ ◦ φ)([a]) corresponds to the class of [V aV ∗] in Cu(K(HA)G). Denote by e : L2(G) → L2(G) the projection onto the constant functions. With the presentation of HC used above, it is clear that Cu(θ) maps [V aV ∗] to the class of a ⊗ e ∈ (K ⊗ A ⊗ K(L2(G)))G ∼= (K ⊗ K(L2(G)) ⊗ A)G. Since κ is induced by the embedding (K(L2(G)) ⊗ A)G → K ⊗ (K(L2(G)) ⊗ A)G as the upper left corner, and since ψ is equivariant (see Proposition 5.2), it is now not difficult to check that σ(φ([a])) agrees with [ca] in Cu(A ⋊α G). (cid:3) We recall a version of the Atiyah-Segal completion theorem that is convenient for our purposes. If a compact group G acts on a compact Hausdorff space X, then there exists a canonical map K ∗(X/G) → K ∗ G(X) obtained by regarding a vector bundle on X/G as a G-vector bundle on X (using the trivial action). For a compact group G, we denote by IG the augmentation ideal in R(G). That is, IG is the kernel of the dimension map R(G) → Z. Theorem 7.4. (Atiyah-Segal). Let X be a compact Hausdorff space and let a compact Lie group G act on X. The the following statements are equivalent: (1) The action of G on X is free. (2) The natural map K ∗(X/G) → K ∗ (3) The natural map K 0(X/G) → K 0 G(X) is an isomorphism. G(X) is an isomorphism. Proof. That (1) implies (2) is proved in Proposition 2.1 in [Seg68]. That (2) implies (3) is obvious. Let us show that (3) implies (1), so assume that the natural map K 0(X/G) → K 0 G(X) is an isomorphism. G(X) = K 0 G(X) ⊕ K 1 An inspection of the proof of the implication (4) ⇒ (1) in Proposition 4.3 of [AS69] shows that, in our context, there exists n ∈ N such that I n 0 (X) = 0. Now, the R(G)-module K ∗ G(X) is in fact an R(G)-algebra, where multiplication is given by tensor product (with diagonal G-actions). In this algebra, the class of the trivial G-bundle over X is the unit, so it belongs to K 0 G(X). In particular, I n G(X), and hence it annihilates all of K ∗ ∗ (X) is discrete in the IG-adic topology. The implication (1) ⇒ (4) in Proposition 4.3 of [AS69] now shows that the G-action is free. (cid:3) ∗ (X) = 0. In other words, K G G annihilates the unit of K ∗ G(X), that is, I n G · K G G · K G We mention here that the implication (1) ⇒ (2) holds even if G is not a Lie group, and even if X is merely locally compact (this is essentially due to Rieffel; see the proof of Theorem 7.6 below for a similar argument). However, the equivalence between (2) and (3) may fail if X is not compact: the trivial action on R is a counterexample. This can happen even for free actions. 50 EUSEBIO GARDELLA AND LUIS SANTIAGO Recall that a unital C∗-algebra A is said to be finite if u ∈ A and u∗u = 1 imply uu∗ = 1. A nonunital C∗-algebra is finite if its unitization is. Finally, a C∗- algebra A is stably finite if Mn(A) is finite for all n ∈ N. Commutative C∗-algebras and AF-algebras are stably finite, as are the tensor products of these, and their subalgebras. Remark 7.5. By Theorem 3.5 in [BC09], if A is a stably finite C∗-algebra, then the set of compact elements in Cu(A) can be naturally identified with the Murray-von Neumann semigroup V (A) of A. In the next theorem, for an abelian semigroup V , we denote by G(V ) its Gro- thendieck group. Also, if α : G → Aut(A) is an action of a compact group on a C∗- algebra A, we denote by V G(A) the semigroup of equivariant Murray-von Neumann equivalence classes of projections in (A ⊗ K(Hµ))G, where µ : G → U(H) is a finite dimensional unitary representation. With this notation, the equivariant K-theory K G ∗ (A) of A is, by definition, the group G(V G(A)); see Chapter 2 in [Phi87]. Theorem 7.6. Let X be a locally compact, metric space and let a compact group G act on X. Consider the following statements: (1) The action of G on X is free. (2) The canonical map φ : Cu(C0(X/G)) → CuG(C0(X)) is a Cu-isomorphism. Then (1) implies (2). If G is a Lie group and X is compact, then the converse is also true. Proof. Assume that the action of G on X is free, and denote by α : G → Aut(C0(X)) the induced action. By Proposition 7.3, under the identification CuG(C0(X), α) ∼= Cu(C0(X) ⋊α G) provided by Theorem 5.3, the canonical map φ becomes the map at the level of the (ordinary) Cuntz semigroup induced by the map c : C0(X)G = C0(X/G) → C0(X) ⋊α G defined before Proposition 7.3. Denote by e ∈ K(L2(G)) the projection onto the constant functions. By the main theorem in [Ros79], we have c(C0(X)G) = e(C0(X) ⋊α G)e. Since Cu is a stable functor, it is enough to show that e is a full projection in C0(X) ⋊α G. For a, b ∈ C0(X), denote by fa,b : G → C0(X) the function given by fa,b(g)(x) = a(x)b(g−1 · x) for g ∈ G and x ∈ X. It is an easy consequence of the Stone-Weierstrass theorem that the set {fa,b : a, b ∈ C0(X)} has dense linear span in C0(X) ⋊ G. Denote by I the ideal in C0(X) ⋊ G generated by e. Let (aλ)λ∈Λ be an approx- imate identity for C0(X). Upon averaging over G, we may assume that aλ belongs to C0(X)G for all λ ∈ Λ. Let a, b ∈ C0(X). Then fa,aλb = caλfa,b for λ ∈ Λ, and hence fa,b = lim λ∈Λ fa,aλb = lim λ∈Λ (caλ fa,b) , so fa,b belongs to I. We conclude that C0(X) ⋊ G = I, as desired. Assume now that G is a compact Lie group, that X is compact, and that the canonical map Cu(C(X/G)) → CuG(C(X)) is an isomorphism in Cu. We claim that the canonical map K 0(X/G) → K 0 G(X) is an isomorphism. Under the natural identification given by Theorem 5.3, the canonical inclusion c : C(X/G) → C(X) ⋊ G induces an isomorphism Cu(c) at the level of the Cuntz 51 semigroup (see also Proposition 7.3). The algebra C(X/G) is clearly stably finite. On the other hand, C(X) ⋊α G is also stably finite because it is a subalgebra of the stable finite C∗-algebra C(X) ⊗ K(L2(G)). By Remark 7.5, the restriction of the isomorphism Cu(c) to the compact elements of C(X/G) yields an isomorphism ψ : V (C(X/G)) → V (C(X) ⋊ G) between the respective Murray-von Neumann semigroups of projections. By taking the Grothendieck construction, one gets an isomorphism ϕ : G(V (C(X/G))) → G(V (C(X) ⋊ G)) between the respective Grothendieck groups. We want to conclude from this that ϕ induces an isomorphism between the K0-groups of these C∗-algebras. Since C(X/G) is unital, we have G(V (C(X/G))) = K0(C(X/G)). However, C(X) ⋊ G is not unital unless G is finite, and it is even not clear whether it (or its stabilization) has an approximate identity consisting of projections. (This would also imply that its K0-group is obtained as the Grothendieck group of its Murray- von Neummann semigroup.) Instead, we appeal to Julg's theorem for equivariant K-theory. Indeed, the proof given in Theorem 2.6.1 in [Phi87] shows that if A is a unital C∗-algebra, G is a compact group, and α : G → Aut(A) is a continuous action, then there exists a canonical isomorphism of semigroups V G(A) ∼= V (A ⋊ G). Since K G for K0(A ⋊α G). In our context, this shows that ϕ induces an isomorphism 0 (A) is the Grothendieck group of V G(A), it follows that the same is true θ : K 0(X/G) → K G 0 (C(X)) ∼= K 0 G(X). Now, the implication (3) ⇒ (1) in Theorem 7.4 shows that the action of G on (cid:3) X is free. 8. Classification of actions using CuG In this section, we classify a class of actions of finite abelian groups on certain stably finite C∗-algebras using the equivariant Cuntz semigroup; see Theorem 8.1. We describe the general strategy first. Let G be a finite abelian group, and let α and β be actions of G on C∗-algebras A and B. A CuG-homomorphism ρ : CuG(A, α) → CuG(B, β) can be regarded, via Theorem 5.14, as a Cu-homomorphism Cu(A ⋊α G) → Cu(B ⋊β G) that is to a class of C∗-algebras for which the Cuntz semigroup classifies homomorphisms (up to approximate unitary equivalence), then we can obtain a ∗-homomorphism equivariant with respect to the dual actionsbα and bβ. If A ⋊α G and B ⋊β G belong ϕ : A ⋊α G → B ⋊β G such that ϕ ◦bατ is approximately unitarily equivalent to bβτ ◦ ϕ for all τ ∈ bG. When the dual actions have the Rokhlin property, results from [GS16] can be used to replace ϕ with an approximately unitarily equiva- lent ∗-homomorphism θ which is actually equivariant. If θ moreover satisfies a scale condition (which can be phrased in terms of ρ), one can essentially restrict ψ (using Theorem 8.1) to an equivariant homomorphism ψ : A → B which satisfies CuG(ψ) = ρ. Finally, when ρ is an isomorphism, we can choose θ and ψ to be ∗-isomorphisms. 52 EUSEBIO GARDELLA AND LUIS SANTIAGO The following result will allow us to go from equivariant ∗-homomorphisms be- tween double crossed products to equivariant ∗-homomorphisms between the orig- inal dynamical systems. Recall that - stands for Cuntz subequivalence (see the beginning of Section 2.1). Theorem 8.1. Let G be a finite group, let A and B be C∗-algebras such that B has stable rank one, and let α : G → Aut(A) and β : G → Aut(B) be actions. Denote by λ : G → U(ℓ2(G)) the left regular representation, and let e ∈ K(ℓ2(G)) be the projection onto the constant functions on G. Let sA and sB be G-invariant strictly positive elements in A and B, respectively. (1) Every equivariant ∗-homomorphism ϕ : (A ⊗ K(ℓ2(G)), α ⊗ Ad(λ)) → (B ⊗ K(ℓ2(G)), β ⊗ Ad(λ)) satisfying ϕ(sA ⊗ e) - sB ⊗ e in (B ⊗ K(ℓ2(G)))β⊗Ad(λ) induces an equivariant ∗-homomorphism ψ : (A, α) → (B, β), and con- versely. (2) The actions α and β are conjugate if and only if there exists an equivariant isomorphism ϕ : (A ⊗ K(ℓ2(G)), α ⊗ Ad(λ)) → (B ⊗ K(ℓ2(G)), β ⊗ Ad(λ)) such that ϕ(sA ⊗ e) is Cuntz equivalent to sB ⊗ e in (B ⊗ K(ℓ2(G)))β⊗Ad(λ). Proof. (1). If ψ : A → B is an equivariant ∗-homomorphism, then ϕ = ψ ⊗ idK(ℓ2(G)) : A ⊗ K(ℓ2(G)) → B ⊗ K(ℓ2(G)) is also equivariant and satisfies ϕ(sA ⊗ e) - sB ⊗ e in (B ⊗ K(ℓ2(G)))β⊗Ad(λ). Conversely, let ϕ be an equivariant ∗-homomorphism as in the statement. Using that B has stable rank one together with Theorem 3 in [CEI08], choose x ∈ (B ⊗ K(ℓ2(G)))β⊗Ad(λ) such that x∗x = ϕ(sA ⊗ e) and xx∗ ∈ Her(sB ⊗ e). Let x = vx be the polar decomposition of x in the bidual of (B ⊗ K(ℓ2(G)))β⊗Ad(λ). Then conjugation by v defines a ∗-isomorphism Ad(v) : Her(ϕ(sA ⊗ e)) → Her(xx∗). Using that v is fixed by the extension of β ⊗ Ad(λ) to the bidual of B ⊗ K(ℓ2(G)), we get Ad(v) ◦ (β ⊗ Ad(λ)) ◦ Ad(v∗)Her(ϕ(sA⊗e)) = Ad(vv∗) ◦ (β ⊗ Ad(λ))Her(ϕ(sA⊗e)). Now, since vv∗ is a unit for Her(xx∗), the above composition equals β ⊗ Ad(λ) on Her(ϕ(sA ⊗ e)). It follows that Ad(v) is equivariant with respect to β ⊗ Ad(λ). Define ψ : A → B to be the following composition: A ∼= Her(sA ⊗ e) ϕ / Her(ϕ(sA ⊗ e)) Ad(v) / Her(sB ⊗ e) ∼= B. Then ψ is easily seen to be equivariant, and this finishes the proof. (2). Use Proposition 2.5 in [CES11] to choose x ∈ (B ⊗ K(ℓ2(G)))β⊗Ad(λ) with x∗x = ϕ(sA ⊗ e) and Her(xx∗) = Her(sB ⊗ e). Keeping the notation from the previous part, it follows that Ad(v) is an (equivariant) isomorphism between / / Her(ϕ(sA ⊗ e)) and Her(sB ⊗ e). We conclude that ψ = Ad(v) ◦ ϕ determines an equivariant isomorphism A → B, as desired. (cid:3) Next, we introduce the class of actions we will focus on. 53 Definition 8.2. An action α of a finite group G on a C∗-algebra A is said to be locally representable if there exist an increasing sequence (An)n∈N of subalgebras An, and of A such that An is α-invariant for all n ∈ N and such that A = Sn∈N unitary representations u(n) : G → U(M (An)) of G, for n ∈ N, such that such that αg(a) = Ad(u(n) g )(a) for all g ∈ G, for all a ∈ An, and for all n ∈ N. Let B be a class of C∗-algebras and let A be a C∗-algebra in B. If α is a locally representable action of G on A such that the subalgebras An as above can be chosen to belong to B, then we say that α is locally representable in B. Actions that are locally representable in the class of AF-algebras were stud- ied and classified by Handelman and Rossmann in [HR85]. The invariant they used is easily seen to be equivalent to the equivariant K-theory of the actions. In Theorem 8.4, we use the equivariant Cuntz semigroup to classify actions that are locally representable actions in a class of stably finite algebras containing all AI-algebras. We will need the following easy preservation result. Lemma 8.3. Let B be a class of C∗-algebras that is closed under countable direct limits and under direct sums. Let A be a C∗-algebra in B, let G be a finite group and let α be an action of G on A. Assume that α is locally representable in B. Then A ⋊α G belongs to B. Proof. Choose an increasing sequence (An)n∈N of subalgebras of A that belong to An, and unitary representations u(n) : G → U(M (An)) of G such B with A = lim −→ that such that αg(a) = Ad(u(n) g )(a) for all g in G, for all a in An, and for all n in N. Using continuity of the crossed product functor in the first step and the fact that inner actions are cocycle equivalent to the trivial action in the second step, we conclude that A ⋊α G ∼= lim −→ An ⋊ Ad(u(n)) G ∼= lim −→ (An ⊗ C∗(G)) ∼= lim −→ An ⊕ · · · ⊕ An. Now, the assumptions on the class B imply that A ⋊α G belongs to B. (cid:3) The following is the main result of this section. For a finite group G, we denote by e ∈ K(ℓ2(G))G the projection onto the constant functions. If α : G → Aut(A) is an action and sA ∈ AG is a strictly positive element in A, the element sA ⊗e belongs to (A ⊗ K(ℓ2(G)))G, and hence it has a well-defined class [sA ⊗ e]G in CuG(A, α). We refer the reader to [Rob12] for the definition of 1-dimensional NCCW-com- plexes, as well as for the classification of certain direct limits of such C∗-algebras. In the next theorem, R will denote the class of unital C∗-algebras that can be written as inductive limits of one-dimensional NCCW-complexes with trivial K1-groups. (Unitality of the algebras can be dropped if one instead assumes the existence of an approximate unit consisting of projections.) Algebras in R are classified by their Cuntz semigroup, by the main result in [Rob12]. Theorem 8.4. Let G be a finite abelian group, let A and B be C∗-algebras in R. Let α : G → Aut(A) and β : G → Aut(B) be locally representable in R. Let sA and sB be strictly positive G-invariant elements in A and B, respectively. 54 EUSEBIO GARDELLA AND LUIS SANTIAGO (1) For every Cu(G)-semimodule morphism ρ : CuG(A, α) → CuG(B, β) satis- fying ρ([sA]G) ≤ [sB]G in CuG(B, β), there exists an equivariant ∗-homomorphism ψ : (A, α) → (B, β) with CuG(ψ) = ρ. Moreover, ψ is unique up to approx- imate unitary equivalence with G-invariant unitaries. (2) The actions α and β are conjugate if and only if there exists a Cu(G)- semimodule isomorphism ρ : CuG(A, α) → CuG(B, β) with ρ([sA]G) = [sB]G in CuG(B, β). Proof. (1). By Lemma 8.3, the crossed product A ⋊α G is again inductive limit of one-dimensional NCCW complexes with trivial K1-group. Moreover, the dual action bβ has the Rokhlin property by part (ii) of Proposition 4.4 in [Naw12]. Let ρ : CuG(A, α) → CuG(B, β) be a Cu(G)-semimodule homomorphism as in the state- ment. Using Theorem 5.14 and Proposition 5.16, ρ can be regarded as an equivari- ant morphism ρ : (Cu(A ⋊α G), Cu(bα)) → (Cu(B ⋊β G), Cu(bβ)). Since ρ([sA]G) ≤ [sB]G, we use part (i) in Theorem 3.13 of [GS16] to deduce that the morphism ρ lifts to a bG-equivariant ∗-homomorphism θ : (A ⋊α G,bα) → (B ⋊β G,bβ) that satisfies Cu(θ) = ρ. (Although Theorem 3.13 of [GS16] is stated for the functor Cu∼, the assumptions on A imply that we can replace Cu∼ with Cu.) Applying the crossed product functor, we obtain an equivariant ∗-homomorphism which, by Takai duality, is equivalent to an equivariant ∗-homomorphism ϕ : A ⊗ K(ℓ2(G)) → B ⊗ K(ℓ2(G)). bθ : (A ⋊α G) ⋊ bα bG → (B ⋊β G) ⋊ bβ bG, Observe that since e ∈ K(ℓ2(G)) is a rank-one projection, the element sA ⊗ e ∈ (A ⊗ K(ℓ2(G)))G represents the same class in CuG(A, α) as sA (see Corollary 5.7), and similarly for sB ⊗ e. It follows that ρ([sA ⊗ e]G) ≤ [sB ⊗ e]G. Under the canonical identification of B ⋊β G with the fixed point algebra of B ⊗ K(ℓ2(G)), we then get θ(sA ⊗ e) - sB ⊗ e in (B ⊗ K(ℓ2(G)))β⊗Ad(λ). Part (1) in Theorem 8.1 shows that there exists an equivariant ∗-homomorphism ϕ : (A, α) → (B, β) which induces ψ, and clearly CuG(ϕ) = ρ, as desired. The uniqueness statement follows from part (ii) of Theorem 3.13 in [GS16]. (2). The proof of this part is similar to the proof of the first part. Let ρ : CuG(A, α) → CuG(B, β) be a Cu(G)-semimodule homomorphism as in the statement. Using Theorem 5.14 and Proposition 5.16, ρ can be regarded as an equivariant morphism Since ρ([sA]G) = [sB]G, we use part (i) in Theorem 3.14 of [GS16] to deduce that ρ : (Cu(A ⋊α G), Cu(bα)) → (Cu(B ⋊β G), Cu(bβ)). that satisfies Cu(θ) = ρ. We obtain an equivariant ∗-isomorphism the morphism ρ lifts to a bG-equivariant ∗-isomorphism θ : (A ⋊α G,bα) → (B ⋊β G,bβ) bθ : (A ⋊α G) ⋊ bα bG → (B ⋊β G) ⋊ bβ bG, 55 which, by Takai duality, is equivalent to an equivariant ∗-isomorphism ϕ : A ⊗ K(ℓ2(G)) → B ⊗ K(ℓ2(G)). It follows that θ(sA⊗e) ∼ sB⊗e in (B⊗K(ℓ2(G)))β⊗Ad(λ). Part (2) in Theorem 8.1 shows that there exists an equivariant ∗-isomorphism ϕ : (A, α) → (B, β) which in- duces ψ, and clearly CuG(ϕ) = ρ, as desired. (cid:3) The assumptions in the above theorem can be relaxed to obtain more general conclusions: (1) For the conclusion in (1), one only needs to assume that B is separable, proof is in fact identical in this case. that B ⋊β G has stable rank one and that bβ has the Rokhlin property. The (2) If the conditions on ρ([sA⊗e]G) and [sB ⊗e]G are omitted, then one can pro- duce a β-cocycle ω : G → U(M (B)) and an equivariant ∗-homomorphism ϕ : (A, α) → (B, βω). (3) If the conditions on ρ([sA]G) and [sB]G are omitted, one has to replace the dynamical system (A, α) with (A ⊗ K(ℓ2(N)), α ⊗ idK(ℓ2(N))), and similarly with (B, β). References [ABP13] R. Antoine, J. Bosa, and F. Perera, The Cuntz semigroup of continuous fields, Indiana Univ. Math. J. 62 (2013), no. 4, 1105 -- 1131. [APS11] R. Antoine, F. Perera, and L. Santiago, Pullbacks, C(x)-algebras, and their Cuntz semi- group, J. Funct. Anal. 260 (2011), no. 10, 2844 -- 2880. [APT14] R. Antoine, F. Perera, and H. Thiel, Tensor products and regularity properties of Cuntz semigroups, Memoirs Amer. Math. Soc., to appear. (Preprint, arXiv:1410.0483.) 2014. [AS69] M. Atiyah and G. Segal, Equivariant K-theory and completion, J. Differential Geometry 3 (1969), 1 -- 18. [BTZ16] J. Bosa, G. Tornetta, and J. Zacharias, A Bivariant Theory for the Cuntz Semigroup, preprint arXiv:1602.02043, 2016. [BC09] N. Brown and A. Ciuperca, Isomorphism of Hilbert modules over stably finite C ∗- algebras, J. Funct. Anal. 257 (2009), no. 1, 332 -- 339. [BP08] N. Brown and A. Perera, F. Toms, The Cuntz semigroup, the Elliott conjecture, and dimension functions on C ∗-algebras, J. Reine Angew. Math. 621 (2008), 191 -- 211. [CEI08] K. Coward, G. Elliott, and C. Ivanescu, The Cuntz semigroup as an invariant for C ∗- algebras, J. Reine Angew. Math. 623 (2008), 161 -- 193. [CES11] A. Ciuperca, G. Elliott, and L. Santiago, On inductive limits of type-I C ∗-algebras with one-dimensional spectrum, Int. Math. Res. Not. IMRN 11 (2011), 2577 -- 2615. [CRS10] A. Ciuperca, L. Robert, and L. Santiago, The Cuntz semigroup of ideals and quotients and a generalized Kasparov stabilization theorem, J. Oper. Theory 64 (2010), no. 1, 155 -- 169. [Cun78] J. Cuntz, Dimension functions on simple C ∗-algebras, Math. Ann. 233 (1978), no. 2, 145 -- 153. [ERS11] G. Elliott, L. Robert, and L. Santiago, The cone of lower semicontinuous traces on a C ∗-algebra, Am. J. Math. 133 (2011), no. 4, 969 -- 1005. [Fol95] G. Folland, A course in abstract harmonic analysis, Studies in Advanced Mathematics, CRC Press, 1995. [Gar14a] E. Gardella, Circle actions on UHF-absorbing C ∗-algebras. Houston J. of Math, to appear. (Preprint, arXiv:1406.4198.) 2014. [Gar14b] , Classification theorems for circle actions on Kirchberg algebras, I, preprint, arXiv:1405.2469, 2014. [Gar14c] , Crossed products by compact group actions with the Rokhlin property. J. Non- comm. Geom., to appear. (Preprint, arXiv:1408.1946.) 2014. [Gar14d] , Rokhlin dimension for compact group actions, Indiana Univ. Math., to appear. (Preprint, arXiv:1407.1277.) 2014. 56 EUSEBIO GARDELLA AND LUIS SANTIAGO [Gar15a] [Gar15b] , Compact group actions with the Rokhlin property., In preparation, 2015. , Regularity properties and Rokhlin dimension for compact group actions, Hous- ton J. of Math., to appear. (Preprint, arXiv:1407.5485.) 2015. [GHS16] E. Gardella, I. Hirshberg, and L. Santiago, Rokhlin dimension: tracial properties and [GS16] crossed products, in preparation, 2016. E. Gardella and L. Santiago, Equivariant *-homomorphisms, Rokhlin constraints and equivariant UHF-absorption, J. Funct. Anal. 270 (2016), no. 7, 2543-2590. [HR85] D. Handelman and W. Rossmann, Actions of compact groups on AF C ∗-algebras, Illinois J. Math. 29 (1985), 51 -- 95. [Jac13] B. Jacelon, A simple, monotracial, stably projectionless C ∗-algebra, J. Lond. Math. Soc. [Jul81] (2) 87 (2013), no. 2, 365 -- 383. P. Julg, K-th´eorie ´equivariante et produits croises., C. R. Acad. Sci., Paris, S´er. I 292 (1981), 629 -- 632. [Kas80] G. Kasparov, Hilbert C ∗-modules: theorems of Stinespring and Voiculescu, J. Operator Theory 4 (1980), no. 1, 133 -- 150. [Kat85] Y. Katayama, Takesaki's duality for a non-degenerate co-action, Math. Scand. 55 (1985), 141 -- 151. [KOQ15] S. Kaliszewski, T. Omland and J. Quigg, Three versions of categorical crossed-product duality, New York J. Math. 22 (2016) 293 -- 339. [KR02] E. Kirchberg and M. Rørdam, Infinite non-simple C ∗-algebras: absorbing the Cuntz algebra O∞, Adv. Math. 167 (2002), no. 2, 195 -- 264. [Lan95] E. Lance, Hilbert C ∗-modules, London Mathematical Society Lecture Note Series, vol. 210, Cambridge University Press, 1995, A toolkit for operator algebraists. [Naw12] N. Nawata, Finite group actions on certain stably projectionless C ∗-algebras with the Rokhlin property, Trans. Amer. Math. Soc. 368 (2016), no. 1, 471-493. [Ped99] Gert K. Pedersen, Pullback and pushout constructions in C ∗-algebra theory, J. Funct. Anal. 167 (1999), no. 2, 243 -- 344. [Phi87] N. C. Phillips, Equivariant K-theory and freeness of group actions on C ∗-algebras, [PT07] Lecture Notes in Mathematics, 1274. Berlin etc.: Springer-Verlag. VIII, 1987. F. Perera and A. Toms, Recasting the Elliott conjecture, Math. Ann. 338 (2007), no. 3, 669 -- 702. [Raz02] S. Razak, On the classification of simple stably projectionless C*-algebras, Canad. J. Math. 54 (2002), 138 -- 224. [Rob12] L. Robert, Classification of inductive limits of 1-dimensional NCCW complexes, Adv. [Ros79] Math. 231 (2012), no. 5, 2802 -- 2836. J. Rosenberg, Appendix to O. Bratteli's paper on "Crossed products of UHF algebras", Duke Math. J. 46 (1979), no. 10, 25 -- 26. [Seg68] G. Segal, Equivariant K-theory, Publ. Math. Inst. Hautes Etudes Sci. 34 (1968), 129 -- 151. [Tom08] A. Toms, On the classification problem for nuclear C ∗-algebras, Annals Math. 167 (2008), 1029 -- 1044. [Tor16] G. Tornetta, An equivariant theory for the bivariant Cuntz semigroup, preprint, 2016. [ZZ08] G. Zhang and R. Zhang, Equivariant vector bundles on quantum homogeneous spaces, Math. Res. Lett. 15 (2008), no. 2, 297 -- 307. Westfalische Wilhelms-Universitat Munster, Fachbereich Mathematik, Einstein- strasse 62, 48149 Munster, Germany E-mail address: [email protected] Institute of Mathematics, University of Aberdeen, Fraser Noble Building, Aberdeen AB24 3UE, UK E-mail address: [email protected] URL: http://homepages.abdn.ac.uk/lmoreno/pages/index.html
1503.02708
1
1503
2015-03-09T21:50:40
Hilbert modules over a planar algebra and the Haagerup property
[ "math.OA", "math.CT", "math.FA", "math.QA" ]
Given a subfactor planar algebra P and a Hilbert P-module of lowest weight 0 we build a bimodule over the symmetric enveloping inclusion associated to P. As an application we prove diagrammatically that the Temperley-Lieb-Jones standard invariants have the Haagerup property. This provides a new proof of a result due to Popa and Vaes.
math.OA
math
HILBERT MODULES OVER A PLANAR ALGEBRA AND THE HAAGERUP PROPERTY by Arnaud Brothier1 and Vaughan Jones 2 Abstract. Given a subfactor planar algebra P and a Hilbert P-module of lowest weight 0 we build a bimodule over the symmetric enveloping inclusion associated to P. As an application we prove diagrammatically that the Temperley-Lieb-Jones standard invariants have the Haagerup property. This provides a new proof of a result due to Popa and Vaes. 1. Introduction and main results Popa initiated the study of approximation properties of subfactors in [Pop86, Pop94a, Pop94b]. To any finite index subfactorof type II1 one can associate a combinatorial object called the stan- dard invariant. This invariant has been axiomatized as a paragroup, a λ-lattice, and a planar algebra respectively by Ocneanu, Popa, and the second author [Ocn88, Pop95, Jonb]. An ana- logue of quantum doubles for a subfactor was introduced by Ocneanu, Longo and Rehren, and Popa( [Ocn88, LR95, Pop94b]). The latter construction is called the symmetric enveloping in- clusion. For the construction of subfactors of Guionnet et al. in [CJS14] gave a diagrammatic description of the symmetric enveloping inclusion. in [GJS10], Curran et al. Recently, Popa and Vaes introduced a representation theory for subfactors and standard invariants [PV]. They defined the Haagerup property for a subfactor and showed that it depends only on its standard invariant. They then showed that the Temperley-Lieb-Jones standard invariants have the Haagerup property. (Note, this result was already announced in [Pop06, Remark 3.5.5].) Their proof uses previous work on discrete quantum groups and the equivalence between the bimodule category associated to the Temperley-Lieb-Jones standard invariant and the representation category of the quantum group PSUq(2) [DCFY14]. Here we give another proof: Theorem 1.1. The Temperley-Lieb-Jones standard invariant has the Haagerup property for any loop parameter δ ∈ {2 cos π n , n > 3} ∪ [2 : ∞). Our proof only uses planar algebra technology. The idea is that the lowest weight zero annular representations of the planar algebra immediately give "compact" bimodules (which are obvious in the Curran et al. pictures), which tend to the trivial bimodule in the way required by the Popa-Vaes definiton of the Haagerup property. Acknowledgement. The first author thanks Dietmar Bisch for many encouragements. He also thanks Jesse Peterson and Jean-Louis Lhuillier for their patience and guidance. 2. Preliminaries 2.1. The symmetric enveloping inclusion associated to a subfactor planar algebra. We refer to [Jonb] for more details about planar algebras. We recall the construction of [CJS14, Section 2]. Note, we define the symmetric enveloping inclusion via the product introduced in [CJS14, Section 2.1] that we call the Bacher product. Let P = (P ± n , n > 0) be a subfactor 1Vanderbilt University, Department of Mathematics, 1326 Stevenson Center Nashville, TN, 37212, USA, [email protected] 2Vanderbilt University, Department of Mathematics, 1326 Stevenson Center Nashville, TN, 37212, USA, [email protected] 1 2 HILBERT MODULES OVER A PLANAR ALGEBRA AND THE HAAGERUP PROPERTY planar algebra. For any k, n, m > 0, let Dk(n, m) be a copy of the vector space P + n+m+2k. We decorate strings with natural numbers to indicate that they represent a given number of parallel strings. The distinguished interval of a box is decorated by a dollar sign if it is not at the top left corner. We will omit unnecessary decorations. Consider the direct sum that we equipped with the Bacher product: GrkP ⊠ GrkP := M n,m>0 Dk(n, m) x ⋆k y = min(2n,2i) min(2m,2j) X a=0 X b=0 2k x a 2k b y 2k where x ∈ Dk(n, m) and y ∈ Dk(i, j). Let † : GrkP ⊠ GrkP −→ GrkP ⊠ GrkP be the anti-linear involution that sends Dk(n, m) to itself and satisfies $ x† = $ x∗ , for any x ∈ Dk(n, m). Consider the linear form τ : GrkP ⊠ GrkP −→ C, which is zero unless n = m = 0 and sends the unit of Dk(0, 0) to 1. We have that (GrkP ⊠ GrkP, ⋆k, †, τ ) is an associative ∗-algebra with a faithful tracial state [CJS14, Corollary 2.3]. Further, GrkP ⊠ GrkP acts by bounded operators on the Gelfand-Naimark-Segal Hilbert space for τ [CJS14, Theorem 2.1]. Let Mk ⊠ Mk be its Gelfand-Naimark-Segal completion which is a factor of type II1. Let Mk be the von Neumann subalgebra of Mk ⊠ Mk generated by elements of the form ∈ Dk(n, 0), n > 0. x k and commutes with Mk. We identify M op Observe, the von Neumann subalgebra of Mk ⊠Mk generated by the family of sets Dk(0, m), m > 0 is isomorphic to M op k with this von Neumann k subalgebra. Note, we have a unital inclusion of Mk−1 in Mk by adding two horizontal strings under elements of Mk−1. By [GJS10], Mk−1 ⊂ Mk is a subfactor of type II1 with standard invariant isomorphic to the subfactor planar algebra P or its opposite depending on the parity of k. The von Neumann subalgebra of Mk ⊠ Mk generated by Mk and M op is isomorphic to Mk⊗M op k and the inclusion Mk ∨ M op k ⊂ Mk ⊠ Mk k is isomorphic to Popa's symmetric enveloping inclusion associated to the subfactor Mk−1 ⊂ Mk for any k > 1 [CJS14]. We denote by M ⊗M op ⊂ M ⊠ M the inclusion M0 ∨ M op 0 ⊂ M0 ⊠ M0 and call it the symmetric enveloping inclusion associated to P. Similarly, we write D0(n, m) = D(n, m) for any n, m > 0. 2.2. Hilbert modules over a subfactor planar algebra. We introduce notations and termi- nology regarding Hilbert modules over a subfactor planar algebra. We refer to [Jona] for more de- tails. Let us fix a subfactor planar algebra P. An annular tangle α is a tangle in P with the choice of a distinguished internal disc. We write AnnP((m, ε), (n, ǫ)) the complex vector space spanned by annular tangles with 2n (resp. 2m) boundary points on its internal (resp. external) disc and HILBERT MODULES OVER A PLANAR ALGEBRA AND THE HAAGERUP PROPERTY 3 where the dollar sign is in a region with shading ǫ (resp. ε). A tangle in AnnP((m, ε), (n, ǫ)) is called a ((m, ε), (n, ǫ))-annular tangle. Let AP = (AP((m, ε), (n, ǫ)), n, m > 0, ǫ, ε ∈ ±) be the annular algebroid associated to P. We denote by α 7−→ α† the anti-linear involution which sends a ((m, ε), (n, ǫ))-annular tangle to a ((n, ǫ), (m, ε))-annular tangle by reflection in a circle half way between the inner and outer boundaries. A Hilbert P-module is a graded vector space V = (V ± n is a finite dimensional Hilbert space, AP acts on V , and the inner product is compatible with this action. It means that if α ∈ AP((m, ε), (n, ǫ)) then it defines a linear map from V ǫ n , n > 0), where each V ± n to V ε m such that hα(v), wi = hv, α†(w)i, for any v ∈ V ǫ n , w ∈ V ε m. The lowest weight of a Hilbert P-module V is the smallest natural number n such that V + n 6= {0}. 2.3. Hilbert T LJ -modules of lowest weight 0. Consider the Temperley-Lieb-Jones planar algebra P = T LJ with loop parameter δ > 2. Irreducible Hilbert T LJ-modules of lowest weight 0 have been fully classified in [Jona] and in [GL98] for the unshaded case. For any 0 < t 6 δ there exists a Hilbert T LJ-module V (t) = (V (t)± 0 is one dimensional and spanned by a unit vector ξ(t) which satisfies n , n > 0) such that V (t)+ hα(ξ(t)), β(ξ(t))i = δct2d, where α, β are annular tangles, c is the number of contractible circles in the (±, ±)-annular tangle β† ◦ α and d is half the number of non-contractible ones. Those Hilbert T LJ-modules will be used to construct unital completely positive maps on the symmetric enveloping inclusion associated to the Temperley-Lieb-Jones planar algebra. 3. Hilbert P -modules give (M ⊗M op ⊂ M ⊠ M )-bimodules Let V = (V ± n , n > 0) be a Hilbert P-module of lowest weight 0. For i, j > 0, let Hi,j be a copy of the Hilbert space V + i+j. Let H = Li,j>0 Hi,j be the Hilbert space equal to the direct sum of the Hi,j. In particular, Hi+1,j−1 is orthogonal to Hi,j in H. Consider the dense pre-Hilbert subspace K ⊂ H spanned by the union of all Hi,j. We put π0(x)ξ = X a,b x a b ξ , for any x ∈ D(n, m) ⊂ GrP ⊠ GrP and ξ ∈ Hi,j. This defines a representation where L(K) is the algebra of endomorphism of the vector space K. π0 : GrP ⊠ GrP −→ L(K), Proposition 3.1. For any x ∈ GrP ⊠ GrP, π0(x) defines a bounded operator on H. Further, the representation π0 extends to a normal ∗-representation π : M ⊠ M −→ B(H). Proof. Consider x in GrP ⊠ GrP. We can prove that π0(x) defines a bounded operator by following a similar argument than [JSW10, Theorem 3.3]. We continue to denote by π0(x) its extension to H. Let ξ ∈ H0,0 be a unit vector and let ωξ be its associated vector state. Note, ωξ ◦ π0(x) = τ (x) for any x ∈ GrP ⊠ GrP, where τ is the unique normal tracial state on M ⊠ M . Therefore, π0 extends to a normal ∗-representation π : M ⊠ M −→ B(H). (cid:3) Recall, if T ⊂ S is an inclusion of von Neumann algebras, then a Hilbert (T ⊂ S)-module is a couple (H, ξ) such that H is a Hilbert S-module and ξ is a T -central vector of H. 4 HILBERT MODULES OVER A PLANAR ALGEBRA AND THE HAAGERUP PROPERTY Corollary 3.2. Let V be a Hilbert P-module of lowest weight 0. Consider the Hilbert space H constructed above and let ξ ∈ H0,0 be a unit vector. Then, (H, ξ) has a structure of Hilbert (M ⊗M op ⊂ M ⊠ M )-bimodule where the left action is given by π and the right action is defined similarly. Proof. Proposition 3.1 implies that H is a M ⊠ M -bimodule with the action described above. Consider x⊗yop ∈ GrP ⊗GrP op, where GrP ⊗GrP op = M ⊗M op ∩GrP ⊠GrP. Since ξ ∈ H0,0, we have (x ⊗ yop) · ξ = x y ξ and ξ · (x ⊗ yop) = ξ x y . Those two pictures are isotopic to each other. Therefore, (x ⊗ yop) · ξ = ξ · (x ⊗ yop). By density of GrP ⊗ GrP op inside M ⊗M op, we obtain that ξ is a M ⊗M op-central vector. (cid:3) 4. The Temperley-Lieb-Jones standard invariant has the Haagerup property In this article, any inclusion of tracial von Neumann algebras will be supposed to be unital and tracial. We recall the definition of the relative Haagerup property due to Boca [Boc93]. Note, Popa defined a very similar property [Pop06]. Those two definitions coincide in the context of Definition 4.2. Definition 4.1. Consider an inclusion of tracial von Neumann algebras N ⊂ (M, τ ). A completely positive approximation of the identity (CPAI) for N ⊂ (M, τ ) is a sequence of normal N -bimodular trace-preserving unital completely positive maps (ϕl : M −→ M, l > 0) such that kϕl(x) − xk2 −→l 0, for any x ∈ M, and the unique continuous extension Θl ∈ B(L2(M, τ )) of ϕl to L2(M, τ ) is in the compact ideal space of hM, eN i. If such a sequence exists we say that N ⊂ (M, τ ) has the relative Haagerup property. Definition 4.2. [PV] A subfactor N ⊂ M has the Haagerup property if its symmetric en- veloping inclusion has the relative Haagerup property. A standard invariant G has the Haagerup property if there exists a subfactor N ⊂ M with standard invariant isomorphic to G which has the Haagerup property. Recall that if two subfactors have isomorphic standard invariants, then one of them has the Haagerup property if and only if the other one has the Haagerup property, see [Pop06, Remark 3.5.5] or [PV]. Lemma 4.3. Let P be a subfactor planar algebra. Then P has the Haagerup property if and only if its associated symmetric enveloping inclusion M ⊗M op ⊂ M ⊠ M has the relative Haagerup property. Proof. Consider the subfactor M0 ⊂ M1 defined in Section 2. Its planar algebra is equal to P. Popa's symmetric enveloping inclusion associated to M0 ⊂ M1 is isomorphic to M1∨M op 1 ⊂ M1⊠ M1. Consider the inclusion M0 ∨M op 0 ⊂ M1 ⊠M1. Let e be the Jones projection e = 1 Note, the compression e(M0 ∨ M op 0 )e ⊂ e(M1 ⊠ M1)e is isomorphic to M ⊗M op ⊂ M ⊠ M . Therefore, by [Pop06, Proposition 2.3 and Proposition 2.4], M ⊗M op ⊂ M ⊠ M has the relative Haagerup property if and only if M1 ∨ M op 1 ⊂ M1 ⊠ M1 has the relative Haagerup property. (cid:3) . δ HILBERT MODULES OVER A PLANAR ALGEBRA AND THE HAAGERUP PROPERTY 5 Lemma 4.4. Let T LJ be the Temperley-Lieb-Jones planar algebra with a loop parameter δ > 2 and let M ⊗M op ⊂ M ⊠ M be its associated symmetric enveloping inclusion. Consider the 2nth-Jones-Wenzl idempotent gn ∈ TLJ+ 2n that we identity with its associated element in D(n, n) ⊂ M ⊠ M. Let Ln ⊂ L2(M ⊠ M ) be the M ⊗M op-bimodule generated by gn. Then Ln op, where Xn is the irreducible M0-bimodule corresponding the the 2nth is isomorphic to Xn⊗Xn vertex in the principal graph of the subfactor M0 ⊂ M1. Further, L2(M ⊠ M ) is equal to the direct sum of the bimodule Ln. Proof. We follow an argument in [CJS14, pp. 120-122]. Let us show that Ln is orthogonal to Lm if n 6= m. This is equivalent to show that for any x, y ∈ T LJ, we have xgny ⊥ gm in the planar algebra T LJ. But this is obvious. Observe, the ∗-algebra GrP ⊠ GrP is generated by the set of Jones-Wenzl idempotents and GrP ⊗ GrP op. Therefore, L2(M ⊠ M ) is equal to the direct sum of the bimodules Ln. Consider the M -bimodule Xn ⊂ L2(Mn) equal to the image of gn viewed as an element of T LJ + 2n = M ′ ∩ M2n ⊂ B(L2(Mn)). We have an isomorphism from Xn⊗Xn onto Ln given by the tangle which connects the 2n side strings of an elements of Xn op (resp. Xn ) to the top strings of gn (resp. the bottom strings of gn). (cid:3) op Theorem 4.5. Let T LJ be the Temperley-Lieb-Jones planar algebra with any loop parameter δ ∈ {2 cos( π n ), n > 3} ∪ [2 : ∞). Then T LJ has the Haagerup property. Proof. If δ = 2 cos( π n ) for some n > 3, then T LJ has finite depth. Therefore, its symmetric enveloping inclusion is a subfactor of finite index. This implies that T LJ has the Haagerup property. We assume that δ > 2. We write T = M ⊗M op and S = M ⊠ M . Consider 0 < t < δ and the pointed Hilbert T LJ-module (V (t), ξ(t)) of section 2.3 where ξ(t) ∈ V (t)+ 0 is a unit vector. Let (H t, ξt) be its associated (T ⊂ S)-bimodule as constructed in section 3. Let Zt : L2(S) −→ H t be the continuous linear map densely defined as follows Zt(xΩ) = ξt ·x, for any x ∈ S. Define the normal T -bimodular unital completely positive map φt : S −→ S by the formula φt(x) = Z ∗ t πt(x)Zt, where πt : S −→ B(H t) is the left action of S on H t. We will show that the net (φt, 0 < t < δ) is the desired approximation of the identity. Note, the T -bimodules Ln are isomorphic to Xn⊗X op n for any n > 0. Hence, they are irreducible and pairwise non-isomorphic. By Schur's Lemma, there exists a scalar valued function ct : N −→ C such that Θt = Pn>0 ct(n)sn, where Θt is the unique continuous extension of φt to L2(S) and sn is the orthogonal projection from L2(S) onto Ln. We have the formula ct(n) = hφt(gn), gni hgn, gni , for any n > 0. Let τ2n be the non-normalized trace of the C∗-algebra T LJ + 0. Let q be the unique real number bigger than 1 satisfying q + q−1 = δ. It is well known that τ2n(gn) = [2n + 1]q, where 2n. Remark, τ2n(gn) = hgn, gni, for any n > [2n + 1]q = q2n+1 − q−2n−1 q − q−1 is the 2n+1th quantum integer with parameter q [Jon83, Section 5.1]. We claim that (1) hφt(gn), gni = [2n + 1]ω, if n > 1, 6 HILBERT MODULES OVER A PLANAR ALGEBRA AND THE HAAGERUP PROPERTY where ω is a complex number satisfying ω + ω−1 = t. Observe, hφt(gn), gni = hgn · ξt, ξt · gni = hgn · ξt · gn, ξti = h gn ξt gn , ξti = h gn ξt , ξti. Hence, proving the equality (1) is a routine computation using the induction formula of [Jon83, Section 5.1] or [Wen87]. Therefore, Observe, ct(n) = [2n + 1]ω [2n + 1]q for any n > 0. ct(n) −→ 1, as t → δ, for any n > 0, and ct(n) −→ 0, as n → ∞, for any 0 < t < δ. Note, τ ◦ φt = a0τ = τ. Hence, any sequence of real numbers (0 < tn < δ, n > 0) that converges to δ defines a CPAI (φtn , n > 0). Therefore, T ⊂ S has the relative Haagerup property. (cid:3) References [Boc93] [CJS14] F. Boca. On the method of constructing irreducible finite index subfactors of Popa. Pacific. J. Math, 161(2):201 -- 231, 193. S. Curran, V.F.R. Jones, and D. Shlyakhtenko. On the symmetric enveloping algebra of planar algebra subfactors. Trans. Amer. Math. Soc, 366(1):113 -- 133, 2014. [DCFY14] K. De Commer, A. Freslon, and M. Yamashita. CCAP for universal discrete quantum groups. Comm. [GJS10] [GL98] Math. Phys., 331:677 -- 701, 2014. A. Guionnet, V.F.R. Jones, and D. Shlyakhtenko. Random matrices, free probability, planar algebras and subfactor. Quanta of maths: Non-commutative Geometry Conference in Honor of Alain Connes, in Clay Math. Proc., 11:201 -- 240, 2010. J.J. Graham and G.I. Lehrer. The representation theory of affine Temperley-Lieb algebras. Enseign. Math., 44(2):173 -- 218, 1998. V.F.R. Jones. The annular structure of subfactors. Preprint. arXiv:0105.071. V.F.R. Jones. Planar algebras I. Preprint. arXiv:9909.027. V.F.R. Jones. Index for subfactors. Invent. Math, 72:1 -- 25, 1983. [Jona] [Jonb] [Jon83] [JSW10] V.F.R. Jones, D. Shlyakhtenko, and K. Walker. An orthogonal approach to the subfactor of a planar [LR95] [Ocn88] [Pop86] [Pop94a] [Pop94b] [Pop95] [Pop06] [PV] algebra. Pacific J. Math., 246:187 -- 197, 2010. R. Longo and K. Rehren. Nets of subfactors. Rev. Math. Phys., 7:567 -- 597, 1995. A. Ocneanu. Quantized groups, string algebras and Galois theory for algebras. Operator algebras and applications, London Math. Soc. Lecture Note Ser., 136:119 -- 172, 1988. S. Popa. Correspondences. INCREST, 1986. S. Popa. Classification of amenable subfactors of type II. Acta. Math., 172:163 -- 255, 1994. S. Popa. Symmetric enveloping algebras, amenability and AFD properties for subfactors. Math. Res. Lett., 1:409 -- 425, 1994. S. Popa. An axiomatization of the lattice of higher relative commutants of a subfactor. Invent. Math., 120(3):427 -- 445, 1995. S. Popa. On a class of type II1 factors with Betti numbers invariants. Ann. of Math., 163:809 -- 889, 2006. S. Popa and S. Vaes. Representation theory for subfactors, λ-lattices and C∗-tensor categories. Preprint. arXiv:1412.2732. [Wen87] H. Wenzl. On sequences of projections. C.R. Math. Acad. Sci. Soc. R. Can., 9(1):5 -- 9, 1987.
1707.09257
2
1707
2017-10-06T19:46:03
Classification of $L^p$ AF algebras
[ "math.OA" ]
We define spatial $L^p$ AF algebras for $p \in [1, \infty) \setminus \{ 2 \}$, and prove the following analog of the Elliott AF algebra classification theorem. If $A$ and $B$ are spatial $L^p$ AF algebras, then the following are equivalent: 1) $A$ and $B$ have isomorphic scaled preordered $K_0$-groups. 2) $A \cong B$ as rings. 3) $A \cong B$ (not necessarily isometrically) as Banach algebras. 4) $A$ is isometrically isomorphic to $B$ as Banach algebras. 5) $A$ is completely isometrically isomorphic to $B$ as matrix normed Banach algebra. As background, we develop the theory of matrix normed $L^p$ operator algebras, and show that there is a unique way to make a spatial $L^p$ AF algebra into a matrix normed $L^p$ operator algebra. We also show that any countable scaled Riesz group can be realized as the scaled preordered $K_0$-group of a spatial $L^p$ AF algebra.
math.OA
math
CLASSIFICATION OF Lp AF ALGEBRAS N. CHRISTOPHER PHILLIPS AND MARIA GRAZIA VIOLA Abstract. We define spatial Lp AF algebras for p ∈ [1, ∞) \ {2}, and prove the following analog of the Elliott AF algebra classification theorem. If A and B are spatial Lp AF algebras, then the following are equivalent: • A and B have isomorphic scaled preordered K0-groups. • A ∼= B as rings. • A ∼= B (not necessarily isometrically) as Banach algebras. • A is isometrically isomorphic to B as Banach algebras. • A is completely isometrically isomorphic to B as matrix normed Banach algebras. As background, we develop the theory of matrix normed Lp operator algebras, and show that there is a unique way to make a spatial Lp AF algebra into a matrix normed Lp operator algebra. We also show that any countable scaled Riesz group can be realized as the scaled preordered K0-group of a spatial Lp AF algebra. 1. Introduction In a well known paper [7] of 1976, Elliott gave a complete classification of ap- proximately finite dimensional (AF) C*-algebras. He showed that two AF C*- algebras A1 and A2 are isomorphic if and only if their scaled preordered K0-groups (cid:0)K0(A1), K0(A1)+, Σ(A1)(cid:1) and (cid:0)K0(A2), K0(A2)+, Σ(A2)(cid:1) are isomorphic. More- over, the work of Effros, Handelman, and Shen showed (see [5] and [6]) that any countable scaled Riesz group (G, G+, Σ) can be realized as the scaled preordered K0-group of an AF C*-algebra. In a series of papers (see [16], [17], [18], and [19]), the first author introduced and studied Lp analogs of the uniformly hyperfinite (UHF) algebras and Lp analogs of the Cuntz algebras. One result of [17] is that two spatial Lp UHF algebras are isomorphic if and only if they have the same supernatural number. This result is analogous to the result of Glimm [10], that two UHF C*-algebras are isomorphic if and only if they have the same supernatural number. (This is a special case, done earlier, of Elliott's AF classification theorem.) It is therefore natural to ask if there are Lp analogs of AF algebras which can be classified by their scaled preordered K0 groups. In this paper, we show that the algebras that we call the spatial Lp AF algebras provide a positive answer In Theorem 10.20, we show that two spatial Lp AF algebras to this question. are completely isometrically isomorphic (as matricial Lp operator algebras) if and only if their scaled ordered K0 groups are isomorphic. We further show that, Date: 31 July 2017. 2000 Mathematics Subject Classification. Primary 47L10; Secondary 46L35. The first author was partially supported by the US National Science Foundation under Grants DMS-1101742 and DMS-1501144. The second author was supported by a Natural Sciences and Engineering Research Council Discovery Grant. 1 2 N. CHRISTOPHER PHILLIPS AND MARIA GRAZIA VIOLA as in the C*-algebra case, given any scaled countable Riesz group (G, G+, Σ), there exists a spatial Lp AF algebra A such that the scaled preordered K0 group (cid:0)K0(A), K0(A)+, Σ(A)(cid:1) is isomorphic to (G, G+, Σ). We also show that spatial Lp AF algebras have unique Lp matrix norms. We don't list examples. Theorem 10.17 and Theorem 10.22 show that for each p ∈ [1, ∞) there is a one to one correspondence between isomorphism classes of AF C*-algebras and spatial Lp AF algebras, so the examples are "the same". Although we don't address this issue here, constructions like the C*-algebra of a locally finite discrete abelian group, which give AF C*-algebras, give Lp operator algebras which are AF in some sense but are not spatial Lp AF algebras. In a forthcoming paper we will prove that the ideal structure of a spatial Lp AF algebra is determined by K-theory in the same way as for an AF C*-algebra. We will prove that, like a C*-algebra, a spatial Lp AF algebra is incompressible in the sense that any contractive homomorphism to some other Banach algebra can be factored as a quotient map followed by an isometric homomorphism. (In particular, contractive injective homomorphisms from spatial Lp AF algebras are isometric.) We will also study the isometries and automorphisms of a spatial Lp AF algebra. The results will be quite different from what happens with AF C*-algebras. A spatial Lp AF algebra is the direct limit of a direct system of semisimple finite dimensional Lp operator algebras in which the connecting maps are contractive ho- momorphisms having the property that the image of the identity is a spatial partial isometry in the sense of Definition 6.4 of [16]. In the context of Lp operator alge- bras, where in general we do not require the homomorphisms between Lp operator algebras to be unital, that is the best possible form for our maps. To make sense of uniqueness of Lp matrix norms and completely isometric iso- morphism, we develop the basics of the theory of matrix normed Banach algebras and matricial Lp operator algebras. The arguments used for the classification of spatial Lp AF algebras are similar to the ones used for the classification of AF C*-algebras. However, to be able to carry out these arguments, background material needs to be developed. Much of it is fairly elementary, and for this part the novelty is putting it together in the right way. There are several somewhat more substantial ingredients, including a structure theorem for contractive representations of C(X) on an Lp space (Theorem 4.5), the recognition that, in connection with nonunital maps between unital algebras, idempotents must be required to be hermitian (contractivity is not good enough; see Section 6), and what to require of approximate identities of idempotents in order to get a unique suitable norm on the unitization (see Proposition 9.9). We also have to prove that the direct limit of Lp operator algebras is again an Lp operator algebra. The paper is organized as follows. In Section 2 we recall Lp operator algebras and give some preliminary results on their representations on Lp-spaces. In Section 3 we introduce matricial (matrix normed) Lp operator algebras and discuss their representations on Lp spaces. This material is needed to define Lp operator algebras that have unique Lp matrix norms, which we examine in Section 4. Most of the Lp operator algebras in this article have unique Lp matrix norms, including the matrix algebra M p n and the algebra C(X) for a compact metric space X. Sections 5 and 7 deal with direct sums and direct limits of (matricial) Lp operator algebras, while Section 6 contains material on hermitian idempotents, including a CLASSIFICATION OF Lp AF ALGEBRAS 3 characterization of hermitian idempotents in an Lp operator algebra in terms of multiplication operators. In Section 8 we introduce our building blocks (the spatial semisimple finite di- mensional Lp operator algebras), and the appropriate homomorphisms between them, the spatial homomorphisms. We characterize spatial homomorphisms in terms of block diagonal homomorphisms. In Section 9 we define spatial Lp AF al- gebras, show that every spatial Lp AF algebra is an Lp operator algebra as in [16], and that it has unique Lp matrix norms. Section 10 contains our main result. We give a complete classification of spatial Lp AF algebras using the scaled preordered K0 group, and show that, as in the C*-algebra case, any countable scaled Riesz group can be realized as the scaled preordered K0 group of a spatial Lp AF algebra. Shortly after posting this paper on the airXiv, E. Gardella informed us of his work with Lupini on the uniqueness of the matricial norm structure for Lp analogs of groupoid C*-algebras. (See [9].) Spatial Lp AF algebras are examples of such algebras; see Subsection 7.2 in [9]. We were unaware of the work of Gardella and Lupini while preparing this manuscript and we refer the reader to their paper for a different proof of the uniqueness of the Lp matrix norms. We use the following standard notation throughout the paper. Notation 1.1. If E is a Banach space, then L(E) denotes the Banach algebra of all bounded linear operators on E, with the operator norm. Notation 1.2. If (X, B, µ) is a measure space, and E ⊂ X is measurable, then µE denotes the measure on E gotten by restricting µ to the σ-algebra of measurable subsets of E. We also recall that an idempotent in a ring is an element e satisfying e2 = e. 2. Lp operator algebras In this section we define Lp operator algebras, and state some of the standard results about Lp operator algebras and their representations. These results are basic for the rest of the paper. The following definitions are based on Definition 1.1 and Definition 1.17 of [19]. Definition 2.1. Let p ∈ [1, ∞). An Lp operator algebra is a Banach algebra such that there exists a measure space (X, B, µ) and an isometric isomorphism from A to a norm closed subalgebra of L(Lp(X, µ)). Definition 2.2. Let p ∈ [1, ∞). (1) A representation of an Lp operator algebra A (on Lp(Y, ν)) is a continuous homomorphism π : A → L(Lp(Y, ν)) for some measure space (Y, C, ν). (2) The representation π is contractive if kπ(a)k ≤ kak for all a ∈ A, and isometric if kπ(a)k = kak for all a ∈ A. (3) We say that the representation π : A → L(Lp(Y, ν)) is separable if Lp(Y, ν) is separable, and that A is separably representable if it has a separable isometric representation. (4) We say that π is σ-finite if ν is σ-finite, and that A is σ-finitely representable if it has a σ-finite isometric representation. (5) We say that π is nondegenerate if π(A)Lp(Y, ν) = span(cid:0)(cid:8)π(a)ξ : a ∈ A and ξ ∈ Lp(Y, ν)(cid:9)(cid:1) 4 N. CHRISTOPHER PHILLIPS AND MARIA GRAZIA VIOLA is dense in Lp(Y, ν). We say that A is nondegenerately (separably) repre- sentable if it has a nondegenerate (separable) isometric representation, and nondegenerately σ-finitely representable if it has a nondegenerate σ-finite isometric representation. The following fact about the restriction of an operator looks obvious (and the proof is easy), but it is the sort of statement that should not be taken for granted outside of the context of C*-algebras. The condition kf k = 1 is necessary; see Example 2.5. Lemma 2.3. Let E be a Banach space, let a ∈ L(E), and let f ∈ L(E) be an idempotent with kf k = 1 such that af = a. Then kaf Ek = kak. Proof. It is obvious that kaf Ek ≤ kak. For the reverse inequality, let ε > 0, choose ξ ∈ E such that kξk ≤ 1 and kaξk > kak − ε, and set η = f ξ. Then η ∈ f E, kηk ≤ 1, and kaηk = kaξk > kak − ε. (cid:3) The main application of Lemma 2.3 is the next result, which we will use repeat- edly in the following sections. Corollary 2.4. Let A be a unital Banach algebra in which k1k = 1. Let E be a Banach space, and let π : A → L(E) be a nonzero representation. Set F = π(1)E. Then there is a unital representation π0 : A → L(F ) such that π0(a)ξ = π(a)ξ for all a ∈ A and ξ ∈ F . If kπ(1)k = 1, then π0 is contractive if and only if π is contractive and π0 is isometric if and only if π is isometric. Proof. The existence of π0 follows from the equation π(1)π(a)π(1) = π(a) for all If π is contractive then kπ(1)k ≤ 1. If kπ(1)k = 1, taking f = π(1) in a ∈ A. Lemma 2.3 gives kπ0(a)k = kπ(a)k for all a ∈ A. (cid:3) Example 2.5. Lemma 2.3 fails without kf k = 1 and Corollary 2.4 fails without k1k = 1. Take f ∈ L(E) to be any idempotent with kf k > 1. For example, take 0 0 ). Then kf k = 21−1/p. For p ∈ (1, ∞), take E to be C2 with k ·kp, and take f = ( 1 1 Lemma 2.3 take a = f . Then af E is the identity operator, so kaf Ek = 1 < kak. For Corollary 2.4 take A = Cf ⊂ L(E) with the operator norm and take π to be the identity representation. Then π is isometric but π0 is not. Proposition 2.6 (Proposition 1.25 of [19]). Let p ∈ [1, ∞), and let A be a separable Lp operator algebra. Then A is separably representable. If A is nondegenerately representable, then A is separably nondegenerately representable. Since we will only use separable Lp operator algebras in this paper, we need only deal with separable Lp spaces. Lemma 2.7 implies that we can always assume that the measures are σ-finite. Lemma 2.7. Let p ∈ [1, ∞). Let (X, B, µ) be a measure space such that Lp(X, µ) is separable. Then there exists a σ-finite measure space (Y, C, ν) such that Lp(X, µ) is isometrically isomorphic to Lp(Y, ν). Proof. See the Corollary to Theorem 3 in Section 15 of [13]. (cid:3) The following result will be used often enough that we restate it here. Proposition 2.8. Let p ∈ [1, ∞) \ {2}. Let (X, B, µ) be a measure space and let e ∈ L(Lp(X, µ)) be an idempotent. Then kek ≤ 1 if and only if there exists a CLASSIFICATION OF Lp AF ALGEBRAS 5 measure space (Y, C, ν) and an isometric bijection from Lp(Y, ν) to the range of e. Moreover, if Lp(X, µ) is separable, then ν can be chosen to be σ-finite. Proof. This is part of Theorem 3 in Section 17 of [13]. (cid:3) Proposition 2.9. Let A be a unital Lp operator algebra in which k1k = 1. Then A has an isometric unital representation on an Lp space. If A is separable then the Lp space can be chosen to be separable and to be the Lp space of a σ-finite measure space. Proof. Let (X, B, µ) be a measure space such that there is an isometric represen- tation ρ : A → L(Lp(X, µ)). Then e = ρ(1) is an idempotent in L(Lp(X, µ)) with kek = 1. Set E = ran(e). Then ρ induces an isometric unital homomorphism ρ0 : A → L(E) by Corollary 2.4. By Proposition 2.8, there is a measure space (Y, C, ν) such that E is isometrically isomorphic to Lp(Y, ν). The first part of the conclusion follows. For the second part, if A is separable then we may require that Lp(X, µ) be separable by Proposition 2.6. Then E must be separable. Proposi- tion 2.8 implies that E is isometrically isomorphic to the Lp space of a σ-finite measure space. (cid:3) 3. Matrix normed algebras and matricial Lp operator algebras We will mostly work with ordinary Lp operator algebras, but for some results we will need the matrix normed version introduced here (Definition 3.18). We also need the analogs of Proposition 2.6 and Proposition 2.9 for matricial Lp operator algebras; see Proposition 3.19 and Proposition 3.20. Matrix normed spaces (operator spaces of various kinds) are well known, but we have not seen a general definition of a matrix normed algebra. We therefore give one here (Definition 3.2). The conditions on the matrix norms seem to be the minimal "reasonable" conditions. Condition (1) essentially says that submatrices have smaller norm. We first describe our (fairly standard) notation for matrices. Notation 3.1. Let n ∈ Z>0. Then Mn denotes the algebra of n×n complex matri- ces (without any specific norm being assumed). For j, k ∈ {1, 2, . . . , n}, we let ej,k denote the corresponding standard matrix unit of Mn. For any complex algebra A, n we identify the algebra Mn(A) with Mn ⊗ A via (aj,k)1≤j,k≤n 7→ ej,k ⊗ aj,k. Xj,k=1 For x ∈ Mn and a ∈ Mn(A), the products xa and ax are defined in the obvious way, so that x(y ⊗ b) = xy ⊗ b and (y ⊗ b)x = yx ⊗ b for y ∈ Mn and b ∈ A. Definition 3.2. A matrix normed algebra is a complex algebra A equipped with algebra norms k · kn on Mn(A) for all n ∈ Z>0, satisfying the following: (1) For any m, n ∈ Z>0 with m ≤ n, any injective functions σ, τ : {1, 2, . . . , m} → {1, 2, . . . , n}, and any a = (aj,k)1≤j,k≤n ∈ Mn(A), we have (cid:13)(cid:13)(aσ(j),τ (k))1≤j,k≤m(cid:13)(cid:13)m ≤ kakn. s = diag(λ1, λ2, . . . , λn) ∈ Mn, then (2) For any n ∈ Z>0, any a ∈ Mn(A), and any λ1, λ2, . . . , λn ∈ C, if we set kaskn, ksakn ≤ max(cid:0)λ1, λ2, . . . , λn(cid:1)kakn. 6 N. CHRISTOPHER PHILLIPS AND MARIA GRAZIA VIOLA (3) For any m, n ∈ Z>0, a ∈ Mm(A), and b ∈ Mn(A), we have kdiag(a, b)km+n = max(kakm, kbkn). We abbreviate k · k1 to k · k. If A is complete in k · k, we call A a matrix normed Banach algebra. Remark 3.3. In Definition 3.2, if A is unital and k1k = 1, or even if A has an approximate identity which is bounded by 1, condition (2) follows from condition (3) and submultiplicativity of k · kn. Remark 3.4. In Definition 3.2, the inequality kdiag(a, b)km+n ≥ max(kakm, kbkn) in condition (3) follows from condition (1). Lemma 3.5. Let A be a matrix normed algebra, let n ∈ Z>0, and let a = (aj,k)1≤j,k≤n ∈ Mn(A). Then max 1≤j,k≤n kaj,kk ≤ kakn ≤ n Xj,k=1 kaj,kk. Proof. The first inequality follows from Definition 3.2(1). We prove the second inequality. First, two applications of condition (1), taking σ and τ there to be permutations, show that permuting the rows and also permuting the columns of a matrix does not change its norm. Using this fact at the second step and condi- tion (3) at the first step, we get kaj,kk = kdiag(aj,k, 0, . . . , 0)kn = kej,jaek,kkn. Apply this to the relation a = m Xj,k=1 ej,jaek,k to complete the proof. (cid:3) Corollary 3.6. Let A be a matrix normed Banach algebra. Then Mn(A) is com- plete for all n ∈ Z>0. Proof. This is immediate from Lemma 3.5. (cid:3) For clarity, we state the standard definitions related to completely bounded maps. Definition 3.7. Let A and B be matrix normed algebras, and let ϕ : A → B be a linear map. For n ∈ Z>0, write ϕ(n) or idMn ⊗ ϕ for the map Mn(A) → Mn(B) determined by (aj,k)1≤j,k≤n 7→ (ϕ(aj,k))1≤j,k≤n. Then: (1) We set kϕkcb = sup n∈Z>0 bounded . kϕ(n)k. If kϕkcb < ∞, we say that ϕ is completely (2) We say that ϕ is completely contractive if kϕkcb ≤ 1. (3) We say that ϕ is completely isometric if ϕ(n) is isometric (not necessarily surjective) for all n ∈ Z>0. (4) We say that ϕ is a completely isometric isomorphism if ϕ is completely isometric and bijective. Definition 3.8. Let A be a matrix normed algebra. (1) Let B be a subalgebra of A. For n ∈ Z>0, we define the norm k · kn on Mn(B) to be the restriction to Mn(B) of the given norm on Mn(A). (2) Let J ⊂ A be a closed ideal. For n ∈ Z>0, we define the norm k · kn on Mn(A/J) to be the quotient norm coming from the obvious identification of Mn(A/J) with Mn(A)/Mn(J). CLASSIFICATION OF Lp AF ALGEBRAS 7 Lemma 3.9. Let A be a matrix normed algebra. (1) Let B ⊂ A be a subalgebra. Then the norms in Definition 3.8(1) make B a matrix normed algebra, and the inclusion map is completely isometric. (2) Let J ⊂ A be a closed ideal. Then the norms in Definition 3.8(2) make A/J a matrix normed algebra, and the quotient map is completely contractive. Proof. Part (1) is immediate. We prove part (2). Let π : A → A/J be the quotient map. Complete con- tractivity of π is immediate. For Definition 3.2(1), let m, n ∈ Z>0 with m ≤ n, let σ, τ : {1, 2, . . . , m} → {1, 2, . . . , n} be injective functions, and let x = (xj,k)1≤j,k≤n ∈ Mn(A/J). Let ε > 0. Choose a = (aj,k)1≤j,k≤n ∈ Mn(A) such that π(n)(a) = x and kakn < kxkn + ε. Since π(m) is contractive, we have (cid:13)(cid:13)(xσ(j),τ (k))1≤j,k≤m(cid:13)(cid:13)m ≤(cid:13)(cid:13)(aσ(j),τ (k))1≤j,k≤m(cid:13)(cid:13)m ≤ kakn < kxkn + ε. The proofs of Definition 3.2(2) and the inequality kdiag(a, b)km+n ≤ max(kakm, kbkn) in Definition 3.2(3) are similar. Equality in Definition 3.2(3) now follows from Remark 3.4. (cid:3) Definition 3.10. Let n ∈ Z>0. A matrix s ∈ Mn is a permutation matrix if there n exists a bijection σ : {1, 2, . . . , n} → {1, 2, . . . , n} such that s = eσ(j), j. The Xj=1 n matrix s is a complex permutation matrix if there exist a bijection σ : {1, 2, . . . , n} → {1, 2, . . . , n} and λ1, λ2, . . . , λn ∈ S1 = {z ∈ C : z = 1} such that s = λjeσ(j), j. Xj=1 The complex permutation matrices form a group. Lemma 3.11. Let A be a matrix normed algebra and fix n ∈ Z>0. Let a ∈ Mn(A), and let s ∈ Mn be a complex permutation matrix. Interpret as and sa as in Notation 3.1. Then kaskn = ksakn = kakn. Proof. Since s−1 is also a complex permutation matrix, it suffices to prove that kaskn ≤ kakn and ksakn ≤ kakn. Since a complex permutation matrix is a product of a permutation matrix and a diagonal matrix with diagonal entries in S1, it suffices to prove these inequalities for these two kinds of matrices separately. For the first kind, apply Definition 3.2(1). For the second kind, apply Definition 3.2(2). (cid:3) Definition 3.12. Let m, n ∈ Z>0 and let σ : {1, 2, . . . , m} × {1, 2, . . . , n} → {1, 2, . . . , mn} be a bijection. We let θσ : Mm ⊗ Mn → Mmn be the unique algebra isomorphism such that for i, j ∈ {1, 2, . . . , m} and k, l ∈ {1, 2, . . . , n}, we have θσ(ei,j ⊗ ek,l) = eσ(i,k), σ(j,l). The standard choice of bijection is the one given by σ(j, l) = j + m(l − 1) for j = 1, 2, . . . , m and l = 1, 2, . . . , n. 8 N. CHRISTOPHER PHILLIPS AND MARIA GRAZIA VIOLA Definition 3.13. Let A be a matrix normed algebra and let m ∈ Z>0. We define matrix norms on Mm(A) as follows. For n ∈ Z>0, choose some bijection σn : {1, 2, . . . , n} × {1, 2, . . . , m} → {1, 2, . . . , nm}, and use it (and Notation 3.1) to get the isomorphism θσn ⊗ idA : Mn(Mm(A)) → Mnm(A). For a ∈ Mn(Mm(A)), we then use the matrix norms on A to define kakn = k(θσn ⊗ idA)(a)knm. Lemma 3.14. In Definition 3.13, the matrix norms are independent of the choice of (σn)n∈Z>0, and make Mm(A) a matrix normed algebra. Proof. Independence of (σn)n∈Z>0 follows from Lemma 3.11, and the fact that one gets a matrix normed algebra follows easily from Definition 3.2. (cid:3) Definition 3.15. Let (X, B, µ) be a measure space, and let A ⊂ L(Lp(X, µ)) be a closed subalgebra. We equip A with the matrix norms coming from the identification of Mn(A) with a closed subalgebra of L(cid:0)Lp({1, 2, . . . , n} × X, ν × µ)(cid:1), in which ν is counting measure on {1, 2, . . . , n}. d with Md in the standard way. For a ∈ M p d(cid:1) with the usual operator norm, d = lp(cid:0){1, 2, . . . , d}(cid:1). We further let M p Notation 3.16. For any set S and any p ∈ [1, ∞], we give lp(S) the usual meaning (using counting measure on S), and we set (as usual) lp = lp(Z>0). For d ∈ Z>0 we let lp and we algebraically identify M p d , we write the norm as kakp. We equip M p d with the matrix norms as in Definition 3.15. Lemma 3.17. Let (X, B, µ) be a measure space, and let A ⊂ L(Lp(X, µ)) be a closed subalgebra. Then A is a matrix normed algebra with the matrix norms of Definition 3.15. Moreover for n ∈ Z>0, a ∈ Mn(A), and x ∈ Mn, with products as in Notation 3.1, we have kaxkn, kxakn ≤ kxkpkakn. Furthermore, for m ∈ Z>0 the matrix norms on Mm(A) from Definition 3.13 agree with those gotten from the obvious inclusion d = L(cid:0)lp in which ν is counting measure on {1, 2, . . . , m}. Mm(A) → L(cid:0)Lp({1, 2, . . . , m} × X, ν × µ)(cid:1), Proof. All parts are easy. (cid:3) Definition 3.18. Let p ∈ [1, ∞). A matricial Lp operator algebra is a ma- trix normed Banach algebra A such that there exists a measure space (X, B, µ) and a completely isometric isomorphism from A to a norm closed subalgebra of L(Lp(X, µ)). Using the terminology from Definition 2.2, we say that a matricial Lp opera- tor algebra A is separably representable if it has a separable completely isometric representation. We say that A is σ-finitely representable if it has a σ-finite com- pletely isometric representation. We say that A is nondegenerately (separably) representable if it has a nondegenerate (separable) completely isometric representa- tion, and nondegenerately σ-finitely representable if it has a nondegenerate σ-finite completely isometric representation. Proposition 3.19. Let p ∈ [1, ∞), and let A be a separable matricial Lp operator algebra. Then A is separably representable. If A is nondegenerately representable, then A is separably nondegenerately representable. CLASSIFICATION OF Lp AF ALGEBRAS 9 Proof. For n ∈ Z>0 let νn be counting measure on {1, 2, . . . , n}. Let S be a count- able dense subset of A, and for n ∈ Z>0 define which is a countable dense subset of Mn(A). Sn =(cid:8)b ∈ Mn(A) : bj,k ∈ S for j, k = 1, 2, . . . , n(cid:9), By hypothesis, there exist a measure space (X, B, µ) and a completely isometric representation ρ : A → L(Lp(X, µ)), which we can take to be nondegenerate when A is nondegenerately representable. For any m, n ∈ Z>0 and b ∈ Sn, choose such that ξn,b,m =(cid:0)ξ(j) n,b,m(cid:1)1≤j≤n ∈ Lp(cid:0){1, 2, . . . , n} × X, νn × µ(cid:1) kξn,b,mkp = 1 and (cid:13)(cid:13)(idMn ⊗ ρ)(b)ξn,b,m(cid:13)(cid:13) > kbk − 1 m . n,b,m, ξ(2) n,b,m, . . . , ξ(2) By the argument used in the proof of Proposition 1.25 of [19] there exists a sepa- rable closed sublattice Fn,b,m of Lp(X, µ) containing ξ(1) n,b,m and such that ρ(A)Fn,b,m ⊂ Fn,b,m. Moreover, Fn,b,m is isomorphic to Lp(Yn,b,m, νn,b,m) for some measure space (Yn,b,m, νn,b,m). Furthermore, if ρ is nondegenerate then Fn,b,m can be chosen to satisfy span(ρ(A)Fn,b,m) = Fn,b,m. The map defined by πn,b,m(a) = ρ(a)Fn,b,m is a completely contractive representation of A on a separa- ble Lp-space, which is nondegenerate if ρ is nondegenerate. Since Fn,b,m contains ξ(1) n,b,m, ξ(2) m for every m ∈ Z>0. Now let π be the Lp direct sum of the representations πn,b,m for m, n ∈ Z>0 and b ∈ Sn, as in Definition 1.23 of [19]. Then π is a completely contractive representa- tion on a separable Lp space. We have k(idMn ⊗ π)(b)k = kbk for all n ∈ Z>0 and b ∈ Sn, so density of Sn in Mn(A) implies that π is completely isometric. Moreover, by Lemma 1.24 in [19], π is nondegenerate if ρ is nondegenerate. (cid:3) n,b,m, we get k(idMn ⊗πn,b,m)(b)k > kbk− 1 n,b,m, . . . , ξ(2) Proposition 3.20. Let p ∈ [1, ∞), and let A be a unital matricial Lp operator algebra in which k1k = 1. Then A has a completely isometric unital representation on an Lp space. If A is separable then the Lp space can be chosen to be separable and to come from a σ-finite measure space. Proof. Let (X, B, µ) be a measure space such that there is completely isometric rep- resentation ρ0 : A → L(Lp(X, µ)). Then e = ρ0(1) is an idempotent in L(Lp(X, µ)), and kek = 1. By Proposition 2.8 there exists a measure space (Y, C, ν) such that ran(e) is isometrically isomorphic to Lp(Y, ν). Thus, ρ0 gives a completely isomet- ric unital homomorphism ρ : A → L(Lp(Y, ν)). Moreover, if A is separable, then Lp(X, µ) can be chosen to be separable, which implies that ran(e) is also separable. To get σ-finiteness, use Lemma 2.7. (cid:3) 4. Unique matrix norms We consider uniqueness of matrix norms on Lp operator algebras. Most of the Lp operator algebras we deal with will have unique Lp operator matrix norms, in the sense of Definition 4.1 below. The basic examples are M p d and C(X). We will show in Corollary 9.12 below that all spatial Lp AF algebras have unique Lp operator matrix norms. The proof that C(X) has unique Lp operator matrix norms uses a structure theorem (Theorem 4.5) for contractive unital representations of C(X) on Lp spaces, which also plays a key role later. To avoid technical issues, we restrict our discussion to the separable case. 10 N. CHRISTOPHER PHILLIPS AND MARIA GRAZIA VIOLA Definition 4.1. Let p ∈ [1, ∞) \ {2}. Let A be a separable Lp operator algebra. We say that A has unique Lp operator matrix norms if whenever (X, B, µ) and (Y, C, ν) are σ-finite measure spaces such that Lp(X, µ) and Lp(Y, ν) are separable, π : A → L(Lp(X, µ)) and σ : A → L(Lp(Y, ν)) are isometric representations, and π(A) and σ(A) are given the matrix normed structures of Definition 3.15, then σ ◦ π−1 : π(A) → σ(A) is completely isometric. When A is unital and k1k = 1, in Definition 4.1 we need only consider unital isometric representations. Lemma 4.2. Let p ∈ [1, ∞) \ {2}. Let A be a unital separable Lp operator algebra in which k1k = 1. Assume that whenever (X, B, µ) and (Y, C, ν) are σ-finite measure spaces such that Lp(X, µ) and Lp(Y, ν) are separable, π : A → L(Lp(X, µ)) and σ : A → L(Lp(Y, ν)) are unital isometric representations, and π(A) and σ(A) are given the matrix normed structures of Definition 3.15, then σ ◦ π−1 : π(A) → σ(A) is completely isometric. It follows that A has unique Lp operator matrix norms. Proof. Let (X, B, µ) and (Y, C, ν) be σ-finite measure spaces such that Lp(X, µ) and Lp(Y, ν) are separable, and let π : A → L(Lp(X, µ)) and σ : A → L(Lp(Y, ν)) be isometric representations. The operator e = π(1) is an idempotent in L(Lp(X, µ)) with kek = 1. Set E = ran(e). Then π induces a unital homomorphism π0 : A → L(E), which is isometric by Corollary 2.4. By Proposition 2.8, there is a measure space (X0, B0, µ0) such that E is isometrically isomorphic to Lp(X0, µ0). Since E is separable, we may require that µ0 be σ-finite. Similarly, ran(σ(1)) is isometrically isomorphic to a separable Lp space Lp(Y0, ν0) in which ν0 is σ-finite, and σ induces an isometric unital homomorphism σ0 : A → L(Lp(Y0, ν0)). In particular, σ0 ◦ π−1 : π0(A) → σ0(A) is isometric. 0 Let n ∈ Z>0. We take the norms on Mn(π(A)), Mn(π0(A)), Mn(σ(A)), and Mn(σ0(A)) to be as in Definition 3.15. Define ρ0 = σ0 ◦ π−1 0 : π0(A) → σ0(A) ⊂ L(Lp(Y0, ν0)) ρ = σ ◦ π−1 0 : π0(A) → σ(A) ⊂ L(Lp(Y, ν)). and Then so k(idMn ⊗ ρ)(1)k = 1. The hypothesis implies that idMn ⊗ ρ0 is isometric. So Corollary 2.4 implies that idMn ⊗ ρ is isometric. (idMn ⊗ ρ)(1) = 1Mn ⊗ σ(1) ∈ L(cid:0)lp n ⊗p Lp(Y, ν)(cid:1), Similarly, the map idMn ⊗ (π ◦ π−1 0 ) : Mn(π0(A)) → Mn(π(A)) is isometric. Therefore idMn ⊗ (σ ◦ π−1) : Mn(π(A)) → Mn(σ(A)) is isometric. This completes the proof. (cid:3) Proposition 4.3. Let p ∈ [1, ∞) \ {2}. Let M p every nonzero contractive unital representation of M p completely isometric. d be as in Notation 3.16. Then d on a separable Lp space is Proof. Let (X, B, µ) be a measure space such that Lp(X, µ) is separable, and let ρ : M p d → L(Lp(X, µ)) be a contractive unital representation. By Lemma 2.7, we can assume that (X, B, µ) is σ-finite. Theorem 7.2 of [16] provides a σ-finite measure space (Z, C, λ) and a bijective isometry u : lp d ⊗ Lp(Z, λ) → Lp(X, µ) CLASSIFICATION OF Lp AF ALGEBRAS 11 such that for all a ∈ M p see that d we have ρ(a) = u(a ⊗ 1)u−1. For n ∈ Z>0, it is easy to 1Mn ⊗ u : lp n ⊗ lp d ⊗ Lp(Z, λ) → lp n ⊗ Lp(X, µ) is a bijective isometry such that (1Mn ⊗ ρ)(b) = (1Mn ⊗ u)(b ⊗ 1)(1Mn ⊗ u)−1 d ). It is now immediate that ρn is isometric. for all b ∈ Mn(M p Corollary 4.4. Let p ∈ [1, ∞) \ {2}. The algebra M p Lp operator matrix norms. (cid:3) d of Notation 3.16 has unique Proof. Combine Lemma 4.2 and Proposition 4.3. (cid:3) Next, we give a structure theorem for any contractive unital representation of C(X) on an Lp space. Theorem 4.5. Let p ∈ [1, ∞) \ {2}. Let X be a compact metrizable space, let (Y, C, ν) be a σ-finite measure space, and let π : C(X) → L(Lp(Y, ν)) be a contrac- tive unital homomorphism. Let µ : L∞(Y, ν) → L(Lp(Y, ν)) be the representation of L∞(Y, ν) on Lp(Y, ν) by multiplication operators. Then there exists a unital C*-algebra homomorphism ϕ : C(X) → L∞(Y, ν) such that π = µ ◦ ϕ. Proof. We claim that the range of π is contained in the range of µ. It suffices to prove that if f ∈ C(X) is real valued and satisfies kf k < π, then π(f ) is in the range of µ. Let f be such a function. For λ ∈ R, the function wλ = exp(iλf ) is invertible in C(X) and satisfies kwλk = kw−1 λ k = 1. Therefore π(wλ) is a bijective isometry in L(Lp(Y, ν)). By Lemma 6.16 of [16], the operator π(wλ) is a spatial isometry in the sense of Definition 6.4 of [16]. In particular, it has a spatial system (Eλ, Fλ, Sλ, gλ) as there. By Lemma 6.22 of [16], we have Sλ = S0 for all λ ∈ R. Now π(w0) = 1, so, in the notation of Definition 6.3 of [16] and Definition 5.4 of [16], the operator π(w0) has the spatial system (Y, Y, idC/N (ν), 1). The uniqueness statement in Lemma 6.6 of [16] now implies that Sλ = idC/N (ν) for all λ ∈ R. Therefore π(wλ) is a multiplication operator; in fact, π(wλ) = µ(gλ) for all λ ∈ R. Now let log be the holomorphic branch which is real on (0, ∞) and defined on C \ (−∞, 0]. We have sp(g1) = sp(π(w1)) ⊂ sp(w1) ⊂ C \ (−∞, 0] and π(f ) = π(−i log(w1)) = −i log(π(w1)) = −i log(µ(g1)) = µ(−i log(g1)). Thus π(f ) is in the range of µ, as claimed. It follows that there is a contractive homomorphism ϕ : C(X) → L∞(Y, ν) such that π = µ ◦ ϕ. Obviously ϕ is unital. It follows from Proposition A.5.8 of [2] that ϕ is a C*-algebra homomorphism. (cid:3) We don't need the following proposition, but it is an interesting result which follows from the machinery we have developed. Proposition 4.6. Let X be a compact metrizable space, and let p ∈ [1, ∞) \ {2}. Then C(X) has unique Lp operator matrix norms. They are given as follows. Let a ∈ Mn(C(X)). Interpret a as a continuous function a : X → Mn. Equip Mn = M p n with the norm k · kp from Notation 3.16. Then kakn = sup x∈X ka(x)kp. 12 N. CHRISTOPHER PHILLIPS AND MARIA GRAZIA VIOLA Proof. Choose a σ-finite Borel measure µ on X such that µ(U ) > 0 for every nonempty open set U ⊂ X. Represent C(X) on Lp(X, µ) as multiplication opera- tors. It is then easy to check that the matrix norms from Definition 3.15 are equal to the matrix norms in the statement of the proposition. In view of Lemma 4.2, it remains to show that if (Y, C, ν) is a σ-finite measure space such that L(Lp(Y, ν)) is separable, and π : C(X) → L(Lp(Y, ν)) is an iso- metric unital homomorphism, then π is completely isometric. Let ρ : L∞(Y, ν) → L(Lp(Y, ν)) be the representation given by multiplication operators. Then ρ is iso- metric, so we can identify L∞(Y, ν) with its image under ρ, and thus make L∞(Y, ν) a matricial Lp operator algebra using the matrix norms of Definition 3.15. For n ∈ Z>0, identify Mn(L∞(Y, ν)) with the algebra of L∞ functions from Y to Mn. It is easy to check that the norms on Mn(L∞(Y, ν)) are given by kakn = ess sup ka(y)kp. y∈Y Now let Z be the maximal ideal space of L∞(Y, ν), and let γ : L∞(Y, ν) → C(Z) be the Gelfand transform, which is an isomorphism. Define matrix norms on C(Z) in the same way as on C(X) in the statement of the proposition. For every f ∈ L∞(Y, ν), the essential range of f is the range of γ(f ). It follows that for every a ∈ Mn(L∞(Y, ν)), the essential range of a is equal to the range of (idMn ⊗ γ)(a) ∈ Mn(C(Z)) = C(Z, Mn). Therefore γ is completely isometric. Apply Theorem 4.5 to π. We get a unital C*-algebra homomorphism ϕ : C(X) → L∞(Y, ν) such that π = ρ ◦ ϕ. Moreover, ϕ is injective. There is a continuous function h : Z → X such that (γ ◦ ϕ)(f ) = f ◦ h for all f ∈ C(X). Injectivity It is now immediate that γ ◦ ϕ is completely of γ ◦ ϕ implies surjectivity of h. isometric. Since γ and ρ are completely isometric, we conclude that π is completely isometric. (cid:3) In this section we show that direct sum of a family of (matricial) Lp operator algebras is also a (matricial) Lp operator algebra. 5. Direct sums Definition 5.1. If(cid:0)(Xi, Bi, µi)(cid:1)i∈I is a family of measure spaces, then the measure space (X, B, µ) =ai∈I (Xi, Bi, µi) is determined by taking X =ai∈I B =(cid:8)E ⊂ X : E ∩ Xi ∈ Bi for all i ∈ I(cid:9), µi(E ∩ Xi) for E ∈ B. Xi, and µ(E) =Xi∈I Definition 5.2. Whenever N ∈ Z>0 and A1, A2, . . . , AN are Banach algebras, we N Mk=1 make norm Ak a Banach algebra by giving it the obvious algebra structure and the for a1 ∈ A1, a2 ∈ A2, . . . , aN ∈ AN . If A1, A2, . . . , AN are matrix normed Banach k(a1, a2, . . . , aN )k = max(cid:0)ka1k, ka2k, . . . , kaN k(cid:1) algebras, we define matrix norms on Ak by N Mk=1 for n ∈ Z>0 and a1 ∈ Mn(A1), a2 ∈ Mn(A2), . . . , aN ∈ Mn(AN ). k(a1, a2, . . . , aN )kn = max(cid:0)ka1kn, ka2kn, . . . , kaN kn(cid:1) CLASSIFICATION OF Lp AF ALGEBRAS 13 Lemma 5.3. Let N ∈ Z>0. Let A1, A2, . . . , AN be matrix normed Banach alge- bras. Then N Mk=1 Ak, as in Definition 5.2, is a matrix normed Banach algebra. Proof. The proof is easy, and is omitted. (cid:3) Lemma 5.4. Let the notation be as in Definition 5.2. Let B be a Banach al- gebra, and for k = 1, 2, . . . , N let ϕk : B → Ak be a homomorphism. Define N Mk=1 ϕ : B → Ak by ϕ(b) = (cid:0)ϕ1(b), ϕ2(b), . . . , ϕN (b)(cid:1) for b ∈ B. Then ϕ is con- tractive if and only if ϕk is contractive for k = 1, 2, . . . , N . If A1, A2, . . . , AN are matrix normed Banach algebras, then ϕ is completely contractive if and only if ϕk is completely contractive for k = 1, 2, . . . , N . Proof. The proof is immediate. (cid:3) Lemma 5.5. Let the notation be as in Definition 5.2. Let S ⊂ {1, 2, . . . , N }. Then Ak is an ideal in Mk∈S N Mk=1 Ak, and the obvious map N Mk=1 Ak.Mk∈S Ak →Mk6∈S Ak is completely isometric when the quotient is given the matrix norms of Definition 3.8(2). Proof. The proof is easy, and is omitted. (cid:3) Lemma 5.6. Let A be a matrix normed Banach algebra. Let n ∈ Z>0, and let n ϕ : A → Mn(A) be the map ϕ(a1, a2, . . . , an) = diag(a1, a2, . . . , an) for a1, a2, . . . , an ∈ Mk=1 A. Then ϕ is completely isometric. Proof. Let r ∈ Z>0. Let σ be the standard bijection of Definition 3.12, with r in n place of n, and let θσ be as there. For a ∈ A the matrix (cid:2)(θσ ⊗ idA) ◦ (idMr ⊗ ϕ)(cid:3)(a) is block diagonal. So iteration of condition (3) in Definition 3.2 shows that (θσ ⊗ idA) ◦ (idMr ⊗ ϕ) is isometric. Lemma 3.14 implies that θσ ⊗ idA is isometric. So idMr ⊗ ϕ is isometric. (cid:3) Mk=1 Lemma 5.7. Let p ∈ [1, ∞). In Definition 5.2, if A1, A2, . . . , AN are Lp opera- N tor algebras, then so is A = Ak. If A1, A2, . . . , AN are matricial Lp operator Mk=1 algebras, then A is a matricial Lp operator algebra. Proof. We give the proof for Lp operator algebras; the matricial case is essentially the same. Suppose that ρk : Ak → L(Lp(Xk, µk)) is an isometric representation for k = 1, 2, . . . , N . Let X = N ak=1 Xk and µ be as in Definition 5.1. Then Lp(X, µ) is 14 N. CHRISTOPHER PHILLIPS AND MARIA GRAZIA VIOLA the Lp direct sum N Mk=1 Lp(Xk, µk). Define ρ : Ak → L(Lp(X, µ)) by N Mk=1 ρ(a1, a2, . . . , aN ) = ρ1(a1) ⊕ ρ2(a2) ⊕ · · · ⊕ ρN (aN ) for a1 ∈ A1, a2 ∈ A2, . . . , aN ∈ AN . Clearly ρ is an isometric representation of N Ak. Mk=1 (cid:3) 6. Hermitian idempotents The right kind of idempotent to consider in an Lp operator algebra for p 6= 2 is what might be called a "spatial idempotent", that is, one which is a spatial partial isometry in the sense of Definition 6.4 of [16]. We develop some of the basic theory in this section. Such idempotents can be characterized as those which are hermitian in the sense of Definition 6.4 below, a much older notion (see [20]). Although we will not need the general theory of hermitian elements of a Banach algebra, it seems appropriate to make the connection with the older concept. We formalize the following terminology for idempotents. Definition 6.1 (Definition 4.1.1 of [1]). Let A be a ring (not necessarily unital), and let e, f ∈ A be idempotents. We say that f dominates e, written f ≥ e or e ≤ f , if f e = ef = e. We say that e and f are orthogonal if ef = f e = 0. Even if A is a C*-algebra, the notation e ≤ f need not agree with the usual C*-algebraic order. Among other things, e and f need not be selfadjoint. If e and f happen to be projections in a C*-algebra, then our notation does agree with the usual C*-algebraic order. Orthogonality need not be the same as the version of orthogonality for C*-algebras implicit in the remark after Definition 4.1.1 of [1]. Definition 6.2 (Definition 2.6.1 of [15]). Let A be a unital Banach algebra in which k1k = 1. Let a ∈ A. Then the numerical range W (a) is the set of all numbers ω(a) ∈ C for linear functionals ω on A such that kωk = ω(1) = 1. Theorem 6.3 (Theorem 2.6.7 of [15]). Let A be a unital Banach algebra in which k1k = 1, and let a ∈ A. Then the following are equivalent: (1) W (a) ⊂ R. (2) k exp(iλa)k = 1 for all λ ∈ R. (3) k exp(iλa)k ≤ 1 for all λ ∈ R. (4) With the limit being taken over λ ∈ R, lim λ→0 λ−1(cid:0)k1 − iλak − 1(cid:1) = 0. Proof. The equivalence of conditions (1), (2), and (4) is in Theorem 2.6.7 of [15]. That (2) implies (3) is trivial. That (3) implies (2) follows from k1k = 1 and exp(iλa)−1 = exp(−iλa). (cid:3) Definition 6.4 (see Definition 2.6.5 of [15] and the preceding discussion). Let A be a unital Banach algebra in which k1k = 1, and let a ∈ A. We say that a is hermitian if a satisfies the equivalent conditions of Theorem 6.3. If a is also an idempotent, we call it a hermitian idempotent . CLASSIFICATION OF Lp AF ALGEBRAS 15 Remark 6.5. Let A be a unital Banach algebra in which k1k = 1. Then clearly 0 and 1 are hermitian idempotents. Also, in any unital Banach algebra, if e is a hermitian idempotent, then so is 1 − e. Indeed, if λ ∈ R then k exp(iλ(1−e))k = ke+exp(iλ)(1−e)k = k exp(iλ) exp(−iλe)k = k exp(−iλe)k = 1, as desired. A hermitian idempotent in a C*-algebra is simply a projection. (See Proposi- tion 3.3.3 in [4], observing that a nonzero hermitian idempotent has norm 1 by Lemma 6.6 below.) The following result gives the characterization we use most often. Lemma 6.6. Let A be a unital Banach algebra in which k1k = 1. Let e ∈ A be an idempotent. Define a homomorphism βe : C⊕C → A by βe(λ1, λ2) = λ1e+λ2(1−e) for λ1, λ2 ∈ C. Then e is hermitian in the sense of Definition 6.4 if and only if, when C ⊕ C is normed as in Definition 5.2, the homomorphism βe is contractive. Proof. We use the characterization (3) of Theorem 6.3. First suppose that βe is contractive. Then for λ ∈ R we have k exp(iλe)k = kβ((exp(iλ), 1))k ≤ k(exp(iλ), 1)k = 1. For the converse, suppose e is hermitian, and let λ1, λ2 ∈ C. We need to prove (6.1) kβe((λ1, λ2))k ≤ max(λ1, λ2). This relation is trivial if λ1 = λ2 = 0. Next, suppose λ1 ≤ λ2 and λ2 6= 0. Multiplying by λ−1 2 , we reduce to the case λ2 = 1. Write λ1 = ρ exp(iθ) with θ ∈ R and 0 ≤ ρ ≤ 1. Define α1 = θ + arccos(ρ) and α2 = θ − arccos(ρ). Then one checks that (λ1, 1) = 1 kβe((λ1, 1))k =(cid:13)(cid:13)(cid:13)(cid:13) which is (6.1). 1 2(cid:2) exp(iα1e) + exp(iα2e)(cid:3)(cid:13)(cid:13)(cid:13)(cid:13) 2(cid:2)(exp(iα1), 1) + (exp(iα2), 1)(cid:3). So 1 ≤ 2(cid:0)k exp(iα1e)k+k exp(iα2e)k(cid:1) ≤ 1, Finally, suppose λ2 ≤ λ1 and λ1 6= 0. Using Remark 6.5, we can apply the case of (6.1) already done to 1 − e, with (λ2, λ1) in place of (λ1, λ2). This gives (6.1) for e and (λ1, λ2). (cid:3) Lemma 6.7. Let A and B be unital Banach algebras such that k1Ak = 1 and k1Bk = 1. Let ϕ : A → B be a contractive unital homomorphism, and let e ∈ A be a hermitian idempotent. Then ϕ(e) ∈ B is a hermitian idempotent. Proof. The proof is immediate from Lemma 6.6. (cid:3) Lemma 6.8. Let N ∈ Z>0, and let A1, A2, . . . , AN be unital Banach algebras N whose identities have norm one. Set A = Ak, equipped with the norm in Def- inition 5.2, and for k = 1, 2, . . . , N let ek be a hermitian idempotent in Ak. Then (e1, e2, . . . , eN ) is a hermitian idempotent in A. Proof. The proof is immediate from Lemma 6.6. (cid:3) Mk=1 16 N. CHRISTOPHER PHILLIPS AND MARIA GRAZIA VIOLA We are interested in hermitian idempotents in Lp operator algebras. Given p ∈ [1, ∞) \ {2}, the following lemma gives a characterization of hermitian idempotents in L(Lp(X, µ)), for a σ-finite measure space (X, B, µ). Lemma 6.9. Let p ∈ [1, ∞) \ {2}. Let (X, B, µ) be a σ-finite measure space, and let e ∈ L(Lp(X, µ)) be an idempotent. Then the following are equivalent: (1) e is a hermitian idempotent. (2) e is a spatial partial isometry in the sense of Definition 6.4 of [16]. (3) There is a measurable subset E ⊂ X such that e is multiplication by χE. Proof. Lemma 6.18 in [16] shows that (2) and (3) are equivalent. It is obvious that (3) implies (1). For the converse, assume that e is a hermitian idempotent. Thus the homomorphism βe : C ⊕ C → L(Lp(X, µ)) of Lemma 6.6 is unital and contractive. Set Y = {0, 1}, and define f ∈ C(Y ) by f (0) = 1 and f (1) = 0. Let ρ be the representation of L∞(X, µ) on Lp(X, µ) by multiplication operators. By Theorem 4.5, there exists a unital *-homomorphism ϕ : C(Y ) → L∞(X, µ) such that βe = ρ ◦ ϕ. Since ϕ(f ) is an idempotent in L∞(X, µ), there is a measurable set E ⊂ X such that ϕ(f ) = χE. (cid:3) Corollary 6.10. Let p ∈ [1, ∞) \ {2}, let d ∈ Z>0, and let e ∈ M p d be an idem- potent. Then e is hermitian if and only if e is a diagonal matrix with entries in {0, 1}. Proof. Identify M p of (1) and (3) in Lemma 6.9. d). Then the statement is immediate from the equivalence (cid:3) d = L(lp We now give several counterexamples. It isn't enough to require that kek ≤ 1 and k1−ek ≤ 1 to get a hermitian idempotent, even if A is a σ-finitely representable unital Lp operator algebra. There is an example in M p 2 . Lemma 6.11. Let p ∈ [1, ∞). Define e ∈ M p 2 ( 1 1 k1 − ekp = 1, but if p 6= 2 then e is not a hermitian idempotent. Proof. Let α, β ∈ C. Let q > 1 be such that 1 to (1, 1) ∈ lq q = 1. Applying Holder's inequality 1 1 ). Then kekp = 2 by e = 1 2 and (α, β) ∈ lp 2, we get p + 1 Use this inequality at the third step, to get α + β ≤ 21−1/p(αp + βp)1/p. 2 (α + β), 1 Since α, β ∈ C are arbitrary, this shows that kekp ≤ 1. Obviously, we have kekp ≥ 1 because e2 = e and e 6= 0. 2 (α + β)(cid:1)(cid:13)(cid:13)p = 21/p(cid:12)(cid:12)(cid:12)(cid:12) ke(α, β)kp =(cid:13)(cid:13)(cid:0) 1 The same argument applies to 1 − e. (Or else take s =(cid:0) 1 0 that s is an invertible isometry with ses−1 = 1 − e.) 0 −1(cid:1) and use the fact ≤ (αp+βp)1/p = k(α, β)kp. It follows from Corollary 6.10 that e is not hermitian. (cid:3) α + β 2 (cid:12)(cid:12)(cid:12)(cid:12) One can also explicitly show that e is not hermitian. For example, suppose p < 2. Letting βe be as in Lemma 6.6, one can explicitly show that kβe((1, i))(1, 0)kp = 21/p−1/2 > 1 = k(1, 0)kp. We don't define hermitian idempotents in a nonunital Banach algebra, since whether an idempotent is hermitian depends on the norm used on the unitization, even for Lp operator algebras, as is shown by the following example. CLASSIFICATION OF Lp AF ALGEBRAS 17 Example 6.12. Let p ∈ [1, ∞) \ {2}. Let X be a second countable locally compact Hausdorff space with a nontrivial compact open set K. Let µ a measure on X with full support. Denote by ψ the representation of C0(X) on Lp(X, µ) given by multiplication operators. Let e be the idempotent in Lemma 6.11. Define ρ : C0(X) → L(cid:0)lp unitization the subagebra ρ(C0(X)) ⊕ C1 of L(cid:0)lp 2 ⊗p Lp(X, µ)(cid:1) by ρ(f ) = e ⊗ ψ(f ). Then ρ is isometric. Take as 2 ⊗p Lp(X, µ)(cid:1). Let χK ∈ C0(X) be the characteristic function of K. Then ψ(χK) is a hermitian idempotent in L(Lp(X, µ)), but ρ(χK ) is not a hermitian idempotent because e is not a hermitian idempotent. Lemma 6.13. Let p ∈ [1, ∞) \ {2}. Let A and B be unital σ-finitely representable Lp operator algebras with k1Ak = 1 and k1Bk = 1, and let ψ : A → B be a contractive homomorphism such that ψ(1) is a hermitian idempotent in B. Let e ∈ A be a hermitian idempotent. Then ψ(e) is a hermitian idempotent in B. We don't know to what extent the hypotheses can be weakened. But some hypothesis is necessary. Let p ∈ [1, ∞)\{2}. By Lemma 6.11 there is a nonhermitian idempotent e ∈ M p 2 defined by λ 7→ λe is contractive but sends the hermitian idempotent 1 to the nonhermitian idempotent e. 2 such that kekp = 1. The homomorphism C → M p Proof of Lemma 6.13. We may assume that there is a σ-finite measure space (Y, C, ν) such that B is a unital subalgebra of L(Lp(Y, ν)). Lemma 6.9 provides a measur- able subset E ⊂ X such that ψ(1) is multiplication by χE. By Corollary 2.4, we may view ψ as a unital contractive homomorphism from A to L(Lp(E, νE)). Let βe : C ⊕ C → A be as in Lemma 6.6. Then βe is contractive, so ψ ◦ βe is contractive. By Lemma 6.7, it follows that ψ(e) is a hermitian idempotent in L(Lp(E, νE)). Lemma 6.9 provides a measurable subset F ⊂ E such that ψ(e) is multiplication by χF . Another application of Lemma 6.9 implies that ψ(e) is a hermitian idempotent in L(Lp(Y, ν)). So ψ(e) is hermitian in B by Lemma 6.6. (cid:3) Corollary 6.14. Let p ∈ [1, ∞) \ {2}. Let (X, B, µ) be a σ-finite measure space, let N ∈ Z>0, and let e1, e2, . . . , eN ∈ L(Lp(X, µ)) be orthogonal hermitian idem- potents. Then: (1) There exist disjoint measurable sets E1, E2, . . . , EN ⊂ X such that ek is multiplication by χEk for k = 1, 2, . . . , N . N ek is a hermitian idempotent. (2) Xk=1 (3) For every ξ ∈ Lp(X, µ), we have kξkp p = kekξkp p. N Xk=1 (4) The map β : CN → L(Lp(X, µ)), given by β(λ1, λ2, . . . , λN ) = λ1, λ2, . . . , λN ∈ C, is a contractive homomorphism. λkek for N Xk=1 Proof. Let ρ : L∞(X, µ) → L(Lp(X, µ)) be the representation by multiplication operators. Lemma 6.9 provides measurable sets F1, F2, . . . , FN ⊂ X such that ek = ρ(χFk ) for k = 1, 2, . . . , N . Since ejek = 0 for j 6= k, we have µ(Fj ∩ Fk) = 0 for j 6= k. So there exist disjoint measurable sets Ek ⊂ Fk for k = 1, 2, . . . , N such that µ(Fk \ Ek) = 0. Then ρ(χEk ) = ρ(χFk ). This proves (1). Part (2) follows 18 N. CHRISTOPHER PHILLIPS AND MARIA GRAZIA VIOLA from Lemma 6.9 because setting E = Ek gives N [k=1 N Xk=1 ek = ρ(χE). Part (3) is N immediate. For (4), define β0 : CN → L∞(X, µ) by β0(λ1, λ2, . . . , λN ) = λkχEk Xk=1 for λ1, λ2, . . . , λN ∈ C. Then β0 and ρ are contractive homomorphisms, and β = ρ ◦ β0. (cid:3) Lemma 6.15. Let p ∈ [1, ∞)\{2}. Let A be a separable unital Lp operator algebra, let N ∈ Z>0, and let e1, e2, . . . , eN ∈ A be orthogonal hermitian idempotents such that N Xk=1 ek = 1. Assume a ∈ A satisfies aek = eka for k = 1, 2, . . . , N . Then kak = max(cid:0)ke1ae1k, ke2ae2k, . . . , keN aeN k(cid:1). Proof. We prove this when N = 2, e1 = e, and e2 = 1 − e. The general case follows by induction using Corollary 6.14. We may assume that e 6= 0. It is immediate from Lemma 6.6 that kek = 1. Proposition 2.9 provides an isometric unital representation π of A on a separable Lp space Lp(X, µ). By Lemma 2.7 we may assume that µ is σ-finite. Lemma 6.7 implies that π(e) is a hermitian idempotent, so Lemma 6.9 provides a measurable set E ⊂ X such that π(e) is multiplication by χE. Thus π(a) commutes with multiplication by χE. With respect to the Lp direct sum decomposition Lp(X, µ) = Lp(E, µ) ⊕p Lp(X \ E, µ), we get π(a) = π(eae) ⊕ π(cid:0)(1 − e)a(1 − e)(cid:1). So kak = kπ(a)k = max(cid:0)kπ(eae)k, kπ((1 − e)a(1 − e))k(cid:1) = max(cid:0)keaek, k(1 − e)a(1 − e)k(cid:1). This completes the proof. (cid:3) It follows that if A is a separable unital Lp operator algebra and e ∈ A is a central hermitian idempotent, then A = eAe ⊕ (1 − e)A(1 − e), normed as in Definition 5.2. We don't know whether this is true for more general Banach algebras. Lemma 6.16. In Definition 5.2, suppose that A1, A2, . . . , AN are separable unital Lp operator algebras whose identities have norm 1, and that they have unique Lp operator matrix norms. Then A = unique Lp operator matrix norms. N Mk=1 Ak, normed as in Definition 5.2, has Proof. For k = 1, 2, . . . , N choose some unital isometric representation of Ak on a separable Lp space of a σ-finite measure, and equip Ak with the matrix norms on its image under this representation as in Definition 3.15. Then make A a matrix normed algebra as in Definition 5.2. In view of Lemma 4.2, it suffices to prove that if (X, B, µ) is a σ-finite measure space with Lp(X, µ) separable, n ∈ Z>0, and π : A → L(Lp(X, µ)) is a unital isometric representation, then idMn ⊗ π : Mn(A) → L(cid:0)lp n ⊗p Lp(X, µ)(cid:1) is isometric. For k = 1, 2, . . . , N , we identify Ak with its image in A, and we let fk be the identity of Ak. Set Z = {1, 2, . . . , N }. Let ϕ : C(Z) → A be the unital It is obvious homomorphism determined by ϕ(χ{k}) = fk for k = 1, 2, . . . , N . CLASSIFICATION OF Lp AF ALGEBRAS 19 N [k=1 from the definition of the norm on A that ϕ is isometric. Then π ◦ ϕ is isomet- ric, from which it easily follows that (π ◦ ϕ)(χ{k}) is a hermitian idempotent for k = 1, 2, . . . , N . Let E1, E2, . . . , EN ⊂ X be the disjoint sets corresponding to k=1 as in Corollary 6.14(1). Since π ◦ ϕ is unital, the idempotents (cid:0)(π ◦ ϕ)(χ{k})(cid:1)N we can assume without loss of generality that Ek = X. For k = 1, 2, . . . , N , let πk : Ak → L(Lp(Ek, µ)) be the unital representation gotten as in Corollary 2.4 from πAk . Corollary 2.4 and the definition of the norm on A imply that πk is isometric. It is immediate that lp n ⊗p Lp(X, µ) is the Lp direct sum of the spaces lp n ⊗p Lp(Ek, µ) for k = 1, 2, . . . , N . It follows that if a = (a1, a2, . . . , aN ) ∈ then k(idMn ⊗ π)(a)k Mn(Ak) = Mn(A), N Mk=1 = max(cid:0)k(idMn ⊗ π1)(a1)k, k(idMn ⊗ π2)(a2)k, . . . , k(idMn ⊗ πN )(aN )k(cid:1). The hypotheses imply that idMn ⊗ πk is isometric for k = 1, 2, . . . , N . Definition 5.2 therefore implies that idMn ⊗ π is isometric. (cid:3) The following lemma will be used in connection with representations of nonunital spatial Lp AF algebras. Lemma 6.17. Let p ∈ [1, ∞) \ {2}. Let (X, B, µ) be a σ-finite measure space such that Lp(X, µ) is separable. Let e1, e2, . . . ∈ L(Lp(X, µ)) be idempotents, and take e0 = 0. Assume that, for all n ∈ Z>0, kenk ≤ 1 and en−1 is a hermitian idempotent in enL(Lp(X, µ))en. Then there are an idempotent e ∈ L(Lp(X, µ)), a σ-finite measure space (Y, C, ν) such that Lp(Y, ν) is separable, an isometric linear map s : Lp(Y, ν) → Lp(X, µ), and measurable subsets Y1, Y2, . . . ⊂ Y , such that: enLp(X, µ). ∞ [n=1 (1) kek ≤ 1. (2) eLp(X, µ) = (5) Y = Yn. ∞ an=1 (3) For every n ∈ Z>0, en is a hermitian idempotent in eL(Lp(X, µ))e. (4) ran(s) = eLp(X, µ). (6) For every n ∈ Z>0, sLp(Yn, νYn ) = (en − en−1)Lp(X, µ). We do not assume that en is a hermitian idempotent in L(Lp(X, µ)), and the conclusion does not claim that e is a hermitian idempotent in L(Lp(X, µ)). Proof of Lemma 6.17. For n ∈ Z>0 define En = enLp(X, µ) ⊂ Lp(X, µ). Set ∞ E = En. For n ∈ Z>0, use Proposition 2.8 to find a σ-finite measure space (Zn, Dn, λn) such that Lp(Zn, λn) is isometrically isomorphic to En. Also, define πn : L(En) → L(Lp(X, µ)) by πn(a)ξ = aenξ for a ∈ L(En) and ξ ∈ Lp(X, µ). Since enaenξ = aenξ for a ∈ L(En), one checks easily that πn is a (nonunital) homomorphism. An application of Corollary 2.4 shows that πn is isometric. Thus, L(cid:0)Lp(Zn, λn)(cid:1) is isometrically isomorphic to enL(Lp(X, µ))en. [n=1 20 N. CHRISTOPHER PHILLIPS AND MARIA GRAZIA VIOLA For m, n ∈ Z>0 with m ≤ n, a similar argument shows that the analogous map πn,m : L(Em) → L(En) is an isometric homomorphism. It follows from Remark 6.5 that em − em−1 is a hermitian idempotent in emL(Lp(X, µ))em, and then it follows from Lemma 6.13 and induction on n that em − em−1 is a hermitian idempotent in enL(Lp(X, µ))en. For n ∈ Z>0, Corollary 6.14(3), applied to e1 − e0, e2 − e1, . . . , en − en−1, shows that for every ξ ∈ En we have (6.2) kξkp p = n Xk=1 k(ek − ek−1)ξkp p. For n ∈ Z>0, use Proposition 2.8 to find a σ-finite measure space (Yn, Cn, νn) and an isometric isomorphism sn : Lp(Yn, λn) → (en − en−1)Lp(X, µ). Following ∞ the notation of Definition 5.1, set (Y, C, ν) = (Yn, Cn, νn). By (6.2), the map from Lp n ak=1 an=1 (Yk, Ck, νk)! to enLp(X, µ), given by Xk=1 (η1, η2, . . . , ηn) 7→ n sk(ηk), is an isometric isomorphism. Combining these for n ∈ Z>0 and extending by continuity, we get an isometric linear map s : Lp(Y, ν) → Lp(X, µ), whose range must be equal to E. We now have the objects s, (Y, C, ν), and Y1, Y2, . . . ⊂ Y of the conclusion, as well as parts (5) and (6). If we use E in place of eLp(X, µ), we also have parts (2) and (4) of the conclusion. Now let ξ ∈ Lp(X, µ). For n ∈ Z>0 we use kenk ≤ 1 and (6.2) to get kξkp p ≥ kenξkp p = k(ek − ek−1)ξkp p. n Xk=1 k(ek − ek−1)ξkp p converges. For m, n ∈ Z>0 with m ≤ n, we get n Therefore ∞ Xk=1 k(en − em)ξkp p = k(ek − ek−1)ξkp p, so (enξ)n∈Z>0 is a Cauchy sequence in Lp(X, µ). Thus enξ exists. Call this limit eξ. By taking suitable limits, one checks that e is linear, kek ≤ 1, e2 = e, and ran(e) = E. We now have all the required objects for the conclusion, and all the conditions except (3). lim n→∞ For (3), use continuity to get kenξkp p for all ξ ∈ E and n ∈ Z>0. From this, it is easy to see that the map (λ1, λ2) 7→ λ1en + λ2(e − en) is a contraction from C ⊕ C to eL(Lp(X, µ))e. Apply Lemma 6.6. (cid:3) p + k(e − en)ξkp p = kξkp 7. Direct limits The main result in this section is that the direct limit of matricial Lp operator algebras is also a matricial Lp operator algebra. Moreover, if each algebra in the system has unique Lp operator matrix norms, and the connecting maps of the direct Xk=m+1 CLASSIFICATION OF Lp AF ALGEBRAS 21 system are isometric, then the direct limit also has unique Lp operator matrix norms. Definition 7.1. Let I be an infinite directed set. A (completely) contractive direct system of Banach algebras indexed by I is a pair (cid:0)(Ai)i∈I , (ϕj,i)i≤j(cid:1) consisting of a family (Ai)i∈I of (matrix normed) Banach algebras and a family (ϕj,i)i≤j of (completely) contractive homomorphisms ϕj,i : Ai → Aj for i, j ∈ I with i ≤ j, such that ϕi,i = idAi for all i ∈ I and ϕk,j ◦ ϕj,i = ϕk,i whenever i, j, k ∈ I satisfy i ≤ j ≤ k. We say that the system is unital if Ai is unital for all i ∈ I and ϕj,i is unital for all i, j ∈ I with i ≤ j. In the contractive case, the direct limit lim −→ i Ai of this direct system is the Banach algebra direct ("inductive") limit, as constructed in Section 3.3 of [1]. In the completely contractive case, for n ∈ Z>0 we use Lemma 3.5 to identify Mn(Ai) up to isomorphism of topological algebras. Then Ai(cid:17) with the norm obtained by applying the contractive case Mn(Ai). Lemma 7.2 below shows that we do indeed get a matrix normed i i −→ Ai(cid:17) with lim Mn(cid:16) lim we equip Mn(cid:16) lim −→ −→ i to lim −→ i Banach algebra this way. Lemma 7.2. Let (cid:0)(Ai)i∈I , (ϕj,i)i≤j(cid:1) be a completely contractive direct system of matrix normed Banach algebras. Then lim −→ Ai is a matrix normed Banach algebra, and for every j ∈ I the standard homomorphism ϕj : Aj → lim −→ Ai is completely i i contractive. Proof. The statement about complete contractivity follows from the identification of the matrix norms. For every n ∈ Z>0 and i ∈ I identify B(n) j,i = idMn ⊗ ϕj,i : B(n) ϕ(n) maps induced by ϕj,i and ϕi. Set B(n) = [i∈I i → B(n) and ϕ(n) j i = Mn(Ai) with Mn ⊗ Ai, and let i → Mn(A) be the i = idMn ⊗ ϕi : B(n) ϕ(n) i (cid:0)B(n) i (cid:1). We claim that the norms on B(n), for n ∈ Z>0, obtained by viewing B(n) as a subalgebra of the Banach algebra direct limit lim −→ , are a system of matrix norms as in Definition 3.2. B(n) i i Let b ∈ B(n), and choose i0 ∈ I and a ∈ B(n) i0 such that ϕ(n) i0 (a) = b. Let σ and τ be injective functions as in Definition 3.2(1). Since Mm(cid:16) lim −→ Definition 3.2(1) in Mn(Ai) = B(n) i we have B(m) i , using (cid:13)(cid:13)(bσ(j),τ (k))1≤j,k≤m(cid:13)(cid:13)m = lim ≤ lim i i −→ Ai(cid:17) = lim i,i0(cid:0)(aσ(j),τ (k))1≤j,k≤m(cid:1)(cid:13)(cid:13)m i,i0(cid:0)(aj,k)1≤j,k≤n(cid:1)(cid:13)(cid:13)n = kbkn. i (cid:13)(cid:13)ϕ(m) i (cid:13)(cid:13)ϕ(n) Moreover, given λ1, λ2, . . . , λn ∈ C, if we set s = diag(λ1, λ2, . . . , λn), we have ϕi,i0 (sa) = sϕi,i0 (a), so, using Definition 3.2(2) on Mn(Ai), ksbkn = lim i kϕi,i0 (sa)kn = lim i ksϕi,i0 (a)kn ≤ max(cid:0)λ1, λ2, . . . , λn(cid:1) lim i kϕi,i0 (a)kn = max(cid:0)λ1, λ2, . . . , λn(cid:1)kbkn. 22 N. CHRISTOPHER PHILLIPS AND MARIA GRAZIA VIOLA Similarly kbskn ≤ max(cid:0)λ1, λ2, . . . , λn(cid:1)kbkn. Lastly, given b1 ∈ B(m) and b2 ∈ B(n), there exist i0 ∈ I, a1 ∈ B(m) (a1) and b2 = ϕ(n) , and i0 (a2). Therefore, diag(b1, b2) = i0 a2 ∈ B(n) i0 ϕ(m+n) (diag(a1, a2)) and such that b1 = ϕ(m) i0 i0 i,i0 = lim (cid:13)(cid:13)diag(b1, b2)(cid:13)(cid:13)m+n = lim i (cid:13)(cid:13)ϕ(m+n) i (cid:13)(cid:13)diag(cid:0)ϕ(m) max(cid:0)(cid:13)(cid:13)ϕ(m) Since B(n) is dense in Mn(cid:16) lim (diag(a1, a2))(cid:13)(cid:13)m+n (a2)(cid:1)(cid:13)(cid:13)m+n i,i0 (a1)(cid:13)(cid:13)m, (cid:13)(cid:13)ϕ(n) i,i0 (a2)(cid:13)(cid:13)n(cid:1) = max(kb1km, kb2kn). Ai(cid:17) for all n ∈ Z>0, the conditions of Defini- This completes the proof of the claim. (a1), ϕ(n) i,i0 = lim −→ i,i0 i i (cid:3) Ai follow by continuity. tion 3.2 for lim −→ i system of Lp operator algebras. Then lim −→ i Theorem 7.3. Let p ∈ [1, ∞). Let (cid:0)(Ai)i∈I , (ϕj,i)i≤j(cid:1) be a contractive direct Theorem 7.4. Let p ∈ [1, ∞). Let (cid:0)(Ai)i∈I , (ϕj,i)i≤j(cid:1) be a completely contrac- tive direct system of matricial Lp operator algebras. Then lim −→ i Ai is an Lp operator algebra. Ai is a matricial Lp operator algebra. The proofs are essentially the same. We prove Theorem 7.3 here. We describe the modifications for the proof of Theorem 7.4 afterwards. Proof of Theorem 7.3. The statement is trivial if I has a largest element. Other- wise, let U0 be the collection of all subsets of I of the form (cid:8)i ∈ I : i ≥ i0(cid:9) for i0 ∈ I. These sets are nonempty. Since I is directed, the intersection of any finite collection of them contains another one. Since I has no largest element, for every i ∈ I there is S ∈ U0 such that i 6∈ S. Therefore there is a free ultrafilter U on I which contains U0. By definition, for every i ∈ I there exists a measure space (Xi, Bi, µi) and an isometric representation ρi : Ai → L(Lp(Xi, µi)). Let M be the Banach space ul- traproduct(cid:16)Yi∈I Lp(Xi, µi)(cid:17).U (Definition 2.1 of [12]). By Theorem 3.3(ii) of [12], there exists a measure space (X, B, µ) such that M is isometrically isomorphic to Lp(X, µ). So it suffices to find an isometric representation of lim −→ i Ai on M . Let B be the algebraic direct limit of the algebras Ai, and for i ∈ I let ϕi : Ai → B be the homomorphism associated to the direct system. Equip B with the direct limit seminorm, and let A = lim Ai be the completion of B/{b ∈ B : kbk = 0}, with −→ the obvious isometric map κ : B → A. i We will construct an isometric representation γ of B on M . (It will not be injective; rather, its kernel will be {b ∈ B : kbk = 0}.) Let x ∈ B. Choose i ∈ I and a ∈ Ai such that ϕi(a) = x. We give an associated operator yl ∈ L(Lp(Xl, µl)) for each l ∈ I. If l ≥ i, set yl = ρl(ϕl,i(a)). Otherwise, set yl = 0. Clearly kylk ≤ kak for all l ∈ I, so the ultraproduct of operators (Definition 2.2 of [12]) gives an CLASSIFICATION OF Lp AF ALGEBRAS 23 operator y = (yl)U ∈ L(M ) such that kyk = lim U cofinal in I, we have kylk. Since J = {l ∈ I : l ≥ i} is lim l∈I kylk = lim l∈J kρl(ϕl,i(a))k = lim l∈J kϕl,i(a)k = kxk. The choice of U ensures that lim U kylk = lim l∈I kylk. Therefore kyk = kxk. We claim that y does not depend on the choices of i and a ∈ Ai. To prove this, suppose that j ∈ I and b ∈ Aj also satisfy ϕj (b) = x. Let zl ∈ L(Lp(Xl, µi)) for l ∈ I, and z = (cid:0)zl(cid:1)U ∈ L(M ), be defined in the same way as yl above, but using j and b in place of i and a. Choose k ∈ I such that k ≥ i, k ≥ j, and ϕk,i(a) = ϕk,j (b). Then zl = yl for all l ∈ I with l ≥ k, and {l ∈ I : l ≥ k} ∈ U, so z = y. The claim is proved. It follows that there is a well defined isometric map γ : B → L(M ) such that if x ∈ B and i ∈ I and a ∈ Ai satisfy ϕi(a) = x, then γ(x) is the element y constructed above. Using directedness of I, it is easy to prove that γ is a homomorphism. Since γ is isometric, we have γ(x) = 0 whenever kxk = 0, and there exists a unique isometric homomorphism ρ : A → L(M ) such that ρ(κ(x)) = γ(x) for all x ∈ B. The existence of ρ shows that A is an Lp operator algebra. (cid:3) Proof of Theorem 7.4. We describe the differences from the proof of Theorem 7.3. We choose the maps ρi in the proof of Theorem 7.3 to be completely isometric, not just isometric. Let m ∈ Z>0 and let ν be counting measure on {1, 2, . . . , m}. One can check that the obvious map gives an isometric isomorphism Yi∈I Mm(cid:16) lim Lp({1, 2, . . . , m} × Xi, ν × µi)(cid:1)!.U → Lp({1, 2, . . . , m} × X, ν × µ)(cid:1). Ai(cid:17) ∼= lim (This is a direct computation from the definitions.) Using the standard isomorphism Mm(Ai), the argument used in the proof of Theorem 7.3 to show that γ is isometric now shows that idMm ⊗ γ is isometric. Since this is true for all m ∈ Z>0, we conclude that γ, hence also ρ, is completely isometric. (cid:3) −→ i −→ i Proposition 7.5. Let p ∈ [1, ∞), and let(cid:0)(Ai)i∈I , (ϕj,i)i≤j(cid:1) be as in Theorem 7.3. Suppose further that I is countable, that for all i ∈ I the algebra Ai is separable and has unique Lp operator matrix norms, and that for all i, j ∈ I with i ≤ j, Ai has unique Lp operator matrix norms. the map ϕj,i is isometric. Then A = lim −→ They are obtained by equipping Ai with its unique Lp operator matrix norms for i ∈ I and, for each n ∈ Z>0, giving Mn(A) the norm coming from the contractive case of Definition 7.1 applied to lim −→ Mn(Ai). i i Proof. The hypotheses imply that A is separable. For i ∈ I, equip Ai with its unique Lp operator matrix norms. The hypotheses imply that if j ∈ I and j ≥ i, then ϕj,i is completely isometric. Equip A with the matrix norms in the statement. Then A is an Lp operator algebra by Theorem 7.4. Let (X, B, µ) be a σ-finite measure space with Lp(X, µ) separable, and let π : A → L(Lp(X, µ)) be isometric. We show that π is completely isometric. Let n ∈ Z>0. For i ∈ I, let ϕi : Ai → A be the homomorphism coming from the direct system. Then ϕi is isometric, so π ◦ ϕi is isometric. The hypothesis on Ai implies that 24 N. CHRISTOPHER PHILLIPS AND MARIA GRAZIA VIOLA idMn ⊗ (π ◦ ϕi) is isometric. Since [i∈I is isometric. ϕi(Ai) is dense in A, it follows that idMn ⊗ π (cid:3) 8. Spatial semisimple finite dimensional Lp operator algebras In this section we introduce our setup by giving the definitions of spatial semisim- ple finite dimensional Lp operator algebras and spatial homomorphisms. We also give a characterization of spatial homomorphisms between spatial semisimple fi- nite dimensional Lp operator algebras in terms of block diagonal homomorphisms (Lemma 8.20), and show that any spatial semisimple finite dimensional Lp operator algebra has unique Lp operator matrix norms. Definition 8.1. Let B be a unital Banach algebra, and let b, c ∈ B. We say that b and c are isometrically similar if there is an invertible isometry s ∈ B such that c = sbs−1. If A is also a Banach algebra, and ϕ, ψ : A → B are linear maps, we say that ϕ and ψ are isometrically similar if there is an invertible isometry s ∈ B such that ψ(a) = sϕ(a)s−1 for all a ∈ A. We state some immediate properties. Proposition 8.2. Let A and B be Banach algebras, with B unital, and let ϕ, ψ : A → B be isometrically similar linear maps. (1) If ϕ is contractive then so is ψ. (2) If ϕ is isometric then so is ψ. (3) If A and B are matrix normed (Definition 3.2) then idMn ⊗ ϕ and idMn ⊗ ψ are isometrically similar. (4) If A and B are matrix normed and ϕ is completely bounded, then so is ψ. (5) If A and B are matrix normed and ϕ is completely contractive, then so is ψ. (6) If A and B are matrix normed and ϕ is completely isometric, then so is ψ. Proof. The only part requiring proof is (3). For this part, we use Definition 3.2(3) to see that if s ∈ B is an invertible isometry, then so is 1 ⊗ s ∈ Mn ⊗ B for any n ∈ Z>0. (cid:3) Proposition 8.3. Let B be a unital Banach algebra, and let e, f ∈ B be isomet- rically similar idempotents. If e is hermitian (Definition 6.4), so is f . Proof. The corresponding homomorphisms in Lemma 6.6 are isometrically similar. Apply Proposition 8.2(1). (cid:3) Definition 8.4. Let p ∈ [1, ∞) \ {2}. Let A be a unital σ-finitely representable Lp operator algebra, let d ∈ Z>0, and let ϕ : M p d → A be a homomorphism (not necessarily unital). We say that ϕ is spatial if ϕ(1) is a hermitian idempotent (Definition 6.4) and ϕ is contractive. The zero homomorphism is allowed as a choice of ϕ. It isn't enough to merely require that ϕ be contractive. Example 8.5. Let p ∈ [1, ∞) \ {2}. Then there is a contractive homomorphism ϕ : C → M p 2 which is not spatial. To construct one, let e be the nonhermitian idempotent from Lemma 6.11. Define ϕ(λ) = λe for λ ∈ C. Then ϕ is clearly contractive, but not spatial because e is not hermitian. CLASSIFICATION OF Lp AF ALGEBRAS 25 Lemma 8.6. Let p ∈ [1, ∞) \ {2}, let d ∈ Z>0, and let s ∈ M p d . Then s is an in- vertible isometry if and only if s is a complex permutation matrix (Definition 3.10). Proof. It is obvious that complex permutation matrices are invertible isometries. Conversely, assume that s is an invertible isometry. It follows from Lemma 6.16 of [16] that s is spatial, and it is easily seen from the definitions (Definition 6.3 and Definition 6.4 of [16]) that s is a complex permutation matrix. (cid:3) The next lemma shows that every spatial homomorphism between two matrix algebras is isometrically similar to a block diagonal homomorphism. Lemma 8.7. Let p ∈ [1, ∞) \ {2}. Let d, m ∈ Z>0, and let ψ : M p d → M p homomorphism (not necessarily unital). Then the following are equivalent: m be a (1) ψ is spatial. (2) There exist k ∈ Z>0 with 0 ≤ kd ≤ m such that ψ is isometrically similar to the homomorphism a 7→ diag(a, a, . . . , a, 0), the block diagonal matrix in which a occurs k times and 0 is the zero element of M p m−kd. Proof. It is easy to check that (2) implies (1). So assume (1). First assume that ψ is unital. Then m = kd for some k ∈ Z>0. The implication from (4) to (8) in Theorem 7.2 in [16] provides a σ-finite measure space (Y, C, ν) and a bijective isometry u : lp d ⊗p Lp(Y, ν) → lp kd such that for all a ∈ M p have dimension k, so it is isometrically isomorphic to lp d we have ψ(a) = u(a ⊗ 1)u−1. The space Lp(Y, ν) must k. There is a bijection {1, 2, . . . , d} × {1, 2, . . . , k} → {1, 2, . . . , m} d ⊗p lp k) with L(lp m such that sψ(a)s−1 = diag(a, a, . . . , a) for all a ∈ M p d . such that the corresponding isomorphism of L(lp m) sends a ⊗ 1 to diag(a, a, . . . , a). This allows us to identify s = u−1 with an invertible isometry in L(lp m) ∼= M p Now consider the general case. Assume that ψ is spatial. So ψ is contractive and ψ(1) is spatial. By Lemma 6.9, there exists a measurable subset E ⊂ {1, 2, . . . , m} such that ψ(1) is multiplication by χE on lp m. By conjugating by a permutation matrix, we can assume that E = {1, 2, . . . , n} for some n ∈ {1, 2, . . . , m}. Identify lp n with lp(E) ⊂ lp n). (In matrix form, this is ι(b) = diag(b, 0).) There is a homomorphism ϕ : L(lp n) such that ψ(a) = ι(ϕ(a)) = ϕ(a) ⊕ 0 for all a ∈ L(lp d), and ϕ is a unital homomorphism from M p n which is contractive by Corollary 2.4. By the case done above, there exist k ∈ Z>0 such that n = kd and an invertible isometry s0 ∈ M p n such that s0ϕ(a)s−1 d). Then s = diag(s0, 1), with 1 being the identity of M p d) we have sψ(a)s−1 = diag(a, a, . . . , a, 0). (cid:3) m−n, is an invertible isometry such that for all a ∈ L(lp 0 = diag(a, a, . . . , a) for all a ∈ L(lp m, and let ι : L(lp n) → L(lp m) be ι(b) = b ⊕ 0 for b ∈ L(lp d) → L(lp d to M p To define a spatial Lp AF algebra, we need to first define its building blocks, the spatial semisimple finite dimensional Lp operator algebras. Definition 8.8. Let p ∈ [1, ∞) \ {2}. A Banach algebra A is called a spa- tial semisimple finite dimensional Lp operator algebra if there are N ∈ Z>0 and d1, d2, . . . , dN ∈ Z>0 such that A is isometrically isomorphic to with the norm as in Definition 5.2. M p dk , endowed N Mk=1 26 N. CHRISTOPHER PHILLIPS AND MARIA GRAZIA VIOLA Remark 8.9. Let p ∈ [1, ∞) \ {2}. To simplify the notation in our proofs, if A is a spatial semisimple finite dimensional Lp operator algebra, we will omit the isometric isomorphism and simply write A = M p dk N Mk=1 with N, d1, d2, . . . , dN ∈ Z>0. Lemma 8.10. Let p ∈ [1, ∞)\{2}. Let A be a spatial semisimple finite dimensional Lp operator algebra, and let J ⊂ A be an ideal. Then A/J, with the quotient norm, is a spatial semisimple finite dimensional Lp operator algebra. Proof. Use the notation in Remark 8.9. Then there is a subset S ⊂ {1, 2, . . . , N } . The quotient norm agrees with the norm M p dk such that, as an algebra, A/J =Mk∈S on Mk∈S by Lemma 5.5. M p dk (cid:3) M p dk be a spatial semisimple N Mk=1 Lemma 8.11. Let p ∈ [1, ∞) \ {2}, and let A = finite dimensional Lp operator algebra. Let s = (s1, s2, . . . , sN ) ∈ A. Then s is an invertible isometry if and only if sk is a complex permutation matrix for k = 1, 2, . . . , N . Proof. This is immediate from Lemma 8.6 and the definition of the norm on A. (cid:3) Lemma 8.12. Let p ∈ [1, ∞) \ {2}, and let A be a spatial semisimple finite dimen- sional Lp operator algebra. Then A is an Lp operator algebra with unique Lp op- erator matrix norms, obtained by combining Definition 5.2 and Definition 3.15. Proof. That A is an Lp operator algebra follows from Lemma 5.7. That A has unique Lp operator matrix norms follows from Corollary 4.4 and Lemma 6.16. (cid:3) The maps we will consider between spatial semisimple finite dimensional Lp op- erator algebras are the spatial homomorphisms. N Definition 8.13. Let p ∈ [1, ∞)\{2}, and let A = M p dk be a spatial semisimple finite dimensional Lp operator algebra. Let B be a σ-finitely representable unital Lp operator algebra, and let ϕ : A → B be a homomorphism. We say that ϕ is spatial if for k = 1, 2, . . . , N , the restriction of ϕ to the summand M p is spatial in dk the sense of Definition 8.4. Mk=1 Lemma 8.14. Let p ∈ [1, ∞) \ {2}, and let A be a spatial semisimple finite dimen- sional Lp operator algebra. Let B be a σ-finitely representable unital Lp operator algebra, and let ϕ : A → B be a homomorphism. Then ϕ is spatial if and only if ϕ(1) is a hermitian idempotent (Definition 6.4) and ϕ is contractive. Proof. We can assume without loss of generality that B is a unital subalgebra of L(Lp(Y, ν)) for some σ-finite measure space (Y, C, ν). Assume also that A = For k = 1, 2, . . . , N , let ιk : M p dk → M p dl N Ml=1 be the inclusion of the k-th summand M p dl . N Ml=1 CLASSIFICATION OF Lp AF ALGEBRAS 27 into A. Let ρ : L∞(Y, ν) → L(Lp(Y, ν)) be the representation by multiplication operators. Suppose that ϕ is spatial. For k = 1, 2, . . . , N , the homomorphism ϕMdk is spa- )) is a hermitian idempotent. The idempotents e1, e2, . . . , eN tial, so ek = ϕ(ιk(1Mdk are clearly orthogonal, so Corollary 6.14(2) implies that ϕ(1) = idempotent. ek is a hermitian N Xk=1 Corollary 6.14(1) provides disjoint measurable sets E1, E2, . . . , EN ⊂ Y such that ek = ρ(χEk ) for k = 1, 2, . . . , N . Set E = N Ek. We can identify Lp(Y, ν) N [k=1 with the Lp direct sum Lp(Y \ E, ν) ⊕p Lp(Ek, ν). For l = 1, 2, . . . , N , let ϕl : M p dl for k = 1, 2, . . . , N , then → L(Lp(Y, ν)) be ϕl(a) = ρ(χEl )ϕ(ιl(a))ρ(χEl ) for a ∈ M p dl . If ak ∈ M p dk Mk=1 ϕ(a1, a2, . . . , aN ) = = ϕ(ιk(ak)) = N Xk=1 ρ(χEk )ϕ(ιk(ak))ρ(χEk ) = ϕ(ιk(1Mdk ))ϕ(ιk(ak))ϕ(ιk(1Mdk )) ϕk(ak). N Xk=1 N N Xk=1 Xk=1 Mk=1 Since the sets Ej are disjoint, kϕ(a1, a2, . . . , aN )k = max 1≤k≤N kϕk(ak)k ≤ max 1≤k≤N kakk = k(a1, a2, . . . , aN )k, so ϕ : A → B is contractive. Conversely assume that ϕ is contractive and ϕ(1) is a hermitian idempotent. For ) is a k = 1, 2, . . . , N , it is obvious from Definition 5.2 and Lemma 6.6 that ιk(1Mdk N hermitian idempotent in M p dk . Therefore ϕ(ιk(1Mdk )) is a hermitian idempotent in B by Lemma 6.13. Also, ϕ ◦ ιk is contractive because ιk and ϕ are. So ϕMdk spatial. is (cid:3) Corollary 8.15. Let p ∈ [1, ∞) \ {2}. Let A be a spatial semisimple finite dimen- sional Lp operator algebra, let B and C be unital σ-finitely representable Lp oper- ator algebras, let ϕ : A → B be a spatial homomorphism, and let ψ : B → C be a contractive homomorphism such that ψ(1) is a hermitian idempotent in C. Then ψ ◦ ϕ is spatial. Proof. Lemma 8.14 implies that ψ ◦ ϕ is contractive. It follows from Lemma 8.14 and Lemma 6.13 that (ψ ◦ ϕ)(1) is a hermitian idempotent in C. So ψ ◦ ϕ is spatial by Lemma 8.14. (cid:3) Corollary 8.16. Let p ∈ [1, ∞) \ {2}. Let A be a spatial semisimple finite dimen- sional Lp operator algebra, let B be a unital σ-finitely representable Lp operator algebra, and let ϕ, ψ : A → B be isometrically similar homomorphisms. Then ϕ is spatial if and only if ψ is spatial. Proof. Use Lemma 8.14, Proposition 8.2(1), and Lemma 8.3. (cid:3) 28 N. CHRISTOPHER PHILLIPS AND MARIA GRAZIA VIOLA The following definition is standard, but is given here for reference. M Definition 8.17. Let A = Mcj be a finite direct sum of full matrix algebras. Mj=1 (1) Let d ∈ Z>0, and let ϕ : A → Md be a homomorphism. Then for j = 1, 2, . . . , M the j-th partial multiplicity of ϕ is defined to be mj(ϕ) = rank(ϕ(1Mcj ))/cj. N (2) Let B = Mdk be another finite direct sum of full matrix algebras, and let ϕ : A → B be a homomorphism. For j = 1, 2, . . . , M and k = 1, 2, . . . , N , we denote by mk,j(ϕ) the j-th partial multiplicity of the composition of ϕ with the projection map B → Mdk . We call m(ϕ) = (mk,j(ϕ))k,j the partial multiplicity matrix of ϕ. We use analogous notation for direct sums indexed by finite sets not of the form {1, 2, . . . , M }. Mk=1 Next we define block diagonal homomorphisms between finite direct sums of full matrix algebras. Definition 8.18. Let p ∈ [1, ∞) \ {2}. Let A = full matrix algebras. Mcj be a finite direct sum of M Mj=1 (1) A unital homomorphism ϕ : A → Md is said to be block diagonal if there n exist n ≥ 1 and r(1), r(2), . . . , r(n) ∈ {1, 2, . . . , M }, satisfying cr(k) = d, (2) A nonunital homomorphism ϕ : A → Md is block diagonal if its unitization M Mj=1 Mcj ⊕ C → Md is block diagonal. N Mj=1 (3) Let B = Mdj be a finite direct sum of full matrix algebras. Then a homomorphism ϕ : A → B is block diagonal if for k = 1, 2, . . . , N the homomorphism ϕk : A → Mdk , given by the composition of ϕ and the projection map B → Mdk, is block diagonal. We list some properties of block diagonal homomorphisms. Lemma 8.19. Let p ∈ [1, ∞) \ {2}. (1) Let A, B, and C be finite direct sums of full matrix algebras, and let ϕ : A → B and ψ : B → C be homomorphisms. Then m(ψ ◦ ϕ) = m(ψ)m(ϕ). such that ϕ(a1, a2, . . . , aM ) = ar(1) 0 0 ... 0 0 ar(2) 0 ... 0 0 0 ar(3) ... 0 0 0 0 ... 0 · · · · · · · · · · · · 0 0 0 ... ar(n)   Xk=1  .  CLASSIFICATION OF Lp AF ALGEBRAS 29 (2) Let A and B be spatial semisimple finite dimensional Lp operator algebras. Then every block diagonal homomorphism ϕ : A → B is spatial. (3) If ϕ is as in Definition 8.18(1), then (4) Let A = Mdk be finite direct sums of full matrix al- M mj(ϕ) = card(cid:0)(cid:8)k ∈ {1, 2, . . . , n} : r(k) = j(cid:9)(cid:1). Mj=1 Mcj and B = N Mk=1 gebras, and let m be an N × M matrix with entries in Z≥0. Then the following are equivalent: (a) There exists a block diagonal homomorphism ϕ : A → B such that m(ϕ) = m. (b) There exists a homomorphism ϕ : A → B such that m(ϕ) = m. M (c) For k = 1, 2, . . . , N , we have mk,j cj ≤ dk. Xj=1 (5) The composition of two block diagonal homomorphisms is block diagonal. (6) Let A1, A2, B1, B2 be finite direct sums of full matrix algebras, and let ϕ1 : A1 → B1 and ϕ2 : A2 → B2 be block diagonal homomorphisms. Then ϕ1 ⊕ ϕ2 : A1 ⊕ A2 → B1 ⊕ B2 is block diagonal. M N (7) Let A = Mcj and B = Mdk be finite direct sums of full matrix al- Mj=1 Mk=1 gebras, let ϕ : A → B be a block diagonal homomorphism, and let r ∈ Z>0. M N Make the identifications Mr ⊗ A = Mrcj and Mr ⊗ B = Mrdk, by Mj=1 Mk=1 using on each summand the isomorphism θσ of Definition 3.12 with σ taken to be the standard choice of bijection as given there. Then idMr ⊗ϕ is block diagonal. (8) Let A and B be spatial semisimple finite dimensional Lp operator algebras, and let ϕ : A → B be block diagonal. Then ϕ is completely contractive. N Proof. We first prove (2). Write B = M p dk , and for k = 1, 2, . . . , N let πk : B → be the projection map. Block diagonal maps to M p dk M p are clearly contractive, so dk πk ◦ ϕ is contractive. Thus ϕ is contractive by Lemma 5.4. For k = 1, 2, . . . , N , the matrix (πk ◦ ϕ)(1) is diagonal with entries in {0, 1}. So (πk ◦ ϕ)(1) is a hermitian idempotent by Corollary 6.10. Now ϕ(1) is a hermitian idempotent by Lemma 6.8. Use Lemma 8.14. Part (8) follows from part (7) and part (2). Everything else is either well known or immediate. L N Lemma 8.20. Let p ∈ [1, ∞) \ {2}, let A = M p cj and B = (cid:3) M p dk be spatial Mk=1 Mj=1 Mk=1 semisimple finite dimensional Lp operator algebras, and let ϕ : A → B be a ho- momorphism. Then ϕ is spatial if and only if ϕ is isometrically similar to a block diagonal homomorphism. Proof. If ϕ is isometrically similar to a block diagonal homomorphism, then ϕ is spatial by Lemma 8.19(2) and Corollary 8.16. 30 N. CHRISTOPHER PHILLIPS AND MARIA GRAZIA VIOLA Conversely, assume that ϕ is spatial. Since the projection map πk : B → M p dk is contractive and πk(1) is a hermitian idempotent, Corollary 8.15 implies that πk ◦ ϕ is spatial. Therefore, it is enough to prove the claim when B = M p d for some d ∈ Z>0. For j = 1, 2, . . . , M let ιj : M p cj → A be the inclusion map. Since ϕ ◦ ιj is spatial (by Lemma 8.14 and Corollary 8.15), it follows from Corollary 6.14 that there are disjoint subsets E1, E2, . . . , EM ⊂ {1, 2, . . . , d} ) is multiplication by χEj for j = 1, 2, . . . , M . Let ρ be the such that (ϕ ◦ ιj)(1M p cj representation of C({1, 2, . . . , d}) on lp d by multiplication operators. Set d0 = 0 and choose a permutation σ of {1, 2, . . . , d} and numbers d1, d2, . . . , dM ∈ {1, 2, . . . , d} such that for j = 1, 2, . . . , M we have σ(Ej ) = (dj−1, dj] ∩ Z. Let s0 ∈ M p d be the corresponding permutation matrix, satisfying s0ρ(χEj )s−1 0 = ρ(χσ(Ej )) for j = 1, 2, . . . , M . Corollary 8.16 implies that the map a 7→ s0ϕ(a)s−1 0 is spatial. For 1 ≤ j ≤ M make the obvious identification ρ(χ(dj−1, dj ]∩Z)M p d ρ(χ(dj−1, dj]∩Z) = M p dj−dj−1 . Since s0(ϕ ◦ ιj)(1)s−1 0 = ρ(χ(dj−1, dj ]∩Z), by Corollary 2.4 the formula s0(ϕ ◦ ιj)(·)s−1 phism ψj : M p dj−dj−1 permutation matrix sj ∈ M p homomorphism from M p cj → M p dj−dj−1 cj to M p dj−dj−1 0 defines a contractive unital homomor- . It follows from Lemma 8.7 that there is a complex is a block diagonal such that a 7→ sjψj(a)s−1 j for 1 ≤ j ≤ M . Set s = [diag(s1, s2, . . . , sM , 1d−dM )] · s0, which is a complex permutation matrix in M p d . Since s0ϕ(a1, a2, . . . , aM )s−1 for a = (a1, a2, . . . , aM ) ∈ L 0 = diag(cid:0)ψ1(a1), ψ2(a2), . . . , ψM (aM ), 0d−dM(cid:1) Mj=1 cj , it follows that a 7→ sϕ(a)s−1 is block diagonal. M p (cid:3) 9. Spatial Lp AF algebras We define spatial Lp AF algebras and show that any spatial Lp AF algebra is a separable nondegenerately representable Lp operator algebra. Definition 9.1. Let p ∈ [1, ∞)\{2}. A spatial Lp AF direct system is a contractive direct system with index set Z≥0 (that is, a pair (cid:0)(Am)m∈Z≥0, (ϕn,m)0≤m≤n(cid:1) as in Definition 7.1), which satisfies the following additional conditions: (1) For every m ∈ Z≥0, the algebra Am is a spatial semisimple finite dimen- sional Lp operator algebra (Definition 8.8). (2) For all m, n ∈ Z≥0 with m ≤ n, the map ϕn,m is a spatial homomorphism (Definition 8.13). We further say that a Banach algebra A is a spatial Lp AF algebra if it is isomet- rically isomorphic to the direct limit of a spatial Lp AF direct system. CLASSIFICATION OF Lp AF ALGEBRAS 31 Definition 9.2. Let p ∈ [1, ∞) \ {2}. Let(cid:0)(Am)m∈Z≥0, (ϕn,m)0≤m≤n(cid:1) be a spatial Lp AF direct system (Definition 9.1). We make A = lim −→ m Lp operator algebra via Lemma 8.12 and Theorem 7.4. (Am, ϕn,m) into a matricial The matrix norms on A a priori depend on how A is realized as a direct limit. We will show in Theorem 9.12 that in fact they are independent of the realization. Lemma 9.3. Let p ∈ [1, ∞)\{2}. Let r ∈ Z>0 and let(cid:0)(Am)m∈Z≥0, (ϕn,m)0≤m≤n(cid:1) be a spatial Lp AF direct system. Then is a spatial Lp AF direct system. (cid:0)(Mr(Am))m∈Z≥0 , (idMr ⊗ ϕn,m)0≤m≤n(cid:1) d ) is isometrically isomorphic to M p Proof. Using Definition 3.15 and Definition 3.13, for any d ∈ Z>0 we see that Mr(M p rd, via a map as in Definition 3.12. There- fore Mr(Am) is a spatial semisimple finite dimensional Lp operator algebra for all m ∈ Z>0. Lemma 8.20 implies that ϕn,m is isometrically similar to a block diag- onal homomorphism. It follows from Lemma 8.19(7) and Proposition 8.2(3) that the maps idMr ⊗ ϕn,m are isometrically similar to block diagonal homomorphisms. Now use Lemma 8.20. (cid:3) Corollary 9.4. Let p ∈ [1, ∞) \ {2}. Let (cid:0)(Am)m∈Z≥0 , (ϕn,m)0≤m≤n(cid:1) be a spatial Lp AF direct system. Let A = lim −→ m (Am, ϕn,m) be the direct limit, equipped with the matricial Lp operator algebra structure of Definition 9.2. Let r ∈ Z>0. Then Mr(A) is a spatial Lp AF algebra. Proof. This is immediate from Lemma 9.3. (cid:3) Lemma 9.5. Let p ∈ [1, ∞) \ {2}. Let N ∈ Z>0 and for k = 1, 2, . . . , N let m )m∈Z≥0, (ϕ(k) (cid:0)(A(k) Then is a spatial Lp AF direct system. Proof. Obviously A(k) ϕ(k) m !m∈Z≥0 n,m)0≤m≤n(cid:1) be a spatial Lp AF direct system (Definition 9.1).   N Mk=1 Mk=1 m is a spatial semisimple finite dimensional Lp operator n,m!0≤m≤n , N Mk=1 N A(k)   algebra for all m ∈ Z>0. By Lemma 8.20, a direct system of spatial semisimple finite dimensional Lp operator algebras is a spatial Lp AF direct system if and only if its maps are all isometrically similar to block diagonal maps. It follows from Lemma 8.19(6) that the direct sum of maps isometrically similar to block diagonal maps is again isometrically similar to a block diagonal map. (cid:3) Corollary 9.6. Let p ∈ [1, ∞) \ {2}. Then the direct sum of finitely many spatial Lp AF algebras is again a spatial Lp AF algebra. Proof. This is immediate from Lemma 9.5. (cid:3) Definition 9.7. Let A be a Banach algebra, and let e = (en)n∈Z>0 be a sequence of idempotents in A which is nondecreasing, that is, for n ∈ Z>0 we have en ≤ en+1 in the sense of Definition 6.1. Set e0 = 0 (by convention), and let θe : Cc(Z>0) → A 32 N. CHRISTOPHER PHILLIPS AND MARIA GRAZIA VIOLA be the unique homomorphism such that θe(χ{n}) = en − en−1 for all n ∈ Z>0. We equip Cc(Z>0) with the norm k · k∞, and when we refer to kθek, or demand that θe be contractive or bounded, we use this norm. Proposition 9.8. Let p ∈ [1, ∞) \ {2}. Let A be a separable Lp operator algebra, and let e = (en)n∈Z>0 and θe be as in Definition 9.7. Assume that this sequence is an approximate identity for A, and that θe is contractive. Then there are a σ-finite measure space (Y, C, ν), with Lp(Y, ν) separable, and an isometric nondegenerate representation π : A → L(Lp(Y, ν)), such that π(en) is a hermitian idempotent in L(Lp(Y, ν)) for all n ∈ Z>0. Proof. Use Proposition 2.6 to find a σ-finite measure space (X, B, µ) such that Lp(X, µ) is separable and an isometric representation ρ of A on Lp(X, µ). It is clear from contractivity of θe and Lemma 6.6 that if n ∈ Z>0 then kenk = 1 and (taking e0 = 0) that en−1 is a hermitian idempotent in enAen. Apply Lemma 6.17 to the idempotents ρ(en) for n ∈ Z>0. In the rest of the proof, we use the notation of Lemma 6.17. Set E = eLp(X, µ). Then s is an invertible isometry from Lp(Y, ν) to E. Moreover, the map a 7→ ρ(a)E defines a homomorphism from A to L(E), with kρ(a)Ek = kρ(a)k by Lemma 2.3. Now the representation π : A → L(Lp(Y, ν)), defined by π(a) = s−1[ρ(a)E]s, is nondegen- erate and isometric. Moreover, for n ∈ Z>0, the operator π(en) is multiplication n by the characteristic function of L(Lp(Y, ν)). [k=1 Yn and is hence a hermitian idempotent in (cid:3) Under the hypotheses of Proposition 9.8, an Lp operator algebra has a canonical norm on its unitization. Proposition 9.9. Let p ∈ [1, ∞)\ {2}. Let A be a separable nonunital Lp operator algebra, and let e = (en)n∈Z>0 be as in Definition 9.7. Assume that this sequence is an approximate identity for A, and that the homomorphism θe of Definition 9.7 is contractive. Then there is a unique norm k · k on the unitization A+ of A satisfying the following conditions: (1) k · k agrees with the given norm on A ⊂ A+. (2) k · k is equivalent to the usual norm on the unitization. (3) A+ is an Lp operator algebra. (4) Identify C(Z>0 ∪ {∞}) with C0(Z>0)+, and give it the usual supremum e : C(Z>0 ∪ {∞}) → A+ be the unitization norm on C(Z>0 ∪ {∞}). Let θ+ of θe. Then θ+ e is contractive. Proof. We first prove existence. Let π : A → L(Lp(Y, ν)) be as in Proposition 9.8. Extend this homomorphism to a homomorphism π+ : A+ → L(Lp(Y, ν)). Then π+ is injective because A is not unital. Define kak = kπ+(a)k for a ∈ A+. Conditions (1), (2), and (3) are immediate. It remains to prove condition (4). By density, it suffices to prove that for n ∈ Z>0, if we set K = {n + 1, n + 2, . . . , ∞}, and take any function f ∈ C(Z>0 ∪ {∞}) vanishing on K and any λ ∈ C, then (9.1) k(π+ ◦ θ+ e )(f + λχK )k ≤ max(kf k, λ). [n=1 CLASSIFICATION OF Lp AF ALGEBRAS 33 Since Proposition 9.8 implies that π(en) is a hermitian idempotent, by Lemma 6.9 there is a measurable subset E ⊂ Y such that π(en) is multiplication by χE on Lp(Y, ν)). Then (π+ ◦ θ+ e )(f ) acts on Lp(E, ν) and is zero on Lp(Y \ E, ν), while (π+ ◦ θ+ e )(λχK ) is multiplication by λ on Lp(Y \ E, ν) and zero on Lp(E, ν). So (9.1) holds. Now we prove uniqueness. Let k · k be a norm as in the statement. For n ∈ Z>0, a ∈ enAen, and λ ∈ C we prove that (9.2) ka + λ · 1k = max(ka + λenk, λ). Since the right hand side of (9.2) depends only on the norm on A, and since ∞ enAen is dense in A, uniqueness will follow. It follows from (4) that en is a hermitian idempotent in A+. Also, en commutes with a + λ · 1 and k1 − enk = kθ+ e (χK)k ≤ 1. So Lemma 6.15 implies that ka+λ·1k = max(cid:0)ken(a+λ·1)enk, k(1−en)(a+λ·1)(1−en)k(cid:1) = max(ka+λenk, λ), which is (9.2). (cid:3) Proposition 9.10. Let p ∈ [1, ∞) \ {2}, and let A = lim −→ m (Am, ϕn,m) be a spatial Lp AF algebra, expressed as a direct limit as in Definition 9.1, and with canonical maps ϕn : An → A for n ∈ Z>0. Then A is a separable nondegenerately repre- sentable Lp operator algebra. Moreover, e = (ϕn(1An))n∈Z>0 is a nondecreasing approximate identity of idempotents such that the corresponding homomorphism θe of Definition 9.7 is contractive. Proof. Theorem 7.3 and Lemma 8.12 imply that A is an Lp operator algebra. Sep- arability is obvious. We prove the statement about the approximate identity. By (Am, ϕn,m) is nondegenerately repre- Proposition 9.8, this will imply that A = lim −→ m sentable. For n ∈ Z>0, write fn for the identity of An, and set en = ϕn(fn). It is clear that kenk ≤ 1 (with equality unless An = 0), and that en ≤ en+1. Set e = (en)n∈Z>0, as in the statement of the theorem. Then we have lim n→∞ enϕm(a) = lim n→∞ ϕm(a)en = ϕm(a) for every m ∈ Z>0 and a ∈ Am. Since [m∈Z>0 for all n ∈ Z>0, a standard ε for A. ϕm(Am) is dense in A and kenk ≤ 1 3 -argument shows that e is an approximate identity It remains to prove that θe is contractive. We prove by induction on n that, with ϕn,0(f0) taken to be zero, the idempotents ϕn,j(fj)−ϕn,j−1(fj−1) are hermitian for j = 1, 2, . . . , n. For n = 1, this is just the assertion that the identity is a hermitian idempotent in A1. If the statement is known for n, then for j = 1, 2, . . . , n we have ϕn+1, j(fj) − ϕn+1, j−1(fj−1) = ϕn+1, n(cid:0)ϕn,j(fj) − ϕn,j−1(fj−1)(cid:1), which is a hermitian idempotent by Lemma 6.13. Also, ϕn+1, n+1(fn+1) − ϕn+1, n(fn) = 1An+1 − ϕn+1, n(fn) is hermitian because ϕn+1, n(fn) is hermitian. This completes the induction. 34 N. CHRISTOPHER PHILLIPS AND MARIA GRAZIA VIOLA Corollary 6.14(4) now implies that θeC({1,2,...,n}) is contractive for all n ∈ Z>0. (cid:3) It follows that θe is contractive. Proposition 9.11. Let p ∈ [1, ∞) \ {2}, and let A be a spatial Lp AF algebra. Then A is isometrically isomorphic to the direct limit of a spatial Lp AF direct system in which all the connecting maps are injective. Proof. Let(cid:0)(Am)m∈Z≥0, (ϕn,m)m≤n(cid:1) be a spatial Lp AF direct system such that A Mj=1 is isometrically isomorphic to lim −→ m Am. Then for m ∈ Z>0 we can write Am = N (m) M p d(m,j) with N (m) ∈ Z≥0 and d(m, 1), d(m, 2), . . . , d(m, N (m)) ∈ Z>0. ∞ Set Jm = Ker(ϕn,m). Then Jm is a closed ideal in Am, and Am/Jm is a [n=m+1 spatial semisimple finite dimensional Lp operator algebra by Lemma 8.10. Let m ∈ Z≥0. For n ∈ Z≥0 with n ≥ m, Corollary 8.15 shows that the induced homomorphism Am → An/Jn is spatial. Lemma 8.14 can then be used to show that the induced homomorphism ϕn,m : Am/Jm → An/Jn is spatial. Clearly ϕn,m is injective. Now (cid:0)(Am/Jm)m∈Z≥0, (ϕn,m)0≤m≤n(cid:1) is a spatial Lp AF direct system whose direct limit is isometrically isomorphic to A. Corollary 9.12. Let p ∈ [1, ∞) \ {2}, and let A be a spatial Lp AF algebra. Then A has unique Lp operator matrix norms. (cid:3) Proof. By Proposition 9.11, there is a spatial Lp AF direct system(cid:0)(Am)m∈Z≥0, (ϕn,m)0≤m≤n(cid:1) with injective maps such that A is isometrically isomorphic to lim −→ m Am. For m, n ∈ Z≥0 with m ≤ n, Lemma 8.20 implies that ϕn,m is isometrically similar to a block di- agonal homomorphism. An injective block diagonal homomorphism is isometric, so ϕn,m is isometric. The result now follows from Lemma 8.12 and Proposition 7.5. (cid:3) 10. Classification of spatial Lp AF algebras In this section we prove our main result, the classification of spatial Lp AF al- gebras based on their scaled preordered K0 groups. Moreover, as for AF algebras, we show that every countable scaled Riesz group can be realized as the scaled pre- ordered K0 group of a spatial Lp AF algebra. Given the theory already developed, the proofs are now essentially the same as in the C* algebra case. The original C* theory of AF algebras is mainly due to Bratteli [3], Elliott [7], and Effros, Handelman, and Shen [6]. As references for the entire theory, we rely on Chapter 3 of [1] and on [5]. For a more detailed discussion of Riesz groups than is needed here, see [11]. We only state the classification theorem in terms of scaled preordered K-theory. We don't discuss the connection with Bratteli diagrams, since the relation between Riesz groups and Bratteli diagrams is well known and the generalization of the AF algebra classification to spatial Lp AF algebras introduces nothing new here. We begin by describing the relevant K-theoretic background. Definition 10.1. A preordered abelian group is a pair (G, G+) in which G is an abelian group and G+ is a subset of G such that 0 ∈ G+ and G+ + G+ ⊂ G+. For η, µ ∈ G we write η ≤ µ to mean that µ − η ∈ G+. CLASSIFICATION OF Lp AF ALGEBRAS 35 A scaled preordered abelian group is a triple (G, G+, Σ) such that (G, G+) is a preordered abelian group, and Σ (the scale) is a subset of G+ such that 0 ∈ Σ. If (H, H+) is another preordered abelian group, and f : G → H is a homomor- phism, then f is positive if f (G+) ⊂ H+. If Σ ⊂ G+ and Γ ⊂ H+ are scales, we say that f is contractive if f (Σ) ⊂ Γ. The definitions are weak because they are supposed to accommodate K0(A) for any Banach algebra A. For example (using the notation of Definition 10.3 below), for the algebras C0(R2), O∞, and On we get the following results for (cid:0)K0(A), K0(A)+, Σ(A)(cid:1): (Z, {0}, {0}), (Z, Z, Z), and (Z/(n − 1)Z, Z/(n − 1)Z, Z/(n − 1)Z). The scale need not be hereditary if A does not have cancellation. We will take the K0-group of a Banach algebra to be as in Section 5 of [1]. (We will make very little use of the K1-group, and we don't recall its definition.) To start, we recall one of the standard equivalence relations on idempotents. It is called algebraic equivalence in Definition 4.2.1 of [1]. Definition 10.2. Let A be a Banach algebra. Let e and f be idempotents in A. We say that e is algebraically Murray-von Neumann equivalent to f , denoted by e ∼ f , if there exist x, y ∈ A such that xy = e and yx = f . Definition 10.3. Let A be a ring. We define M∞(A) to be the (algebraic) di- rect limit of the matrix rings Mn(A) under the embeddings a 7→ diag(a, 0). (See Definition 5.1.1 of [1].) We define V (A) to be the abelian semigroup of algebraic Murray-von Neumann equivalence classes of idempotents in M∞(A). (See Defini- tion 5.1.2 of [1] and the discussion afterwards.) When A is a Banach algebra, we define (cid:0)K0(A), K0(A)+, Σ(A)(cid:1) as follows. We take K0(A) to be the usual K0-group of A, as in, for example, Definition 5.5.1 of [1]. (There is trouble if one uses the definition there for more general rings.) For n ∈ Z>0 and an idempotent e ∈ Mn(A), we write [e] for its class in K0(A). We take K0(A)+ to be the image of V (A) in K0(A) under the map coming from 5.5.2 and Definition 5.3.1 of [1]. We take Σ(A) to be the image under this map of the subset of V (A) consisting of the classes of idempotents in A ⊂ M∞(A). We warn that [e] is sometimes used for the class of e in V (A). Since V (A) → K0(A) need not be injective, this is not the same as the class of e in K0(A). Remark 10.4. We can rewrite the definitions of K0(A)+ and Σ(A) as and K0(A)+ =(cid:8)[e] : e is an idempotent in M∞(A)(cid:9). Σ(A) =(cid:8)[e] : e is an idempotent in A(cid:9). Proposition 10.5. Let A be a Banach algebra. Then (cid:0)K0(A), K0(A)+, Σ(A)(cid:1) is a scaled preordered abelian group in the sense of Definition 10.1. Proof. This is immediate. (cid:3) Direct limits of direct systems of scaled preordered abelian groups are con- structed in the obvious way. 36 N. CHRISTOPHER PHILLIPS AND MARIA GRAZIA VIOLA Lemma 10.6. Let I be a directed set. For every i ∈ I let (Gi, (Gi)+, Σi)i∈I be a scaled preordered abelian group, and for i, j ∈ I with i ≤ j let gj,i : Gi → Gj be a positive contractive homomorphism. Let G be the direct limit lim Gi as abelian −→ groups and for i ∈ I let gi : Gi → G be the canonical map. Set G+ = [i∈I gi((Gi)+) and Σ = [i∈I i gi(Σi). Then (G, G+, Σ) is a scaled preordered abelian group and (G, G+, Σ) is the direct limit of (Gi, (Gi)+, Σi)i∈I in the category of scaled preordered abelian groups and positive contractive homomorphisms. Proof. Without the scales, see Proposition 1.15 in [11]. The additional work for scaled preordered abelian groups is easy, and is omitted. (cid:3) Theorem 10.7. The assignment A 7→ (cid:0)K0(A), K0(A)+, Σ(A)(cid:1) is a functor from Banach algebras and homomorphisms to scaled preordered abelian groups and con- tractive positive homomorphisms which commutes with direct limits in which the maps are contractive. Proof. Functoriality of K0(A) is stated after Definition 5.5.1 of [1]. Functoriality of the other two parts is clear. (Also see 5.2.1 of [1].) The fact that K0(A) commutes with direct limits is Theorem 6.4 of [19]. The statement for K0(A)+ follows from that for V (A), which is 5.2.4 of [1]. The state- ment for Σ(A) follows by the same proof, which is Propositions 4.5.1 and 4.5.2 of [1]. (cid:3) For the part about the scale in the following definition, we refer the reader to the beginning of Chapter 7 of [5]. Riesz groups are sometimes called dimension groups, for example in the definition at the beginning of Chapter 3 of [11]. Definition 10.8. Let (G, G+) be a preordered abelian group. We say that G is an unperforated ordered group if: (1) G+ − G+ = G. (2) G+ ∩ (−G+) = {0}. (3) Whenever η ∈ G and n ∈ Z>0 satisfy nη ∈ G+, then η ∈ G+. We say that (G, G+) is a Riesz group if, in addition: (4) Whenever η1, η2, µ1, µ2 ∈ G satisfy ηj ≤ µk for j, k ∈ {1, 2}, then there exists λ ∈ G such that ηj ≤ λ ≤ µk for j, k ∈ {1, 2}. Let (G, G+, Σ) be a scaled preordered abelian group. We say that G is a scaled Riesz group if (G, G+) is a Riesz group, and in addition: (5) For every η ∈ G+ there are n ∈ Z>0 and µ1, µ2, . . . , µn ∈ Σ such that η = µ1 + µ2 + · · · + µn. (6) Whenever η, µ ∈ G satisfy 0 ≤ η ≤ µ and µ ∈ Σ, then η ∈ Σ. (7) For all η, µ ∈ Σ there is λ ∈ Σ such that η ≤ λ and µ ≤ λ. We recall for reference some standard definitions and facts. A few are restated for the Lp case. Definition 10.9. For N ∈ Z>0 we make ZN a Riesz group by taking (ZN )+ =(cid:8)(η1, η2, . . . , ηN ) ∈ ZN : ηk ≥ 0 for k = 1, 2, . . . , N(cid:9). CLASSIFICATION OF Lp AF ALGEBRAS 37 Remark 10.10. The possible scales on ZN are exactly the following sets. Take d = (d1, d2, . . . , dN ) ∈ ZN with dk > 0 for k = 1, 2, . . . , N , and define [0, d] =(cid:8)(µ1, µ2, . . . , µN ) ∈ (ZN )+ : µk ≤ dk for k = 1, 2, . . . , N(cid:9). See page 43 of [5]. M Remark 10.11. Let p ∈ [1, ∞). Let A = M p cj be a spatial semisimple finite dimensional Lp operator algebra. Set c = (c1, c2, . . . , cM ). As in the C* algebra case (see pages 55 -- 56 of [5]), using the notation from Definition 10.9 and Remark 10.10, we have (cid:0)K0(A), K0(A)+, Σ(A)(cid:1) ∼=(cid:0)ZM , (ZM )+, [0, c](cid:1). The map K0(A) → ZM sends the class of an idempotent (e1, e2, . . . , eM ) ∈ Mj=1 Mn(M p cj ) M Mj=1 to If B = N Mj=1 (cid:0)rank(e1), rank(e2), . . . , rank(eN )(cid:1). is another spatial semisimple finite dimensional Lp operator algebra, M p dj and ϕ : A → B is a homomorphism, then ϕ∗ : ZM → ZN is given by the partial multiplicity matrix m(ϕ) of Definition 8.17(2). M N Mj=1 Mk=1 Lemma 10.12. Let A = M p cj and B = M p dk be spatial semisimple finite dimensional Lp operator algebras. Let f a positive contractive homomorphism from (cid:0)K0(A), K0(A)+, Σ(A)(cid:1) to (cid:0)K0(B), K0(B)+, Σ(B)(cid:1). Then there exists a spatial homomorphism ϕ : A → B such that ϕ∗ = f . Moreover, ϕ is unique up to isometric similarity, and it can be chosen to be block diagonal. Proof. The homomorphism f from to (cid:0)K0(A), K0(A)+, Σ(A)(cid:1) ∼=(cid:0)ZM , ZM (cid:0)K0(B), K0(B)+, Σ(B)(cid:1) ∼=(cid:0)ZN , ZN + , [0, [1A]](cid:1) + , [0, [1B]](cid:1) is given by an N × M matrix m = (mk,j)1≤k≤N, 1≤j≤M with entries in Z. (See Remark 10.11). One checks that positivity implies that the entries are in Z≥0 M Xj=1 and that contractivity implies that mk,jcj ≤ dk for k = 1, 2, . . . , N . Lemma 8.19(4) implies that there is a block diagonal homomorphism ϕ : A → B such that m(ϕ) = m, and it is clear that ϕ∗ = f (Remark 10.11). It follows from Lemma 8.19(2) that ϕ is spatial. Now suppose ψ : A → B is another spatial homomorphism such that ψ∗ = f . Lemma 8.20 implies that ψ is isometrically similar to a block diagonal homomor- phism. Therefore we may assume that both ϕ and ψ are block diagonal homomor- phisms. It is easy to check that if n ∈ Z>0 then two block diagonal homomorphisms from A to Mn with the same partial multiplicities are similar via a permutation matrix, and are thus isometrically similar. It is now immediate that m(ϕ) = m(ψ) implies that ϕ and ψ are isometrically similar. (cid:3) 38 N. CHRISTOPHER PHILLIPS AND MARIA GRAZIA VIOLA Theorem 10.13. Let (G, G+, Σ) be a countable scaled preordered abelian group. Then (G, G+, Σ) is a scaled Riesz group if and only if for n ∈ Z≥0 there are scaled Riesz groups (Gn, (Gn)+, Σn), each isomorphic to a group as in Remark 10.10, and positive contractive homomorphisms fn+1, n : Gn → Gn+1, such that, if for m, n ∈ Z≥0 with m ≤ n we set fn,m = fn, n−1 ◦ fn−1, n−2 ◦ · · · ◦ fm+1, m, then (G, G+, Σ) ∼= lim −→(cid:0)(Gn, (Gn)+, Σn)n∈Z≥0, (fn,m)0≤m≤n(cid:1). Proof. The statement for a countable preordered abelian group (G, G+) is Theo- rem 2.2 in [6]. Using Lemma 7.1 of [5] one shows that if (G, G+, Σ) is a scaled Riesz group then the homomorphisms in the commutative diagram in Lemma 2.1 of [6] can be chosen to be positive and contractive. With this choice of homomorphisms in Theorem 2.2 of [6], one obtains the result for a scaled Riesz group. (cid:3) Corollary 10.14. Let p ∈ [1, ∞)\ {2}, and let A be a spatial Lp AF algebra. Then (cid:0)K0(A), K0(A)+, Σ(A)(cid:1) is a scaled Riesz group. Proof. Use Theorem 10.7, Remark 10.11, and Theorem 10.13. (cid:3) For completeness, we also state the result for K1. Proposition 10.15. Let p ∈ [1, ∞) \ {2}, and let A be a spatial Lp AF algebra. Then K1(A) = 0. Proof. By Definition 9.1, there is a spatial Lp AF direct system(cid:0)(Am)m∈Z≥0, (ϕn,m)0≤m≤n(cid:1) N (m) N (m) Am. For m ∈ Z≥0, write Am = such that A ∼= lim −→ m and since K1(M p n) = 0 for every n ∈ Z>0 by Example 8.1.2(a) in [1], we obtain K1(Am) = 0. Since K1 commutes with Banach algebra direct limits (Remark 8.1.5 in [1]), the conclusion follows. (cid:3) cm,j . Since K1(Am) = Mj=1 M p K1(M p cj ), Mj=1 Lemma 10.16 (Proposition 5.5.5 of [1]). Let A be a Banach algebra which has an approximate identity consisting of idempotents. Then K0(A) is naturally isomor- phic to the Grothendieck group of V (A). We can now give the main classification results. Theorem 10.17. Let p ∈ [1, ∞). Let (G, G+, Σ) be a countable scaled Riesz group. Then there exists a spatial Lp AF algebra A such that (cid:0)K0(A), K0(A)+, Σ(A)(cid:1) ∼= (G, G+, Σ). Proof. Choose a direct system as in Theorem 10.13. For n ∈ Z≥0, Remark 10.11 shows that there is a spatial semisimple finite dimensional Lp operator algebra An such that (cid:0)K0(An), K0(An)+, Σ(An)(cid:1) ∼= (Gn, (Gn)+, Σn). Lemma 10.12 provides a block diagonal homomorphism ϕn+1, n : An → An+1 such that (ϕn+1, n)∗ = fn+1, n. For m, n ∈ Z≥0 with m ≤ n set ϕn,m = ϕn, n−1 ◦ ϕn−1, n−2 ◦ · · · ◦ ϕm+1, m : Am → An. Then ϕn,m is spatial by Corollary 8.15 and Lemma 8.19(2). CLASSIFICATION OF Lp AF ALGEBRAS 39 Define A = lim and −→(cid:0)(An)n∈Z≥0 , (ϕn,m)0≤m≤n(cid:1). Then A is a spatial Lp AF algebra, by Theorem 10.7. (cid:0)K0(A), K0(A)+, Σ(A)(cid:1) ∼= (G, G+, Σ) (cid:3) The proof of Elliott's Theorem has two steps, the first of which is entirely about the category of scaled Riesz groups and positive contractive homomorphisms (and which is exactly the same in every category of algebras), and the second of which transfers the result to algebras in the appropriate category. The first step is the following lemma. Even though it is a key step in the proof, and the proof appears in a number of books, we haven't found an explicit statement of this result in the literature. Lemma 10.18. For every m ∈ Z≥0 let (Gm, (Gm)+, Σm) and (Hm, (Hm)+, Tm) be scaled Riesz groups, each isomorphic to a group as in Remark 10.10, and for m, n ∈ Z≥0 with m ≤ n let gn,m : Gm → Gn and hn,m : Hm → Hn be positive contractive homomorphisms satisfying gn,m ◦ gm,k = gn,k and hn,m ◦ hm,k = hn,k whenever 0 ≤ k ≤ m ≤ n. Set (G, G+, Σ) = lim −→ n (Gn, (Gn)+, Σn) and (H, H+, T ) = lim −→ n (Hn, (Hn)+, Tn), and let f : G → H be an isomorphism of scaled ordered groups. Then there exist m0, m1, . . . , n0, n1, . . . ∈ Z≥0 such that m0 < m1 < · · · and n0 < n1 < · · · , and positive contractive homomorphisms rk : Gmk → Hnk and sk : Hnk → Gmk+1 for k ∈ Z≥0, such that the following diagram commutes: Gm0 r0 Hn0 gm1 ,m0 / Gm1 <③ gm2 ,m1 / Gm2 <③ gm3 ,m2 / Gm3 <③ ③ ③ ③ s0 ③ ③ ③ ③ ③ ③ ③ r1 Hn1 ③ ③ ③ ③ ③ ③ s1 ③ ③ ③ ③ ③ ③ s2 ③ ③ ③ ③ r2 Hn2 ③ ③ ③ ③ hn1 ,n0 hn2 ,n1 hn3 ,n2 / · · · · · · <① ① ① gm4 ,m3 / s3 ① ① ① ① ① · · · · · · hn4,n3 r3 Hn3 ① ① ① ① / G f / H, and such that f is the direct limit of the maps rk for k ∈ Z≥0 and f −1 is the direct limit of the maps sk for k ∈ Z≥0. Proof. This result is contained, in slightly different language, in the proof of The- orem 7.3.2 of [1] (starting with the second diagram there). (cid:3) We also want the one sided version of Lemma 10.18. Lemma 10.19. Let (G, G+, Σ) and (H, H+, T ) be as in Lemma 10.18, and let f : G → H be a positive contractive homomorphism. Then there exist m0, m1, . . ., n0, n1, . . . ∈ Z≥0 such that m0 < m1 < · · · and n0 < n1 < · · · , and positive contractive homomorphisms rk : Gmk → Hnk for k ∈ Z≥0, such that the following diagram commutes: gm1 ,m0 / gm3 ,m2 / gm2 ,m1 / gm4 ,m3 / / · · · · · · / G Gm0 Gm1 Gm2 Gm3 r0 r1 r2 r3 Hn0 hn1 ,n0 / Hn1 / Hn2 hn2 ,n1 hn3 ,n2 / Hn3 hn4,n3 / · · · · · · f / H, /   /   /     /   / / < / / < / / < / / < / /   /   /     /   / / / / / 40 N. CHRISTOPHER PHILLIPS AND MARIA GRAZIA VIOLA and such that f is the direct limit of the maps rk for k ∈ Z≥0. Proof. The proof is very similar to, but slightly simpler than, that of Lemma 10.18. (cid:3) Theorem 10.20. Let p ∈ [1, ∞). Let A and B be spatial Lp AF algebras, and let f : K0(A) → K0(B) define an isomorphism from (cid:0)K0(A), K0(A)+, Σ(A)(cid:1) to (cid:0)K0(B), K0(B)+, Σ(B)(cid:1). Then there is a completely isometric isomorphism ϕ : A → B such that ϕ∗ = f . Proof. By definition, we can write A and B as direct limits of spatial Lp AF direct systems, A = lim −→ m (cid:0)(Am)m∈Z≥0, (αn,m)0≤m≤n(cid:1) and B = lim m (cid:0)(Bm)m∈Z≥0, (βn,m)0≤m≤n(cid:1). For m ∈ Z≥0 let αm : Am → A and βm : Bm → B be the canonical maps. Apply Lemma 10.18 with −→ (Gm, (Gm)+, Σm) =(cid:0)K0(Am), K0(Am)+, Σ(Am)(cid:1) (Hm, (Hm)+, Tm) =(cid:0)K0(Bm), K0(Bm)+, Σ(Bm)(cid:1) for m ∈ Z≥0 (see Remark 10.11), with gm,n = (αn,m)∗ and hm,n = (βn,m)∗ when- ever m, n ∈ Z≥0 with n ≥ m, and with and and (G, G+, Σ) =(cid:0)K0(A), K0(A)+, Σ(A)(cid:1) (H, H+, T ) =(cid:0)K0(B), K0(B)+, Σ(B)(cid:1) (justified by Theorem 10.7). Let m0 < m1 < · · · , n0 < n1 < · · · , rk : Gmk → Hnk , and sk : Hnk → Gmk+1 be as in Lemma 10.18, making the diagram there commute. We construct by induction on k spatial homomorphisms ϕk : Amk → Bnk and ψk : Bnk → Amk+1 such that (ϕk)∗ = rk and (ψk)∗ = sk for k ∈ Z≥0, and such that the diagram αm1 ,m0 / Am1 <③ αm2 ,m1 / Am2 <③ αm3 ,m2 / Am3 <③ ③ ③ ψ0 ③ ③ ③ ③ ③ ③ ③ ③ βn1 ,n0 ϕ1 Bn1 ③ ③ ③ ③ ③ ψ1 ③ ③ ③ ③ ③ βn2,n1 ϕ2 Bn2 ③ ③ ψ2 ③ ③ ③ ③ ③ ③ ③ ③ βn3 ,n2 ϕ3 Bn3 ① ① ① ① / · · · · · · <① ① αm4 ,m3 / ① ① ψ3 ① ① ① ① βn4 ,n3 / · · · · · · / A / B (10.1) Am0 ϕ0 Bn0 commutes. For the initial step, use the existence statement in Lemma 10.12 to choose a spatial homomorphism ϕ0 : Am0 → Bn0 such that (ϕ0)∗ = r0. Use the existence statement in Lemma 10.12 to choose a spatial homomorphism ψ(0) : Bn0 → Am1 such that (cid:0)ψ(0) Lemma 10.12 to choose an invertible isometry t ∈ Am1 such that 0 (cid:1)∗ = s0, and use Corollary 8.15 and the uniqueness statement in 0 for all a ∈ Am0 . Define ψ0 by ψ0(b) = tψ(0) 0 (b)t−1 for b ∈ Bn0. For the induction step, suppose we have ϕk and ψk. Use the existence statement in Lemma 10.12 to t(cid:0)ψ(0) 0 ◦ ϕ0(cid:1)(a)t−1 = αm1,m0(a) /   /   /     / / / < / / < / / < / < / CLASSIFICATION OF Lp AF ALGEBRAS 41 choose a spatial homomorphism ϕ(0) Use the uniqueness statement in Lemma 10.12 to choose an invertible isometry v ∈ Bnk+1 such that k+1 : Amk+1 → Bnk+1 such that (cid:0)ϕ(0) k+1(cid:1)∗ = rk+1. v(cid:0)ϕ(0) k+1 ◦ ψk(cid:1)(b)v−1 = βnk+1,nk (b) for all b ∈ Bnk , and define ϕk+1 by ϕk+1(a) = vϕ(0) k+1(a)v−1 for a ∈ Amk+1 . The construction of ψk+1 is now the same as the construction of ψ0 in the initial step. For k ∈ Z≥0, the map ϕk is completely contractive by Lemma 8.20, Lemma 8.19(8), and Proposition 8.2(5). Commutativity of the diagram (10.1) therefore implies the existence of a contractive homomorphism ϕ : A → B such that ϕ ◦ αmk = βnk ◦ ϕk for all k ∈ Z≥0, and ϕ must in fact be completely contractive. Similarly, we get a completely contractive homomorphism ψ : B → A such that ψ ◦ βnk = αmk+1 ◦ ψk for all k ∈ Z≥0. Using the universal property of direct limits, we find that ϕ ◦ ψ = idB and ψ ◦ ϕ = idA. Therefore ϕ and ψ are completely isometric. It is clear that ϕ∗ = f . (cid:3) Theorem 10.21. Let p ∈ [1, ∞). Let A and B be spatial Lp AF algebras, and let f : K0(A) → K0(B) define a positive contractive homomorphism from (cid:0)K0(A), K0(A)+, Σ(A)(cid:1) to (cid:0)K0(B), K0(B)+, Σ(B)(cid:1). Then there is a completely contractive homomorphism ϕ : A → B such that ϕ∗ = f . Proof. The proof is a one sided version of the proof of Theorem 10.20, using Lemma 10.19 in place of Lemma 10.18. (cid:3) Theorem 10.22. Let p ∈ [1, ∞). Let A and B be spatial Lp AF algebras. Then the following are equivalent: (1) (cid:0)K0(A), K0(A)+, Σ(A)(cid:1) ∼=(cid:0)K0(B), K0(B)+, Σ(B)(cid:1). (2) A ∼= B as rings. (3) A is isomorphic to B (not necessarily isometrically) as Banach algebras. (4) A is isometrically isomorphic to B as Banach algebras. (5) A is completely isometrically isomorphic to B as matrix normed Banach algebras. Proof. It is trivial that (5) implies (4), that (4) implies (3), and that (3) implies (2). Since V (A) depends only on the ring structure of A, Lemma 10.16 and Proposi- tion 9.10 show that K0(A) depends only on the ring structure of A. It now fol- lows directly from the definitions that K0(A)+ and Σ(A) depend only on the ring structure of A. Thus (2) implies (1). The implication from (1) to (5) is Theo- rem 10.20. (cid:3) References [1] B. Blackadar, K-Theory for Operator Algebras, 2nd ed., MSRI Publication Series 5, Cam- bridge University Press, Cambridge, New York, Melbourne, 1998. [2] D. P. Blecher and C. Le Merdy, Operator Algebras and their Modules -- an Operator Space Approach, London Mathematical Society Monographs, New Series, no. 30, Oxford Science Publications, The Clarendon Press, Oxford University Press, Oxford, 2004. [3] O. Bratteli, Inductive limits of finite dimensional C*-algebras, Trans. Amer. Soc. 171 (1971), 195 -- 234. [4] J. B. Conway A Course in Functional Analysis, 2nd ed., Springer-Verlag Graduate Texts in Math. no. 96, Springer-Verlag, New York, Berlin, etc., 1990. [5] E. G. Effros, Dimensions and C*-Algebras, CBMS Regional Conf. Ser. in Math. No. 46, Amer. Math. Soc., Providence RI, 1981. 42 N. CHRISTOPHER PHILLIPS AND MARIA GRAZIA VIOLA [6] E. G. Effros, D. E. Handelman, and C. L. Shen, Dimension groups and their affine represen- tations, Amer. J. Math. 102 (1980), 385 -- 407. [7] G. A. Elliott, On the classification of inductive limits of sequences of semi-simple finite dimensional algebras, J. Algebra 38 (1976), 29 -- 44. [8] R. J. Fleming and J. E. Jamison, Isometries on Banach Spaces, Vol. 1: Function Spaces, Chapman & Hall/CRC Monographs and Surveys in Pure and Applied Mathematics, no. 129, Chapman & Hall/CRC, Boca Raton FL, 2003. [9] E. Gardella and M. Lupini, Representations of ´etale groupoids on Lp spaces, Adv. Math. 318 (2017), 233 -- 278.. [10] J. G. Glimm, On a certain class of operator algebras, Trans. Amer. Soc. 95 (1960), 318 -- 340. [11] K. R. Goodearl, Partially Ordered Abelian Groups with Interpolation, Math. Surveys and Monographs no. 20, Amer. Math. Soc., Providence RI, 1986. [12] S. Heinrich, Ultraproducts in Banach space theory, J. reine angew. Math. 313 (1980), 72 -- 104. [13] H. E. Lacey, The Isometric Theory of Classical Banach Spaces, Die Grundlehren der math- ematischen Wissenchaften no. 208, Springer-Verlag, Berlin, Heidelberg, New York, 1974. [14] J. Lamperti, On the isometries of certain function-spaces, Pacific J. Math. 8 (1958), 459 -- 466. [15] T. W. Palmer, Banach Algebras and the General Theory of *-Algebras. Vol. I. Algebras and Banach algebras, Encyclopedia of Mathematics and its Applications no. 49, Cambridge University Press, Cambridge, 1994. [16] N. C. Phillips, Analogs of Cuntz algebras on Lp spaces, preprint (arXiv:1201.4196v1 [math.FA]). [17] N. C. Phillips, Simplicity of UHF and Cuntz algebras on Lp spaces, preprint (arXiv:1309.0115.4196v1 [math.FA]). [18] N. C. Phillips, Isomorphism, nonisomorphism, and amenability of Lp UHF algebras, preprint (arXiv:1309.36941 [math.FA]). [19] N. C. Phillips, Crossed products of Lp operator algebras and the K-theory of Cuntz algebras on Lp spaces, preprint (arXiv:1309.6406 [math.FA]). [20] I. Vidav, Eine metrische Kennzeichnung der selbstadjungierten Operatoren, Math. Zeitschr. Bd. 66 (1956), 121 -- 128. Department of Mathematics, University of Oregon, Eugene OR 97403-1222, USA, and Department of Mathematics, University of Toronto, Room 6290, 40 St. George St., Toronto ON M5S 2E4, Canada. E-mail address: [email protected] Lakehead University Orillia, Orillia ON L3V 0B9, Canada, and Fields Institute, 222 College Street, Toronto, ON M5T 3J1, Canada E-mail address: [email protected]
1110.2691
3
1110
2011-11-07T20:25:42
An Analogue of Hin\u{c}in's Characterization of Infinite Divisibility for Operator-Valued Free Probability
[ "math.OA", "math.FA", "math.PR" ]
Let $B$ be a finite, separable von Neumann algebra. We prove that a $B$-valued distribution $\mu$ that is the weak limit of an infinitesimal array is infinitely divisible. The proof of this theorem utilizes the Steinitz lemma and may be adapted to provide a nonstandard proof of this type of theorem for various other probabilistic categories. We also develop weak topologies for this theory and prove the corresponding compactness and convergence results.
math.OA
math
AN ANALOGUE OF HIN CIN'S CHARACTERIZATION OF INFINITE DIVISIBILITY FOR OPERATOR-VALUED FREE PROBABILITY. JOHN D. WILLIAMS Abstract. Let B be a finite, separable von Neumann algebra. We prove that a B-valued distribution µ that is the weak limit of an infini- tesimal array is infinitely divisible. The proof of this theorem utilizes the Steinitz lemma and may be adapted to provide a nonstandard proof of this type of theorem for various other probabilistic categories. We also develop weak topologies for this theory and prove the corresponding compactness and convergence results. 1. Introduction. As a starting point, we recall a theorem, due to Hincin in classical proba- bility theory. Let M denote the collection of all probability measures on R. Given µ, ν ∈ M, we denote by µ ∗ ν the distribution of X + Y where X and Y are independent random variables with distribution µ and ν respectively. For r ∈ R we denote by δr ∈ M the Dirac mass at r. Theorem 1.1. Let {µij}i∈N,j=1,...,ni and µ be probility measures on R. As- sume that the following properties hold: (1) For every ǫ > 0, we have that limi↑∞ µij([−ǫ, ǫ]) → 1. (2) There exists a sequence {ri}i∈N ⊂ R such that δri ∗µi1 ∗· · · ∗µin1 → µ weakly. Then, for every N ∈ N, there exists a probability measure µ1/N such that µ = µ1/N ∗ · · · ∗ µ1/N , where the convolution on the right hand side is N -fold. The conclusion of the above Theorem is the assertion that µ is infinitely divisible with respect to the convolution operation. Free probability was developed in the 80's by Voiculescu as a method for encoding free product phenomenon in operator algebras in a probabalistic setting. In [13], Voiculescu introduced free indepence which is a noncommu- tative analogue of classical independence. The corresponding convolution operation, free additive convolution (in symbols, ⊞) was also introduced. In particular, if X and Y are freely independent random variables with respec- tive distributions µ and ν, then we denote by µ ⊞ ν the distribution of the random variable X + Y . In this noncommutative setting, analogues of Theorems 1.1 have been developed. Indeed, it was shown by Bercovici and Pata in [4] that the same 1 2 JOHN D. WILLIAMS result is true if classical independence is replaced by free independence and classical infinite divisibility is replace by ⊞-infinite divisibility. Belinschi and Bercovici proved in [3] that this characterization of infinite divisibility held for multiplicative free convolution, which arises when taking the product of free random variables. The primary focus of this paper will be an operator valued generalization of this theorem for additive convolution. In order to address amalgamated free product phenomenon in operator algebras with probabilistic methods, operator valued versions of free proba- bility theory were developed in [14], [16] and [17].The purpose of this paper is to prove Theorem 6.1, which is a version of Theorem 1.1 for operator valued free probability. A weaker version of this type of theorem was proven in [2] as a necessary intermediate lemma in their proof of an operator valued generalization of the Bercovici-Pata bijection. This paper is organized as follows. Section (2) contains preliminaries for B-valued distributions. Section (3) is devoted to the preliminaries of operator valued free probability. Section (4) contains preliminary results related to those distributions that arise from tracial von Neumann algebras. Section (5) is devoted to the Steinitz lemma which is the underlying tool in our approach to this theorem. Section (6) contains the main result and section (7) includes concluding remarks and acknowledgements. 2. B-Valued Distributions. Let B denote a unital C ∗-algebra and BhXi the space of noncommutative polynomials over B. We say that a map µ : BhXi → B is completely positive if for any finite set of elements P1(X), . . . , Pn(X) ∈ BhXi we have that the matrix µ(P ∗ i (X)Pj (X))n i,j=1 is a positive element of Mn(B). A map µ : BhXi → B is B-bimodular if µ(bP (X)b′) = bµ(P (X))b′ for all b, b′ ∈ B and P (X) ∈ BhXi. We denote by Σ the space of all B-bimodular, completely positive, B valued maps. Let µ ∈ Σ. We say that µ is exponentially bounded by M if for all elements b1, . . . , bn ∈ B, we have that kµ(Xb1X · · · XbnX)k ≤ M n+1kb1k · · · kbnk. We denote by Σ0 the set of all exponentially bounded elements in Σ and by Σ0,M the set of all such elements with a bound of M . The elements µ ∈ Σ which arise as the distribution of elements of in a C ∗-probability space are of primary interest in this paper. We defer the development of the theory of these distributions to the next section. We endow Σ with the topology of pointwise weak convergence (that is µλ → µ if an only if µλ(P (X)) → µ(P (X)) in the weak topology on B for all P (X) ∈ BhXi). See [9] for a study of norm topologies on Σ. Lemma 2.1. If we further assume that B is a W ∗-algebra, then the space Σ0,M is compact in the topology of pointwise weak convergence for all M > 0. Proof. We follow the proof of Alaoglu's theorem. We assume without loss of generality that M = 1 OPERATOR-VALUED INFINITE DIVISIBILITY. 3 Let {Pi}i∈I ⊂ BhXi be a family of monomials with B-bilinear span equal to BhXi that possess the property that each Pi may be written as a product Pi = Xb1Xb2 · · · XbnX with kbik ≤ 1 for i = 1, . . . , n. Let K = Qi∈I B1, where B1 denotes the unit ball in B, and endow this space K with the product of the weak toplogies on B. By Alaoglu and Tychonoff's theorems, K is compact. Let Φ : Σ0,1 → K by letting Φ(µ) = (µ(Pi))i∈I . Our exponential bound condition implies that the image is indeed in K. Since these elements Pi have dense linear span, this map is injective. Observe that if µλ → µ in Σ0,1 then for any σ ⊂ I finite, we have that µλ(Pi) → µ(Pi) uniformly over i ∈ σ. Thus, Φ is continuous. Similarly, if Φ(µλ) → Φ(µ), we have that µλ(Pi) → µ(Pi) weakly so that Φ−1 is continuous. Thus, Φ is a homeomorphism of Σ0,1 onto its image. Our lemma will follow when we show that Φ(Σ0,1) is a closed subset of K. Assume that Φ(µλ) → (bi)i∈I . We define a map µ : BhXi → B by letting µ(Pi) = bi and then extending this map by linearity. Clearly, µ is well defined. We claim that µ ∈ Σ0,1. To show this we must show that µ is B-bimodular, completely positive and has exponential bound equal to 1. First, observe that for φ ∈ B∗ and b, b′ ∈ B, we have φ(bµ(P (X))b′) − φ(µ(bP (X)b′)) = lim λ (φ(bµ(P (X))b′) − φ(bµλ(P (X))b′) + (φ(µλ(bP (X)b′)) − φ(µ(bP (X)b′)) The right hand side of the equation is 0 so that µ is B-bimodular. Next, note that for a finite collection {fj}k i fj)]k i,j=1 with Mk(B) endowed with the weak j=1 ⊂ BhXi, the matrix [µ(f ∗ i fj)]k i,j=1 is the weak limit of [µλ(f ∗ toplogy. As the positive cone is weakly closed, the fact that [µλ(f ∗ is positive implies the same for [µ(f ∗ i,j=1 i,j=1. Thus, µ is completely positive. Lastly, for any monomial P (X) = b1Xb2X · · · bnX in BhXi, we have that µλ(P (X)) ≤ kb1k · · · kbnk. Alaoglu's theorem implies that µ(P (X)) has the same bound. Thus, our lemma holds. (cid:3) i fj)]k i fj)]k 3. Operator Valued Free Probability. Let (A, φ, B) be a triple with B ⊂ A an inclusion of C ∗-algebras and φ : A → B a B-bimodular, completely positive map. We shall refer to such a triple as a B-valued probabliity space. Given an element a ∈ A we denote by µa : BhXi → B the B-valued distribution of a, defined by the equation µa(P (X)) = φ(P (a)) for all P (X) ∈ BhXi. We say that a family of subalgebras {Ai}i∈I are B-freely independent if φ(a1 · · · an) = 0 whenever aj ∈ Aij satisfies ij 6= ij+1 for all j = 1, . . . , n − 1 and φ(aℓ) = 0 for all ℓ = 1, . . . , n. Given elements ρ, ν ∈ Σ0 that arise as the B-valued distributions of elements x ∈ A and y ∈ A′, there exists a larger C ∗-algebra A ∗B A′ the contains copies of A and A′ as B-freely independent subalgebras (see [16] for an example of this construction). We shall denote 4 JOHN D. WILLIAMS by ρ ⊞ ν := µx+y as the additive B-free convolution of ρ and ν (here x + y denotes the sum of these elements in this larger algebra). We say that an element µ ∈ Σ is infinitely divisible with respect to B-valued additive free convolution if for every n ∈ N there exists an elements µ1/n ∈ Σ such that µ = µ1/n ⊞ · · · ⊞ µ1/n, where the convolution on the right hand side is n-fold (this may be defined more generally using free cumulants but this level of generality is enough for our present consideratinos, see [11]). Let M + n (B) = {x ∈ Mn(B) : ∃ǫ > 0 s.t. ℑx > ǫIn} where In denotes n (B). We shall n (B) and the the indentity element in Mn(B). Let H +(B) = F∞ refer to this sets as the noncommutative upper half-plane ( M − noncommutative lower half-plane are defined analagously). n=1 M + We shall define a transform the encodes a distribution as a function on H+(B). Indeed, given a self-adjoint a ∈ A with distribution µ, we define the function Gµ : H+(B) → H−(B) by defining, for each n, a function G(n) µ : Mn(B)+ → Mn(B)− where G(n) µ (b) = µ((X ⊗ In − b)−1) Since a is self adjoint, a ⊗ In − b is indeed invertible. Consider the series expansion G(n) µ (b) = b µ((b−1X ⊗ In)n) ∞ X n=0 and observe that for µ with an exponential bound of M , we have that this series is convergent for all b such that kb−1k < M . We shall refer to the function Gµas the Cauchy transfrom. As is well known, the distribution µ may be recovered from its Cauchy transform. We refer to [16] for the noncommutative function theory associated to B-valued free probability. µ = (G(n) We next define a map Fµ : H+(B) → H+(B) by letting F (n) µ )h−1i where the superscript denotes the multiplicative inverse of this element. For each distribution µ ∈ Σ0, there exists a set Γn, which is a neigh- n (B), where F (n) borhood of ∞ in Mn(B) intersected with M + is invert- ible. For each n we define the function ϕ(n) : Mn(B)+ → Mn(B)− by µ letting ϕ(n) µ )−1(b) − b (the superscript without the brackets refers to the inverse with respect to composition). We refer to the col- lection of all such maps over n as the Voiculescu transform of µ (in symbols, ϕµ : H +(B) → H −(B)). The fact that the image of this map lies in the lower half plane is a consequence of the fact that ℑF (n) µ (b) ≥ ℑb for all b ∈ M + n (B) (see [2] for proof of this fact). Given distributions µ and ν we have that µ (b) = (F (n) µ ϕ(n) µ⊞ν(b) = ϕ(n) µ (b) + ϕ(n) ν (b) for all b ∈ Mn(B)+ in the common domain of the two right hand functions. 1 Mk(B) → Mk(B) denote the cumulant functions. These functions are studied extensively in [11] and satisfy the following properties: Let κµ,n,k : Nn (1) κµ⊞ν,n,k(b1, . . . , bn) = κµ,n,k(b1, . . . , bn) + κν,n,k(b1, . . . , bn) OPERATOR-VALUED INFINITE DIVISIBILITY. 5 µ (b) = P∞ (2) ϕ(k) (3) κµ,n,k(b1, . . . , bn) ≤ M (4M )nkb1k · · · kbnk for µ ∈ Σ0,M n=1 κµ,n,k(b−1, . . . , b−1) Remark 3.1. Property (3) above implies that for µ ∈ Σ0,M , the power series in equation (2) is convergent for all b ∈ Mk(B)+ such that kb−1k < 1/4M . This, as well as the fact that G(k) µ (b) may be written as a convergent series for all b ∈ Mk(B)+ satisfying kb−1k < 1/M , will be exploited in the following lemma. Lemma 3.2. Let {µλ}λ∈Λ ∈ Σ0,M . The following are equivalent. µ (b) weakly for every b ∈ Mk(B)+ such that kb−1k < µ (b) weakly for every b ∈ Mk(B)+ such that kb−1k < µλ (b) → G(k) (1) µλ → µ in the pointwise weak topology. (2) G(k) 1/M . µλ (b) → F (k) (3) F (k) 1/M . µ and ϕ(k) (4) ϕ(k) such that kb−1k < 1/4M . Furthermore, ϕ(k) all b in this set. µλ have analytic extension to the set of all b ∈ Mk(B)+ µ (b) weakly for µλ (b) → ϕ(k) Proof. The equivalence of (2) and (3) is obvious since the map sending an operator to its multiplicative inverse is continuous in this topology. If we assume (1), then parts (2) and (4) follow immediately. Indeed, if we consider G(k) µλ (b) = b N X n=0 µλ((b−1X ⊗ Ik)n) + b ∞ X n=N +1 µλ((b−1X ⊗ Ik)n) for N large enough, the second term on the right hand side of equation has norm less than ǫ (this may be done uniformly over λ since all of the µλ are elements of Σ0,M ). Since the moments µλ((b−1X ⊗ Ik)n) converge weakly to µ((b−1X ⊗ Ik)n) uniformly for 0 ≤ n ≤ N , this implies that G(k) µ (b) weakly. As the above proof only relies on the fact that the transform may be written as a convergent series in the moments of our distributions, (4) follows in a similar manner. µλ (b) → G(k) Regarding (2) ⇒ (1), recall that we may recover a distribution from either its Voiculescu or Cauchy transform when considered as a fully matricial function (that is, consider G(n) for all n ∈ N, see [16]). Compactness of µ Σ0,M implies that {µλ}λ∈Λ has cluster points in this set. Assuming (2), if ν is any cluster point, by the argument in the previous paragraph, Gν is equal to Gµ for those b for which kb−1k < 1/M . Our claim follows from analytic continuation and the fact that we may recover a distribution from its Cauchy transform. The proof for (4) ⇒ (1) is similar. (cid:3) 4. W ∗-Algebras. Let (A, τ ) be a tracial W ∗-algebra and B ⊂ A a W ∗-subalgebra. There is a natural B-valued probability space (A, EB, B) where EB : A → B 6 JOHN D. WILLIAMS is the canonical conditional expectation (we shall refer to this triple as a tracial W ∗-probability space). We refer to [10] for an introduction to these constructions. We isolate the following facts for easy reference. Lemma 4.1. Let (A, EB, B) be as above. The expectation EB has the fol- lowing properties: (1) EB is a contraction with EB(1) = 1 (2) EB(bac) = bEB(a)c for all a ∈ A and b, c ∈ A. (3) τ (EB(x)y) = τ (xEB(y)) = τ (EB(x)EB(y)) for all x, y ∈ A (4) EB is a normal, completely positive map. Moreover, EB is the unique trace preserving map that satisfies property (2) Let Στ 0 ⊂ Σ0 denote those elements µ so that µ(P (X)) = τ (EB(P (a))) where (A, EB, B) form a tracial W ∗-probability space. These distributions were studied extensively in [16] and [17]. Note that this space is closed under the ⊞ operation through amalgamated free product constructions. The primary purpose of this section is to show that Στ 0 ∩ Σ0,M is compact in the pointwise weak toplogy (Corollary 4.6). We begin our study of these distributions with the following lemma. Lemma 4.2. Let (A, EB, B) be a tracial W ∗-probability space. Assume that X, Y ∈ A are B-free and EB(Y ) = 0. Then kXk ≤ kX + Y k. Proof. Let EBhXi : A → BhXi denote the canonical condition expectation. Observe that τ (EBhXi(Y )P (X)) = τ (Y P (X)) = τ (Y P (X)1) = τ (Y P (X)EB(1)) = τ (EB(Y P (X))) = τ (EB(Y )EB(P (X))) = 0 for all P (X) ∈ BhXi. The first two equalities follow from the properties of the expectation and the next to last equality follows from B-freeness. This implies that EBhXi(Y ) = 0 so that EBhXi(X + Y ) = X. As this map is a contraction, we have that kXk ≤ kX + Y k (cid:3) The following subordination result was originally proven in [6]. A simple approach to this theorem utilizing the structure of bialgebras was developed by Voiculescu [15]. We also refer to [7] for an extension of this theorem to free compression semigroups. Theorem 4.3. Let X, Y ∈ A be B-free random variables. Then, there exists a holomorphic map Φ(n) : M + n (B) such that n (B) → M + EMn(B)hXi([(X + Y ) ⊗ In − b]−1) = (X ⊗ In − Φ(n)(b))−1 This theorem implies that, for distributions µ and ν, the above holomor- phic map satisfies F (n) µ⊞ν (b) = F (n) µ (Φ(n)(b)) for b ∈ Mn(B)+. The following lemma is a simple consequence of subor- dination and is a slightly more general version of remark 3.1 for this class OPERATOR-VALUED INFINITE DIVISIBILITY. 7 of distributions. We follow [18] which addresses the scalar-valued case. For µ, ν, ρ ∈ Στ 0 satisfying µ = ν ⊞ ρ, we shall refer to the distributions ν and ρ as factors of µ. Lemma 4.4. Given a B-valued distribution µ there exists an open set Γ ⊂ n (B) such that (F (n) M + )−1 has analytic continuation to Γ for all factors ν or µ. Moreover, ℑ(ϕ(n) ν ν )(b) ≤ 0 for all b ∈ Γ. ρ Proof. Fix n ∈ N. In what follows, we will drop the n and refer to the functions of the form F (n) as Fρ. As is well know (see [16]), there exists subsets Γ1 and Γ2 of the form {b ∈ Mn(B)+ : ℑb > α, ℑb > βℜb} so that F −1 are respectively defined and have positive imaginary part. Utilizing Theorem 4.3, we have that Φ(n) ◦ F −1 (b) for b ∈ Γ1 ∩ Γ2. Since the left hand side may be continued to Γ1, the same must be true of F −1 µ (b) = F −1 and F −1 µ ν ν . ν With respect to the negativity claim, observe that b = Fν (F −1 (b)) = Fν(Φ(n) ◦ F −1 µ (b)) for b ∈ Γ1 ∩ Γ2 and that, through continuation, this is true for b ∈ Γ1. Recall that Fν satisfies ℑFν (b) ≥ ℑb. Thus, abusing notation by letting ϕ(n) ν denote the extension of the Voiculescu transfrom to Γ1, we have the following: ν ν (b) = Φ(n) ◦ F −1 ϕ(n) µ (b) − b = Φ(n) ◦ F −1 µ (b) − Fν (Φ(n) ◦ F −1 µ (b)) and our claim follows. (cid:3) We close the section with a theorem providing necessary and sufficient conditions that a distribution arises from a conditional expectation of tracial von Neumann algebras. Theorem 4.5. Let µ ∈ Σ0 where B is assumed to be a tracial von Neumann algebra. Then, µ ∈ Στ 0 if and only if the following conditions hold for all P (X), Q(X) ∈ BhXi: (1) τ (µ(P ∗(X)XP (X))) ≤ M τ (µ(P ∗(X)P (X))) . (2) τ (µ(P ∗(X)Q(X)))2 ≤ τ (µ(P ∗(X)P (X)))τ (µ(Q∗(X)Q(X))) (3) τ (µ(P (X)Q(X))) = τ (µ(Q(X)P (X))) Proof. Assume that µ(P (X)) = EB(P (a)). Condition (1) follows from pos- itivity of EB. Condition (2) follows from the fact that EB preserves τ and the Cauchy-Schwarz inequality. Condition (3) follows from the fact that EB preserves τ and that τ is a trace. We next assume conditions (1), (2) and (3) above. Let N = {N (X) ∈ BhXi : τ (µ(N (X)∗N (X))) = 0}. Condition (3) implies that N is closed un- der the adjoint operation. Conditions (2) and (3) implies that τ (N (X)P (X)) = τ (P (X)N (X)) = 0 for all P (X) ∈ BhXi. Thus, N is a 2-sided ideal in BhXi. We consider BhXi/N . Observe that τ (µ((Q∗(X)+N (X))(P (X)+N ′(X)))) = τ (µ(Q∗(X)P (X))) for all P (X), Q(X) ∈ BhXi and N (X), N ′(X) ∈ N . Thus, we have a well defined inner product hP (X) + N , Q(X) + N i = 8 JOHN D. WILLIAMS τ (µ(Q∗(X)P (X))) so that BhXi/N is a pre-Hilbert space. We denote by H its completion with respect to the norm defined by this inner product (in symbols, k · kH is the inner product norm and k · k is the norm on B). We define an action of BhXi/N on H through left multiplication. Ob- serve that, for b ∈ B, we have that kbk2 − b∗b is a positive element in B. Thus, kbk2 − b∗b = c∗c for some c ∈ B. As our notion of positiv- ity of µ is purely algebraic, we have that µ(P ∗(X)c∗cP (X)) ≥ 0 so that kbk2µ(P ∗(X)P (X)) ≥ µ(P ∗(X)b∗bP (X)) for all b ∈ B and P (X) ∈ BhXi. Therefore, given a monomial Xb1X · · · XbnX ∈ BhXi, we have the follow- ing: k((Xb1X · · · XbnX + N ) · (P (X) + N )k2 H = τ (µ(P ∗(X)Xb∗ nX · · · Xb∗ 1XXb1X · · · XbnXP (X)) ≤ M 2τ (µ(P ∗(X)Xb∗ ≤ M 2kb1k2τ (µ(P ∗(X)Xb∗ nX · · · Xb∗ nX · · · b∗ 1b1X · · · XbnXP (X)) 2XXb2 · · · XbnXP (X)) By induction, we have that k((Xb1X · · · XbnX + N ) · (P (X) + N )kH ≤ M n+1kb1k · · · kbnkkP (X) + N kH As this holds for all P (X) + N ∈ BhXi/N which is dense in H, we have that the monomials are bounded operators on this Hilbert space. Extending through linearity, we may imbed BhXi/N into B(H). Let A denote the weak closure of its image. The map µ : BhXi/N → B is well defined since, for N (X) ∈ N and b ∈ B, we have that τ (µ(N (X))b) = τ (µ(N (X)b)) = 0 by (2). This implies that µ(N (X)) = 0. To complete our proof, we must extend µ to all of A and show that this extension is positive, faithful, B-bimodular and satisfies condition (3). First, for each b ∈ B define ξb = b + N ∈ H. Let {Pλ(X) + N }λ∈Λ form a weakly Cauchy net in A. This implies that h(Pλ(X) + N )ξ1, ξ(b′b)∗i = τ (µ(Pλ(X))bb′) is Cauchy in C for all b, b′ ∈ B. Since functionals of this type induce the weak operator toplogy on B (with respect to the standard represenation), we have that the set {µ(Pλ(X) + N )}λ∈Λ is Cauchy in the weak operator topology so that we have a well defined extension with µ(a) = limλ(µ(Pλ(X) + N )) in this topology. In order to prove positivity, we may further assume that the net {Pλ(X)+ N }λ∈Λ converges to a ∈ A in the strong operator topology. Since products are continuous in this topology and the positive cone is weakly closed, we have that µ(a∗a) = limλ µ((P ∗ λ (X)Pλ(X)) ≥ 0. To prove that our extension is faithful, we again assume that Pλ(X) + N → a in the strong operator toplogy on A. Assuming that µ(a∗a) = limλ µ(P ∗ λ (X)Pλ(X)) = 0 where the limit is in the weak operator topology on B, we have that, for Q(X), R(X) ∈ OPERATOR-VALUED INFINITE DIVISIBILITY. 9 BhXi, h(Pλ(X) + N ) · (Q(X) + N ), (R(X) + N )i = τ (µ(Pλ(X)Q(X)R∗(X))) and the right hand side goes to 0 by condition (2) and weak continuity of τ . As elements of this type are dense in H, this implies that a = 0. Bimodularity and condition (3) follow through similar methods. To finish the proof, we define a trace on A by letting τ ′(a) = τ (µ(a)). Note that τ ′(µ(a)) = τ (µ(µ(a))) = τ (µ(1)µ(a)) = τ ′(a) so that µ is trace preserving. B-bimodularity of µ implies, by Theorem 4.1, that µ is the canonical conditional expectation. Lastly observe that µ(P (X) + N ) = µ(P (X)) so that our distribution arises from this expectation. (cid:3) Corollary 4.6. The set µτ vergence. In particular, Στ 0 is closed in the topology of pointwise weak con- 0 ∩ Σ0,M is compact for all M ∈ R+. Proof. Observe that, since τ is weakly continuous, conditions (1), (2), and (3) are closed under pointwise weak limits. Thus, Στ 0 ∩ Σ0,M is a closed subset of a compact set. (cid:3) 5. The Steinitz Lemma. The following theorem was originally proven by Steinitz in [12]. Lemma 5.1. Let {vi}k i=1 ⊂ RN be a set of elements in the unit ball, where RN is equipped with the Euclidean metric. Then, there exists a permutation σ of {1, . . . , k} such that for all 1 ≤ j ≤ k we have that Pj i=1 vσ(i) ≤ N . We refer to [8] for a modern proof of this lemma. We refer to [1] for a survey of its history and applications to convex geometry. The following simple corollary of this fact is singled out for easy reference. i=1 ⊂ RN such that vi ≤ ǫ and j=1 vi = v. Then, for each t ∈ (0, 1), there exists a subset σ ⊂ {1, . . . , n} Corollary 5.2. Consider vectors {vi}n Pn such that Pi∈σ vi − tv ≤ N ǫ. Proof. We assume that v = (v, 0, . . . , 0), vi = (ti1, ti2, . . . , tiN ) and wi = (ti2, ti3, . . . , tiN ) ∈ RN −1. Observe that Pn i=1 wi = 0 and wi ≤ ǫ. By the Steinitz lemma, we may assume that Pℓ i=1 wi ≤ (N − 1)ǫ for all ℓ = 1, . . . , n. This implies that Pℓ i=1 vi is contained in a tube about the line passing through v and the origin of radius (N −1)ǫ for all ℓ = 1, . . . , n. Since each of the vi has magnitude bounded by ǫ, the intermediate value property implies that there exists an m ∈ {1, . . . , n} such that Pm i=1 ti1 − tv ≤ ǫ/2. For this m, we have that Pm (cid:3) i=1 vi − tv ≤ N ǫ, proving our result. The following is easily derived from Corollary 5.2. The details are left to the reader. 10 JOHN D. WILLIAMS Corollary 5.3. Let t ∈ (0, 1) and {vij}i∈N,j=1,...,ni ⊂ RN satisfy kvijk → 0 uniformly over j as i ↑ ∞ and k(vi1 + · · · + vini) − vk → 0 for some v ∈ RN . Then, there exists a sequence of subsets σi ⊂ {1, 2, . . . , ni} such that k Pj∈σi Remark 5.4. Note that, through a trivial approximation argument, we may replace Euclidean space in the preceding corollary with Hilbert space. vij − tvk → 0 as i ↑ ∞. 6. Main Results We now formulate and prove our main result. We assume throughout that B ⊂ B(H) with H separable. For an elements b ∈ B we denote by δb the distribution defined by the equation δb(P (X)) = P (b) for all P (X) ∈ BhXi. Theorem 6.1. Consider µ, {µij}i∈N,j=1,...,ni ⊂ Στ {bi}i∈N ⊂ B satisfying the following properties: 0 and self adjoint elements (1) µi = µi1 ⊞ µi2 ⊞ · · · µini (2) µi → µ in the pointwise weak topology. (3) µij → δ0 in the pointwise weak topology, uniformly over j = 1, . . . , ni. ⊞ · · · ⊞ µ1/n ⊞ δbi ∈ Σ0,M for all i ∈ N. Then, for each n ∈ N, there is a µ1/n ∈ Στ where the convolution on the right hand side is n-fold. 0 such that µ = µ1/n Proof. Fix t = 1/p for p ∈ N. Let {ξk}k∈N ⊂ H denote a separable basis. We assume without loss of generality that each of the µij satisfies µij(X) = 0. Observe that lemma 4.2 implies that {bi}∞ i=1 are bounded in norm so that, by 4.6, we may assume that {δbi }i∈N converges in the pointwise weak topology (note that the limit point is an element of Στ 0 but need not be of the form δb for b ∈ B). Also observe that, by lemma 4.2, ⊞j∈σµij ⊞ δbi ∈ Σ0,M for any subset σ ⊂ {1, 2, . . . , ni}, so that ⊞j∈σµij ∈ Σ0,2M . Now, for a Voiculescu transfrom ϕν = (ϕ(ℓ) ν )ℓ∈N, we restrict our attention to ϕ(1) ν . Let {dn}n∈N ⊂ B be a family of self adjoint elements with dense linear span. Consider cn = dn +iλI where I is the unit in B and λ > 16M so that kc−1 are defined on {cn}n∈N for all i ∈ N and σ ⊂ {1, . . . , ni}, and that this function is completely determined by its values on this countable set. n k < 1/16M . Note that ϕ(1) µ and ϕ(1) ⊞j∈σµij The idea of the proof is to use the Steinitz lemma to construct a sequence of decompositions µi = νi⊞ρi so that νi subconverges to µt. Since this lemma is for finite dimensional spaces, we must truncate the Voiculescu transform. Towards this end, let PM : H → R2M be defined by PM (P∞ k=1 αkξk) = (ℑα1, ℜα1, . . . , ℑαM , ℜαM ). We then define a map ΦK,M,N : Σ0,M → R2KM N as follows: ΦKM N (µ) := K Y k=1 N Y n=1 PM (ϕ(1) µ (cn) · ξk) The purpose of this construction is that the relevant transforms are com- pletely determined by their values on {cn}n∈N. Since φ(1) µ (cn) ∈ B, these OPERATOR-VALUED INFINITE DIVISIBILITY. 11 elements are completely determined by their action on this basis for H. Thus, ϕ(1) µ may be recovered from {ΦKM N (µ)}K,M,N ∈N. Observe that ni X j=1 ΦKM N (µij) = ΦKM N (µi1 ⊞ · · · ⊞ µini) → ΦKM N (µ) − ΦKM N (δbi) Further note that the assumption that µij → δ0 in the pointwise weak topology, as well as lemma 3.2, implies that ΦKM N (µij) → (0, . . . , 0) uni- formly over i. If v is the limit point of ΦKM N (µ) − ΦKM N (δbi ), by corol- lary 5.3, there exists a sequence of subsets σi ⊂ {1, 2, . . . , ni} such that ΦKM N (µij) → tv. Thus, up to truncation, any cluster point of the Pj∈σi sequence of measures {⊞j∈σiµij ⊞ δtbi}i∈N will have Voiculescu transform equal to ϕµ1/p. Observe that the above proof also works for ϕ(ℓ) µ for all ℓ ∈ N. Thus, if we diagnolize over ℓ, K, M and N , we obtain a sequence of subsets σi ⊂ {1, . . . , ni} such that ϕ⊞j∈σi µij ⊞δtbi µ in the pointwise weak topology on an open set Γ ⊂ H +. As we saw in the opening comments, the sequence ⊞j∈σiµij ⊞ δtbi 0. By Corollary 4.6, this set is compact in the pointwise weak topology. By lemma 3.2, any cluster point of this sequence will have the required distribution, so our theorem holds. (cid:3) → tϕℓ is contained in Σ0,2M ∩ Στ 7. Conculusion and Acknowledgements. We begin by noting that, while the above proof may seem quite com- plex due to the correspondingly complex machinery, the underlying idea is quite simple. Indeed, if a measure µ is the limit of an infinitesimal array {µij}i∈N,j=1,...,ni, then, after taking the appropriate transforms and utilizing the Steinitz lemma, we we may construct a sequence of subsets σi ⊂ {1, 2, . . . , ni} so that ⊞j∈σiµij converges to µ1/n. In RN , this is pre- cisely the Steinitz lemma, so that the whole approach is to come up with a family of maps into Rn that allow us to exploit this lemma. Observe that the proof of our main result may be adapted to other prob- abilistic settings. Indeed, if we are to consider the classical theorem due to Hincin, the above proof may be adapted with the logarithm of the Fourier transform replacing the Voiculescu transform. Furthermore, utilizing re- mark 5.4, one would expect this proof to work for vector valued probability distributions. This is a somewhat more intuitive construction than the trad- tional function theoretic approach to these theorems since the more classical approach does not include the observation that a subset of the infinitesimal array actually converges to the distribution µ1/n. Lastly, this project has raised questions about the suitability of various weak topologies to the theory. In this paper, we develop a theory of pointwise weak convergence which, although slightly unnatural in an operator algebra, has the desirable properties that the unit ball is compact and that conver- gence in this topology corresponds to convergence of our various transforms. 12 JOHN D. WILLIAMS However, it is unclear whether weak topologies that are more intrinsic to these operator algebras behave well with respect to the transformations. In particular, the question arises as to whether weak convergence of elements in an operator algebra A corresponds to some type of convergence for their Voiculescu transforms. In the scalar valued case, weak convergence is equiva- lent to uniform convergence of the Voiculescu transforms on certain compact subsets in the complex upper half space (see [5]). Is there a corresponding theorem in the more general operator valued case? Acknowledgements: I would like to thank Hari Bercovici, Michael An- shelevitch, Michael Hartglass and Ken Dykema for their helpful advice. I would also like to thank Imre B´ar´any for referring me to the relevant liter- ature on the Steinitz lemma. References [1] I. B´ar´any. On the power of linear dependencies. In Building bridges, volume 19 of Bolyai Soc. Math. Stud., pages 31 -- 45. Springer, Berlin, 2008. [2] S. Belinschi, M. Popa, and V. Vinnikov. Infinite divisibility and a non-commutative boolean-to-free bercovicipata bijection. Journal of Functional Analysis, (0): -- , 2011. [3] S. T. Belinschi and H. Bercovici. Hincin's theorem for multiplicative free convolution. Canad. Math. Bull., 51(1):26 -- 31, 2008. [4] H. Bercovici and V. Pata. A free analogue of Hincin's characterization of infinite divisibility. Proc. Amer. Math. Soc., 128(4):1011 -- 1015, 2000. [5] H. Bercovici and D. Voiculescu. Free convolution of measures with unbounded sup- port. Indiana Univ. Math. J., 42(3):733 -- 773, 1993. [6] P. Biane. Processes with free increments. Math. Z., 227(1):143 -- 174, 1998. [7] S. Curran. Analytic Subordination for Free Compression. ArXiv e-prints, Mar. 2008. [8] V. S. Grinberg and S. V. Sevastjanov. Value of the Steinitz constant. Funktsional. Anal. i Prilozhen., 14(2):56 -- 57, 1980. [9] M. Popa and V. Vinnikov. Non-Commutative Functions and Non-Commutative Free Levy-Hincin Formula. ArXiv e-prints, July 2010. [10] A. M. Sinclair and R. R. Smith. Finite von Neumann algebras and masas, volume 351 of London Mathematical Society Lecture Note Series. Cambridge University Press, Cambridge, 2008. [11] R. Speicher. Combinatorial theory of the free product with amalgamation and operator-valued free probability theory. Mem. Amer. Math. Soc., 132(627):x+88, 1998. [12] E. Steinitz. Bedingt konvergente reihen und konvexe systeme. Journal fr die reine und angewandte Mathematik (Crelle's Journal), 143:128 -- 175, 1913. [13] D. Voiculescu. Addition of certain noncommuting random variables. J. Funct. Anal., 66(3):323 -- 346, 1986. [14] D. Voiculescu. Operations on certain non-commutative operator-valued random variables. Ast´erisque, (232):243 -- 275, 1995. Recent advances in operator algebras (Orl´eans, 1992). [15] D. Voiculescu. The coalgebra of the free difference quotient and free probability. Internat. Math. Res. Notices, (2):79 -- 106, 2000. [16] D. Voiculescu. Free analysis questions. I. Duality transform for the coalgebra of ∂X : B. Int. Math. Res. Not., (16):793 -- 822, 2004. [17] D.-V. Voiculescu. Free analysis questions II: the Grassmannian completion and the series expansions at the origin. J. Reine Angew. Math., 645:155 -- 236, 2010. OPERATOR-VALUED INFINITE DIVISIBILITY. 13 [18] J. D. Williams. A khintchine decomposition for free probability. Annals of Probability, Accepted for publication. Texas A&M University, Dept. of Mathematics,, Mail Stop 3368 , College Station, TX 77843-3368, [email protected]
1811.00546
1
1811
2018-11-01T16:42:27
Non-commutative Stein inequality and its applications
[ "math.OA", "math.FA" ]
The non-commutative Stein inequality asks whether there exists a constant $C_{p,q}$ depending only on $p, q$ such that \begin{equation*} \left\| \left(\sum_{n} |\mathcal{E}_{n} (x_n) |^{q}\right)^{\frac{1}{q}} \right\|_p \leq C_{p,q} \left\| \left(\sum_{n} | x_n |^q \right)^{\frac{1}{q}}\right \|_p\qquad \qquad (S_{p,q}), \end{equation*} for (positive) sequences $(x_n)$ in $L_p(\mathcal{M})$. The validity of $(S_{p,2})$ for $1 < p < \infty$ and $(S_{p,1})$ for $1 \leq p < \infty$ are known. In this paper, we verify (i) $(S_{p,\infty})$ for $1 < p \leq \infty$; (ii) $(S_{p,p})$ for $1 \leq p < \infty$; (iii) $(S_{p,q})$ for $1 \leq q \leq 2$ and $q<p<\infty$. We also present some applications.
math.OA
math
NON-COMMUTATIVE STEIN INEQUALITY AND ITS APPLICATIONS ALI TALEBI AND MOHAMMAD SAL MOSLEHIAN Abstract. The non-commutative Stein inequality asks whether there exists a con- stant Cp,q depending only on p, q such that (Sp,q), for (positive) sequences (xn) in Lp(M). The validity of (Sp,2) for 1 < p < ∞ and (Sp,1) for 1 ≤ p < ∞ are known. In this paper, we verify (i) (Sp,∞) for 1 < p ≤ ∞; (ii) (Sp,p) for 1 ≤ p < ∞; (iii) (Sp,q) for 1 ≤ q ≤ 2 and q < p < ∞. We also present some applications. En(xn)q! (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xn 1 q(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p ≤ Cp,q(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xn xnq! 1 q(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p 1. Introduction Throughout this note, (M, τ ) denotes a non-commutative probability space, that is, a von Neumann algebra M equipped with a normal faithful finite trace τ with τ (1) = 1, where 1 stands for the identity of M. For 1 ≤ p < ∞ the space Lp(M) is the completion of M with respect to the p-norm kxkp := τ (xp)1/p, where x = (x∗x)1/2 is the absolute value of x. If p = ∞, Lp(M) is M itself with the operator norm. We denote by L+ p (M) the positive cone in Lp(M) consisting of those elements, which are limits of sequences of positive elements in M. Assume that N is a von Neumann subalgebra of M. Then there exists a map EN : M −→ N , named the conditional expectation, which satisfies the following properties: (i) EN is normal positive contractive projection from M onto N ; (ii) EN (axb) = aEN (x)b for every x ∈ M and a, b ∈ N ; (iii) τ ◦ EN = τ . Moreover, EN is the unique map verifying (ii), (iii). It is known that EN can be extended to a contraction, denoted by the same EN , from Lp(M) into Lp(N ). By a filtration we mean an increasing sequence (Mn)n≥0 of von Neumann subalgebras Mn generates M in the w∗-topology. A sequence (xn)n≥0 in Lp(M) of M such that Sn≥0 2010 Mathematics Subject Classification. 46L53, 47A30, 60E15. Key words and phrases. Stein inequality; non-commutative probability space; trace; conditional expectation. 1 2 A. TALEBI, M.S. MOSLEHIAN is said to be adapted to (Mn)n≥0 if xn ∈ Lp (Mn) (n ≥ 0). The reader is referred to [14, 16, 17, 18] for more information on non-commutative probability spaces. Let (Ω, F , P) be a probability space, (Fn)∞ n=0 be a filtration on it and (Xn)∞ n=1 be a stochastic process adapted to (Fn)∞ n=0. The so-called Stein inequality (1.1) ( EFn−1Xnq) ∞ Xn=1 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 1 q(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p ≤ Cp,q(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ( Xnq) ∞ Xn=1 1 q(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p (Sp,q) was proved by Stein [15] for 1 < p < ∞ and q = 2 and by Asmar and Montgomery- Smith [1] for any 1 < p < ∞ and 1 ≤ q ≤ ∞. Using the duality between the martingale Hardy spaces and BMO-spaces, Lepingle [8] verified inequality (1.1) for adapted process with q = 2, p = 1 and Cp,q = 2. In a paper of Bourgain ([4], proposition 5) inequality (1.1) was obtained with a constant 3. It is evident that non- commutative probability theory is inspired by classical probability theory and quantum mechanics. Several mathematicians investigated and applied inequality (1.1) in the non-commutative setting. In 1997, Pisier and Xu [11] proved a non-commutative Stein inequality for 1 < p < ∞ and q = 2. They indeed proved that there is a constant Cp depending on p such that for any finite sequence (xn)N n=1 in Lp(M), N Xn=1 En(xn)2! (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 1 2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p ≤ Cp(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) N Xn=1 xn2! , 1 2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p See also [13] for another proof. In 2002, Junge [6] verified the non-commutative ana- logue of (1.1) for 1 ≤ p < ∞ and q = 1. In 2014, Qiu [12] employ a duality argument mentioned to show that if (Mn)N n=0 is an n=1 is an adapted sequence increasing filtration of von Neumann subalgebras and (xn)N in L1(M), then (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) N Xn=1 En−1(xn)2! 1 2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)1 ≤ 2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) N Xn=1 xn2! , 1 2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)1 where En denotes the conditional expectation with respect to Mn. The aim of this note is to obtain some versions of the non-commutative Stein inequal- ity (1.1) for other values p and q. More precisely, we verify (i) (Sp,∞) for 1 < p ≤ ∞ and q = ∞ in Theorem 2.8; (ii) (Sp,p) for 1 ≤ p < ∞ in Proposition 2.11; (iii) (Sp,q) for 1 ≤ q ≤ 2 and q < p < ∞ in Corollary 2.15. NON-COMMUTATIVE STEIN INEQUALITY 3 2. Stein inequality in noncommutative settings Suppose that (X, F , µ) is a measure space. The Banach space of all sequences f = (fn)n≥1 of functions such that the norm kf kp,q := ZX Xn p q fn(x)q! 1 p dµ(x)  is finite, is denoted by Lp(ℓq), in the case that 1 ≤ p, q < ∞. In the case that q = ∞, we set kf kp,∞ :=(cid:18)ZX(cid:18)sup n 1 p fn(x)(cid:19)p dµ(x)(cid:19) If p = ∞, we adopt the natural definition of essential supremum norm. The behavior of Lp(ℓq) is very similar to Lp(X, µ). For instance, the dual space of Lp(ℓq) (1 ≤ p, q < ∞) is Lp′(ℓq′), where 1 p + 1 p′ = 1 q + 1 q′ = 1; cf. [15]. Remark 2.1. Difficulty of having a non-commutative analogue of the space Lp(ℓ∞) is the lack of a non-commutative analogue of maximum. It is evident that there may be no maximum of two positive matrices. So we lead to the notion of Lp(M, ℓ∞); cf. [6]. We recall that the non-commutative spaces Lp(M, ℓ1) and Lp(M, ℓ∞) playing an essential role in non-commutative analysis. The results presented in this section come from [2, 6, 18]. Definition 2.2. The space of all sequences x = (xn)n∈N in Lp(M), which can be decomposed as xn = aynb for each n ∈ N, where a, b ∈ L2p(M) and y = (yn)n∈N is in M, is denoted by Lp(M, ℓ∞). It is known that Lp(M, ℓ∞) is a Banach space equipped with the norm kxkLp(M,ℓ∞) := inf(cid:26)kak2p sup n≥1 kynk∞ kbk2p(cid:27) , where the infimum runs over all possible decompositions of x as above. The element kxkLp(M,ℓ∞) is denoted by k sup+ n xnkp. The reader should notice that n xnkp is a notation for supn xn. Thus we may consider the space Lp(M, ℓ∞) as a k sup+ non-commutative analogue of the classical space Lp(ℓ∞). • Let Lp (M, ℓ1) be the space of all sequences x = (xn)n∈N in Lp(M), which admits kndkn (n ∈ N), where (ckn)k,n≥1 and (dkn)k,n≥1 It is known that Lp (M, ℓ1) equipped with the norm a factorization of the form xn =Pn∈N c∗ are sequences in L2p(M) such that Pk,n c∗ knckn and Pk,n d∗ kxkLp(M,ℓ1) := inf knckn(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) kndkn(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xk,n Xk,n  d∗ c∗ p 1 2 1 2 p ,   kndkn are in Lp(M). 4 A. TALEBI, M.S. MOSLEHIAN where the infimum runs over all possible decompositions of x as above, is a Banach space. It is proved in [18] that the space Lp (M, ℓ1) (1 ≤ p < ∞) is the predual of Lp′(M, ℓ∞) (where p′ is the conjugate to p) under the duality hx, yi = Pn∈N τ (xnyn) for x ∈ Lp (M, ℓ1) and y ∈ Lp′(M, ℓ∞). Remark 2.3. For any positive sequence x = (xn) in Lp (M, ℓ1) (i.e., xn ≥ 0 for all n) we have kxkLp(M,ℓ1) = kPn∈N xnkp. Therefore one may consider the space Lp (M, ℓ1) as a generalization of the space Lp(ℓ1) in the non-commutative setting. In the sequel, we present some useful properties of the spaces Lp (M, ℓ1) and Lp(M, ℓ∞). Theorem 2.4. ([18, Proposition 2.12]) Let 1 ≤ p ≤ ∞. (1) Each element in the unit ball of Lp(M, ℓ1) (resp. Lp(M, ℓ∞)) is a sum of eight (resp. sixteen) positive elements in the same ball. (2) For any x ∈ Lp(M, ℓ∞) it holds that + sup k n Moreover, if x is positive, xnkp = sup(Xn∈N xnkp = sup(Xn∈N + sup n k τ (xnyn) : y ∈ Lp′ (M, ℓ1) and kykLp′ (M,ℓ1) ≤ 1) . τ (xnyn) : yn ∈ L+ p′ and kXn∈N ynkp′ ≤ 1) . Some important spaces of sequences in Lp(M) can be formed via the row and column spaces, which are related to Burkholder -- Gundy non-commutative inequalities. Definition 2.5. Let 1 ≤ p ≤ ∞ and x = (xn)N two norms n=1 be a finite sequence in Lp(M). Set 1 spectively. kxkLp(M;ℓC 1 x∗ n2! xn2! and kxkLp(M;ℓR 2 ) :=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) N Xn=1 2 ) :=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p N Xn=1 The corresponding completion spaces are denoted by Lp(cid:0)M; ℓC 2(cid:1) and Lp(cid:0)M; ℓR 2(cid:1), re- Remark 2.6. Note that Lp(cid:0)M; ℓC 2(cid:1)) is isometric to the column (resp. 2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p 2(cid:1) (resp. Lp(cid:0)M; ℓR row) subspace of Lp (M ⊗ B(ℓ2)) via the following maps, where B(ℓ2) is the algebra of bounded linear maps on ℓ2 with its usual trace T r and M ⊗ B(ℓ2) denotes the von Neumann tensor product equipped with the (semifinite faithful) tensor trace τ ⊗ T r: . x = (xn)n≥0 7→  x0 0 . . . x1 0 . . . ... ...   (resp. x = (xn)n≥0 7→  . . . . . . x0 x1 0 0 ... ... ).   NON-COMMUTATIVE STEIN INEQUALITY 2(cid:1)∗ [11, p. 670]) that Lp(cid:0)M; ℓC q = 1 (1 ≤ p < ∞). p + 1 = Lq(cid:0)M; ℓC 2(cid:1)∗ 2(cid:1) and Lp(cid:0)M; ℓR 5 = It is known (cf. Lq(cid:0)M; ℓR 2(cid:1), whenever 1 For notational convenience, we recall the space CRp [Lp(M)]. • For 1 ≤ p ≤ ∞, let us define the space CRp [Lp(M)] in two cases as follows: • If p ≥ 2, with the following norm: CRp [Lp(M)] = Lp(cid:0)M; ℓC 2(cid:1) ∩ Lp(cid:0)M; ℓR 2(cid:1) k(xn)kCRp[Lp(M)] = maxnk(xn)kLp(M;ℓC 2 ), k(xn)kLp(M;ℓR • If p < 2, equipped with the following norm: CRp [Lp(M)] = Lp(cid:0)M; ℓC k(xn)kCRp[Lp(M)] = infnk(an)kLp(M;ℓC 2(cid:1) + Lp(cid:0)M; ℓR 2(cid:1) 2 ) + k(bn)kLp(M;ℓR 2 )o . 2 )o , where the infimum runs over all possible decompositions xn = an + bn with an and bn in Lp(M). In [7] Junge and Xu defined an interesting complex interpolation space between . The reader is Lp(M, ℓ∞) and Lp(M, ℓ1) as Lp(M, ℓq) := [Lp(M, ℓ∞), Lp(M, ℓ1)] 1 referred to [3] for interpolation theory and to [7] for some useful properties of the space Lp(M, ℓq). Note that if M is injective, this definition is a special case of Pisier's vector-valued non-commutative Lp-space theory [10]. q In the sequel, let (Mn)∞ n=0 be an increasing sequence of von Neumann subalgebras of M and En denote the conditional expectation of M with respect to Mn. We consider its extension, denoted by the same En from Lp(M) to Lp(Mn). In [6], a noncommutative Doob's inequality is obtained by the following dual version of Doob's inequality: Lemma 2.7. Let 1 ≤ p < ∞. Then for every sequence of positive elements (xn) in Lp(M) , (1 ≤ p < ∞) (DDp) The next result is a non-commutative Stein inequality for the case when 1 ≤ p < ∞ and q = ∞, which is a consequence of DDp. We state it for the sake of completeness. (2.1) (2.2) Xn∈N (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) En(xn)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p xn(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p ≤ Cp(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xn∈N En(x)(cid:13)(cid:13)(cid:13)(cid:13)p (cid:13)(cid:13)(cid:13)(cid:13) sup n in which Cp is a positive constant depending only on p. By a duality argument, it is deduced from DDp (see [6]) that ≤ Cp′kxkp. 6 A. TALEBI, M.S. MOSLEHIAN Theorem 2.8 (Sp,∞ for 1 < p ≤ ∞). Let 1 ≤ p < ∞ and x = (xn)n∈N be an arbitrary sequence in Lp(M). Then (cid:13)(cid:13)(En(xn))n∈N(cid:13)(cid:13)Lp(M,ℓ∞) ≤ C k(xn)n∈NkLp(M,ℓ∞) for some positive constant C depending only on p. (Sp,∞) Proof. It follows from Theorem 2.4(1) that any element in the unit ball of Lp(M, ℓ∞) is a finite linear combination of positive elements in the same ball. Hence it is enough to deduce the result for positive sequences. Using Theorem 2.4(2) we have τ (xnEn(yn)) : yn ∈ L+ p′ and kXn∈N ynkp′ ≤ 1) τ(cid:18)xn En(yn) Cp′ (cid:19) : yn ∈ L+ p′ and kXn∈N ynkp′ ≤ 1) τ (xnzn)) : zn ∈ L+ p′ and kXn∈N znkp′ ≤ 1) (by Theorem 2.4(2)) + sup n (cid:13)(cid:13)(cid:13)(cid:13) En(xn)(cid:13)(cid:13)(cid:13)(cid:13)p = sup(Xn∈N = Cp′ sup(Xn∈N ≤ Cp′ sup(Xn∈N xn(cid:13)(cid:13)(cid:13)(cid:13)p = Cp′(cid:13)(cid:13)(cid:13)(cid:13) + sup n To get the inequality above, we note that En is a positive map for any n and use the inequality DDp′ (2.1). (cid:3) Remark 2.9. If we put xn = x for all n in Theorem 2.8, then (Sp,∞) implies (2.2). Corollary 2.10. Let 1 ≤ p < ∞ and x = (fn)n∈N be an arbitrary sequence in Lp(Ω, F , P). Then for some positive constant C depending only on p. sup n∈N (cid:13)(cid:13)(cid:13)(cid:13) En(fn)(cid:13)(cid:13)(cid:13)(cid:13)p ≤ C(cid:13)(cid:13)(cid:13)(cid:13) sup n∈N fn(cid:13)(cid:13)(cid:13)(cid:13)p The next Proposition provides a non-commutative Stein inequality for the case when 1 ≤ p = q < ∞. We should notify that the case p = q = ∞ is proved in Theorem 2.8. Proposition 2.11 (Sq,q for 1 ≤ q < ∞). Let (xn)n∈N be an arbitrary sequence in Lq(M). Then (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xn∈N En−1(xn)q! Xn∈N xnq! 1 q(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)q ≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 1 q(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)q (Sq,q) for every real number q ≥ 1. NON-COMMUTATIVE STEIN INEQUALITY 7 Proof. We show the desired inequality by some properties of trace and conditional expectation as follows. 1 q q N N N ∈N N Xn=1 = sup N ∈N τ (En−1(xn)q) En−1(xn)q! (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) q(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xn∈N = τ ∞ En−1(xn)q! = τ sup Xn=1 Xn=1 Xn=1 τ N Xn=1 = (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xn∈N xnq! Xn=1 xnq! = τ ∞ xnq! Xn=1 q(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) kEn−1(xn)kq q ≤ sup N ∈N = sup N ∈N = sup N ∈N . q 1 q N kxnkq q En−1(xn)q!! (by the normality of τ ) (by the contractivity of En) (again by the normality of τ ) (cid:3) In [11], the non-commutative Stein inequality Sp,2 is shown for 1 < p < ∞ by apply- ing Burkholder -- Gundy inequality. Next, Qiu [12] obtained inequality S1,2 for adapted sequences. The next corollary provides Stein inequality in the spaces CRp [LP (M)]. Corollary 2.12. Let 1 < p < ∞ and x = (xn)n∈N be an adapted sequence in Lp(M). Then k(En−1(xn))nkCRp[LP (M)] ≤ Cp k(xn)nkCRp[LP (M)] , where Cp is the constant appeared in Sp,2. Proof. Suppose that p ≥ 2. By replacing xn by x∗ n in inequality Sp,2 we get k(En−1(xn))nkLp(M;ℓR Xn∈N En−1(x∗ n)2! 1 2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 1 2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p 2 ) =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xn∈N ≤ Cp(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) En−1(xn)∗2! 2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p n2! 2 ) . x∗ 1 = Cp k(xn)nkLp(M;ℓR Xn∈N (by the ∗-preserving property of En) (by Sp,2) 8 A. TALEBI, M.S. MOSLEHIAN Employing again inequality Sp,2, we infer that k(En−1(xn))nkCRp[Lp(M)] ≤ Cp k(xn)nkCRp[LP (M)] . Now let 1 < p < 2. Assume that xn = an + bn with an and bn in Lp (M). Then En−1(xn) = En(an) + En(bn) is a factorization of En−1(xn) for all n ∈ N. By applying inequality Sp,2 to (an)n and (bn)n, we have k(En−1(an))nkLp(M;ℓC 2 ) ≤ Cp k(an)nkLp(M;ℓC 2 ) and Hence k(En−1(bn))nkLp(M;ℓR 2 ) ≤ Cp k(bn)nkLp(M;ℓR 2 ) . ≤ k(En−1(an))nkLp(M;ℓC k(En−1(xn))nkCRp[LP (M)] = infnk(cn)nkLp(M;ℓC ≤ Cp(cid:16)k(an)nkLp(M;ℓC 2(cid:1) and (dn)n ∈ Lp(cid:0)M; ℓR 2(cid:1). with (cn)n ∈ Lp(cid:0)M; ℓC 2 ) + k(dn)nkLp(M;ℓR 2 ) + k(En−1(bn))nkLp(M;ℓR 2 ) 2 ) + k(bn)nkLp(M;ℓR 2 )o 2 )(cid:17) , where the infimum runs over all possible decompositions (En−1(xn))n = (cn)n + (dn)n Whence by taking infimum over all decompositions as above, we conclude the required inequality. (cid:3) Remark 2.13. Due to the dual version of Doob's inequality and noting that each element in the unit ball of Lp(M, ℓ1) is a sum of eight positive elements in the same ball (see 2.4(1)), we deduce that the linear map φ : Lp (M; ℓ1) → Lp (M; ℓ1) with φ ((xn)) = (En(xn)) is bounded. However, Junge obtained it with the bound Cp as same as the constant, which is obtained from inequality DDp (2.1). Moreover, φ is a bounded linear map on Lp (M; ℓ∞) by Theorem 2.8. Applying interpolation theorem one may deduce that φ : Lp (M; ℓq) → Lp (M; ℓq) is bounded. It seems that there is a problem to get the general Stein inequality. The problem is whether the equality k(xn)kLp(M;ℓq) =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xn xq n! 1 q(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p holds for any positive sequence (xn) or not. However we establish inequality Sp,q for 1 ≤ q ≤ 2 and p ≥ q in the following theorem. Theorem 2.14. Suppose that 1 ≤ q ≤ 2, p ≥ q, (yn) is a sequence of isometries in M, and (xn) in L+ p (M) is an arbitrary positive sequence. Then Xn∈N (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (En(y∗ n xn yn))q! 1 q(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p ≤ C(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xn∈N y∗ n xq n yn! , 1 q(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p NON-COMMUTATIVE STEIN INEQUALITY 9 where C is a positive constant depending only on p and q. Proof. We have (En(y∗ n xn yn))q! Xn∈N (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 1 q(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p (by the Choi -- Davis -- Jensen inequality (see e.g. [5, 9]) and the operator monotonicity of tr (0 < r ≤ 1)) (En ((y∗ n xn yn)q)! 1 q(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p ≤ (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xn∈N ≤ (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xn∈N q (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xn∈N q (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xn∈N ≤ C p = C p En (y∗ n xq q p2 p q q p2 n yn)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) n yn(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) n yn! p q y∗ n xq y∗ n xq (by the Jensen operator inequality) (by DD p q (2.1)) . 1 q(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p (cid:3) Corollary 2.15 (Sp,q for 1 ≤ q ≤ 2 and q < p < ∞). Suppose that 1 ≤ q ≤ 2, q < p < ∞ and (xn) in L+ p (M) is an arbitrary positive sequence. Then (2.3) (En(xn))q! Xn∈N (xn)q! 1 q(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p ≤ C(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xn∈N , 1 q(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p where C is a positive constant depending only on p and q. (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Proof. Apply Theorem 2.14 with yn = 1 for all n ∈ N. (cid:3) The following results give some applications of Sp,q (2.3), which is interesting on its own right. Corollary 2.16 (Semi-noncommutative case). Let (Ω, F , P) be a probability space, and (Fn)∞ n=0 be a filtration on it. If 1 ≤ q ≤ 2, q < p < ∞, then there exists a positive constant C such that for every positive stochastic process (fn)∞ n=1 with values in Lp(M) (En(fn))q! Xn∈N (fn)q! 1 q(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p ≤ C(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xn∈N , 1 q(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p in which En = En ⊗ idLp(M), where En is the usual conditional expectation with respect to Fn. Proof. Set N := L∞(Ω, F , P) ⊗ M and ν := E ⊗ τ , and apply (2.3). (cid:3) 10 A. TALEBI, M.S. MOSLEHIAN Corollary 2.17. If 1 ≤ q ≤ 2, q < p < ∞, then there exists a positive constant C, depending only on p and q, such that for every sequence of mutually orthogonal projections (rn)n≥1 in Lp(M) . (En(rn))q! Xn∈N (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≤ C 1 q(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p Corollary 2.18 (Matrix algebra case). If 1 ≤ q ≤ 2, q < p < ∞, then there exists a positive constant C such that for every sequence of positive matrices (Xn)n in Sp, the Schatten p-class on l2, equipped with the usual trace it holds that (En Xn En))q! (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xn∈N 1 q(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p ≤ C(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xn∈N (Xn)q! , 1 q(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p where En is the operator projecting a sequence in l2 into its n first coordinates. 2 ))n∈N. The corresponding conditional expectation from B(l2) onto B(ln Proof. Regarding B(ln 2 ) as a subalgebra of B(l2), we have an increasing filtration (B(ln 2 ) is En such that En(X) = EnXEn (X ∈ B(l2)), which extends to a contractive projection from Sp onto Sn (cid:3) p (see [18]). Now, apply (2.3). References [1] N. Asmar and S. Montgomery-Smith, Littlewood -- Paley theory on solenoids, Colloq. Math. 65 (1993), 69 -- 82. [2] T. N. Bekjan, Z. Chen and A. Osekowski, Noncommutative maximal inequalities associated with convex functions, Trans. Amer. Math. Soc. 369 (2017), no. 1, 409 -- 427. [3] J. Bergh and J. Lofstrom, Interpolation spaces, Springer-Verlag, Berlin, 1976. [4] J. Bourgain, Embedding L1 in L1/H 1, Trans. Amer. Math. Soc. 278 (1983), no. 2, 689 -- 702. [5] M. D. Choi, Some assorted inequalities for positive linear maps on C ∗-algebras, J. Operator Theory, 4 (1980), 271 -- 285. [6] M. Junge, Doob's inequality for non-commutative martingales, J. Reine Angew. Math. 549 (2002), 149 -- 190. [7] M. Junge and Q. Xu, Noncommutative Burkholder/Rosenthal inequalities II: Applications, Israel J. Math. 167 (2008), 227 -- 282. [8] D. Lepingle, Une in´egalit´e de martingales, (French) S´eminaire de Probabilities, XII (Univ. Stras- bourg, Strasbourg, 1976/1977), pp. 134 -- 137, Lecture Notes in Math., 649 (1978), Springer, Berlin. [9] M. S. Moslehian and H. Najafi, Around operator monotone functions, Integral Equations Operator Theory 71 (2011), no. 4, 575 -- 582. [10] G. Pisier, Non-commutative vector valued Lp-spaces and completely p-summing maps, Ast´erisque No. 247, 1998. [11] G. Pisier and Q. Xu, Non-commutative martingale inequalities, Comm. Math. Phys. 189 (1997), no. 3, 667 -- 698. [12] Y. Qiu, A non-commutative version of Lepingle -- Yor martingale inequality, Statist. Probab. Lett. 91 (2014), 52 -- 54. NON-COMMUTATIVE STEIN INEQUALITY 11 [13] N. Randrianantoanina, Non-commutative martingale transforms, J. Funct. Anal. 194 (2002), no. 1, 181 -- 212. [14] Gh. Sadeghi and M. S. Moslehian, Noncommutative martingale concentration inequalities, Illinois J. Math. 58 (2014), no. 2, 561 -- 575. [15] E. M. Stein, Topics in harmonic analysis related to the Littlewood-Paley theory, Annals of Math- ematics Studies, No. 63 Princeton University Press, Princeton, N.J.; University of Tokyo Press, Tokyo 1970. [16] A. Talebi, M. S. Moslehian and Gh. Sadeghi, Etemadi and Kolmogorov inequalities in noncom- mutative probability spaces, Michigan Math. J. (to appear), arXiv:1709.08044. [17] A. Talebi, M. S. Moslehian and Gh. Sadeghi, Noncommutative Blackwell -- Ross martingale in- equality, Infin. Dimens. Anal. Quantum Probab. Relat. Top. (to appear), arXiv:1705.07122. [18] Q. Xu, Operator spaces and non-commutative Lp, Lectures in the Summer School on Banach spaces and Operator spaces, Nankai University China, 2007. Department of Pure Mathematics, Ferdowsi University of Mashhad, P.O. Box 1159, Mashhad 91775, Iran. E-mail address: [email protected] E-mail address: [email protected] and [email protected]
1604.05717
1
1604
2016-04-19T10:54:58
Positive maps which are bijective on the set of rank projections
[ "math.OA", "quant-ph" ]
Extending Wigner's theorem we give a characterization of positive maps of $B(H)$ into itself which map the set of rank k projections onto itself.
math.OA
math
POSITIVE MAPS WHICH MAP THE SET OF RANK K PROJECTIONS ONTO ITSELF ERLING STØRMER Abstract. Extending Wigner's theorem we give a characterization of positive maps of B(H) into itself which map the set of rank k projections onto itself One form of the celebrated Wigner's theorem [5] is that if φ is a linear map of the bounded operators B(H) on a Hilbert space H into itself with the property that it maps the set of rank 1 projections bijectively onto itself, then φ is of the form (*) φ(a) = U aU ∗ or φ(a) = U atU ∗, where at is the transpose of a with respect to a fixed orthonormal basis for H, and U is a unitary operator. In the paper [2] Sarbicki, Chruscinski and Mozrzymas generalized this to the case when H is of finite dimension n with n a prime number, and the set of rank 1 projections is replaced by rank k projections, where k is a natural number strictly smaller than n. They gave a counter example to the conclusion (*) when n is not a prime. In that case φ is no longer a positive map. In the present note we make the extra assumption that φ is a positive unital map. Then for any Hilbert space we obtain the conclusion (*). Closely related results have been obtained by Molnar [1]. Recall that an atomic masa in B(H) is a maximal abelian subalgebra A generated by the rank 1 projections corresponding to the vectors in an orthonormal basis for H. Thus if H is finite dimensional each maximal abelian subalgebra is atomic. We start with a lemma. See also [1], Lemma 2.1.5. Lemma 1. Let p ∈ B(H) be a rank 1 projection and A an atomic masa in B(H) containing p. Let k be a natural number, k < dimH. Then there exist k + 1 projections P1, ..., Pk+1 in A such that p = 1 k k+1 X j=2 Pj − k − 1 k P1. Proof. Let p1 = p, p2, ..., pk+1 be mutually orthogonal rank 1 projections in A. Let Pj = k+1 X i=1,i6=j pi, j = 1, ...k + 1. Then Pj is a projection of rank k, and pi ≤ Pj for all j 6= i, so pi ≤ Pj for k of the projections Pj. It is therefore an easy computation to show the above formula. The proof is complete. Theorem 2. Let φ be a positive unital map of B(H) into itself such that φ maps the set of projections of rank k in B(H), k < dimH, onto itself. Then φ is of the form (*). Proof. Since each projection of rank k in B(H) is in the image under φ of a rank k projection, it follows from Lemma 1 that the rank 1 projections are in image of φ, hence each finite rank operator Date: 18-4-2016. 1 2 ERLING STØRMER is in the image of the finite rank operators. By continuity of φ it follows that φ when restricted to the compact operators C(H), maps C(H) onto a norm dense subset of itself. The definite set D of φ is the set of self-adjoint operators a such that φ(a2) = φ(a)2. Let Q be a projection of rank k; then P = φ(Q) is a projection of rank k, hence φ(Q2) = φ(Q) = P = P 2 = φ(Q)2, so that Q ∈ D. By [4], Proposition 2.1.7, D is a norm closed Jordan subalgebra of B(H), so by the same argument as above D ∩C(H) = C(H)sa , the self-adjoint operators in C(H). Furthermore, the restriction of φ to D is a Jordan homomorphism. Since C(H) is irreducible, by [3], Corollary 3.4, φ is either a homomorphism or an anti-homomorphism on C(H). But C(H) is a simple C*-algebra, so φ is either an automorphism or an anti-automorphism of C(H). Let now ωx be a vector state on B(H). Then if p is the rank 1 projection onto the 1-dimensional subspace of H generated by x , then for a ∈ B(H), ωx(a) = (ax, x) = T r(pap). Since φ is a Jordan automorphism of C(H) there is a unit vector y such that if q is the rank 1 projection onto the subspace spanned by y, then φ(q) = p. Thus for a ∈ B(H), since q ∈ D, we have, using [4], Proposition 2.1.7, ωx(φ(a)) = T r(pφ(a)p) = T r(φ(q)φ(a)φ(q)) = T r(φ(qaq)) = T r(φ(ωy(a))q) = ωy(a)T r(p) = ωy(a). We have thus shown that each vector state composed with φ is a vector state. Hence by [4] ,Theorem 3.3.2 φ is of the desired form (*). The proof is complete. References [1] L.Molnar, Selected preserver problems on algebraic structures of linear operators and on function spaces, Lecture Notes in Mathematics, 1895. Springer (2007). [2] G.Sarbicki, D.Chruscinski, and M.Mozrzymas, Generalising Wigner's theorem, arXiv: 1602.04968v1 (quant-ph), 16 Feb 2016. [3] E.Størmer, On the Jordan structure of C*-algebras, Trans.Amer.Math.Soc. (120) (1965), 438-447. [4] E.Størmer, Positive linear maps of operator algebras, Springer Monographs in Mathematics, Springer (2013). [5] E.P.Wigner, Group theory, Academic Press. New York (1959) Department of Mathematics, University of Oslo, 0316 Oslo, Norway. e-mail [email protected]
1909.06784
1
1909
2019-09-15T11:33:38
Controlled continuous g-Frames in Hilbert $C^{\ast}$-Modules
[ "math.OA", "math.FA" ]
Frame Theory has a great revolution in recent years, this Theory have been extended from Hilbert spaces to Hilbert $C^{\ast}$-modules. The purpose of this paper is the introduction and the study of the concept of Controlled Continuous g-Frames in Hilbert $C^{\ast}$-Modules. Also we give some properties.
math.OA
math
Controlled continuous g-Frames in Hilbert C ∗-Modules H. LABRIGUI1 ∗, A. TOURI1, and S. KABBAJ1 Abstract. Frame Theory has a great revolution in recent years, this Theory have been extended from Hilbert spaces to Hilbert C ∗-modules. The purpose of this paper is the introduction and the study of the concept of Controlled Continuous g-Frames in Hilbert C ∗-Modules. Also we give some properties. 1. Introduction and preliminaries The concept of frames in Hilbert spaces has been introduced by Duffin and Schaeffer [5] in 1952 to study some deep problems in nonharmonic Fourier series, after the fundamental paper [4] by Daubechies, Grossman and Meyer, frames theory began to be widely used, particularly in the more specialized context of wavelet frames and Gabor frames [7]. Hilbert C ∗-module arose as generalizations of the notion Hilbert space. The basic idea was to consider modules over C ∗-algebras instead of linear spaces and to allow the inner product to take values in the C ∗-algebras [18]. Continuous frames defined by Ali, Antoine and Gazeau [19]. Gabardo and Han in [8] called these kinds frames, frames associated with measurable spaces. For more details, the reader can refer to [15], [16] and [17]. Theory of frames have been extended from Hilbert spaces to Hilbert C ∗- modules [9], [11], [12], [13], [14]. In the following we briefly recall the definitions and basic properties of C ∗- algebra, Hilbert A-modules. Our reference for C ∗-algebras is [3, 6]. For a C ∗- algebra A if a ∈ A is positive we write a ≥ 0 and A+ denotes the set of positive elements of A. Definition 1.1. [3]. Let A be a unital C ∗-algebra and H be a left A-module, such that the linear structures of A and H are compatible. H is a pre-Hilbert A-module if H is equipped with an A-valued inner product h., .iA : H × H → A, such that is sesquilinear, positive definite and respects the module action. In the other words, (i) hx, xiA ≥ 0 for all x ∈ H and hx, xiA = 0 if and only if x = 0. (ii) hax + y, ziA = ahx, yiA + hy, ziA for all a ∈ A and x, y, z ∈ H. (iii) hx, yiA = hy, xi∗ A for all x, y ∈ H. 2010 Mathematics Subject Classification. 41A58, 42C15. Key words and phrases. Continuous g-Frames, Controlled continuous g-frames, C ∗-algebra, Hilbert A-modules. ∗ Corresponding author. 1 2 H. LABRIGUI∗, A. TOURI, S. KABBAJ 1 1 2 and the A-valued norm on H is defined by x = hx, xi For x ∈ H, we define x = hx, xiA 2 . If H is complete with ., it is called a Hilbert A-module or a Hilbert C ∗-module over A. For every a in C ∗-algebra A, we have a = (a∗a) A for all x ∈ H. Let H and K be two Hilbert A-modules, A map T : H → K is said to be adjointable if there exists a map T ∗ : K → H such that hT x, yiA = hx, T ∗yiA for all x ∈ H and y ∈ K. A(H,K) for the set of all adjointable operators from H to K and End∗ A(H,H) is abbreviated to End∗ We reserve the notation End∗ A(H). 1 2 The following lemmas will be used to prove our mains results Lemma 1.2. [10]. Let H be Hilbert A-module. If T ∈ End∗ A(H), then hT x, T xi ≤ kTk2hx, xi ∀x ∈ H. Lemma 1.3. [1]. Let H and K two Hilbert A-modules and T ∈ End∗(H,K). Then the following statements are equivalent: (i) T is surjective. (ii) T ∗ is bounded below with respect to norm, i.e., there is m > 0 such that (iii) T ∗ is bounded below with respect to the inner product, i.e., there is m′ > 0 kT ∗xk ≥ mkxk for all x ∈ K. such that hT ∗x, T ∗xi ≥ m′hx, xi for all x ∈ K. Lemma 1.4. [2]. Let H and K two Hilbert A-modules and T ∈ End∗(H,K). Then: (i) If T is injective and T has closed range, then the adjointable map T ∗T is invertible and (ii) If T is surjective, then the adjointable map T T ∗ is invertible and k(T ∗T )−1k−1 ≤ T ∗T ≤ kTk2. k(T T ∗)−1k−1 ≤ T T ∗ ≤ kTk2. 2. Controlled continuous g-Frames in Hilbert C ∗-Modules Let X be a Banach space, (Ω, µ) a measure space, and function f : Ω → X a measurable function. Integral of the Banach-valued function f has defined Bochner and others. Most properties of this integral are similar to those of the integral of real-valued functions. Because every C ∗-algebra and Hilbert C ∗- module is a Banach space thus we can use this integral and its properties. Let (Ω, µ) be a measure space, let U and V be two Hilbert C ∗-modules, {Vw}w∈Ω A(U, Vw) is the collection of all adjointable is a sequence of subspaces of V, and End∗ A-linear maps from U into Vw. We define ⊕w∈ΩVw = (cid:26)x = {xw}w∈Ω : xw ∈ Vw,(cid:13)(cid:13)(cid:13)(cid:13) ZΩ xw2dµ(w)(cid:13)(cid:13)(cid:13)(cid:13) For any x = {xw}w∈Ω and y = {yw}w∈Ω, if the A-valued inner product is defined by hx, yi = RΩhxw, ywidµ(w), the norm is defined by kxk = khx, xik 2 , the ⊕w∈ΩVw < ∞(cid:27) . 1 CONTROLLED CONTINUOUS G-FRAMES IN HILBERT C ∗-MODULES 3 is a Hilbert C ∗-module. Let GL+(U) be the set for all positive bounded linear invertible operators on U with bounded inverse. Definition 2.1. [20] We call {Λw ∈ End∗ for Hilbert C ∗-module U with respect to {Vw : w ∈ Ω} if: A(U, Vw) : w ∈ Ω} a continuous g-frame measurable; • for any x ∈ U, the function x : Ω → Vw defined by x(w) = Λwx is • there exist two strictly nonzero elements A and B in A such that (2.1) Ahx, xi ≤ ZΩhΛwx, Λwxidµ(w) ≤ Bhx, xi,∀x ∈ U. The elements A and B are called continuous g-frame bounds. If A = B we call this continuous g-frame a continuous tight g-frame, and if A = B = 1A it is called a continuous Parseval g-frame. If only the right- hand inequality of (2.1) is satisfied, we call {Λw : w ∈ Ω} a continuous g-Bessel sequence for U with respect to {Vw : w ∈ Ω} with Bessel bound B. The contnuous g-frame operator S on U is : Sx = ZΩ Λ∗ ωΛωxdµ(ω) The frame operator S is a bounded, positive, selfadjoint, and invertible (see [20]) Theorem 2.2. [20] Let Λ = {Λw ∈ End∗ A(U, Vw) : w ∈ Ω}, then Λ be a con- tinuous g-frame for U with respect to {Vw : w ∈ Ω} if and only if there exist a constants A and B such that for any x ∈ U : Akxk2 ≤ kZΩhΛwx, Λwxidµ(w)k ≤ Bkxk2 Definition 2.3. Let C, C ′ ∈ GL+(U), we call Λ = {Λw ∈ End∗ A(U, Vw) : w ∈ Ω} a (C−C ′)-controlled continuous g-frame for Hilbert C ∗-module U with respect to {Vw : w ∈ Ω} if Λ is continuous g-Bessel sequence and there exist two constants A > 0 and B < ∞ such that : (2.2) Ahx, xi ≤ ZΩhΛwCx, ΛwC xidµ(w) ≤ Bhx, xi,∀x ∈ U. ′ A and B are called the (C − C ′)-controlled continuous g-frames bounds. If C ′ = I then we call Λ a C-controlled continuous g-frames for U with respect to {Vw : w ∈ Ω}. A(U, Vw) : w ∈ Ω} be a continuous g-frames for U with Let Λ = {Λw ∈ End∗ respect to {Vw : w ∈ Ω}. The bounded linear operator TCC ′ : l2({Vw}w∈Ω) → U given by TCC ′ ({yw}w∈Ω) = ZΩ ωyωdµ(w) (CC 2 Λ∗ ) ′ 1 ∀{yw}w∈Ω ∈ l2({Vw}w∈Ω) is called the synthesis operator for the (C − C ′)-controlled continuous g-frame {Λw}w∈Ω. 4 H. LABRIGUI∗, A. TOURI, S. KABBAJ The adjoint operator T ∗ CC ′ : U → l2({Vw}w∈Ω) given by T ∗ CC ′ (x) = {Λω(C 2 x}ω∈Ω C) 1 ′ ∀x ∈ U (2.3) is called the analysis operator for the (C − C ′)-controlled continuous g-frame {Λww ∈ Ω}. When C and C ′ commute with each other, and commute with the operator Λ∗ ωΛω for each ω ∈ Ω, then the (C − C ′)-controlled continuous g-frames operator: wΛwCxdµ(w) SCC ′ : U −→ U is defined as: SCC ′ x = TCC ′ T ∗ From now on we assume that C and C ′ commute with each other, and commute with the operator Λ∗ Proposition 2.4. The (C − C ′)-controlled continuous g-frames operator SCC ′ is bounded, positive, sefladjoint and invertible. ωΛω for each ω ∈ Ω CC ′ x = RΩ C ′Λ∗ Proof. . We show that SCC ′ is a bounded operator: kSCC ′k = sup x∈U,kxk≤1khSCC ′ x, xik = sup Λ∗ wΛwCxdµ(w)k ≤ B From the (C − C ′)-controlled continuous g-frames identity (2.2), we have: C ′ x∈U,kxk≤1ZΩ Ahx, xi ≤ hSCC ′ x, xi ≤ Bhx, xi so A.IdU ≤ SCC ′ ≤ B.IdU Where IdU is the identity operator in U. We clearly see that SCC ′ is a positive operator. Thus the (C − C ′)-controlled continuous g-frames operator SCC ′ is bounded and invertible In other hand we know every positive operator is self adjoint. A(U, Vw) : w ∈ Ω} is (C − C ′)-controlled Theorem 2.5. Let Λ = {Λw ∈ End∗ continuous g-bessel sequence for U, then Λ is a (C − C ′)-controlled continuous g-frames for U with respect to {Vw : w ∈ Ω} if and only if there exist a positive constants A and B such that : (cid:3) (2.4) Akxk2 ≤ kZΩhΛwCx, ΛwC ′ xidµ(w)k ≤ Bkxk2 ∀x ∈ U. Proof. Let {Λw}w∈Ω be a (C − C ′)-controlled continuous g-frames for U with respect to {Vw : w ∈ Ω} with bounds A and B. Hence, we have (2.5) Ahx, xi ≤ ZΩhΛwCx, ΛwC ′ xidµ(w) ≤ Bhx, xi ∀x ∈ U. Since 0 ≤ hx, xi, ∀x ∈ U , then we can take the norme in the left, middle and right termes of the above inequality (2.5). Thus we have: kAhx, xik ≤ kZΩhΛwCx, ΛwC ′ xidµ(w)k ≤ kBhx, xik ∀x ∈ U. CONTROLLED CONTINUOUS G-FRAMES IN HILBERT C ∗-MODULES 5 So, Akxk2 ≤ kZΩhΛwCx, ΛwC ′ xidµ(w)k ≤ Bkxk2 ∀x ∈ U. Conversely, suppose that (2.4) holds,we have: (2.6) 1 2 CC ′ x, S hS 2 1 CC ′ xi = hSCC ′ x, xi = ZΩhΛwCx, ΛwC ′ xidµ(w) using (2.6) in (2.5), we get for all x ∈ U: √Akxk ≤ kS by lemma 1.3 ∃m, M > 0 such that : mhx, xi ≤ hS CC ′ x, S 1 1 1 2 2 CC ′ xk ≤ √Bkxk 2 CC ′ xi ≤ Mhx, xi Therefore {Λw : w ∈ Ω} is a (C − C ′)-controlled continuous g-frames for U with respect to {Vw : w ∈ Ω} Theorem 2.6. Let C ∈ GL+(U), the sequence Λ = {Λw ∈ End∗ A(U, Vw) : w ∈ Ω} is a continuous g-frame for U with respect to {Vw : w ∈ Ω} if and only if Λ is a (C − C)-controlled continuous g-frames for U with respect to {Vw : w ∈ Ω} Proof. Suppose that {Λw : w ∈ Ω} is (C − C)-controlled continuous g-frames with bounds A and B, then : (cid:3) Akxk2 ≤ kZΩhΛwCx, ΛwCxidµ(w)k ≤ Bkxk2 ∀x ∈ U. For any x ∈ U, we have: Akxk2 = AkCC −1xk2 ≤ AkCk2kC −1xk2 ≤ kCk2kZΩhΛwCC −1x, ΛwCC −1xidµ(w)k = kCk2kZΩhΛwx, Λwxidµ(w)k hence, (2.7) on the other hand : AkCk−2kxk2 ≤ kZΩhΛwx, Λwxidµ(w)k kZΩhΛwx, Λwxidµ(w)k = kZΩhΛwCC −1x, ΛwCC −1xidµ(w)k (2.8) kZΩhΛwx, Λwxidµ(w)k ≤ BkC −1xk2 ≤ BkC −1k2kxk2 From (2.7) and (2.8) and theorem2.2 we conclude that {Λw, w ∈ Ω} is a contin- uous g-frame with bounds AkCk−2 and BkC −1k2 Conversely, let {Λw, w ∈ Ω} is a continuous g-frame with bounds E and F , then 6 H. LABRIGUI∗, A. TOURI, S. KABBAJ for all x ∈ U we have : Ehx, xi ≤ ZΩhΛwx, Λwxidµ(w) ≤ Fhx, xi So, for all x ∈ U, Cx ∈ U, and : (2.9) ZΩhΛwCx, ΛwCxidµ(w) ≤ FhCx, Cxi ≤ FkCk2hx, xi Also, for all x ∈ U, Ehx, xi = EhC −1Cx, C −1Cxi ≤ EkC −1k2hCx, Cxi then, (2.10) Ehx, xi ≤ kC −1k2ZΩhΛwCx, ΛwCxidµ(w) From (2.9) and (2.10), we have: EkC −1k−2hx, xi ≤ ZΩhΛwCx, ΛwCxidµ(w) ≤ FkCk2hx, xi Hence Λ is a (C − C)-controlled continuous g-frames with bounds EkC −1k−2 and FkCk2 (cid:3) Proposition 2.7. Let {Λw, w ∈ Ω} is a continuous g-frame for U with respect to {Vw : w ∈ Ω} and S the continuous g-frame operator associated. Let C, C ′ ∈ GL+(U), then {Λw, w ∈ Ω} is (C − C ′)-controlled continuous g-frames Proof. Let {Λw, w ∈ Ω} is a continuous g-frame with bounds A and B. by theorem (2.2) we have: Akxk2 ≤ kZΩhΛwx, Λwxidµ(w)k ≤ Bkxk2 ∀x ∈ U =⇒ Akxk2 ≤ khSx, xik ≤ Bkxk2 ∀x ∈ U (2.11) and (2.12) kZΩhΛwCx, ΛwC ′ xidµ(w)k = kCkkC ′ kkZΩhΛwx, Λwxidµ(w)k = kCkkC ′ kkhSx, xik From (2.11) and (2.12), we have : AkCkkC ′ kkxk2 ≤ kZΩhΛwCx, ΛwC ′ xidµ(w)k ≤ BkCkkC ′ kkxk2 ∀x ∈ U we conclude by theoreme 2.5 that {Λw, w ∈ Ω} is (C − C ′)-controlled continuous g-frames with bounds AkCkkC ′k and BkCkkC ′k (cid:3) CONTROLLED CONTINUOUS G-FRAMES IN HILBERT C ∗-MODULES 7 A(U, Vω) and let C, C ′ ∈ GL+(U) so that Theorem 2.8. Let {Λw, w ∈ Ω} ⊂ End∗ C, C ′ commute with each other and commute with Λ∗ ωΛω for all ω ∈ Ω. Then the following are equivalent : (1) the sequence {Λw, w ∈ Ω} is a (C − C ′)-controlled continuous g-Bessel se- quence for U with respect {Vω}ω∈Ω with bounds A and B (2) The operator TCC ′ : l2({Vw}w∈Ω) → U given by TCC ′ ({yw}w∈Ω) = Zw∈Ω (CC ′ ) 1 2 Λ∗ ωyωdµ(w) ∀{yw}w∈Ω ∈ l2({Vw}w∈Ω) is well defined and bounded operator with kTCC ′k ≤ √B Proof. (1) =⇒ (2) Let {Λw, w ∈ Ω} be a (C − C ′)-controlled continuous g-Bessel sequence for U with respect {Vω}ω∈Ω with bound B. From theorem 2.5 we have : (2.13) kZΩhΛwCx, ΛwC ′ xidµ(w)k ≤ Bkxk2 ∀x ∈ U. For any sequence {yw}w∈Ω ∈ l2({Vω}ω∈Ω) kTCC ′ ({yw}w∈Ω)k2 = sup ′ ′ ) (CC 1 2 Λ∗ ) 1 2 Λ∗ = sup = sup = sup ωyω, xidµ(w)k2 ωyωdµ(w), xik2 x∈U,kxk=1khTCC ′ ({yw}w∈Ω), xik2 x∈U,kxk=1khZΩ x∈U,kxk=1kZΩh(CC x∈U,kxk=1kZΩhyω, Λω(CC 2 xidµ(w)k2 x∈U,kxk=1kZΩhyω, yωidµ(w)kkZΩhΛω(CC ≤ sup x∈U,kxk=1kZΩhyω, yωidµ(w)kkZΩhΛωCx, ΛωC x∈U,kxk=1kZΩhyω, yωidµ(w)kBkxk2 = Bk{yω}ω∈Ωk2 ≤ sup = sup ′ 1 ) ′ 1 ) 2 x, Λω(CC ′ ) 1 2 xidµ(w)k ′ xidµ(w)k Then we have kTCC ′ ({yw}w∈Ω)k2 ≤ Bk{yω}ω∈Ωk2 =⇒ kTCC ′k ≤ √B ′ is well defined and bounded we conclude the operator TCC (2) =⇒ (1) Let the operator TCC ′ is well defined, bounded and kTCC ′k ≤ √B For any x ∈ U and finite subset Ψ ⊂ Ω, we have: ZΨhΛwCx, ΛwC ′ xidµ(w) = ZΨhC ′ Λ∗ wΛwCx, xidµ(w) 8 H. LABRIGUI∗, A. TOURI, S. KABBAJ 1 ′ ′ 1 ) ) 2 Λ∗ wΛw(CC = ZΨh(CC = hTCC ′ ({yw}w∈Ψ), xi ≤ kTCC ′kk({yw}w∈Ψ)kkxk 2 x, xidµ(w) 1 Where: yw = Λw(CC ′) Therefore, ZΨhΛwCx, ΛwC 2 x if ω ∈ Ψ and yw = 0 if ω /∈ Ψ ′k(ZΨ kΛw(CC xidµ(w) ≤ kTCC ′ ′ ) 1 2 xk2dµ(w)) 1 2kxk = kTCC ′k(ZΨhΛwCx, ΛwC ′ xidµ(w)) 1 2kxk Since Ψ is arbitrary, we have: ZΩhΛwCx, ΛwC ′ xidµ(w) ≤ kTCC ′k2kxk2 =⇒ ZΩhΛwCx, ΛwC ′ xidµ(w) ≤ Bkxk2 as : kTCC ′k ≤ √B Therfore {Λw, w ∈ Ω} is a (C − C ′)-controlled continue g-Bessel sequence for U with respect to {Vω}ω∈Ω Proposition 2.9. Let Λ = {Λw ∈ End∗ A(U, Vw) : w ∈ Ω} and Γ = {Γw ∈ A(U, Vw) : w ∈ Ω} be two (C − C ′)-controlled continue g-Bessel sequence for End∗ U with respect to {Vω}ω∈Ω with bounds E1 and E2 respectively. Then the operator LCC ′ : U −→ U given by: (2.14) wΛωCxdµ(w) Γ∗ C (cid:3) ′ LCC ′ (x) = ZΩ is well defined and bounded with kLCC ′k ≤ √E1E2. Also its adjoint operator is CC ′ (g) = RΩ C ′Λ∗ L∗ Proof. for any x ∈ U and Ψ ⊂ Ω, we have : kZΨ Γ∗ wΛωCxdµ(w), yik2 C Γ∗ wΛωCxdµ(w)k2 = sup wΓωCxdµ(w) C ′ ∀x ∈ U since Ψ is arbitrary, RΨ C ′Γ∗ kLCC ′k = kZΨ ′ C Γ∗ wΛωCxdµ(w)k ≤ pE1E2 ′ ′ = sup yidµ(w)k2 y∈U,kyk=1khZΨ y∈U,kyk=1kZΨhΛωCx, ΓwC y∈U,kyk=1kZΨhΛωCx, ΛωCxidµ(w)kkZΨhΓwC ≤ sup ≤ kZΨhΛωCx, ΛωCxidµ(w)kE2 ≤ E1E2kxk2 wΛωCxdµ(w) converge in U and ′ ′ y, ΓwC yidµ(w)k CONTROLLED CONTINUOUS G-FRAMES IN HILBERT C ∗-MODULES 9 In other hand, we have: hLCC ′ ′ C ′ x, yi = hZΨ = ZΨhC = ZΨhx, CΛ∗ = hx,ZΨ CΛ∗ Γ∗ wΛωCxdµ(w), yi Γ∗ wΛωCx, yidµ(w) wΓωC ′ yidµ(w) wΓωC ′ ydµ(w)i (cid:3) CC ′ (g) = RΩ C ′Λ∗ Thus L∗ wΓωCxdµ(w) A(U, Vw) : w ∈ Ω} be a (C − C ′)-controlled Theorem 2.10. Let Λ = {Λw ∈ End∗ continue g-frames for U with respect to {Vω}ω∈Ω and Γ = {Γw ∈ End∗ A(U, Vw) : w ∈ Ω} be a (C − C ′)-controlled continue g-Bessel sequence for U with respect to {Vω}ω∈Ω. Assume that C and C ′ commute with each other and commute with Γ∗ wΓw. If the operator LCC ′ defined in (2.14) is surjective then Γ = {Γw : w ∈ Ω} is also a (C − C ′)-controlled continue g-frames for U with respect to {Vω}ω∈Ω Proof. Let Λ = {Λw ∈ End∗ g-frames for U with respect to {Vω}ω∈Ω. by theorem 2.8, the operator TCC ′ : l2({Vw}w∈Ω) → U given by A(U, Vw) : w ∈ Ω} be a (C − C ′)-controlled continue TCC ′ ({yw}w∈Ω) = Zw∈Ω (CC ′ ) 1 2 Λ∗ ωyωdµ(w) ∀{yw}w∈Ω ∈ l2({Vw}w∈Ω) is well defined and bounded operator. By (2.3) its adjoint operator T ∗ (2.15) CC ′ : U → l2({Vw}w∈Ω) given by T ∗ CC ′ (x) = {Λω(C 1 ′ C) 2 x}ω∈Ω Since Γ = {Γw : w ∈ Ω} is also a (C−C ′)-controlled continue g-Bessel sequence for U with respect to {Vω}ω∈Ω. Again by theorem 2.8, the operator KCC ′ : l2({Vw}w∈Ω) → U given by ′ ({yw}w∈Ω) = Zw∈Ω ∀{yw}w∈Ω ∈ l2({Vw}w∈Ω) is well defined and bounded operator. Again its adjoint operator is given by ωyωdµ(w) KCC (CC 2 Γ∗ ) 1 ′ K ∗ CC ′ (x) = {Γω(C 1 ′ C) 2 x}ω∈Ω ∀x ∈ U Hence for any x ∈ U, the operator defined in (2.14) can be written as : LCC ′ (x) = ZΩ ′ C Γ∗ wΛωCxdµ(w) = KCC ′ T ∗ CC ′ x ′ is surjective then for any x ∈ U, there exists y ∈ U such that: Since LCC x = LCC ′ x = KCC ′ T ∗ CC ′ x ∈ l2({Vw}w∈Ω) This implies that KCC ′ is surjective. As a result of lemma1.4, we have K ∗ bounded below, that is there exists m > 0 such that: CC ′ x and T ∗ CC ′ is CC ′ x, K ∗ hK ∗ CC ′ xi ≥ mhx, xi ∀x ∈ U 10 H. LABRIGUI∗, A. TOURI, S. KABBAJ =⇒ hKCC 2 Γ∗ (CC ) ′ 1 ′ K ∗ CC ′ x, xi ≥ mhx, xi ∀x ∈ U wΓω(CC ′ ) 1 2 xdµ(ω), xi ≥ mhx, xi ∀x ∈ U =⇒ hZΩ =⇒ ZΩhΓwCx, ΓwC ′ xidµ(ω) ≥ mhx, xi ∀x ∈ U Hence Γ = {Γw : w ∈ Ω} is also a (C − C ′)-controlled continue g-frames for U with respect to {Vω}ω∈Ω (cid:3) References Amer. Math. Soc. 135 (2007) 469-478. 1. L. Arambasi´c, On frames for countably generated Hilbert C∗-modules, Proc. 2. A.Alijani,M.Dehghan, ∗-frames in Hilbert C∗modules,U. P. B. Sci. Bull. Se- 3. J.B.Conway ,A Course In Operator Theory,AMS,V.21,2000. 4. I. Daubechies, A. Grossmann, and Y. Meyer, Painless nonorthogonal expan- ries A 2011. sions, J. Math. Phys. 27 (1986), 1271-1283. 5. R. J. Duffin, A. C. Schaeffer, A class of nonharmonic fourier series, Trans. Amer. Math. Soc. 72 (1952), 341-366. 6. F. R. Davidson, C∗-algebra by example,Fields Ins. Monog. 1996. 7. D. Gabor, Theory of communications, J. Elec. Eng. 93 (1946), 429-457. 8. J. P. Gabardo and D. Han, Frames associated with measurable space, Adv. Comp. Math. 18 (2003), no. 3, 127-147. 9. M. Frank, D. R. Larson, Frames in Hilbert C ∗-modules and C ∗-algebras, J. Oper. Theory 48 (2002), 273-314. 10. W. Paschke, Inner product modules over B∗-algebras, Trans. Amer. Math. Soc., (182)(1973), 443-468. modules, Ann. Univ. Paedagog. Crac. Stud. Math. 17 (2018), 15-24. 11. M. Rossafi and S. Kabbaj, ∗-g-frames in tensor products of Hilbert C ∗- 12. M. Rossafi and S. Kabbaj, ∗-K-g-frames in Hilbert C ∗-modules, Journal of 13. M. Rossafi, A. Touri, H. Labrigui and A. Akhlidj, Continuous ∗-K-G-Frame in Hilbert C ∗-Modules, Journal of Function Spaces, vol. 2019, Article ID 2426978, 5 pages, 2019. Linear and Topological Algebra Vol. 07, No. 01, 2018, 63-71. 14. M. Rossafi and S. Kabbaj, Generalized Frames for B(H,K), accepted for publication in Iranian Journal of Mathematical Sciences and Informatics. 15. M.Rahmani, On Some properties of c-frames, J.Math.Res.Appl, Vol. 37(4), (2017),466-476. 16. M.Rahmani, Sum of c-frames, c-Riesz Bases and orthonormal mapping, U.P.B.Sci. Bull, Series A,Vol.77(3), (2015),3-14 17. A.Rahmani, A.Najati, and Y.N.Deghan, Continuous frames in Hilbert spaces, methods of Functional Analysis and Topology Vol. 12(2), (2006),170- 182. 18. E.C. Hilbert C ∗-modules, A Toolkit for Operator Algebraists: University of Leeds, Cambridge University Press, (1995). CONTROLLED CONTINUOUS G-FRAMES IN HILBERT C ∗-MODULES 11 19. S.T.Ali, J.P.Antoine and J.P.Gazeau, continuous frames in Hilbert spaces, Annals of physics 222 (1993), 1-37 20. Mehdi Rashidi Kouchi1 and Akbar Nazari2, Hindawi Publishing Corpora- tion Abstract and Applied Analysis Volume 2011, Article ID 361595, 20 pages. 1Department of Mathematics, University of Ibn Tofail, B.P. 133, Kenitra, Morocco E-mail address: hlabrigui75@gmail; [email protected];[email protected]
1903.05342
1
1903
2019-03-13T07:13:32
Quantization of Yang--Mills metrics on holomorphic vector bundles
[ "math.OA" ]
We investigate quantization properties of Hermitian metrics on holomorphic vector bundles over homogeneous compact K\"ahler manifolds. This allows us to study operators on Hilbert function spaces using vector bundles in a new way. We show that Yang--Mills metrics can be quantized in a strong sense and for equivariant vector bundles we deduce a strong stability property which supersedes Gieseker-stability. We obtain interesting examples of generalized notions of contractive, isometric, and subnormal operator tuples which have geometric interpretations related to holomorphic vector bundles over coadjoint orbits.
math.OA
math
Quantization of Yang -- Mills metrics on holomorphic vector bundles Andreas Andersson Email: [email protected] Chalmers University of Technology, Mathematical Sciences, Maskingrand 2, 412 58 Gothenburg, Sweden Mathematics Subject Classification 2010 -- Primary: 47L80; Secondary: 47B10, 32L10 March 14, 2019 Abstract We investigate quantization properties of Hermitian metrics on holomorphic vector bundles over homogeneous compact Kahler manifolds. This allows us to study operators on Hilbert function spaces using vector bundles in a new way. We show that Yang -- Mills metrics can be quantized in a strong sense and for equivariant vector bundles we deduce a strong stability property which supersedes Gieseker-stability. We obtain interesting examples of generalized notions of contractive, isometric, and subnormal operator tuples which have geometric interpretations related to holomorphic vector bundles over coadjoint orbits. Contents 1 3 Introduction 3 1.1 Quantization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6 1.2 Yang -- Mills metrics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7 1.3 Balanced metrics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8 1.4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9 1.5 The nature of ς(PE) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.6 Guo-stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9 1.7 Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10 Into Hardy space 2 Multivariable operator theory of G/K 2.2 2.3 2.4 11 2.1 Preliminaries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 2.1.1 The first-row algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 2.1.2 Haar orthogonality relations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13 Subnormality and spherical expansivity . . . . . . . . . . . . . . . . . . . . . . . . . . 13 Schatten-class membership . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15 (d + 1)-isometries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15 2.4.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15 2.4.2 Result . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16 SOT-Toeplitz operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 2.5.1 The L∞ Toeplitz algebra C∗(ς(L∞(S))) . . . . . . . . . . . . . . . . . . . . . 17 2.5.2 . . . . . . . . . . . . . . . . 20 2.5.3 Vector-valued essential normality . . . . . . . . . . . . . . . . . . . . . . . . . 22 Projections onto vector-valued quotient modules 2.5 1 3 Cowen -- Douglas bundles of quotient modules 3.1 3.2 Algebraic aspects 3.3 Extension and boundary values Serre sheaf versus Cowen -- Douglas sheaf 23 Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23 3.1.1 The reference Hermitian line bundle . . . . . . . . . . . . . . . . . . . . . . . 23 3.1.2 Higher-rank Cowen -- Douglas bundles . . . . . . . . . . . . . . . . . . . . . . . 26 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 3.2.1 Graded A-modules and coherent sheaves . . . . . . . . . . . . . . . . . . . . . 27 . . . . . . . . . . . . . . . . . . . . . 29 3.2.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31 3.3.1 Abel convergence: ς versus ςB . . . . . . . . . . . . . . . . . . . . . . . . . . . 31 3.3.2 Extension of vector bundles . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34 3.3.3 The reproducing kernel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35 3.4 The Cowen -- Douglas projection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36 3.4.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . 37 3.4.2 The boundary limit of ΠE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38 3.4.3 Geometric interpretation of Em . . . . . . . . . . . . . . . . . . . . . . . . . . 39 3.5 Nullstellensatz . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40 Some alternating projections 4 The Toeplitz part of a superharmonic projection 4.2 4.1.1 Nonexisting lifts 4.1.2 Into Hardy space 4.2.1 4.2.2 4.2.3 42 4.1 Lifts of projections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44 dim Ran ς (m)(P E) versus χ(E(m)) . . . . . . . . . . . . . . . . . . . . . . . . . 45 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45 Subnormality with algebraic relations . . . . . . . . . . . . . . . . . . . . . . 46 Similarity to a spherical isometry . . . . . . . . . . . . . . . . . . . . . . . . . 47 Identification of KE N . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48 4.3 Hidden Szego expansion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49 . . . . . . . . . . . . . . . . . . . . . . . . . . 49 Interpretation of the hidden Szego expansion . . . . . . . . . . . . . . . . . . 50 ς E(m)(A−1 E,m) expansion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51 4.3.1 Geometric meaning of [S∗ 4.3.2 4.3.3 E, SE] 5 Lifts of Yang -- Mills metrics 53 5.1 Balanced metrics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53 Superharmonic lifts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55 5.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57 5.3 Subharmonic lifts 5.4 Direct sums . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58 5.5 The nature of ς(PE) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58 5.6 Gieseker-stability and superharmonic lifts . . . . . . . . . . . . . . . . . . . . . . . . . 60 5.7 From balance to Gieseker-stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60 6 Equivariant vector bundles 62 6.1 Quotient modules with equivariant Cowen -- Douglas sheaves . . . . . . . . . . . . . . . 62 Equivariant Cowen -- Douglas metrics are balanced . . . . . . . . . . . . . . . . 62 6.1.1 E 6.1.2 l,m is unital . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63 6.1.3 The (d + 1)-isometry SE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64 6.1.4 Characterization of equivariance . . . . . . . . . . . . . . . . . . . . . . . . . 66 6.2 Guo-stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67 References 67 2 1 Introduction 1.1 Quantization It is well established that every smooth projective variety M ⊂ CPn−1 can be quantized in the sense that there is a sequence H• = (Hm)m∈N0 of finite-dimensional Hilbert spaces such that the matrix algebra B(Hm) of all bounded operators on Hm becomes arbitrarily close, as a C∗-algebra, to the C∗-algebra C0(M) of continuous functions on M when m goes to infinity. As a vector space one has Hm = H 0(M;Lm) := the space of global holomorphic sections of the mth power of a positive line bundle L over M. If we fix a Kahler form ω on M in the class c1(L) then the choice of inner product on Hm is a choice of Hermitian metric hm on L and the relation between hm and ω is not arbitrary as m gets large. If ω has constant scalar curvature then the quantization is slightly more well-behaved than in general. The strongest possible quantization (a "regular" quantization) can obtained only when M is a coadjoint orbit G/K under some Lie group G and ω is G-invariant. Here in precise terms a regular quantization means a quantization where two kinds of natural positive maps (Toeplitz and covariant symbol maps) are both unital. In [An6] we took an operator-theoretic and operator-algebraic approach to quantization. The physical motivation for this is outlined in [An4, An5]. In this paper we shall study the quantization of Hermitian metrics on holomorphic vector bundles over a coadjoint orbit M = G/K ⊂ CPn−1. The case M = CPn−1 is already interesting. Inspired by noncommutative geometry, and in particular [Hawk1, Hawk2], a quantization of a vector bundle E will be a sequence of modules B(Hm, Em) over the matrix algebras B(Hm) such that when m is getting large B(Hm, Em) becomes arbitrarily close to the C0(M)-module Γ0(M;E) of global continuous sections of E. Actually there is more to it. When we quantized C0(M) we had inner products whose associated norms approximated the norm on C0(M). A Hermitian metric hE on the vector bundle E is the same thing as a C0(M)-valued inner product on Γ0(M;E); when the C0(M)-module Γ0(M;E) is endowed with such a C0(M)-valued inner product it is called a Hilbert C∗-module and we denote it by Γ0(M;E, hE). Therefore we would like to have a Hilbert C∗-module structure on B(Hm, Em) (or, what is the same, an inner product on the Hilbert space Em) which approximates Γ0(M;E, hE) in the limit m → ∞. Observe that we have fixed the inner product on Hm and therefore the C∗-structure on B(Hm) and the Hilbert C∗-module structure on B(Hm, Em), so that we are looking for the possibility to quantize the Hermitian metric hE with respect to the given quantization H• of the manifold M. In [An6] we used a particular choice of quantization H• of M which gives C0(M) as a kind of inductive limit of the matrix algebras B(Hm) provided by unital completely positive maps ιm,m+1 : B(Hm) → B(Hm+1). On the Hilbert space HN := Lm∈N0 of generalized Toeplitz operators with symbol in C0(M), and one has a short exact sequence of C∗-algebras acts a C∗-algebra T (0) H (1.1) 0 → Mm∈N0 B(Hm) → T (0) H ς→ C0(M) → 0 which is split by a positive linear Toeplitz-type map ς : C0(M) → T (0) H . The map ς is the adjoint of the Toeplitz map and called the covariant Berezin symbol map. The operators in T (0) H preserve the N0- grading on HN and its elements can therefore be regarded as sequences (Am)m∈N0 with Am ∈ B(Hm). The ideal Lm∈N0 B(Hm) consists of those compact operators on HN which preserve the grading. This is the quantization of the manifold which we will fix, and we will describe the quantization of vector bundles, as defined above, as follows. For any N ∈ N we can extend ς to a map on the algebra C0(M)⊗MN (C) of N ×N -matrices with entries in C0(M) by applying ς to each entry; we use the same notation ς for all N . If P E is an idempotent in the algebra C0(M)⊗ MN (C) for some N , in which case we say that P E is an idempotent over C0(M), then P E defines a continuous vector bundle E over M. Indeed, since P E is continuous the rank of P E(x), say r, is the same for all x ∈ M, and we can view P E as a topological embedding of M in the Grassmannian manifold Grr(CN ) of rank-r projections acting on CN . We then obtain a vector bundle E over M by pulling back the universal vector bundle over Grr(CN ) via P E. Another way of viewing is that P E defines a projective C0(M)-module Γ0(M;E, P E) := P E(C0(M) ⊗ CN ) 3 which by Swan's theorem determines a C0 vector bundle E uniquely up to isomorphism. Having presented a vector bundle E using an idempotent P E, a quantization of E is an idempotent PE over T (0) H which lifts P E in the sense that ς(PE) = P E. Since ς splits the Toeplitz short exact sequence (1.1), this means that there is a compact operator KE such that PE = ς(P E) + KE. PE,m(B(Hm) ⊗ CN ) = B(Hm, Em) If we let PE,m be the component of PE in B(Hm) ⊂ B(HN) then we can indeed define vector spaces Em via the identification (1.2) of left B(Hm)-modules. By a purely algebraic argument we can show that the C0(M)-module is obtained as the set-theoretical inductive limit of the B(Hm)-modules B(Hm, Em). Such a lift PE always exists (see §4.1). Since we have fixed the C∗-structure on B(Hm), the quantization is really with respect to H•. If we take P E to be a projection (meaning a selfadjoint idempotent) then ς(P E) is a positive operator, PE must be a projection and Em becomes a Hilbert space under the canonical inner product obtained from the identification (1.2). And an inner product on the vector space Em is the same datum as a B(Hm)-valued inner product on B(Hm, Em). The standard C0(M)-valued inner product on C0(M) ⊗ CN gives a C0(M)-valued inner product on the module Γ0(M;E, P E) (i.e. a Hermitian metric on E) when P E is a projection. Moreover, up to isomorphism every Hermitian vector bundle E is obtained like this for some projection P E. Therefore, we shall typically only consider a vector bundle E up to smooth or C0 isomorphism and refer to a projection P E with Γ0(M;E) ∼= P E(C0(M) ⊗ CN ) as a metric on E. However, the B(Hm)-valued inner product on B(Hm, Em) may not approximate the given Her- mitian metric P E on E. We would like to have some better compatibility between the Hilbert C∗-modules B(Hm, Em) for different m's or, what is the same, a better behavior of the sequence E• with respect to the sequence H•. That ς(P E) itself is a projection, so that we could take PE to be just ς(P E), happens only if all Chern classes of E vanish. So this is a too strong requirement. As mentioned, we shall in this paper assume that M is a homogeneous space G/K under some Lie group G. This ensures that the Toeplitz map ς is unital. The Toeplitz operators are then precisely the fixed points under an explicit unital completely positive map Ψ : B(HN) → B(HN). Therefore the Toeplitz the theory of noncommutative operators will also be referred to as Ψ-harmonic operators (cf. Poisson boundaries [Iz1, INT1]). So we have Ψ(ς(P E)) = ς(P E), and we said that it is too strong to assume that the lift PE is fixed by Ψ. As a second best thing we could look for a lift PE which is superharmonic under Ψ, Ψ(PE) ≤ PE. (1.3) We shall investigate the geometric meaning of such a lift. Let us first look at its meaning in operator theory. This is in fact very easy to describe. The Hilbert space HN is a reproducing kernel Hilbert space of analytic functions on a complex-analytic submanifold B of the unit ball Bn in Cn. More H 0(M;Lm), so that precisely, if V ⊂ Cn denotes the spectrum of the graded algebra A := Lm∈N0 (V \ {0})/C× = M, then B := V ∩ Bn. The reproducing kernel of HN has the complete Nevanlinna -- Pick property. Let us give two important characterizations of this property. The first is that HN can be identified with a subspace of the Drury -- Arveson space H 2 n, whose tuple of multiplication operators by the coordinate functions z1, . . . , zn on Bn is the universal model for pure row contractions. This means that the multiplication tuple S = (S1, . . . , Sn) on HN, defined by (Sαψ)(w) := wαψ(w), ∀ψ ∈ HN, w ∈ B, 4 α=1 SαS∗ satisfies the row contraction property Pn α = 1 − p0, where p0 is the projection onto the constant functions, and is pure in the sense that SOT− limp→∞ Φp(1) = 0, where Φ is the con- tractive completely positive map Φ(X) := Pn α on B(HN). When restricted to the algebra Qm B(Hm) of grading-preserving operators the map Φ is the adjoint of Ψ with respect to the state Pm φm, where φm is the tracial state on B(HN). Explicitly, α=1 SαXS∗ Ψ(X) = T ∗ αXTα, n Xα=1 ∀X ∈ B(HN) with T = (T1, . . . , Tn) acting just as S but weighted to ensure Ψ(1) = 1. The operator tuples S and T have the same set of invariant graded subspaces and the same set of coinvariant graded subspaces. The Ψ-superharmonicity condition (1.3) means precisely that the range EN of PE as an operator on HN ⊗ CN is invariant under the operators S∗ n. This leads us to the second characteristic property of the complete Nevanlinna -- Pick space HN. Namely, that (1.3) is equivalent to the existence of a matrix ΘE of analytic multipliers of HN such that 1 , . . . , S∗ PE = 1 − MΘE M ∗ ΘE , (1.4) where MΘE is the operator of multiplication by ΘE. Note that MΘE must then be a partial isometry. We know from [GRS2, Thm. 4.3] or [BhSa1, Thm. 6.1] that ΘE has an L∞ extension to the boundary S of Bsuch that ΘE(ζ) is a partial isometry for almost every ζ ∈ S. Therefore (1.4) is regarded as a multivariable analogue of the Beurling representation for model spaces on the unit disk D, although in the one-variable setting the graded situation is trivial. A multivariable Beurling representation holds also for not necessarily grading-preserving projections PE ∈ B(HN ⊗ CN ) with (1.3) but we will focus on the graded ones here. We shall see that (1.3) guarantees that PE has entries in a von Neumann algebra of Toeplitz operators with L∞ symbols. We will also show that if PE has entries in the Toeplitz C∗-algebra T (0) H then the real-analytic matrix-valued function ςB(PE)(v) := 1 − ΘE(v)ΘE(v)∗, ∀v ∈ B has a continuous extension to a projection-valued function on S, which is equivariant under the action on S by the circle group U(1) and hence descends to the projective variety M = S/ U(1). We shall see that the latter is precisely the symbol ς(PE) of the Ψ-superharmonic projection PE. Thus, starting from a Ψ-superharmonic projection PE over T (0) H we obtain in a canonical way a continuous vector bundle over M, namely the vector bundle E defined by the projection ς(PE) over C0(M). The compression of the shift on HN ⊗ CN to the range of PE is again a pure row contraction SE. The rank of the vector bundle E is precisely the so-called Arveson curvature of SE which is studied e.g. In the present paper we shall investigate further the geometric in [Arv7a, Arv7b, GRS1, GRS2]. properties of E and their relation to the operator tuple SE. There is a more well-studied way of associating a vector bundle to an operator tuple such as SE, provided that one can show that SE satisfies certain conditions. Cowen and Douglas invented an approach to classify operator tuples with uncountably many eigenvalues [CoDo1, CoDo2]. If Ω is a subset of Cn and r ≥ 1 is an integer, a tuple W = (W1, . . . , Wn) of commuting operators on a Hilbert space H is said to belong to the Cowen -- Douglas class Br(Ω) if (i) Ran(W − w1) is a closed subspace of H for all w ∈ Ω, (ii) dim Ker(W − w1) = r for all w ∈ Ω, where Ker(W − w1) :=Tn (iii) ∨w∈Ω Ker(W − w1) = H. If W is in class Br(Ω) then the family of Hilbert subspaces Ker(W − w1) parameterized by w ∈ Ω can be given the structure of form an Hermitian holomorphic vector bundle over Ω. It is not hard to show that when S is the shift on HN as before then the backward shift S∗ is in class B1(B), with the domain B := Bn ∩ SpecA mentioned above. The associated holomorphic line bundle over B is isomorphic to the trivial one but its Hermitian metric is still quite interesting. α=1 Ker(Wα − wα1), and 5 In view of our quantization problem it is natural to look at graded subspaces EN of HN ⊗ E0 for some finite-dimensional Hilbert space E0. Compressing the shift S on HN ⊗ E0 down to the subspace EN we obtain a commutative operator tuple SE which we also refer to as the shift on EN. When the projection PE onto EN is Ψ-superharmonic, which happens iff EN is invariant under S∗, we call EN a graded quotient module. We will prove: Theorem 1.1 (Theorem 3.23). Let EN ⊂ HN ⊗ E0 be a graded quotient module and suppose that the projection PE onto EN has entries in the Toeplitz C∗-algebra T (0) H . Then the operator tuple SE is in class BrE (B \ {0}), where rE = lim m→∞ dim Em dim Hm is the Arveson curvature of the pure row contraction SE. Typically ECD does not extend to a vector bundle over all of B, i.e. it is necessary to remove the E of HN⊗ CN is in the Cowen -- Douglas origin. By [DKKS2, Thm. 3.3], a quotient module EN = Ker Θ∗ class B∗ r (B) for some r if MΘE has a left inverse (see also [DFS1, Cor. 4.4]). From Theorem 3.23 one can then deduce obstructions to the existence of a left inverse MΘE : Such a left inverse cannot exist if the "Serre sheaf" of the graded A-module EN (see §3.2.1) is not locally free on all of B. The left-invertibility of MΘE is a kind of corona condition (see [Doug1]). 1.2 Yang -- Mills metrics We just discussed the vector bundles associated with a Ψ-superharmonic projection PE over T (0) H . If we go back to our starting point, with an arbitrary projection P E over C0(M), it turns out that there does not exist a Ψ-superharmonic lift of P E. But we shall prove: Theorem 1.2 (Lemma 5.4, Theorem 5.5). Let P E be a projection over C∞(M) defining a Hermitian Yang -- Mills vector bundle E over M. Then the Ψ-superharmonic projection Ran ς(P E) is a lift of P E: ς(Ran ς(P E)) = P E. Moreover, the Cowen -- Douglas sheaf ECD of HE to the pullback EB\{0} of E; as Hermitian holomorphic vector bundles we have N := Ran ς(P E) is analytically isomorphic over B\{0} ECD = OCD ⊗ EB\{0} (1.5) where EB\{0} is endowed with the Hermitian metric given by pullback of P E to B \ {0}. The Yang -- Mills equation under consideration is trω ΘE = µ(E)1E, where ω is the Kahler form in class c1(L) associated with the G-invariant volume form, ΘE is the Chern curvature of P E and the given holomorphic structure on E, and trω : A2(M)⊗ End Γ∞(M;E) → End Γ∞(M;E) is the operator of taking trace against ω (with A2(M) denoting the space of differential 2-forms on M). The Yang -- Mills condition depends on the Kahler metric ω, which for our coadjoint orbit M = G/K is encoded in the quantization H•. The difference between the Toeplitz operator ς(P E) and its range projection PE is a measure on how well P E can be quantized with respect to the chosen quantization H• of the manifold. We have said that ς (m)(P E) cannot itself be expected to be a projection for all m unless the vector bundle E defined by P E has trivial Chern character. The next best thing would be that ς (m)(P E) is a scalar multiple of its range projection PE,m for each m. This holds for very special metrics: Proposition 1.3 (Proposition 6.1). Suppose that P E is a projection over C∞(M) defining an ir- reducible G-equivariant Hermitian vector bundle E over M = G/K. Then for all m ≫ 0 we have (1.6) ς (m)(P E) = cE,mPE,m 6 with the scalar cE,m := nm rankE χ(E(m)) where χ(E(m)) is the Euler characteristic of the vector bundle E(m) := Lm ⊗ E. Recall that for m large enough, χ(E(m)) is just the dimension of the vector space H 0(M;E(m)) of global holomorphic sections of E(m). 1.3 Balanced metrics For more general vector bundles E, which are not G-equivariant, there does not exist a metric P E satisfying (1.7) for each m ≫ 0. But suppose that we have a sequence (BE m)m≫0 of projections over C∞(M) which all define vector bundles smoothly isomorphic to E and such that ς (m)(BE m) = cE,mPE,m, (1.7) where PE,m is a projection in B(Hm)⊗MN (C) for some N . If we want the BE m's to give a quantization of E then we need to require that the BE m's for different m are related in some way. One way would be to require the projection PE :=Pm PE,m to belong to the C∗-algebra T (0) H ⊗ MN (C), so that the symbol ς(PE) is continuous. Let us instead assume that E admits a holomorphic structure. Then the vector space EN := Mm∈N0 H 0(M;E(m)) H 0(M;Lm). Up to finite-dimensional vector is a graded module over the graded algebra A :=Lm∈N0 spaces we may assume that EN for some N is the quotient of A ⊗ CN by some graded submodule IN. Then a natural condition on the BE m's is to require that PE is the projection of HN ⊗ CN onto the orthogonal complement EN of IN. Indeed, the graded A-module structure on EN defined by the compressed shift SE is isomorphic to that on EN. The subspace EN ⊂ HN ⊗ CN is invariant under S∗, so PE is Ψ-superharmonic. The Toeplitz map ς (m) has an explicit expression in terms of ω and the Fubini -- Study metric FS(Hm) of the line bundle Lm. To find the meaning of (1.7) we will make use of frame theory for Hilbert C∗-modules (see [FrLa3] for background). A frame for a Hilbert C∗-module will be referred to as a C∗-frame in order to distinguish it from the usual frames for Hilbert spaces, which we will also need (see [Chri1, HaLa1] for background on these). As will be discussed in more detail in §5.1, the condition (1.7) says that there is a Parseval C∗-frame for the Hilbert C0(M)-module Γ0(M;E(m), FS(Hm) ⊗ BE m) which is at the same time an orthonormal basis for the vector space H 0(M;E(m)) endowed with the L2-inner product of ω and the metric FS(Hm)⊗ BE m on E(m). Briefly, (1.7) says that FS(Hm) ⊗ BE m is a balanced metric on E(m) in the sense of [Wang1]. One can m as saying that the Hilbert C∗-struture on Γ0(M;E(m), FS(Hm) ⊗ view the balance of FS(Hm) ⊗ BE BE m) (i.e. the metric FS(Hm) ⊗ BE m) is completely encoded in the finite-dimensional Hilbert space H 0(M, ω;E(m), FS(Hm) ⊗ BE m). Balanced metrics are related to Yang -- Mills metrics via Wang's theorem, here reformulated using the Toeplitz maps: Lemma 1.4 ([Wang2]). Let P E be a metric on a holomorphic vector bundle E over M. Then P E is Yang -- Mills if and only if there exists a sequence (BE m)m≫0 of metrics on E such that the metric FS(Hm) ⊗ BE m on E(m) is balanced and in the topology of C∞(M). lim m→∞ BE m = P E The connection between the Yang -- Mills condition and the existence of a Ψ-superharmonic lift stated in Theorem 1.2 can now be better understood: If P E is a Yang -- Mills then P E is the limit of some projections BE m satisfying (1.7). We shall see that PE,m for m ≫ 0 coincides with the range projection of ς (m)(P E) and therefore the convergences limm→∞ BE m = 1 give that ς(P E) differs from its range projection modulo compacts. m = P E and limm→∞ cE 7 1.4 Into Hardy space Given a metric P E on a holomorphic vector bundle E(m) we can also look at the Hardy space H 0(S, ω; P E) of the pullback of P E to the circle bundle S = ∂B. It is natural to ask for a relation between H 0(S, ω; P E) and the quotient module EN := Ran ς(P E). First observe that the Hardy space H 0(S, ω; P E) is invariant under the tuple of operator on L2(S, ω; P E) acting by multiplication with the coordinate functions Z = (Z1, . . . , Zn). The multiplication tuple VE on H 0(S, ω; P E) is therefore subnormal. In fact, VE is a spherical isometry in the sense that n V ∗ EVE := V ∗ E,αVE,α = 1, Xα=1 and this algebraic relation directly implies subnormality by [Atha3, Prop. 2]. In contrast, the shift SE on EN is not subnormal, i.e. it does not have a normal extension. Even if we take the spherical isometry T on HN⊗ CN and consider its compression TE to the coinvariant subspace EN, the operator T ∗ ETE is typically not the identity on EN. But while SE belongs to something that one could call a multivariable analogue of the class C·0 of contractions, TE is a contraction of class C1· if we assume that the projection PE has entries in the Toeplitz algebra T (0) H . And for contractions TE of class C1· there is a canonical way of turning TE into an isometry by applying a similarity transformation, namely the Fredholm operator AE := lim m→∞ Xk=m T ∗ E,kTE,k, where TE,k := TE,k1 ··· TE,km. The inverse of A−1 subspace, and we may ignore this subspace for present purposes. The tuple A1/2 spherical isometry and in particular subnormal. Theorem 1.5 (Theorem 4.12). The operator tuple A1/2 the multiplication tuple VE on H 0(S, ω; P E). E of AE is well-defined outside a finite-dimensional is a on EN is unitarily equivalent to E TEA−1/2 E E TEA−1/2 E Another way of viewing this is to say that the positive invertible operator AE can be used to change the inner product on EN to that of H 0(S, ω; P E). Now observe that AE is just the restriction to EN of the Toeplitz operator ς(P E). Let us discuss the geometric interpretation of ς(P E). We have already mentioned that ς (m)(P E) equals PE,m only if there is a Parseval frame for the Hilbert C0(M)-module Γ0(M;E(m), FS(Hm) ⊗ P E) which is also a Parseval frame for the Hilbert space H 0(M, ω;E(m), FS(Hm)⊗ P E). In fact ς (m)(P E) can be identified with the frame operator of a Parseval frame for the Hilbert C0(M)-module Γ0(M;E(m), FS(Hm)⊗ P E) regarded as a frame for the Hilbert space H 0(M;E(m), FS(Hm) ⊗ P E). For this reason it is interesting to note that the diagonal of the Szego kernel (aka Bergman kernel), studied e.g. in [Catl1, MaMa3], is in the present context the frame operator ΣE(m) of a Parseval frame for the Hilbert space H 0(M;E(m), FS(Hm) ⊗ P E) regarded as a frame for the Hilbert C0(M)-module Γ0(M;E(m), FS(Hm)⊗ P E). Thus finding out how ς (m)(P E) differs from the identity operator on Em ∼= H 0(M, ω;E(m), FS(Hm) ⊗ P E) is analogous to finding the error between ΣE(m) and the identity endomorphism. The latter is given in the Szego expansion [Wang2, Thm. 5.2] χ(E(m)) vol(M,L) rankE ΣE(m) = md1E + md−1(trω ΘE − sω/2)1E + O(md−2), where sω is the scalar curvature of the Kahler metric ω. For ς (m)(P E) the expansion would be one of operators on the Hilbert space Em. Since AE is the limit of Ψp(PE) as p goes to infinity, a first approximation to the compact operator PE − AE is given by n n Xα=1 Xα=1 (id −Ψ)(PE) = = 8 [PE, Tα]∗[PE, Tα] [T ∗ E,α, TE,α] + n [T ∗ α, Tα], Xα=1 E,α, TE,α] and α, Tα] are operator analogues of the mean curvature trω ΘE and the scalar curvature sω which may be viewed as an operator "second fundamental form", while Pn Pn respectively. For PE − AE itself we have: Theorem 1.6 (Theorem 4.15). In the setting of Theorem 1.5 there is an injective completely positive map ς (m) VE α=1[T ∗ α=1[T ∗ the space of operators on Em into End Γ0(M;E) such that ς (m) VE (PE,m − AE,m) = m−1(trω ΘE − µ(E))1E + O(m−2). Here the curvature ΘE is that of the Chern connection of P E and the holomorphic structure on E coming from the graded A-module underlying EN, which by Theorem 1.5 is the same as the holomorphic structure on E that we started with if we take P E to be real-analytic. 1.5 The nature of ς(PE) So for a projection P E which does not define a Yang -- Mills metric, what is the meaning of the symbol ς(PE) of the superharmonic projection PE := Ran ς(P E)? The Uhlenbeck -- Yau theorem says that a holomorphic vector bundle E admits a Yang -- Mills metric if and only if E is slope-stable [Koba1, UhYa1]. Any holomorphic vector bundle E has a filtration by slope-semistable subsheaves, and each of these slope-semistable subsheaves has a filtration by slope-stable subsheaves. Taking successive quotients of the members of these filtrations and summing up one obtains a torsionfree sheaf Gr(E) which is a direct sum of slope-stable vector bundles (see [Jaco2, §2.1] for details). Each of these slope-stable summand of Gr(E) admits a Yang -- Mills metric, and the direct sum of these metrics will be referred to as the Yang -- Mills metric on Gr(E). We prove: Theorem 1.7 (Theorem 5.11). Let E be a holomorphic vector bundle over M and suppose that Gr(E) H 0(M;E(m)) in is locally free. Let EN be the quotient module obtained by completing EN :=Lm∈N0 some embedding into HN ⊗ CN , and let PE be the projection onto EN. Then the projection ς(PE) defines Gr(E) as smooth vector bundle, and ς(PE) is the Yang -- Mills metric on Gr(E). We conjecture that in general, when Gr(E) is merely torsionfree, ς(PE) still gives a singular Yang -- Mills metric on Gr(E). To prove that one would need a generalization of Lemma 1.4 to singular Yang -- Mills metrics and balanced metrics on torsionfree sheaves. When Gr(E) is locally free the assumptions of Theorem 1.6 thus holds. The term m−1(trω ΘE − µ(E))1E in the statement of Theorem 1.6 need not be zero even though ς(PE) is a Yang -- Mills on Gr(E), because ΘE is the Chern connection for ς(PE) and the holomorphic structure on E, and not the holomorphic structure on Gr(E). 1.6 Guo-stability A slightly weaker notion of stability than slope-stability is Gieseker-stability. Also this notion of stability has been characterized by Wang using balanced metrics [Wang1]. Namely, a holomorphic vector bundle E is Gieseker-stable if and only if for all m ≫ 0 there exists a balanced metric on E(m). The distinction between Gieseker- and slope-stability is thus the convergence of the balanced metrics. We will give an operator-theoretic proof of half of Wang's theorem (see Theorem 5.17) and we will consider the failure of the convergence of the balanced metrics in Proposition 5.15. Recall that a holomorphic vector bundle E over M is called Gieseker-semistable if for all proper analytic quotients E → F → 0 one has χ(F (m)) rankF ≥ χ(E(m)) rankE , ∀m ≫ 0. (1.8) If we observe that rankF / rankE = liml→∞ χ(F (l))/χ(E(l)) then we can rewrite (1.8) as χ(F (m)) χ(E(m)) ≥ lim l→∞ χ(F (l)) χ(E(l)) , ∀m ≫ 0. 9 It is known [ACKi1, Remark 2.2] that (1.8) is equivalent to χ(F (m)) χ(E(m)) ≥ χ(F (l)) χ(E(l)) , ∀l ≫ m ≫ 0. (1.9) We say that a holomorphic vector bundle E is Guo-semistable for each proper analytic quotient sheaf F the condition (1.9) holds for all l ≥ 0 (not just for l sufficiently large compared to m): χ(F (m)) χ(E(m)) ≥ χ(F (l)) χ(E(l)) , ∀l ≥ m ≫ 0. (1.10) As usual we replace semistable by stable when strict inequality holds in (1.10) for all F . We use this terminology because it was shown in [Guo3, Prop. 2.3] that the trivial line bundle on CPn−1 is Guo-stable in this sense. We shall prove: Theorem 1.8 (Theorem 6.4). Let E be an irreducible G-equivariant vector bundle over the coadjoint orbit M = G/K. Then E is Guo-stable. We thereby see that the operator-theoretic result as it is stated in [Guo3, Prop. 2.3] extends to a much wider range of reproducing kernel Hilbert spaces. In the end of the paper we will discuss some natural questions and open problems building on this work that would be interesting to investigate in the future. Acknowledgment. I thank Robert Berman, Daniel Persson, Magnus Goffeng, Erlend Fornaess- Wold, Tuyen Troung, Hakan Samuelsson Kalm, Mats Andersson, Bo Berndtsson, Martin Sera, Ramiz Reza, Dennis Eriksson, and Ulrik Enstad for discussions on the topic of the paper. This work was initiated when I was a postdoc at the University of Oslo, supported by ERC (grant 307663-NCGQG). 1.7 Notation If E is a holomorphic vector bundle on a smooth projective variety M then we denote by H 0(M;E) the vector space of global holomorphic sections of E. If P E(m) is a Hermitian metric on E(m) then we denote by H 0(M, ω;E(m), P E(m)), or just H 0(ω, P E(m)), the vector space H 0(M;E) endowed with the inner product hφψiω,P E(m) := χ(E(m)) rankE ω((φψ)P E(m) ), ∀φ, ψ ∈ Γ0(M;E(m)). (1.11) The Euler characteristic of a coherent analytic sheaf E over M is denoted by χ(E) and defined as the integer d χ(E) := (−1)p dim H p(M;E), Xp=0 here H p(M;E) is the pth sheaf cohomology group of the OM-module E. If E is a holomorphic vec- tor bundle then χ(E(m)) depends only on the isomorphism class of the topological vector bundle underlying E. Indeed, the Hirzebruch -- Riemann -- Roch theorem says that χ(E) is the pairing of the fundamental class of M with the Chern character of E wedged with the Todd class of the tangent bundle of M. We can take the latter formula as the definition of χ(E) for an arbitrary smooth vector bundle E which need not admit any holomorphic structure. With L = OM(m) a fixed very ample line bundle on M, we set E(m) := Lm ⊗ E for each m ∈ N0. The Hilbert polynomial of a smooth vector bundle E is the polynomial N0 ∋ m → χ(E(m)). While χ(E(m)) for a given m only depends on the topological structure of E(m), the choice of line bundle L is affected by the holomorphic structure of M since we want L to be very ample. 10 We will often write cE,m for the constant cE,m := nm rankE χ(E(m)) Let ω : C0(M) → C be a state, i.e. a functional with ω(1) = 1. where as always nm := dim Hm is the Hilbert polynomial of the trivial line bundle OM. If P E is a projection in C0(M) ⊗ MN (C) then we denote by L2(ω, P E) the completion of Γ0(M; P E) := P E(C0(M) ⊗ CN ) in the L2-inner product of P E and ω, hφψiL2(ω,P E ) := ω((φψ)Γ0(M;P E )), ∀φ, ψ ∈ Γ0(M; P E). χ(E(m)) rankE Remark 1.9 (ω on matrices). For an element B of L∞(M, ω) ⊗ B(H) for some separable Hilbert space H we can canonically define an operator ω(B) ∈ B(H) by defining ω(f ⊗ X) := ω(f )X on simple tensors f ⊗ X and extending ω by C-linearity. A choice of basis (or more generally a countable Parseval frame) for H gives a representation of B as an L∞(M, ω)-valued matrix (see [Bala1]), and the above definition just means that we apply ω to each entry in such a matrix. So even if the Parseval frame has many more element than the dimension of H, applying ω to each entry in the frame matrix of B gives the same as if we apply ω to each entry in a matrix of size dim H representing the action of B in an orthonormal basis for H. In particular, ω commutes with taking the trace over H, ω(TrH(B)) = TrH ω(B), ∀B ∈ L∞(M, ω) ⊗ B(H). Sometimes we write (ω ⊗ id)(B) for ω(B) for clarity when B is in L∞(M, ω)⊗B(H), with id standing for the identity map on B(H). 2 Multivariable operator theory of G/K 2.1 Preliminaries Let n ≥ 2 be an integer. In this section we recall from [An6, §6] how to associate a graded quotient HN of H 2 n to a compact matrix group G ⊂ U(n). 2.1.1 The first-row algebra Throughout the rest of the paper, G is a compact matrix group, i.e. G is a closed subgroup of the group U(H) of unitray transformations of some finite-dimensional Hilbert space H. The C∗-algebra C0(G) of continuous functions on G is generated by the matrix coefficients of a unitary matrix u ∈ B(H) ⊗ C0(G). Set n := dim(H) and fix an orthonormal basis e1, . . . , en of H so that H ∼= Cn, and let uα,β be the matrix coefficients of u in this basis. Definition 2.1. The first-row algebra of G is the C∗-algebra C0(S) generated by the first row Z1 := u1,1, . . . , Zn := u1,n. This defines the homogeneous space S. Since G is a Lie group, S is a smooth manifold. There is a Z-grading on C0(S) obtained by letting the Zα's have degree 1 while their adjoints are given degree −1. We write the decomposition into spectral subspaces for the corresponding U(1)-action as Definition 2.2. We define the homogeneous space G/K as the manifold corresponding to the C∗- subalgebra of fixed points in C0(S) for the U(1)-action: k·k C0(S)(k) . C0(S) =Mk∈Z C0(G/K) := C0(S)(0). 11 It is clear that C0(G/K) is generated by the n2 elements {Z ∗ αZβ}n α,β=1. Example 2.3. If G = U(n) is the whole unitary group then S is the unit sphere S2n−1 in Cn while G/K is the complex projective space CPn−1. Here we obtain K = U(1) × SU(n − 1). By the above example we see that, in general, S ⊂ S2n−1 and G/K ⊂ CPn−1. Since G is a Lie group, the space S, and hence also G/K = S/ U(1), is a smooth manifold and the action of G on C0(G/K) restricts to an action on the subalgebra C∞(G/K) of smooth functions. The space S is a smooth principal U(1)-bundle over the smooth manifold G/K. We denote by Irrep G the set of equivalence classes of irreducible unitray representations of G. We choose a representative Hλ for each λ ∈ Irrep G, i.e. Hλ is a (necessarily finite-dimensional) Hilbert space which carries an irreducible representation of G in class λ. An important part in the theory of compact groups is the Peter -- Weyl decomposition [Seg1, Cor. 9.14] C0(G) = Mλ∈Irrep G B(Hλ)∗ (2.1) of the vector space underlying the C∗-algebra C0(G), where B(Hλ)∗ denotes the dual of B(Hλ) with respect to the trace. Recall also that C0(G) has a unique state ω (the Haar state) which is invariant under the left and right translation action of G on C0(G). The completion L2(G) of C0(G) in the inner product defined by the Haar state decomposes into irreducibles as well, because of (2.1) and the G-invariance of ω. Lemma 2.4. [An6, Lemma 6.18] The first-row algebra C0(S) carries an ergodic action of G which contains every irreducible representation of G with multiplicity one. The unique G-invariant state on C0(S) is the restriction to C0(S) of the Haar state ω on C0(G). We write ω also for the G-invariant states on C0(S) and C0(G/K). We have already seen that the topological space G/K is contained in CPn−1. Now observe that the algebra A := Alg(Z1, . . . , Zn) M := Proj(A) ⊂ CPn−1. generated by Z1, . . . , Zn is the quotient of the polynomial algebra C[z1, . . . , zn] by some homogeneous ideal. Therefore A is the homogeneous coordinate ring of some projective variety That is, if L = OM(1) denotes the restriction to M of the hyperplane bundle on CPn−1 then A = Mm∈N0 H 0(M;Lm). The completion HN of A in the inner product of the symmetric Fock space H∨N is a graded quotient module. The elements of HN are analytic functions on the manifold where V := SpecA. B := Bn ∩ V, It is shown in [An6] that the C∗-algebra C0(M) of continuous functions on M is the inductive limit of the finite-dimensional matrix algebras B(Hm) as m goes to infinity, and that C0(M) and C0(G/K) coincide in such a way that the generators Z1, . . . , Zn of C0(S) become the homogeneous coordinates on M, Thus G/K is given the structure of a complex projective variety. M = G/K. The state ω : C0(M) → C defines a unique Kahler 2-form, also denoted by ω, in the cohomology class c1(L) via ω(f ) = 1 vol(M,L) M f (x)eω(x), ∀f ∈ C0(M), where vol(M,L) = ´M eω(x) = limm→∞ dim Hm/mdim M is the volume of (M,L). 12 2.1.2 Haar orthogonality relations Recall that the Haar orthogonality relations [Seg1, Thm. 9.7(iii)] say that if K is an irreducible representation of G and e1, . . . , enK is an orthonormal basis for K then where fα ∈ C0(G) is the function fα(a) := heαa · eαi. That is, in addition to the C0 Peter -- Weyl decomposition δαβ = heαeβiK = (dim K)ω(f ∗ αfβ), C0(G) = Mκ∈Irrep GB(Kκ) (as vector spaces) one has that the Haar state ω restricts to the normalized trace on B(Kκ), viz. the L2 Peter -- Weyl decomposition (as Hilbert spaces). The first-row algebra takes the form L2(G, ω) = Mκ∈Irrep GB(Kκ) C0(S) = Mκ∈Irrep G Kκ, where the irreducible representations Hm appear as special cases of the Kκ's, namely as the subspaces spanned by the products of m elemens of the generating set {Z1, . . . , Zn}. To obtain an arbitrary irreducible representation Kκ one has to use also products with elements of {Z ∗ For the special case of Kκ = Hm for m ∈ N0, the Haar orthogonality relations give the following: Lemma 2.5. Let e1, . . . , en be an orthonormal basis for H1. Then for all m ∈ N0 and all j, k ∈ F+ with j = m = k we have 1 , . . . , Z ∗ n}. n ω(ZjZ ∗ k) = hejpmeki dim Hm . Here F+ n denotes the set of multi-indices k = k1 ··· km of finite length k := m ∈ N0, and pm : H⊗m → Hm is the orthogonal projection. 2.2 Subnormality and spherical expansivity We are given our coadjoint orbit M = G/K ⊂ CPn−1 and the associated graded quotient HN of the symmetric Fock space H∨N. The shift on H∨N can be compressed to the subspace HN to give a tuple S = (S1, . . . , Sn) of mutually commuting operators. If e1, . . . , en denotes an orthonormal basis for H1, this means Sαψ = pm(eα ⊗ ψ), ∀ψ ∈ HN, α ∈ {1, . . . , n}, where pm : H⊗m → Hm is the orthogonal projection, and if the vectors in HN are identified with analytic functions on B, then each Sα becomes a multiplication operator: ∀ψ ∈ HN, w ∈ B. We shall also refer to HN as the Fock space and call S the shift on HN. (Sαψ)(w) := wαψ(w), It can be helpful to view the shift S as a quantization of the generating tuple Z = (Z1, . . . , Zn) of the C∗-algebra C0(S). Indeed, the graded algebra A ⊂ C0(S) generated by Z1, . . . , Zn is isomorphic to the graded algebra generated by S1, . . . , Sn. While the Zα's commute with their adjoints, [S∗ α, Sβ] is nonzero. One has n SS∗ := SαS∗ α = 1 − p0, where p0 ∈ B(HN) denotes the projection onto the 1-dimensional subspace H0 spanned by the constant functions. This is a remnant of the sphere condition Z ∗Z = ZZ ∗ = 1 satisfied by Z. So what about the operator S∗S :=Pn α=1 S∗ αSα? Xα=1 13 Lemma 2.6. The shifts S1, . . . , Sn satisfy n Xα=1 S∗ αSα = Xm∈N0 dim Hm+1 dim Hm pm. (2.2) Proof. Let ω be the restriction of the Haar state on C0(G) to the subalgebra C0(S) and let L2(S, ω) be the GNS Hilbert space of ω. Let H 0(S, ω) be the closure of A in L2(S, ω). The operators of multiplication by Z1, . . . , Zn on L2(S, ω) leave the subspace H 0(S, ω) invariant. Denote by T1, . . . , Tn the restrictions of the multiplication operators Z1, . . . , Zn to H 0(S, ω) and let P be the orthogonal projection of L2(S, ω) onto H 0(S, ω). Since the representation of C0(G) is a ∗-homomorphism, the multiplication operators satisfy Pn Xk=1 αZα =Pn Xk=1 α = 1 and hence T ∗ k Tk = P α=1 ZαZ ∗ α=1 Z ∗ (2.3) Z ∗ n n k Zk(cid:12)(cid:12)H0(S,ω) = 1. By Lemma 2.5 the inner product h··i on Fock space HN is a simple scaling of that of L2(S, ω), hφψiL2(S,ω) = 1 nmhφψi, ∀φ, ψ ∈ Am = span{Zk k ∈ F+ n , k = m}. It follows that the tuple T1, . . . , Tn is unitarily equivalent to the operator tuple T1, . . . , Tn on the Fock space HN defined by TαHm :=r nm nm+1 SαHm, ∀m ∈ N0, α, β ∈ {1, . . . , n}. if S1, . . . , Sn are the standard shifts on HN. From (2.3) we get Sα = S Tα, ∀α ∈ {1, . . . , n}. with S :=pPn k=1 S∗ kSk. The formula (2.2) then follows from (2.3) and the definition of T1, . . . , Tn. We thus see that the Haar orthogonality relations ensure that the tuple S = (S1, . . . , Sn) is a simple quasi-affine transform of the spherical isometry T = (T1, . . . , Tn) acting on the Hardy-type space H 0(S, ω). Recall that T being a spherical isometry means T ∗T := Pn αTα = 1, and that by [Atha3, Prop. 2] this is equivalent to saying that T is subnormal with normal extension having joint spectrum in S2n−1 (in the present case the normal spectrum is S ⊂ S2n−1). as an operator tuple acting on HN. In the rest of the paper we will not distinguish between T and T , so T will sometimes be regarded α=1 T ∗ Corollary 2.7. The operator tuple S = (S1, . . . , Sn) is a spherical expansion, i.e. Pn 1. Equivalently (since Pn α = 1 − p0), the operator α=1 SαS∗ n α=1 S∗ αSα ≥ [S∗, S] := [S∗ α, Sα] Xα=1 is positive. Question 2.8. We thus have [S∗, S] ≥ 0 when the quotient module HN comes from a coadjoint orbit. Does that hold for a general quotient module? One could also ask whether each of the operators Sα is hyponormal, i.e. whether [S∗ α, Sα] ≥ 0 holds for all α ∈ {1, . . . , n}. The hyponormality of each Sα appears a bit optimistic even for coadjoint orbits (although it is true for HN = H 2 n [Arv6c, §5]). 14 The scalar curvature of the Kahler metric associated with the state ω on the coadjoint orbit M = G/K is a constant function sω = sω1. For any quotient module HN of H 2 n, i.e. for any projective variety M ⊂ CPn−1 and any Kahler metric in the class c1(L) of the line bundle L = OM(1), the average scalar curvature sω := ω(sω) appears, by Lemma 2.6 and Hirzebruch -- Riemann -- Roch, as a contribution to the traces Tr([S∗, S]pm), φm([S∗, S]) = nm+1 − nm nm = m−1sω/2 + O(m−1). From this behavior of the traces one may expect that [S∗, S] is a quantization of the scalar curvature sω in some sense. 2.3 Schatten-class membership In [An6] it was shown that the commutators [S∗ α, Sβ] of the shift operators are compact. Here we give a simpler proof in our special case of coadjoint orbits, and we also obtain sharp estimates for membership in the Schatten classes Lp: Theorem 2.9. Let n ∈ N, let HN =Lm∈N0 Hm be the graded quotient module of the Drury -- Arveson n associated with a coadjoint orbit M = G/K ⊂ CPn−1, let d := dimC M, and let S1, . . . , Sn be space H 2 the compressions to HN of the shift operators on H 2 n. Then for all α, β ∈ {1, . . . , n} we have [S∗ α, Sβ] ∈ Lp ⇐⇒ p > d + 1. α, Sα] is positive. We shall α, Sβ] is in Lp for all α, β ∈ {1, . . . , n} if and only if the operator [S∗, S] is in Lp α=1[S∗ use the fact that [S∗ (see [Arv8, Thm. 4.3] for a proof). Proof. Recall that Corollary 2.7 says that the operator [S∗, S] :=Pn Since Tr(pm) = dim Hm, for m ≥ 1 we have from Lemma 2.6 that Tr([S∗, S]pm) = dim Hm+1 − dim Hm, and N0 ∋ m → dim Hm+1 − dim Hm is a polynomial of degree d − 1 (because N0 ∋ m → dim Hm ∈ N is a polynomial of degree d). So for the normalized trace φm(·) := Tr( · pm)/ Tr(pm) we have φm([S∗, S]) = O(m−1). Thus the largest (and only) eigenvalue of [S∗, S]pm grows as O(m−1). The eigenvalue of [S∗, S]ppm then grows as O(m−p). So we have Tr([S∗, S]p) ∼Pm∈N(dim Hm)/mp < ∞ if and only if Tr([S∗, S]p) ∼ Xm∈N 1 mp−d < ∞, which is the case if and only if p > d + 1. 2.4 (d + 1)-isometries 2.4.1 Background For the moment let S = (S1, . . . , Sn) be an arbitrary commutating tuple of opertors on a Hilbert αXSα. For each α=1 S∗ space H and consider the map Φ∗ : B(H) → B(H) defined by Φ∗(X) := Pn p ∈ N0 we define With the convention (cid:0)m p(cid:1) := 0 for p > m we have [GlRi1, Lemma 2.2] Bp(S) := (id−Φ∗)p(1). Φm ∗ (1) = Xp∈N0 (−1)p(cid:18)m p(cid:19)Bp(S). 15 (2.4) Definition 2.10 ([GlRi1, §2]). Let q ∈ N. A commuting operator tuple S = (S1, . . . , Sn) is a q-isometry if Bq(S) = 0. A q-isometry S is strict if Bq−1(S) 6= 0. Using the tautological relation Bp(S) = Bp−1(S) − Φ∗(Bp−1(S)) we can equivalently say that a tuple S is a q-isometry if the operator Bq−1(S) is a fixed point of Φ∗, Φ∗(Bq−1(S)) = Bq−1(S). By (2.4) (see also [HoMa1, Thm. 3.1]), S is a q-isometry if and only if there exists a degree-(q− 1) polynomial χS(m) = Cq−1mq−1 + Cq−2mq−2 + ··· + C0 with operator coefficients Cp ∈ B(H) such that A 1-isometry is what is usually called a spherical isometry. χS(m) = Φm ∗ (1), ∀m ∈ N0. 2.4.2 Result Theorem 2.11. The n-tuple S = (S1, . . . , Sn) on HN is a (d + 1)-isometry. Proof. Recall that by Lemma 2.6, √S∗S is the central operatorPm wmpm determined by the weight sequence (wm)m∈N0 given by wm :=r nm+1 nm . Recall also that nm is a polynomial in m ∈ N0 of degree d := dimC M. Motivated by the proof of [BMN2, Prop. 3.2], [AbLe1, Thm. 1] we can deduce that S is a (d + 1)-isometry. ∗(1))r∈N0 First recall that to check the (d+1)-isometric property we need to consider the powers (Φr of the map Φ∗ applied to the identity. We have Φr ∗(1)pm = nm+r nm pm. Since p the tuple S is a q-isometry iff Bp(S) = q (−1)r(cid:18)p r(cid:19)Φr ∗(1), Xr=0 (−1)r(cid:18)q r(cid:19)Φr ∗(1). 0 = Xr=0 This is to say 0 =Pq r=0(−1)r(cid:0)q nm r(cid:1) nm+r pm for all m ∈ N0, which is equivalent to (−1)r(cid:18)q Xr=0 ∀m ∈ N0 q r(cid:19)nm+r, and holds iff nm is a polynomial in m of degree ≤ q − 1. 0 = There are other properties of S encoded in the operators Bp(S). We expect that when HN is a coadjoint orbit, as assumed here, S is a complete hyperexpansion, i.e. but we do not know. When we replace S by the 1-isometry T we have: Bp(S) ≤ 0, ∀p ∈ N, 16 Proposition 2.12. Let S = ST be the polar decomposition of the shift compressed to HN. Then T is completely hypercontractive, i.e. Bp(T ) ≥ 0, ∀p ∈ N. But Bp(T ) = 0 for all p since T is a 1-isometry!! Proof. By definition T is a 1-isometry, hence subnormal with spectrum in clB ⊂ clBn. The result is now given by [?, Prop. 3.4]. As a special case of Proposition 6.7 later in the paper we have for each p ∈ N0 that Tr(Bp(S)pm) = p Xr=0 (−1)r(cid:18)p r(cid:19)nm+r, and in particular For p ≥ d + 1 we have, from the fact that nm is a polynomial of degree d, that Tr(B1(S)pm) = nm+1 − nm Tr(Bp(S)) = 0. Since B1(S)pm is a scalar we obtain recursively from the relation Bp(S) = (id−Φ∗)(Bp−1(S)) that Bp(S)pm is a scalar also for p ≥ 1, for each m ∈ N0. This gives another proof of the vanishing Bp(S) = 0 for p ≥ d + 1. 2.5 SOT-Toeplitz operators 2.5.1 The L∞ Toeplitz algebra C ∗(ς(L∞(S))) Again we consider the polar decomposition S = ST of the shift. By Lemma 2.2 the tuples S and T have the same invariant graded subspaces. The fact that T is a spherical isometry says that the completely positive map Ψ : B(HN) → B(HN) defined by Ψ(X) := n Xα=1 T ∗ αXTα, ∀X ∈ B(HN) is unital. Let B(HN)Ψ be the fixed-point set of the map Ψ. We say that an operator in B(HN)Ψ, i.e. an operator X with Ψ(X) = X, is Ψ-harmonic. And if X ∈ B(HN) satisfies Ψ(X) ≤ X then we say that X is Ψ-superharmonic. If X is Ψ-superharmonic and SOT− limp→∞ Ψp(X) = 0 then X is called a Ψ-potential (or pure Ψ-superharmonic). By [GaKu1, Thm. 3.3] or [Pop7, Thm. 3.1], if X is Ψ-superharmonic then it has a Riesz decomposition into the sum of a unique Ψ-harmonic operator X1 and a unique Ψ-potential X2, X = X1 + X2. Consider the von Neumann algebra generated by the operators in B(HN)Ψ (the L∞ Toeplitz algebra) The following lemma can then be deduced directly from [Prun1, Thm. 1.2]: Lemma 2.13. There is a short exact sequence of C∗-algebras L := C∗(B(HN)Ψ). 0 → SC(L) → L ς→ L∞(S) → 0 (2.5) with a unital completely positive splitting ς : L∞(S) → L 17 whose image equals B(HN)Ψ. Here SC(L) = Ker ς is the semicommutator ideal in L, i.e. the two- sided ideal generated by the operators [ς(f ), ς(g)) := ς(f )ς(g) − ς(f g) with f, g ∈ L∞(M). So we have a direct sum of vector spaces L = ς(L∞(S)) + SC(L) = B(HN)Ψ + SC(L). Endowed with the Choi -- Effros multiplication the operator system B(HN)Ψ becomes a von Neumann algebra isomorphic via ς to L∞(S), SOT− lim m→∞ Ψm(ς(f )ς(g)) = ς(f g), ∀f, g ∈ L∞(S). For every X ∈ L there is a unique Ψ-harmonic operator ς(fX ) ∈ B(HN)Ψ such that SOT− lim m→∞ Ψm(X) = ς(fX ) (2.6) and the map ς can be described as ς(X) = fX . If U : HN → H 0(S, ω) is the unitary which intertwines the multiplication tuple T on the Hardy space H 0(S, ω) with the weighted shift T = U −1T U on HN as in the proof of Lemma 2.6 then ς(f ) = U −1(P fH0(S,ω))U, ∀f ∈ L∞(S) where P is the orthogonal projection of L2(S, ω) onto H 0(S, ω) and we identify L∞(S) in its ∗- representation on L2(S, ω). By analogy with [BaHa1, Fein1] we say that an operator X ∈ B(HN) is SOT-asymptotic Toeplitz if the limit SOT− limm→∞ Ψm(X) exists. In this case SOT− limm→∞ Ψm(X) is fixed by Ψ so it must be of the form ς(fX ) for some fX ∈ L∞(S). Then fX is the symbol of X. Lemma 2.13 says that every element of L is SOT-asymptotic Toeplitz. Proposition 2.14. The SOT-asymptotic Toeplitz symbol map ς SOT coincides with ς when restricted to L. Proof. We have L = ς(L∞(S))+Ker ς and the conditional expectation onto ς(L∞(S)) just kills Ker ς. Thus Ker ς is the set of elements C of L with ς SOT(C) = 0. Corollary 2.15. Suppose that ς(f ) is a Toeplitz operator with SOT-Toeplitz symbol zero, SOT− lim m→∞ Ψm(ς(f )) = 0. Then f = 0. Note that Ker ς SOT is not contained in L however, and in particular there are SOT-asymptotic Toeplitz operators on HN which do not belong to L. Proposition 2.16. Let MH the WOT-closed algebra generated by 1, S1, . . . , Sn (this is the multiplier algebra of the Hilbert function space HN) and let MHM∗ H be the set of operators of the form ξη∗ with ξ, η ∈ MH. Then spanWOT C MHM∗ H = spanC{X ∈ B(HN) X ≥ 0, Φ(X) ≤ X} and this is a von Neumann algebra equal to L, L = spanWOT C MHM∗ H. 18 (2.7) (2.8) Proof. Since Φ is pure we have (by [Arv7c, Prop. 1.6] or [Pop7, Cor. 3.10]) that an operator X ∈ B(HN) satisfies X ≥ 0 and Φ(X) ≤ X if and only if X is "factorable" in the sense that X = LL∗ with some A-module map L from HN ⊗ M and HN, for some separable Hilbert space M. Suppose then that X is factorable like this. The fact that L is a module map means precisely that if (ej)j∈J is an orthonormal basis for M then the vectors belong to MH. Moreover, we have (2.9) ∀j ∈ J ξj := L(1 ⊗ ej), X = SOT−Xj∈J H. Conversely, for X ∈ spanWOT ξjξ∗ j so that X belongs to spanWOT an expansion (2.9) and this gives Φ(X) ≤ X since Φ(1) ≤ 1 gives C MHM∗ C MHM∗ H with X ≥ 0 we have Φ(cid:16)Xj∈J ξj ξ∗ j(cid:17) = n Xα=1Xj∈J So Equality (2.7) holds. Sαξj ξ∗ j S∗ α =Xj∈J ξjΦ(1)ξ∗ j ≤Xj∈J ξjξ∗ j . We next prove that (2.8) holds. The quotient L/ Ker ς is isomorphic to L∞(S), which is commu- tative. Therefore every element of L can be normally ordered modulo Ker ς, i.e. every element of L belongs to spanWOT HMH of anti- normally ordered products of multipliers is precisely the set of fixed points under Ψ, since Ψ(1) = 1. Therefore H up to some term in Ker ς. Now observe that the set M∗ C MHM∗ M∗ HMH = B(HN)Ψ = Ran ς = (Ker ς)⊥, C MHM∗ H. Therefore spanWOT and so Ker ς is contained in spanWOT a von Neumann algebra which contains B(HN)Ψ, whence it must equal L = C∗(B(HN)Ψ). Remark 2.17 (Uniform Toeplitz operators). Again by analogy with [Fein1] we say that an operator X ∈ B(HN) is uniform asymptotic Toeplitz if the limit limm→∞ Ψm(X) exists in the norm topology on B(HN). As a generalization of pre-existing proofs in the literature (cf. [CuLe1, Lemma 3.1]) we can deduce that every compact operator K is uniform asymptotic Toeplitz with symbol H is an algebra, and thus C MHM∗ First observe that, since the polynomials are dense in HN and since every compact operator is a norm-limit of finite-rank operators, it sufficies to show that fK = 0. m→∞kΨm(ξihη)k = 0. lim for any rank-1 operator K = ξihη where ξ, η ∈ A ⊂ HN are polynomials. Next, Ψm(K) is a finite sum of operators of the form T ∗ k KTk for multi-indices k ∈ F+ n with k = m. Now k ξihT ∗ k η. kξihηTk = T ∗ T ∗ If deg ξ = m0 then T ∗ obtain Ψm(ξihη) = 0, as desired. k ξ = 0 for all k ∈ F+ n with k > m0. Thus, for all m ≥ min{deg ξ, deg η} we From [Prun1, Thm. 1.2] and [An6] we also obtain a short exact sequence of C∗-algebras 0 → K(HN) → TH ς→ C0(S) → 0 (2.10) where K(HN) is the ideal of compact operators on the Fock space HN, and this sequence is also linearly split by the Toeplitz map ς. That is, every element of TH is of the form ς(f ) + K with f ∈ C0(S) and K compact. This shows that every element of TH is uniform asymptotic Toeplitz. 19 Remark 2.18 (Fixed points). As mentioned, the (d + 1)-isometric property of S can be expressed as Φ∗(Bd(S)) = Bd(S). Each operator Bp(S) is compact. While Ψ has no compact fixed points, Φ∗ thus has. Xmpm with Xm ∈ B(Hm) for each m. The algebra Γb of grading-preserving bounded operators on HN can therefore be identified as If X is an operator on HN that preserves the grading then one has X = Pm∈N0 φm. The grading on HN gives rise to an Γb = Ym∈N0 B(Hm). It is a von Neumann algebra admitting a finite trace Pm∈N0 action of U(1) on B(HN) whose fixed-point subalgebra is given by Γb. B∞ := B(HN)Ψ ∩ Γb In [An6] we only discussed the U(1)-invariant part of the operator system B(HN)Ψ, and we showed that B∞ is completely isometrically isomorphic to L∞(M) via an explictly Toeplitz-type map ς : L∞(M) → B∞. The notation for the map ς : L∞(S) → B(HN) appearing in Lemma 2.6 was chosen because its defining property (2.6) shows that it restricts to the Toeplitz map ς : L∞(M) → Γb. Therefore there is also no confusion if we refer to ς : L∞(S) → B(HN) as the Toeplitz map. This is in accordance with [Prun1, Prun2] and the single-variable theory of "generalized Toeplitz operator". Let N := C∗(B∞) be the C∗-algebra generated by the operators in B∞. Then N is a von Neumann subalgebra of L which we call the L∞ Toeplitz core. Taking the "U(1)-invariant part" of Proposition 2.16 we obtain a presentation of N in terms of factorable or superharmonic elements, N = spanWOT C (MHM∗ H) ∩ Γb = spanC{X ∈ Γb X ≥ 0, Φ(X) ≤ X}. With these extra facts the following is implicit in [An6]: Corollary 2.19. The von Neumann algebra N := C∗(B∞) fits into the short exact sequence of C∗-algebras (2.11) where the kernel [B∞,B∞) = Ker(ς) is the semicommutator ideal in N . The Toeplitz map ς : L∞(M) → N gives a completely positive splitting, so that 0 → [B∞,B∞) → N ς→ L∞(M) → 0 N = B∞ + [B∞,B∞), and ς is the adjoint of the Toeplitz map ς : L∞(M) → Γb with respect to the inner product of L2(M, ω) and the inner product on Γb defined by the state Pm φm. 2.5.2 Projections onto vector-valued quotient modules We now look at graded subspaces HE For ease of notation we write N of the graded Hilbert space HN ⊗ CN for some integer N ∈ N. Sα := Sα ⊗ 1N for the diagonal representation of the shift operators on HN ⊗ CN . We shall characterize coinvariance N in terms of the orthogonal projection PE of HN ⊗ CN onto HE of HE N . It will often be convenient to regard PE and other operators on HN ⊗ CN as N × N -matrices with entries in B(HN). Then we can apply the maps Φ and Ψ entrywise to such operators. Thus we define Φ(X) := (Φ ⊗ idN )(X), ∀X ∈ B(HN) ⊗ MN (C) = B(HN ⊗ CN ) 20 and so on. Recall that Γb ⊂ B(HN) is the von Neumann algebra of grading-preserving operators on HN. If φ is a bounded multiplier of HN then we denote by Mφ the associated bounded operator on HN, (Mφψ)(v) := φ(v)ψ(v), ∀ψ ∈ HN, v ∈ B. Lemma 2.20. For a graded projection PE acting on HN ⊗ CN , the following are equivalent: (a) 1− PE has a Beurling factorization, i.e. there is a multiplier ΘE : HN ⊗ ℓ2(N0) → HN ⊗ CN such that 1 − PE = MΘE M ∗ ΘE = SOT−Xj∈N Mφj M ∗ φj with φj ∈ Amj ⊗ CN a homogeneous polynomial for each j ∈ N (where φj = 0 is allowed). N of PE is invariant under S∗. (b) Φ(1 − PE) ≤ 1 − PE, in which case SOT− limm→∞ Φm(1 − PE) = 0. (c) The image HE (d) Ψ(PE) ≤ PE. (e) PE = ς(P E) + CE where P E is a projection over L∞(M) and CE is a Ψ-potential. In this case CE has entries in N (since PE has, by (a)). Proof. (a) ⇐⇒ (b) is shown in [Arv7c, Prop. 1.6] or [Pop7, Cor. 3.10], using SOT− limm→∞ Φm(1) = 0. More precisely, that the φj's can be taken to be homogeneous polynomials is in [Zhao1, Lemma 2.2] (the proof there generalizes immediately to arbitrary N ; see also [Arv7b, Cor. 2 to Prop. 8.13] for N = 1). If Φ(1− PE) ≤ 1− PE then HE Conversely, if HE N is invariant under S∗ by [Pop7, Lemma 4.1] (using Φ(1) ≤ 1), so N is coinvariant then we have Φ(1 − PE) ≤ 1 − PE (see [Arv7c, Eq. (2.4]). So Since Ψ(1) = 1 we have the equivalence of (c) and (d) by [Pop7, Cor. 4.4]. Combining, we have shown that (a) -- (c) is equivalent to (d). Assuming that this holds, i.e. that Ψ(PE) ≤ PE, we obtain a Riesz decomposition PE = ς(P E) + CE where ς(P E) is in B∞ and CE is a Ψ-potential. By (a) we have PE over N and since N = B∞ + Ker Ψ∞ we get that both ς(P E) and CE are over N as well. Thus P E is over L∞(M). Conversely, if (e) holds then Ψ(PE) ≤ PE and so we are done. (c) =⇒ (b). (b) =⇒ (c). Remark 2.21. For the equivalence of (a) and (c), see also [Sark3, Cor. 3.3]. In the one-variable setting (n = 1), every submodule of H 2 1 = H 0(T) is the range of an inner multiplier, by the Beurling theorem. There is a Beurling decomposition of every submodule of H 2 n also for n ≥ 2. There are no nontrivial graded submodules of H 0(T), so there is no surprise that when we look at graded submodules we encounter new phenomena, namely that the multipliers in the Beurling decomposition can be chosen to be homogeneous polynomials. In the same fashion we obtain the ungraded version of Lemma 2.20: Lemma 2.22. For a projection PE on HN ⊗ CN , the following are equivalent: (a) 1 − PE has a Beurling factorization, i.e. 1 − PE = SOT−Xj∈N φjφ∗ j with φj ∈ MH a multiplier for each j ∈ N (where φj = 0 is allowed). (b) Φ(1 − PE) ≤ 1 − PE, in which case SOT− limm→∞ Φm(1 − PE) = 0. (c) The image of PE is invariant under S∗. (d) Ψ(PE) ≤ PE. 21 (e) PE = ς(P E) + CE where P E is a projection over L∞(S) and CE is a Ψ-potential. In this case CE has entries in L (since PE has, by (a)). Remark 2.23. For the Fock space HN associated with an arbitrary projective variety M one has Ψ(1) ≤ 1 iff Ψ(1) = 1. So in case Ψ is not unital, the superharmonicity property Ψ(PE) ≤ PE is not equivalent to the coinvariance property Φ(1 − PE) ≥ 1 − PE. If we let PE be a Ψ-superharmonic projection and write its Riesz decomposition as PE = ς(P E) + CE then CE is a potential. The "charge" of the potential CE is, in the terminology of [GaKu1, §3], given by i.e. XE := (id−Ψ)(CE) = (id−Ψ)(PE) ≥ 0, CE = SOT− Xm∈N0 Ψm(XE) = SOT− Xm∈N0 (Ψm − Ψm+1)(PE). We can see (Ψm − Ψm+1)(PE) as the (m + 1)th order obstruction of PE from being harmonic. So CE contains not only the first-order obstruction XE = (id−Ψ)(PE), although this is the leading term. The potential CE need not have the same range projection as its charge XE. However, there is always a Ψ-summable element XE whose range projection is equal to that of CE [GaKu1, Lemma 4.1]. 2.5.3 Vector-valued essential normality Following Barr´ıa -- Halmos [BaHa1] we observe that since both Pn αSα − 1 are compact (i.e. S is essentially a spherical unitary), the essential commutant of S can be described as αSα − 1 and Pn α=1 S∗ α=1 S∗ {S}ec = {X ∈ B(HN) (id−Ψ)(X) ∈ K(HN)}. Since the essential commutant is a C∗-algebra which contains the set B(HN)Ψ of all Ψ-harmonic elements, it must contain L = C∗(B(HN)Ψ), L ⊂ {S}ec. This observation leads to: Theorem 2.24. Suppose that PE is a projection acting on HN⊗CN with Ψ(PE) ≤ PE and preserving the N0-grading. Then the shift tuple SE on the graded quotient module HE N := Ran PE is essentially normal. Proof. Write (id−Ψ)(PE)HE N = 1 − Xm∈N0 nm nm+1 n Xα=1 S∗ E,αSE,αpm = 1 − Λ−1S∗ ESE nm+1 where Λ :=Pm∈N0 pm does not depend on the projection PE. By Proposition 2.16, the assump- tions of the theorem imply that PE is a projection in N ⊗ MN (C). The entries of (id−Ψ)(PE)HE are compact since PE has entries in N ⊂ {S}ec. Since Γ0 := K(HN) ∩ Γb is an ideal in Γb and Λ is bounded, we have nm N S∗ ESE − Λ = Λ(id−Ψ(PE)HE N ∈ Γ0 ⊗ MN (C). Since SES∗ of degree d we easily get that SES∗ This gives E = 1− PE,0 is a finite-rank perturbation of the identity operator and nm is a polynomial E − Λ is compact. E − Λ is in Lp iff p > d + 1, so in particular SES∗ E = (S∗ ESE − Λ) − (SES∗ E − Λ) ∈ Γ0 ⊗ MN (C) as desired. ESE − SES∗ S∗ 22 The proof relies on the assumption that PE is graded, since only then does PE commute with pm for all m. And indeed there are ungraded counterexamples to essential normality [GRS1, §4.1]. Example 2.25. If N = 1 then graded submodules of HN are in bijection with homogeneous ideals in the coordinate ring A. The quotient modules HE n, are thus orthogonal complements of ideals and equal the completions of the coordinate rings of analytic subvarieties ME ⊂ M. N as PE = ς(P E) + CE then we get If we write the projection PE of HN onto such a quotient module HE N of HN, which are also quotients of H 2 ω(P E) = lim m→∞ φm(PE,m) = lim m→∞ dim HE m nm = 0 unless dim ME = dim M (i.e. unless ME = M) since dim HE Thus, if ME is not all of M, m is a polynomial in m of degree dim ME. ς(P E) = 0. In other words PE is in the kernel of the symbol map ς. We see that the kernel of ς is much larger than the ideal of compact operators. We also see that the projection P I onto the orthogonal complement N of HE HI N has symbol ς(PI ) = ς(1− PE) = 1. So the projection ς(PI ) does not recover the ideal sheaf I of ME, even if I is locally free. This is not surprising since, even if I is locally free, the subsheaf I ⊂ OM is not a subbundle. Remark 2.26 (Generalization to general M). Submodules of HN ⊗ CN are semi-invariant under the backward Drury -- Arveson shift M ∗: they are submodules of a quotient module. The projection onto such a subspace is of the form PE − PF where both PE and PF ≤ PE are coinvariant. Since Theorem 2.24 says that M -coinvariant projections are essentially normal we get that all submodules of vector- valued complete Nevanlinna -- Pick-modules are essentially normal. So for essential normality one gets nothing new by dropping the assumption about having a coadjoint orbit. If Ψ(X) ≤ X and we write the Riesz decomposition of X as X = ς(fX ) + CX then we get Hence, if X is over N , the pure Ψ-superharmonic part always satisfies (id−Ψ)(X) = (id−Ψ)(CX ) ≥ 0. (id −Ψ)(CX ) ∈ Γ0. But even in the case of a graded projection X = PE it need not be that CX itself is compact. 3 Cowen -- Douglas bundles of quotient modules 3.1 Setup 3.1.1 The reference Hermitian line bundle Recall that B ⊂ Bn is defined as the zero set in the unit ball of the ideal in C[z1, . . . , zn] corresponding to the embedding M ֒→ CPn−1. In other words, if A is the homogeneous coordinate ring of M then B = Spec(A) ∩ Bn. The Fock space HN is a graded completion of A and its elements are analytic functions on B. In the following we write Ker(S∗ − v1) := \α=1 for the subspace of joint eigenvectors of S∗ 1 , . . . , S∗ know that Ker(S∗ − v1) is nonzero for each v ∈ B, Ker(S∗ α − vα1) ⊂ HN n with joint eigenvalue v = (v1, . . . , vn) ∈ B. We σp(S∗) = B. 23 Indeed, Ker(S∗ − v1) is 1-dimensional and spanned by the reproducing vector K¯v (or its normalized version k¯v := K¯v/kK¯vk) at ¯v, Ker(S∗ − v1) = Ck¯v. Since dim Ker(S∗ − v1) = 1 for all v ∈ B and the map B ∋ v → k¯v is holomorphic, the vector spaces OCD(v) := Ker(S∗ − v1) form a holomorphic line bundle over B which we denote by OCD and call the Cowen -- Douglas bundle of the Hilbert module HN (or of the operator tuple S). We have a global holomorphic section v → k¯v of OCD which vanishes nowhere. So OCD is a trivial holomorphic line bundle (it is not algebraic though, since k¯v is not algebraic): OCD ∼= OB. Sometimes we will only need the restriction of OCD to B \ {0}, and we denote this restriction again by OCD. Definition 3.1. The Cowen -- Douglas projection of HN is the projection CD(HN) ∈ C0(B \ {0}) ⊗ B(HN) which defines the line bundle OCD, i.e. CD(HN)(v) := k¯vihk¯v = projection onto Ker(S∗ − v1). For ψ ∈ HN and v ∈ B we have hK¯vψi = ψ(¯v) and hence (CD(HN)ψ)(v) = CD(HN)(v)ψ = kK¯vk−1ψ(¯v)k¯v. Since kK¯vk−1ψ(¯v)k¯v belongs to the subspace Ker(S∗ − v1) of HN for each v ∈ B, the function B ∋ v → CD(HN)(v)ψ ∈ HN is a section CD(HN)ψ ∈ Γ0(B;OCD) of the line bundle OCD. Moreover, the section CD(HN)ψ is holomorphic because both v → k¯v and ψ are holomorphic. If CD(HN)ψ is the zero section then ψ belongs to(cid:16)∨v∈B Ker(S∗− v1)(cid:17)⊥ into the space of global holomorphic sections of OCD. then we can regard CD(HN) as a mapping If we identify the projective space P[HN] with the manifold of rank-1 projections acting on HN = {0}. Thus HN embeds CD(HN) : B → P[HN], which is a holomorphic embedding since kv is never zero. If (ψj )j∈N is any orthonormal basis for HN then from hψjKvi = ψj(v) we see that one can express the coherent vectors as kv = kKvk−1Xj∈N ψj(v)ψj , ∀v ∈ B. Therefore, whatever orthonormal basis (ψj)j∈N for HN chosen for the identification of P[HN] with the infinite-dimensional projective space CP∞, the embedding CD(HN) can therefore be described as where we denote by [z1 : z2 : ··· ] the homogeneous coordinates on CP∞. CD(HN)(v) = [ψ1(¯v) : ψ2(¯v) : ··· ], 24 In the literature on Kahler geometry it is more common to consider the slightly different holo- morphic embedding FS(HN) : B → P[HN], FS(HN)(v) := [ψ1(v) : ψ2(v) : ··· ]. The notation FS(HN) makes sense since FS(HN) depends only on the Hilbert space and not the choice of orthonormal basis; FS stands for "Fubini -- Study". We shall not need to distinguish between FS(HN)(v) and CD(HN)(v) and we identify them as the same projection acting on HN. More generally if H is the Hilbert space of global holomorphic sections of some globally generated vector bundle then we denote by FS(H) the embedding of the base manifold into the Grassmannian defined in the same way as above by an orthonormal basis for H. As a special case, the embedding M ⊂ CPn−1 is correspond to the projection FS(H1), after we have the identification H1 = Cn so that P[H1] = CPn−1. We are also interested in the embeddings FS(Hm) of M into P[Hm] where Hm as before is the vector space H 0(M;Lm) of global holomorphic sections of the line bundle Lm endowed with the inner product of the tensor product (H1)⊗m. We can regard FS(Hm) as a projection in the C∗-algebra C0(M) ⊗ B(Hm). Any choice of Parseval frame (ψj )j∈J for Hm gives an expansion of FS(Em) as Xj,k∈J with coefficients FS(Hm)j,k ∈ C0(M) given by FS(Hm) = nm FS(Hm)j,kψjihψk For the particular choice (ψj )j∈J = (Zk)k=m we obtain FS(Hm)j,k(x) = ψj(x)ψk(x), ∀x ∈ M. FS(Hm) = Xj=m=k ZjZ ∗ k ⊗ SjS∗ kpm (3.1) where Sj := Sj1 ··· Sjm for a multi-index j = j1 ··· jm. The m-fold tensor product of the projection FS(H1) is given by FS(H1)⊗m = Xj=m=k ZjZ ∗ k ⊗ ejihek, where e1, . . . , en is an orthonormal basis for Hm and ej := ej1 ⊗ ··· ⊗ ejm . Since the Zα's satisfy the relations of the ideal defining M, we can express this as FS(H1)⊗m = Xj=m=k ZjZ ∗ k ⊗ pmejihpmek. and be regarded as a function on M with values in the subalgebra B(Hm) ⊂ B(H⊗m). Now pmejihpmek = SjS∗ identified with FS(Hm). kpm. Thus the Fock inner product on Hm is precisely the one such that FS(H1)⊗m is naturally If we write v = rζ ∈ B with r ∈ (0, 1) and ζ ∈ S then we have ¯vjSjΩ = (1 − r2)1/2 Xm∈N0 kv = (1 − v2)1/2 Xj∈F+ CD(HN)(rζ) = kr ¯ζihkr ¯ζ = (1 − r2) Xm∈N0 rm Xj=m r2m Xj=m=k n ¯ζjSjΩ. ζj ¯ζkSjS∗ k. Thus While Hm ∩ Ckv = {0}, the projection pmkv of kv onto Hm spans a 1-dimensional subspace of Hm; the projection onto this subspace is CD(Hm)(v) := pmk¯vihpmk¯v kpmkvk2 . 25 It depends only on the class [v] of v in the quotient M = (B\{0})/D×. We write k(m) so that CD(Hm)(x) = k(m) ψ ∈ Hm the reproducing property of Kv gives hk(m) := pmkv/kpmkvk, for all x ∈ M. We have kpmkvk2 = (1 − v2)−1v2m. For all ψi = ψ(x), ihk(m) [v] ¯x ¯x x where x → ψ(x) is the function on M induced by the homogeneous degree-m polynomial ψ. Therefore, for any Parseval frame (ψj)j∈J for Hm we can expand the B(Hm)-factor of CD(Hm) (cf. [Bala1]) to obtain CD(Hm) = Xj,k∈J with coefficients CD(Hm)j,k ∈ C0(M) given by CD(Hm)j,kψjihψk CD(Hm)j,k(x) = hψjk(m) ¯x ihk(m) ¯x ψki = ψj(¯x)ψk(¯x), ∀x ∈ M. Thus CD(Hm) coincides with FS(Hm). If we take the Parseval frame for Hm given by (ψj )j∈J = (ek)k=m where e1, . . . , en is the standard basis for H1 = Cn then we see that CD(HN)(rζ) = Xm∈N0 r2m Xj=m=k FS(Hm)([ζ]). 3.1.2 Higher-rank Cowen -- Douglas bundles Throughout this section, HE HE N . The backward shift S∗ on HN ⊗ CN restricts to a row contraction S∗ HE N . We shall study the commuting operator tuple S∗ amounts to looking at the family of eigenspaces of S∗ E, E − v1), N is a graded quotient module of HN ⊗ CN and PE is the projection onto E on the Hilbert subspace E with the Cowen -- Douglas approach. This ECD(v) := Ker(S∗ and how they vary with v. If S∗ E were in the Cowen -- Douglas class Br(B) for some r ≥ 1 then the ECD(v)'s would be the fibers of a holomorphic vector bundle on B. However, we shall see that this is too strong an assumption if we want to use operator theory to study vector bundles over M. The "correct" condition for projective geometry is instead to ask for membership in the Cowen -- Douglas class Br(B \ {0}), i.e. we have to allow ECD to be singular at 0. ξ ∈ CN , The eigenvectors of the backward shift S∗ on HN ⊗ CN are of the form kv ⊗ ξ with v ∈ B and ∀v ∈ B, S∗(kv ⊗ ξ) = ¯vkv ⊗ ξ. The joint spectrum of the tuple S∗ is equal to the joint point spectrum, which is σ(S∗) = σp(S∗) = B. The multiplicity of each eigenvalue is equal to N . Since S∗ E is just the restriction of S∗, the vector spaces Ker(S∗ E − v1) = Ran(SE − ¯v1)⊥ is a subspace of k¯v ⊗ CN for each v ∈ B, and so finite-dimensional. In particular, Ran(SE − ¯v1) is a closed subspace of HE E is in class Br(B \ {0}) is the same as asking for the consancy of the function dim Ker(S∗ E − v1) in v ∈ B \ {0}. N . Clearly the eigenvectors of S∗ N . So the condition that S∗ E span HE N is an S∗-invariant subspace, it is of the form Since HE for some closed subset E ⊂ B × CN . But since the coherent vectors kv are not orthogonal, there are many such sets E. We will will adopt a special notation for the largest possible choice of E, namely HE N = span{kv ⊗ ξ (v, ξ) ∈ E} Define ECD := {(v, ξ) ∈ B × CN S∗ E(k¯v ⊗ ξ) = vk¯v ⊗ ξ}. ECD(v) := {ξ ∈ CN, (v, ξ) ∈ ECD}, ECD(v). Since Ker(SE − v1) ⊂ HE N we must thus have ECD(v) = span{k¯v ⊗ ξ ξ ∈ so that ECD =Fv∈B ECD(v)}. Let us record this fact: 26 Proposition 3.2. For each v ∈ B there is a vector space ECD(v) ⊂ CN such that ECD(v) = Ck¯v ⊗ ECD(v). Let ΘE : B× ℓ2(N0) → B× CN be any multiplier such that the associated multiplication operator N . The adjoint operator ΘE acts on coherent vectors as MΘE ∈ B(HN ⊗ CN ) has range equal to the orthogonal complement of HE M ∗ M ∗ ΘE (kv ⊗ ξ) = kv ⊗ ΘE(v)∗ξ. (3.2) So we have M ∗ Ker M ∗ ΘE gives ΘE (kv ⊗ ξ) = 0 if and only if ΘE(v)∗ξ = 0. On the other hand, the relation EN = Thus (3.2) gives and, for each v ∈ B, (cf. [KwTr1, Remark 1.2]). k¯v ⊗ ξ ∈ Ker M ∗ Θ ⇐⇒ (v, ξ) ∈ ECD, ECD(v) = Ker Θ∗ E(¯v), Ker(S∗ E − v1) = Ck¯v ⊗ Ker Θ∗ E(¯v) The analyticity of ΘE and the finite-dimensionality of ECD(v) for each v ensure that the ECD(v)'s form the holmorphic linear space of a coherent analytic sheaf over B in the sense of [Fisc1, §1.6]. Therefore also the ECD(v)'s form a coherent analytic sheaf ECD over B. We call ECD the Cowen -- Douglas sheaf of the quotient module HE N . 3.2 Algebraic aspects 3.2.1 Graded A-modules and coherent sheaves H 0(M;Lm) be the graded coordinate ring of the embedded variety M ⊂ CPn−1. The Let A =Lm∈N0 Fock space HN is the completion of A in the inner product of H∨N and H 0(M;Lm) is the vector space underlying the Hilbert space Hm for each m ∈ N0. Given a quotient module HE N of HN ⊗ CN as before, define N ∩ (A ⊗ CN ). EN := HE Then EN is the graded A-module whose completion in the inner product of HN ⊗ CN is equal to HE N . Note that EN is isomorphic to a quotient of A ⊗ CN and thus finitely generated. Recall that for each A-module one can associate in a canonical fashion an algebraic sheaf on V := SpecA [Serr2]: Definition 3.3. Let E be an A-module. The Serre sheaf of E is the OV-module EV := E ⊗A OV. If E = EN is a graded A-module we can also define an OM-module E (also referred to as the Serre sheaf of EN) by E := EN ⊗A OM. We shall denote by EV\{0} the restriction of EV to V \ {0}. If we for α ∈ {1, . . . , n} let Uα ⊂ V be the open set where the coordinate function Zα ∈ A is nonzero then OV(Uα) = AZα is the localization of the ring A at Zα. So EV(Uα) = E ⊗A OV(Uα) = E ⊗A AZα = EZα 27 is the the module of fractions of E with denominator Zα. If we denote by Uα ⊂ M also the projec- tivization of Uα ⊂ V \ {0} then OM(Uα) = A(Zα) is the homogeneous localization of the ring A at Zα. So the Serre sheaf E on M of a graded module EN can be described as E(Uα) = EN ⊗A OM(Uα) = EN ⊗A A(Zα) = (EN)(Zα), i.e. by taking homogeneous localizations of the graded module EN. Let EN be a graded A-module with Serre sheaves E and EV\{0} on M and V\{0} respectively, and consider the quotient map π : V \ {0} → M. Evidently π−1(EN ⊗A OM) ⊗π−1OM OV\{0} = EN ⊗A OV\{0}. That is, Lemma 3.4 ([Serr2]). The Serre sheaf of a (graded) A-module E is a coherent as OV-module (or OM-module) if and only if E is finitely generated. The module E identifies with the module of global holomorphic sections of its Serre sheaf, π∗E = EV\{0}. H 0(V;EV) = E. If EN is a graded A-module then modulo finite-dimensional A-modules we have H 0(V \ {0};EV\{0}) = Mm∈N0 H 0(M;E(m)) ∼= EN as graded A-modules. Conversely, if we start with a coherent OM-module E then the Serre sheaf of graded A-module EN :=Lm∈N0 Thus replacing EN by EN :=Lm∈N0 H 0(M;E(m)) we get a Serre sheaf E on M which is isomorphic to E. However, EV := EN ⊗A OV can differ from EV at the origin 0 ∈ V where the finite-dimensional distinction between EN and EN is still significant. The Abelian category coh M of coherent algebraic sheaves on M can be identified with the quotient H 0(M;E(m)) is isomorphic to E as OM-module. category qgr(A) = gr(A)/ tors(A), where gr(A) is the category of of finitely generated graded A-modules and tors(A) is the subcategory of modules which are finite-dimensional as vector spaces over C [Serr2, §59]. The quotient functor gr(A) → qgr(A) is exact, while the global section functor qgr(A) ∋ E → Lm H 0(M;E(m)) ∈ gr(A) is only left-exact. Thus if 0 → I → E → F → 0 is a short exact sequence of quasicoherent sheaves then we get an exact sequence 0 → IN → EN → FN of graded A-modules by applying the global section functor. So even if F is globally generated, i.e. if we have a surjection OM ⊗ CN → F → 0 for some N , we cannot conclude that FN := Lm H 0(M;F (m)) is a graded quotient of A ⊗ CN . There is however a graded A-module quotient A ⊗ CN → FN → 0 with Fm = Fm for m ≫ 0 and the Serre sheaf of FN equals F . If EN is a graded quotient module of HN ⊗ CN then the shift SE on EN satisfies SES∗ E := n Xα=1 SE,αS∗ E,α = PE n Xα=1 = PE(1 − p0 ⊗ 1N )EN SαS∗ α(cid:12)(cid:12)EN so that SES∗ N = 1) one can use this fact to construct explicit isometric embeddings (cf. Eq. (6.3)) E restricted to Em equals the identity operator on Em for all m 6= 0. As in [An6] (where The "subproduct" structure (3.3) is a generalization of the subproduct property of H•, which reads El ֒→ Em ⊗ Hl−m, ∀l ≥ m ≥ 0. (3.3) Hl ֒→ Hm ⊗ Hl−m, ∀l ≥ m ≥ 0. 28 H 0(M;E(m)) is a graded quotient of A ⊗ CN ; in this case from (3.3) we have canonical embeddings of vector spaces (3.4) For an arbitrary coherent OM-module E it is not necessarily the case that Lm∈N0 ∀l ≥ m ∈ N0, H 0(M;E(l)) ֒→ H 0(M;E(m)) ⊗ H 0(M;Ll−m), and this is a characteristic of sheaves E which are regular in the sense of Castelnuovo -- Mumford [Laza1, Thm. 1.8.3]. These are in particular globally generated. 3.2.2 Serre sheaf versus Cowen -- Douglas sheaf Let EN be a graded quotient of the standard Hilbert module HN ⊗ E0 for some finite-dimensional Hilbert space E0. Let SE be the shift on EN. As mentioned, the vector spaces Ker(S∗ E − v1) form a coherent analytic sheaf ECD. Let mv be the ideal of functions in A vanishing at v. The vector space Ker(S∗ E − v1) is linearly isomorphic to the annihilator of Ran(SE − v1) in the dual space of EN, and the latter is linearly isomorphic to (EN/mvEN)∗, so Ker(S∗ E − v1) ∼= (EN/mvEN)∗. Let EN be the graded A-module whose completion equals EN, i.e. EN = EN ∩ (A ⊗ E0). The fibers of the Serre sheaf EB are given by EB(v) = EN ⊗A Cv ∼= EN/mvEN. In general EN/mvEN and EN/mvEN may not be isomorphic. In this section we compare the Cowen -- Douglas sheaf of EN with the Serre sheaf of EN. Let IN be the orthogonal complement of EN; thus IN is a graded submodule of HN ⊗ E0 and equals the completion of a graded A-submodule IN of A ⊗ E0. Lemma 3.5 (cf. [Fang4, p. 1690]). Define the fiber space over v ∈ B of the Hilbert module IN to be the vector space Then the map IN(v) := {ψ(v) ∈ E0 ψ ∈ IN}. IN/(IN ∩ (mvHN ⊗ E0)) → IN(v) induced by evaluation of functions at v is an isomorphism, and we have a short exact sequence of vector spaces. 0 → IN(v)∗ → OCD(v) ⊗ E0 → ECD(v) → 0 (3.5) (3.6) Proof. Let v ∈ B and let ev : HN ⊗ E0 → E0 be the evaluation at v, i.e. ev(ψ) := ψ(v). Clearly the restriction of ev to IN is onto the fiber space IN(v). Moreover, if ψ(v) = 0 then hψkv ⊗ ξiHN⊗E0 = hξψ(v)iE0 = 0 so ψ belongs to Ker(S∗ − ¯v1)⊥ ⊗ E0 = mvHN ⊗ E0. Hence the map (3.5) is an isomorphism. The short exact sequence 0 → IN/(IN ∩ (mvHN ⊗ E0)) → (HN/mvHN) ⊗ E0 → EN/mvEN → 0 then gives (3.6). Let I be the Serre sheaf of IN, so that rankI = limm→∞ dim Im/ dim Hm. From [GRS1, Thm. 1.2] or [Fang4, Lemma 16] we have, for each v ∈ B, dim IN(v) = rankI =⇒ dim(IN/mvIN) = rankI. 29 For φ ∈ IN we have φ(v) = 0 if and only if φ belongs to IN ∩ mv(HN ⊗ E0), so if and only if we can write φ = (zα − vα)ψα n Xα=1 with ψ1, . . . , ψn ∈ HN ⊗ E0. By definition, IN is Gleason solvable at v if and only if for each φ ∈ IN ∩ mv(HN ⊗ E0) we can take the ψα's to belong to IN. Since EN is not necessarily invariant under Sα − vα1, there could at the same time be possible to choose ψα to not belong to IN. But we see that IN is Gleason solvable at v if and only if the inclusion is an equality. We have dim IN(v) < rankI iff (3.7) is a proper inclusion. Let us now look at the algebraic analogues of the above. We have a short exact sequence mvIN ⊂ IN ∩ mv(HN ⊗ E0). (3.7) 0 → IN/(IN ∩ (mv ⊗ E0)) → A/mv ⊗ E0 → EB(v) → 0. The vector space IN/(IN ∩ (mv ⊗ E0)) is isomorphic to IN(v) := {φ(v) φ ∈ IN} so there is also a short exact sequence 0 → IN(v) → A/mv ⊗ E0 → EB(v) → 0. Thus EB is locally free at v iff dim IN(v) does not drop from its maximal value rankI. Proposition 3.6. Let EN be a graded quotient of HN ⊗ E0 with underlying graded A-module EN, and suppose that the Serre sheaf E of EN is locally free. Then the Cowen -- Douglas sheaf ECD of EN is locally free on B \ {0} and isomorphic to the dual of the pullback EB\{0} of E to B \ {0}, ECD ∼= E ∗ B\{0}. Proof. Since EB\{0} is locally free, dim IN(v) = rankI for all v ∈ B\{0}. The inclusion IN(v) ⊂ IN(v) is thus an equality for each v, so that ECD is locally free on B \ {0} as well. The algebraic vector bundle Sv∈B\{0} IN/mvIN is isomorphic to Sv∈B\{0} IN(v). Since equality holds in (3.7) for each v ∈ B \ {0}, the vector bundle Sv∈B\{0} IN(v) is analytically isomorphic to the holomorphic vector bundleSv∈B\{0} IN/mvIN via the maps induced by the evaluation homomorphisms ev : IN → E0. We shall later see that when ECD is locally free it is D×-equivariant (up to a factor of OCD), just as EB\{0}. When ECD and EB\{0} are isomorphic, this need not be by a D×-equivariant isomorphism however. Remark 3.7 (Germ model). Another sheaf associated to a submodule IN is studied in [BMP1], namely the sheaf IBMP whose stalk at v ∈ B is given by IBMP,v :=n k Xj=1 (φj{v})ψj(cid:12)(cid:12)(cid:12) φ1, . . . , φk ∈ IN, ψ1, . . . , ψk ∈ Ov, k ∈ No. The sheaf IBMP coincides with the "germ model" of IN in the sense of Cheng -- Fang [ChFa1, §4]. The Cowen -- Douglas sheaf EN ⊗A OB is instead the restriction to B ⊂ V of the sheaf model of [ChFa1, §4], EN ⊗A OV = (EN ⊗ OV)/((SE ⊗ 1 − 1 ⊗ Z)(EN ⊗ OV)) = OV(EN)/(SE − Z1)OV(EN). By definition IBMP is a subsheaf of OB ⊗ E0. [BMP1, Eq. (1.1)]) It coinides with the image under the map (cf. ICD = IN ⊗O(B) OB → (HN ⊗ E0) ⊗O(B) OB = OB ⊗ E0. As in [BMP1, Eq. (1.3)] this gives a surjection of analytic sheaves ICD → IBMP → 0 30 mvIN = {0} \v∈U ∨v∈U Ker(S∗ I − v1) = IN, (3.8) (3.9) and surjections on fibers Ker(S∗ I − v1) → IBMP(v) → 0. Thus, as soon as the dimensions of the fibers coincide the two sheaves will be analytically isomorphic. Since dim Ker(S∗ I − v1) ≥ dim IBMP(v) ≥ rankI, the sheaves ICD and IBMP are thus analytically isomorphic over B \ singC(IN). Remark 3.8 (Spanning eigenvectors). For an arbitrary graded submodule IN we have mvIN ⊂ IN ∩ mv(HN ⊗ E0). This implies for all open subsets U ⊂ B. Indeed, for φ ∈ Tv∈U mvIN we have φ(v) = 0 for all v ∈ U and so φ is zero on all of B by the identity theorem of holomorphic functions [Kaup1, §0.6]. We can rewrite (3.8) as which says that the eigenvectors of S∗ coinvariant, since typically no eigenvectors of S∗ Remark 3.9 (Reducing subspaces and subbundles). Let S be the shift on HN ⊗ E0 and let SI and SE be its compression to IN and EN. The equality I are of the form kv ⊗ ξ for some (v, ξ) ∈ B × E0. I span the whole space IN. This does not imply that IN is Ker(S∗ − w1) = Ker(S∗ I − w1) ⊕ Ker(S∗ E − w1), ∀w ∈ B \ {0}. (3.10) can fail dramatically. For instance, if IN ⊂ M is the closure of an ideal in A defining an analytic subvariety E of V then for v ∈ E we have dim Ker(S∗ − v1) = 1 = Ker(S∗ E − v1) while [BMP1, Cor. 2.12] dim Ker(S∗ − v1) = n − dimC E. This comes from the failure of IN to be Gleason solvable at v. We claim that Ker(S∗ I − v1) is a subspace of Ker(S∗ − v1) if and only if IN is reducing. To see this, define the matrix-valued kernels KI and KE by KI (v, w)ξ := (PI (kw ⊗ ξ))(v), KE(v, w)ξ := (PE(kw ⊗ ξ))(v). For all ξ in the subspace E(w) ⊂ E0 we have S∗KE(·, w)ξ = ¯wK(·, w)ξ. On the other hand, S∗ αKI (·, w)ξ = ( ¯wα + [S∗ (3.11) Therefore, while PI (kw ⊗ ξ) is in the kernel of S∗ I − ¯w1 for all ξ ∈ E0, it is not necessarily in the kernel of S∗ − ¯w1. Indeed, (3.11) shows that Ker(S∗ I − w1) ⊂ Ker(S∗ − w1) only if [S∗ α, PI ] = 0 for all α ∈ {1, . . . , n}. The latter is to say that IN is a reducing subspace under S. So (3.10), which says that we have a holomorphic direct sum OCD ⊗ E0 = ICD ⊕ ECD, holds if and only if IN is reducing. α, PI ])KI (·, w)ξ. 3.3 Extension and boundary values 3.3.1 Abel convergence: ς versus ςB Since HN has a reproducing vector kv at each v ∈ B we can associate in a standard fashion [Bere2] to each operator A ∈ B(HN) a function which we call the B-Berezin symbol of A. holomorphic in the first variable and antiholomorphic in the second variable, It can extended to a function on ¯B × B which is ςB(A)(v) := hkvAkvi, ∀v ∈ B, ςB(A)(v, w) = hkvAkwi hkvkwi , ∀v, w ∈ B, 31 but we will usually just consider the diagonal values. More generally, for A ∈ B(HN ⊗ E0) for some Hilbert space E0 we define ςB(A)(v) := (TrHN ⊗ id)(A(kvihkv ⊗ 1E0)), ∀v ∈ B, where kvihkv is the rank-1 projection onto the subspace Ckv. Thus ςB(A) is a B(E0)-valued function on B. In [Kara4] the Berezin symbol for the unit disk was used to prove a theorem of Abel, namely that if a sequence (am)m∈N0 of complex numbers is convergent to a ∈ C then a = lim r→1− Xm∈N0 amrm. Here we shall use Abel's theorem to show that the Berezin symbol map ς gives the boundary limits of the B-Berezin symbol map ςB when restricted to the algebra of grading-preserving Toeplitz operators with continuous symbol. We shall use some facts from [An6], namely that C0(M) is the norm closure of a union of subspaces ς (m)(B(Hm)) where ς (m) : B(Hm) → C0(M) ⊂ L2(M, ω) is the adjoint of the Toeplitz map ς (m) : L∞(M) → B(Hm). The characteristic property of ς (m) is that it maps normally ordered products of shift operators directly to their classical limits, ς (m)(pmSjS∗ kHm) = ZjZ ∗ k, whenever j = m = k. So if we express an operator A ∈ B(Hm) as a matrix (Aj,k)j=m=k in the frame (pmek)k=m, which is to say that we write A =Pj=m=k Aj,kSjS∗ kHm, then (3.12) ς (m)(A) = Xj=m=k Aj,kZjZ ∗ k. We can extend ς (m)(A) to a U(1)-equivariant function on S, which takes the simple form ς (m)(A)(ζ) = Xj=m=k Aj,kζjζ∗ k, ∀ζ ∈ S. (3.13) We shall now compare ς (m)(A) with ςB(A). For all j, k ∈ F+ ςB(SjS∗ kpm)(v) = hS∗ j kvS∗ kpmkvi = (1 − v2)hΩΩivjv∗ = (1 − r2)r2mζjζ∗ k n (m) and v ∈ B we have k = (1 − v2)vjv∗ k where we write v = rζ with r ∈ (0, 1) and ζ ∈ S. So for a finite-rank grading preserving operator A ∈ B(Hm) ⊂ B(HN) we get, using the formula (3.12), that ςB(A)(rζ) = (1 − r2)r2mς (m)(A)(ζ), ∀rζ ∈ B. (3.14) In the following lemma we are not using the assumption that M is a coadjoint orbit. Lemma 3.10. Let A = (Am)m∈N0 be an operator in the Toeplitz core T (0) regard its symbol ς(A) ∈ C0(M) as a U(1)-equivariant function on S. Then r2mς (m)(Am)(ζ), ς(A)(ζ) = lim r→1− ςB(A)(rζ) = lim r→1− (1 − r2) Xm∈N0 H ⊂ Qm∈N0 B(Hm) and ∀ζ ∈ S. In the special case of a Toeplitz operator A = ς(f ) with f ∈ C0(M) we have fS(ζ) = ς(ς(f ))(ζ) = lim r→1− (1 − r2) Xm∈N0 r2mnm Xj=m=k ω(Z ∗ j Zkf )ζjζ∗ k, ∀ζ ∈ S. 32 Proof. Let ζ ∈ S be given. Since the sequence (ς (m)(Am))m∈N0 converges in the norm of C0(M) to the function ς(A) ∈ C0(M), the sequence (ς (m)(Am)(ζ))m∈N0 ∈ cb(N0) converges to the complex number ς(A)(ζ). By Abel's theorem, we get ς(A)(ζ) = lim m→∞ ς (m)(Am)(ζ) = lim r→1− (1 − r2) Xm∈N0 r2mς (m)(Am)(ζ). The proof is complete by the formula (3.14). Using the expression (3.1) for FS(Hm) we can rewrite the formula (3.12) for the symbol ς (m)(A) of an operator A ∈ B(Hm) as ς (m)(A)(x) = Tr(FS(Hm)(x)A), ∀x ∈ M. Since the Toeplitz map ς (m) is the adjoint of ς (m) with respect to ω and φm, we obtain for each f ∈ C0(M) the formula ς (m)(f ) = nm(ω ⊗ id)(FS(Hm)(f ⊗ pm)), (3.15) which will be useful later in the paper. Remark 3.11. Let H be an reproducing kernel Hilbert space of functions on a set B, and suppose that B sits inside a topological space and has nonempty topological boundary S = ∂B. Denote by kv the normalized reproducing kernel of H. Then H is standard if the sequence (kvm )m∈N converges weakly to zero for every sequence (vm)m∈N of points in B converging to a point in S. For the space H = HN we have hkvfi = (1 − v2)1/2f (v), ∀f ∈ HN, so it is clear that limv→1−hkvfi = 0 for all f ∈ A. Since A is dense in HN we get limv→1−hkvψi = 0 for all ψ ∈ HN. Thus the reproducing kernel Hilbert space HN is standard. It follows that every compact operator C on HN has B-Berezin symbol vanishing on the boundary, lim r→1− ςB(C)(rζ) = 0, ∀ζ ∈ S. For an operator A in TH we have A = ς(ς(A)) + C with C ∈ Γ0, and hence ς (m)(ς (m)(ς(A))). ς(A) = lim m→∞ ς (m)(Apm) = lim m→∞ since ς ◦ ς = id. Therefore lim r→1− ςB(A)(rζ) = lim r→1− ςB(ς(ς(A)))(rζ), ∀ζ ∈ S. Recall that the unique G-invariant state ω on C0(M) coincides with the limit ω = limm→∞ φm of the normalized traces φm : B(Hm) → C. Moreover, ω extends to the unique G-invariant state ωS on C0(S), which coincides with the normalized surface measure when S is regarded as the boundary of a domain B in Cn. Using these facts we have an L∞ version of Lemma 3.10: Proposition 3.12. Let A = (Am)m∈N0 be an operator in the L∞ Toeplitz core N ⊂Qm∈N0 B(Hm) and regard its symbol ς(A) ∈ L∞(M) as a U(1)-equivariant function on S. Then r2mς (m)(Am)(ζ) ς(A)(ζ) = lim r→1− ςB(A)(rζ) = lim r→1− (1 − r2) Xm∈N0 for ωS-almost all ζ ∈ S. We saw in Proposition 2.14 that the SOT-asymptotic Toeplitz symbol map ς SOT also coincides with ς when restricted to N . So: Corollary 3.13 (Asymptotic Toeplitz symbols versus boundary limits). The map coincides with ς SOT. N ∋ X → lim r→1− ςB(X)(r ·) ∈ L∞(M) 33 3.3.2 Extension of vector bundles Denote by C× = GL(1, C) the multiplicative group of nonzero complex numbers and consider the semigroup D× := {λ ∈ C× 0 < λ < 1}. As before we denote by U(1) the circle group (the unitary group of dimension 1), identified with the subgroup of C× consisting of complex numbers of modulus 1. In this subsection we work with an arbitrary smooth projective variety M ⊂ CPn−1. Let π : Cn \ {0} → CPn−1 be the natural surjection associated with the C×-action on Cn \ {0}, and set V := π−1(M) ∪ {0}, B := V ∩ Bn, S := ∂B = V ∩ S2n−1, where Bn ⊂ Cn is the unit ball and S2n−1 = ∂Bn is the unit sphere. The manifold M can be obtained by quoting out actions on the manifolds V \ {0}, B \ {0}, and S: M = (V \ {0})/C× = (B \ {0})/D× = S/ U(1). If E is an OM-module then its inverse image under π, denoted by π−1E and defined on open subsets U ⊂ V by is a π−1OM-module. The sheaf π∗E defined on V \ {0} by (π−1E)(U) := E(π(U)), π∗E := π−1E ⊗π−1OM OV\{0} is an OV\{0}-module called the pullback (or analytic inverse image) of E under π. coordinate ring A of M. The subsheaf The space of global holomorphic sections of the structure sheaf OV\{0} is isomorphic to the π−1OM ⊂ π∗OM = OV\{0} has no nonconstant global holomorphic sections. Denoting by (OV\{0})C× the structure sheaf OV\{0} we have the C×-invariant part of So for any OM-module E we have (π∗E)C× This gives (OV\{0})C× = π−1OM. = (π−1E ⊗π−1OM OV\{0})C× = π−1E. (π∗π∗E)C× := π∗(π∗E)C× = E. Definition 3.14. An OV\{0}-module EV\{0} is C×-equivariant if C× acts on EV\{0} compatibly with the OV\{0}-module structure on EV\{0} and the C×-action on OV\{0}. Since C× acts freely on V \ {0}, an OV\{0}-module EV\{0} is C×-equivariant if and only if EV\{0} equals π∗E for some OM-module E [KKT1, Prop. 4.2]. Lemma 3.15. A coherent analytic sheaf EB\{0} on B \ {0} is U(1)-equivariant if and only if there exists a coherent analytic sheaf E on M such that EB\{0} = π∗EB\{0}. In this case E is unique, and E is locally free iff EB\{0} is locally free. Proof. We use that the complex Lie group C× is the complexification of the compact Lie group U(1). The manifold V \ {0} is the "complexification" C× · S of the U(1)-space S in the sense of [Hein1]. As a special case of [HaHe1], we obtain that EB\{0} has a unique extension to a C×-equivariant coherent analytic sheaf EV\{0} on V \ {0}. Since (V \ {0})/C× = M, we have EV\{0} = π∗E for some coherent analytic sheaf E on M, as asserted. The uniqueness of E follows from the uniqueness of EV\{0}. The result about locally free sheaves is also in [HaHe1]. 34 Every C0 vector bundle over B admits a unique holomorphic structure by the Oka principle. However, this holomorphic structure is not D×-equivariant in general, since otherwise every C0 vector bundle over M would admit a holomorphic structure (which is not true). The above lemma says that D×-equivariance of the holomorphic structure is the same as U(1)-equivariance. Any U(1)-equivariant vector bundle EB on B\{0} therefore also admits a Hermitian metric which is the pullback of a Hermitian metric on the induced bundle E on M. Note however that EB also admits Hermitian metrics which are not D×-equivariant (even if they are U(1)-equivariant). The relevance of this discussion for the present paper is that the Cowen -- Douglas sheaf ECD of a graded quotients EN of HN ⊗ CN is easily seen to be U(1)-equivariant and so by Lemma 3.15 it descends to coherent sheaves M. It comes with a Hermitian metric, the Cowen -- Douglas metric, which is also U(1)-equivariant but not always D×-equivariant (i.e. not always a pullback of a metric on the induced bundle over M). For the reference space HN, even though CD(HN) is not D×-equivariant it defines a line bundle OCD which is isomorphic to the pullback of a line bundle on M, viz. OCD = OB\{0} = π∗OMB\{0}. 3.3.3 The reproducing kernel Let K Hn K(z, w) = Kw(z) be the reproducing kernel for the quotient module HN. projection of H 2 (z, w) = (1 − hw, zi)−1 be the reproducing kernel for the Drury -- Arveson space H 2 n onto the subspace HN then we get n and let If P is the orthogonal Kw(z) = (P K Hn w )(z) = hKzP K Hn w i = hK Hn z P K Hn w i. Recall that HN = span{Kw w ∈ B} ⊂ H 2 therefore get n. So P K Hn w = K Hn w for w ∈ B. For (z, w) ∈ B × B we Kw(z) = K Hn w (z) = (1 − hw, zi)−1. The orthogonal complement of HN is not invariant under the backward shifts S∗ Since we usually regard HN as a space of functions on the subset B ⊂ Bn, we see that the reproducing kernel for HN is just the restriction of the kernel for H 2 w however makes sense for all w ∈ Bn as a function on all of Bn. n, so it cannot be equal to span{Kv v ∈ Bn \ B}. Therefore P Kw can be nonzero also for w ∈ Bn \ B, and P Kw 6= Kw always happens for such w's. Θ gives a formula for the extension of Kw from B to the whole unit ball Bn, The Beurling factorization P = 1 − MΘM ∗ n. The projected kernel P K Hn 1 , . . . , S∗ Using M ∗ PE as defined in §3.3.1, ΘE (kv ⊗ ξ) = kv ⊗ ΘE(v)∗ξ we see that the numerator is precisely the B-Berezin symbol of K E(z, w) K(z, w) = 1 − ΘE(v)ΘE(w)∗ = ςB(PE)(v, w). The function ςB(PE)(v, w) has been studied for quotient modules of various reproducing kernel Hilbert spaces under the names "core function" and "defect function". It was observed in [Arv7b, Cheng1, Fang5, GRS1] that the boundary value of the restriction of the core function to the diagonal exists as an element of L∞(S) ⊗ MN (C) and is idempotent. This boundary function is sometimes called the "Arveson curvature function" of the quotient module. 35 Consider now the more general case of a vector-valued quotient module HE ΘE shows that the reproducing kernel K E N of HN ⊗ CN . The N has w (z) for HE Beurling representation of PE = 1− MΘE M ∗ the form Kw(z) = 1 − Θ(z)Θ(w)∗ 1 − hw, zi , ∀z, w ∈ Bn. K E(z, w) = 1 − ΘE(z)ΘE(w)∗ 1 − hw, zi1 ∈ MN (C). If PE is a projection over T (0) continuous vector bundle E on M. Applying Lemma 3.10 we see that H then P E := ς(PE) is a projection over C0(M) which thus defines a ςS(PE)(ζ) := lim r→1− ςB(PE)(rζ), ∀ζ ∈ S defines a continuous projection-valued function on S which is U(1)-equivariant and descends to M = S/ U(1) as the projection ς(PE) = lim m→∞ ς (m)(PE,m) ∈ C0(M) ⊗ MN (C). H . But ςB(PE) itself is not a projection in general. Note that ςB(PE) is always real-analytic. Only the continuity of its boundary value requires PE to be over T (0) Note also that ς (m)(PE,m) is a matrix over the image of B(Hm) under the symbol map ς (m), and therefore ς (m)(PE,m) is real-algebraic and in particular C0. This does not rely on PE having entries in T (0) H . Corollary 3.16. Let EN be a graded quotient module and let PE be the projection onto EN. Then the boundary value of the diagonal of the core function (i.e. the Arveson curvature function) of EN coincides with the covariant symbol ς(PE) = SOT− limm ς (m)(PE,m) of the projection PE. Remark 3.17 (Arveson curvature). Let E be the Serre sheaf of the graded quotient module EN. Its rank is by definition the leading coefficient of the Hilbert polynomial N ∋ m → χ(E(m)) = dim Em of the globally generated analytic sheaf E. The projection ς(PE) defines a sheaf of C0 M-modules with the same rank as E. Indeed one gets (ω ⊗ TrN )(ς(PE)) = lim m→∞ (φm ⊗ TrN )(PE,m) = lim m→∞ dim Em nm = rankE, (3.16) The integer rankE coincides with the "Arveson curvature" of the pure row contraction SE; see [Arv7a, Arv7b, GRS2]. 3.4 The Cowen -- Douglas projection Let EN ⊂ HN ⊗ CN be a quotient module. Definition 3.18. The Cowen -- Douglas projection of EN is the projection CD(EN) ∈ L∞(B \ {0}) ⊗ B(EN) defined by CD(EN)(v) := projection onto Ker(S∗ E − v1), ∀v ∈ B \ {0}. The sheaf ECD is thus locally free over B \ {0} if and only if CD(EN) belongs to the subalgebra C0(B \ {0}) ⊗ B(EN), in which case one could regard CD(EN) as a continuous (in fact real-analytic) map from B \ {0} into the Grassmannian of r-planes in EN, where r := rankECD. Then ECD is the vector bundle obtained by pulling back the universal rank-r vector bundle over the Grassmannian using the map CD(EN), and the Hermitian metric on ECD is obtained by pulling back the universal metric on the universal bundle via CD(EN). Recall that Ker(S∗ E − v1) = Ckv ⊗ ECD(v) where ECD(v) is a rankECD-dimensional subspace of CN for each v ∈ B \ {0}. Therefore we have a factorization CD(EN)(v) = CD(HN)(v) ⊗ ΠE(v) where ΠE(v) ∈ MN (C) is a projection onto the subspace ECD(v) ⊂ CN . The dimension of ECD(v) is ≥ rankECD. Denote by PE the projection of HN ⊗ E0 onto EN. Note that CD(EN) equals CD(HN ⊗ E0) ∧ PE, where for two projections P and Q acting on the same Hilbert space we denote by P ∧ Q 36 Similarly, for each m ∈ N0 we can define In this subsection we shall look at the geometric meaning of CD(Em) and CD(EN) and their relation to the symbols ς (m)(PE,m) and ςB(PE). CD(Em) := CD(Hm ⊗ E0) ∧ PE,m. (3.17) 3.4.1 Some alternating projections Recall [Halm1, Problem 122] that if P and Q are two projections acting on a Hilbert space then powers of the compression P QP converges to the infimum P ∧ Q of P and Q, P ∧ Q = lim p→∞ (P QP )p. This is useful for us because then we can compare the infimum P ∧ Q with the compression P QP . And for our choices of P and Q the compression P QP is going to equal CD(Hm) ⊗ ς (m)(PE,m): Proposition 3.19. Define a projection P E m ∈ L∞(M) ⊗ MN (C) by writing Then for each x ∈ M we have CD(Em) = CD(Hm) ⊗ P E m . P E m(x) = lim p→∞ ς (m)(PE,m)(x)p (3.18) x ⊗ CN . Note that Ck(m) in the norm of MN (C). Proof. Fix x ∈ M and let A := CD(Hm ⊗ CN )(x)PE,m CD(Hm ⊗ CN ) be the compression of PE,m to the subspace Ck(m) x ⊗ CN has finite dimension N . With that in mind we get from [Deut1, Lemma 9.38] that the positive operators Ap converge in norm to CD(Hm⊗CN )(x)∧PE,m as p goes to inifnity. Observe that CD(Hm⊗CN )(x)PE,m CD(Hm⊗CN )(x) = k(m) The proof is then complete, since by definition we have CD(Em) = CD(Hm ⊗ CN ) ∧ PE,m. Remark 3.20 (Another compression). Instead of the compression CD(Hm ⊗ E0)(x)PE,m CD(Hm ⊗ E0)(x) we can instead look at PE,m CD(Hm ⊗ E0)(x)PE,m. The powers of this operator also converge to CD(Em)(x). For all rζ ∈ B we have i = CD(Hm)(x)⊗ς (m)(PE,m)(x). PEk(m) ⊗hk(m) x ihk(m) x x x PE CD(HN)(rζ)HE N N = PE(krζihkrζ ⊗ 1N )HE = (1 − r2)PE Xm∈N0 = (1 − r2) Xm∈N0 = (1 − r2) Xm∈N0 r2m Xj=m=k r2m Xj=m=k r2m Xj=m=k ζjζ∗ k(SjS∗ k ⊗ 1N )(cid:12)(cid:12)HE N ζjζ∗ kSE,jS∗ E,k ς (m)(SjS∗ k)(ζ)SE,jS∗ E,k ∈ B(EN). Let us apply ω to elements of L∞(M)⊗B(HN⊗ CN ), producing elements of B(HN⊗ CN ). Then clearly ω(PE,m CD(Hm ⊗ CN )PE,m) = PE,mω(CD(Hm ⊗ CN ))PE,m = We have ς (m)ς (m) = id so (ς (m)ς (m))(PE,m) = PE,m holds trivially. Therefore 1 nm PE,m. nmω(CD(Hm ⊗ CN )PE,m CD(Hm ⊗ CN )) = nmω(ς (m)(PE,m) ⊗ FS(Hm)) = PE,m = nmω(PE,m CD(Hm ⊗ CN )PE,m). That is, the two choices of compressions have the same ω-integrals. 37 Lemma 3.21. Let EN ⊂ HN ⊗ CN be a graded quotient module and suppose that the projection PE onto EN has entries in T (0) H . Then CD(Em) is C0 for all m ≫ 0. That is, limp→∞ ς (m)(PE,m)p exists in C0(M) for all m ≫ 0. Proof. Since kς (m)(PE,m)k ≤ kPE,mk = 1, we can write ς (m)(PE,m) = P E positive operator of norm kCE m where CE m ⊕ CE m is a Moreover, mk ≤ 1. PE ∈ T (0) H ⊗ B(E0) ⇐⇒ lim m→∞ CE m = 0, H . Then for m ≫ 0 we have kς(PE) − P E m 's to ς(PE) iff PE is over T (0) i.e. we have uniform convergence of the P E over T (0) unitarily equivalent for all x ∈ M [Ols, Prop. 5.2.6]; thus the rank of P E rankE, which is to say that P E limp→∞(CE m)p = 0 uniformly by Dini's theorem. m is continuous. And CE m is C0 iff P E m is C0. If CE H . So assume that PE is m(x) are m is constantly equal to m is C0 then we have mk < 1. This gives that ς(PE)(x) and P E 3.4.2 The boundary limit of ΠE We shall now investigate how far ς (m)(PE,m) is from a projection in the case PE is Ψ-superharmonic. Define a projection ΠE ∈ L∞(B \ {0}) ⊗ MN (C) by writing CD(EN) = CD(HN) ⊗ ΠE. In the same way as Proposition 3.19 one deduces that for each v ∈ B we have ΠE(v) = lim p→∞ ςB(PE)(v)p. (3.19) Proposition 3.22. For almost every ζ ∈ S we have m ([ζ]) = lim m→∞ ΠE(rζ) = lim m→∞ P E lim r→1− ς (m)(PE,m)([ζ]) = ς(PE)([ζ]). The function limr→1− ΠE(rζ) descends to a projection over L∞(M) which concides with ς(PE). If PE is over T (0) Proof. Since limr→1− ςB(PE)(rζ) is a projection ς(PE)([ζ]) for almost every ζ ∈ S we have from (3.19) that H then limr→1− ΠE(r ·) coincides with ς(PE) as a projection over C0(M). lim r→1− (ΠE − ςB(PE))(rζ) = 0. Similarly, since P E equals limm→∞ ς (m)(PE,m) = ς(PE) as element of L∞(M) ⊗ MN (C). m is the limit of the powers of ς (m)(PE,m) we have that limm→∞ P E m exist and Consider the vector space also characterized by Em(v) := {ξ ∈ E0 k(m) v ⊗ ξ ∈ Em}, Ck(m) v ⊗ Em(v) = CD(Em)(v)Em = (Ck(m) v ⊗ E0) ∩ Em, i.e. Em(v) = Ran P E m ([v]). The grading on EN gives k(m) v ⊗ E(v) ⊂ Em so we have We have equality E(v) = Em(v) iff the dimensions are equal. So when P E ΠE, and they coincide. m is continuous then so is E(v) ⊂ Em(v). Note that k(m) 0 ⊗ ξ = pm(1 ⊗ ξ) = 0 for all m 6= 0 and all ξ ∈ E0, so Em(0) = E0 holds for all m. We typically only consider P E m for v 6= 0 however. 38 Theorem 3.23. Suppose that P E := ς(PE) is C0. Then P E is real-analytic and ECD is locally free on B \ {0}. In fact, for large enough m we have P E m = P E, and ΠE is the pullback of P E to B \ {0}, ΠE(rζ) = P E([ζ]). Proof. The coinvariance property ιm,l(PE,m) ≥ PE,l gives that ς (m)(PE,m) = (ιm,l(PE,m))l≥m + Γ0 = P E + CE m m ∈ L∞(M) ⊗ B(E0) with kCE for some CE kς (m)(X)k ≤ kXk for all X). By assumption we have norm-convergence limm→∞ CE is strictly less that 1 for m large. So for m ≫ 0, mk ≤ 1 and P ECE m = 0 = CE mP E (here we use that m = 0 so kCE mk P E m := lim p→∞ ς (m)(PE,m)p = P E, m is C0. As remarked before the lemma, when P E and thus P E m = ΠE. So ΠE is C0 and for v 6= 0 the projection ΠE(v) depends only on the coset [v] ∈ M = (B × {0})/D×. Since limr→1− ΠE(rζ) = P E([ζ]), we obtain ΠE(rζ) = P E([ζ]) for all ζ. Since ΠE is automatically real-analytic when C0 we see that the same is true for P E. m is continuous we have P E Corollary 3.24. If ς(PE) is C0 then we have a factorization of Hermitian vector bundles ECD = OCD ⊗ EB\{0} where EB\{0} is the Hermitian vector bundle over B \ {0} defined by the pullback of ς(PE) to B \ {0}. In this corollary, the holomorphic structure on ECD gives a holomorphic structure on EB\{0}. And since ςB(PE) is U(1)-equivariant, so is ΠE = limp→∞ ςB(PE)p and hence the holomorphic structure on EB\{0} is U(1)-equivariant. By Lemma 3.15 this means that there is a unique holomorphic structure on the smooth vector bundle E defined by ς(PE) which pulls back to that of EB\{0}. Remark 3.25 (Characteristic function isometric a.e. on S). Recall the Beurling factorization PE = 1 − Θ∗ EΘE where ΘE is the characteristic function of the pure finite-rank row contraction SE. It is shown in [GRS2, Thm. 4.3] and [BhSa1, Thm. 6.1] that ΘE becomes a partial isometry a.e. at the boundary S. Thus Proposition 3.22 is not surprising. Indeed, every element of the tensor algebra AH has a continuous extension to S and so it is easy to see that if PE has entries in TH = spanAHA∗ (i.e. when ς(PE) is continuous) then ΘE(rζ) becomes a partial isometry as r → 1− for every point ζ on S. H Note that Ran ς (m)(PE,m) is not equal to P E m , and limm→∞ Ran ς (m)(PE,m) need not equal ς(PE). Similarly, the boundary limit of Ran ςB(PE) need not equal ς(PE). We shall see in later sections that even if ς(PE) is continuous and even if we assume ECD to be locally free on B \ {0}, the vector bundle over M defined by ς(PE) need not be C0-isomorphic to ECD when pulled back to B \ {0}. 3.4.3 Geometric interpretation of Em Let E be a holomorphic vector bundle over M. Let m be large enough so that E(m) is globally generated and let Em be H 0(M;E(m)) endowed with some inner product. In §3.1.1 we defined a projection FS(Em) in C0(M) ⊗ B(Em) geometrically. Let us now give a more algebraic definition. Since Em is a Hilbert space we have a standard C0(M)-valued inner product on the C0(M)-module C0(M) ⊗ Em, defined on simple tensors by (f ⊗ ξg ⊗ η)C 0(M)⊗Em := f ∗ghξηiE0 , ∀f, g ∈ C0(M), ξ, η ∈ Em. Equivalently, this means that (φψ)C 0(M)⊗Em (x) = hφ(x)ψ(x)iEm , ∀φ, ψ ∈ C0(M) ⊗ Em. (3.20) 39 FS(Em)(x) = Xj,k∈J hψj(x)ψk(x)iEmψjihψk ∈ B(Em). (3.21) Definition 3.26. Define FS(Em) to be the projection acting on C0(M)⊗Em whose range is isomorphic as C0(M)-module to Γ0(M;E(m)) and the subspace FS(Em)(1 ⊗ Em) ⊂ Γ0(M;E(m)) identifies with H 0(M;E(m)). This uniquely determines FS(Em). Indeed, we obtain ψ(x) = FS(Em)(x)ψ, ∀ψ ∈ Em, x ∈ M. Therefore, if (ψj)j∈J is any Parseval frame for the Hilbert space Em then we can expand FS(Em)(x) as By (3.20) this is equivalent to saying that any Parseval frame (ψj)j∈J for Em is a Parseval C∗-frame for FS(Em). This property thus characterizes FS(Em). Note that FS(Em) exists precisely when E(m) is globally generated (as we assume here) because that is when there exists a holomorphic C∗-frame for Γ0(M;E(m)). Proposition 3.27. Let E be aglobally generated holomorphic vector bundle over M, let E0 be an inner product on H 0(M;E) and let FS(E0) be the projection over C0(M) with image Γ0(M; E) and with a Parseval C∗-frame given by an orthonormal basis for E0. Assume that m is large enough so that E(m) is Castelnuovo -- Mumford regular (see (3.4)). Let Em be H 0(M;E(m)) endowed with the inner product of Hm ⊗ E0. Then FS(Hm) ⊗ FS(E0) ∈ C0(M) ⊗ B(Hm ⊗ E0) can be regarded as an element of C0(M) ⊗ B(Em) and it has a Parseval C∗-frame given by an orthonormal basis for Em, i.e. FS(Hm) ⊗ FS(E0) = FS(Em). Proof. We can dispense with the tensor products in the C0(M)-factor of FS(Hm)⊗ FS(E0), and since ψZk is a holomorphic section of E(m) for all k ∈ F+ n (m) and ψ ∈ E0 we obtain that FS(Hm)⊗ FS(E0) is in C0(M) ⊗ B(Em) and has a Parseval C∗-frame ψ obtain by applying the multiplication map to the tensor product of the Parseval C∗-frames for FS(E0) and FS(Hm). Since E(m) is assumed to be Castelnuovo -- Mumford regular we have that Em is the image of the multiplication map Hm⊗E0 → Em. If PE,m is the projection of Hm ⊗ E0 onto Em then ψ is the image under PE,m of the Parseval C∗- frame for FS(Hm) ⊗ FS(E0), which is a Parseval frame for Hm ⊗ E0. Hence ψ is a Parseval frame for Em. Hence FS(Hm)⊗ FS(E0) has a Parseval C∗-frame given by a Parseval frame for Em (and this gives a natural representation of FS(Hm) ⊗ FS(E0) as a matrix over C0(M) of size N nm). Therefore any Parseval frame for Em gives a Parseval C∗-frame FS(Hm) ⊗ FS(E0), i.e. FS(Hm) ⊗ FS(E0) coincides with FS(Em). The following gives the geometric meaning of the Fock inner product Em on H 0(M;E(m)): Corollary 3.28. Let E be a holomorphic vector bundle over M and let EN be the completion of H 0(M;E(m)) when represented as a quotient of A ⊗ CN for some N . Then for l ≥ m ≫ 0 Lm∈N0 we have FS(El) = FS(Em) ⊗ FS(Hl−m). 3.5 Nullstellensatz Definition 3.29. Let E be a coherent analytic sheaf over M. The affine linear space of E is H 0(M;E(m)). the holomorphic linear space E of the Serre sheaf EV of the A-module EN := Lm∈N0 Typically we consider only the restriction of E to B ⊂ V and call this also the affine linear space of E. Recall that E and EB determine each other up to natural isomorphisms (see [Fisc1]). Since every coherent analytic sheaf over B is globally generated, E always appears as a holomorphic linear subspace E ⊂ B × CN 40 for some integer N . If E is generated by holomophic sections globally over M then the embedding E ⊂ B× CN is D×-equivariant outside 0 ∈ B and hence descends to an embedding E ⊂ M× CN where E now denotes the holomorphic linear space of E. Lemma 3.30. Let E be a coherent analytic sheaf over M and let E ⊂ B × CN be its affine linear space. Consider the submodule J(E) := {f ∈ A ⊗ CN hf (v)ξiCN = 0 for all (v, ξ) ∈ E} and the quotient Hilbert module E := span{kv ⊗ ξ (v, ξ) ∈ E}. Suppose that E is locally free and moreover that the Cowen -- Douglas sheaf ECD of E is locally free on B \ {0}. Then the following are equivalent: (a) J(E) is a graded submodule. (b) E is a graded quotient Hilbert module. v ⊗ ξ (v, ξ) ∈ E} for all m ∈ N0. (c) E ∩ (Hm ⊗ E0) = span{k(m) (d) E is globally generated (so we can take the embedding E ⊂ B × CN to be D×-equivariant). We thus see that E can be graded even if EN :=Lm H 0(M;E(m)) does not fit into a short exact sequence 0 → IN → A ⊗ CN → EN → 0 of graded A-modules (as the latter is slightly stronger than (d)). Proof. Let f ∈ A ⊗ CN and (v, ξ) ∈ B × CN . Observe that So we can describe J(E) as hf (v)ξiCN = hkv ⊗ ξfiHN⊗CN . J(E) = {f ∈ A ⊗ CN hkv ⊗ ξfiHN⊗CN = 0 for all (v, ξ) ∈ E}. In other words, (HN ⊗ CN ) ⊖ J(E) = E := span{kv ⊗ ξ (v, ξ) ∈ E}. Hence (a) is equivalent to (b). The quotient module E is graded if and only if (pm ⊗ 1E0 )E ⊂ E, Since (pm ⊗ 1E0)(kv ⊗ ξ) is a scalar multiple of k(m) ∀m ∈ N0. v ⊗ ξ we can write this as {k(m) v ⊗ ξ (v, ξ) ∈ E} ⊂ E, ∀m ∈ N0. (3.22) v ⊗ ξ (v, ξ) ∈ E, m ∈ N0} = E. Therefore (3.22) and (c) are equivalent. Clearly span{k(m) Suppose that E(v) = E([v]). Then the Cowen -- Douglas projection CD(E) = CD(HN)⊗ΠE satisfies ΠE(v) = ΠE([v]). So ΠE is the pullback of its boundary limit ς(PE) and PE must be over N , i.e. the projection PE onto E must preserve the grading. This gives the lemma. Theorem 3.31 (Nullstellensatz). Let E be a coherent analytic sheaf over M = (B \ {0})/D× and let E ⊂ B × CN be its affine linear space Let J(E) := {f ∈ A ⊗ CN hf (v)ξiCN = 0 for all (v, ξ) ∈ E} and for any subset J ⊂ A ⊗ CN define V(J) := {(v, ξ) ∈ B ⊗ CN hf (v)ξiCN = 0 for all f ∈ J}. 41 Let IN be the kernel of the surjection of A ⊗ CN onto EN := Lm H 0(M;E(m)). Finally let [EN] be the closure of in the Fock inner product of HN ⊗ CN , define E := span{kv ⊗ ξ (v, ξ) ∈ E}. and let ECD be the Cowen -- Douglas sheaf of E. Then for m ≫ 0 we have Em = E ∩ (Hm ⊗ E0) and Im = J(E) ∩ (Hm ⊗ E0), and the following are equivalent: (a) EB ∼= ECD (b) V(J(E)) = E If E is globally generated then J(E) = JN(E) is a graded submodule and E = EN is a graded quotient module. Finally, suppose that the A-module maps in the short exact sequence 0 → IN → A ⊗ CN → EN → 0 preserve the grading, and that E is locally free. Then (a) and (b) hold, and we have an identification [EN] = EN and IN = JN(E), of graded A-modules. Proof. We first show that the C-linear span of the k(m) x ⊗ ξ's with (x, ξ) ∈ E is precisely the vector space Em for m ≫ 0. That would give [EN] = E (up to finite-dimensional vector spaces). Clearly k(m) x ⊗ ξ for (x, ξ) ∈ E is mapped to Em under the multiplication map Am ⊗ CN → Em. In the proof of Lemma 3.30 we saw that E∩ (Hm ⊗ E0) equals span{k(m) v ⊗ ξ (v, ξ) ∈ E} for m ≫ 0. So for m ≫ 0 the map Am ⊗ CN → Em is surjective and we have indeed Em = Em. We have V(J(E)) = ECD := {(v, ξ) ∈ B ⊗ CN S∗ E(kv ⊗ ξ) = ¯vkv ⊗ ξ}, and E ⊂ ECD. Note that the fiber of ECD is given by kv ⊗ ECD. So the two conditions V(J(E)) = E and EB ∼= ECD are equivalent. That (a) holds when E is locally free follows from Proposition 3.6. The last statements follow from Lemma 3.30. Remark 3.32 (Grading and D×-equivariance). In general it is not clear if ECD is D×-equivariant iff E is. If the inclusion E ⊂ ECD is proper then E(v) = E([v]) need not imply ΠE(v) = ΠE([v]). That is, ΠE might fail to be D×-equivariant and hence not equal the pullback of ς(PE). The linear space E is by definition locally over an open subset U ⊂ B the kernel of an M -tuple of holomorphic functions for some integer M ≥ 1. But the Nullstellensatz above shows that if we take M = +∞ then in this algebraic setting we can describe ECD globally as a zero-set when E is locally free: Corollary 3.33. Suppose that ECD = OCD ⊗ EB\{0} where EB\{0} is the pullback of a holomorphic vector bundle E over M and let ΘE : B × ℓ2(N0) → B × CN be any multiplier with Ker M ∗ ΘE = span{kv ⊗ ξ (v, ξ) ∈ E}. Then E(v) = Ker Θ∗ E(¯v). 4 The Toeplitz part of a superharmonic projection 4.1 Lifts of projections We are now going to investigate the possibility of quantizing smooth Hermitian vector bundles over M = G/K. From the results of [Hawk1] and [Wang1, Wang2] we expect that a Hermitian metric on a vector bundle can be quantized in a stronger sense if the vector bundle is G-equivariant or satisfies some kind of stability condition. 42 Recall the Toeplitz short exact sequence 0 → Γ0 → T (0) H ς→ C0(M) → 0 (4.1) where T (0) H is the C∗-algebra of Toeplitz operators with symbol in C0(M) acting on the Fock space HN and Γ0 is the ideal in T (0) H consisting of compact operators which preserve the grading on HN. By the Swan theorem, a C0 vector bundle E over M is the same datum as an idempotent P E over C0(M), which if taken selfadjoint also defines a Hermitian metric on E. The symbol map ς was discussed extensively in the last two sections. The Toeplitz map ς : C0(M) → T (0) gives a positive linear splitting of (4.1). Following ideas of noncommutative geometry, and in particular [Hawk1, Hawk2], a quantization of P E is a projection PE over T (0) H which is equal to ς(P E) modulo Γ0, i.e. such that H ς(PE) = P E. As we have seen, there is a special class of projections over T (0) H whose ranges are quotient modules, namely the Ψ-superharmonic projections with symbol in C0(M). A natural question is thus whether our given P E admits a Ψ-superharmonic quantization. We shall see that the answer is "no" in general. Yet for any P E there is a natural candidate to a Ψ-superharmonic quantization. Indeed, if P E belongs to C0(M) ⊗ MN (C) then the Toeplitz operator ς(P E) acts on HN ⊗ CN and the range projection of ς(P E) is Ψ-superharmonic since ς(P E) is Ψ-superharmonic (indeed Ψ-harmonic). If we let Ran ς(P E) denote the range projection of ς(P E), we would thus like to know if Ran ς(P E) is a quantization of P E. As mentioned, in general Ran ς(P E) is not a quantization of P E. Let us look at the basic operator aspects of this lifting problem. Since ς(ς(P E)) = P E, we know that ς(P E) is a projection modulo Γ0 or, what is the same since ς(P E) preserves the grading on HN ⊗ CN , we know that ς(P E) is a projection modulo compact operators. By Brown -- Douglas -- Fillmore theory [Davi2, §IX], every projection modulo compacts is of the form normal plus compact. But recall that even more is true: every projection modulo compacts is projection plus compact (this is a particular case of [Olse1]; see a quick proof in [Weav1, Prop. 3.1]). So there is a projection QE acting on HN ⊗ CN such that QE = ς(P E) + CE with CE compact. Since T (0) exists a lift of P E to a projection over T (0) H = ς(C0(M)) + Γ0, we see that QE belongs to T (0) H . So there always of the same matrix size, as can also be shown by operator-algebraic reasoning applied to the sequence (4.1). H However, a projection modulo compacts is rarely equal to its range projection modulo compacts. For a simple example, let K be a compact operator. Then K is equal to zero modulo compacts (hence K is a projection modulo compacts) but the range projection PK of K need not be of finite rank (it is not unless K is of finite rank, by definition) so PK is not zero modulo compact in general (a projection is compact iff it has finite rank). Thus, for the range projection of ς(P E), in general we have Still, since ς(P E) has entries in T (0) von Neumann algebra. In fact: P E = ς(ς(P E)) 6= ς(Ran ς(P E)). H ⊂ N we know that Ran ς(P E) has entries in N , because N is a Proposition 4.1. Let P E be a projection over C0(M). Then Ran ς(P E) has entries in T (0) H ⊂ N . Proof. As observed in [DRS1, Thm. 10.4], the proof of [Arv6c, Lemma 1.13]generalizes so as to show that the C∗-algebra TH is the C∗-envelope of the operator system generated by S. This gives that TH is an injective C∗-algebra, hence an AW ∗-algebra, hence every element of TH has its range projection belonging to TH (see [SaWr1, Lemma 2.1.5]). 43 Thus the range projection of ς(P E) is a coinvariant projection over T (0) H , hence with symbol in C0(M), but in order to obtain Ran ς(P E) from ς(P E) one may have to do more than just adding compacts. Given a coinvariant projection PE, there is a unique projection P E over L∞(M) such that the Toeplitz operator ς(P E) is the Ψ-harmonic part of PE. Indeed, it is given by P E = ς(PE). Let us now discuss uniqueness of a coinvariant lift: Proposition 4.2 (Uniqueness of superharmonic lift). Suppose that P E is a projection over C0(M) with a coinvariant lift PE. Then Ran ς(P E) is also a coinvariant lift of PE. In fact, PE equals Ran ς(P E) up to finite-rank operators. Proof. The assumption ς(PE) = P E gives PE = ς(P E) + CE with a pure Ψ-superharmonic operator CE. Moreover, CE is compact since PE is over T (0) H . We have ς(P E) = limq→∞ Ψq(PE) ≤ Ψp(PE) ≤ PE for all p ∈ N, so ς(P E) preserves the range of PE, as does the pure superharmonic part CE of PE. Since ς(P E) equals PE modulo compacts, this N of PE. Thus ς(P E) is invertible gives that ς(P E) restricts to a Fredholm operator on the range HE modulo finite-rank operators as an operator on HE N , so that ς (m)(P E) is invertible as operator on HE m for large enough m. Thus, up to finite-rank operators, PE equals the range projection of ς (m)(P E). The proof of Proposition 4.2 breaks down if CE is not compact. Question 4.3. Can we drop the continuity assumption in Proposition 4.2? That is, if P E is a projection over L∞(M) with a coinvariant lift PE, is then Ran ς(P E) also a coinvariant lift of PE? Remark 4.4 (Continuous symbol). As for the uniqueness of Ran ς(P E) up to finite-rank operators as coinvariant lift of P E, we shall see that the continuity of P E is a necessary assumption. Here it is very important to distinguish between P E being continuous and P E having entries in the embedded subalgebra C0(M) ⊂ L∞(M), since an element of the latter is just the almost everywhere equivalence class of a continuous functions. The map ς : N → L∞(M) can take values in C0(M) even on elements that do not belong to T (0) H . Indeed, its values on ς(C0(M)) + Γω are in C0(M). Therefore we can have a lift and coinvariance of PE just says that PE = ς(P E) + CE where CE is merely pure Ψ-superharmonic. So the Ψ-superharmonic lift would not be unique. PE ∈ (ς(C0(M)) + Γω) ⊗ B(E0), 4.1.1 Nonexisting lifts If P E is a projection over C∞(M) defining a smooth vector bundle E which is not holomorphic, what is the geometric meaning of the graded quotient module Ran ς(P E)? What is the coherent OM-module ECD and how is it related to the Serre sheaf of the graded A-module underlying Ran ς(P E)? Note that ECD cannot be the pullback of E even as smooth vector bundle because then ECD would have to be locally free, contradicting the assumption that E does not admit a holomorphic structure. Indeed, M with E o a coherent OM-module then E is if E is a coherent C∞ locally free if and only if E o is locally free. To see this, note that we can define a ¯∂-operator on E by M with ¯∂ the operator on C∞(M) defined by the complex-analytic setting ¯∂E := 1 ⊗ ¯∂ on E o ⊗OM C∞ structure on M. Then ¯∂E is integrable, i.e. a holomorphic structure, since ¯∂ is. The kernel of ¯∂E, which is precisely E o, is locally free by the Koszul -- Malgrange theorem [KoMa1]. M -module of the form E = E o ⊗OM C∞ This observation and Theorem 3.23 give us: Corollary 4.5. Let P E be a projection over C∞(M) and suppose that the vector bundle defined by P E does not admit a holomorphic structure. Then and the vector bundles defined by ς(Ran ς(P E)) and P E are not topologically isomorphic. ς(Ran ς(P E)) 6= P E. 44 4.1.2 dim Ran ς (m)(P E) versus χ(E(m)) The projection onto any quotient module EN is of the form PE = ς(P E)+CE where P E is a projection over L∞(M) and ς(P E) ≤ PE. If we assume that CE is compact then the restriction of ς(P E) to EN is invertible modulo compact, which is the same as being invertible modulo finite-rank operators. So when CE is compact we have for m ≫ 0 that dim Ran ς (m)(P E) = Tr(PE,m) = χ(E(m)), where E is the Serre sheaf of the graded A-module underlying EN. So P E has a lift with "correct dimensions". But this P E was special since it was the symbol of a superharmonic projection. In general we can ask: Question 4.6. Let E be a smooth vector bundle over M and let P E be a smooth Hermitian metric on E. Does it follow in this generality that dim Ran ς (m)(P E) equals χ(E(m)) for m ≫ 0? We have a result in this direction: Proposition 4.7. Let E be a holomorphic vector bundle over M and suppose that P E is a real- analytic Hermitian metric on E with a Parseval C∗-frame given by a basis for H 0(M;E). Then dim Ran ς (m)(P E) = χ(E(m)) for m ≫ 0. Proof. For each m ≥ 0 we have that FS(Hm) ⊗ P E has a Parseval C∗-frame given by a basis for H 0(M;E(m)). Since ς (m)(P E) = nm(ω ⊗ id)(FS(Hm) ⊗ P E) (see (3.15)), this gives the result. We saw in the last section (Theorem 3.31) that we can associate a (graded) quotient module EN to every (globally generated) holomorphic vector bundle E over M with dim Em = χ(E(m)) for m ≫ 0. However, the symbol ς(PE) of this quotient module EN need not be continuous. Also, in the setting of Proposition 4.7 in general the projection PE := Ran ς(P E) has symbol and therefore the metric P E is of "less value" for quantization purposes. ς(PE) 6= P E, 4.2 Into Hardy space Let PE be a Ψ-superharmonic projection acting on HN ⊗ CN . As before we write PE uniquely as PE = ς(P E) + CE with P E a projection over L∞(M) and SOT− limm→∞ Ψm(CE) = 0. If we let T = (T1, . . . , Tn) be the Kraus operators of Ψ, i.e. we have Ψ(X) = Pn αXTα for all X ∈ B(HN ⊗ CN ), then T has the same invariant and coinvariant subspaces as the shift S. Therefore, if HE N is the range of PE then the restriction αEN = (PETαHE ∀α = 1, . . . , n T ∗ E,α := T ∗ α=1 T ∗ )∗, N preserves EN. From Ψ(PE) ≤ PE we get ΨE(1) := n Xα=1 T ∗ E,αTE,α ≤ PE where PE is now playing the role of identity operator on HE on HE N . The limit N . That is, TE is a spherical contraction ATE := SOT− lim m→∞ Ψm E (1) therefore exists, and indeed we see that ATE = ς(P E)EN . 45 Note that ATE and ς(P E) are practically the same since ς(P E) acts by zero outside HE N . If we assume that CE is compact then HE N coincides with Ran ς(P E) up to finite-dimensional subspaces, and ATE is a Fredholm operator in the sense that it is a positive operator with finite-dimensional kernel and cokernel. The "asymptotic limit" ATE of a contraction TE has been widely studied in the case (n = 1) of a single operator [Gehe1, Gehe2, Kerc10, Kubr1] but occasionally also for tuples [Pop7]. Using ATE one changes the inner product on the Hilbert space to make TE an isometry, provided that ATE is invertible. Here we shall use ATE for this purpose and moreover find the geometric meaning of It turns out that, if P E defines a vector bundle, the rather nonstandard the new inner product. quantization HE • transforms via ATE to the more familiar quantization taking place on the Hardy space of P E. 4.2.1 Subnormality with algebraic relations Recall that S ⊂ S2n−1 is the principal U(1)-bundle over M ⊂ CPn−1 associated with the hyperplane bundle restricted to M. Let IM be the ideal in C[z1, . . . , zn] which defines H•. In other words, IM is the ideal such that the homogeneous coordinate ring of M is given by A = C[z1, . . . , zn]/IM. Definition 4.8. An n-tuple T = (T1, . . . , Tn) of commuting operators on a Hilbert space is H•- subnormal if T is jointly subnormal with minimal normal extension M = (M1, . . . , Mn) satisfying the relations of the ideal IM in the sense that In this case, T is called an S-isometry if f (M ) = 0, ∀f ∈ IM. σ(M ) ⊂ S. S-isometry is a spherical isometry. By Athavale's theorem [Atha3, Prop. 2], an S2n−1-isometry is the same as a commutative spher- αTα = 1. So every ical isometry, i.e. an operator tuple T of commuting operators with Pn The following was inspired by [Feld1, Thm. 2.1] and [Pop7, §2]: Lemma 4.9. Let T = (T1, . . . , Tn) be a tuple of commuting operators on a Hilbert space H. Then T is an S-isometry if and only if there exists a unital completely positive map α=1 T ∗ : C0(S) → B(H) with (Zα) = Tα and (Z ∗ αZβ) = T ∗ αTβ, ∀α, β ∈ {1, . . . , n}. (4.2) Proof. Suppose that such a map exists. As for any unital completely positive map, we have a Stinespring representation π : C0(S) → B(H) of , i.e. a ∗-homomorphism such that (f ) = V ∗ π(f )V, ∀f ∈ C0(S) with an isometry V : H → H into some Hilbert space H. Since π is a ∗-algebra homomorphism, the tuple M = (M1, . . . , Mn) defined by Mα := π(Zα), ∀α ∈ {1, . . . , n} consists of normal operators on H satisfying the relations of the ideal in C[z1, . . . , zn] which defines H•. Since π is a ∗-algebra homomorphism we also have σ(M ) ⊂ S. If we can show that the subspace V(H) ⊂ H is invariant under M1, . . . , Mn then the tuple T will be an S-isometry. For that we use the assumption (4.2). For each α ∈ {1, . . . , n} it says T ∗ αTα = V ∗ π(Z ∗ αZα)V = V ∗ M ∗ αMαV, 46 and if we denote by P the orthogonal projection of H onto V(H) then writing V ∗ M ∗ αMαV = P M ∗ αMαV(H) αTα + P M ∗ = T ∗ α(1 − P )MαV(H) we conclude that T ∗ αTα = T ∗ αTα + P M ∗ α(1 − P )MαV(H), so that (1 − P )MαV(H) = 0, i.e. V(H) is invariant under Mα, as desired. To prove the converse, note that C0(S) is generated by a commuting tuple Z = (Z1, . . . , Zn) of normal operators satisfying Pn αZα = 1 (sphere condition) and the relations of H• but no other relation. Therefore, if we suppose that T is an S-isometry on H, with minimal normal extension M thus satisfying the relations of H• and having spectrum in S, then there is a ∗-representation π of C0(S) with π(Zα) = Mα. We let V be the isometric embedding of H into the Hilbert space on which M acts, and we define (f ) := V ∗ M π(f )VM for all f ∈ C0(S). Then, since H is invariant under each Mα, we have the property (4.2). α=1 Z ∗ 4.2.2 Similarity to a spherical isometry In [An6] we showed that C0(S) can be identified with the "Cuntz -- Pimsner algebra" of the sub- product system H•, namely the C∗-algebra OH defined as the quotient of the Toeplitz algebra TH = C∗(S1, . . . , Sn) by the ideal of compact operators. This fact can be useful as one can of- ten adopt known constructions involving the Cuntz algebra On to a more general Cuntz -- Pimsner algebra. Here is an example: Corollary 4.10. Let HE Define a commutative operator n-tuple VE by N be a graded quotient module such that the positive operator ATE is invertible. Then VE is an S-isometry. VE,α := A1/2 TE TE,αA−1/2 TE , ∀α ∈ {1, . . . , n}. Proof. The tuple VE satisfies the same relations as does TE, since they are similar. We have E,αVE,β = A−1/2 V ∗ TE E,αATE TE,βA−1/2 T ∗ TE . Since ΨE(ATE ) :=Pα=1 T ∗ E,αATE TE,α = ATE , it is clear that VE is a spherical isometry. We need to show that the minimal normal extension of VE also satisfies the relations of H•. But we can construct as in [Pop7, Thm. 2.3] a unital completely positive linear map E : OH → B(HE N ) with ∀α, β ∈ {1, . . . , n}. By Lemma 4.9 this is precisely the statement that VE is an S-isometry. E,αVE,β, αZβ) = V ∗ E(Z ∗ Of course, after observing that VE is a spherical isometry an application of Athavale's theorem immediately gives that VE is subnormal. What is important is however that we obtain a unital com- pletely positive linear map E : OH → B(HE N ) whose Stinespring dilation gives a ∗-representation πE of C0(S) = OH on a Hilbert space containing HE N as a subspace invariant under πE(Z1), . . . , πE(Zn). We shall see next that VE is unitarily equivalent to the multiplication tuple on the Hardy space associated to ς(PE) and ω, and that the normal dilation is the multiplication tuple on ambient L2-space. TE,αA−1/2 Remark 4.11. The commutative spherical isometry VE given as above by VE,α := A1/2 TE is not the spherical isometry WE appearing in the polar decomposition TE = WETE of the column operator TE : HE E (1). We shall need to use VE because ATE is a Toeplitz operator (i.e. ΨE(ATE ) = ATE ) while TE2 is not. Moreover, ATE is invertible (modulo finite-rank operators) under the natural assumption that ς(PE) is C0, while TE2 might not be so. N )⊕n. It is the distinction TE2 = ΨE(1) versus ATE = limm Ψm N → (HE TE 47 If E is a Hilbert space and A is a positive invertible operator on E, we denote by A−1/2E the vector space E endowed with the inner product hφψiA−1/2 E := hA−1/2φA−1/2ψiE, ∀φ, ψ ∈ E. We can then view A1/2 as a unitary operator from E to A−1/2E, When HE N is a graded quotient module such that ATE is invertible, we write KE N := A−1/2 TE HE N and dentote by A1/2 identified with A1/2 E : HE E TEA−1/2 N → KE E acting on KE N . N the associated unitary operator. The tuple VE will sometimes be N Identification of KE 4.2.3 In the following we denote by FS(H1)⊗(−m) the transpose of FS(H1)⊗m for each m ∈ N. Thus FS(H1)⊗(−m) is a projection over C∞(M) which defines the line bundle OM(−m). If P E is a projection over C∞(M) defining a smooth vector bundle E then the vector space Γ∞(S;ES, P E) of global sections of the pullback of E to S splits as C∞(S)-module into Γ∞(S;ES, P E) =Mk∈Z Γ∞(M;E(m), FS(H1)⊗k ⊗ P E). Theorem 4.12. Let P E be a projection over C0(M) defining a smooth vector bundle E and assume that ς(Ran ς(P E)) = P E. Then the limit operator ATE of the spherical contraction TE on the quotient module HE N into a subspace of the L2-space of P E and ω, N := Ran ς(P E) can be used to map HE KE N := A−1/2 N ⊂ L2(S, ω; P E) =Mk∈Z and there is a holomorphic structure on E such that KE (up to a finite-dimensional subspace), HE TE L2(ω; FS(H1)⊗k ⊗ P E), N identifies with the Hardy space of P E and ω KE N = H 0(S, ω; FS(HN) ⊗ P E) = Mm∈N0 H 0(M, ω; FS(Hm) ⊗ P E). . E E TEA−1/2 N is invariant under action of the generators Z1, . . . , Zn ∈ C0(S) acting in the multiplication N identifies with the So KE representation on L2(S, ω;E, P E) and the restriction of Z = (Z1, . . . , Zn) to KE S-isometry VE := A1/2 Proof. By assumption P E is an element of C0(M)⊗B(E0) for some finite-dimensional Hilbert space E0. We may identify E0 as a vector subspace of Γ0(M;E) via ψ(x) = P E(x)ψ for ψ ∈ E0. Then P E has a Parseval C∗-frame consisting of an orthonormal basis for E0, symbolically P E = FS(E0), but note that so far we have not shown that E0 consists of holomorphic sections of E for any holomorphic structure on E (we did not assume that E admits a holomorphic structure). The projection FS(Hm) ⊗ P E has a Parseval C∗-frame ψ = (ψ(m) )j∈J given by elements of Hm ⊗ E0, and we denote by Em the C-linear span in Hm ⊗ E0 of that C∗-frame ψ. The tensor product of the Parseval C∗-frames for FS(Hm) and P E is a Parseval frame for Hm ⊗ E0. Thus, if PE,m is the projection of Hm ⊗ E0 onto Em then ψ is the image under PE,m of a Parseval frame for Hm ⊗ E0. Hence ψ is a Parseval frame for Em [HaLa1, Example A]. So we have (4.3) j FS(Hm) ⊗ P E = FS(Em). Since FS(Em)(x) belongs to B(Em) ⊂ B(HN ⊗ E0) for each x ∈ M we have that nmω(FS(Em)) = nmω(FS(Hm) ⊗ P E) = ς (m)(P E) 48 is naturally an operator on Em. Moreover, ω(FS(Em)) is invertible on Em since the elements of ψ span Em by definition. Thus HE m := Ran ς (m)(P E) = Em for all m. N coming from compression of the shift on HN ⊗ E0. The multiplication map HN ⊗ E0 → EN makes EN a graded quotient module, and this is the same Let EN denote the graded vector space underlying EN but endowed with the inner product of as the action of A on HE L2(S, ω; P E) =Lk∈Z L2(ω; FS(H1)⊗k ⊗ P E). The actions of the generators Z1, . . . , Zn of C0(S) are the same on EN ⊂ L2(S, ω; P E) as on EN, given just by multiplication on sections of E over S (the adjoints of the Zα's act differently on EN compared to EN however). Therefore EN is an A-invariant subspace of L2(S, ω; P E) and the multiplication tuple on L2(S, ω; P E) is a normal extension of the shift tuple on EN. From the fact that Em generates Γ0(M;E(m)) a C0(M)- module for each m we obtain moreover that the normal extension is the minimal one. E,mς (m)(P E) ∈ B(Em) is the Gram matrix of ψ as frame for Em. This means that if we use the analysis operator of the frame ψ to identify Em and Em as vector spaces, the operator E,mς (m)(P E) is the frame operator of ψ as frame for Em (see [Bala1, Example 3.1.1]). Therefore c−1 E,mς (m)(P E)−1/2ψ will be a Parseval frame for Em. We know that ψ is a Parseval frame for Em. c−1 So the inner product on Em is obtained from that of Em by applying ς (m)(P E)−1/2. This gives Em = KE m. The operator c−1 The assumption that P E is continuous and ς(Ran ς(P E)) = P E implies that ATE is Fredholm. So the Hilbert space KE N is well-defined up to finite-dimensional subspaces. From Theorem 3.23 we have that HE Cowen -- Douglas projection CD(EN) = CD(HN) ⊗ P E identifies HE sections of the Cowen -- Douglas bundle. Since KE are also holomorphic sections. The proof is thus complete. N endows E with a canonical holomorphic structure. The N with a space of holomorphic N , the elements of KE N is the same vector space as HE N 4.3 Hidden Szego expansion When P E is a projection over C∞(M) we discussed the condition that Ran ς(P E) is a lift of P E. That is, the condition that Ran ς(P E) and ς(P E) differ by a compact operator. Let us now investigate the geometric meaning of this compact operator which is the obstruction to the idempotency of the Toeplitz operator ς(P E). E, SE] 4.3.1 Geometric meaning of [S ∗ Let EN ⊂ HN ⊗ E0 be a quotient module and suppose that the projection PE onto EN has entries in T (0) H . Then P E := ς(PE) defines a holomorphic vector bundle E over M with Hilbert polynomial given by χ(E(m)) = dim Em for large enough m (see §4.1.2). Using φm ◦ Ψl−m = φl we get from Hirzebruch -- Riemann -- Roch that (Tr ⊗ TrE0)((id −Φ∗)(PE,m)) = χ(E(m)) − χ(E(m + 1)) = md−1M(cid:0) trω ΘE + sω/2(cid:1)eω + O(md−2), (4.4) where ΘE is the curvature 2-form of the Chern connection of the metric P E and trω ΘE is its trace against the Kahler 2-form ω. Using the coinvariance of PE we can write n (id−Φ∗)(PE) = [PE, Sα]∗[PE, Sα] = [S∗ E, SE] + [S∗, S]. Xα=1 (4.5) We saw in §2.2 that φm([S∗, S]) = m−1 vol(M,L) M (sω/2)eω + O(m−2), 49 so from (4.4) and (4.5) we obtain φm([S∗ E, SE]) = m−1 vol(M,L) M trω ΘEeω + O(m−2), (4.6) When we discussed the spherical expansion S in §2.2 we observed that ς (m)(m[S∗, S]pm) ap- proximates the (constant) scalar curvature. Now (4.6) and the relation φm = ω ◦ ς (m) suggest that ς (m)(m[S∗ E, SE]pm) should play the role of trω ΘE. Since ς(P E) is the limit of the sequence (Ψp(PE))p∈N0 , and since Ψp = p Xq=0(cid:18)p q(cid:19)(id −Ψ)q, we see that the operator in (4.5) is a kind of first-order approximation to the compact operator CE := PE − ς(P E). In the next subsection we discuss the geometric meaning of CE, from which one could hint a relation between [S∗ E, SE] and trω ΘE. Interpretation of the hidden Szego expansion 4.3.2 Let E be a globally generated holomorphic vector bundle over M and let P E be a Hermitian metric on E. For each m ∈ N0 one can consider the "Szego endomorphism" of the Hilbert space H 0(ω, FS(Hm)⊗ P E), which is the C0(M)-linear endomorphism ΣE(m) of Γ0(M;E(m), FS(Hm)⊗P E) defined as follows. is an orthonormal basis for the Hilbert space H 0(ω, FS(Hm)⊗ P E) then one can If ψ = (ψj )j=1,...,nE view it as a sequence of elements in Γ0(M;E(m)), and this sequence ψ is a C∗-frame for the Hilbert C0(M)-module Γ0(M;E(m), FS(Hm) ⊗ P E) since we assume that E(m) is globally generated. The Szego endomorphism ΠE m is defined to be the frame operator of the C∗-frame ψ, m ΣE(m)φ := nE m Xj=1 (ψjφ)FS(Hm)⊗P E ψj, ∀φ ∈ Γ0(M;E(m)), where (ψj(x)φ(x))FS(Hm)⊗P E (x) = (ψj (x)φ(x))FS(Hm)⊗P E is the inner product on the fiber E(x) obtained from the projection FS(Hm) ⊗ P E. In §5.1 we will discuss the notion of "balanced" metrics. By definition, the Hermitian metric FS(Hm)⊗ P E is ω-balanced if ΣE(m) is the identity endomorphism. That is, if an orthonormal basis for the Hilbert space H 0(ω, FS(Hm) ⊗ P E) is at the same time a Parseval C∗-frame for the Hilbert C0(M)-module Γ0(M;E(m), FS(Hm)⊗ P E). The Szego expansion of P E (and the reference metrics P L and ω) is a large-m expansion of ΣE(m) which shows how the obstruction to FS(Hm)⊗ P E being balanced disappears as m grows (see e.g. [MaMa3]). We can also go in the opposite direction: we can start with a Parseval C∗-frame ψ for the Hilbert module Γ0(M;E(m), FS(Hm) ⊗ P E) given by a basis for the vector space H 0(M;E(m)) and we can ask whether it is an orthonormal basis for H 0(ω, FS(Hm)⊗ P E) or not. The obstruction is the frame operator of ψ regarded as a sequence of vectors in H 0(ω, FS(Hm) ⊗ P E). But the same information is contained in the Gram matrix of ψ regarded as a sequence of vectors in H 0(ω, FS(Hm) ⊗ P E). More precisely, from [Bala1] we have a ∗-algebra monomorphism from B(H 0(ω, FS(Hm) ⊗ P E)) to B(HN ⊗ CN ) sending the frame operator of ψ to the operator c−1 E,mς (m)(P E). Now, this Gram matrix is precisely χ(E(m)) rankE ω((ψψ)FS(Hm)⊗P E ) = c−1 E,mς (m)(P E). Thus a large-m expansion of the operators c−1 expansion" ("hidden", as it has not been studied so far). E,mς (m)(P E) will go under the name "hidden Szego 50 4.3.3 ς E(m)(A−1 E,m) expansion Under our assumption that PE is a superharmonic projection over T (0) H , the restriction ATE of ς(P E) to the range EN of PE is Fredholm. For present purposes we may assume ATE is invertible. If we regard A1/2 TE as a unitary operator A1/2 E : EN → KE N N := H 0(S, ω;E, P E) then the 1-isometry VE = A1/2 onto the Hardy space KE is precisely the multiplication tuple on KE N (recall Theorem 4.12). Each endomorphism f of E gives rise to an endomorphism of the pullback of E to S, and then to a grading-preserving multiplication operator on L2(S, ω;E, P E) =Lk∈Z L2(M, ω;E(m), FS(H1)⊗k⊗P E). Restricting such multiplication operators to the Hardy space and following them by compression back to KE N gives us grading-preserving Toeplitz operators, which we denote by E TEA−1/2 E ςVE (f ) = Xm∈N0 ς (m) VE (f ) ∈ B(KE N ) VE (f ) is the component of ςVE acting on the graded piece KE with f ∈ End Γ0(M;E, P E). Here ς (m) H 0(M, ω;E(m), FS(Hm) ⊗ P E). Lemma 4.13. The Toeplitz operators ςVE (f ) with f ∈ End Γ0(M;E, P E) are fixed-points of the unital map ΨVE (X) :=Pn N ). Moreover, for each f ∈ End Γ0(M;E, P E) we have (4.7) E,αXVE,α acting on X ∈ B(KE ς (m)(f )A−1/2 ςVE (f ) = A−1/2 α=1 V ∗ m = . E E Proof. Since VE is a commutative spherical isometry it is deduced as in [Prun1] that the Toeplitz operators ςVE (f ) are the fixed points of the map ΨVE . To deduce (4.7), note that n n ΨVE (X) := E,αXVE,α = A−1/2 V ∗ E E,αA1/2 T ∗ E XA1/2 E TE,αA−1/2 E = A−1/2 E ΨE(A1/2 E XA1/2 E )A−1/2 E , Xα=1 Xα=1 α=1 T ∗ where ΨE(Y ) := Pn E,αY TE,α. An operator Y is fixed under ΨE if and only if Y i a Toeplitz operator ς (m)(f ) with symbol f ∈ L∞(M) ⊗ B(E0) such that Y is zero outside EN. So X = A−1/2 is a fixed point of ΨVE for each f ∈ End Γ0(M;E, P E). ς (m)(f )A−1/2 E E Define a state ωE : End Γ0(M;E, P E) → C by specifying it on rank-1 endomorphisms to be ωE(ψ)(φ) := ω((φψ)P E ), ∀φ, ψ ∈ Γ0(M;E), 1 rankE where (··)P E is the C0(M)-valued inner product on Γ0(M;E, P E) and ψ)(φ is the rank-1 endomor- phism acting as ψ)(φψ′ := (φψ′)P E ψ for ψ′ ∈ Γ0(M;E, P E). Also, denote by φE m : B(KE m) → C the tracial state, be the adjoint of ς (m) VE Let ς (m) VE End Γ0(M;E), From (4.7) we have with respect to φE ω(ς (m) VE (X)f ) := φE φE m(X) := . Tr(X) dim Em m and ωE. That is, for X ∈ B(Em) and f ∈ m(X ς (m) VE (f )). m(X ς (m) φE VE (f )) = φE E XA−1/2 φm(A−1/2 ς (m)(f )) E XA−1/2 E E m(A−1/2 cE,m rankE = = cE,mωE(ς (m)(A−1/2 E XA−1/2 E )f ) ς (m)(f )) 51 so that ς (m) VE (X) = cE,mς (m)(A−1/2 E XA−1/2 E ). (4.8) Finally, let ΘE be the curvature of the Chern connection of the metric P E on the holomorphic ω be the Bochner Laplacian acting on End Γ∞(M;E) (see [MaMa3, Eq. vector bundle E, and let ∆E (1.3.19)]). Lemma 4.14. Let ς (m) VE its adjoint with respect to φE be the mth component of the Toeplitz map on H 0(S, ω;E, P E) and let ς (m) be m and ωE. Then for all f ∈ End Γ0(M;E) we have the Berezin transform (f )) = f + m−1({trω ΘE, f}/2 − µ(E)f + ∆E ω f ) + O(m−2) VE ς (m) VE (ς (m) VE where {f, g} := f g + gf . In particular, ς (m) VE (ς (m) VE (1E)) = 1E + m−1(trω ΘE − µ(E)1E) + O(m−2). If the scalar curvature sω were not constant we would have ς (m) VE (ς (m) VE (1E)) = 1E + m−1(trω ΘE − µ(E) − sω/2 + sω)1E + O(m−2) where sω := ω(sω). Proof. We are going to show that ΣE(m) := c−1 E,mς (m) VE (ς (m) VE (1E)) is the Szego kernel of KE from a rescaling of that in [Wang2, Thm. 5.2] Let (ψj )j∈J be an orthonormal basis for KE m and let m := H 0(M, ω;E(m), FS(Hm)⊗P E). Then we will obtain the desired expansion FS†(KE m) ∈ End Γ0(M;E(m), FS(KE m)) ⊗ B(KE m) be the matrix whose (j, k)th entry is the rank-1 endomorphism ψj)(ψk. Using the isomorphism End Γ0(M;E(m)) ∼= End Γ0(M;E) we shall always regard FS†(KE m) as an element of End Γ0(M;E) ⊗ B(KE m). We want to prove the formula ς (m) VE (X) = cE,m(id ⊗ Tr)((1 ⊗ X) FS†(KE m)). (4.9) (PE,m) would be the sum of the diagonal elements of FS†(KE m), which is indeed the E,mς (m) Then c−1 Szego kernel of KE m. VE Since A−1/2 E,m Em = KE m we have E,m FS†(Em)A−1/2 A−1/2 E,m =Xj E,m φj)(A−1/2 A−1/2 E,m φj =Xj ψj )(ψj where (φj )j∈J is an orthonormal basis for Em, where (ψj)j∈J = (A−1/2 basis for KE That is, E,m φj)j∈J is an orthonormal m and where ψj)(ψj is the associated rank-1 operator on Γ0(M;E(m), FS(Hm) ⊗ P E). A−1/2 E,m FS†(Em)A−1/2 E,m = FS†(KE m). From (4.8) we have ς (m) VE (X) = cE,mς (m)(A−1/2 E,m XA−1/2 E,m ) = cE,m(id⊗ Tr)((1 ⊗ A−1/2 E,m XA−1/2 E,m ) FS†(Em)) = cE,m(id⊗ Tr)((1 ⊗ X)A−1/2 E,m FS†(Em)A−1/2 E,m )) and so (4.9) holds. 52 From [Wang2, Thm. 5.2] we have χ(E(m)) vol(M,L) rankE ΣE(m) = md1E + md−1(trω ΘE − sω/2)1E + O(md−2), from which we obtain the result for f = 1E using χ(E(m)) vol(M,L) rankE = md + md−1(µ(E) − sω/2) + ··· . The same argument gives that the expansion of ς (m) VE in [KMS1, Prop. 3.6]. (ς (m) VE (f )) for general f follows from the expansion The following gives a geometric meaning to the compact operator PE − ATE : Theorem 4.15. ς (m) VE (PE,m − ATE ,m) = m−1(trω ΘE − µ(E)1E) + O(m−2). Proof. Taking X = PE,m in (4.8) yields ς (m) VE P E. We obtain (PE,m) = ς E(m)(A−1 TE ,m) and ς (m) VE (AE) = ς E(m)(PE,m) = ς (m) VE (PE,m − ATE ,m) = ς (m) VE = ς (m) VE = m−1 trω ΘE + O(m−2). (PE,m) − P E (ς (m) VE (P E) − P E 5 Lifts of Yang -- Mills metrics 5.1 Balanced metrics The notion of balanced metrics on holomorphic vector bundles was introduced in [Wang1]. In this section we shall reformulate it in terms of Hilbert modules and frames. Reall from §3.4.3 that if Em is an inner product on H 0(M;E(m)) then FS(Em) denotes the pro- jection over C∞(M) with a Parseval C∗-frame consisting of an orthonormal basis for Em. Definition 5.1. Let E be a holomorphic vector bundle over M with Aut(E) = C1E. Let m ∈ N be large enough so that dim H 0(M;E) = χ(E(m)). A Hermitian metric BE(m) ∈ C0(M) ⊗ B(H 0(ω, BE(m))) on E(m) is ω-balanced if BE(m) = FS(H 0(ω, BE(m))). In other words, BE(m) is ω-balanced if (i) BE(m) has a Parseval C∗-frame consisting of a basis for the vector space H 0(M;E(m)) and (ii) if PE,m denotes the identity operator on the vector space H 0(M;E(m)) then ω(BE(m)) = rankE χ(E(m)) PE,m. (5.1) If Aut(E) is nontrivial then a Hermitian metric BE(m) on E(m) is weakly ω-balanced if BE(m) is the direct sum of ω-balanced metrics on the summands in some decomposition of E(m) into a direct sum of simple holomorphic vector bundles. If the summands have the same reduced Hilbert polynomials then BE(m) is ω-balanced. 53 In other words, BE(m) is weakly ω-balanced if it has a Parseval C∗-frame consisting of an orthonor- m := H 0(M, ω;E(m), BE(m)): in symbols BE(m) = FS(H 0(ω, BE(m))). mal basis for the Hilbert space KE The defining properties (i) and (ii) of an ω-balanced metric are thus the same no matter what Aut(E) is, since (ii) requires the summands in any decomposition of E to have the same reduced Hilbert poly- nomial, viz. χ(E(m))/ rankE. Indeed, (i) in m on H 0(M;E(m)) while (ii) says Definition 5.1 says that BE(m) = FS(KE that KE When BE(m) is balanced we also say that the inner product (or Hilbert space) H 0(ω, BE(m)) is balanced. Thus, an inner product Em on H 0(M;E(m)) is balanced iff the L2-inner product of the Hermitian metric FS(Em) coincides with Em. By Remark 1.9, our notion of balanced metric coincides with that of [Wang1]. m) for some inner product KE m = H 0(ω, BE(m)). Remark 5.2. A G-equivariant vector bundle need not admit an ω-balanced metric, since the ir- reducible summands need not have the same reduced Hilbert polynomials, but it always admis a weakly ω-balanced metric in the above sense (see [Moss4]). Remark 5.3 (The normalization constant). Since BE(m) must have fiber trace equal to rankE, the only inner product on H 0(M;E(m)) which can support a Parseval C∗-frame for BE(m) as orthonormal basis is the one in which the natural L2(BE(m), ω) inner product. If this inner product is scaled by χ(E(m))/ rankE then the frame bound is scaled by χ(E(m))/ rankE and the balancedness condition becomes to existence of a tight frame with frame constant χ(E(m))/ rankE. The balance condition reads BE(m) = FS(H 0(ω, BE(m))). Recall from §3.4.3 that this means that there is a basis (ψj)j∈J for the vector space H 0(M;E(m)) such that BE(m)(x) = Xj,k∈J hψj (x)ψk(x)iH0(ω,BE(m))ψjihψk ∈ End(H 0(M;E(m))). If we have either a surjection A⊗CN → EN → 0 or an embedding EN ֒→ A⊗CN of graded A-modules then we could regard BE(m)(x) as an element BE(m)(x) ∈ End(Am ⊗ CN ) = B(Hm ⊗ CN ). The assumption that the surjection (or embedding) is a map of graded A-modules ensures that we still have BE(m) = FS(H 0(ω, BE(m))) for the same holomorphic vector bundle E(m), up to holomorphic isomorphism. Let us now investigate the consequences of this representation of the balanced metrics, first in the case of a surjection A ⊗ CN → EN → 0. Given the holomorphic structure on E there is an m0 ∈ N0 such that E(m) is Castelnuovo -- H 0(M;E(m)) is a Mumford regular for all m ≥ m0. Thus, the graded vector space E≥m0 :=Lm≥m0 finitely generated graded A-module and isomorphic to a graded quotient of A≥m0 ⊗ H 0(M;E(m0)). If we let Em0 be H 0(M;E(m0)) Idenfity E≥m0 with a vector subspace of A≥m0 ⊗ H 0(M;E(m0)). endowed with some given inner product then the closure E≥m0 of E≥m0 in H≥m0 ⊗ Em0 is a quotient module in the sense of Hilbert modules, if we endow it with the action given by compressing the A-action on H≥m0 ⊗ Em0 , and the underlying graded A-module is isomorphic to E≥m0 . Now BE(m) can be viewed as a positive idempotent element of C0(M) ⊗ B(Hm ⊗ Em0). The operator PE,m in (5.1) is then the projection of Hm ⊗ Em0 onto the Hilbert subspace Em. Define a projection BE m ∈ C0(M) ⊗ B(Em0) by writing BE(m) = FS(Hm) ⊗ BE m. Then BE m defines E as smooth vector bundle for each m. Using the explicit formula (3.15) for the Toeplitz map ς (m) we can rewrite (5.1) as ς (m)(BE m) = cE,mPE,m (5.2) 54 with the constant cE,m := If we have a balanced metric BE(m) = FS(Hm) ⊗ BE . nm rankE χ(E(m)) m on E(m) for each m ≥ m0 then c−1 E,mς (m)(BE m) = PE, Xm≥m0 where PE is the orthogonal projection of H≥m0 ⊗ Em0 onto E≥m0. 5.2 Superharmonic lifts We have seen hat if P E = FS(E0) is a projection in C0(M) ⊗ MN (C) which defines a holomorphic subbundle E ⊂ OM ⊂ CN then there is a natural associated quotient module EN with FS(Hm)⊗P E = FS(Em) for all m (Proposition 3.27). If the projection PE onto EN has continuous symbol ς(PE) then we have shown that ς(PE) = P E (Theorem 3.23). But we shall see later that ς(PE) is rarely continuous (Theorem 5.15). Also, as we discussed, for any N × N -projection P E over L∞(M) the Toeplitz range Ran ς(P E) is a projection onto a graded module quotient of HN ⊗ CN . And ς(Ran ς(P E)) is continuous when P E is. But even if P E is over C∞(M) and even if the vector bundle E defined by P E is holomorphic it cannot hold that Ran ς(P E) lifts P E unless possibly if E is globally generated. However, there is a special class of Hermitian metrics which do have coinvariant lifts: Lemma 5.4. Let E be a slope-stable holomorphic vector bundle over M = G/K, and let P E be the Yang -- Mills metric on E. Then the Ψ-superharmonic projection Ran ς(P E) is a lift of P E: ς(Ran ς(P E)) = P E. Proof. From [Wang2] we know that there exists a sequence of metrics (BE FS(Hm) ⊗ BE m is an ω-balanced metric on E(m) for large enough m and m)m∈N0 on E such that in C0 (actually in C∞). lim m→∞ BE m = P E We can represent EN as a quotient of A⊗ E0 which is a graded quotient up to a finite-dimensional subspace. Therefore we can embed EN as a vector subspace of HN⊗E0 whose closure is coinvariant up to a finite-dimensional subspace. We may choose the dimension of E0 large enough so that P E is an m⊗FS(Hm) on E(m) has a Parseval C∗-frame given element of C0(M)⊗B(E0). The balanced metric BE by elements of Em and hence can be regarded as an element of C0(M) ⊗ B(Hm ⊗ E0) as determined m is an element of C0(M) ⊗ B(E0). The by our chosen embedding Em ⊂ Hm ⊗ E0. The projection BE balance condition then says that c−1 m) equals the projection PE,m of Hm ⊗ E0 onto the embedded copy of Em. E,mς (m)(BE Now kς (m)(f )k ≤ kfk for all f ∈ C0(M), so m→∞kς (m)(BE lim m − P E)k ≤ lim m→∞kBE m − P Ek = 0, E,mς (m)(BE m) differ by a compact operator. In turn this gives that ς(P E) and i.e. ς(P E) and Pm ς (m)(BE m) also differ by a compact, since c−1 PE = Pm c−1 E,m = 1 + O(m−1). So PE is a lift of P E, i.e. ς(PE) = P E. In particular PE is a projection over T (0) H . We want to show that PE differs from the range projection of ς(P E) only by a finite-rank operator. But PE is Ψ-superharmonic (up to finite-rank operators), so ς(PE) = P E gives its Ψ-harmonic part as ς(P E), so we are done by the uniquness of superharmonic lifts (Proposition 4.2). 55 N := ς(P E)−1/2HE ECD = OCD ⊗ EB\{0} N := Ran ς(P E). Then the Cowen -- Douglas sheaf ECD of HE Theorem 5.5. Let P E be a projection over C∞(M) defining a Hermitian Yang -- Mills vector bundle E over M and set HE N is analytically isomorphic over B \ {0} to the pullback EB\{0} of E; as Hermitian holomorphic vector bundles we have (5.3) where EB\{0} is endowed with the Hermitian metric given by pullback of P E. Moreover, the subnormal tuple on the space KE N is unitarily equivalent (up to finite-rank operators) to the multiplication tuple on the Hardy space H 0(S, ω, FS(HN) ⊗ P E). Proof. We saw in the proof of Lemma 5.4 that HE N is (up to finite-dimensional subspaces) a completion N coincides with P E. of EN, and that the symbol ς(PE) of the projection PE := Ran ς(P E) onto HE So PE is a projection over T (0) H . Theorem 3.23 gives that the Cowen -- Douglas metric on ECD is given by CD(EN) = CD(HN)⊗ P E, where P E is pulled back to a D×-equivariant function on B×{0}. This gives (5.3) and in particular the Cowen -- Douglas bundle of HE m identifies via FS(Hm) ⊗ P E with the C-linear span Em of a C∗-frame ψ for FS(Hm) ⊗ P E. The space EN is thus a graded A-module whose algebraic part is precisely EN. (We stress that FS(Hm) ⊗ P E acting on C1 ⊗ ⊗Em ⊂ C0(M) ⊗ Em does not give holomorphic sections of E(m) but just some other C∗-frame for FS(Hm) ⊗ P E. Thus, when FS(HN) ⊗ P E acts on C0(S) ⊗ E0 it does not projects HE m ⊂ Γ0(M;E(m), FS(Hm) ⊗ P E) endowed with the inner product of L2(ω, FS(Hm) ⊗ P E). We saw that the subnormal tuple on the space KE N is unitarily equivalent (up to finite-rank operators) to the multiplication tuple on the subspace EN of L2(S, ω; FS(HN) ⊗ P E). So we only need to show that multiplication tuples on H 0(S, ω; FS(HN) ⊗ P E) and EN are unitarily equivalent. But we now that the underlying graded A-modules are isomorphic, and since H 0(S, ω, FS(HN)⊗ P E) and EN sits as Hilbert subspaces of the same Hilbert space we can extend this isomorphism to a unitary operator from H 0(S, ω, FS(HN)⊗P E) to EN. As in the proof of Theorem 4.12 we denote by Em the space EE N to the copy of EN sitting in Γ0(S;E).) N is isomorphic to the pullback of E. In the proof of Theorem 4.12 we saw that HE Corollary 5.6. Let E be a slope-stable vector bundle over M, and denote its associated graded A- H 0(M;E(m)). Represent EN as a quotient A-module where the surjection module by EN := Lm∈N0 A ⊗ E0 → EN is grading-presering up to finite-dimensional vector spaces, and let [EN] be its closure of EN in HN ⊗ E0. Then the Cowen-Douglas sheaf ECD of [EN] is locally free boundary limit of the Cowen -- Douglas projection ΠE for [EN] is the Yang -- Mills metric on E. Remark 5.7 (Balanced metric versus CD(Em)). Let E be a slope-stable holomorphic vector bundle over M = G/K and let EN := Ran ς(P E)) be the quotient module from Lemma 5.4. Recall that m is the limit of the powers of the positive operator CD(Hm) ⊗ ς (m)(PE,m). CD(Em) = CD(Hm) ⊗ P E If we use that same notation BE m as in the proof of Lemma 5.4 for the metrics on E converging to the Yang -- Mills metric P E then we have c−1 m)) = ς (m)(PE,m) and hence E,mς (m)(ς (m)(BE lim p→∞ (c−1 E,mς (m)(ς (m)(BE m)))p = P E m . (5.4) But ς (m)(ς (m)(BE The constant c−1 m, so we have ς (m)(ς (m)(BE m)) = BE m + O(m−1). m)) is the Berezin transform of BE E,m is of the form 1 + O(m−1). Therefore c−1 E,mς (m)(ς (m)(BE m)) = BE m + K E m with limm kK E mk = 0. We also know that, for all m ≫ 0, P E m = lim p ς (m)(PE,m)p = P E. The coinvariance of PE gives that, for m ≫ 0, ς (m)(PE,m) = P E ⊕ CE m on Ran P E ⊕ Ker P E m (5.5) 56 m ≥ 0. In contrast, K E m need not be positive and need not be zero on the range of BE with CE m or the range of P E. For large m we know that kP E − BE mk < 1 in the norm of C0(M) so that ς(PE)(x) and BE m(x) are unitarily equivalent uniformly in x [Ols, Prop. 5.2.6]. The analytic embeddings of E(m) into OM ⊗ E0 given by FS(Hm)⊗ P E and BE m are thus arbitrarily close as m gets large but they need not coincide for any finite m. 5.3 Subharmonic lifts When the dual E ∗ of E is globally generated we have an embedding E ֒→ OM ⊗ H 0(M;E ∗) of holo- morphic vector bundles, and hence an inclusion EN ⊂ A ⊗ H 0(M;E ∗) of graded A-modules. Still there is an m0 ∈ N such that E(m) is regular for all m ≥ m0. So we can represent E≥m0 as a graded quotient of A⊗ H 0(M;E(m0)) as before and we obtain a graded quotient E≥m0 of HN⊗ H 0(M;E(m0)) by identifying E≥m0 with a vector subspace of A ⊗ H 0(M;E(m0)) and completing it in the inner product of the Fock space. The coinvariant subspace E≥m0 of HN ⊗ H 0(M;E(m0)) will have the same A-action as the completion of E≥m0 in A ⊗ H 0(M;E ∗) up to graded A-module isomorphism but as Hilbert A-modules they are very different: one is a quotient module and one is a submodule and this is a very important difference for Hilbert modules. For instance, if PE is the projection onto a quotient module then (id −Φ)(PE) is a finite-rank operator, while if PE projects out a submodule (id−Φ)(PE) is not of finite rank except in trivial cases [Guo3]. m ⊗ FS(Hm−m0) on E(m) for each m ≥ m0. Since [EN] ⊂ HN ⊗ H 0(M;E ∗) is an embedding of graded A-modules we can, up to graded A-module isomorphism (or equivalently without changing the isomorphism class of the holomorphic vector bundle E) regard BE(m) as an element of C0(M)⊗B(Hm ⊗ H 0(M;E ∗)). The balance condition then becomes (5.6) Suppose now that we have a balanced metric BE(m) = BE ς (m)(BE m) = cE,mIE,m, where IE = Pm IE,m is the projection onto the graded submodule [EN] ⊂ HN ⊗ H 0(M;E ∗). Note that IE is Ψ-subharmonic, in contrast to the Ψ-superharmonic PE obtained from the presentation of EN as quotient module. We shall see that, in case the sequence (BE m)m≥m0 converge, both ς(IE) and ς(PE) are Yang -- Mills metrics on holomorphic vector bundles isomorphic to E(m0), and for all practical purposes they coincide if we allow an analytic isomorphism to act on E so that they are metrics on the same vector bundle. Lemma 5.8. Let G be a smooth vector bundle over M = G/K admitting a slope-stable holomorphic H 0(M;G(m)). Suppose that G is a subbundle G ⊂ E of some Castelnuovo -- structure GN = Lm∈N0 Mumford regular holomorphic vector bundle E and choose an embedding GN ⊂ EN as graded A- module. Let IG be the projection onto the SE-invariant subspace [GN] ⊂ [EN]. Then the symbol of the ΨE-subharmonic projection IG defines G as smooth vector bundle as well as a Hermitian metric on G which is Yang -- Mills with respect to a holomorphic structure isomorphic to GN. Proof. Given our embedding GN ⊂ EN we obtain the projection IG onto [GN] ⊂ [EN] ⊂ HN ⊗ E0 as (5.7) ς (m)(BG m) = cE,mIG,m, ∀m ≫ 0 for the unique ω-balancing metrics FS(Hm)⊗ BG GN ⊂ EN. Let P G := limm→∞ BG the proof of Lemma 5.4 we see that m on G(m) with respect to the holomorphic structure m be the associated Yang -- Mills metric on G. In the same way as in m→∞ kς (m)(BG lim m − P G)k ≤ lim m→∞kBG m − P Gk = 0, i.e. ς(P G) and Pm ς (m)(BG IG = Pm c−1 ς(IG) = P G. In particular IG is a projection over T (0) H . m) differ by a compact operator. In turn this gives that ς(P G) and m) differ by a compact, since cG,m − 1 = O(m−1). So IG is a lift of P G, i.e. G,mς (m)(BG 57 Note that the subspace [GN] ⊂ [EN] ⊂ HN ⊗ E0 is merely semi-invariant under the shift S on HN ⊗ E0. Corollary 5.9. Let G be a smooth vector bundle over M with a slope-stable holomorphic structure H 0(M;G(m)). Suppose that G∗ is globally generated. Then the Yang -- Mills metric P G GN =Lm∈N0 admits a Ψ-subharmonic lift. Proof. Since G∗ is globally generated is globally generated we can take E in Lemma 5.8 to be a trivial holomorphic vector bundle. This means that [GN] is a submodule of HN ⊗ G0 for some Hilbert space G0 and hence IG will be Ψ-subharmonic. 5.4 Direct sums If E = F⊕ G is a direct sum of Hilbert A-modules then for the Serre sheaves we have [GoWe1, Prop. 7.14(3)] E ∼= F ⊕ G as OM-modules. Also, as in Remark 3.9, FN ⊂ EN is a reducing submodule if and only if for the Cowen -- Douglas sheaves we have ECD ∼= FCD ⊕ GCD as Hermitian holomorphic vector bundles. In other words, iff CD(F ⊕ G) ∼= CD(F) ⊕ CD(G). Proposition 5.10 (Direct sums of Yang -- Mills metrics). Suppose that P E is a projection over C∞(M) defining a holomorphic vector bundle E = F ⊕ G with slope-stable summands F and G, and that P E = P F + P G with P F and P G Yang -- Mills metrics on F and G. Then ς(Ran ς(P E)) = P E. Proof. From Lemma 5.4 we get ς(PF ) = P F and ς(PG) = P G for superharmonic projections PF and PG onto quotients FN and GN, and the Serre sheaves of these quotients are isomorphic to F and G as holomorphic vector bundles. Setting EN := FN ⊕ GN we obtain E as the Serre sheaf of EN. We can present EN as a quotient of HN ⊗ E0 where E0 = F0 ⊕ G0. The ranges of ς(P F ) and ς(P G) are orthogonal so, Ran ς(P E) = Ran ς(P F + P G) is the same as the projection PE := PF + PG onto EN. Thus P E = ς(PF ) + ς(PG) = ς(PF + PG) = ς(PE) and the proposition holds. 5.5 The nature of ς(PE) If E is a torsionfree OM-module, denote by Gr(E) the torsionfree OM-module obtained by summing the successive quotients in the Harder -- Narasimhan -- Seshadri filtration of E (see [Jaco2, §2.1] for details). Thus Gr(E) is a direct sum of slope-stable torsionfree sheaves on M, and the summands have the same slope iff E is slope-semistable. If E is a holomorphic vector bundle over M, a locally free subsheaf G ⊂ E is a subbundle iff the quotient E/G is locally free. So if Gr(E) is locally free then all the subsheaves in the Harder -- Narasimhan -- Seshadri filtration are by subbundles and this splits smoothly. Hence for an arbitrary holomorphic vector bundle E we have that Gr(E) ∼= E smoothly if and only if Gr(E) is locally free. Theorem 5.11. Let E be a holomorphic vector bundle over M = G/K, represent EN :=Lm∈N0 as a quotient of A ⊗ E0 for some Hilbert space E0, and let EN be the completion of EN in HN ⊗ E0. Let PE be the projection of HN ⊗ E0 onto a quotient module EN. Suppose that Gr(E) is locally free and let QE be the direct sum of the Yang -- Mills metrics on the stable summands of Gr(E). Then ς(PE) = QE. In particular ς(PE) is a matrix over C0(M) ⊂ L∞(M). H 0(M;E(m)) 58 Proof. We consider first the case when E is slope-semistable. Assume that E has a filtration 0 → G → E with G a stable subbundle (a filtration with several subbundles can be treated by the same argument). The direct sum Gr(E) of stable quotients is then Gr(E) = (E/G) ⊕ G. 5.4). The stable vector bundles G and F := E/G admit Yang -- Mills metrics ς(PG) and ς(PF ) obtained from Ψ-superharmonic projections PG and PF over T (0) H . These are the projections onto the completions of GN :=Lm H 0(M;G(m)) and FN :=Lm H 0(M;F (m)) realized as quotient modules (recall Lemma The image of the projection Q := PG + PF is the completion of the direct sum QN = GN ⊕ FN in a Fock inner product and ς(Q) is a Yang -- Mills metric on Gr(E) as we saw in Proposition 5.10. Here QN = GN ⊕ FN is a direct sum of graded A-modules by definition. We note that GN ⊕ FN equals EN as graded vector space (up to finite-dimensional subspaces), but GN is merely a submodule of EN and need not have an A-module complement. Therefore QN = (EN/GN) ⊕ GN need not be A-isomorphic to EN as graded A-module (under the present local freness assumption Gr(E) ∼= E we know that the pullbacks of Gr(E) and E are analytically isomorphic as vector bundles over B \ {0} but this need not be D×-equivariantly). We realize EN as a quotient module EN and denote by IG the projection of EN onto the submodule GN := [GN]. Let also PF denote the projection of EN onto the quotient module EN ⊖ GN = [FN]. Then ς(PE) = ς(IG + PF ) = ς(IG) + ς(PF ) is a direct sum of Yang -- Mills metrics on the holomorphic direct sum G ⊕ F . Indeed, ς(IG) can be identified with ς(PG) by Lemma 5.8. Now if E is not semistable we have a Harder -- Narasimhan filtration by semistable subsheaves (assumed to be subbundles here). For the proof of the theorem we may assume it is given by 0 → G → E. Set F := E/G. We have PE = PG + PF and since F is semistable we know that ς(PF ) is the direct sum of Yang -- Mills metrics on Gr(F ), and similarly for G. This gives the result. Corollary 5.12. Let EN be a range of a Ψ-superharmonic projection PE over T (0) H and let ECD,M denote the holomorphic vector bundle over M such that ECD = OCD ⊗ ECD,M. Then ς(PE) is a Yang -- Mills metric on ECD,M. Proof. Since ECD is locally free (see Theorem 3.23), the Serre sheaf E of EN is locally free and ς(PE) is a real-analytic metric on ECD,M. We have a factorization CD(EN) = CD(HN) ⊗ ς(PE) so ς(PE) defines ECD,M and its holomorphic structure. By Theorem 5.11 we have that ς(PE) is a Yang -- Mills metric on Gr(E), so Gr(E) is analytically isomorphic to ECD,M and ς(PE) is a Yang -- Mills metric on ECD,M. The next step would be to obtain a generalization Wang's theorem (Lemma 1.4), namely that every torsionfree slope-stable sheaf has a "singular Yang -- Mills metric" coming from a sequence of "singular balanced metrics". That would prove: Conjecture 5.13. Theorem 5.11 is true without the assumption that Gr(E) is locally free, i.e. ς(PE) is always a metric on Gr(E) which is the direct sum of singular Yang -- Mills metrics on the simple summands of Gr(E). If this conjecture is true then we have: Corollary 5.14. For arbitrary holomorphic vector bundle E, letting PE be the projection onto a Fock H 0(M;E(m)) the symbol ς(PE) has entries C0(M) ⊂ L∞(M) if and only completion of EN :=Lm∈N0 if Gr(E) is locally free. 59 5.6 Gieseker-stability and superharmonic lifts With L∞(M) acting as multiplication operators on L2(M, ω) one can consider limits of sequences in L∞(M) ⊗ MN (C) in the strong operator topology. Proposition 5.15. Let E be a Gieseker-stable vector bundle and let (BE balanced metrics. Then the projection m)m≫0 be its sequence of P E := SOT− lim m→∞ BE m exists as a matrix over L∞(M) and it coincides with ς(PE), where PE is the projection onto the Proof. We have completion of EN :=Lm H 0(M;E(m)) in Fock space. BE m = SOT− lim = SOT− lim = SOT− lim = ς(PE), SOT− lim m→∞ m→∞ m→∞ m→∞ ς (m)(ς (m)(BE m)) c−1 E,mς (m)(ς (m)(BE ς (m)(PE,m) m)) where we used the balance of each BE m for m ≫ 0 in the penultimate line. For a vector bundle E which is Gieseker-stable but not slope-stable the limit limm→∞ BE m does not exist in C∞ by [Wang2]. If it also fails to exist in C0 then the projection PE onto [EN] must be of the form PE = ς(P E) + CE with CE noncompact. Conjecture 5.16. Let E be a Gieseker-stable vector bundle and let (P E balanced metrics. Then the projection m )m≫0 be its sequence of P E := SOT− lim defines Gr(E) and a singular Yang -- Mills metric on Gr(E). m→∞ P E m 5.7 From balance to Gieseker-stability We now give a more direct proof of one implication in the main theorem of [Wang1]: Theorem 5.17. Let E be a holomorphic vector bundle over M and suppose that E(m) admits an ω-balanced Hermitian metric for each m ≫ 0. Then E is Gieseker-polystable. Proof. Fix m ≫ 0. By assumption we have a balanced inner product HE (ψj )j∈J be an orthonormal basis for HE obtained from the projection FS(HE by the matrix whose (j, k)th entry is the function (ψjψk) ∈ C∞(M). Let m)) ⊗ B(HE m) m on H 0(M;E(m)). Let m. Let (··) be the Hilbert C∗-structure on Γ∞(M;E(m)) m) can be represented m) ∈ End Γ∞(M;E(m), FS(HE m) ∈ C∞(M) ⊗ B(HE m). Recall that FS(HE FS†(HE be the matrix of rank-1 operators ψj )(ψk on Γ∞(M;E(m), FS(HE C∞(M) we have the isomorphism m)). Using End Γ∞(M;Lm) ∼= and we shall identify FS†(HE End Γ∞(M;E(m)) ∼= End Γ∞(M;E), m) with an endomorphism of Γ∞(M;E). Recall that the state ωE : End Γ∞(M;E) → C is defined on rank-1 endomorphisms by (5.8) ωE(ψ)(φ) := 1 rankE ω((φψ)). 60 So balance of HE m says that ωE(ψj )(ψk) = 1 χ(E(m)) (5.9) for all j, k = 1, . . . , dim HE for Γ∞(M;E(m), FS(HE m = χ(E(m)). Recall also that balance says that (ψj)j∈J is a Parseval frame m)), which under the isomorphism (5.8) means that ψj)(ψj = 1E, (5.10) Xj where 1E is the identity in End Γ∞(M;E). For X ∈ B(HE m), define an endomorphism of Γ∞(M;E) by ς E m(X) := (id ⊗ TrHE m )((1 ⊗ X) FS†(HE m)) = χ(E(m))(id ⊗φE m)((1 ⊗ X) FS†(HE m)). Then ς E by balance (5.10). m(1) is the frame operator of (ψj )j regarded as an element of End Γ∞(M;E), which equals 1E We also define, for each f ∈ End Γ0(M;E), a Toeplitz-type operator on HE m)). ς E(m)(f ) = χ(E(m))(ωE ⊗ id)((f ⊗ 1HE ) FS†(HE m m by The resulting map ς E(m) : End Γ0(M;E) → B(HE m) is then unital by balance (5.9). The maps ς E(m) and ς E(m) are adjoints with respect to ωE and φE m, )((f ⊗ X) FS†(HE m)) )((1 ⊗ X) FS†(HE φE m(ς E(m)(f )X) = (ωE ⊗ TrHE = ωE(f (id⊗ TrHE m m m))), and since they are both unital they therefore both intertwine the states ωE and φE m, φE m(ς E(m)(f )) = ωE(f ) and ωE(ς E m(X)) = φE m(X). Let G ⊂ E be an analytic subsheaf. Then GN := Lm H 0(M;G(m)) is a graded submodule of EN :=Lm H 0(M;E(m)), and we denote by PG the projection of [EN] onto [GN]. If (ψj)j∈J is our balanced frame as before then there is a subset JG ⊂ J such that each ψG with j ∈ JG will be in Gm := H 0(M;G(m)) and these ψG projection acting on Γ∞(M;E) with image Γ∞(M;G) then P GψG we obtain χ(E(m)) rankE j := ψj j 's form a basis for Gm. So if P G is the j for all j ∈ JG. So for j, k ∈ JG χ(E(m))ωE(P Gψj)(P Gψk) = ω((ψjψk)) = δj,k by balance. Let FN be the quotient EN/GN. Since (ψG ς E(m)(P G) is always positive there is a positive operator CG,m on Fm such that j ψk) can be nonzero even for k /∈ JG but j = ψG So we obtain ς E(m)(P G) = PG,m ⊕ CG,m on Em = Gm ⊕ Fm. This gives, for m large enough so that dim Em = χ(E(m)) and dim Gm = χ(G(m)), that Equality χ(E(m))/ rankE = χ(G(m))/ rankG occurs iff ς E(m)(P G) = PG,m and this happens iff j )j∈JG as Parseval C∗-frame, so that E splits FS(HE holomorphically as E = F ⊕ G. This gives the statement. m) splits into P F ⊕ P G with P G having (ψG ς E(m)(P G) ≥ PG,m. rankG rankE = ωE(P G) = φE ≥ φE = m(ς E(m)(P G)) m(PG,m) χ(G(m)) χ(E(m)) . 61 6 Equivariant vector bundles In this section we specialize to the case of equivariant vector bundles over M = G/K. The quantization of these were initiated in [Hawk1], which we now build on. A new thing here is that we add new data: inner products on the quantum side and Hermitian metrics on the classical side. It turns out that these go very well with the quantization because of the equivariance, and in fact the existence of a Hermitian metric' which"quantizes perfectly ' is likely to chararacterize equivariant vector bundles, in a sense made more precise below. Given any K-representation K we can form the vector bundle E = G ×K K associated with the principal K-bundle G → M → 0 and the represetation K. Such a vector bundle E is called G-equivariant [Snow1]. Every G-equivariant vector bundle comes with a natural holo- morphic structure [Rama1, §3.2]. The space H 0(M;E) of global holomorphic sections then carries a representation of G which is said to be induced from the representation K of the subgroup K [Bott1]. A G-equivariant vector bundle E = G ×K K is irreducible if K is an irreducible K-representation. Then H 0(M;E(m)) is an irreducible G-representation for all m ≥ 0. 6.1 Quotient modules with equivariant Cowen -- Douglas sheaves 6.1.1 Equivariant Cowen -- Douglas metrics are balanced If E is an irreducible G-equivariant vector bundle over M = G/K then each fiber E(x) carries an irreducible K-representation and a unique K-invariant inner product. This gives us a unique G- equivariant Hermitian metric P E on E. We also have a unique G-invariant inner product E0 on the G-representation H 0(M;E). The evaluation H 0(M;E) ∋ ψ → ψ(x) ∈ E(x) is G-equivariant so if it is surjective then E(x) sits as a Hilbert subspace of E0. The G-invariant Hermitian metric on E is then given by P E = FS(E0). It is known that P E is ω-Yang -- Mills [Koba4]. Proposition 6.1. Let P E be a projection over C0(M) defining a globally generated irreducible G- equivariant Hermitian vector bundle E, so that P E = FS(E0) for the G-invariant inner product E0 on H 0(M;E). Then FS(Em) = FS(Hm) ⊗ P E is balanced on E(m) for each m ∈ N0. Moreover, the Cowen -- Douglas metric of the quotient module EN := Ran ς(P E) is of the form CD(EN)(v) = CD(HN)(v) ⊗ P E([v]), ∀v ∈ B \ {0}, and the graded A-module underlying EN is precisely EN :=Lm H 0(M;E(m)). Proof. As noticed in Proposition 3.27, we have FS(Em) = FS(Hm) ⊗ P E for all m. The Haar orthogonality relations (see §2.1.2) say precisely that ω(P E µ,ν) = 1 dim E0 δµ,ν, ∀µ, ν = 1, . . . , dim E0 (6.1) i.e. the Hilbert space H 0(ω, FS(E0)) coincides with E0. In other words, the metric P E = FS(E0) is ω-balanced. More generally, for each m ∈ N0, the metric FS(Em) = FS(Hm) ⊗ P E on E(m) is G-equivariant and hence balanced by the above argument. We thus have ς (m)(P E) = cE,mPE,m, ∀m ∈ N0. It follows that PE is the range projection of ς(P E) and that ς(P E) equals PE modulo compacts. Therefore (6.1) holds by Theorem 3.23. As in the proof of Theorem 4.12 the vector space Em := Ran ς (m)(P E) coincides with the C-linear span Em of a Parseval frame for FS(Hm) ⊗ P E. 62 6.1.2 E l,m is unital Recall our unital completely positive maps ιm,l : B(Hm) → B(Hl) defined by ∀A ∈ B(Hm) m,l(A ⊗ 1l−m)Vm,l, ιm,l(A) := V ∗ (6.2) where Vm,l is the isometric embedding of Hl into Hm ⊗ Hl−m. These can be generalized to vector- valued quotient modules. To simplify the notation we write ιm,l also for the induced map from B(Hm ⊗ CN ) to B(Hl ⊗ CN ), where we identify operators on HN ⊗ CN with N × N -matrices of operators on HN as before. Let now PE be the projection of HN ⊗ CN onto a coinvariant subspace EN. Coinvariance gives m,l(PE,m)PE,l is thus an operator on El which is ≥ PE,l. Since m,l(PE,m) ≥ PE,l. The operator PE,lιE ιE ιE m,l is unital it is contractive, so that we must have ιE m,l(PE,m)PE,l = PE,l = PE,lιE m,l(PE,m). The map ιE m,l : B(Em) → B(El) defined by ιE m,l(A) := PE,lιm,l(A)El , is therefore unital. From (6.2) we obtain the formula m,l (A ⊗ 1l−m)V E m,l, ιE m,l(A) = V E∗ ∀A ∈ B(Em) ∀A ∈ B(Em), (6.3) m,l = PE,lV ∗ m,l is unital, V E m,lEm⊗Hl−m. m,l := Vm,lEl. Since ιE where V E The adjoint is given by V E∗ m,l is an isometric embedding of El into Em ⊗ Hl−m. The Hilbert space Hm carries an irreducible unitary representation of G. Endow Hm ⊗ CN with the unitary G-representation where G trivially on the factor CN . Proposition 6.2. Suppose that each Hilbert space Em ⊂ Hm ⊗ E0 is invariant under the G-action on Hm ⊗ E0. Then the isometries V E Proof. As already noticed above, coinvariance ensures that V E m,l is the restriction of Vm,l to Em. Since Em and El are G-invariant they carry G-representations by restriction of the G-representations on Hm ⊗ CN and Hl ⊗ CN respectively. So we obtain the proposition from the fact that Vm,l intertwines the G-representation on Hm ⊗ CN with that on Hl ⊗ CN . m,l : El → Em ⊗ Hl are intertwiners of G-representations. Let φE m be the normalized trace on B(Em), φE m(A) := (dim Em)−1 Tr(A), ∀A ∈ B(Em), and define the completely positive map to be the adjoint of ιE m,l with respect to φE l . Basically by definition this means that E l,m : B(El) → B(Em) m and φE E l,m(B) := (φE l−m ⊗ id)(Vm,lBV ∗ m,l), ∀B ∈ B(El). Proposition 6.3. Suppose that the Cowen -- Douglas sheaf of EN is of the form ECD = OCD ⊗ EB\{0} where EB\{0} is the pullback to B\{0} of a G-equivariant vector bundle E over M. Then E l,m is unital. Proof. By assumption (6.1) holds and then, as before, the graded A-module underlying EN is precisely EN :=Lm H 0(M;E(m)). Thus each vector space Em is a G-representation. As explained in [Hawk1, §C], for all l ≥ m ≥ 0 the representation Em is contained as a vector space in the tensor product El ⊗ H∗ l−m is obtained by averaging over the group G using the Haar state ω, and equals the Fock inner product on El ⊗ H∗ l−m l−m. The G-invariant inner product on El ⊗ H∗ 63 up to a scalar factor on each irreducible direct summand. Therefore there is an isometric embedding of Em into El ⊗ H∗ l−m, l,m : Em → El ⊗ H∗ W E l−m. By assumption the G-representation on E0 is irreducible and this gives that the G-representation on Em = H 0(M;E(m)) is irreducible for each m. Then W E m,l are unique and from [Hawk1, Eq. (5.9)] we have l,m and V E The rightmost formula shows that E E l,m(B) = (φE l−m ⊗ id)(Vm,lBV ∗ m,l) = W E∗ l,m is unital, as asserted. l,m(B ⊗ 1H∗ l−m )W E l,m, . l,m : Em → El ⊗ H∗ Remark 6.4 (Reversed time evolution). Recall that the conjugate of an irreducible representation u of a compact group (or more generally, compact quantum group) G is an irreducible representation ¯u such that u⊗ ¯u contains the trivial representation. The existence of a conjugate to each irreducible representation is one of the structural properties that characterize representation categories of com- pact quantum groups. Therefore we do not expect isometries W E l−m to exist unless the Em's are representations of some compact quantum group. In [An4] we interpreted the existence of the backward maps E l,m as a "quantum symmetry" in case the Hilbert spaces Em models the environment of some physical quantum system. If Em is not irreducible then we again let φE m be the trace on B(Em) corresponding to the Haar state, i.e. the direct sum of the normalized traces on the irreducible direct summands. Then the adjoint E is again unital. If all irreducible summands of E have the same reduced Hilbert polynomial χ(E(m))/ rankE then these two definitions of E 6.1.3 The (d + 1)-isometry SE Recall that if EN ⊂ HN ⊗ E0 is a graded subspace then we denote by SE = (SE,1, . . . , SE,n) the shift compressed to EN and consider the grading-preserving positive operator S∗ E,αSE,α. We let SE be the positive square root of S∗ ESE. In this section we obtain a generalization of some results in §2 (which are recovered by taking E to be the trivial line bundle). First of all, Proposition 6.3 gives: ESE := Pn m,l with respect to φE l,m coincide. m and φE l l,m of ιE α=1 S∗ Corollary 6.5. Let SE be the shift on EN. Suppose that the Cowen -- Douglas sheaf of EN descends to a G-equivariant vector bundle E on M = G/K and that all irreducible summands of E have the same reduced Hilbert polynomial χ(E(m))/ rankE. Then SEPE,m is a scalar for each m ∈ N0, viz. SEPE,m =s χ(E(m + 1)) χ(E(m)) PE,m. Proof. The operator S∗ ESEPE,m E m,m+1(PE,m+1) = equals the identify PE,m by Proposition 6.3. χ(E(m)) χ(E(m + 1)) Recall the maps Ψ and Φ∗ on B(HN) defined by Ψ(X) :=Pn where S = ST is the shift on HN. Proposition 6.6. For any X =Pm Xm ∈ Γb =Qm B(Hm) we have r(cid:19)φm+r(X) φm((id −Ψ)p(X)) = (−1)r(cid:18)p α=1 T ∗ p Xr=0 αXTα and Φ∗(X) :=Pn α=1 S∗ αXSα, and for all m, p ∈ N0. Tr((id −Φ∗)p(Xm)) = p Xr=0 (−1)r(cid:18)p r(cid:19) Tr(Xm+r) 64 Proof. We have p (id−Ψ)p = (−1)r(cid:18)p r(cid:19)Ψr Xr=0 and Tr(Φr and φm ◦ Ψr = φm+r. Hence the first result. The second follows from (id−Φ∗)p =Pp r(cid:1)Φr If EN is a quotient module then we have the shift tuple SE acting on EN, and as in §2.4.1 we can ∗(X)pm) = Tr(XΦr(pm)) = Tr(Xpm+r). r=0(−1)r(cid:0)p ∗ consider the operators where ΦE,∗(X) := Pn preserves the subspace EN we have α=1 S∗ E,αXSE,α for all X ∈ B(EN). Since the backward shift S∗ on HN ⊗ E0 Bp(SE) := (id −ΦE,∗)p(1), Bp(SE) := (id −Φ∗)p(PE), where as usual Φ∗(X) :=Pn Proposition 6.7. For any graded quotient module EN and each p ∈ N0 the operator Bp(SE) satisfies α=1 S∗ αXSα. Tr(Bp(SE)pm) = p Xr=0 (−1)r(cid:18)p r(cid:19) Tr(PE,m+r) for all m ∈ N0. For p ≥ d + 1 we have Tr(Bp(SE)) = 0. Proof. From Proposition 6.6 we have Tr((id−Φ∗)p(PE,m)) = Pp Serre sheaf E of EN is Castelnuovo -- Mumford regular, so r=0(−1)r(cid:0)p r(cid:1) Tr(PE,m+r). Now the Tr(PE,m+r) = dim Em+r = dim H 0(M;E(m + r)) = χ(E(m + r)) for all m + r ∈ N0. Let δp is the pth iterate of the difference operator δ acting on sequences a = (a(m))m∈N0 as We have p (δa)(m) := a(m) − a(m + 1). (−1)r(cid:18)p r(cid:19) Tr(PE,m+r) = (δpχE)(m) Xr=0 where χE(m) := χ(E(m)). The Euler characteristic χ(E(m)) is a polynomial in m of degree d, which is the same as saying that δpχE = 0 for p ≥ d + 1. So for p ≥ d + 1 we obtain Tr(Bp(SE)pm) = 0, ∀m ∈ N0. Remark 6.8. We could also consider the operators Bp(TE) of the rescaled tuple TE; these are given by Bp(TE) = (id−Ψ)p(PE). The normalized trace φm(PE) = χ(E(m))/nm is not a polynomial in m. Therefore Proposition 6.6 does not imply that Tr(Bp(TE)pm) is zero for any finite p. There- fore TE is not a p-isometry for any p and ς(P E) = Ψ∞(PE) cannot be expressed as a finite sum Pp q=0(cid:0)p Proposition 6.9. Let EN be a quotient module such that ECD = OCD ⊗ EB\{0} where EB\{0} is the pullback to B \ {0} of a G-equivariant vector bundle E over M. Then SE is a strict (d + 1)-isometry: Bd(SE) 6= 0 and q(cid:1)(id−Ψ)q(PE) for any p. Bp(SE) = 0, ∀p ≥ d + 1. 65 Moreover, if E is irreducible then Bp(SE) acts as a scalar on each graded piece Em ⊂ EN for each p = 0, . . . , d, and we have the Scatten-class estimate Bp(SE) = Xm∈N0 φE m(Bp(SE))PE,m, Bp(SE) ∈ Lq ⇐⇒ q > d + 2 − p. Proof. Suppose first that E is irreducible. Then ΦE,∗(PE) = SE2 acts as a scalar on each graded piece Em. It follows that B1(SE)PE,m = (id −ΦE,∗)(PE)PE,m is a scalar, viz. φE m(B1(SE))PE,m. Hence also B2(SE)PE,m = (id−ΦE,∗)(B1(SE))PE,m is a scalar, and so on. From this and Proposition 6.7 we see directly that Bp(SE) = 0 for p ≥ d+1 when E is irreducible. For arbitrary equivariant E we have that EN is a direct sum of SE-reducing subspaces corresponding to the irreducible summands of E, so SE is a (d + 1)-isometry for all equivariant vector bundles E. From Proposition 6.7 we have the estimate from which the stated Schatten-class estimate on Bp(SE) follows. φE m(Bp(SE)) = O(m−p) We see that Bp(SE) is zero precisely when Bp(SE) is trace-class. 6.1.4 Characterization of equivariance Proposition 6.10. Let EN be a quotient module with continuous symbol P E = ς(PE). Assume that the shift SE on EN has no reducing subspaces. Then the following are equivalent: (a) ς (m)(P E) = cE,mPE,m for all m ≫ 0. (b) Em = cE,mH 0(ω, FS(Hm) ⊗ P E) for all m ≫ 0. (c) S∗ ESEpm = dim Em+1/ dim Em for all m ≫ 0. (d) FS(H 0(ω, FS(Hm) ⊗ P E)) = FS(Hm) ⊗ P E is ω-balanced for all m ≫ 0. Proof. The operator ς (m)(P E) compares the inner products Em and H 0(ω, FS(Hm) ⊗ P E) (cf. The- orem 4.12 and its proof). So (a) is equivalent to (b). Clearly (a) ⇐⇒ (d). E,mς (m) : Γ∞(M;E) → B(Em) as l goes to infinity. So (c) implies (a). If (b) holds then the fact that the multiplication tuple on H 0(S, ω; P E) is a spherical isometry l,m's are unital for all l ≥ m then so is their limit c−1 If the E ensures that (c) holds (cf. the proof of Lemma 2.6). This gives the proposition. We know that all conditions in Proposition 6.10 hold when EN comes from an equivariant vector bundle E. We expect (c) in Proposition 6.10 to hold only if the vector bundle defined by P E is G-equivariant (cf. Remark 6.4). Therefore the equivalent conditions in Proposition 6.10 are likely to characterize the quotient modules which give rise to G-equivariant vector bundles over M. 66 6.2 Guo-stability In this section we show, as asserted in the Introduction, that every G-equivariant vector bundle over G/K is a direct sum of vector bundles satisfying a stability condition (which we call Guo-stability) that is stronger than Gieseker-stability. This seems to be a new result. Theorem 6.11. Let E be a G-equivariant vector bundles over M and suppose that all irreducible summands of E have the same reduced Hilbert polynomial χ(E(m))/ rankE. Then for every quotient sheaf E → F → 0 and all m ≫ 0 we have χ(F (m)) χ(F (l)) ≥ χ(E(m)) χ(E(l)) ∀l ≥ m (6.4) , with equality iff F is a subbundle (in which case F is a direct summand of E). In particular, E is Gieseker-polystable. Proof. By replacing E with E(m0) for large enough m0 if necessary we may assume that EN := H 0(M;E(m)) is a graded quotient of A ⊗ CN . Similarly, whenever E → F → 0 is a quotient Lm∈N0 H 0(M;F (m)) is a graded quotient of EN. Let EN and FN sheaf we may assume that FN := Lm∈N0 be the completions of EN and FN in the inner product of HN ⊗ CN . Then FN is an invariant subspace for the backward shift S∗ E-invariant subspace FN of EN, and conversely every S∗ E-invariant subspace gives rise to a quotient of E (see §3.2.1). E on EN. We have thus encoded the quotient sheaf F as an S∗ Consider the unital completely positive map map ΨE(X) := (S∗ ESE)−1/2 n Xα=1 S∗ E,αXSE,α(S∗ ESE)−1/2, ∀X ∈ Γb. By Corollary 6.5 ΨE restricts to E l,m : B(Em+1) → B(Em): For B ∈ B(Em+1) we have ΨE(B) := χ(E(m + 1)) χ(E(m)) n Xα=1 S∗ E,αBSE,α. E,α(S∗ E,α and S∗ We have seen that S∗ that S∗ Since ΨE is unital, the assumption that FN is invariant under S∗ with equality iff FN is reducing. For X ∈ B(El) we have Ψl−m terms of the maps E m ◦ E φE ESE commutes with every grading-preserving operator on EN, and this ensures ESE)−1/2 have the same invariant graded subspaces for each α ∈ {1, . . . , n}. E is thus equivalent to ΨE(PF ) ≤ PF E (X) = l,m(X) for all m ≤ l. In l,m(PF,l) ≤ PF,m for all l ≥ m. Using l,m the inequality ΨE(PF ) ≤ PF reads E l we then obtain l,m = φE χ(F (l)) χ(E(l)) = φE l (PF,l) = φE m(E l,m(PF,l)) ≤ φE m(PF,m) = χ(F (m)) χ(E(m)) for all l ≥ m, with equality iff F≥m is reducing. Most likely the Guo-stability condition (6.4) is stronger than even slope-stability, and is a char- acteristic of G-equivariance (cf. Proposition 6.10). References [AbLe1] Abdullah B, Le T. The structure of m-isometric weighted shift operators. Oper. Matrices. Vol 10, Issue 2, pp. 319-334 (2016). [ACKi1] ´Alvarez-C´onsul L, King A. A functorial construction of moduli of sheaves. Invent. Math. Vol 168, pp. 613-666 (2007). 67 [An4] Andersson A. Andersson A. Detailed balance as a quantum-group symmetry of Kraus opera- tors. J. Math. Phys. Vol 59, Issue 2, 022107 (2018). [An5] Andersson A. Dequantization via quantum channels. Lett. Math. Phys. Vol 106, Issue 10, pp. 1397-1414 (2016). [An6] Andersson A. Berezin quantization of noncommutative projective varieties. arXiv: 1506.01454 (2015). [Arv6c] Arveson W. Subalgebras of C∗-algebras III: Multivariable operator theory. Acta Math. Vol 181, pp. 159-228 (1998). [Arv7a] Arveson W. The curvature of a Hilbert module over C[z1, . . . , zd]. Proc. Natl. Acad. Sci. USA. Vol 96, 11096-11099 (1999). [Arv7b] Arveson W. The curvature invariant of a Hilbert module over C[z1, . . . , zd]. J. Reine Angew. Math. Vol 522, pp. 173-236 (2000). [Arv7c] Arveson W. The curvature invariant of a Hilbert module over C[z1, . . . , zd]. arXiv:math/9808100v1 (1998). [Arv8] Arveson W. p-Summable commutators in dimension d. J. Operator Theory. Vol 54, pp. 101- 117 (2005). [Arv9] Arveson W. Quotients of standard Hilbert modules. Trans. Amer. Math. Soc. Vol 359, pp. 6027-6055 (2007). [Atha3] Athavale A. On the intertwining of joint isometries. J. Oper. Theory. Vol 23, pp. 339-350 (1990). [Bala1] Balazs P. Matrix representation of operators using frames. Sampl. Theory Signal Image Process. Int. J. Vol 7, Issue 1, pp. 39-54 (2008). [BaHa1] Barr´ıa J. Halmos PR. Asymptotic Toeplitz operators. Trans. Amer. Math. Soc. Vol 273, Issue 2, pp. 621-630 (1982). [Bere2] Berezin FA. General concept of quantization. Comm. Math. Phys. Vol 40, pp. 153-174 (1995). [BMN2] Berm´udez T, Martin´on A, Negr´ın E. Weighted shift operators which are m-isometries. In- tegr. Equ. Oper. Theory. Vol 68, pp. 301-312 (2010). [BhSa1] Bhattacharyya T, Sarkar J. Characteristic function for polynomially contractive commuting tuples. J. Math. Anal. Appl. Vol 321, pp. 242-259 (2006). [BMP1] Biswas S, Misra G, Putinar M. Unitary invariants for Hilbert modules of finite rank. J. Reine Angew. Math. Issue 662, pp. 165-204 (2012). [Bott1] Bott R. Homogeneous vector bundes. Ann. Math. Vol 66, Issue 2, pp. 203-248 (1957). [CDT1] Carmeli C, De Vito E, Toigo A. Vector valued reproducing kernel Hilbert spaces of integrable functions and Mercer theorem. Anal. Appl. (Singap.) Vol 4, Issue 4, pp. 377-408 (2006). [Catl1] Catlin DW. The Bergman kernel and a theorem of Tian. In: Analysis and geometry in several complex variables. pp. 1-23. Birkhauser (1999). [ChDo2] Chen L, Douglas RG. A local theory for operator tuples in the Cowen -- Douglas class. Adv. Math. Vol 307, pp. 754-779 (2017). [Cheng1] Cheng G. An additive invariant on the vector-valued Hardy space over the ball. J. Operator Theory. Vol 67, Issue 1, pp. 21-31 (2012). [ChFa1] Cheng G, Fang X. An additive formula for Samuel multiplicities on Hilbert spaces of analytic functions. J. Funct. Anal. Vol 260, Issue 7, pp. 2027-2042 (2011). [Chri1] Christensen O. An introduction to frames and Riesz bases. Birkahuser (2003). [CoDo1] Cowen MJ, Douglas RG. Complex geometry and operator theory. Acta Math. Vol 141, pp. 187-261 (1978). 68 [CoDo2] Cowen MJ, Douglas RG. Operators possessing an open set of eigenvalues. Proceedings, Fejer-Riesz Conference, Colloq. Math. Sot. Jdnos Bolyai. Vol 35, pp. 323-341 (1980). [CuLe1] Cuckovi´c Z, Le T. Toeplitzness of composition operators in several variables. Complex Var. Elliptic Equ. Vol 59, Issue 10, pp. 1351-1362 (2014). [Davi2] Davidson KR. C∗-algebras by example. Fields Institute Monographs (1996). [DRS1] Davidson KR, Ramsey C, Shalit OM. The isomorphism problem for some universal operator algebras. Adv. Math. Vol 228, Issue 1, pp. 167-218 (2011). [DRS2] Davidson KR, Ramsey C, Shalit OM. Operator algebras for analytic varieties. Trans. Amer. Math. Soc. Vol 367, Issue 2, pp. 1121-1150 (2014). [Deut1] Deutsch F. Best approximation in inner product spaces. Springer (2001). [Doug1] Douglas RG. Operator theory and complex geometry. Extracta Math. Vol 24, Issue 2, pp. 135-165 (2009). [DFS1] Douglas RG, Foias C, Sarkar J. Resolutions of Hilbert modules and similarity. J. Geom. Anal. Vol 22, pp. 471-490 (2012). [DKKS2] Douglas RG, Kim Y-S, Kwon H-K, Sarkar J. Curvature invariant and generalized canonical operator models -- II. J. Funct. Anal. Vol 266, pp. 2486-2502 (2014). [DoPa1] Douglas RG, Paulsen VI. Hilbert Modules over function algebras. Longman Research Notes. Vol 217, Springer (1989). [Fang4] Fang X. The Fredholm index of a pair of commuting operators, II. J. Funct. Anal. Vol 256, Issue 6, pp. 1669-1692 (2009). [Fang5] Fang X. Additive invariant on the Hardy space over the polydisc. J. Funct. Anal. Vol 253, pp. 359-372 (2007). [Fein1] Feintuch A. On asymptotic Toeplitz and Hankel operators. In: The Gohberg anniversary collection, Vol II (pp. 241-254). Oper. Theory Adv. Appl. Vol 41, Birkhauser, Basel (1989). [Feld1] Feldman NS. Essentially subnormal operators. Proc. Amer. Math. Soc. Vol 127, Issue 4, pp. 1171-1181 (1999). [Fisc1] Fischer G. Complex analytic geometry. Lect. Notes Math. Vol 538, Springer 1976. [FrLa3] Frank M, Larson DR. A module frame concept for Hilbert C∗-modules. Contemp. Math. Vol 247, pp. 207-233 (1999). [Gehe1] Geh´er GP. Positive operators arising asymptotically from contractions. Acta Sci. Math. (Szeged). Vol 79, pp. 273-287 (2013). [Gehe2] Geh´er GP. Characterization of Ces`aro and L-asymptotic limits of matrices. Linear Multilin- ear Algebra. Vol 63, Issue 4, pp. 788-805 (2015). [GlRi1] Gleason J, Richter S. m-isometric commuting tuples of operators on a Hilbert space. Integr. Equ. Oper. Theory. Vol 56, Issue 2, pp. 181-196 (2006). [GRS1] Gleason J, Richter S, Sundberg C. On the index of invariant subspaces in spaces of analytic functions. J. Reine. Angew. Math. Vol 587, pp. 49-76 (2005). [GRS2] Greene D, Richter S, Sundberg C. The structure of inner multipliers on spaces with complete Nevanlinna Pick kernels. J. Funct. Anal. Vol 194, pp. 311-331 (2002). [Guo3] Guo K. Defect operators for submodules of H 2 n. J. Reine Angew. Math. Vol 573, pp. 181-209 (2004). [GaKu1] Gartner A, Kummerer B. A coherent approach to recurrence and transience for quantum Markov operators. arXiv:1211.6876 (2012). [GoWe1] Gortz U, Wedhorn T. Algebraic geometry: Part I: Schemes. With examples and exercises. Vieweg + Teubner Verlag (2010). [Halm1] Halmos PR. A Hilbert space problem book. Second Edition. Springer (1982). 69 [HaLa1] Han D, Larson D. Frames, bases and group representations. Mem. Amer. Math. Soc. Vol 697 (2000). [HaHe1] Hausen J, Heinzner P. Actions of compact groups on coherent sheaves. Transform. Groups. Vol 4, Issue 1, pp. 25-34 (1999). [Hawk1] Hawkins E. Quantization of equivariant vector bundles. Comm. Math. Phys. Vol 202, Issue 3, pp. 517-546 (1999). [Hawk2] Hawkins E. Geometric quantization of vector bundles and the correspondence with defor- mation quantization. Comm. Math. Phys. Vol 215, Issue 2, pp. 409-432 (2000). [Hein1] Heinzner P. Geometric invariant theory on Stein spaces. Math. Ann. Vol 289, pp. 631-662 (1991). [HoMa1] Hoffmann PHW, Mackey M. (m, p)-isometric and (m,∞)-isometric operator tuples on normed spaces. Asian-Eur. J. Math. Vol 8, Issue 2 p. 1550022 (32 pages) (2015). [Iz1] Izumi M. Non-commutative Poisson boundaries and compact quantum group actions. Adv. Math. Vol 169, Issue 1, pp. 1-57 (2002). [Jaco2] Jacob A. The Yang -- Mills flow and the Atiyah -- Bott formula on compact Kahler manifolds. Amer. J. Math. Vol 138, Issue 2, pp. 329-365 (2016). [KaSc1] Karabegov AV, Schlichenmaier M. Identification of Berezin -- Toeplitz deformation quantiza- tion. J. Reine Angew. Math. Issue 540, p.49-76 (2001). [Kara4] Karaev MT. Functional analysis proofs of Abel's theorems. Proc. AMS, Vol 132, Issue 8, pp. 2327-2329 (2004). [Kaup1] Kaup L, Kaup B. Holomorphic functions in several complex variables. Walter de Gruyter (1983). [KMS1] Keller J, Meyer J, Seyyedali R. Quantization of the Laplacian operator on vector bundles. Math. Ann. Vol 366, Issue 3, pp. 865-907 (2016). [Kerc10] K´erchy L. A description of invariant subspaces of C11-contractions. J. Operator Theory. Vol 15, pp.327-344 (1986). [KKT1] Knop F, Kraft H, Vust T. The Picard group of a G-variety. In Algebraische Transforma- tionsgruppen und Invarianten theorie Algebraic Transformation Groups and Invariant Theory (pp. 77-87). Birkhauser, Basel (1989). [Koba1] Kobayashi S. Differential geometry of complex vector bundles. Princeton University Press (1987). [Koba4] Kobayashi S. Homogeneous vector bundles and stability. Nagoya Math. J. Vol 101, pp. 37-54 (1986). [KoMa1] Koszul JL, Malgrange B. Sur certaines structures fibr´ees complexes. Arch. Math. Vol 9, Issue 1, pp. 102-109 (1958). [Kubr1] Kubrusly CS. An introduction to models and decompositions in operator theory. Birkhauser (1997). [KwTr1] Kwon HK, Treil S. Similarity of operators and geometry of eigenvector bundles. Publ. Mat. Vol 53, Issue 2, pp. 417-438 (2009). [Laza1] Lazarsfeld R. Positivity in algebraic geometry, I and II. Ergeb. Math. Grenzgeb. Vols 48-49. Springer-Verlag (2004). [MaMa3] Ma X, Marinescu G. Holomorphic Morse inequalities and Bergman kernels. Birkhauser (2007) [Moss4] Mossa R. Balanced metrics on homogeneous vector bundles. Int. J. Geom. Methods Mod. Phys. Vol 8, Issue 7, p. 1433-1438 (2011). [INT1] Izumi M, Neshveyev S, Tuset L. Poisson boundary of the dual of SUq(n). Comm. Math. Phys. Vol 262, Issue 2, pp. 505-531 (2006). 70 [Olse1] Olsen CL. A structure theorem for polynomially compact operators. Amer. J. Math. Vol 93, pp. 686-698 (1971). [Ols] Olsen W. K-theory of C∗-algebras - a friendly approach. Oxford University Press (1993). [Pop7] Popescu G. Similarity and ergodic theory of positive linear maps. J. Reine Angew. Math. Vol 561, pp. 87-129 (2003). [Prun1] Prunaru B. Some exact sequences for Toeplitz algebras of spherical isometries. Proc. Am. Math. Soc. Vol 135, pp. 3621-3630 (2007). [Prun2] Prunaru B. Toeplitz operators associated to commuting row contractions. J. Funct. Anal. Vol 254, Issue 6, pp. 1626-1641 (2008). [Rama1] Ramanan S. Holomorphic vector bundles on homogeneous spaces. Topology. Vol 5, Issue 2, pp. 159-177 (1966). [SaWr1] Saito K, Wright JM. Monotone complete C∗-algebras and generic dynamics. Springer (2015). Hilbert spaces. I. J. Operator Theory. Vol 73, Issue 2, pp. 433-441 (2015). [Sark3] Sarkar J. An invariant subspace theorem and invariant subspaces of analytic reproducing kernel Hilbert spaces - II. J. Complex Anal. Oper. Theory. Vol 10, Issue 4, pp. 769-782 (2016). [Seg1] Segal G. Lectures on Lie groups. In: Lectures on Lie groups and Lie algebras. Cambridge University Press (1995). [Serr2] Serre JP. Faisceaux alg´ebriques coh´erents. Ann. of Math. pp. 197-278 (1955). [Snow1] Snow D. Homogeneous vector bundles. In: Group actions and invariant theory (Montreal, PQ, 1988), pp. 193-205 (1989). [TrWi1] Treil S, Wick BD. Analytic projections, corona problem and geometry of holomorphic vector bundles. J. Amer. Math. Soc. 22, Issue 1, pp. 55-76 (2009). [UhYa1] Uhlenbeck K, Yau ST. On the existence of Hermitian-Yang -- Mills connections in stable vector bundles. Commun. Pure Appl. Math. Vol 39, pp. S257-S293 (1986). [Wang1] Wang X. Balance point and stability of vector bundles over a projective manifold. Math. Res. Lett. Vol 9, Issues 2-3, pp. 393-411 (2002). [Wang2] Wang X. Canonical metrics on stable vector bundles, Comm. Anal. Geom. Vol 13, Issue 2, pp. 253-285 (2005). [Weav1] Weaver N. Set theory and C∗-algebras. Bull. Symb. Logic. Vol 13, pp. 1-20 (2007). [Zhao1] Zhao C. Approximate representation of Bergman submodules. Chin. Ann. Math. Vol 37B, Issue 2, pp. 221-234 (2016). 71
1207.1930
3
1207
2013-01-23T02:00:27
Picard groups of certain stably projectionless C*-algebras
[ "math.OA" ]
We compute Picard groups of several nuclear and non-nuclear simple stably projectionless C*-algebras. In particular, the Picard group of Razak-Jacelon algebra W_2 is isomorphic to a semidirect product of Out(W_2) with R_+^\times. Moreover, for any separable simple nuclear stably projectionless C*-algebra with a finite dimensional lattice of densely defined lower semicontinuous traces, we show that Z-stability and strict comparison are equivalent. (This is essentially based on the result of Matui and Sato, and Kirchberg's central sequence algebras.) This shows if A is a separable simple nuclear stably projectionless C*-algebra with a unique tracial state (and no unbounded trace) and has strict comparison, the following sequence is exact: [{CD} {1} @>>> \mathrm{Out}(A) @>>> \mathrm{Pic}(A) @>>> \mathcal{F}(A) @>>> {1} {CD}] where $\mathcal{F}(A)$ is the fundamental group of A.
math.OA
math
PICARD GROUPS OF CERTAIN STABLY PROJECTIONLESS C∗-ALGEBRAS NORIO NAWATA Abstract. We compute Picard groups of several nuclear and non-nuclear sim- ple stably projectionless C∗-algebras. In particular, the Picard group of the Razak-Jacelon algebra W2 is isomorphic to a semidirect product of Out(W2) with R× +. Moreover, for any separable simple nuclear stably projectionless C∗-algebra with a finite dimensional lattice of densely defined lower semicon- tinuous traces, we show that Z-stability and strict comparison are equivalent. (This is essentially based on the result of Matui and Sato, and Kirchberg's central sequence algebras.) This shows if A is a separable simple nuclear sta- bly projectionless C∗-algebra with a unique tracial state (and no unbounded trace) and has strict comparison, the following sequence is exact: 1 −−−−−→ Out(A) −−−−−→ Pic(A) −−−−−→ F (A) −−−−−→ 1 where F (A) is the fundamental group of A. 1. Introduction Let A be a C∗-algebra. Brown, Green and Rieffel introduced the Picard group Pic(A) of A in [5]. We say that an automorphism α of A is inner if there exists a unitary element u in the multiplier algebra M (A) of A such that α(a) = uau∗ for any a ∈ A. Let Inn(A) denote the set of inner automorphisms of A, and let Out(A) = Aut(A)/Inn(A). They showed that if A is σ-unital, then Pic(A) is isomorphic to Out(A ⊗ K). Kodaka computed Picard groups of several unital C∗-algebras in [21], [22] and [23]. In particular he computed the Picard groups of the irrational rotation algebras Aθ. If θ is not quadratic irrational number, then Pic(A) is isomorphic to Out(Aθ) and if θ is a quadratic number, then Pic(Aθ) is isomorphic to Out(Aθ) ⋊ Z. Kodaka considered the following set FP/ ∼= {[p] p is a full projection in A ⊗ K such that p(A ⊗ K)p ∼= A} where [p] is the Murray-von Neumann equivalence class of p and showed that if Out(A) is a normal subgroup of Out(A ⊗ K) and A is unital, then FP/ ∼ has a suitable group structure and the following sequence is exact: 1 −−−−→ Out(A) −−−−→ Pic(A) −−−−→ FP/ ∼ −−−−→ 1. Note that there exists a simple unital AF algebra B with a unique tracial state such that FP/ ∼ of B does not have any suitable group structure. If A is unital, K-theoretical method enables us to show that Out(A) is a normal subgroup of Out(A ⊗ K) (see [21, Proposition 1.5]). The set of FP/ ∼ is similar to the fundamental group F (M ) of a II1 factor M introduced by Murray and von Neumann in [28]. Watatani and the author introduced the fundamental group F (A) of a simple unital C∗-algebra A with a unique tracial state τ based on Kodaka's results. The fundamental group F (A) 2010 Mathematics Subject Classification. Primary 46L05, Secondary 46L08; 46L35. Key words and phrases. Picard group; Fundamental group; Stably projectionless C∗-algebra; Cuntz semigroup; Kirchberg's central sequence algebra. The author is a Research Fellow of the Japan Society for the Promotion of Science. 1 2 NORIO NAWATA is defined as the set of the numbers τ ⊗ Tr(p) for some projection p ∈ Mn(A) such that pMn(A)p is isomorphic to A. We showed that F (A) is a multiplicative subgroup of R× + and computed fundamental groups of several C∗-algebras in [31]. Moreover we showed that any countable subgroup of R× + can be realized as the fundamental group of a separable simple unital C∗-algebra with a unique tracial state in [32]. Note that the fundamental groups of separable simple unital C∗- algebras are countable. Furthermore the author introduced the fundamental group of a simple stably projectionless C∗-algebra with unique (up to scalar multiple) densely defined lower semicontinuous trace τ in [29]. If τ is a tracial state and A is σ-unital, then the fundamental group of F (A) of A is defined as the set of the numbers dτ (h) for some positive element h ∈ A ⊗ K such that h(A ⊗ K)h is isomorphic to A where dτ is the dimension function defined by τ . Note that if A is unital, then this definition coincides with the previous definition and there exist separable simple stably projectionless C∗-algebras such that their fundamental groups are equal to R× +. The fundamental group of a II1 factor M is equal to the set of trace-scaling constants for automorphisms of a II∞ factor M ⊗ B(H). This characterization shows that the fundamental groups of II1 factors are related to the structure theorem for type IIIλ factors where 0 < λ ≤ 1 (see [43] and [44]). We have a similar characterization, that is, if A is σ-unital, then the fundamental group of A is equal to the set of trace scaling constants for automorphisms of A ⊗ K. We denote by Z the Jiang-Su algebra constructed in [14]. The Jiang-Su algebra Z is a unital separable simple infinite-dimensional nuclear C∗-algebra whose K- theoretic invariant is isomorphic to that of complex numbers. We may regard Z as the stably finite analogue of the Cuntz algebra O∞. We say that a C∗-algebra A is Z-stable if A is isomorphic to A ⊗ Z. It has recently become important to study regularity properties in Elliott's classification program for nuclear C∗-algebras. In particular, Toms and Winter conjectured that for simple separable nuclear non-type I unital C∗-algebras, the properties of (i) finite nuclear dimension, (ii) Z-stability and (iii) strict comparison of positive elements are equivalent (see, for example [46] and [49]). It is known that (i) implies (ii) and (ii) implies (iii) due to work of Winter [49] and Rørdam [40] respectively. Recently, Matui and Sato showed that (iii) implies (ii) in the case of finitely many extremal tracial states in [25]. In this paper we shall compute Picard groups of several nuclear and non-nuclear simple stably projectionless C∗-algebras. In the case of stably projectionless C∗- algebras, the theory of the Cuntz semigroup enables us to compute Picard groups of several examples. We shall show that if A is a separable simple exact Z-stable stably projectionless C∗-algebra with a unique tracial state τ and no unbounded trace, then the following sequence is exact: 1 −−−−→ Out(A) −−−−→ Pic(A) −−−−→ F (A) −−−−→ 1. Since there exists a unital simple Z-stable algebra A with a unique tracial state such that Out(A) is not a normal subgroup of Pic(A), Z-stable stably projectionless C∗-algebras are in this sense more well-behaved than unital stably finite Z-stable C∗-algebras. Let W2 be the Razak-Jacelon algebra studied in [13], [37], which has trivial K-groups and a unique tracial state and no unbounded trace. Then W2 is Z-stable, and hence the sequence above is exact in this case. Moreover we shall show that the exact sequence above splits. Therefore Pic(W2) is isomorphic to Out(W2) ⋊ R× +. Based on the result of Matui and Sato, and Kirchberg's central sequence algebras, for any separable simple infinite-dimensional non-type I nuclear C∗-algebra with a finite dimensional lattice of densely defined lower semicontinuous traces, we shall PICARD GROUPS OF CERTAIN STABLY PROJECTIONLESS C∗-ALGEBRAS 3 show that Z-stability and strict comparison are equivalent. consider property (SI).) (It is important to In particular, if A is a simple C∗-algebra with a finite dimensional lattice of densely defined lower semicontinuous traces in the class of Robert's classification theorem ([37, Corollary 6.2.4]), then A is Z-stable. Moreover we see that there are many examples that the sequence above is exact. But we do not know whether the exact sequence above splits in this case. This question is related to the existence of a one parameter trace scaling automorphism group of A ⊗ K. In the final part of this paper we shall give some remarks and a reason of the notation of W2. Some results show every separable simple Z-stable stably projectionless C∗-algebra A with a unique tracial state has similar properties of (McDuff) II1 factors. 2. The Picard group In this section we shall review basic facts on the Picard groups of C∗-algebras introduced by Brown, Green and Rieffel in [5] and some results in [29]. Let A be a C∗-algebra and X a right Hilbert A-module, and let H(A) denote the set of isomorphic classes [X ] of countably generated right Hilbert A-modules. We denote by LA(X ) the algebra of the adjointable operators on X . For ξ, η ∈ X , a "rank one operator" Θξ,η is defined by Θξ,η(ζ) = ξhη, ζiA for ζ ∈ X . We denote by KA(X ) the closure of the linear span of "rank one operators" Θξ,η and by K the C∗-algebra of compact operators on an infinite-dimensional separable Hilbert space. Let XA be a right Hilbert A-module A with the obvious right A- action and ha, biA = a∗b for a, b ∈ A. Then there exists a natural isomorphism of KA(XA) to A, where A acts on XA by left multiplication. Hence if A is unital, then KA(XA) = LA(XA). A multiplier algebra, denote by M (A), of a C∗-algebra A is the largest unital C∗-algebra that contains A as an essential ideal. It is unique up to isomorphism over A and isomorphic to LA(XA). Let HA denote the standard nxn converges in A} with an A-valued inner nyn. Then there exists a natural isomorphism of Hilbert module {(xn)n∈N xn ∈ A,P x∗ product h(xn)n∈N, (yn)n∈Ni = P x∗ A ⊗ K to KA(HA). Let A and B be C∗-algebras. An A-B-equivalence bimodule is an A-B-bimodule F which is simultaneously a full left Hilbert A-module under a left A-valued inner product Ah·, ·i and a full right Hilbert B-module under a right B-valued inner product h·, ·iB, satisfying Ahξ, ηiζ = ξhη, ζiB for any ξ, η, ζ ∈ F . We say that A is Morita equivalent to B if there exists an A-B-equivalence bimodule. There exists an isomorphism ϕ of A to KB(F ) such that ϕ(Ahξ, ηi) = Θξ,η for any ξ, η ∈ F . The standard Hilbert module HA can be regard as an A ⊗ K-A-equivalence bimodule. A dual module F ∗ of an A-B-equivalence bimodule F is a set {ξ∗; ξ ∈ F } with the operations such that ξ∗ + η∗ = (ξ + η)∗, λξ∗ = (λξ)∗, bξ∗a = (a∗ξb∗)∗, Bhξ∗, η∗i = hη, ξiB and hξ∗, η∗iA = Ahη, ξi. The bimodule F ∗ is a B-A-equivalence bimodule. We refer the reader to [35] and [36] for the basic facts on equivalence bimodules and Morita equivalence. For A-A-equivalence bimodules E1 and E2, we say that E1 is isomorphic to E2 as an equivalence bimodule if there exists a C-linear one-to-one map Φ of E1 onto E2 with the properties such that Φ(aξb) = aΦ(ξ)b, AhΦ(ξ), Φ(η)i = Ahξ, ηi and hΦ(ξ), Φ(η)iA = hξ, ηiA for a, b ∈ A, ξ, η ∈ E1. The set of isomorphic classes [E] of the A-A-equivalence bimodules E forms a group under the product defined by [E1][E2] = [E1 ⊗A E2]. We call it the Picard group of A and denote it by Pic(A). The identity of Pic(A) is given by the A-A-bimodule E := A with Aha1, a2i = a1a∗ 1a2 for a1, a2 ∈ A. The inverse element of [E] in the Picard group of A is the dual module [E ∗]. Let α be an automorphism of A, and let E A α = A with the obvious left A-action and the obvious A-valued inner product. We define the right A-action on E A α by ξ · a = ξα(a) for 2 and ha1, a2iA = a∗ 4 NORIO NAWATA α . Then E A α ⊗ E A α and a ∈ A, and the right A-valued inner product by hξ, ηiA = α−1(ξ∗η) any ξ ∈ E A for any ξ, η ∈ E A α is an A-A-equivalence bimodule. For α, β ∈ Aut(A), E A α is isomorphic to E A β if and only if there exists a unitary u ∈ M (A) such that α = ad u ◦ β. Moreover, E A α◦β. Hence we obtain an homomorphism ρA of Out(A) to Pic(A). Note that for any α ∈ Aut(A), E A α is isomorphic to XA as a right Hilbert A-module. Conversely we have the following proposition. β is isomorphic to E A Proposition 2.1. Let E be an A-A-equivalence bimodule such that E is isomorphic to XA as a right Hilbert A-module. Then there exists an automorphism α of A such that E is isomorphic to E A α as an A-A-equivalence bimodule. Proof. Let Φ be a right Hilbert A-module isomorphism of XA to E, and let ψ be an isomorphism of KA(E) to KA(XA) induced by Φ. Since KA(XA) is naturally isomorphic to A, we may regard ψ as an isomorphism of KA(E) to A. There exists an isomorphism ϕ of A to KA(E) such that ϕ(Ahξ, ηi) = Θξ,η for any ξ, η ∈ E because E is an A-A-equivalence bimodule. Put α := (ψ ◦ ϕ)−1, and define a map F of E to E A α by F (Φ(a)) := α(a) for any a ∈ A. Note that we have AhΦ(a), Φ(b)i = ϕ−1(ΘΦ(a),Φ(b)) = ϕ−1 ◦ ψ−1(ab∗) = α(ab∗) and a · Φ(b) = ϕ(a)Φ(b) = Φ(ψ ◦ ϕ(a)b) = Φ(α−1(a)b) for any a, b ∈ A. Therefore it can easily be checked that F is an A-A-equivalence bimodule isomorphism. (cid:3) An A-B-equivalence bimodule F induces an isomorphism Ψ of Pic(A) to Pic(B) by Ψ([E]) = [F ∗ ⊗ E ⊗ F ] for [E] ∈ Pic(A). Therefore if A is Morita equivalent to B, then Pic(A) is isomorphic to Pic(B). Brown, Green and Rieffel showed that if A is σ-unital, then Pic(A) is isomorphic to Out(A ⊗ K) (see [5, Theorem 3.4 and Corollary 3.5]). Indeed a homomorphism ρA⊗K of Aut(A ⊗ K) to Pic(A ⊗ K) induces an isomorphism of Out(A ⊗ K) onto Pic(A ⊗ K). i=1 ξihξi, ηiA in norm for any η ∈ X . X if η = P∞ {PN A sequence {ξi}iN of a right Hilbert A-module X is called countable basis of If KA(X ) is σ-unital, then X has a countable basis. A sequence {ξi}iN is a countable basis if and only if i=1 Θξi,ξi}N ∈N is an approximate unit for KA(X ). See [15], [16], [29] and [50] for details of bases of Hilbert modules. We denote by T (A) the set of densely defined lower semicontinuous traces on A and T1(A) the set of tracial states on A. We have the following proposition. Proposition 2.2. ([29, Proposition 2.4]) Let A be a simple σ-unital C∗-algebra and X a countably generated Hilbert A- module, and let τ be a densely defined lower semicontinuous trace on A. For x ∈ KA(X )+, define T rX τ (x) := ∞X i=1 τ (hξi, xξiiA) i=1 is a countable basis of X . Then T rX where {ξi}∞ τ does not depend on the choice of basis and is a densely defined (resp. strictly densely defined) lower semicontinuous trace on KA(X ) (resp. LA(X )). The following proposition is [29, Remark 2.5]. Moreover it is well-known (see for example [6]). But we include the proof for completeness. PICARD GROUPS OF CERTAIN STABLY PROJECTIONLESS C∗-ALGEBRAS 5 Proposition 2.3. Let A be a simple σ-unital C∗-algebra and X a countably gener- ated Hilbert A-module. Then there exists a bijective correspondence between T (A) and T (KA(X )). Proof. Since a right Hilbert A-module X is a KA(X )-A-equivalence bimodule, X ∗ is an A-KA(X )-equivalence bimodule. Let {ξj}j∈N be a countable basis of X and {η∗ i }i∈N a countable basis of X ∗. For any a ∈ A+ and τ ∈ T (A), we have T rX ∗ T rX τ (a) = lim n→∞ = lim n→∞ = lim n→∞ = lim n→∞ = lim n→∞ = lim n→∞ = lim n→∞ i=1 i=1 i=1 nX nX nX nX nX nX nX i=1 i=1 i=1 i=1 T rX τ (hη∗ i , aη∗ i iKA(X )) T rX τ (KA(X )hηia 1 2 , ηia 1 2 i) T rX τ (Θ ηia 1 2 ,ηia ) 1 2 j=1 j=1 ∞X ∞X ∞X ∞X j=1 j=1 τ (hξj , ηia 1 2 hηia 1 2 , ξjiAiA) τ (hξj , ηia 1 2 iAhηia 1 2 , ξjiA) τ (hηia 1 2 , ξjiAhξj, ηia 1 2 iA) τ (hηia 1 2 , ξjhξj , ηia 1 2 iAiA). Since {ξj}j∈N is a countable basis of X , we see that hηia 1 2 , mX j=1 Hence ξjhξj , ηia 1 2 iAiA ր hηia 1 2 , ηia 1 2 iA (m → ∞). lim n→∞ nX i=1 ∞X j=1 τ (hηia 1 2 , ξjhξj , ηia 1 2 iAiA) = lim n→∞ nX i=1 τ (hηia 1 2 , ηia 1 2 iA) by the lower semicontinuity of τ . Since Θη∗ i ,η∗ i is corresponding to hηi, ηiiA, we see i=1hηi, ηiiA}n∈N is an approximate unit for A. Therefore we have that {Pn nX lim n→∞ τ (hηia 1 2 , ηia 1 2 iA) = lim n→∞ 1 τ (a 2 ( hηi, ηiiA)a 1 2 ) = τ (a) nX i=1 by the lower semicontinuity of τ . Consequently T rX ∗ T rX τ i=1 = τ . Since (X ∗)∗ is naturally isomorphic to X as a KA(X )-A-equivalence bimodule, = τ for any τ ∈ T (KA(X )) as above. Hence we obtain the (cid:3) we see that T rX conclusion. T rX ∗ τ The following Corollary is folklore. Corollary 2.4. Let A be a simple σ-unital C∗-algebra and h be a non-zero positive element in A. Then every densely defined lower semicontinuous trace on hAh is a restriction of some densely defined lower semicontinuous trace on A. Proof. Put X = hA. Since KA(hA) is naturally isomorphic to hAh, it is enough to show that T rX τ (τ ) = τ hAh for any τ ∈ T (A) by Proposition 2.3. Let {an}n∈N be 6 NORIO NAWATA a countable basis of hA. (Note that the norm of Hilbert A-module hA is equal to the norm of C∗-algebra A.) Since τ (a∗ 2 ana∗ 2 ) for any x ∈ hAh+ and nxan) for any N ∈ N. Therefore (cid:3) τ ∈ T (A), we have PN nxan) ≤ PN +1 τ (x) = τ (x) by the lower semicontinuity of τ . nxan) = τ (x n=1 τ (a∗ we see that T rX n=1 τ (a∗ 1 1 nx For τ ∈ T (A), define a map Tτ of H(A) to [0, ∞] by Tτ ([X ]) := ∞X n=1 τ (hξn, ξniA) where {ξn}n∈N is a countable basis of X . This map is well-defined map and does not depend on the choice of basis. Moreover we have Tτ (X ) = T rX τ k. n ) for h ∈ (A ⊗ K)+. Then dτ is a dimension Put dτ (h) = limn→∞ τ ⊗ Tr(h τ (1LA(X )) = kT rX 1 function. We have the following proposition. Proposition 2.5. [29, Proposition 3.1] Let A be a simple σ-unital C ∗-algebra with unique (up to scalar multiple) densely defined lower semicontinuous trace τ and h a positive element in A ⊗ K. Then Tτ (hHA) = dτ (h). The following proposition is an immediate corollary of [29, Proposition 3.3] Proposition 2.6. Let A be a simple σ-unital C∗-algebra with a unique tracial state τ and no unbounded trace. Then for every right Hilbert A-module X and every A-A-equivalence bimodule E, Tτ ([X ⊗ E]) = Tτ ([X ]) Tτ ([E]). If A is σ-unital, then for any A-A-equivalence bimodule E there exists a positive element h in A ⊗ K such that E is isomorphic to hHA as a right Hilbert A-module. Note that h(A ⊗ K)h is isomorphic to A and hHA has a suitable structure as an A-A-equivalence bimodule in this case. (See, for example, [29, Proposition 2.3].) The following proposition is a key proposition in this paper. Proposition 2.7. Let A be a simple σ-unital C∗-algebra with a unique tracial state τ and no unbounded trace. Define a map T of Pic(A) to R× + by T ([hHA]) = dτ (h). Then T is a well-defined multiplicative map and T ([E A α ]) = 1 for any α ∈ Aut(A). Moreover Im(T ) is equal to the set {dτ (h) ∈ R× + h is a positive element in A ⊗ K such that A ∼= h(A ⊗ K)h}. Proof. Let [hHA] ∈ Pic(A). Then dτ (h) = Tτ (hHA) = kT rhHA 2.5. Since KA(hHA) ∼= A has no unbounded trace, dτ (h) = kT rhHA we see that T is well-defined map and Im(T ) is equal to the set τ τ k by Proposition k < ∞. Hence {dτ (h) ∈ R× + h is a positive element in A ⊗ K such that A ∼= h(A ⊗ K)h} by an argument above. Proposition 2.6 implies that T is a multiplicative map. It is easy to see that E A α is isomorphic to sA = A as a right Hilbert A-module where s is a strictly positive element in A. Since τ is a tracial state, T ([E A α ]) = dτ (s) = 1. (cid:3) Put F (A) = Im(T ). We call F (A) the fundamental group of A, which is a multiplicative subgroup of R× + by the proposition above. Let A a simple C∗-algebra with unique (up to scalar multiple) densely defined lower semicontinuous trace τ . For any α ∈ Aut(A ⊗ K), τ ⊗ Tr ◦ α is a densely defined lower semicontinuous trace on A ⊗ K. Hence there exists a positive number λ such that τ ⊗Tr◦α = λτ ⊗Tr. Define a map S of Out(A⊗ K) to R× + by S([α]) = λ PICARD GROUPS OF CERTAIN STABLY PROJECTIONLESS C∗-ALGEBRAS 7 where τ ⊗ Tr ◦ α = λτ ⊗ Tr. Then S is a well-defined multiplicative map and Im(S) is equal to the set S(A) := {λ ∈ R× + τ ⊗ Tr ◦ α = λτ ⊗ Tr for some α ∈ Aut(A ⊗ K) }. The following proposition is a strengthened version of [29, Proposition 4.20]. Proposition 2.8. Let A be a simple σ-unital C∗-algebra with a unique tracial state τ and no unbounded trace. Then there exists an isomorphism Ψ of Out(A ⊗ K) to Pic(A) such that S([α])−1 = T ◦ Ψ([α]) for any [α] ∈ Out(A ⊗ K), and hence F (A) = S(A). Proof. Define Ψ([α]) := [(HA)∗ ⊗ E A⊗K α ⊗ HA]. Since HA is an A ⊗ K-A-equivalence bimodule and ρA⊗K induces an isomorphism of Out(A ⊗ K) to Pic(A ⊗ K) by [5, Corollary 3.5], Ψ is an isomorphism of Out(A ⊗ K) to Pic(A). Note that (HA)∗ is naturally isomorphic to (s ⊗ e11)(A ⊗ K) as an A-A ⊗ K-equivalence bimodule where s is a strictly positive element in A and e11 is a rank one projection in K. It is easy to see that for any element ζ in an algebraic tensor product (s ⊗ e11)(A ⊗ K) ⊙ E A⊗K ⊙ HA, there exists an element ξ in HA such that α ζ = (s 1 2 ⊗ e11) ⊗ (s 1 4 ⊗ e11) ⊗ α−1(s 1 4 ⊗ e11)ξ. Therefore it can easily be checked that (HA)∗ ⊗ E A⊗K α α−1(s ⊗ e11)HA as a right Hilbert A-module. We have ⊗ HA is isomorphic to dτ (α−1(s ⊗ e11)) = lim n→∞ τ ⊗ Tr(α−1((s ⊗ e11) 1 n )) = S([α])−1 since τ is a tracial state on A. Hence we obtain the conclusion. (cid:3) 3. The Cuntz semigroup In this section we shall review basic facts of the Cuntz semigroup and some results in [7], [10], [38] and [40]. See, for example, [2] for details of the Cuntz semigroup. Let A be a C∗-algebra. For positive elements a, b ∈ A we say that a is Cuntz smaller than b, written a - b, if there exists a sequence {xn}n∈N of A such that kx∗ nbxn − ak → 0. Positive elements a and b are said to be Cuntz equivalent, written a ∼ b, if a - b and b - a. Define the Cuntz semigroup Cu(A) as the set of Cuntz equivalence classes of positive elements in A ⊗ K endowed with the order [a] ≤ [b] if a is Cuntz smaller than b, and the addition [a] + [b] = [a′ + b′] where a ∼ a′, b ∼ b′ and a′b′ = 0. Note that this definition is different from the original definition W(A) in [8]. (We have Cu(A) = W(A ⊗ K).) The Cuntz semigroup Cu(A) is also defined using right Hilbert A-modules (see [7]). For positive elements a, b ∈ A ⊗ K we say that a is compactly contained in b, written a ≪ b if whenever [b] ≤ supn∈N[bn] for an increasing sequence {[bn]}n∈N, then there exists a natural number n such that [a] ≤ [bn]. Coward, Elliott and Ivanescu [7] showed that Cu(A) has the following properties: (1) every increasing sequence in Cu(A) has a supremum, (2) for any element [a] in Cu(A) there exists an increasing sequence {[an]}n∈N of Cu(A) such that [an] ≪ [an+1] for any n ∈ N and [a] = sup[an], (3) the operation of passing to the supremum of an increasing sequence and the relation ≪ are compatible with addition. Moreover they showed that Cu(A) is a functor which is continuous with respect to inductive limits ([7, Theorem 2]). For a positive element a ∈ A ⊗ K and ǫ > 0 we denote by (a − ǫ)+ the element f (a) in A ⊗ K where f (t) = max{0, t − ǫ}, t ∈ σ(a). Then we have (a − ǫ)+ ≪ a. Following the definition in [40], the Cuntz semigroup Cu(A) is said to be almost unperforated if (k + 1)[a] ≤ k[b] for some k ∈ N implies that [a] ≤ [b]. Rørdam 8 NORIO NAWATA showed that if A is Z-stable, then Cu(A) is almost unperforated (see [40, Theorem 4.5]). If A is a simple exact C∗-algebra with traces, then Cu(A) is almost unper- forated if and only if A has strict comparison, that is, if a, b ∈ (A ⊗ K)+ with dτ (a) < dτ (b) < ∞ for any τ ∈ T (A), then [a] ≤ [b]. (See [10, Proposition 4.2, Remark 4.3 and Proposition 6.2] and [40, Proposition 3.2 and Corollary 4.6].) Lemma 3.1. Let A be a simple C∗-algebra and a a non-zero positive element in A ⊗ K. Then for any positive element b in A ⊗ K, [b] ≤ supn∈N n[a]. ∞ Proof. Let B := (a + b)(A ⊗ K)(a + b). Then B is a σ-unital hereditary subalgebra of A ⊗ K. By a variant of Brown's theorem (see for example [27, Theorem 1.9]), aHB is isomorphic to HB as a right Hilbert module. Since bHB ⊆ HB, we see that [b] ≤ supn∈N n[a] in Cu(B). We obtain the conclusion because B is a hereditary subalgebra of A ⊗ K. (cid:3) The following proposition is an immediate corollary of [10, Theorem 6.6]. (Note that they considered the more general case.) But we shall give a self-contained proof based on their arguments (see also [10, Proposition 6.4]). Proposition 3.2. Let A be a simple exact C∗-algebra, and let a and b be posi- tive elements in A ⊗ K. Assume that Cu(A) is almost unperforated and 0 is an accumulation point of the spectrum σ(a) of a. Then if dτ (a) ≤ dτ (b) < ∞ for any τ ∈ T (A), then a is Cuntz smaller than b. Proof. Let a and b be positive elements in A ⊗ K such that dτ (a) ≤ dτ (b) for any τ ∈ T (A). We may assume that kak = kbk = 1. For any k ∈ N we have dτ (diag( a, .., a)) = kdτ (a) ≤ kdτ (b) < (k + 1)dτ (b) = dτ (diag( k z } { k+1 z}{ b, .., b)). Hence k[a] ≤ (k + 1)[b] for any k ∈ N because A has strict comparison. Let ǫ > 0, and choose a positive function cǫ on σ(a) such that cǫ(t) > 0 on t ∈ (0, ǫ) and cǫ(t) = 0 on σ(a) \ (0, ǫ). Then we have [cǫ(a)] + [(a − ǫ)+] ≤ [a]. Note that for any ǫ > 0, cǫ(a) is a nonzero positive element because 0 is an accumulation point of σ(a). Hence we have 2[a] ≤ supn∈N n[cǫ] by Lemma 3.1. There exists a natural number m such that 2[(a − ǫ)+] ≤ m[cǫ(a)] since 2[(a − ǫ)+] ≪ 2[a]. Therefore we have (m + 2)[(a − ǫ)+] ≤ m[(a − ǫ)+] + m[cǫ(a)] ≤ m[a] ≤ (m + 1)[b]. By the assumption that Cu(A) is almost unperforated, we see that [(a − ǫ)+] ≤ [b] for any ǫ > 0, and hence we have [a] ≤ [b]. (cid:3) Corollary 3.3. Let A be a simple exact stably projectionless C∗-algebra, and let a and b be positive elements in A ⊗ K. Assume that Cu(A) is almost unperforated. Then if dτ (a) = dτ (b) < ∞ for any τ ∈ T (A), then a is Cuntz equivalent to b. Proof. For any nonzero positive element a in A ⊗ K, 0 is an accumulation point of σ(a) because A is a stably projectionless C∗-algebra. Hence we obtain the conclu- sion by Proposition 3.2. (cid:3) Based on the result in [38], we say that a C∗-algebra A has almost stable rank one if for every σ-unital hereditary subalgebra B ⊆ A ⊗ K we have B ⊆ GL(eB). Robert showed that if A is a simple Z-stable stably projectionless C∗-algebra, then A has almost stable rank one (see [38, Corollary 4.5] and [40]). The following proposition is [38, Proposition 4.7]. See [7, Theorem 3] for the proof. Proposition 3.4. Let A be a simple σ-unital C∗-algebra such that A has almost stable rank one and a and b positive elements in A ⊗ K. Then a is Cuntz smaller than b if and only if there exists a right Hilbert A-module X ⊆ bHA such that X PICARD GROUPS OF CERTAIN STABLY PROJECTIONLESS C∗-ALGEBRAS 9 is isomorphic to aHA as a right Hilbert A-module, and a is Cuntz equivalent to b if and only if aHA is isomorphic to bHA as a right Hilbert A-module. Corollary 3.3 and Proposition 3.4 are important in the proof of our main result. These propositions show that every separable simple Z-stable stably projection- less C∗-algebra A with a unique tracial state has similar properties of II1 factors (Murray-von Neumann comparison theory). Moreover we have the following propo- sition. Proposition 3.5. Let A be a simple exact σ-unital stably projectionless C∗-algebra with a unique tracial state τ and no unbounded trace. Assume that Cu(A) is almost unperforated, A has almost stable rank one and F (A) = R× +. Then every nonzero hereditary subalgebra of A is isomorphic to A. Proof. Let B be a non-zero hereditary subalgebra of A. Then B is isomorphic to h0Ah0 for some non-zero positive element h0 in A. Since dτ (h0) ∈ R× + = F (A), there exists a positive element h in A ⊗ K such that dτ (h) = dτ (h0) and h(A ⊗ K)h is isomorphic to A. But then h ∼ h0 by Corollary 3.3 and so hHA is isomorphic to (h0 ⊗ e11)HA by Proposition 3.4. Hence A ∼= KA(hHA) ∼= KA((h0 ⊗ e11)HA) ∼= B. (cid:3) 4. Main result The following theorem is the main result in this paper. See [21, Corollary 4.8] and [31, Proposition 3.26] for the unital case. Theorem 4.1. Let A be a simple exact σ-unital stably projectionless C∗-algebra with a unique traical state τ and no unbounded trace. Assume that Cu(A) is almost unperforated and A has almost stable rank one. Then the following sequence is exact: 1 −−−−→ Out(A) ρA−−−−→ Pic(A) T−−−−→ F (A) −−−−→ 1. Proof. It is clear that T is onto by definition of F (A). We see that ρA is one-to-one and Im(ρA) ⊆ Ker(T ) by [5, Corollary 3.2] and Proposition 2.7 respectively. We shall show that Ker(T ) ⊆ Im(ρA). Let [E] ∈ Ker(T ). Then Corollary 3.3 and Proposition 3.4 imply E is isomorphic to (s ⊗ e11)HA as a right Hilbert A-module where s is a strict positive element in A and e11 is a rank one projection in K because we have dτ (s ⊗ e11) = 1 by kτ k = 1. Since (s ⊗ e11)HA is isomorphic to XA as a right Hilbert A-module, there exists some automorphism α such that [E] = [E A (cid:3) α ] by Proposition 2.1. Hence [E] ∈ Im(ρA). Corollary 4.2. Let A be a simple exact separable Z-stable stably projectionless C∗-algebra with a unique tracial state τ and no unbounded trace. Then the follow- ing sequence is exact: 1 −−−−→ Out(A) ρA−−−−→ Pic(A) T−−−−→ F (A) −−−−→ 1. Proof. This is an immediate consequence of [40, Theorem 4.5], [38, Corollary 4.5] and Theorem 4.1. (cid:3) Remark 4.3. There exists a unital simple AF algebra A with a unique tracial state such that Out(A) is not a normal subgroup of Pic(A). (See [30].) Of course A is a unital stably finite Z-stable C∗-algebra. Therefore the corollary above shows that Z-stable stably projectionless C∗-algebras are in this sense more well-behaved than unital stably finite Z-stable C∗-algebras. 10 NORIO NAWATA We shall show some examples. Let W2 be the Razak-Jacelon algebra studied in [13], [37] and [38], which has triv- ial K-groups and a unique tracial state and no unbounded trace. The Razak-Jacelon algebra W2 is constructed as an inductive limit C∗-algebra of Razak's building block in [34], that is, k z } { c ∈ Mn(C) k+1 c, .., c), z } { c, .., c, 0n), f (1) = diag( f ∈ C([0, 1]) ⊗ Mm(C) f (0) = diag( A(n, m) =     where n and m are natural numbers with nm and k := m n − 1. Let O2 denote the Cuntz algebra generated by 2 isometries S1 and S2. For every λ1, λ2 ∈ R there exists by universality a one-parameter automorphism group α of O2 given by αt(Sj) = eitλj Sj. Kishimoto and Kumjian showed that if λ1 and λ2 are all nonzero of the same sign and λ1 and λ2 generate R as a closed subgroup, then O2 ⋊α R is a simple stable projectionless C∗-algebra with unique (up to scalar multiple) densely defined lower semicontinuous trace in [19] and [20]. Moreover Robert [37] showed that W2 ⊗ K is isomorphic to O2 ⋊α R for some λ1 and λ2. In particular, W2 ⊗ K has a one parameter trace scaling automorphism group σ (see [19]). (See also [9].) Theorem 4.4. The Picard group of Razak-Jacelon algebra W2 is isomorphic to a semidirect product of Out(W2) with R× +. Moreover if A is a simple exact σ-unital C∗-algebra with a unique tracial state τ and no unbounded trace, then the Picard group of A ⊗ W2 is isomorphic to a semidirect product of Out(A ⊗ W2) with R× +. Proof. Note that we see that A ⊗ W2 is stably projectionless C∗-algebra because A ⊗ W2 ⊗ K has a one parameter trace scaling automorphism group id ⊗ σ. Since W2 is Z-stable, we have the following exact sequence: 1 −−−−→ Out(A ⊗ W2) ρA−−−−→ Pic(A ⊗ W2) T−−−−→ F (A ⊗ W2) −−−−→ 1 by Corollary 4.2. By Proposition 2.8, we see that F (A ⊗ W2) = R× + and the exact sequence above splits because A ⊗ W2 ⊗ K has a one parameter trace scaling automorphism group. Consequently Pic(A ⊗ W2) is isomorphic to Out(A ⊗ W2) ⋊ R× +. (cid:3) Remark 4.5. (i) Note that we have Out(W2 ⊗ K) ∼= Out(W2) ⋊ R× +. (ii) We do not assume that A is nuclear in the theorem above. Hence we have Pic(W2 ⊗ C ∗ r (Fn)) ∼= Out(W2 ⊗ C ∗ r (Fn)) ⋊ R× + r (Fn). where Fn is a non-amenable free group with n generators. Moreover Proposition 3.5 shows that every nonzero hereditary subalgebra of W2 ⊗ C ∗ r (Fn) is isomorphic to W2 ⊗ C ∗ (iii) Let B be a simple unital AF algebra with two extremal tracial states. Then W2 ⊗ B is a simple stably projectionless C∗-algebra with two extremal tracial states and in the class of Robert's classification theorem [37]. It can be checked that Out(W2 ⊗ B) is not a normal subgroup of Pic(W2 ⊗ B) by Robert's classification theorem and a similar proposition as [21, Proposition 1.5]. (We need to replace the K0-groups with the trace spaces.) PICARD GROUPS OF CERTAIN STABLY PROJECTIONLESS C∗-ALGEBRAS 11 5. Z-stability of stably projectionless C∗-algebras In this section we shall generalize the result of Matui and Sato in [25] to stably projectionless C∗-algebras. Note that our arguments are essentially based on their arguments. We shall review some results of Kirchberg's central sequence algebra in [17]. We denote by A the unitization algebra of A. Note that we consider A = A when A is unital. For a separable C∗-algebra A, set c0(A) := {(an)n∈N ∈ ℓ∞(N, A) lim n→∞ kank = 0}, A∞ := ℓ∞(N, A)/c0(A). Let B be a C∗-subalgebra of A. We identify A and B with the C∗-subalgebras of A∞ consisting of equivalence classes of constant sequences. Put A∞ := A∞ ∩ A′, Ann(B, A∞) := {(an)n ∈ A∞ ∩ B′ (an)nb = 0 for any b ∈ B}. Then Ann(B, A∞) is an closed two-sided ideal of A∞ ∩ B′, and define F (A) := A∞/Ann(A, A∞). It is easy to see that [(hn)n] is a unit in F (A). We call F (A) the central sequence algebra of A. A sequence (an)n is said to be central if limn→∞ kana − aank = 0 for all a ∈ A. A central sequence is a rep- resentative of an element in A∞. Since A is separable, A has a countable ap- proximate unit {hn}n∈N. If A is unital, then F (A) = A∞. Moreover we see that F (A) is isomorphic to M (A)∞ ∩ A′/Ann(A, M (A)∞) since for any (yn)n ∈ M (A)∞ ∩ A′, (ynhn)n is a central sequence in A and [(yn)n] = [(ynhn)n] in M (A)∞ ∩ A′/Ann(A, M (A)∞). Let {eij}i,j∈N be the standard matrix units of K. Define a map ϕ of F (A) to i=1 eii)n]. Then it is easily seen that ϕ is a well- defined injective homomorphism. A similar argument as above shows any element i,j=1 xn,i,j ⊗ ei,j)n] for some sequence {xn,i,j}n∈N in A. Using matrix units and the centrality of sequence, we can show that if i 6= j, then limn→∞ xn,i,ja = 0 for any a ∈ A and limn→∞(xn,i,i − xn,j,j)a = 0 for any i, j ∈ N and a ∈ A. Since M∞(A) is dense in A ⊗ K, it can be checked that ϕ is surjective. Hence F (A) is isomorphic to F (A ⊗ K). (See [17, Proposition 1.9] for more general cases.) F (A ⊗ K) by ϕ([(xn)n]) = [(xn ⊗Pn in F (A ⊗ K) is equal to [(Pn We denote by I(k, k + 1) the prime dimension drop algebra {f ∈ C([0, 1]) ⊗ Mk(C) ⊗ Mk+1(C) f (0) ∈ Mk(C) ⊗ idk+1, f (1) ∈ idk ⊗ Mk+1(C)} for k ∈ N. The Jiang-Su algebra Z is constructed as an inductive limit C∗-algebra of prime dimension drop algebras in [14]. We shall show the following proposition (which is based on [48, Proposition 2.2]) by a similar way as in [39, Theorem 7.2.2]. See [17, Proposition 4.11] for more general cases. Proposition 5.1. Let A be a separable C∗-algebra. There exist a unital homo- morphism of the prime dimension drop algebra I(k, k + 1) to F (A) for any k ∈ N if and only if A is Z-stable. Proof. Assume that there exist a unital homomorphism of the prime dimension drop algebra I(k, k + 1) to F (A) for any k ∈ N. By a similar argument as in [48, Proposition 2.2] and the construction of Z in [14], we see that there exists a unital homomorphism α of Z to F (A). Let ϕ be an injective homomorphism of A to A ⊗ Z defined by ϕ(a) = a ⊗ 1Z , and put C := M (A ⊗ Z)∞ ∩ ϕ(A)′/Ann(ϕ(A), M (A ⊗ Z)∞). Then we can regard α as a unital homomorphism of Z to C since F (A) is isomorphic to M (A)∞ ∩ A′/Ann(A, M (A)∞). Define a unital homomorphism of β of Z to M (A ⊗ Z)∞ ∩ ϕ(A)′ by β(x) = (1M(A) ⊗ x)n, and let [β] : Z → C be the quotient homomorphism of β. Then we see that C∗(α(Z), [β](Z)) in C is isomorphic to Z ⊗ Z. Since Z has 12 NORIO NAWATA approximately inner flip and is K1-injective (see [47, Proposition 1.13]), there exists a sequence {wm}m∈N of unitary elements in C such that limm→∞ w∗ m[β](x)wm = α(x) for any x ∈ Z and wm is in the connected component of 1C in U (C) for any m ∈ N. Since wm is in the connected component of 1C in U (C), there exists a unitary element um in M (A ⊗ Z)∞ ∩ ϕ(A)′ such that [um] = wm for any m ∈ N. For any a ∈ A, x ∈ Z and all y ∈ M (A ⊗ Z)∞ ∩ ϕ(A)′ such that [y] = α(x), we have yϕ(a) = lim m→∞ u∗ mβ(x)umϕ(a) = lim m→∞ u∗ mβ(x)ϕ(a)um = lim m→∞ u∗ m(a ⊗ x)um by [y] = limm→∞[u∗ mβ(x)um] and the definition of Ann(ϕ(A), M (A ⊗ Z)∞). Since [y] = α(x), we can take y ∈ M (ϕ(A))∞ ∩ ϕ(A)′ ⊆ M (A ⊗ Z)∞ ∩ ϕ(A)′. Hence we m(a⊗x)um is an element in ϕ(A)∞. Therefore for any z ∈ A⊗Z, see that limm→∞ u∗ limm→∞ d(u∗ mzum, ϕ(A)∞) = 0. We see that A is Z-stable by a similar argument as in [39, Proposition 2.3.5 and Proposition 7.2.1]. Conversely assume that A is Z-stable. Then A is isomorphic to A ⊗ (⊗∞ k=1Z). Since M (A) is the largest unital C∗-algebra that contains A as an essential ideal, A⊗ (⊗∞ k=1Z)). Hence there exists a unital homomorphism of Z to M (A)∞ ∩ A′. Therefore we see that there exists a unital homomorphism of the prime dimension drop algebra I(k, k + 1) to F (A) for any k ∈ N because F (A) is isomorphic to M (A)∞ ∩ A′/Ann(A, M (A)∞). (cid:3) k=1Z) is a unital subalgebra of M (A⊗ (⊗∞ We denote by Ped(A) the Pedersen ideal of A. The Pedersen ideal Ped(A) is a minimal dense two-sided ideal of A. Hence every densely defined lower semicontin- uous trace τ on A is finite on Ped(A) because τ is finite on a dense two-sided ideal. Moreover for any positive element h in Ped(A), hAh is contained in Ped(A). We refer the reader to [3, II 5.2.4] and [33, Section 5.6] for details of the Pedersen ideal. If A is unital, every densely defined lower semicontinuous trace on A is bounded. Hence if A is simple and A ⊗ K has a nonzero projection, then there exists a full hereditary subalgebra B of A such that every densely defined lower semicontinuous trace on B is bounded. In general, we have the following proposition. Proposition 5.2. Let A be a σ-unital simple C∗-algebra. Then there exists a full hereditary subalgebra B of A such that every densely defined lower semicontinuous trace on B is bounded. Proof. Let h be a nonzero positive element in Ped(A). Then any τ ∈ T (A) restricts to a bounded trace on hAh because every positive liner functional is automatically bounded. We obtain the conclusion by Corollary 2.4. (cid:3) If A is separable, then A is Z-stable if and only if some full hereditary subalgebra is Z-stable by Proposition 5.1 and Brown's theorem in [4] since F (A) is isomorphic to F (A ⊗ K). (See also [47].) Therefore we may assume that A has no unbounded trace by the proposition above. Note that if A has strict comparison and no un- bounded trace, then for any a, b ∈ A+ satisfying dτ (a) < dτ (b) for all τ ∈ T1(A), we have a - b. Proposition 5.3. Let A be a separable C∗-algebra such that T1(A) is a non-empty compact set, and let {hm}m∈N be a countable approximate unit for A and ǫ > 0. Then there exists a natural number N such that max τ ∈T1(A) τ (fn) − τ (hmfn) < ǫ for any m ≥ N and for any sequence (fn)n∈N of positive contractions in A. particular, we have In lim n→∞ max τ ∈T1(A) τ (hnfn) − τ (fn) = 0. PICARD GROUPS OF CERTAIN STABLY PROJECTIONLESS C∗-ALGEBRAS 13 Proof. For any τ ∈ T1(A), we have τ (hm) ≤ τ (hm+1) and lim τ (hm) = 1. By Dini's theorem, there exists a natural number N such that max τ ∈T1(A) 1 − τ (hm) < ǫ for any m ≥ N . For any sequence (fn)n∈N of positive contractions in A, max τ ∈T1(A) τ (fn) − τ (hmfn) = max τ ∈T1(A) τ ((1 − hm)1/2fn(1 − hm)1/2) for m ≥ N . We recall some definitions. ≤ max τ ∈T1(A) 1 − τ (hm) < ǫ (cid:3) Definition 5.4. Let A be a separable C∗-algebra with no unbounded trace. Assume that T1(A) is a non-empty compact set. We say that A has property (SI) if for any central sequences (en)n and (fn)n of positive contractions in A satisfying lim n→∞ max τ ∈T1(A) τ (en) = 0, lim m→∞ lim inf n→∞ min τ ∈T1(A) τ (f m n ) > 0, there exists a central sequence (sn)n in A such that lim n→∞ ks∗ nsn − enk = 0, lim n→∞ kfnsn − snk = 0. For a completely positive map ϕ of A to A, we say that ϕ can be excised in small central sequences in A if for any central sequences (en)n and (fn)n of positive contractions in A satisfying the property above, there exists a sequence (sn)n∈N in A such that lim n→∞ ks∗ nasn − ϕ(a)enk = 0 for any a ∈ A, lim n→∞ kfnsn − snk = 0. Remark 5.5. In the definition above, it is important that en and fn are elements in A. We see that if id A can be excised in small central sequences in A, then A has property (SI) (see [25, Proof of (iii)⇒(iv) of Theorem1.1]). We shall generalize [24, Lemma 4.6] and [25, Lemma 2.4] to non-unital C∗- algebras. Lemma 5.6. Let c be a positive element in a separable C∗-algebra A such that T1(A) is a non-empty compact set, and let θ ∈ R. For any central sequence (fn)n of positive contractions in A, we have lim sup n→∞ max τ ∈T1(A) τ (cfn) − θτ (fn) ≤ 2 max τ ∈T1(A) τ (c) − θ. Proof. Let {hm}m∈N be a countable approximate unit for A. Replacing fn and θ in [24, Lemma 4.6] with hmfn and θhm respectively, the same argument in the proof of [24, Lemma 4.6] shows that lim sup n→∞ max τ ∈T1(A) τ (cfn) − θτ (hmfn) ≤ 2 max τ ∈T1(A) τ (c) − θτ (hm) for any m ∈ N. By Proposition 5.3, we have lim sup n→∞ max τ ∈T1(A) τ (cfn) − θτ (fn) ≤ 2 max τ ∈T1(A) τ (c) − θ. (cid:3) Lemma 5.7. Let A be a separable simple C∗-algebra such that T1(A) is a non- empty compact set, and let a be a nonzero positive element in A. If (fn)n is a central sequence of positive contractions in A such that lim m→∞ lim inf n→∞ min τ ∈T1(A) τ (f m n ) > 0, 14 then NORIO NAWATA lim m→∞ lim inf n→∞ min τ ∈T1(A) τ (f m/2 n af m/2 n ) > 0. Proof. Put R := a1/2A. Since A is simple, R is a right ideal of A such that R∗R = AaA is a dense ideal of A. Therefore there exists a sequence {vj}j∈N in A j avj}n∈N is an approximate unit for A by a similar argument as in [4, Lemma 2.3]. By Proposition 5.3, there exists a natural number N such that such that {Pn j=1 v∗ lim m→∞ lim inf n→∞ min τ ∈T1(A) τ ( NX j=1 j avj f m v∗ n ) > 0. We have lim m→∞ lim inf n→∞ min τ τ ( NX j=1 j avjf m v∗ n ) = lim m→∞ lim inf n→∞ min τ = lim m→∞ lim inf n→∞ min τ j=1 NX NX j=1 τ (v∗ j a1/2f m n a1/2vj) τ (f m/2 n a1/2vjv∗ j a1/2f m/2 n ≤ NX j=1 kvjk2 lim m→∞ lim inf n→∞ min τ τ (f m/2 n af m/2 n ). Hence we obtain the conclusion. ) (cid:3) Let A be a separable simple C∗-algebra, and let τ be a tracial state on A. Consider the GNS representation (πτ , Hτ , ξτ ) associated with τ . Then πτ (A)′′ is a finite von Neumann algebra and πτ (A) is strongly dense subalgebra of πτ (A)′′ in general. (Indeed, every approximate unit for πτ (A) is strongly convergent to 1Hτ .) In particular, Kaplansky density theorem shows that for any positive contraction H ∈ πτ (A)′′ there exists a sequence {an}n∈N of positive contractions in A such that π(an) is strongly converge to H. We can identify C∗(πτ (A), 1Hτ ) in B(Hτ ) with its unitization algebra A. Therefore the same proof as [42, Lemma 2.1] shows the following lemma. See also [26, Proposition 3.5 and Theorem 4.3]. Lemma 5.8. ([42, Lemma 2.1]) Let A be a separable simple nuclear C∗-algebra, and let τ be a tracial state on A. For any sequence {Hn}n∈N of positive contractions in πτ (A)′′ such that k[Hn, x]kτ → 0 for all x ∈ πτ (A)′′, there exists a central sequence (cn)n of positive contractions in A such that kcn − Hnkτ → 0. Maybe someone considers that [42, Lemma 2.1] depends on a unit for the ap- plication of Haagerup's theorem ([12, Theorem 3.1]); see for example [11, Theorem 2.1] for details. But we can check that the same proof of [42, Lemma 2.1] works for non-unital C∗-algebras because A is a two-sided ideal of A and for any positive contraction H ∈ πτ (A)′′ there exists a sequence {an}n∈N of positive contractions in A such that π(an) is strongly converge to H. If τ is an extremal tracial state on a separable simple infinite-dimensional nuclear C∗-algebra A, then πτ (A)′′ is the AFD II1 factor in general. Therefore Lemma 5.8 and the same proof as [25, Lemma 3.3] show the following lemma. Lemma 5.9. ([25, Lemma 3.3]) Let A be a separable simple infinite-dimensional nuclear C∗-algebra with finitely many extremal tracial states. For any k ∈ N, there exist central sequences (ci,n)n in A , i = 1, 2, .., k such that c1,n is a positive contraction for any n ∈ N, (ci,nc∗ j,n)n = PICARD GROUPS OF CERTAIN STABLY PROJECTIONLESS C∗-ALGEBRAS 15 δi.j(c2 1,n)n and for any m ∈ N. lim n→∞ max τ ∈T1(A) τ (cm 1,n) − 1 k = 0 Note that we need to consider unitaries in A in the proof above. But it is also no problem because A is a two-sided ideal of A. Let ω be a pure state on A. Then we can uniquely extend ω to a pure state ω on A. Moreover if A is a separable simple non-type I C∗-algebra, then πω(A)∩K(Hω) = {0}. Therefore the same proof as [25, Lemma 3.1] shows that every completely positive map of A to A can be approximated in the pointwise norm topology by completely positive map ϕ of the form ϕ(x) = NX l=1 NX i,j=1 ω(d∗ i xdj )c∗ l,icl,j, x ∈ A i xdj)c∗ i,j=1 ω(d∗ where cl,i, di ∈ A. For 1 ≤ l ≤ N , let ϕl(x) = PN l,icl,j. Then ϕ = ϕ1 + ... + ϕN . Using Lemma 5.7 instead of [25, Lemma 2.4], we can prove a version of [25, Proposition 2.2], i.e. that each ϕl can be excised in small central sequences in A. (See the proof of [25, Lemma 2.5], which is where [25, Lemma 2.4] gets used; note also this where we need strict comparison.) We can check that [25, Lemma 3.4] holds without the assumption of a unit by using Lemma 5.6 and Lemma 5.9 instead of [24, Lemma 4.6] and [25, Lemma 3.3] respectively. By this lemma, we see that a sum of completely positive maps A → A, each of which can be excised in small central sequences in A, can itself be excised in small central sequences in A. Therefore we obtain the following theorem. Theorem 5.10. Let A be a separable simple infinite-dimensional nuclear C∗- algebra with finitely many extremal tracial states and no unbounded trace. If A has a strict comparison, then any completely positive map of A to A can be excised in small central sequences in A. The following theorem is the main theorem in this section. Theorem 5.11. Let A be a separable simple infinite-dimensional non-type I nu- clear C∗-algebra with a finite dimensional lattice of densely defined lower semicon- tinuous traces. Then A has strict comparison if and only if A is Z-stable. Proof. Rørdam showed that if A is Z-stable, then A has strict comparison (see [40, Corollary 4.6]). We shall show the only if part. By Proposition 5.2, we may assume that A has no unbounded trace. Hence A has property (SI) by Remark 5.5 and Theorem 5.10. For any k ∈ N, there exist central sequences (ci,n)n in A , i = 1, 2, .., k such that c1,n is a positive contraction , (ci,nc∗ j,n)n = δi.j(c2 1,n)n and lim n→∞ max τ ∈T1(A) τ (cm 1,n) − 1 k = 0 for any m ∈ N by Lemma 5.9. Let {hn}n∈N be an approximate unit for A. Taking a suitable subsequence of {hn}n∈N, we may assume that (hn)n(c1,n)n = (c1,n)n(hn)n. Define central sequences (fi,n)n in A, i = 1, .., k by (fi,n)n := (ci,nh1/2 i,nfi,n)n. Then we may assume that (en)n is a central sequence of positive contractions in A. Proposition 5.3 implies limn→∞ maxτ τ (f ∗ n )n, and put (en)n := (hn − Pk i,nci,n) = 0 for any 1 ≤ i ≤ k, and hence we have i,nfi,n − c∗ i=1 f ∗ lim n→∞ max τ ∈T1(A) τ (en) = 0. 16 NORIO NAWATA Note that (f1,n)n is a central sequence of positive contractions in A by the assump- tion of (hn)n. Because {h1/2 n }n∈N is also an approximate unit for A, we have lim sup n→∞ max τ kc1,n − c1/2 1,n h1/2 n c1/2 1,n k2 τ = lim sup n→∞ max τ τ ((c1,n − c1/2 1,n h1/2 n c1/2 1,n )2) ≤ lim sup n→∞ max τ = 0 by Proposition 5.3. Hence τ (c1,n − c1/2 1,n h1/2 n c1/2 1,n ) lim sup n→∞ max τ τ (cm 1,n) − τ (f m 1,n) = lim sup n→∞ max τ τ (cm 1,n − (c1,nh1/2 n )m) = lim sup n→∞ max τ ≤ lim sup n→∞ max τ τ (cm 1,n − (c1/2 1,n h1/2 n c1/2 1,n )m) kcm 1,n − (c1/2 1,n h1/2 n c1/2 1,n )m)kτ = 0 for any m ∈ N. Therefore we have lim m→∞ lim inf n→∞ min τ ∈T1(A) τ (f m 1,n) = 1/k > 0. i=1 f ∗ nsn + Pk i,nfi,n)n = (hn)n and (f1,nsn)n = (sn)n. We have [(fi,nf ∗ Since A has property (SI), there exists a central sequence (sn)n in A such that (s∗ j,n)n] = δi.j[(f 2 n )n] is a unit in F (A). It follows from [41, Proposition 2.1] that there exists a unital homomorphism of I(k, k + 1) to F (A). Consequently A is Z-stable by Proposition 5.1. (cid:3) i,nfi,n)n] = 1 in F (A) because [(h1/2 nsn +Pk 1,n)n] and [(s∗ i=1 f ∗ Remark 5.12. Let A be a separable simple infinite-dimensional non-type I nuclear C∗-algebra with a finite dimensional lattice of densely defined lower semicontinuous traces, that has strict comparison. Since A is Z-stable by the theorem above, there exists a unital homomorphism of Z to M (A)∞ ∩ A′. But we do not know that we could show this fact directly without using Kirchberg's central sequence algebras. Note that if A is non-unital, then there exists no unital homomorphism of Z to ( A)∞ ∩ A′ because A is not Z-stable. The following corollary is an immediate consequence of the theorem above and Corollary 4.2. Corollary 5.13. Let A be a separable simple nuclear stably projectionless C∗- algebra with a unique tracial state and no unbounded trace. Assume that A has strict comparison. Then we have the following exact sequence: 1 −−−−→ Out(A) ρA−−−−→ Pic(A) T−−−−→ F (A) −−−−→ 1. We shall consider some examples. We refer the reader to [45] for details of slow dimensional growth for nonunital C∗-algebras. Tikuisis showed that if A is a simple separable approximately subhomogeneous C∗-algebra with slow dimension growth, then Cu(A) is almost unperforated in [45, Corollary 5.9]. The following immediate corollary of this result and Theorem 5.11 is suggested by the referee. Corollary 5.14. Let A be a simple separable non-type I approximately subhomo- geneous C∗-algebra with slow dimension growth and a finite dimensional lattice of densely defined lower semicontinuous traces. Then A is Z-stable. PICARD GROUPS OF CERTAIN STABLY PROJECTIONLESS C∗-ALGEBRAS 17 We say that A is a 1-dimensional NCCW complex if A is a pullback C∗-algebra of the form A yπ1 C([0, 1]) ⊗ F π2−−−−→ E yρ δ0⊕δ1−−−−→ F ⊕ F where E and F are finite-dimensional C∗-algebras and δi is the evaluation map at i. Razak's building block A(n, m) is a 1-dimensional NCCW complex. The Cuntz semigroup of a 1-dimensional NCCW complex was computed in [1]. Every simple inductive limit C∗-algebras of 1-dimensional NCCW complexes is approximately subhomogeneous C∗-algebra with slow dimension growth. The following example is also suggested by the referee. Example 5.15. For n ≥ 2, let On denote the Cuntz algebra generated by n isome- tries S1, ..., Sn. Given λ1, ..., λn ∈ R, there exists by universality a one-parameter automorphism group α of On given by αt(Sj) = eitλj Sj. Kishimoto and Kumjian showed that if λj are all nonzero of the same sign and {λ1, ..., λn} generates R as a closed subgroup, then On ⋊ R is a simple stable projectionless C∗-algebra with unique (up to scalar multiple) densely defined lower semicontinuous trace in [19] and [20]. In particular, On ⋊α R has a one parameter trace scaling automorphism group. Dean showed that there exist many sets of numbers {λ1, .., λn} such that On ⋊α R can be expressed as an inductive limit C∗-algebra of 1-dimensional NCCW- complexes in [9, Theorem 5.1]. Therefore for α defined by such a set of numbers {λ1, .., λn}, On ⋊α R is Z-stable. Moreover it can be checked that for any positive element h in Ped(On ⋊α R), Pic(h(On ⋊α R)h) is isomorphic to a semidirect prod- uct of Out(h(On ⋊α R)h) with R× + by the same argument of the proof in Theorem 4.4. Robert classified inductive limit C∗-algebras of 1-dimensional NCCW complexes with trivial K1-groups in [37]. Corollary 5.16. Let A be a simple stably projectionless C∗-algebra with a unique tracial state and no unbounded trace, that is expressible as an inductive limit C∗-algebra of 1-dimensional NCCW-complexes with trivial K1-groups and B a sep- arable simple C∗-algebra with a unique tracial state and no unbounded trace. Then we have the following exact sequence: 1 −−−−→ Out(A ⊗ B) ρA⊗B−−−−→ Pic(A ⊗ B) T−−−−→ R× + −−−−→ 1. Proof. For any r ∈ (0, 1) there exists a positive element h in A such that dτ (h) = r because A has a positive element with a continuous spectrum. Note that the class of C∗-algebras covered by Robert's classification theorem in [37] is closed under stable isomorphism (see [37, Theorem 1.0.1]). By [37, Proposition 3.1.7], we see that a classifying invariant of the class of C∗-algebras which contains A and hAh is equal to that of [37, Corollary 6.2.4]. Hence we see that A is isomorphic to hAh. Therefore F (A) = R× + and A ⊗ B is separable, A ⊗ B is a stably projectionless C∗-algebra by [29, Corollary 4.10]. Therefore we obtain the conclusion by Corollary 4.2 and Corollary 5.14. (cid:3) +. Since F (A ⊗ B) = R× We do not know whether the exact sequence above splits. This question is related to the existence of a one parameter trace scaling automorphism group of A ⊗ K. For any countable abelian groups G1 and G2, Kishimoto showed that there exists a stable projectionless simple separable nuclear C∗-algebra A with unique (up to scalar multiple) densely defined lower semicontinuous trace with K0(A) = G1 and K1(A) = G2 in [18]. These stably projectionless C∗-algebras are constructed as 18 NORIO NAWATA the crossed products O ⋊α R by certain one parameter automorphism groups α of Kirchberg algebras O and the dual actions of α are trace scaling actions of O ⋊α R. Hence it is natural to believe that there exists a kind of duality between Z-stable stably projectionless C∗-algebras (with unique trace) and O∞-stable C∗-algebras. From this view point, it seems to be possible to introduce the stably projectionless C∗-algebra Wn for any n ≥ 3. Hence we denote by W2 the Razak-Jacelon algebra. On the other hand, Tikuisis [45] constructed a simple separable nuclear stably projectionless C∗-algebra whose Cuntz semigroup is not almost unperforated. Acknowledgments The author would like to thank the Fields Institute, where a part of this work was done, for their hospitality. This travel is supported by the Global COE program "Education and Research Hub for Mathematics-for-Industry" at Kyushu University. He is also grateful to Leonel Robert for informing him about some results in [38] and to Hiroki Matui for useful suggestions. We also thank the referee for his or her careful reading and many valuable suggestions. References [1] R. Antoine, F. Perera and L. Santiago, Pullbacks, C(X)-algebras, and their Cuntz semigroup, J. Funct. Anal. 260 (2011), no. 10, 2844 -- 2880. [2] P. Ara, F. Perera and A. Toms, K-theory for operator algebras. Classification of C∗-algebras, Aspects of operator algebras and applications, 1 -- 71, Contemp. Math., 534, Amer. Math. Soc., Providence, RI, 2011. [3] B. Blackadar, Operator Algebras : Theory of C*-Algebras and von Neumann Algebras, En- cyclopaedia of Mathematical Sciences, 122, Springer, 2006. [4] L. G. Brown, Stable isomorphism of hereditary subalgebras of C ∗-algebras, Pacific J. Math. 71 (1977), 335 -- 348. [5] L. G. Brown, P.Green and M. A. Rieffel, Stable isomorphism and strong Morita equivalence of C∗-algebras, Pacific J. Math. 71 (1977), no. 2, 349 -- 363. [6] F. Combes and H. Zettl, Order structures, traces and weights on Morita equivalent C ∗- algebras, Math. Ann. 265 (1983), no. 1, 67 -- 81. [7] K. Coward, G. A. Elliott and C. Ivanescu, The Cuntz semigroup as an invariant for C∗- algebras, J. Reine Angew. Math. 623 (2008), 161 -- 193. [8] J. Cuntz, Dimension functions on simple C ∗-algebras, Math. Ann. 233 (1978), no. 2, 145 -- 153. [9] A. Dean, A continuous field of projectionless C∗-algebras, Canad. J. Math. 53 (2001), no. 1, 51 -- 72. [10] G. A. Elliott, L. Robert and L. Santiago, The cone of lower semicontinuous traces on a C∗-algebra, Amer. J. Math. 133 (2011), no. 4, 969 -- 1005. [11] H. Futamura, N. Kataoka and A. Kishimoto, Type III representations and automorphisms of some separable nuclear C ∗-algebras, J. Funct. Anal. 197 (2003), no. 2, 560 -- 575. [12] U. Haagerup, All nuclear C∗-algebras are amenable, Invent. Math. 74 (1983), no. 2, 305 -- 319. preprint, [13] B. Jacelon, A simple, monotracial, stably projectionless C*-algebra, arXiv:1006.5397v3 [math.OA], to appear in J. London Math. Soc. [14] X. Jiang and H. Su, On a simple unital projectionless C∗-algebra, Amer. J. Math. 121 (1999), no. 2, 359 -- 413. [15] T. Kajiwata, C. Pinzari and Y. Watatani, Jones index theory for Hilbert C∗-bimodules and its equivalence with conjugation theory, J. Funct. Anal. 215 (2004), no. 1, 1 -- 49. [16] T. Kajiwara and Y. Watatani, Jones index theory by Hilbert C∗-bimodules and K-theory, Trans. Amer. Math. Soc. 352 (2000), no. 8, 3429 -- 3472. [17] E. Kirchberg, Central sequences in C∗-algebras and strongly purely infinite algebras, Operator Algebras: The Abel Symposium 2004, 175 -- 231, Abel Symp., 1, Springer, Berlin, 2006. [18] A. Kishimoto, Pairs of simple dimension groups, Internat. J. Math. 10 (1999), no. 6, 739 -- 761. [19] A. Kishimoto and A. Kumjian, Simple stably projectionless C∗-algebras arising as crossed products, Canad. J. Math. 48 (1996), no. 5, 980 -- 996. [20] A. Kishimoto and A. Kumjian, Crossed products of Cuntz algebras by quasi-free automor- phisms, in Operator algebras and their applications (Waterloo, ON, 1994/1995), 173 -- 192, Fields Inst. Commun., 13, Amer. Math. Soc., Providence, RI, 1997. PICARD GROUPS OF CERTAIN STABLY PROJECTIONLESS C∗-ALGEBRAS 19 [21] K. Kodaka, Full projections, equivalence bimodules and automorphisms of stable algebras of unital C∗-algebras, J. Operator Theory, 37 (1997), no. 2, 357 -- 369. [22] K. Kodaka, Picard groups of irrational rotation C∗-algebras, J. London Math. Soc. (2) 56 (1997), no. 1, 179 -- 188. [23] K. Kodaka, Projections inducing automorphisms of stable UHF-algebras, Glasg. Math. J. 41 (1999), no. 3, 345 -- 354. [24] H. Matui and Y. Sato, Z-stability of crossed products by strongly outer actions, Comm. Math. Phys. 314 (2012), 193 -- 228. [25] H. Matui and Y. Sato, Strict comparison and Z-absorption of nuclear C∗-algebras, Acta Math. 209 (2012), no. 1, 179 -- 196. arXiv:1111.1637v1 [math.OA]. [26] H. Matui and Y. Sato, Z-stability of crossed products by strongly outer actions II , preprint, arXiv:1205.1590v1 [math.OA]. [27] J. A. Mingo and W. J. Phillips. Equivariant triviality theorems for Hilbert C ∗-modules, Proc. Amer. Math. Soc. 91 (1984), no. 2, 225 -- 230. [28] F. Murray and J. von Neumann, On rings of operators IV, Ann. Math. (2) 44, (1943), 716 -- 808. [29] N. Nawata, Fundamental group of simple C∗-algebras with unique trace III, Canad. J. Math. 64, (2012), no. 3, 573 -- 587. [30] N. Nawata, A note on trace scaling actions and fundamental groups of C∗-algebras, preprint:arXiv:1009.1722 [math.OA]. [31] N. Nawata and Y. Watatani, Fundamental group of simple C∗-algebras with unique trace, Adv. Math. 225 (2010), no. 1, 307 -- 318. [32] N. Nawata and Y. Watatani, Fundamental group of simple C∗-algebras with unique trace II, J. Funct. Anal. 260 (2011), no. 2, 428 -- 435. [33] G. K. Pedersen, C ∗-Algebras and Their Automorphism Groups, Academic Press, London- New York-San Francisco, 1979. [34] S. Razak, On the classification of simple stably projectionless C∗-algebras, Canad. J. Math. 54 (2002), no. 1, 138 -- 224. [35] I. Raeburn and D. P. Williams, Morita Equivalence and Continuous-Trace C∗-Algebras, Mathematical Surveys and Monographs, 60, American Mathematical Society, Providence, RI, 1998. [36] M. A. Rieffel, Morita equivalence for operator algebras, Operator algebras and applications, Part I (Kingston, Ont., 1980), pp. 285 -- 298, Proc. Sympos. Pure Math., 38, Amer. Math. Soc., Providence, R.I., 1982. [37] L. Robert, Classification of inductive limits of 1-dimensional NCCW complexes, Adv. Math. 231 (2012), no. 5, 2802 -- 2836. arXiv:1007.1964v2 [math.OA]. [38] L. Robert, Remarks on R, preprint, arXiv:1112.6069v1 [math.OA]. [39] M. Rørdam, Classification of nuclear, simple C∗-algebras, in:Classification of nuclear C ∗- algebras. Entropy in operator algebras, 1 -- 145, Encyclopaedia of Mathematical Sciences, 126, Springer, 2002. [40] M. Rørdam, The stable and the real rank of Z-absorbing C∗-algebras, Internat. J. Math. 15 (2004), no. 10, 1065 -- 1084. [41] Y. Sato, The Rohlin property for automorphisms of the Jiang-Su algebra, J. Funct. Anal. 259 (2010), no. 2, 453 -- 476. [42] Y. Sato, Discrete amenable group actions on von Neumann algebras and invariant nuclear C∗-subalgebras, preprint, arXiv:1104.4339v1 [math.OA]. [43] M. Takesaki, Theory of operator algebras. II, Encyclopaedia of Mathematical Sciences, 125, Springer, 2003. [44] M. Takesaki, Theory of operator algebras. III, Encyclopaedia of Mathematical Sciences, 127, Springer, 2003. [45] A. Tikuisis, Regularity for stably projectionless, simple C∗-algebras, J. Funct. Anal. 263 (2012), no. 5, 1382 -- 1407. [46] A. Toms, Characterizing classifiable AH algebras, C. R. Math. Acad. Sci. Soc. R. Can. 33 (2011), no. 4, 123 -- 126. [47] A. S. Toms and W. Winter, Strongly self-absorbing C∗-algebras, Trans. Amer. Math. Soc. 359 (2007), no. 8, 3999 -- 4029. [48] A. S. Toms and W. Winter, Z-stable ASH algebras, Canad. J. Math. 60 (2008), no. 3, 703 -- 720. [49] W. Winter, Nuclear dimension and Z-stability of pure C∗-algebras, Invent. Math. 187 (2012), no. 2, 259 -- 342. [50] Y. Watatani, Index for C∗-subalgebras, Memoir AMS 424 (1990). 20 NORIO NAWATA Department of Mathematics and Informatics, Graduate school of Science, Chiba University,1-33 Yayoi-cho, Inage, Chiba, 263-8522, Japan E-mail address: [email protected]
1105.1686
1
1105
2011-05-09T14:33:11
Geometry of unitary orbits of pinching operators
[ "math.OA", "math.DG" ]
Let I be a symmetrically-normed ideal of the space of bounded operators acting on a Hilbert space H. Let ${p_i}_1 ^w$ $(1\leq w \leq \infty)$ be a family of mutually orthogonal projections on H. The pinching operator associated with the former family of projections is given by P: I --> I, P(x)=\sum_{i=1}^{w} p_i x p_i. Let UI denote the Banach-Lie group of the unitary operators whose difference with the identity belongs to I. We study several geometric properties of the orbit UI(P)={L_{u} P L_{u^*} : u \in UI}, where L_u is the left representation of UI on the algebra B(I) of bounded operators acting on I. The results include necessary and sufficient conditions for UI(P) to be a submanifold of B(I). Special features arise in the case of the ideal K of compact operators. In general, UK(P) turns out to be a non complemented submanifold of B(K). We find a necessary and sufficient condition for UK(P) to have complemented tangent spaces in B(K). We also show that UI(P) is a covering space of another natural orbit of P. A quotient Finsler metric is introduced, and the induced rectifiable is studied. In addition, we give an application of the results on UI(P) to the topology of the UI-unitary orbit of a compact normal operator.
math.OA
math
Geometry of unitary orbits of pinching operators ∗ Eduardo Chiumiento and Mar´ıa E. Di Iorio y Lucero† Let I be a symmetrically-normed ideal of the space of bounded operators acting on a Hilbert space H. Let { pi }w 1 (1 ≤ w ≤ ∞) be a family of mutually orthogonal projections on H. The pinching operator associated with the former family of projections is given by Abstract P : I −→ I, P (x) = w X i=1 pixpi. Let UI denote the Banach-Lie group of the unitary operators whose difference with the identity belongs to I. We study several geometric properties of the orbit UI(P ) = { LuP Lu∗ : u ∈ UI } , where Lu is the left representation of UI on the algebra B(I) of bounded operators acting on I. The results include necessary and sufficient conditions for UI (P ) to be a submanifold of B(I). Special features arise in the case of the ideal K of compact operators. In general, UK(P ) turns out to be a non complemented submanifold of B(K). We find a necessary and sufficient condition for UK(P ) to have complemented tangent spaces in B(K). We also show that UI (P ) is a covering space of another natural orbit of P . A quotient Finsler metric is introduced, and the induced rectifiable is studied. In addition, we give an application of the results on UI (P ) to the topology of the UI-unitary orbit of a compact normal operator. 1 1 Introduction Let H be an infinite dimensional separable Hilbert space and B(H) the space of bounded linear operators acting on H. We denote by U the group of unitary operators on H. Let Φ be a symmetric norming function and I = SΦ the corresponding symmetrically-normed ideal of B(H) equipped with the norm k . kI. Let UI denote the group of unitaries which are perturbations of the identity by an operator in I, i.e. UI = { u ∈ U : u − 1 ∈ I }. It is a real Banach-Lie group with the topology defined by the metric d(u1, u2) = ku1 − u2kI, and its Lie algebra equals Ish = { x ∈ I : x∗ = −x }, which is the real Banach space of skew-hermitian operators in I (see [5]). Let { pi }w 1 (1 ≤ w ≤ ∞) be a family of mutually orthogonal hermitian projections in B(H). We do not make any assumption on the sum of all the projections of the family, ∗2010 MSC. Primary 46T05; Secondary 47B49, 47B10, 57N20, 58B20. †All authors partially supported by Instituto Argentino de Matem´atica and CONICET. 1Keywords and phrases: pinching operator, left representation, symmetrically-normed ideal, submanifold, covering map, Finsler metric. 1 so we could have that the projection p0 := 1 −Pw associated with { pi }w 1 is defined by i=1 pi is nonzero. The pinching operator P : I −→ I, P (x) = pixpi, w Xi=1 where in case w = ∞ the series is convergent in the uniform norm. Let B(I) denote the Banach algebra of bounded operators acting on I. Left multiplication defines the bounded linear operators Lx : I −→ I, Lx(y) = xy, for x ∈ B(H) and y ∈ I. The left representation of UI on B(I), namely UI −→ B(I), u 7→ Lu, allows us to introduce the following orbit UI(P ) := { LuP Lu∗ : u ∈ UI } . The aim of this paper is to study geometric properties of this orbit. Since every pinching operator is a continuous projection, the present work might be regarded as a modest con- tribution to the vast literature on the differential and metric geometry of unitary orbits of projections in different settings (see e.g. [1, 4, 7, 14, 15, 24]). Despite of some usual geometric properties that have already been studied in the afore-mentioned papers and still hold in this special orbit, we will also show some new special features of UI (P ), especially concerning with its submanifold structure. Pinching operators generalize the so-called notion of pinching of block matrices developed in matrix analysis (see e.g. [8]). In the framework of symmetrically-normed ideals, these operators have been studied in [17, 22]. If I is the trace class ideal, pinching operators arise in quantum mechanics due to a well-known postulate of von Neumann on the measurement of density operators [25]. More recently, they have been shown to be examples of the quantum reduction maps introduced in [20]. Let us describe the contents of the paper. In Section 2 we recall some basic facts on symmetrically-normed ideals, pinching op- erators and submanifolds of Banach manifolds. Section 3 is devoted to the study of the differential structure of UI (P ). For any symmetrically-normed ideal I different from the compact operators, we describe in Theorem 3.9 several equivalent conditions to UI(P ) be a submanifold of B(I). For the ideal K of compact operators many of these conditions are no longer equivalent. In fact, UK(P ) is always a quasi submanifold of B(K), which rarely has complemented tangent spaces in B(K) (see Theorem 3.16). In Section 4 we go further into the topological structure of UI (P ). We show that UI (P ) is a covering space of another natural orbit of P . The methods of this section use those of [4], where a similar situation arises in relation with the unitary orbit of a conditional expectation in von Neumann algebras. The Section 5 is concerned with the metric structure of UI (P ). Motivated by similar results on other homogeneous spaces [2, 12] we study the rectifiable distance induced by quotient Finsler metric on UI (P ). Under the assumption that the quotient topology on UI(P ) coincide with the inherited topology from B(I), we prove that the rectifiable distance defines these topologies. As a by-product we find that UI (P ) is complete with the rectifiable distance. In Section 6 we study the topology of UI-unitary orbits of a compact normal operator. These type of unitary orbits may be endowed with the quotient topology, though there is another quite natural topology, the one defined by the norm of the ideal I. We show that both topologies coincide if and only if the compact operator has finite rank. The proof makes use of the previous results on the topology of UI (P ). This result is related with several works [2, 3, 6, 9], where under different assumptions, the finite rank condition appears as sufficient to the statement on the topologies. 2 2 Preliminaries Symmetrically-normed ideals. We begin with some basics facts on symmetrically- normed ideals. For a deeper discussion of this subject we refer the reader to [17] or [22]. Let H be a Hilbert space. No confusion will arise if k . k denotes the norm of vectors in H and the uniform norm in B(H). For ξ, η ∈ H, let ξ ⊗ η be the rank one operator defined by (ξ ⊗ η)(ζ) = hζ, ηi ξ, for ζ ∈ H. By a symmetrically-normed ideal we mean a two-sided ideal I of B(H) endowed with a norm k . kI satisfying • (I, k . kI) is a Banach space. • kxyzkI ≤ kxkkykIkzk, for x, z ∈ B(H) and y ∈ I. • kξ ⊗ ηkI = kξk kηk, for ξ, η ∈ H. A result that goes back to J. Calkin ([11]) states the inclusions F ⊆ I ⊆ K, where F is the set of all the finite rank operators, I is a two-sided ideal of B(H) and K the ideal of compact operators on H. Symmetrically-normed ideals are closely related to the following class of norms. Let c be the real vector space consisting of all sequences with a finite number of nonzero terms. A symmetric norming function is a norm Φ : c → R satisfying the following properties: • Φ(1, 0, 0, . . .) = 1. • Φ(a1, a2, . . . , an, 0, 0, . . .) = Φ(aj1 , aj2 , . . . , ajn, 0, 0, . . .), where j1, . . . , jn is any per- mutation of the integers 1, 2, . . . , n and n ≥ 1. Any symmetric norming function Φ gives rise to two symmetrically-normed ideals. Indeed, for any compact operator x one may consider the sequence (sn(x))n of its singular values arranged in non-increasing order, and thus define kxkΦ := sup k≥1 Φ(s1(x), s2(x), . . . , sk(x), 0, 0, . . .) ∈ [0, ∞]. It turns out that and the k . kΦ-closure in SΦ of the finite rank operators, that is SΦ := { x ∈ K : kxkΦ < ∞ } S (0) Φ := F k . kΦ , (0) are symmetrically-normed ideals. It is not difficult to show that S Φ = SΦ if and only if SΦ is separable. Moreover, any separable symmetrically-normed ideal coincides with some S (0) Φ (see [17, p. 89]). Pinching operators. Let Φ a symmetric norming function, and I = SΦ. Recall that given a family { pi }w 1 (1 ≤ w ≤ ∞) of mutually orthogonal hermitian projections, i.e. we define the pinching operator associated with the family by pi = p∗ i , pipj = δij, P : I −→ I, P (x) = pixpi, w Xi=1 Notice that we might have w = ∞. Since x is compact, the series, which at first converges in the strong operator topology, turns out to be convergent in the uniform norm. It is also noteworthy that P is well defined in the sense that P (x) ∈ I whenever x ∈ I (see [17, p. 82]). 3 Bellow we need to consider the Banach algebra B(I) of all bounded operators on I with the usual operator norm: for X ∈ B(I), kXkB(I) = sup kykI=1 kX(y)kI. We collect some basic properties of pinching operators in the next proposition. Proposition 2.1. Let Φ a symmetric norming function, and I = SΦ. Let P be the pinching operator associated with a family { pi }w 1 . The following assertions hold: i) P 2 = P . ii) P is a module map over its range. iii) P (x)∗ = P (x∗). iv) P is continuous. In fact, kP kB(I) = 1. Proof. The proofs of i) − iii) are trivial. For a proof of iv) we refer the reader to [17, p. 82]. In the paper we will use different notions of submanifold of a (Banach) Submanifolds. manifold. Since the terminology is not uniform in the literature, we need to mention that we follow Bourbaki [10]. To be precise, let M be a manifold and N a topological space contained in M . Recall that a subspace F of a Banach space E is said to be complemented if F is closed and there exists a closed subspace F1 such that F ⊕ F1 = E. We will use the following definitions: • N is a submanifold of M if for each point x ∈ N there exists a Banach space E and a chart (W, φ) at x, φ : W ⊆ M −→ E, such that φ(W ∩ N ) is a neighborhood of 0 in a complemented subspace of E. • N is a quasi submanifold of M if for each point x ∈ N there exists a Banach space E and a chart (W, φ) at x, φ : W ⊆ M −→ E, such that φ(W ∩ N ) is a neighborhood of 0 in a closed subspace of E. The following criterion will be useful (see [10]). Proposition 2.2. Let M be a manifold, N be a topological space and N ⊆ M . Then N is a submanifold (resp. quasi submanifold) of M if and only if the topology of N coincides with the topology inherited from M and the differential map of the inclusion map N ֒→ M has complemented range (resp. closed range) at every x ∈ N . 3 Differential structure of UI(P ) Throughout this section, let Φ a symmetric norming function, and I = SΦ. Let P be the pinching operator associated with a family of mutually orthogonal projections { pi }w 1 (1 ≤ w ≤ ∞). We first show that UI(P ) has a smooth manifold structure endowed with the quotient topology. Lemma 3.1. Let x ∈ B(H). Then LxP = P Lx if and only if x =Pw Proof. Suppose that LxP = P Lx, which actually means that i=0 pixpi. (pix − xpi)ypi = 0, w Xi=1 4 for all y ∈ I. Let i ≥ 0 and (en)n be a sequence of finite rank projections such that en ≤ pi and en ր pi in the strong operator topology. We first assume that i ≥ 1. Replacing y by en, we get pixen = xen for all n ≥ 1. This gives pixpi = xpi for all i ≥ 1. Thus pjxpi = 0 for all i ≥ 1 and j ≥ 0. In the case in which i = 0 we replace y by p0x∗. Then we see that pixp0 = 0 for all i ≥ 1, and thus x must be block diagonal. The proof of the converse assertion is trivial. Proposition 3.2. Let Φ a symmetric norming function, and I = SΦ. Then UI(P ) is a real analytic homogeneous space of UI. Proof. Note that the isotropy group at P of the natural underlying action of UI is It is a closed subgroup of UI. Its Lie algebra can be identified with G = { u ∈ UI : LuP = P Lu }. G = { z ∈ Ish : LzP = P Lz }. kzkI < π. By the condition on the norm of z, we have z = log(u) =P∞ We will prove that G is a Banach-Lie subgroup of UI. Let u = ez ∈ G, with z ∈ Ish and (−1)n (n+1) (u − 1)n+1. Notice that LuP = P Lu, or Lu−1P = P Lu−1, clearly implies Lr(u−1)P = P Lr(u−1) for any polynomial r ∈ R[X], and by continuity we have LzP = P Lz. Denote by expUI : Ish −→ UI, expUI (z) = ez the exponential map of the Banach-Lie group UI. Hence we have proved that expUI (G ∩ V ) = G ∩ expUI (V ), for any sufficiently small neighborhood V of the origin in Ish. n=0 On the other hand, by Lemma 3.1 we can rewrite the Lie algebra as G =(cid:26) w Xi=0 pizpi : z ∈ Ish(cid:27), which is a real closed subspace of Ish. Moreover, the following subspace M = { z ∈ Ish : pizpi = 0, ∀ i ≥ 0 } =(cid:26) Xi6=j pizpj : z ∈ Ish(cid:27) is a closed supplement for G in Ish. Then, G is a Banach-Lie subgroup of UI, and by [24, Theorem 8.19] we conclude that UI(P ) is a real analytic homogeneous space of UI. 3.1 When is UI(P ) a submanifold of B(I)? The case I 6= K In this section, we discuss the submanifold structure of UI (P ) under the assumption that I 6= K. Recall that given the pinching operator P associated with a family of mutually 1 (1 ≤ w ≤ ∞), we may consider the larger family { pi }w orthogonal projections { pi }w 0 , i=1 pi. However, the pinching operator P is always associated with the 1 . The following estimate will be useful. where p0 = 1 −Pw first family { pi }w Lemma 3.3. Let Φ a symmetric norming function, and I = SΦ. Then kLxP − P LxkB(I) ≥ kpixpj k, for x ∈ I, i ≥ 1, j ≥ 0 and i 6= j. Proof. Consider the Schmidt expansion of the compact operator pixpj, namely sk ξk ⊗ ηk, pixpj = ∞ Xk=1 5 where sk are the singular values of pixpj and (ξk)k, (ηk)k are orthogonal systems of vectors. In particular, there is a vector ξ ∈ R(pj), kξk = 1, such that pixpjξ = kpixpjkξ. Pick any η ∈ R(pi) such that kηk = 1. Then note that (LxP − P Lx)(ξ ⊗ η) = −pix(ξ ⊗ η)pi = −pixpj(ξ ⊗ η) = −kpixpjk ξ ⊗ η. We thus get kLxP − P LxkB(I) ≥ k(LxP − P Lx)(ξ ⊗ η)kI = kpixpjk. The first obstruction for UI(P ) to be a submanifold of B(I) lies in the fact that its tangent spaces may not be closed. The tangent space of UI(P ) at Q (i.e. the derivatives at Q of smooth curves inside UI(P )) is apparently given by (T UI(P ))Q = { LzQ − QLz : z ∈ Ish }. We denote tangent vectors briefly by [Lz, Q]. Lemma 3.4. Assume that I 6= K. Then tangent spaces of UI (P ) are closed in B(I) if and only if w < ∞ and there is only one infinite rank projection in the family { pi }w 0 . Proof. It suffices to prove the statement for the tangent space at P . Indeed, if Q = LuP Lu∗ for some u ∈ UI, then [Lz, Q] = Lu[Lu∗zu, P ]Lu∗. Thus (T UI(P ))Q is closed in B(I) if and only if (T UI(P ))P is closed in B(I). Suppose that (T UI(P ))P is closed in B(I). Let x /∈ I be a compact operator and (en)n be a sequence of finite rank projections such that en ր 1 in the strong operator topology. Since x is compact, the sequence of finite rank operators zn = enxen satisfies kx − znk → 0. Let ℜe( . ) be the real part of an operator, then k [Lℜe(zn), P ] − [Lℜe(x), P ] kB(I) ≤ 2kLℜe(zn) − Lℜe(x)kB(I) = 2kℜe(zn) − ℜe(x)k ≤ 2kzn − xk → 0. Then there exists some z0 ∈ Ish such that [Lz0, P ] = [Liℜe(x), P ]. We can proceed analo- gously with the imaginary part ℑm( . ) to find another operator z1 ∈ Ish such that [Lz1, P ] = [Liℑm(x), P ]. Hence we obtain [Lx, P ] = [Lz, P ] for z = −iz0 + z1 ∈ I. By Lemma 3.1 the latter can be rephrased as x − z = pi(x − z)pi. w Xi=0 Xi=0 w In particular, we see that x − pixpi ∈ I. (1) Recall that I = SΦ for some symmetric norming function Φ. Since I is different from the compact operators, there exists a sequence of positive numbers (an)n such that an → 0 and Φ((an)n) = ∞. Suppose that the family { pi }w 0 has two projections pi, pj, i 6= j, such that both have infinite rank. Let (ξn)n be an orthonormal basis of R(pi) and (ηn)n be an orthonormal basis of R(pj). Consider the following compact operator: x = ∞ Xn=1 an ξn ⊗ ηn. 6 x −Pw From our choice of the sequence (an)n it follows that x /∈ I. Thus we find that x = pixpj = i=0 pixpi /∈ I, which contradicts equation (1). Hence it is impossible to have two different projections with infinite rank in the family { pi }w 0 . It remains to prove that w < ∞. Suppose that there is an infinite number of projections p1, p2, . . .. We can construct an orthonormal system of vectors (ξi)i such that ξi ∈ R(pi). Then we define the following compact operator: x = ∞ Xn=1 an ξn+1 ⊗ ξn. It is easily seen that x = P∞ contradiction with equation (1). n=1 pn+1xpn = x −P∞ i=0 pixpi /∈ I. We thus get again a In order to prove the converse we assume that the family { pi }w 0 satisfies w < ∞ and it has only one projection pi0 with infinite rank. Let (zk)k be a sequence in Ish such that k [Lzk, P ] −XkB(I) → 0, where X ∈ B(I). It is worth noting that by Lemma 3.1 the sequence (zk)k can be chosen satisfying pizkpi = 0 for all k and i = 0, . . . , w. Since ( [Lzk , P ] )k is a Cauchy sequence in B(I), Lemma 3.3 implies that kpi(zk − zr)pjk −→ k,r→∞ 0 for i = 1, . . . , w, j = 0, . . . , w and i 6= j. Note that the rank of the operators pi(zk − zr)pj is uniformly bounded on the subscripts k and r by C := max{ rank(pj) : j = 0, . . . , w, j 6= i0 }. Then we get kpj(zk − zr)pikI ≤ Ckpj(zr − zk)pik −→ k,r→∞ 0. Hence each (pjzkpi)k converges in the ideal norm to some zij ∈ I. We can construct an operator z by defining its matricial blocks with respect to the projections p0, p1, . . . , pw as follows: pizpj :=(cid:26) 0 zij if if i = j, i 6= j. Then z is a skew-hermitian operator in I satisfying kz − zkkI ≤Xi6=j kpjzpi − pjzkpikI =Xi6=j kzij − pjzkpikI → 0. Therefore k [Lzk, P ] − [Lz, P ] kB(I) ≤ 2kLzk − LzkB(I) = 2kzk − zk ≤ 2kzk − zkI → 0. Hence we conclude X = [Lz, P ], and the lemma is proved. We can endow UI(P ) with two natural topologies. According to Proposition 3.2 we have that UI(P ) ≃ UI/G has a real analytic manifold structure in the quotient topology in such way that the map π : UI −→ UI(P ), π(u) = LuP Lu∗ is a real analytic submersion. On the other hand, we can regard UI(P ) as a subset of B(I) with the inherited topology. In this case, we denote the projection map by π : UI −→ UI(P ), π(u) = LuP Lu∗. Note that π is also continuous, and the following diagram commutes π UI UI(P ) ≃ UI/G A A A A A A π A A id UI(P ) ⊆ B(I) 7 / /   Here id stands for the identity map. Note that id is always continuous, but it may not be a homeomorphism. In fact, we will show that the two topologies defined on UI(P ) coincide if and only if tangent spaces are closed. As we will see, the proof of this result depends on the existence of continuous local cross sections for the action. Remark 3.5. Let P be the pinching operator associated with a family { pi}w consider the unitary orbit of each projection pi, i.e. 1 . We will Oi := { upiu∗ : u ∈ UI }. If I is the ideal of Hilbert-Schmidt operators, the above defined orbits are usually known as the connected component of pi in the restricted Grassmannian (see e.g. [21]). Note that Oi ⊆ pi + I, so we may endow each orbit with the subspace topology defined by the metric (upiu∗, vpiv∗) 7→ kupiu∗ − vpiv∗kI. Lemma 3.6. Assume that w < ∞ and there is only one infinite rank projection in the family { pi}w 0 . Then the map Fi : UI(P ) −→ Oi, Fi(LuP Lu∗ ) = upiu∗ is continuous for i = 0, 1, . . . w, when UI(P ) is endowed with the topology inherited from B(I). Proof. We first show that the function Fi is well defined for i = 0, 1, . . . , w. From Lemma i=0 piv∗upi. Then we get v∗upi = 3.1 we know that LuP Lu∗ = LvP Lv∗ implies v∗u = Pw piv∗upi = piv∗u, or equivalently, upiu∗ = vpiv∗. To prove the continuity of Fi we will actually see that Fi is Lipschitz. Since the under- lying actions are isometric, it suffices to estimate the distance from Fi(LuP Lu∗) = upiu∗ to Fi(P ) = pi. For u ∈ UI, set a(u) := kLuP Lu∗ − P kB(I) = k [Lu, P ] kB(I). From Lemma 3.3 it follows that kpiupjk = kpi(u − 1)pjk ≤ a(u), for j = 0, 1, . . . , w, i = 1, . . . , w and i 6= j. The same estimate can be extended for all i 6= j. In fact, we have kpjupik = kpiu∗pjk ≤ a(u∗) = a(u). Let pi0 the unique infinite rank projection in the family { pi }w rank(piupj) ≤ min{ rank(pi) , rank(pj) }, and then we get 0 . For u ∈ UI, we note that max{ rank(pjupi) : i, j = 0, 1, . . . , w, i 6= j } ≤ max{ rank(pj) : j = 0, 1, . . . , w, j 6= i0 } := C. This implies that for i 6= j. Thus we get kpiupjkI ≤ Ckpiupjk kFi(LuP Lu∗) − Fi(P )kI = kupi − piukI ≤ Xj:j6=i kpjupik + Xj:j6=i ≤ C(cid:18) Xj:j6=i kpjupikI + Xj:j6=i kpiupjk(cid:19) ≤ 2wCkLuP Lu∗ − P kB(I), kpiupjkI (2) which shows that F is Lipschitz. 8 Remark 3.7. Let M be the supplement of the Lie algebra defined in Proposition 3.2. Suppose that w = ∞ or there exist two different infinite rank projections in the family { pi }w 0 . Under the assumption that I 6= K, we will construct a sequence (zk)k in M satisfying kzkk → 0 and kzkkI = 1. To this end, put ak := Φ(1, 1, . . . , 1 , 0, 0, . . .), k {z } where Φ is a symmetric norming function such that I = SΦ. Since I 6= K, it follows that Φ is not equivalent to the uniform norm of ℓ∞, so that ak → ∞ (see [17, p. 76]). In the case in which w = ∞, let (ξi)i be an orthonormal system such that ξi ∈ R(pi) for all i ≥ 1. It is not difficult to see that the sequence defined by zk := a−1 2k k Xi=1 ξ2i−1 ⊗ ξ2i − ξ2i ⊗ ξ2i−1 satisfies the required properties. In the case in which there exist two different infinite rank projections pi and pj, let (ξi)i be an orthonormal system such that ξ2k−1 ∈ R(pi) and ξ2k ∈ R(pj) for all k ≥ 1. Then we can define the sequence (zk)k in the same fashion as before. Lemma 3.8. Assume that I 6= K. Then the following conditions are equivalent: i) The quotient topology of UI(P ) coincides with the topology inherited from B(I). ii) w < ∞ and there is only one infinite rank projection in the family { pi}w 0 . Proof. Suppose that the quotient topology of UI(P ) ≃ UI/G coincides with the topology inherited form B(I). Let M be the supplement of the Lie algebra of G defined in Proposition 3.2. Recall that a real analytic atlas of UI(P ) compatible with the quotient topology can be constructed by translation of the homeomorphism ψ : W ⊆ M −→ ψ(W), ψ(z) = (π ◦ expUI )(z) = Lez P Le−z , where W is an open neighborhood of 0 ∈ M and ψ(W) an open neighborhood of P (see for instance [5, Theorem 4.19]). Assume that the family { pi }w 0 does not satisfy the claimed properties. This leads us to consider two cases, namely w = ∞ or there exist two different infinite rank projections in { pi }w 0 . In any case we can find a sequence (zk)k in M such that kzkk → 0 and kzkkI = 1 according to Remark 3.7. Then note that kLez k P Le−z k − P kB(I) = k [Lez k −1, P ] kB(I) ≤ 2kezk − 1k → 0, and using that the quotient topology of UI(P ) coincides with the subspace topology, we arrive at a contradiction: kzkkI = kψ−1(Lez k )kI → 0. k P Le−z To prove the converse, assume that w < ∞ and there is only one infinite rank projection 0 . Clearly, our assertion about the topology of UI(P ) will follow if we in the family { pi}w show that the projection map π : UI −→ UI(P ), π(u) = LuP Lu∗ have continuous local cross sections, when UI(P ) is considered with the relative topology of B(I). To this end, for i = 0, 1, . . . , w, we need to consider the orbits Oi := { upiu∗ : u ∈ UI }. In [1, Proposition 2.2] the authors showed that the maps πi : UI −→ Oi, πi(u) = upiu∗, 9 has continuous local cross sections, when I is the ideal of Hilbert-Schmidt operators. Actu- ally, the same proof works out for any symmetrically-normed ideal I, so we have that there exist continuous maps ψi : { q ∈ Oi : kq − pikI < 1 } ⊆ pi + I −→ UI such that ψi(upiu∗)piψi(upiu∗)∗ = upiu∗ for any u ∈ UI such that kupiu∗ − pikI < 1. Now we can explicitly give the required section for π, namely σ :(cid:8) Q ∈ UI(P ) : kQ − P kB(I) < 1/2wC(cid:9) −→ UI, σ(LuP Lu∗) = ψi(upiu∗)pi. w Xi=0 If Q = LuP Lu∗ lies in the domain of σ, then by the estimate (2) in Lemma 3.6, the operators upiu∗ do lie in the domain of each ψi. Our next task is to show that σ = σ(LuP Lu∗ ) ∈ UI. In fact, we see that σσ∗ =(cid:18) w Xi=0 ψi(upiu∗)pi(cid:19)(cid:18) w Xi=0 piψi(upiu∗)∗(cid:19) = w Xi=0 ψi(upiu∗)piψ(upiu∗)∗ = upiu∗ = 1. w Xi=0 Note that pjψj(upju∗)∗ψi(upiu∗)pi = ψj(upju∗)∗upjpiu∗ψi(upiu∗) = δij, then σ∗σ =(cid:18) w Xi=0 piψi(upiu∗)∗(cid:19)(cid:18) w Xi=0 ψi(upiu∗)pi(cid:19) = w Xi=0 pi = 1. Also we see that σ − 1 = (ψi(upiu∗) − 1)pi ∈ I. w Xi=0 On the other hand, the map σ is actually a section for π: for any y ∈ I, Lσ(LuP Lu∗ )P Lσ(LuP Lu∗ )∗ (y) = = = σ(LuP Lu∗ )piσ(LuP Lu∗ )∗ypi ψi(upiu∗)piψi(upiu∗)∗ypi upiu∗ypi = LuP Lu∗ (y). w w Xi=0 Xi=0 Xi=0 w Finally, to show the continuity of σ, it is enough to remark that σ(LuP Lu∗) = w Xi=0 ψi(Fi(LuP Lu∗ ))pi and use the continuity of each Fi, which has already been proved in Lemma 3.6. Now our main result on the differential structure of UI(P ) follows. Theorem 3.9. Let Φ a symmetric norming function, and I = SΦ. Assume that I 6= K. Let P be the pinching operator associated with a family { pi}w 1 (1 ≤ w ≤ ∞). Then the following assertions are equivalent: i) The quotient topology on UI(P ) coincides with topology inherited from B(I). ii) Tangent spaces of UI(P ) are closed in B(I). 10 iii) w < ∞ and there is only one infinite rank projection in the family { pi }w 0 . iv) UI(P ) is a submanifold of B(I). Proof. Suppose that UI(P ) is a submanifold of B(I). By Proposition 2.2, tangent spaces of UI (P ) has to be closed in B(I). From Lemma 3.4 it follows that the family { pi }w 0 satisfies the stated properties. Now we assume that w < ∞ and there is only one infinite rank projection in the family { pi }w 0 . According to Lemma 3.4 and Lemma 3.8, what is left to prove is that tangent spaces are complemented in B(I). Clearly, it suffices to show that (T UI (P ))P is complemented in B(I). We will divide the proof into two cases according to whether the rank of p0 is infinite or finite. Let us first assume that rank(p0) = ∞, so that rank(pi) < ∞ for all i = 1, . . . , w. Then X(pi) is well defined for any X ∈ B(I), i = 1, . . . , w, and we can set z : B(I) −→ Ish, z(X) = 2iℑm(cid:18) w Xi=1 i−1 Xj=0 pjX(pi)(cid:19). Clearly z is a continuous linear operator. Then we define a bounded linear projection onto the tangent space by E : B(I) −→ (T UI(P ))P , E(X) = [Lz(X), P ]. In order to show that E actually defines a projection we pick X = [Lz, P ] for some z ∈ Ish. Notice that X(pi) = (1 − pi)zpi, for all i = 1, . . . , w, then we get that z(X) = 2iℑm(cid:18) w Xi=1 i−1 Xj=0 pjzpi(cid:19) = z − n Xi=0 pizpi. From Lemma 3.1 we deduce that E(X) = [Lz(X), P ] = X, which proves that E is a projec- tion. Finally, the continuity of z easily implies that of E. Now we consider the case in which the infinite rank projection is not p0. Without loss of generality we may assume that rank(p1) = ∞. Let us point out that the above definition of the operator z(X) does not work in this case for two different reasons: on one hand, since p1 /∈ I we cannot evaluate any X ∈ B(I) at p1, and on the other hand, every tangent vector [Lz, P ] vanishes at p0. In order to solve this case we need to modify the definition of the operator z. Recall that rank(p0) < ∞ since rank(p1) = ∞. Let η1, . . . , ηm be an orthonormal basis of R(p0). Let ξ ∈ R(p1) be a unit vector. Then we define z : B(I) −→ Ish, z(X) = 2iℑm(cid:18) w Xi=2 i−1 Xj=0 pjX(pi) − m Xk=1 X(ηk ⊗ ξ)ξ ⊗ ηk(cid:19), and the projection onto the tangent space is E : B(I) −→ (T UI(P ))P , E(X) = [Lz(X), P ]. It is apparent that E is continuous, so we are left with the task of proving that E is a projection. To this end, let X = [Lz, P ] for some z ∈ Ish. Note that X(ηk ⊗ ξ) = w Xi=1 (zpi − piz)(ηk ⊗ ξ)pi = (zp1 − p1z)(ηk ⊗ ξ)p1 = −p1z(ηk ⊗ ξ), 11 and then Thus we get m Xk=1 z(X) = 2iℑm(cid:18) w Xi=2 X(ηk ⊗ ξ)ξ ⊗ ηk = −p1zp0. i−1 Xj=0 pjzpi + p1zp0(cid:19) = z − n Xi=0 pizpi. Hence we conclude that E([Lz, P ]) = [Lz, P ], and the proof is complete. 3.2 When is UK(P ) a submanifold of B(K)? In this section we turn to the case I = K. The following estimate is a somewhat improved version of Lemma 3.3. Lemma 3.10. Let x ∈ K such that pixpi = 0 for all i ≥ 1. Then kLxP − P LxkB(K) ≥ kx(1 − p0)k, where p0 = 1 −Pw i=1 pi. Proof. To estimate the norm of LxP − P Lx as an operator acting on K we need to consider the following projections: if rank(pi) = ∞, let (pi,k)k be a sequence of finite rank projections satisfying pi,k ≤ pi and pi,k ր pi, and if rank(pi) < ∞, we set pi,k = pi for all k ≥ 1. Now assume that the pinching operator P is associated with a family { pi }w 1 such that w < ∞. i=1 pi,k have finite rank. We thus get Then the projections given by ek =Pn kLxP − P LxkB(K) ≥ k(LxP − P Lx)(ek)k =(cid:13)(cid:13)(cid:13)(cid:13) w Xi=1 (1 − pi)xpi,k(cid:13)(cid:13)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:13)(cid:13) x w Xi=1 where in the last equality we use that pixpi = 0. Using that x ∈ K and pi,k ր pi, we find that , pi,k(cid:13)(cid:13)(cid:13)(cid:13) kLxP − P LxkB(K) ≥ kx(1 − p0)k. In the case where w = ∞, we set en,k = Pn that i=1 pi,k. In the same fashion as above we find Letting k → ∞, we have x kLxP − P LxkB(K) ≥(cid:13)(cid:13)(cid:13)(cid:13) kLxP − P LxkB(K) ≥(cid:13)(cid:13)(cid:13)(cid:13) x n pi,k(cid:13)(cid:13)(cid:13)(cid:13) Xi=1 pi(cid:13)(cid:13)(cid:13)(cid:13) Xi=1 n , for all n ≥ 1. Now letting n → ∞, we get the estimate in this case. Proposition 3.11. Tangent spaces of UK(P ) are closed in B(K). Proof. By the remark at the beginning of the proof of Lemma 3.4, we may restrict, without loss of generality, to verify the statement for the tangent space at P . Let (zk)k be a sequence in Ksh such that pizkpi = 0 for all i ≥ 0 and k ≥ 1. Suppose that k [Lzk, P ] − X kB(K) → 0 for some X ∈ B(K). According to Lemma 3.10, k(zk − zr)(1 − p0)k ≤ k [Lzk−zr , P ] kB(K). Also note that k(zk − zr)p0k = kp0(zk − zr)k = kp0(zk − zr)(1 − p0)k ≤ k [Lzk−zr , P ] kB(K). 12 Therefore (zk)k is a Cauchy sequence and thus has a limit z0 ∈ Ksh. Then we see that k [Lzk, P ] − [Lz0, P ] k ≤ 2kzk − z0k → 0. Thus we conclude that X = [Lz0, P ]. Now we turn to the study of the topology of UK(P ). We will find that the quotient topol- ogy and the topology inherited from B(K) coincide regardless the number or rank of the projections in the family { pi }w 0 . Remark 3.12. Let P be the pinching operator associated with a family { pi }w 1 (1 ≤ i ≤ w). In this subsection we need to consider again the unitary orbit of the projections, which we denote by Oi = { upiu∗ : u ∈ UK }. for i = 0, . . . , w. We claim that the map F0 : UI(P ) −→ Oi, F0(LuP Lu∗) = up0u∗. i=0 piupi we have that is Lipschitz. In fact, according to Lemma 3.10 applied with x = u − 1 −Pw u −Pw kp0u(1−p0)k =(cid:13)(cid:13)(cid:13)(cid:13) piupi(cid:19)(1−p0)(cid:13)(cid:13)(cid:13)(cid:13) piupi(cid:19)(1−p0)(cid:13)(cid:13)(cid:13)(cid:13) ≤(cid:13)(cid:13)(cid:13)(cid:13) (cid:18)u− Replacing u by u∗ we find that w Xi=0 p0(cid:18)u− w Xi=0 i=0 pi(u − 1)pi = ≤ kLuP Lu∗−P kB(K) . k(1 − p0)up0k = kp0u∗(1 − p0)k ≤ kLuP Lu∗ − P kB(K). Thus we get kF0(LuP Lu∗ ) − F0(P )k = kup0u∗ − p0k ≤ k(1 − p0)up0k + kp0u(1 − p0)k ≤ 2kLuP Lu∗ − P k, which proves our claim. Lemma 3.13. Let u, v ∈ UK. Then (cid:13)(cid:13)(cid:13)(cid:13) w Xi=0 upiu∗pi − vpiv∗pi(cid:13)(cid:13)(cid:13)(cid:13) ≤ 3kLuP Lu∗ − LvP Lv∗kB(K), where in the case in which w = ∞ the series on the left side is convergent in the uniform norm. Proof. For each i ≥ 1, let (pi,k)k be a sequence of finite rank projections such that pi,k ≤ pi and pi,k ր pi. In case pi has finite rank, we set pi,k = pi for all k. We will use the orthogonal i=1 pi,k. Put a(u, v) := kLuP Lu∗ − LvP Lv∗kB(K). Then projections defined by ek =Pn w = k(LuP Lu∗ − LvP Lv∗ )(ek)k ≤ a(u, v). Note that for each i ≥ 1, the operator upiu∗ − vpiv∗ is compact. Letting k → ∞, we get that Xi=1 (cid:13)(cid:13)(cid:13)(cid:13) (upiu∗ − vpiv∗)pi,k(cid:13)(cid:13)(cid:13)(cid:13) Xi=1 w (cid:13)(cid:13)(cid:13)(cid:13) (upiu∗ − vpiv∗)pi(cid:13)(cid:13)(cid:13)(cid:13) 13 ≤ a(u, v). Combining this with the Remark 3.12 it gives that w Xi=0 (cid:13)(cid:13)(cid:13)(cid:13) (upiu∗ − vpiv∗)pi(cid:13)(cid:13)(cid:13)(cid:13) Xi=0 ∞ ∞ Xi=0 ≤ 3a(u, v). (3) This finishes the proof for the case w < ∞. If w = ∞, we note that upiu∗pi − vpiv∗pi = upi(u∗ − 1)pi − vpi(v∗ − 1)pi + (u − v)pi. Since the operators u∗ − 1, v∗ − 1 and u − v are compact, this series converges in the uniform norm. Letting w → ∞ in (3), the desired inequality follows. In the following proposition we extend the technique developed in [1] to construct continuous local cross sections. Proposition 3.14. The map π : UK −→ UK(P ) ⊆ B(K), π(u) = LuP Lu∗, has continuous local cross sections, when UK(P ) is considered with the topology inherited from B(K). Proof. Let P be the pinching operator associated with a family { pi }w 1 (1 ≤ w ≤ ∞). Since the action of UK is isometric it will be enough to find a continuous section σ in a neighborhood of P . Also we will restrict ourselves to prove the case w = ∞. The case w < ∞ needs less care, and it can be handled in much the same fashion. We consider the following neighborhood of P to define the cross section, V :=(cid:8) Q ∈ UK(P ) : kQ − P kB(K) < 1/3(cid:9). Given Q = LuP Lu∗ ∈ V, where u ∈ UK, let qi = Fi(Q) = upiu∗ for i ≥ 0. According to the proof of Lemma 3.6 the function Fi is well defined. Then, we set s = s(Q) := qipi . ∞ Xi=0 This series is convergent in the strong operator topology. In fact, we can rewrite the series as qipi = upi(u∗ − 1)pi + (u − 1)pi + pi, ∞ Xi=0 ∞ Xi=0 where the first and second summand on the right are convergent in the uniform norm, while the third is convergent in the strong operator topology. On the other hand, note that Then we get that s is invertible. Moreover, it follows that ks − 1k ≤ 3kQ − P kB(K) < 1. s − 1 = u(cid:18) ∞ Xi=0 pi(u∗ − 1)pi + 1(cid:19) − 1 = u ∞ Xi=0 pi(u∗ − 1)pi + u − 1 ∈ K, i=0 pi(u∗ − 1)pi ∈ K. Now we will show that which is due to the fact that P∞ σ = σ(Q) := ss−1 14 is a continuous local cross section for π. To this end, note that spi = qipi = qis, so that pis2 = s∗qis = s2pi, which implies σpiσ∗ = ss−1pis−1s = spis−2s∗ = spis−1 = qi. This allows us to prove that σ is a section: for any y ∈ K, we have LσP Lσ∗ (y) = σpiσ∗ypi = ∞ Xi=1 ∞ Xi=1 qiypi = Q(y). On the other hand, we have s2 − 1 ∈ K, and consequently, s − 1 = (s2 − 1)(s + 1)−1 ∈ K. Therefore we can conclude σ − 1 = ss−1 − 1 = (s − s)s−1 = (s − 1)s−1 + (1 − s)s−1 ∈ K. Hence σ ∈ UK. Let Gl(H) denote the group of invertible operators on H. In order to prove the continuity of σ we consider the subgroup of Gl(H) given by GlK = { g ∈ Gl(H) : g − 1 ∈ K }. It is a Banach-Lie group endowed with the topology defined by (g1, g2) 7→ kg1 − g2k (see [5]). From Lemma 3.13 the map s : V −→ GlK is continuous. Also note that the map GlK −→ UK, s 7→ ss−1, is real analytic by the regularity properties of the Riesz functional calculus. Thus σ is continuous, being the composition of continuous maps. Our next task in the study of the submanifold structure of UK(P ) is to ask about the existence of a supplement for (T UI(P ))P in B(K). The existence of such supplement is closely related to the fact that for an infinite dimensional Hilbert space H the compact operators are not complemented in B(H). A proof of this result can be found, for instance, in [13]. It is based on the following well known result: c0 (sequences which converges to zero) is not complemented in ℓ∞ (bounded sequences). The reader can find a proof of this latter fact in [26]. Remark 3.15. We will need a slightly modified version of the afore-mentioned result. We first note that Ksh is not complemented in B(H)sh. Otherwise we would have a real bounded projection E : B(H)sh −→ Ksh, then we can define a bounded projection E : B(H) −→ K, E(x) = −iE(iℜe(x)) + iE(iℑm(x)), a contradiction. Let q1, q2 two infinite rank orthogonal projections on H. We claim that q1Kshq2 is not complemented in q1B(H)shq2. In fact, suppose that there exists a real bounded projection E : q1B(H)shq2 −→ q1Kshq2. Let v a partial isometry on H such that v∗v = q1 and vv∗ = q2. Then we have that LvELv∗ : B(q2(H))sh −→ q2Kshq2 is a bounded projection, which is impossible by the previous paragraph. In the following result we collect the above proved properties of UK(P ) and we give a complete characterization of the submanifold structure. Theorem 3.16. Let P be the pinching operator associated with a family {pi}w 1 (1 ≤ w ≤ ∞). Then UK(P ) is a quasi submanifold of B(K). Furthermore, UK(P ) is a submanifold of B(K) if and only if w < ∞ and there is only one infinite rank projection in the family { pi }w 0 . Proof. The first statement about the quasi submanifold structure of UK(P ) has already been proved in Proposition 3.11 and Proposition 3.14. Assume that w < ∞ and there is only one infinite rank projection in the family { pi }w 0 . The same proof of Theorem 3.9 can be carried out to show that (T UK(P ))P is complemented in B(K). Suppose now that UK(P ) is a submanifold of B(K). According to Proposition 2.2, there is a bounded linear projection E : B(K) −→ (T UK(P ))P . Two cases should be considered: 15 first, that there are two infinite rank projections in the family { pi }w w = ∞. q2 ∈ { p1, . . . , pw } \ { q1 } be other infinite rank projection. 0 , and second, that In the first case, let q1 ∈ { p0, p1, . . . , pw } be an infinite rank projection and In the second case, we set k=0 p2k+1. In any case we define the following bounded linear q1 = P∞ k=0 p2k and q2 = P∞ map E : q1B(H)shq2 −→ q1Kshq2, E(q1xq2) = (Lq1E)( [Lq1xq2+q2xq1, P ] )(q2). We claim that E is a projection onto q1Kshq2. In fact, notice that for each x ∈ B(H)sh there is z ∈ Ksh such that E( [Lq1xq2+q2xq1, P ] ) = [Lz, P ]. In the case in which there are two infinite rank projections, note that E(q1xq2) = q1 w Xi=1 (zpi − piz)q2pi = q1(zq2 − q2z)q2 = q1zq2. On the other hand, when w = ∞, E(q1xq2) = q1 ∞ Xi=1 (zpi − piz)q2pi = q1 ∞ Xk=0 (zp2k+1 − p2k+1z)p2k+1 = q1z p2k+1 = q1zq2. ∞ Xk=0 This proves that the range of E is contained in p1Kshp2. Moreover, let x ∈ Ksh, then we have that E( [Lq1xq2+q2xq1 , P ] ) = [Lq1xq2+q2xq1 , P ]. We thus get that E(q1xq2) = q1(q1xq2 + q2xq1)q2 = q1xq2. Hence E is a continuous linear projection onto q1Kshq2. In other words, q1Kshq2 is comple- mented in q1B(H)shq2, but this contradicts Remark 3.15. 4 Covering map For u ∈ UI, consider the inner automorphism given by Adu : I −→ I, Adu(x) = uxu∗. Given a pinching operator P associated with a family { pi }w 1 , there is another orbit of P defined by OI(P ) := { AduP Adu∗ : u ∈ UI }. Note that all the operators in OI(P ) are pinching operators while P is the only pinching operator in UI (P ). The isotropy group of the the co-adjoint action is given by H = { u ∈ UI : AduP Adu∗ = P }. (4) In order to find a characterization of the operators in H we need the following lemma. We make the convention { 0, 1, . . . , ∞} = N0. Lemma 4.1. Let P be the pinching operator associated with a family { pi }w pinching operator associated with another family { qi }v and pi = qσ(i) for some permutation σ of { 0, . . . , w} such that σ(0) = 0. 1 and Q be the 1. Then P = Q if and only if w = v Proof. We first suppose that P = Q. This is equivalent to pixpi = w Xi=1 qjxqj, v Xj=1 (5) for all x ∈ I. j=1 qjpiqj = pi. Then it follows that qjpi = qjpiqj = piqj for all j ≥ 1. If rank(pi) = ∞, we use the same idea with a sequence of projections (en)n such that en ≤ pi, en ր pi, to find that qjen = enqj, which If rank(pi) < ∞, i ≥ 1, we set x = pi to get Pv 16 qjpi = piqj for all i, j ≥ 0. i=1 qj, we can conclude that i=1 pi and q0 = 1 −Pv implies that qjpi = piqj . Since p0 = 1 −Pw end, let ξ ∈ R(pi), ξ 6= 0, and note that piξ = ξ =Pv Now we claim that for each i ≥ 0, we can find a unique σ(i) such that pi = qσ(i). To this j=0 qjξ. This implies that there is some j := σ(i) such that qjξ 6= 0. Then we see that qj ξ = qjpiξ = piqj ξ. Now let η ∈ R(pi) and insert x = η ⊗ qjξ in equation (5). In case i > 0 we find that η ⊗ qjξ = (qjη) ⊗ qj ξ. If j = 0, then η ⊗ qjξ = 0. In particular, if we take η = qjξ 6= 0, we obtain a contradiction. Hence we must have j > 0, so the equation η ⊗ qjξ = (qj η) ⊗ qjξ implies that qjη = η. Since η is arbitrary, we have R(pi) ⊆ R(qj). In a similar way, we may choose η ∈ R(pj) to obtain that R(qj) ⊆ R(pi). Thus pi = qj. In case i = 0, we need to show that p0 = q0. Suppose that there exists some j > 0 such that qjξ 6= 0. By the preceding paragraph we know that qjξ ∈ R(p0). Then we insert x = (qj ξ)⊗qjξ in equation (5) to find that 0 = (qjξ)⊗qj ξ, and hence qjξ = 0, a contradiction. j=0 qjξ = q0ξ, and consequently, R(p0) ⊆ R(q0). Interchanging 0 is a mutually orthogonal family, σ(i) Thus we obtain that ξ =Pv p0 and q0, we can conclude that p0 = q0. Since { qj }v is unique and our claim is proved. In other words, we have proved the existence of a map σ : { 0, . . . , w} → { 0, . . . , v} satisfying pi = qσ(i) and σ(0) = 0. Repeating the previous argument with qj in place of pi, we can construct another map ψ : { 0, . . . , v} → { 0, . . . , w} such that qj = pψ(j) and ψ(0) = 0. But pi = qσ(i) = p(ψσ)(i) and qj = pψ(j) = q(σψ)(j), so we have that σψ = ψσ = 1. Hence, σ is a permutation and w = v. In order to prove the converse, let σ a permutation of { 0, . . . , w }, P be the pinching 1 and Q be the pinching operator associated with i=0 pi. For each operator associated with a family { pi }w { pσ(i) }w x ∈ I, since x is compact, we find that k(1 − ek)xk → 0. Note that for k ≥ 1, 1 . Since the case w < ∞ is trivial, we suppose w = ∞. Set ek =Pk ∞ Xi=1 pixpi(cid:13)(cid:13)(cid:13)(cid:13) pσ(i)ekxpσ(i) = pixpi = k Xi=1 ∞ pσ(i)(1−ek)xpσ(i)− ∞ Xi=1 =(cid:13)(cid:13)(cid:13)(cid:13) piekxpi. Xi=1 pi(1−ek)xpi(cid:13)(cid:13)(cid:13)(cid:13) Xi=1 ∞ ≤ 2k(1−ek)xk → 0, Then we get ∞ Xi=1 pσ(i)xpσ(i)− ∞ Xi=1 (cid:13)(cid:13)(cid:13)(cid:13) which proves that P = Q. Let P be the pinching operator associated with a family { pi }w 1 (1 ≤ w ≤ ∞). Let F be the set of all the permutations σ of { 0, . . . , w } such that σ(i) = i for all but finitely many i ≥ 0. Note that the definition of the set F becomes unnecessary if w < ∞. We will need to consider permutations of a finite number of finite dimensional blocks with the same dimension such that fix zero, i.e. F := { σ ∈ F : σ(0) = 0, rank(pi) = rank(pσ(i)) < ∞ if σ(i) 6= i }. Let (ξi,j(i)) be an orthonormal basis of H such that (ξi,j(i))j(i)=1,...,rank(pi) is a basis of R(pi), where i = 0, . . . , w. For each σ ∈ F , we define the following permutation block operator matrix: rσ(ξi,j(i)) := ξσ(i),j(σ(i)), i = 0, . . . , w, j(i) = 1, . . . , rank(pi). Note that rank(rσ − 1) < ∞, since σ ∈ F . Hence, symmetrically-normed ideal I. it follows that rσ ∈ UI for any 17 Example 4.2. A simple example takes place when H = Cn, rank(pi) = 1 and Pn i=1 pi = 1. Here the set of all the matrices of the form rσ, σ ∈ F , reduces to all the n × n permuta- tion matrices. According to our next result, H has exactly n! connected components in this example. Recall that from the proof of Proposition 3.2 we know that the isotropy group G at P corresponding to the action given by the left representation can be characterized as block diagonal unitary operators, i.e. G =(cid:26) u ∈ UI : w Xi=0 piupi = u(cid:27), where P is the pinching operator associated with a family { pi }w 1 . Lemma 4.3. Let H be the isotropy group defined in (4). Then, H = [σ∈F rσG, where each set in the union is a connected component of H. Proof. Let u ∈ UI such that AduP Adu∗ = P . According to Lemma 4.1 it follows that upiu∗ = pσ(i) for some σ permutation of { 0, . . . , w } such that σ(0) = 0. In particular, note that pjupi = δj,σ(i) pσ(i)u, which actually says that u has only one nonzero block in each row. Since u − 1 ∈ I, we get that σ ∈ F . Hence we can write u = rσrσ−1 u, where rσ−1 u ∈ G. To prove the other inclusion it suffices to note that rσupiu∗rσ−1 = rσpirσ−1 = pσ(i) for any u ∈ G. Then we apply again Lemma 4.1 to obtain that AduP Adu∗ = P . In order to establish the last assertion about the connected components of H, we remark that krσu − rσ′ vkI ≥ krσu − rσ′ vk ≥ 1, whenever σ 6= σ′ and u, v ∈ G. This implies that the distance between any pair of sets that appear in the union is greater than one. On the other hand, it is a well known fact that UI is connected, then so does rσG. Hence the lemma is proved. Remark 4.4. As a consequence of Lemma 4.3, H is a Banach-Lie subgroup of UI. Indeed, the connected components of H are diffeomorphic to the Banach-Lie subgroup G of UI. Hence it follows that OI(P ) ≃ UI/H has a manifold structure endowed with the quotient topology. Theorem 4.5. Let Φ a symmetric norming function, and I = SΦ. Let P be the pinching operator associated with a family {pi}w 1 . If I 6= K assume in addition that w < ∞ and there is only one infinite rank projection in the family { pi }w 0 . Then the map Π : UI(P ) −→ OI(P ), Π(LuP Lu∗) = AduP Adu∗ , is a covering map, when UI(P ) is considered with the topology inherited from B(I) and OI(P ) with the quotient topology. Proof. In the case where I 6= K, under the above hypothesis on the family { pi }w 1 , it was proved in Lemma 3.8 that the quotient topology coincides with the subspace topology on UI(P ). In case I = K both topologies coincide without additional hypothesis by Proposition 3.14. On the other hand, by Lemma 4.3 the quotient H/G is discrete, then H/G is homo- morphic to F . We define an action of F on UI(P ) given by σ · LuP Lu∗ = Lurσ P Lrσ−1 u∗ . Therefore we can make the following identifications: UI(P )/F ≃ UI(P )/(H/G) ≃ (UI/G)/(H/G) ≃ UI/H ≃ OI(P ). 18 Thus we may think of Π as the quotient map UI(P ) −→ UI(P )/F . Hence to prove that Π is a covering map, it suffices to show that F acts properly discontinuous on UI(P ) (see [18]). This means that for any Q ∈ UI(P ), there is an open neighborhood W of Q such that W ∩ σ · W = ∅ for all σ 6= 1. Clearly, there is no loss of generality if we prove this fact for Q = P . To this end, define the open neighborhood by W := { Q ∈ UI(P ) : kQ − P kB(I) < 1/2 }. Suppose that W ∩ σ · W 6= ∅ for some σ 6= 1. Then there are Q, Q ∈ W such that Q = σ · Q. If Q = LuP Lu∗, then we have that Q = Lurσ P Lrσ−1 u∗ . The distance between Q and Q can be estimated as follows kQ − QkB(I) = kP − Lrσ P Lrσ−1 kB(I) ≥(cid:13)(cid:13)(cid:13)(cid:13) (pi − pσ(i))(ξ ⊗ ξ)pi(cid:13)(cid:13)(cid:13)(cid:13)I w Xi=1 where ξ ∈ R(pi) is such that kξk = 1 and σ(i) 6= i. But since Q, Q ∈ W, it follows that kQ − QkB(I) < 1, a contradiction. Hence the action is properly discontinuous, and the proof is complete. = kξ ⊗ ξkI = 1, 5 A complete Finsler metric Let Γ(t), t ∈ [0, 1], be a piecewise C1 curve in UI. One can measure the length of Γ using the norm of the symmetrically-normed ideal, i.e. LUI(Γ) =Z 1 0 k Γ(t)kI dt. Since the tangent space of UI at u can be identified with uIsh (or also with Ishu), the above length functional is well defined. There is rectifiable distance on UI defined in the standard fashion, namely dUI(u0, u1) = inf {LI (Γ) : Γ ⊆ UI, Γ(0) = u0, Γ(1) = u1} . Let P be the pinching operator associated with a family { pi }w mogeneous space, it becomes natural to put a quotient metric on the tangent spaces. Q = LuP Lu∗ for some u ∈ UI, then for [Lz, Q] ∈ (T UI(P ))Q we set 1 . Since UI(P ) is a ho- If k [Lz, Q] kQ = inf{ kz + ykI : y ∈ Ish, AduP Adu∗ (y) = y }. Indeed, the norm on (T UI(P ))Q is the Banach quotient norm of Ish by the Lie algebra of the isotropy group at Q. A standard computation shows that this metric is invariant under the action. We point out that this quotient Finsler metric was already used in several homogeneous spaces. For instance, we refer the reader to [2, 3], where some features of this metric are developed. The quotient Finsler metric on UI (P ) allow us to introduce another length functional, namely LUI(P )(γ) =Z 1 0 k γ(t)kγ, where γ(t), t ∈ [0, 1], is a continuous and piecewise C1 curve in UI (P ). Thus there is an associated rectifiable distance given by dUI(P )(Q0, Q1) = inf{ LUI(P )(γ) : γ ⊆ UI(P ), γ(0) = Q0, γ(1) = Q1 }, when the curves γ considered are continuous and piecewise C1. The next result proves that the rectifiable distance in UI (P ) can be approximated by lifting curves to UI. It is borrowed and adapted from [2]. 19 Lemma 5.1. Let Q0, Q1 ∈ UI (P ). Under the assumptions of Theorem 4.5, dUI(P )(Q0, Q1) = inf(cid:8)LUI(Γ) : Γ ⊆ UI, LΓ(0)Q0LΓ(0)∗ = Q0, LΓ(1)Q0LΓ(1)∗ = Q1(cid:9) , where the curves Γ considered are continuous and piecewise C1. Proof. Clearly it suffices to assume that Q0 = P . Let γ(t) ∈ UI (P ) be a C1 curve joining γ(0) = P and γ(1) = Q1. By Proposition 3.2 the map π : UI → UI (P ) , π(u) = LuP Lu∗ is a submersion when UI(P ) is endowed with the quotient topology, then there exists a continuous piecewise smooth curve Γ in UI such that π(Γ(t)) = γ(t), for all t ∈ [0, 1]. From the definition of the quotient Finsler metric, it is clear that the differential map of π at the identity given by δ : Ish → (T UI (P ))P , δ(z) = LzP − P Lz is contractive. Moreover, since the action is isometric, the differential map of π at any u ∈ UI has to be contractive. Using these facts we find that dUI(P )(P, Q1) ≤ LUI(P )(π(Γ)) ≤ LUI(Γ). To complete the proof, we must show that one can approximate LUI(P )(γ) with lengths of curves in UI joining the fibers of P and Q1. Fix ǫ > 0. Let 0 = t0 < t1 < . . . < tn = 1 be a uniform partition of [0, 1] (∆ti = ti − ti−1 = 1/n) such that the following hold: 1. k γ(s) − γ(s′)kB(I) < ǫ/4 if s, s′ lie in the same interval [ti−1, ti]. L (γ) − < ǫ/2. 2. (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) n−1 Xi=0 k γ (ti)kγ(ti) ∆ti(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) For each i = 0, . . . , n − 1, let xi ∈ Ish be such that δγ(ti) (xi) = γ (ti) and kxikI ≤ k γ (ti)kγ(ti) + ǫ/2. Consider the following curve Γ in UI: Γ(t) = etx0 e(t−t1)x1et1x0 e(t−t2)x2e(t2−t1)x1 et1x0 . . . e(t−tn−1)xn−1 . . . e(t2−t1)x1et1x0 t ∈ [0, t1) , t ∈ [t1, t2) , t ∈ [t2, t3) , . . . t ∈ [tn−1, 1] .   kxikI ∆ti ≤ n−1 Xi=0 Clearly Γ is continuous and piecewise smooth, Γ(0) = 1 and LUI(Γ) = n−1 Xi=0 k γ (ti)kγ(ti) + ǫ/2! ∆ti ≤ LUI(P )(γ) + ǫ. Let us show that π(Γ(1)) lies close to Q1. Indeed, first denote by α(t) = π (etx0) − γ(t), then α(0) = 0 and, using the mean value theorem in Banach spaces, for some s1 ∈ [0, t1]. Explicity, (cid:13)(cid:13)π(cid:0)et1x0(cid:1) − γ (t1)(cid:13)(cid:13)B(I) = kα (t1) − α(0)kB(I) ≤ k α (s1)kB(I) ∆t1, (cid:13)(cid:13)π(cid:0)et1x0(cid:1) − γ (t1)(cid:13)(cid:13)B(I) ≤ kLes1 x0 δQ0 (x0) Le−s1 x0 − γ (s1)kB(I) ∆t1. 20 Note that δ (x0) = γ(0), and that kLes1 x0 γ (0) Le−s1 x0 − γ (s1)kB(I) ≤ kLes1 x0 γ (0) Le−s1 x0 − γ (0)kB(I) + k γ (0) − γ (s1)kB(I) . The second summand is bounded by ǫ/4. The first summand can be bounded as follows kLes1 x0 γ (0) Le−s1 x0 − γ (0)kB(I) = kLes1 x0 γ (0) (Le−s1 x0 − I) + (Les1 x0 − I) γ (0)kB(I) ≤ 2 k γ (0)kB(I) kLes1 x0 − IkB(I) ≤ 2M ∆t1, where M = max t∈[0,1] k γ(t)kB(I) . It follows that (cid:13)(cid:13)π(cid:0)et1x0(cid:1) − γ (t1)(cid:13)(cid:13)B(I) ≤ (2M ∆t1 + ǫ/4) ∆t1. Next estimate (cid:13)(cid:13)π(cid:0)e(t2−t1)x1et1x0(cid:1) − γ (t2)(cid:13)(cid:13)B(I), which by the triangle inequality is less or equal than kLe(t2 −t1 )x1 et1 x0 P Le−t1 x0 e−(t2 −t1)x1 − Le(t2 −t1 )x1 γ (t1) Le−(t2 −t1 )x1 kB(I) + kLe(t2 −t1)x1 γ (t1) Le−(t2 −t1)x1 − γ (t2)kB(I) . The first summand is kLe(t2−t1 )x1 et1 x0 P Le−t1 x0 e−(t2 −t1)x1 − Le(t2−t1 )x1 γ (t1) Le−(t2−t1 )x1 kB(I) = = kLe(t2 −t1 )x1 (Let1 x0 P Le−t1 x0 − γ (t1)) Le−(t2 −t1)x1 kB(I) = = kLet1 x0 P Le−t1 x0 − γ (t1)kB(I) ≤ (2M ∆t1 + ǫ/4) ∆t1. The second can be treated analogously as the first difference above, kLe(t2 −t1)x1 γ (t1) Le−(t2 −t1 )x1 − γ (t2)kB(I) ≤ (2M ∆t2 + ǫ/4) ∆t2. Thus (using that ∆ti = 1/n) (cid:13)(cid:13)(cid:13)π(cid:16)e(t2−t1)x1et1x0(cid:17) − γ (t2)(cid:13)(cid:13)(cid:13)B(I) Inductively, one obtains that ≤ (2M/n + ǫ/4) 2/n. kπ (Γ (tn−1)) − γ (tn−1)kB(I) ≤ 2 (2M/n + ǫ/4) < ǫ/2 choosing n appropriately. According to Lemma 3.8, when I 6= K, or according to Proposition 3.14, when I = K, the map π has continuous local cross sections. Then one can connect Γ(tn−1) with the fiber of Q1 with a curve of arbitrary small length. In order to prove our next theorem, we need to state the next lemma (see [23, p. 109]): Lemma 5.2. Let H be a metrizable topological group, and G be a closed subgroup. If d is a complete distance function on H inducing the topology of H, and if d is invariant under the right translation by G, i.e., d (xg, yg) = d (x, y) for any x, y ∈ H and g ∈ G, then the left coset space H/G = {xG : x ∈ H} is a complete metric space under the metric d given by d (xG, yG) = inf {d (xg1, yg2) : g1, g2 ∈ G} . Moreover, the distance d is a metric for the quotient topology. We will make use of the former lemma with H = UI and G the isotropy group at P . 21 Theorem 5.3. Let Φ a symmetric norming function, and I = SΦ. Let P be the pinching operator associated with a family {pi}w 1 . If I 6= K assume in addition that w < ∞ and there is only one infinite rank projection in the family { pi }w 0 . Let u, v ∈ UI, and let dUI (LuP Lu∗, LvP Lv∗ ) = inf {dUI (uv1, vv2) : v1, v2 ∈ G} . Then, dUI = dUI(P ). In particular, (UI (P ) , dUI(P )) is a a complete metric space and dUI(P ) metricates the quotient topology. Proof. We begin by recalling that (UI, dUI ) is a complete metric space and G is dUI -closed in UI (see e.g. [12, Lemma 2.4]). Thus the quotient distance dUI is well defined. Moreover, since the multiplication by unitaries is isometric, it can be computed as dUI (LuP Lu∗, LvP Lv∗) = inf {dUI (u, vv1) : v1 ∈ G} . To prove one inequality, fix ǫ > 0. By Lemma 5.1 there is a curve Γ ∈ UI satisfying 1. Γ(0) = u, Γ(1) = vv1, with v1 ∈ G, 2. LUI(Γ) < dUI(P ) (LuP Lu∗, LvP Lv∗) + ǫ. Then we have that dUI(LuP Lu∗, LvP Lv∗) ≤ dUI (u, vv1) ≤ LUI (Γ) < dUI(P ) (LuP Lu∗, LvP Lv∗) + ǫ. Since ǫ is arbitrary, we have proved the first inequality. To show the reversed inequality, note that given ǫ > 0, there exists v1 ∈ G such that dUI (u, vv1) < dUI (LuP Lu∗ , LvP Lv∗) + ǫ Then there exists a curve Γ ⊆ UI such that Γ(0) = u, Γ(1) = vv1 and LI (Γ) < dI (u, vv1)+ǫ. So we have that dUI(P )(LuP Lu∗ , LvP Lv∗) ≤ LUI (Γ) < dUI (u, vv1) + ǫ < dUI (LuP Lu∗ , LvP Lv∗) + 2ǫ. We thus get dUI = dUI(P ). The completeness of (UI (P ) , dUI(P )) and the fact that dUI(P ) defines the quotient topology, follow from Lemma 5.2. 6 Application to the unitary orbit of a compact normal operator Let a be a compact normal operator. The question of when the full unitary orbit of a, i.e. U(a) = { uau∗ : u ∈ U }, has the property that the quotient topology coincides with the uniform norm topology was completely solved by L. A. Fialkow [16]. Both topologies coincide if and only if a has finite rank. In this section, we address the same question but with respect to the UI-unitary orbit of a, which is given by UI(a) = { uau∗ : u ∈ UI }. Though the UI-unitary orbit is in general smaller than the full unitary orbit, both orbits are equal if a has finite rank (see [19, Lemma 2.7]). Recall that for u ∈ UI, uau∗ = a + a(u∗ − 1) + (u − 1)au∗ ∈ a + I. 22 Thus one can endow UI(a) with the topology inherited from the affine Banach space a + I. On the other hand, as a homogeneous space, UI(a) may also be endowed with the quotient topology. If I is ideal of the trace class operators, it was proved by P. Bon´a [9] that both topologies coincide when a has finite rank. Later this result was extended to any symmetrically-normed ideal by D. Beltit¸a and T. Ratiu in [6, Theorem 5.10], where they also showed that the UI- unitary orbits are weakly Khaler homogeneous spaces. We will show the converse of this result and we will give a different proof of the already known implication by means of the previous results on the orbits of pinching operators. Our result is also related to the work by E. Andruchow, G. Larotonda and L. Recht [2, 3, 19], where without the assumption of a being compact, several equivalent conditions to the existence of a submanifold structure of the UI-unitary (or full unitary) orbits are described, when I is the ideal of Hilbert-Schmidt or compact operators. In particular, they established sufficient conditions to ensure that both topologies coincide. One of this conditions states that the spectrum of a must be finite. Note that this gives again the sufficient condition, since if a is compact, the spectrum of a is finite if and only if a has finite rank. Remark 6.1. The main idea to link unitary orbits of pinching operators with the UI-unitary orbit of a compact operator is the following. By the spectral theorem we may rewrite the compact normal operator a as a uniform norm convergent series, namely a = w Xi=1 λipi, (6) where 1 ≤ w ≤ ∞, λi are the nonzero distinct eigenvalues of a and { pi }w mutually orthogonal finite rank projections. ker(a − λi). Then we take P to be the pinching operator associated with { pi }w 1 . 1 is a family of Indeed, pi is the orthogonal projection onto Let u ∈ UI such that ua = au. If we use the spectral decomposition of a, we see that u 0 . This says that the isotropy group must be block diagonal with respect to the family { pi }w at a coincides with the isotropy group at P , i.e. { u ∈ UI : ua = au } = { u ∈ UI : LuP = P Lu } = G. Hence it turns out that the quotient topology on UI(a) ≃ UI/G is equal to the quotient topology on UI(P ). Theorem 6.2. Let Φ a symmetric norming function, and I = SΦ. Let a be a compact normal operator. Then the quotient topology on UI(a) coincides with the topology inherited from a + I if and only if rank(a) < ∞. Proof. Suppose that rank(a) < ∞. This is equivalent to state that w < ∞ in the spectral decomposition of a given by equation (6). Under this assumption the family { pi }w 0 has i=1 pi. Indeed, note that p0 is the orthogonal projection onto ker(a). According to Proposition 3.8 when I 6= K, or Proposition 3.14 when I = K, the quotient topology coincides with the topology inherited from B(I) on UI(P ). only one projection of infinite rank, namely p0 = 1 −Pw Since the quotient topology on UI(a) is always stronger than the topology inherited from n − akI → 0 has a + I, it remains to prove that any sequence (un)n in UI satisfying kunau∗ to be convergent to a in the quotient topology. To this end, note that kpiunpjkI ≤ λi − λj−1kuna − aunkI → 0, 23 for all i, j ≥ 0 and i 6= j (where we set λ0 = 0) . Now let x ∈ I such that kxkI = 1. Since (cid:13)(cid:13)(cid:13)(cid:13) we see that w ≤ w Xi=1 kunpi − piunkI ≤ 2Xi6=j (unpi − piun)xpi(cid:13)(cid:13)(cid:13)(cid:13)I Xi=1 n − P kB(I) = kLunP − P LunkB(I) ≤ 2Xi6=j kLunP Lu∗ kpjunpikI, kpjunpikI → 0. By the remarks in the first paragraph of this proof and Remark 6.1, the latter is equivalent to say that unau∗ n → a in the quotient topology. In order to prove the converse we assume that the quotient topology on UI(a) coincides with the topology inherited from a + I. We need to consider two cases. In the first case we suppose that I 6= K. Let M be the supplement of the Lie algebra of G defined in Proposition If rank(a) = ∞, we can construct a sequence (zk)k in M such that kzkk → 0 and 3.2. kzkkI = 1 (see Remark 3.7). i=M+1 λipik ≤ ǫ. Then it follows that kezk ae−zk − akI = k(ezk − 1)a − a(ezk − 1)kI Given ǫ > 0, let M ≥ 1 such that kPw ≤ 2(cid:18) kezk − 1k(cid:13)(cid:13)(cid:13)(cid:13) ≤ 2(cid:18) kezk − 1k(cid:13)(cid:13)(cid:13)(cid:13) M Xi=1 Xi=1 M λipi(cid:13)(cid:13)(cid:13)(cid:13)I λipi(cid:13)(cid:13)(cid:13)(cid:13)I + kezk − 1kI(cid:13)(cid:13)(cid:13)(cid:13) + e ǫ(cid:19). w Xi=M+1 (cid:19) pi(cid:13)(cid:13)(cid:13)(cid:13) Letting k → ∞, we find that ezk ae−zk → a in the norm k · kI, or equivalently, in the quotient topology. By the same argument used at the beginning of Lemma 3.8 we can arrive at kzkkI → 0, a contradiction with our previous choice of (zk)k. Now we turn to the case where I = K. Under the assumption that both topologies coincide on UK(a) we claim that the map Λ : UK(a) −→ UK(P ), Λ(uau∗) = LuP Lu∗ , is continuous, when one endows UK(a) with the topology inherited from K and UK(P ) with the topology inherited from B(K). In fact, by Proposition 3.14 the quotient and the inherited topologies always coincide on UK(P ). Then the map Λ turns out to be the identity map of UK/G, and thus our claim follows. Again we suppose that rank(a) = ∞. We will find a contradiction with the fact that Λ is continuous. Note that there must be an infinite number of finite rank projections in the family { pi }w 1 and the eigenvalues of a satisfy λi → 0. Let (ξi,j(i)) be an orthonormal basis of H such that (ξi,j(i))j(i)=1,...,rank(pi) is a basis of R(pi) for all i ≥ 1. Then take the following sequence of unitary operators: un = ξn+2,1 ⊗ ξn+1,1 + ξn+1,1 ⊗ ξn+2,1 + en, where en is the orthogonal projection onto { ξn+1,1 , ξn+2,1 }⊥. Note that un − 1 has finite rank, then un ∈ UK. Thus we get kunau∗ n − ak = kuna − aunk = k(λn+1 − λn+2) (ξn+2,1 ⊗ ξn+1,1) − (λn+2 − λn+1)(ξn+1,1 ⊗ ξn+2,1 )k ≤ 2λn+1 − λn+2 → 0. 24 On the other hand, note that kLun P Lu∗ n − P kB(K) = ∞ (unpi − piun)xpi(cid:13)(cid:13)(cid:13)(cid:13) ∞ sup Xi=1 kxk=1 , x∈K(cid:13)(cid:13)(cid:13)(cid:13) (unpi − piun)pi(cid:13)(cid:13)(cid:13)(cid:13) ≥(cid:13)(cid:13)(cid:13)(cid:13) Xi=1 = kξn+2,1 ⊗ ξn+1,1 + ξn+1,1 ⊗ ξn+2,1k = 1, which contradicts the continuity of Λ. Hence a must have finite rank, and the theorem is proved. References [1] E. Andruchow, G. Larotonda, Hopf-Rinow Theorem in the Sato Grassmanian, J. Funct. Anal. 255 (2008), no. 7, 1692-1712. [2] E. Andruchow, G. Larotonda, The rectifiable distance in the unitary Fred- holm group, Studia Math. 196 (2010), 151-178. [3] E. Andruchow, G. Larotonda, L. Recht, Finsler geometry and actions of the p-Schatten unitary groups, Trans. Amer. Math. Soc. 362 (2010), 319-344. [4] E. Andruchow, D. Stojanoff, Geometry of conditional expectations and finite index, Intern. J. of Math. 5 (1994), no. 2, 169-178. [5] D. Beltit¸a, Smooth homogeneous structures in operator theory, Chapman and Hall/CRC, Monographs and Surveys in Pure and Applied Mathematics 137, 2006. [6] D. Beltit¸a, T. Ratiu, Symplectic leaves in real banach LiePoisson spaces, Geom. Funct. Anal. 15 (2005), no. 4, 753-779. [7] D. Beltit¸a, T. Ratiu, A. Tumpach,The restricted Grassmannian, Banach Lie- Poisson spaces and coadjoint orbits, J. Funct. Anal. 247 (2007), no. 1, 138-168. [8] R.Bhatia, Pinching, trimming, truncating, and averaging of matrices, Amer. Math. Monthly 107 (2000), 602-608. [9] P. B´ona, Some considerations on topologies of infinite dimensional unitary coadjoint orbits, J. Geom. Phys. 51 (2004), no. 2, 256-268. [10] N. Bourbaki, Vari´et´es diff´erentielles et analytiques. Fascicule de r´esultats. Paragraphes 1 `a 7, Hermann, Paris. [11] J. Calkin, Two-sided ideals and congruences in the ring of bounded oper- ators in Hilbert space, Ann. of Math. 42 (1941), no. 2, 839-873. [12] E. Chiumiento, Metric geometry in infinite dimensional Stiefel manifolds, Differential Geom. Appl. 28 (2010), no. 2, 469-479. [13] J. B. Conway, The compact operators are not complemented in B(H), Proc. Amer. Math. Soc. 32 (1972), 549-550. [14] G. Corach, H. Porta, L. Recht, Differential geometry of systems of projec- tions in Banach algebras, Pacific J. Math. 143 (1990), 209-228. [15] G. Corach, H. Porta, L. Recht, The geometry of spaces of projections in C∗- algebras, Adv. Math. 101 (1993), no. 1, 59-77. [16] L. A. Fialkow, A note on unitary cross sections for operators, Canad. J. Math. 30 (1978), 1215-1227. 25 [17] I. C. Gohberg, M. G. Krein, Introduction to the theory of linear non-self- adjoint operators, Amer. Math. Soc., Providence, R.I., 1960. [18] M. Greenberg, Lectures on algebraic topology, Benjamin, New York, 1967. [19] G. Larotonda, Unitary Orbits in a Full Matrix Algebra, Integral Equations Operator Theory 54 (2006), no. 4, 511-523. [20] A. Odzijewicz, T. Ratiu, Banach Lie-Poisson spaces and reduction, Commun. Math. Phys. 243 (2003), no. 1, 1-54. [21] A. Pressley, G. Segal, Loop Groups, Oxford Mathematical Monographs. Oxford Sci- ence Publications. The Clarendon Press, Oxford University Press, New York, 1986. [22] B. Simon, Trace Ideals and Their Applications, London Mathematical Society Lecture Note Series, vol. 35, Cambridge University Press, Cambridge, 1979. [23] M. Takesaki, Theory of operator algebras III. Springer-Verlag, 1979. [24] H. Upmeier, Symmetric Banach Manifolds and Jordan C∗-Algebras, North- Holland Math. Stud. 104, Notas de Matem´atica 96, North-Holland, Amsterdam, 1985. [25] J. von Neumann, Mathematical Foundations of Quantum Mechanics, Prince- ton Univ. Press, Princeton, N.J., 1955. [26] R. Whitley, Projecting m onto c0, Amer. Math. Monthly 73 (1966), 285-286. Eduardo Chiumiento Departamento de Matem´atica, FCE-UNLP Instituto de Ciencias, UNGS Calles 50 y 115 (1900) La Plata, Argentina e-mail : [email protected] J. M. Gutierrez 1150 (1613) Los Polvorines Buenos Aires, Argentina e-mail : [email protected] Mar´ıa Eugenia Di Iorio y Lucero E. Chiumiento and M. Di Iorio y Lucero Instituto Argentino de Matem´atica "Alberto P. Calder´on", CONICET Saavedra 15 Piso 3 (1083) Buenos Aires, Argentina 26
1511.02246
1
1511
2015-11-05T13:01:13
Lectures on the classical moment problem and its noncommutative generalization
[ "math.OA", "hep-th", "math-ph", "math.FA", "math-ph" ]
These notes contain a presentation of the noncommutative generalization of the classical moment problem introduced in [10] and [12]. They also contain a short summary of the classical moment problem in infinite dimension.
math.OA
math
LECTURES ON THE CLASSICAL MOMENT PROBLEM AND ITS NONCOMMUTATIVE GENERALIZATION Michel DUBOIS-VIOLETTE 1 Abstract These notes contain a presentation of the noncommutative gener- alization of the classical moment problem introduced in [10] and [12]. They also contain a short summary of the classical moment problem in infinite dimension. 1Laboratoire de Physique Th´eorique, UMR 8627, CNRS et Universit´e Paris-Sud 11, Batiment 210, F-91 405 Orsay Cedex [email protected] Contents 1 Introduction 2 Preliminaries on ∗-algebras 2.1 Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2 The GNS construction . . . . . . . . . . . . . . . . . . . . . . 2.3 Tensor ∗-algebras . . . . . . . . . . . . . . . . . . . . . . . . . 3 Cylindrical measures and the classical moment problem 3.1 Polynomials and cylindrical functions . . . . . . . . . . . . . . 3.2 Cylindrical measures . . . . . . . . . . . . . . . . . . . . . . . 3.3 Continuity condition and Minlos theorem . . . . . . . . . . . . 3.4 . . . . . . . . . . . . . . . 3.5 The classical moment problem . . . . . . . . . . . . . . . . . . Integration of cylindrical functions 2 4 4 4 5 5 5 6 7 8 8 4 Noncommutative generalization of the classical moment prob- lem : The m-problem 9 4.1 C ∗-semi-norms . . . . . . . . . . . . . . . . . . . . . . . . . . 10 4.2 A remarkable property of tensor algebras . . . . . . . . . . . . 10 4.3 Properties of the completions . . . . . . . . . . . . . . . . . . 11 4.4 The m-problem . . . . . . . . . . . . . . . . . . . . . . . . . . 11 4.5 Solubility condition for the m-problem . . . . . . . . . . . . . 12 4.6 Convergence . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12 4.7 Homomorphisms . . . . . . . . . . . . . . . . . . . . . . . . . 13 5 Representations, self-ajdointness and determination 13 5.1 Self-adjointness and determination . . . . . . . . . . . . . . . 13 5.2 Normality and topology . . . . . . . . . . . . . . . . . . . . . 14 5.3 Some related results . . . . . . . . . . . . . . . . . . . . . . . 14 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16 5.4 Examples 1 Introduction The aim of these lectures is to give a presentation of the noncommutative generalization of the classical moment problem introduced and studied in [10] and [12] and to compare it with its commutative counterpart namely the classical moment problem. In order to do this we give an appropriate 2 description of the classical moment problem and since here we do not intend to discuss notions of dimension in the noncommutative setting, our descrip- tion should apply to the case of infinite dimensional spaces. This is why we give first a short summary of the relevant part of measure theory and of the classical moment problem in this context. We then describe the noncommu- tative generalization of the classical moment problem called the m-problem. In this generalization the algebra of complex polynomials is replaced by an arbitrary unital ∗-algebra A which is separated by its C ∗-semi-norms, the se- quence of the moment problem is replaced by a linear form on A, the measures are replaced by positive linear forms on a C ∗-algebra canonically associated with A and substitutes for the integration formulas of the classical moment problem are given. The connection between determination of the classical moment problem and the self-adjointness properties in the corresponding (unbounded) representations of the polynomials algebras are generalized. A remarkable property of tensor algebras which generalizes the solubility of the one-dimensional Hamburger's moment problem is pointed out. In the case where A is a locally convex ∗-algebra we introduce a continuity condition on the solutions of the problem which generalizes the continuity condition on cylindrical measures [13] connected with the notion of "scalar cylindrical concentration" [23]. Noncommutative measure theory has a very rich structure with no classi- cal counterpart as shown by Alain Connes (occurrence of canonically associ- ated dynamical systems) [8]. We do not discuss this subject in these lectures in spite of the fact that one meets this structure in the applications of the noncommutative moment problem (the m-problem) to quantum fields where factors of type III1 enter and where the corresponding dynamical systems should get a physical interpretation. For the proofs of the statement concerning the noncommutative moment problem we refer to [10] and [12]. As explained in details in [10] many statements there are easy consequences of powerful results of H.J. Borchers in [3], [4] and [5]. 3 2 Preliminaries on ∗-algebras 2.1 Definitions In the following, a ∗-algebra A is an associative complex algebra A endowed with an antilinear involution x 7→ x∗ such that (xy)∗ = y∗x∗ for any x, y ∈ A. An element x of A is said to be hermitian if x∗ = x. We denote by Ah the real subspace of all hermitian elements of A. A C ∗-algebra is a ∗-algebra B which is a Banach space (i.e. a complete normed space) for a norm x 7→k x k satisfying k xy k≤k x kk y k and k x∗x k=k x k2 for any x, y ∈ B. This implies k x∗ k=k x k [9], [20]. A W ∗-algebra is a C ∗-algebra R which is the dual Banach space (R∗)′ of a Banach space R∗. It can be shown that then the Banach space R∗ is unique, it is called the predual of the W ∗-algebra R [20]. A linear form φ on a ∗-algebra A is said to be positive if one has φ(x∗x) ≥ 0 for any x ∈ A. 2.2 The GNS construction Let A be a unital ∗-algebra. With any positive linear form φ on A (i.e. φ(x∗x) ≥ 0, ∀x ∈ A) is associated a Hausdorff pre-Hilbert space Dφ, an el- ement Ωφ of Dφ and a homomorphism πφ of associative algebras with units of A into the algebra of endomorphisms of Dφ satisfying Dφ = πφ(A)Ωφ, φ(x) = (Ωφπφ(x)Ωφ), (Φπφ(x)Ψ) = (πφ(x∗)ΦΨ) for any x ∈ A and Φ, Ψ ∈ Dφ. Let Hφ be the Hilbert space obtained by completion of Dφ ; the quadru- plet (πφ, Dφ, Ωφ, Hφ) is unique under the above conditions up to a unitary equivalence ; it is a (generally unbounded) ∗-representation of A called the representation associated with φ [18]. 4 2.3 Tensor ∗-algebras Let E be a real vector space and let TC(E) be the tensor algebra over the complexified space of E equipped with its structure of complex algebra with unit and the unique antilinear involution, x 7→ x∗, for which E(⊂ TC(E)) is pointwise invariant and (xy)∗ = y∗x∗ ∀x, y ∈ TC(E). Then TC(E) is a ∗-algebra with unit which we call the tensor ∗-algebra over E. This unital ∗-algebra together with the canonical embedding E ⊂ TC(E)h of E is char- acterized by the following universal property : Let A be a unital ∗-algebra, then any R-linear mapping, α : E → Ah, of E into the real vector space Ah lifts uniquely as an homomorphism TC(α) : TC(E) → A of unital ∗-algebras. Remark. Since x 7→ x∗ is canonically an anti-isomorphism of TC(E) onto the opposite ∗-algebra, the above property is equivalent to the following one : Any α : E → Ah as above lifts uniquely as an anti-homomorphism TC(α) : TC(E) → A. This remark is relevant for quantum field theory [25] because when E is the space of real test functions, the Borchers field algebra [3] is just the completion of TC(E) for a suitable topology and it is known that some space-time symmetries have to be represented there by automorphisms and some others by anti-automorphisms (e.g. TCP) of the Borchers algebra. 3 Cylindrical measures and the classical mo- ment problem 3.1 Polynomials and cylindrical functions . Let E be a real vector space with algebraic dual E ∗. Suppose that instead of working with unital ∗-algebras we are only interested in commutative uni- tal ∗-algebras. Then the analog of TC(E) is the symmetric ∗-algebra over E denoted bt SC(E). This is the complex symmetric algebra over the com- plexified vector space of E equipped with the unique anti-linear involution leaving E pointwise invariant and such that it is a commutative ∗-algebra with unit. SC(E) is also characterized by a universal property. Any R-linear mapping α : E → Ah of E into the real vector space Ah of the hermitian elements of a commutative ∗-algebra with unit A lifts uniquely as an homo- morphism SC(α) : SC(E) → A of commutative ∗-algebras with units. Let SC(E)∧ denote the set of (characters of SC(E)) all the ∗-homomorphisms 5 σ; we call these functions polynomial functions on E ∗ χ of SC(E) into C mapping the unit of SC(E) onto 1 ∈ C (χ(1l) = 1). The restriction to E ⊂ SC(E) maps SC(E)∧ into E ∗ and it follows from the above universal properly applied to the case A = C (so Ah = R) that it is a bijection (χ = SC(χ) ↾ E) of SC(E)∧ onto E ∗. Let p be an ele- ment of SC(E) and ξ an element of E ∗ ; the value at p of SC(ξ) ∈ SC(E)∧ will simply be denoted by p(ξ). Let E ∗ σ be E ∗ equipped with the weak topology σ(E ∗, E). Then ξ 7→ p(ξ) is for each p ∈ SC(E), a continuous function on E ∗ σ. These functions form a ∗-subalgebra with unit of the algebra CE ∗ of all complex functions on E ∗ which is isomorphic to SC(E) (under p 7→ (ξ 7→ p(ξ))). For any p ∈ SC(E) there is a finite family h1, · · · , hn in E and a polyno- mial function P on Rn for which p(ξ) = P (hh1, ξi, · · · , hhn, ξ)) (∀ξ ∈ E ∗). More generally a cylindrical function on E ∗ σ is a function on E ∗ of the form ξ 7→ f (hh1, ξi, · · · , hhn, ξi) for some finite family h1, · · · , hn in E and some complex function f on Rn. These functions also form a ∗-subalgebra with unit of CE ∗. Let h1, · · · , hn be a finite family in E, we denote by C(0)(h1, · · · , hn) the set of cylindrical functions ξ 7→ f (hh1, ξi, · · · , hhn, ξi) when f runs over the C ∗-algebra C(0)(Rn) of complex continuous functions vanishing at infinity on Rn. This is a C ∗-subalgebra of the C ∗-algebra Cb(E ∗ σ) of complex con- σ. Let B(SC(E), E) be the C ∗-subalgebra of tinuous bounded function on E ∗ Cb(E ∗ σ) generated by Sh∈E C(0)(h); it contains ∪C(0)(h1, · · · , hn), where the union is taken over the finite families in E, as a dense ∗-subalgebra. Let us set f (h1, · · · , hn)(ξ) = f (hh1, ξi), · · · , hhn, ξi). 3.2 Cylindrical measures Here E is again a real vector space and we use the above notations. We say that a positive linear form ω on B(SC(E), E) has the property (C) if it satisfies the following condition : ∀h ∈ E, k ω ↾ C(0)(h) k=k ω k (C) (ω ↾ C(0)(h) denotes the restriction of ω to C(0)(h) ⊂ B(SC(E), E)). It can be shown that the property (C) for ω is equivalent to the following (a priori stronger) property (C′) : k ω ↾ C(0)(h1, · · · , hn) k=k ω k (C′) for any finite family (h1, · · · , hn) in E. Thus f 7→ f (h1, · · · , hn) is a ∗- homomorphism of C(0)(Rn) in B(SC(E), E) and therefore it follows that f 7→ 6 ω(f (h1, · · · , hn)) is a positive linear form on C(0)(Rn) for any positive linear form ω on B(SC(E), E). By the Riesz theorem we have 1, ξi, · · · , hh′ 1,··· ,h′ n 1,··· ,h′ labelled by the finite families in E is called a cylindrical measure on E ∗ versely, given a cylindrical measure on E ∗ ω(f (h1, · · · , hn)) = Z f dµh1,··· ,hn for a unique positive bounded measure µh1,··· ,hn on Rn. If furthermore, ω has the property (C), this system of measures is coherent in the follow- ing sense : If f and f ′ are bounded Borel functions on Rn and Rn′ such that we have f (hh1, ξi, · · · , hhn, ξi) = f ′(hh′ n′, ξi) for any ξ ∈ E ∗ then we also have R f dµh1,··· ,hn = R f ′dµh′ ′ , (in particular R dµh1,··· ,hn = R dµh′ n =k ω k). Such a coherent family of positive bounded measure σ. Con- σ, we define a positive linear form ω on ∪C(0)(h1, · · · , hn) by ω(f (h1, · · · , hn)) = R f dµh1,··· ,hn which has norm k ω k= R dµh1,··· ,hn < ∞ and, therefore, extends uniquely into a positive lin- ear form on B(SC(E), E) which obviously has property (C). If µ is a positive bounded regular measure on E ∗ σ (i.e. a Radon measure in the sense of [23]), then ω(f ) = RE n f dµ defines (for f ∈ B(SC(E), E)) a positive linear form ω on B(SC(E), E) which has the property (C). Therefore, with a bounded pos- itive regular measure µ on E ∗ σ for any dense subspace E ′ of E ∗ σ. If E is finite dimensional the converse is also true ; but it is wrong if E has an uncountable basis and the set of bounded positive regular measures on E ∗ σ is (canonically), in this case, a strict subset of the set of cylindrical measures on E ∗ σ. σ) is associated (in an injective way) a cylindrical measure on E ∗ σ (and a fortiori on E ′ 3.3 Continuity condition and Minlos theorem Let Eθ be a real vector space E equipped with a Hausdorff locally convex topology θ. The weak dual E ′ σ of Eθ is canonically a dense topological vector subspace of E ∗ σ. It follows that the restriction f 7→ f ↾ E ′ defines an isomor- phism between B(SC(E), E) and the corresponding C ∗-algebra of continuous functions on E ′ σ and that, more generally, there are no distinctions between continuous functions on E ′ σ. In a dual man- ner, there are no distinctions between cylindrical measures on E ′ σ and on E ∗ σ (although measures on E ′ σ are distinct). There is, however, a very natural way to select by continuity a subset of cylindrical measures associ- ated with the topology θ. Namely we say that a cylindrical measure (on E ∗ σ) σ and continuous functions on E ∗ σ and E ∗ 7 satisfies the θ-continuity condition if, for any integer n and family f1, · · · , fn in C(0)(R), the corresponding positive linear form ω on B(SC(E), E) is such that (h1, · · · , hn) 7→ ω(f1(h1) · · · fn(hn)) is a continuous function on En θ . The theorem of Minlos [17], [13], [23], [7] consists in the following : If Eθ is a nu- clear space then a cylindrical measure on E ∗ σ which satisfies the θ-continuity condition is a positive bounded regular measure on E ′ σ, (notice that if Eθ is a barreled space, every bounded positive regular measure on E ′ σ satisfies this condition). 3.4 Integration of cylindrical functions Let E be a real vector space and (µh1,··· ,hn) be a cylindrical measure on E ∗ σ, we denote by ω the corresponding positive linear form on B(SC(E), E). Let f be a Borel function on Rn which is in L1(dµh1,··· ,hn). Then the integral R f dµh1,··· ,hn does only depend on the function f (h1, · · · , hn) on E ∗ ; we de- note it by ¯ω(f (h1, · · · , hn)) and call it the integral of the cylindrical function f (h1, · · · , hn). If every polynomial p ∈ SC(E) on E ∗ σ is integrable (i.e. all the µh1,··· ,hn are rapidly decreasing measures) we say that (µh1,··· ,hn) is rapidly de- creasing. In this case ¯ω defines a positive linear form on the ∗-algebra P(E ∗ σ) of continuous polynomially bounded cylindrical functions on E ∗ σ. Conversely, if ψ is a positive linear form on P(E ∗ σ), its restriction to B(SC(E), E) has the property (C) and the corresponding cylindrical measure is rapidly decreasing and ψ(f ) = ¯ω(f ) (∀f ∈ P(E ∗ σ)). 3.5 The classical moment problem Let us use the above notations and suppose that (µh1,··· ,hn) is a rapidly de- creasing cylindrical measure on E ∗ σ ; then "its moments" Z t1 · · · tndµh1,··· ,hn(t1, · · · , tn) = ¯ω(h1, · · · hn) exist (∀h1 · · · , hn in E). These moments correspond to a unique linear form φ on SC(E) (φ(p) = ¯ω(p), ∀p ∈ SC(E)) which is positive on the positive valued polynomials on E ∗ σ ; we say that φ is a strongly positive linear form on SC(E). Such a strongly positive linear form is a positive linear form on the ∗-algebra SC(E) but if dim(E) ≥ 2, there are positive linear forms on SC(E) which are not strongly positive linear forms. In the case dim(E) = 1, both concepts coincide but in this case SC(E) = TC(E) and this coincidence 8 will appear as a specific case of a general result on the algebras TC(E) (see below). In the classical moment problem, one starts from a linear form φ on SC(E) and one asks whether the φ(h1 · · · hn) are the moments of some cylindrical measure (rapidly decreasing) on E ∗ σ (notice that if this is the case for a measure µ on E ∗ σ this means φ(h1 · · · hn) = R hh1, ξi · · · hhn, ξidµ(ξ). In order that this should happen, φ must be strongly positive, and it turns out that it is sufficient. To see this, we notice that, by a direct application of the Hahn-Banach theorem, a strongly positive linear form on SC(E) has positive extensions to the cylindrical continuous polynomially bounded functions on E ∗ σ and that (see 3.4) these extensions are canonically rapidly decreasing cylindrical measures on E ∗ σ. We call these cylindrical measures solution of the moment problem for φ. If θ is a Hausdorff locally convex topology on E and if for each integer n, (h1, · · · , hn) 7→ φ(h1 · · · hn) is continuous on En θ then any solution of the moment problem satisfies the θ-continuity condition. It follows (by Minlos theorem) that if θ is a nuclear topology and if φ has the above continuity property then any solution of the moment problem for φ is a bounded positive regular measure on the weak dual E ′ σ of Eθ [15], [24]. If φ is such that there is a unique solution of the moment problem, we say that the moment problem for φ is determined ; there is a classical connection between determination and self-adjointness properties in the GNS representation associated with φ [24], [1] which will be generalized in the noncommutative case (see in Section 5). 4 Noncommutative generalization of the clas- sical moment problem : The m-problem In this section we shall generalize the concepts associated with the pair (SC(E), E) described in the introduction to pairs (A, E) where A is a (non- commutative) ∗-algebra with unit and E is a real vector subspace of Ah which is generating for A (as a unital ∗-algebra, i.e. an ∗-subalgebra of A which contains the unit and E must be identical with A). In particular we shall give a (noncommutative) generalization of cylindrical measures and of the classical moment problem (the m-problem). Special attention will be de- voted to the pair (TC(E), E) because it follows from the universal property of TC(E) that there is a unique surjective homomorphism TC(E) → A of unital ∗-algebras which induces the identity of E onto itself (notice that this is the 9 very reason why these tensor algebras enter into the formulation of quantum field theory [3]); so A is a quotient of TC(E). 4.1 C ∗-semi-norms Let A be a ∗-algebra with unit. A C ∗-semi-norm on A will be a semi-norm q on A which satisfies : q(xy) ≤ q(x)q(y), q(x∗) = q(x), q(x∗x) = q(x)2(∀x, y ∈ A) and q(1l) = 1. In other words q is such that the completion Aq of A/q−1(0) for induced norm is canonically a C ∗-algebra with unit [9]. Notice that any C ∗-semi-norm q on SC(E) is of the form p 7→ q(p) = sup{p(ξ)ξ ∈ Kq} where Kq = {ξ ∈ E ∗ k p(ξ) ≤ q(p), ∀p ∈ SC(E)} is a compact subset of E ∗ σ (every closed bounded subset of E ∗ σ is compact [21]), and that conversely p 7→ sup{p(ξ)ξ ∈ K} is a C ∗-semi-norm on SC(E) for any compact subset K of E ∗ σ. It follows (using the Stone-Weierstrass theorem) that the completion of SC(E) for the locally convex topology generated by its C ∗-semi-norms is canonically the ∗-algebra of all functions on E ∗ σ which are continuous on the compact subsets of E ∗ σ. From this algebra, it is then not hard to extract all the concepts which enter into the formulation of the classical moment problem. It is the noncommutative generalization of this algebraic construction that we shall now describe (for the proofs, we refer to [10]). In this section we shall consider unital ∗-algebras which are separated by their C ∗-semi-norms so the following result is worth noticing. 4.2 A remarkable property of tensor algebras Theorem 1. Let E be a real vector space and let TC(E)+ denote the convex hull in TC(E) of {x∗xx ∈ TC(E)}. Then TC(E)+ is a convex salient cone in TC(E) which is closed for the locally convex topology generated by the C ∗- semi-norms on TC(E). For the proof see [10], Theorem 2. This theorem implies, in particular, that the locally convex topology gen- erated by the C ∗-semi-norms on TC(E) is separated (since the closure of {0} is contained in TC(E)+ ∩ (−TC(E)+) = {0}). But it implies much more ; for instance this is wrong if TC(E) is replaced by SC(E) with dim(E) ≥ 2 because there, the closure of SC(E)+ for the locally convex topology gener- ated by the C ∗-semi-norms is (see above) the set of all the positive valued 10 polynomials on E ∗ squares of absolute values). σ which is strictly bigger than SC(E)+ = (finite sums of 4.3 Properties of the completions Proposition 2. Let A be a ∗-algebra with unit such that the locally convex topology generated by its C ∗-semi-norms is separated and let A be the topo- logical ∗-algebra with unit obtained by completion of A for this topology. We denote, as usual, by Ah the R-vector space of all hermitian elements of A and by A+ the convex hull in A of {x∗xx ∈ A} and for x ∈ A we let Sp(x) be the spectrum of x in A (i.e. Sp(x) = {λ ∈ C(x − λ1l) has no inverse in A}). Then we have the following . a) x ∈ Ah is equivalent to Sp(x) ⊂ R. b) A+ is a closed convex salient cone in A and we have : A+ = {x ∈ ASp(x) ⊂ R+} = {h2h ∈ Ah} c) Every h ∈ Ah has a unique decomposition h = h+ − h− with h+, h− ∈ A+ and h+h− = 0 d) Every h ∈ Ah determines a unique homomorphism f 7→ f (h) of ∗-algebras with units from the ∗-algebra P(R) of all complex continuous polynomially bounded functions on R into the ∗-algebra A such that IdR(h) = h (where IdR denote the identity mapping t 7→ t, of R onto itself ). This homomorphism is (automatically) continuous if P(R) is equipped with the topology of compact convergence on R and f (h) = 0 is equivalent to f ↾ Sp(h) = 0. Notice that B∞ = {x ∈ A k x k= sup{q(x)q ∈ C ∗ − semi-norms on A} ≤ ∞} is canonically a C ∗-algebra and that if h ∈ Ah and if f is a continuous bounded function on R, then f (h) ∈ B∞. 4.4 The m-problem Let A be as above and E be a real vector subspace of Ah which gener- ates A as ∗-algebra with unit. We define the C ∗-algebra B(A, E) asso- ciated with the pair (A, E) to be the C ∗-subalgebra of B∞ generated by {f (h)f ∈ C(0)(R), h ∈ E}. B(A, E) is a dense ∗-subalgebra of A (which gen- erally does not contain the unit) and, for h ∈ E, C(0)(h) = {f (h)f ∈ C(0)(R)} 11 is a C ∗-subalgebra of B(A, E). The notations are coherent with the notations introduced in 3.1 (for the case A = SC(E)) and we say that a positive linear form ω on B(A, E) has the property (C) if k ω ↾ C(0)(h) k=k ω k for any h in E. This is the noncommutative generalization of the notion of cylindrical measure we need to formulate our generalization of the moment problem on (A, E). We say that a linear form φ on A is strongly positive if it is positive valued on the closure A ∩ A+ of A+ in A equipped with the locally convex topology generated by its C ∗-semi-norms. Let φ be a linear form on A; we say that a positive linear form ω on B(A, E) is a solution of the m-problem for φ on (A, E) or of the m(E)-problem for φ if ω has a positive extension ¯ω on some ordered subspace M of A (equipped with its positive cone A+) which contains A and such that φ = ¯ω ↾ A, i.e. φ(h1 · · · hn) = ¯ω(h1 · · · hn) for h1, · · · , hn ∈ E. 4.5 Solubility condition for the m-problem Theorem 3. Let (A, E) be as above, and φ be a linear form on A. Then the m(E)-problem for φ is soluble if and only if φ is strongly positive. The set gφ(E) of all solutions of the m(E)-problem for φ is convex and compact in the weak dual of B(A, E); if φ′ 6= φ then gφ′(E) ∩ gφ(E) is empty. If ω ∈ gφ(E), then ω has the property (C) and, more generally we have : φ(hn) = R tndµh,ω(t) for any h ∈ E and integer n ≥ 0; where the measure µh,ω on R is defined by ω(f (h)) = R f (t)dµh,ω(t), ∀f ∈ C(0)(R). We notice that Theorem 1 means that every positive linear form on TC(E) is strongly positive so the last theorem implies that the m-problem is always soluble for a positive linear form on TC(E). This generalizes the solubility of one-dimensional Hamburger problem for positive moment sequences since C[X] ≃ SD(R) = TC(R) and since it is clear from the introduction that when A = SC(E) then the m(E)-problem reduces to the classical moment problem. If gφ(E) has exactly one element, we say that the m(E)-problem for φ is determined. 4.6 Convergence Proposition 4. Let φα be a net of strongly positive linear forms on A which converges weakly to φ and let ωα ∈ gφα(E) for each α. Then, φ is strongly 12 positive and any weak limit of a subnet of ωα is in gφ(E). If φα(1l) is bounded, then k ωα k= φα(1l) is bounded so the ωα belong to a weakly compact set of positive linear forms on B(A, E) and, therefore in this case there are weakly convergent subnets of ωα. 4.7 Homomorphisms If (Ai, Ei) i = 1, 2 are pairs of ∗-algebras with units Ai and generating subspace Ei ⊂ Ah i , we define a morphism α : (A1, E1) → (A2, E2) to be a homomorphism α of ∗-algebras with unit from A1 into A2 such that α(E1) ⊂ E2. It is then not hard to see that for any morphism α : (A1, E1) → (A2, E2) there is a unique ∗-homomorphism B(α) : B(A1, E1) → B(A2, E2) for which we have B(α)(f (h)) = f (α(h)), ∀h ∈ Eα and f ∈ C(0)(R). Furthermore, this correspondence is functorial. If E is a real vector space, we denote the C ∗-algebra B(TC(E), E) by B0(E). If A : E1 → E2 is a linear mapping from a real vector space E1 into another one E2, then using the universal property of TC(E) we ob- tain a unique ∗-homomorphism B0(A) : B0(E1) → B0(E2) which satisfied B0(A)(f (h)) = f (Ah), ∀h ∈ E1 and f ∈ C(0)(R) (and B0 is a covariant fonc- tor from the category of real vector spaces into the category of C ∗-algebras). However, in view of the remark made at the end of 2.3, there is also, for A as above, a unique anti ∗-homomorphism B0(A) : B0(E1) → B0(E2) satisfying B(A)(f (h)) = f ∗(Ah), ∀h ∈ E1 and f ∈ C(0)(R). 5 Representations, self-ajdointness and de- termination In this section, (A, E) is again a pair which consists of a ∗-algebra with unit A and a real vector subspace E of Ah which generates A as ∗-algebra with unit and it is assumed that A is separated by its C ∗-semi-norms. 5.1 Self-adjointness and determination Theorem 5. Let φ be a strongly positive linear form on A and let ω be a solution of the m(E)-problem for φ (ω ∈ gφ(E)). let (πφ, Dφ, Ωφ, Hφ) be the representation of A associated with φ and (πω, Ωω, Hω) be the representation 13 of B(A, E) associated with ω. Then, Hφ is canonically a Hilbert subspace of Hω with Ωφ = Ωω. If, for every h ∈ E, πφ(h) is essentially self-adjoint (on Dφ) then Hφ = Hω and πω(f (h)) = f (πφ(h)) for f ∈ C0(R) and h ∈ E; so in this case, the m(E)-problem for φ is determined (i.e. ω is unique). This generalizes the connection between self-adjointness and determina- tion in the classical moment problem. For the proof, see reference [12]. 5.2 Normality and topology . Let (A, E) be as above and let θ be a locally convex topology on E. We say that a positive linear form ω on B(A, E) satisfies the θ-continuity condition if, for every integer n and f1, · · · , fn ∈ C(0)(R), (h1, · · · , hn) 7→ ω(f1(h1) · · · fn(hn)) is a continuous function on En θ (Eθ denotes E equipped with θ). We denote by R+ ∗ (A, Eθ) the set of all the positive linear forms on B(A, Eθ) having the property (C) (see in 4.4) and satisfying the θ-continuity condition. Using standard arguments on uniform convergence, it is easily seen that R+ ∗ (A, Eθ) ⇒ ω(x∗(•)x) ∈ R+ ∗ (A, Eθ) ∀x ∈ B(A, E)) convex cone of positive linear forms on B(A, Eθ) in other words it is a folium [14]. So R+ ∗ (A, Eθ) is the cone of positive normal linear form of a W ∗-algebra R(A, Eθ) [20]. ∗ (A, Eθ) is a norm closed invariant (i.e. ω(•) ∈ R+ 5.3 Some related results Theorem 6. Let (A, E, θ) be as above and let φ be a strongly positive linear form on A such that, for each integer n, (h1, · · · , hn) 7→ φ(h1 · · · hn) is a continuous function on En θ and such that πφ(h) is essentially self-adjoint for any h in E. Then the unique solution of the m(E)-problem for φ is in R+ ∗ (A, Eθ). This theorem is a consequence of the following classical results on strong resolvent convergence [16], [19]. If Aα is a net of self-adjoint operators (with domains dom (Aα)), if A is another self-adjoint operator and if there is a core D for A in ∩α dom (Aα) such lim (AαΦ) = AΦ, ∀Φ ∈ D then Aα converges to A in the sense of strong resolvent convergence which turns out to be equivalent with strong convergence of f (Aα) to f (A) for each f ∈ C(0)(R) (and even for each continuous bounded function f on R). 14 The last result shows that the continuity condition is a good one to lock the m(E)-problem with topologies on E. There is a canonical ∗-homomorphism of B(A, E) onto a weakly [20] dense C ∗-subalgebra of R(A, Eθ) which is injective whenever the C ∗-semi-norms on A having θ-continuous restrictions to E separate A. This is the case if A = TC(E) for any Hausdorff locally convex topology θ on E and the following theorem of Borchers [5], [11] shows that, in this case, we have much more. Theorem 7. Let E be a real locally convex vector space with complexified EC = E ⊕ iE and let T (ε) C (E) be the locally convex direct sum of the nth ε EC of EC; T (ε) ε EC), (⊗0EC = C). ε-tensor powers [26] ⊗n Then for each integer N, the continuous C ∗-semi-norms on T (ε) C (E) induce n=0 (⊗nEC) a locally convex topology which coincides with its on T N topology as subspace of T (ε) C (E) = ⊕n≥0(⊗n C (E) = ⊕n=N C (E). It is worth noticing here that as shown in [2] and [27] a similar result holds for certain quadratic algebras which are "partially commutative" quotients of tensor algebras. This includes in particular the quotient of the tensor algebra over the space of test functions by the "locality ideal" [3]. If E is a Hausdorff locally convex real vector space R(TC(E), E) (resp. ∗ (TC(E), E)) will simply be denoted by R0(E) (resp. R+ R+ 0∗(E)). R0(E) contains B0(E) as a weakly dense C ∗-subalgebra (i.e. dense for the weak topology σ(R0(E), R0∗(E)) where R0∗(E) is the predual of R0(E)) and, if A : E1 → E2 is a continuous linear mapping of the locally convex real vector space E1 into another one E2, then B0(A) (resp. B0(A)) extends itself canonically as a W ∗-homomorphism R0(A) : R0(E1) → R0(E2) (resp. W ∗- anti-homomorphism R0(A) : R0(E1) → R0(E2)). Thus R0 is a functor from the category of Hausdorff locally convex real vector spaces and continuous linear mappings in the category of W ∗-algebras and W ∗-homomorphisms, (notice that R0(A2 ◦ A1) = R0(A2) ◦ R0(A1)). Theorem 8. Let E be a Hausdorff locally convex real vector space and F be a subspace of E (equipped with the induced topology). Then R0(F ) is canonically a W ∗-subalgebra of R0(E) which contains the unit of R0(E). If G is another subspace of E then R0(F +G) is generated, as W ∗-subalgebra of R0(E), by R0(F ) ∪ R0(G). If E is the completion of E, we have : R0( E) = R0(E). The above canonical identifications are natural with respect to the 15 functional calculus ; i.e. if h ∈ F ⊂ E and f ∈ C(0)(R). then f (h) ∈ R0(F ) is identified with f (h) ∈ R0(E). 5.4 Examples Let Q be the ∗-algebra generated by the Schrodinger representation of Heisen- berg canonical commutation relations [q, p] = i1l and R(Q) be the von Neu- mann algebra generated by the self-adjoint operators p and q. Let Ω be the ground state of the harmonic oscillator ; then (t1, t2) 7→ t1p+t2q is linear from R2 into Qh, so there is a unique homomorphism of unital ∗-algebra of TC(R2) onto Q, π for which π(t1, t2) = t1p + t2q and the state x 7→ (Ωπ(x)Ω) = φ(x) has the property that πφ(t1, t2) = π(t1, t2) ↾ Dφ are essentially self-ajoint. So there is a unique solution of the m(R2)-problem for φ which (by construc- tion) satisfies the continuity condition. It follows that there is a canonical surjective W ∗-homomorphism from R0(R2) onto R(Q). This shows how to apply the above theory even when there are no C ∗-semi-norm (on Q). Similar considerations apply to self-adjoint quantum fields. This leads to the net O 7→ R(O) = R0(D(O)) of W ∗-algebras where D(O) is the Schwartz space of C ∞-functions with compact supports in O ⊂ R4 [26], [22]. For developments in this direction see for instance [6]. References [1] N.I. Akhiezer. The classical moment problem. Oliver and Boyd Ltd, 1965. [2] J. Alcantara-Bode and J. Yngvason. Algebraic quantum field theory and noncommutative moment problems. I. Ann. Inst. Henri Poincar´e, 48:147 -- 159, 1988. [3] H.J. Borchers. On the structure of the algebra of field operators. Nuovo Cimento, 24:214 -- 236, 1962. [4] H.J. Borchers. Algebraic aspects of Wightman field theory. In R.V. Sen and C. Weil, editors, Statistical mechanics and field theory. Halsted Press, 1972. [5] H.J. Borchers. On the algebra of test functions. In Strasbourg 1973 IRMA/CNRS, editor, RCP 25, volume 15, pages 1 -- 14, 1973. 16 [6] H.J. Borchers and J. Yngvason. From quantum fields to local von Neu- mann algebras. Rev. Math. Phys., 4:15 -- 47, 1992. [7] G. Choquet. Lectures on analysis. Mathematics Lecture Note Series. Benjamin, 1969. [8] A. Connes. Une classification des facteurs de type III. Ann. Scient. ENS 4`eme S´erie, 6:133 -- 252, 1973. [9] J. Dixmier. Les C ∗ alg`ebres et leurs repr´esentations. Gauthier-Villars, 1964. [10] M. Dubois-Violette. A generalization of the classical moment problem on ∗-algebras with applications to relativistic quantum theory, I. Commun. Math. Phys., 43:225 -- 254, 1975. [11] M. Dubois-Violette. On some topological tensor algebras algebras. 1975. [12] M. Dubois-Violette. A generalization of the classical moment problem on ∗-algebras with applications to relativistic quantum theory, II. Commun. Math. Phys., 54:151 -- 172, 1977. [13] I.M. Guelfand and N.Y. Vilenkin. Les distributions, volume 4. Dunod, Paris, 1967. [14] R. Haag, R.V. Kadison, and D. Kastler. Nets of C ∗ algebras and clas- sification of states. Commun. Math. Phys., 16:81 -- 104, 1970. [15] G.C. Hegerfeldt. Extremal decomposition of Wightman functions and of states on nuclear ∗-algebras. Commun. Math. Phys., 45:133 -- 135, 1975. [16] T. Kato. Perturbation theory for linear operators. Number 132 in Grundlehren der mathematischen Wissenschaften. Springer Verlag, 1976. [17] R.A. Minlos. Generalized random processes and their extension to a measure, volume 3 of Selected Trans. Math. Statist. and Prob. Amer. Math. Soc., Providence, RI., 1963. [18] R.T. Powers. Self-adjoint algebras of unbounded operators. I. Commun. Math. Phys., 21:85 -- 124, 1971. 17 [19] M. Reed and B. Simon. Functional analysis. Academic Press, 1972. [20] S. Sakai. C ∗ algebras and W ∗ algebras, volume 60 of Ergebnisse der Mathematik und ihrer Grenzgebiete. Springer Verlag, 1971. [21] H.H. Schaefer. Topological vector spaces, volume 3 of Graduate Texts in Mathematics. Springer Verlag, 1971. [22] L. Schwartz. Th´eorie des distributions. Hermann, Paris, 1966. [23] L. Schwartz. Radon measures on arbitrary topological spaces and cylin- drical measures. Oxford University Press, 1973. [24] J.A. Shohat and J.B. Tamarkin. The problem of moments. Amer. Math. Soc., Providence, RI., 1963. [25] R.F. Streater and A.S. Wightman. PCT, spin and statistics and all that. The mathematical physics monograph series. Benjamin, 1964. [26] F. Treves. Topological vector spaces, distributions and kernels. Academic Press, 1967. [27] J. Yngvason. Algebraic quantum field theory and noncommutative mo- ment problems. II. Ann. Inst. Henri Poincar´e, 48:161 -- 173, 1988. 18
1501.02434
1
1501
2015-01-11T09:55:01
On Defining AW*-algebras and Rickart C*-algebras
[ "math.OA" ]
Let A be a C*-algebra. It is shown that A is an AW*-algebra if, and only if, each maximal abelian self--adjoint subalgebra of A is monotone complete. An analogous result is proved for Rickart C*-algebras; a C*-algebra is a Rickart C*-algebra if, and only if, it is unital and each maximal abelian self--adjoint subalgebra of A is monotone {\sigma}-complete.
math.OA
math
ON DEFINING AW*-ALGEBRAS AND RICKART C*-ALGEBRAS KAZUYUKI SAIT O AND J.D. MAITLAND WRIGHT Abstract. Let A be a C*-algebra. It is shown that A is an AW*-algebra if, and only if, each maximal abelian self–adjoint subalgebra of A is monotone complete. An analogous result is proved for Rickart C*-algebras; a C*-algebra is a Rickart C*-algebra if, and only if, it is unital and each maximal abelian self–adjoint subalgebra of A is monotone σ−complete. 1. AW*-algebras In this note A will be a C*-algebra which is assumed to have a unit element (unless we state otherwise). Let P rojA be the set of all projections in A. Let Asa be the self-adjoint part of A. We recall that the positive cone A+ = {zz∗ : z ∈ A} induces a partial ordering on A. Since each projection is in A+, it follows that the partial ordering of Asa induces a partial ordering on P rojA. Let us recall that a C*-algebra B is monotone complete if each norm bounded, upward directed set in Bsa has a supremum in Bsa. Then, by considering approxi- mate units, it can be shown that B always has a unit element. (Another possible definition is: each upper bounded, upward directed set in Bsa has a supremum in Bsa. For unital algebras these are equivalent but for non-unital algebras they are not the same.) Kaplansky introduced AW*-algebras as an algebraic generalisation of von Neu- mann algebras [17]. Definition 1.1. The algebra A is an AW*-algebra if (i) each maximal abelian self– adjoint subalgebra is (norm) generated by its projections and (ii) each family of orthogonal projections has a least upper bound in P rojA. When A is an AW*-algebra it can be proved that each maximal abelian ∗−subalgebra of A is monotone complete and A is unital. It has been asserted by Wright [29] and Pedersen [19] that, conversely, if each in A is monotone complete then A is an AW*-algebra. It was recently m.a.s.a. pointed out to one of us that no proof of this statement has ever been published; furthermore a straightforward approach does not work. Also some have expressed doubt as to the truth of this assertion. So in this note we repair this omission. By taking the ”correct” definition of monotone complete we can get rid of the assumption that A has a unit. Recent work by Hamhalter [9], Heunen and others, see [12, 15, 18] investigate to what extent the abelian *-subalgebras of a C*-algebra determine its structure. Also a number of interesting new results on AW*-algebras have been discovered; for example Hamhalter [8]; Heunen and Reyes [13] and [14]. So this seems a good moment to justify the assertion. But we should have written this up many years 2010 Mathematics Subject Classification. Primary46L99,37B99. 1 2 KAZUYUKI SAIT O AND J.D. MAITLAND WRIGHT ago. We can only plead, ”The carelessness of youth is followed by the regrets of old age”. The following result is elementary. But since it clarifies the partial ordering of P roj(A), we include a proof. Lemma 1.2. Let p and q be projections. Then p ≤ q if and only if p = qp. Furthermore p ≤ q implies that p and q commute. Proof. Let p ≤ q then (1 − q)p(1 − q) ≤ (1 − q)q(1 − q) = 0. Put z = (1 − q)p and observe that z2 = zz∗ = 0. So p = qp. Since p is self-adjoint, qp = (qp)∗ = pq. That is p and q commute. Conversely, suppose p = qp. By self-adjointness, qp = pq. We have (q − p)2 = q − qp − pq + p = q − p. Since q − p is a projection, it is in (cid:3) A+. So q ≥ p. Lemma 1.3. Let A be a (unital) C*-algebra. Let every maximal abelian ∗-subalgebra of A be monotone complete. Let P be a family of commuting projections. Let L be the set of all projections in A which are lower bounds for P . Then (i) L is upward directed and (ii) P has a greatest lower bound. Proof. (i) Let p and q be in L. Then each c ∈ P commutes with both p and q and hence with p + q. So P ∪ {p + q} is a set of commuting elements. This set is contained in a m.a.s.a. M1. By spectral theory, (( p+q 2 )1/n)(n = 1, 2...) is a monotone increasing sequence whose least upper bound in M1 is a projection f . By operator monotonicity [19] , for each positive integer n, and a, b in Asa, 0 ≤ a ≤ b ≤ 1 implies a1/n ≤ b1/n. For any c in P , c ≥ p and c ≥ q. So c ≥ p+q 2 . So c = c1/n ≥ (cid:18) p + q 2 (cid:19)1/n . Hence c ≥ f . Thus f ∈ L. Also So f ≥ p + q 2 ≥ 1 2 p. 1 2 0 = (1 − f )f ((1 − f ) ≥ (1 − f )p(1 − f ) ≥ 0. Using zz∗ = z2 we find that 0 = (1 − f )p. Thus f ≥ p. Similarly, f ≥ q. So L is upward directed. (ii) Let C be an increasing chain in L. Then C ∪ P is a commuting family of projections. This can be embedded in a m.a.s.a. M2. Let e be the least upper bound of C in M2. Clearly 1 ≥ e ≥ 0. To see that e is a projection we argue as follows. Since e1/2 is an upper bound for C, e1/2 ≥ e. So, by spectral theory, e ≥ e2. Since e commutes with each element of C, by spectral theory, e2 is also an upper bound for C, so e2 ≥ e. It follows that e2 = e. For each p ∈ P , p ≥ e. So e ∈ L. So every chain in L is upper bounded. So, by Zorn’s Lemma, L has a maximal element. Since L is upward directed, a maximal element is a greatest element. In other words, P has a greatest lower bound in P roj(A). (cid:3) ON DEFINING AW*-ALGEBRAS AND RICKART C*-ALGEBRAS 3 Proposition 1.4. Let A be a (unital) C*-algebra. Let every maximal abelian self- adjoint ∗-subalgebra of A be monotone complete. Then A is an AW*-algebra. Proof. Let {eλ}λ∈Λ be a family of orthogonal projections. Let P = {1−eλ : λ ∈ Λ}. Since this is a commuting family of projections, it has a greatest lower bound f in P roj(A). Hence 1 − f is the least upper bound of {eλ}λ∈Λ in P roj(A). Then, by Definition 1.1, A is an AW*-algebra. (cid:3) Theorem 1.5. Let A be a C*-algebra which is not assumed to be unital. Let each m.a.s.a. be monotone complete. Then A is a (unital) AW*-algebra. Proof. All we need to do is show that A has a unit element. Then we can apply Proposition 1.4. Given any x ∈ Asa, there is a m.a.s.a. M which contains x. Then the unit of M is a projection p such that px = x = xp. For any projection q, with p ≤ q, qx = qpx = px = x. Taking adjoints, xq = x. Arguing as in Lemma 1.3(ii), P roj(A) has a maximal element e. Arguing as in Lemma 1.3(i), P roj(A) is upward directed and so e is a largest projection. In particular, p ≤ e. So ex = x = xe. (cid:3) No one has ever seen an AW*-algebra which is not monotone complete. Are all AW*-algebras monotone complete? This is a difficult problem but Christensen and Pedersen made an impressive attack. They showed that every properly infinite AW*-algebra is monotone sequentially complete [5]. In view of Theorem 1.5, this problem could be reformulated as: if every m.a.s.a of a C*-algebra A is monotone complete is A also monotone complete? The following technical lemma will be needed later. It is usually applied with P = P rojB or with P = Bsa. Lemma 1.6. Let B be a unital C*-algebra and let M be a m.a.s.a. in B. Let P be a subset of Bsa such that uP u∗ = P whenever u is a unitary in B. Let Q be a subset of P ∩ Msa which has a least upper bound q in P . Then q is in M . Proof. Let u be any unitary in M . Then for any x in Q, uqu∗ ≥ uxu∗ = x. Then uqu∗ is in P and is an upper bound for Q. So uqu∗ ≥ q. Similarly u∗qu ≥ q, that is q ≥ uqu∗. Thus uqu∗ = q. So q commutes with each unitary in M . But each element of M is a linear combination of at most four unitaries. So q commutes with each element of M . Hence, by maximality, q ∈ M . (cid:3) Let A be an AW*-algebra and let B be a C*-subalgebra of A where B contains the unit of A. Then B is an AW*-subalgebra of A if (i) B is an AW*-algebra and (ii) whenever {eλ : λ ∈ Λ} is a set of orthogonal projections in B then its supremum in P rojB is the same as its supremum in P rojA. By Lemma 1 in [21], or see Exercise 27A of Section 4 page 27 and page 277 in [4], if B is an AW*-subalgebra of A and Q is an upward directed set in P rojB then the supremum of Q in P rojB is the same as it is in P rojA. In any C*-algebra, each abelian C*-subalgebra is contained in a m.a.s.a. 4 KAZUYUKI SAIT O AND J.D. MAITLAND WRIGHT Proposition 1.7. Let A be an AW*-algebra. Let B be a C*-subalgebra of A where B contains the unit of A. Suppose that whenever N is a m.a.s.a. in B, M is a m.a.s.a. in A and N ⊂ M then N is monotone closed in M . Then B is an AW*-subalgebra of A. The converse is also true. Proof. Let N1 be a m.a.s.a. in B then it is a subalgebra of some m.a.s.a. M1 of A. Then M1 is monotone complete because A is an AW*-algebra. By hypothesis N1 is a monotone closed subalgebra of M1. So N1 is monotone complete. Hence B is an AW*-algebra. Let C be a set of commuting projections in B such that C is upward directed. Let p be the supremum of C in P rojA. Let N2 be a m.a.s.a. of B which contains C. Let u be any unitary in N2. Then, for any c ∈ C, upu∗ ≥ c. So the projection upu∗ is an upper bound for C in P rojA. Thus upu∗ ≥ p. On replacing u by u∗, we find that u∗pu ≥ p. So p ≥ upu∗. Thus p = upu∗. So pu = up. Since each element of N2 is the linear combination of four unitaries in N2, it follows that p commutes with each element of N2. So N2 ∪ {p} is contained in a m.a.s.a. M2 of A. Let q be the supremum of C in M2. By spectral theory, q is a projection. Since p is the supremum of C in P rojA, q ≥ p. But p ∈ M2. So q = p. By hypothesis N2 is a monotone closed subalgebra of M2. So p ∈ N2 ⊂ B. So p is the supremum of C in P rojB. Now take {eλ : λ ∈ Λ} to be a set of orthogonal projections in B. Let C = {Pλ∈F eλ : F a finite, non-empty subset of Λ}. It follows from the argument above that B is an AW*-subalgebra of A Conversely suppose that B is an AW*-subalgebra of A. Take any m.a.s.a N in B and any m.a.s.a M in A with N ⊂ M . We shall show that N is monotone closed in M . Let (aα) be any norm bounded increasing net in Nsa such that aα ↑ b in Msa. We shall show b ∈ N . Suppose that kaαk ≤ k for all α. Since N is monotone complete, there exists a ∈ Nsa such that aα ↑ a in Nsa. Clearly b ≤ a. Suppose that a − b 6= 0. By spectral theory, there exist a non-zero projection p in M and a positive real number ε such that εp ≤ (a − b)p. Since aα ↑ a in Nsa, by Lemma 1.1 in [26], there exists an orthogonal family (eγ) of projections in N with supγ eγ = 1 in P roj(N ) and a family {α(γ)} such that k(a − aα)eγk ≤ ε 4 for all α ≥ α(γ) for each γ. Since B is AW*, P rojB is a complete lattice. So (eγ) has a least upper bound e in P rojB. By Lemma 1.6, e ∈ N . But supγ eγ = 1 in P roj(N ). So e = 1. Thus supγ eγ = 1 in P rojB. Since B is an AW*-subalgebra of A, it follows that supγ eγ = 1 in P rojA. Then εp ≤ (a − b)p ≤ (a − aα(γ))eγ p + (a − aα(γ))p(1 − eγ) ≤ ε 4 p + 2k(1 − eγ), that is, εp ≤ ε 4 p + 2k(1 − eγ) for all γ. So peγ ≤ 0. Thus 1 − p ≥ eγ for all γ. So, in P rojA, 1 − p ≥ 1. Thus p = 0. This is a contradiction. So b = a ∈ N . (cid:3) ON DEFINING AW*-ALGEBRAS AND RICKART C*-ALGEBRAS 5 2. Rickart C*-algebras For the purposes of this note a C*-algebra B is monotone σ−complete if each norm bounded, monotone increasing sequence in Bsa has a supremum in Bsa. In general B need not be unital. Rickart C*-algebras are related to monotone σ−complete algebras in a similar way to that of AW*-algebras to monotone complete algebras. In particular every unital monotone σ−complete algebra is well known to be a Rickart C*-algebra; see Corollary 2.6. The converse is suspected to be true but this is a hard problem. However Christensen and Pedersen [5] showed this to be true for properly infinite Rickart C*-algebras. Ara and Goldstein [3] showed that all Rickart C*-algebras are σ−normal which seems a significant step on the way to showing they are monotone σ−complete. See also [23]. Other important results on Rickart C*-algebras can be found in [1] and [2]; [10]; [11]; [6]. In [11], Handelman makes use of embed- ding in regular σ−completions [27]. We remark that normal AW*-algebras were investigated in [28], [24], [7] and [22]. Let A be a unital C*-algebra such that each m.a.s.a. is monotone σ−complete. Call such an algebra pseudo-Rickart. In [25] we obtained a result for such algebras and, without a shred of justification, called them ”Rickart”. So a natural question is : is every pseudo-Rickart C*-algebra also a Rickart C*-algebra? On the one hand, some have stated that a positive answer would be useful for applications to quantum theory [16]. On the other hand, others have expressed scepticism. By modifying the techniques of Section 1, we shall show that the answer is positive. Definition 2.1. A C*-algebra B is Rickart if, for each a ∈ B there is a projection p such that Lemma 2.2. Each Rickart C*-algebra has a unit. {z ∈ B : az = 0} = pB. Proof. In the definition put a = 0. Then B = pB for some projection p. So given any a ∈ B, there exists b, such that a = pb. Since p is a projection, pa = p2b = pb = a. Also ap = (pa∗)∗ = a∗∗ = a. (cid:3) Lemma 2.3. Let B be a C*-algebra, which need not have a unit. Let e ∈ P roj(B) and x ∈ B. Then x∗xe = 0 if, and only if, xe = 0. Proof. If x∗xe = 0 then ex∗xe = 0. So xe2 = 0. Hence xe = 0. The converse is obvious. (cid:3) Lemma 2.4. Let A be a unital C*-algebra such that each m.a.s.a. σ−complete. Let x ∈ A and let is monotone Then P has a largest element. P = {e ∈ P roj(A) : xe = 0}. Proof. It suffices to prove this when x ≤ 1 because, for any strictly positive real number ρ, P is the set of projections (left) annihilated by ρx. First we show that P is upward directed. Let p, q be in P . Then x∗x(p + q) = 0 = (p + q)x∗x. 6 KAZUYUKI SAIT O AND J.D. MAITLAND WRIGHT Let M1 be a m.a.s.a. containing x∗x and (p + q). By spectral theory, the sequence (( p+q 2 )1/n)(n = 1, 2...) is monotone increasing with supremum e in M1. Furthermore e is a projection and x∗xe = 0. So e ∈ P . Also e ≥ 1 2 (p + q). Arguing as in Lemma 1.2(i), it follows that e ≥ p and e ≥ q. Now we show that P has a maximal element. Let C be an increasing chain in P . Then C ∪ {x∗x} is contained in some m.a.s.a. M2. Then, arguing as before, ((x∗x)1/n)(n = 1, 2...) is a monotone increasing sequence with a supremum p in M2, where p is a projection and pc = 0 for each c ∈ C. Also p ≥ x∗x ≥ 0. So (1 − p)x∗x(1 − p) = 0. Hence x(1 − p) = 0. So 1 − p ∈ P . For any c ∈ C, (1 − p)c = c. So C has an upper bound, 1 − p, in P . It now follows from Zorn’s Lemma that P has a maximal element f . Since P is upward directed it follows that f is larger than every other projection in P . (cid:3) Theorem 2.5. Let A be a unital C*-algebra such that each m.a.s.a. is monotone σ−complete. Then A is a Rickart C*-algebra. Proof. Let x ∈ A and let K = {z ∈ A : xz = 0}. Let P be the set of all projections in K. By Lemma 2.4, P has a largest element f . Fix z ∈ K. We shall show that z = f z. It suffices to prove this when z ≤ 1. Since x∗xzz∗ = 0, it follows that x∗x and zz∗ are contained in some m.a.s.a. M3. The monotone increasing sequence ((zz∗)1/n)(n = 1, 2...) has supremum q in M3, where q is a projection. Also x∗xq = 0 and q ≥ zz∗. By Lemma 2.3, q ∈ P . So f ≥ q ≥ zz∗ ≥ 0. Then 0 = (1 − f )f (1 − f ) ≥ (1 − f )zz∗(1 − f ) ≥ 0. Since (1 − f )z2 = (1 − f )zz∗(1 − f ), it follows that z = f z. So K ⊂ f A. Since f ∈ K we also have f A ⊂ K. Thus K = f A. (cid:3) Corollary 2.6. Let A be a Rickart C*-algebra and let x ∈ A. There is a smallest projection q such that x = xq. Furthermore xz = 0 if, and only if, qz = 0. Proof. Since A is Rickart, the algebra is unital and each m.a.s.a. is monotone σ−complete. Let P be as in Lemma 2.4. Let Q = {1 − p : p ∈ P }. Then Q is the set of all projections p for which x = xp. Since f is the largest projection in P , 1 − f is the smallest projection in Q. Furthermore, xz = 0 if, and only if, z = f z. That is, if and only if (1 − f )z = 0. So putting q = 1 − f gives the required projection. (cid:3) Corollary 2.7. Let A be a unital C*-algebra which is monotone σ−complete. Then A is a Rickart C*-algebra. Proof. Let M be any m.a.s.a. in A. Let (an) be a norm-bounded monotone in- creasing sequence in M , with least upper bound a in Asa. By Lemma 1.6, a ∈ M . So M is monotone σ-complete. By Theorem 2.5, A is a Rickart C*-algebra. (cid:3) ON DEFINING AW*-ALGEBRAS AND RICKART C*-ALGEBRAS 7 Example 2.8. Let B(R) be the C*-algebra of all bounded complex valued functions on R. Let A be the subalgebra of all functions f such that {x : f (x) 6= 0} is countable. Then A is a monotone σ-complete C∗-algebra without unit. So A cannot be a Rickart C∗-algebra. But, since A is abelian, the only maximal abelian ∗-subalgebra is A, itself, which is monotone σ-complete. The above example shows that in Theorem 2.5 the hypothesis that A is unital is essential. However we shall tidy up some loose ends in the next section by obtaining results for non-unital algebras . 3. Weakly Rickart C*-algebras Our aim here is to show that each m.a.s.a. of a C*-algebra A is monotone σ- complete if, and only if, A is a weakly Rickart C*-algebra. In [4] (see Section 4 Theorem 1) it is shown that a unital weakly Rickart C*-algebra is a Rickart C*- algebra (and conversely). So the situation for unital C*-algebras has already been dealt with in Section 2. So here we shall suppose that A is a C*-algebra with no unit. Let us adjoin a unit to form A1. Then A is a maximal ideal of A1 and every element of A1 can be written, uniquely, as x + λ1 where x ∈ A and λ ∈ C. Since weakly Rickart C*-algebras may be slightly less familiar than Rickart C*- algebras, we give a brief account of some elementary results we need. The standard reference is [4]. Definition 3.1. [4] Let x be in a C*-algebra A. A projection e ∈ A is an annihi- lating right projection ( abbreviated as ARP according to [4]) for x if xe = x and, whenever y ∈ A satisfies xy = 0, then ey = 0. Since [4] is the standard reference on Rickart C*-algebras, we use his terminology. But ”right support projection for x ” is an alternative name (for ARP) which we find more intuitive. Definition 3.2. A C*-algebra A is weakly Rickart if each x ∈ A has an annihilating right projection e ∈ A. Lemma 3.3. Let M be any maximal abelian ∗-subalgebra of A1. Then M ∩ A is a maximal abelian ∗-subalgebra of A and M = M ∩ A + C1. Conversely, if M0 is a maximal abelian ∗-subalgebra of A and M = M0 + C1 then M is a maximal abelian ∗-subalgebra of A1. Proof. Let x ∈ A such that x commutes with each element of M ∩ A. Take any y ∈ M . Since y = a + λ1 for some a ∈ A and λ ∈ C, we have a = y − λ1 ∈ M ∩ A and so xa = ax. Hence yx = xy. So x commutes with every element of M . Since M is a m.a.s.a. in A1 it follows that x ∈ M ∩ A. So A ∩ M is a maximal abelian ∗-subalgebra of A. Clearly M = A ∩ M + C1. Now suppose M0 is a m.a.s.a. in A. Let a ∈ A such that a + λ1 commutes with each element of M . Then a commutes with each element of M0 and so a ∈ M0. So M is a m.a.s.a. in A1. (cid:3) Lemma 3.4. Let B be a C*-algebra (which may or may not be unital). Let x ∈ B have an ARP p. Then this projection is unique. Let M0 be any m.a.s.a. of B which contains x then p ∈ M0. 8 KAZUYUKI SAIT O AND J.D. MAITLAND WRIGHT Proof. Let f be an ARP of x. Then x(p − f ) = 0. So p(p − f ) = 0. Then p = pf . On taking adjoints, p = f p. Similarly, f = f p. So p = f . Let u be a unitary in M = M0 + C. So xupu∗ = uxpu∗ = uxu∗ = x. Suppose xy = 0. Then xu∗y = u∗xy = 0. So pu∗y = 0. Hence upu∗ is an ARP for x. So p = upu∗. So p commutes with each unitary in M and hence with each element of M . But M is a m.a.s.a. in B1. So p ∈ M . Since p ∈ B, it follows that p ∈ B ∩ M = M0. (cid:3) Lemma 3.5. Let A be a weakly Rickart C∗-algebra. Then each m.a.s.a. in A is monotone σ-complete. Proof. First we observe that A1 is a Rickart C∗-algebra (see [B]). So each m.a.s.a. in A1 is monotone σ-complete. Let M0 be a m.a.s.a. in A. Put M = M0 +C1. Then in A1. So M is monotone σ−complete. Let (an) by Lemma 3.3 M is a m.a.s.a. be a norm bounded increasing sequence in M0. Without loss of generality we may suppose that each an is positive and norm bounded by 1. Since M is monotone σ-complete, there exists a ∈ Msa such that an ↑ a in M . We shall show that a ∈ M0. Let en be the ARP of an in A for each n, that is, anen = an and eny = 0 1 when any = 0. By Lemma 3.4, en ∈ M0. Let x = Pn≥1 2n en. Then x ∈ M0 Let p be the ARP of x in A. Then p ∈ M0 by Lemma 3.4. Then x = xp ≤ xp ≤ p. So 1 2n en ≤ p. Hence (1 − p)en = 0. It follows that an ≤ en ≤ p. So p ≥ a ≥ 0. Since p is in M0 which is an ideal of M , it follows that a ∈ M0. (cid:3) Lemma 3.6. Let A be a non-unital C*-algebra such that each maximal abelian ∗-subalgebra of A is monotone σ-complete. Then A1 is a Rickart C*-algebra. Proof. Let M be any maximal abelian ∗-subalgebra of A1. We shall show that M is monotone σ-complete. By Lemma 2.7, M ∩ A is a maximal abelian ∗-subalgebra of A and so, it is monotone σ-complete. We claim that M = (M ∩ A) + C1, is also monotone σ- complete. Let (an) be any norm bounded monotone increasing sequence in Msa. Then, for each n, we have an = bn + λn1 with bn ∈ M ∩ A and λn ∈ R. Since A is a closed two-sided ideal of A1, (λn) is a bounded increasing sequence in R. Hence there exists λ0 ∈ R such that λn ↑ λ0. Since M ∩ A is monotone σ-complete, there exists a projection p in M ∩ A such that bnp = pbn = bn for all n. Then we have pan = anp ∈ A ∩ M for each n. Since A ∩ M is monotone σ-complete and (anp) is a norm bounded increasing sequence in (A ∩ M )sa, there exists a b ∈ (A ∩ M )sa such that anp ↑ b in (A ∩ M )sa with bp = pb = b. Since an(1 − p) = λn(1 − p) ↑ λ0(1 − p) in Msa, we have an ≤ b + λ0(1 − p) = a(∈ M ) for all n. Take any x ∈ Msa with an ≤ x for all n. Then we have anp ≤ xp for all n and an(1 − p) = λn(1 − p) ≤ x(1 − p) for all n. So ap = bp ≤ xp and λ0(1 − p) ≤ b(1 − p). So we have a ≤ x, that is, an ↑ a in Msa. So M is monotone σ-complete. It now follows from Theorem 2.5 that A1 is a Rickart C∗-algebra. (cid:3) Is the converse of Lemma 3.6 true? The following commutative example shows that it is false. ON DEFINING AW*-ALGEBRAS AND RICKART C*-ALGEBRAS 9 Example 3.7. Let ℓ∞ be the monotone σ−complete C∗-algebra of all bounded complex sequences over N. (By Theorem 2.5, ℓ∞ is Rickart.) The spectrum βN of ℓ∞ is the Stone- Cech compactification of N. Let ω be in βN but not in N. Let A = {f ∈ ℓ∞ : f (ω) = 0}. Then A is a non-unital C∗-algebra which is also a maximal closed ideal of ℓ∞ and A1 = ℓ∞. For each n, define en ∈ ℓ∞ by en = χ{1,2,··· ,n}. Clearly en ∈ P roj(A) and (en) is a norm bounded increasing sequence in Asa. Suppose A is monotone σ−complete. Then (en) has a least upper bound e in A. Then e ∈ ℓ∞. Clearly e(n) ≥ 1 for each n. So e ≥ 1 which implies that e(ω) 6= 0. This is a contradiction. Theorem 3.8. Let A be a non-unital C∗-algebra. Then A is a weakly Rickart C∗-algebra if, and only if, each maximal abelian ∗-subalgebra of A is monotone σ-complete. Proof. Lemma 3.5 gives the implication in one direction. So we now assume that in A is monotone σ−complete and wish to prove that A is weakly each m.a.s.a. Rickart. It suffices to consider x ∈ A with x ≤ 1 and show that x has an annihilating right projection in A. By Lemma 3.6, A1 is a Rickart algebra. So for some projection e ∈ A1 {z ∈ A1 : xz = 0} = (1 − e)A1. Thus e is the ARP for x in A1. By Corollary 2.6, e is the smallest projection in Q = {q ∈ P rojA1 : x = xq}. We have x∗x(1 − e) = 0. So there is a m.a.s.a. M in A1 which contains e and x∗x. By Lemma 3.3, M ∩ A is a m.a.s.a. in A and so monotone σ−complete. Then ((x∗x)1/n) is a monotone increasing sequence in M ∩ A with supremum q in M ∩ A. By spectral theory q is a projection. Also q ≥ x∗x. So 0 = (1 − q)q(1 − q) ≥ (1 − q)x∗x(1 − q) ≥ 0. Hence x(1 − q) = 0. So q ∈ Q. Thus e ≤ q. So e = eq. Since A is an ideal and q is in A, it now follows that e is in A. So x has an annihilating right projection in A. Hence A is a weakly Rickart C*-algebra. (cid:3) It is a pleasure to thank Dr. A.J. Lindenhovius, whose perceptive questions triggered this paper. References [1] P. Ara, Left and right projections are equivalent in Rickart C ∗-algebras, J. Algebra, 120(1989), 433-488. [2] P. Ara and D. Goldstein, A solution of the matrix problems for Rickart C ∗-algebras, Math. Nachr., 164(1993), 259-270. [3] P. Ara and D. Goldstein, Rickart C ∗-algebras are σ-normal, Arch. Math. (Basel), 65(1995), 505-510. [4] S.K. Berberian, Baer∗-rings, Springer Berlin 1972. [5] E. Christensen and G.K. Pedersen, Properly infinite AW ∗-algebras are monotone sequentially complete, Bull. London Math. Soc., 16(1984), 407-410. [6] K.R. Goodearl, D.E. Handelman and J.W. Lawrence, Affine representations of Grothendieck groups and applications to Rickart C ∗-algebras and ℵ0-continuous regular rings, Memoir Amer. Math. Soc., 26(1980), no.234(end of volume). ISSN 0065-9266. 10 KAZUYUKI SAIT O AND J.D. MAITLAND WRIGHT [7] M. Hamana, Regular embeddings of C ∗-algebras in monotone complete C ∗-algebras, J. Math. Soc. Japan, 33(1981), 159-183. [8] J. Hamhalter, Dye’s theorem and Gleason’s theorem for AW ∗-algebras, J. Math. Anal. Appl., 422(2015), 1103-1115. [9] J. Hamhalter, Isomorphisms of ordered structures of abelian C ∗-subalgebras of C ∗-algebras, J. Math. Anal. Appl., 383(2011), 391-399. [10] D. Handelman, Finite Rickart C ∗-algebras and their properties, Studies in Analysis, Advances in Math., Suppl. Studies, 4(1979), 171-196. [11] D. Handelman, Rickart C ∗-algebras II, Adv. Math., 48(1983), 1-15. [12] C. Heunen, many classical faces The of quantum structures, http://www.arxiv.org/abs/1412.2177 [13] C. Heunen and M.L. Reyes, Diagonalizing matrices over AW ∗-algebras, J. Functional Anal- ysis, 264(2013), 1873-1898. [14] C. Heunen and M.L. Reyes, Active lattices determine AW ∗-algebras, J. Math. Anal. Appl., 416(2014), 289-313. [15] C. Heunen and M.L. Reyes, On discretization of C ∗-algebras, http://www. arxiv.org.abs/1412.1721 [16] C. Heunen, N.P. Landsman and B. Spitters, Bohrification of operator algebras and quantum logic, Synthese, 186(2012),719-752 (Springer). [17] I. Kaplansky, Projections in Banach algebras, Ann. of Math., 53(1951), 235-249. [18] A.J. Lindenhovius, Classifying finite dimensional C*-algebras by posets of their commutative C*-algebras, International J. Theoretical Physics (to appear). [19] G.K. Pedersen, C ∗-algebras and their automorphism groups, Academic Press, 1979. [20] C.E. Rickart, Banach algebras with an adjoint operation, Ann. of Math., 47(1946), 528-550. [21] K. Saito, On the embedding as a double commutator in a type I AW ∗-algebra II, Tohoku Math. J., 26(1974), 333-339. [22] K. Saito, On normal AW ∗-algebras, Tohoku Math. J., 33(1981), 567-572. [23] K. Saito, On σ-normal C ∗-algebras, Bull. London Math. Soc., 29(1997), 480-482. [24] K. Saito and J.D.M. Wright, All AW ∗-factors are normal, J. London Math. Soc., 44(1991), 143-154. [25] K. Saito and J.D.M. Wright, C ∗-algebras which are Grothendieck spaces, Rendiconti del Circolo Matematico di Palermo, 52(2003), 141-144. [26] H. Widom, Embedding in algebras of type I, Duke Math. J., 23(1956), 309-324. [27] J.D.M. Wright, Regular σ-completions of C ∗-algebras, J. London Math. Soc., 12(1976), 299- 309. [28] J.D.M. Wright, Normal AW ∗-algebras, Proc. Roy. Soc. Edinburgh, Section A 85(1980), 137- 141. [29] J.D.M. Wright, AW ∗-algebras, Encyclopaedia of Mathematics, Kluwer, Dordrecht 1999. 2-7-5 Yoshinari, Aoba-ku, Sendai, 989-3205, Japan E-mail address: [email protected] Mathematics Institute, University of Aberdeen, Aberdeen AB24 3UE Current address: Christ Church, University of Oxford, Oxford 0X1 1DP E-mail address: [email protected] ; [email protected]
1706.08292
2
1706
2019-09-30T13:39:59
Actions of measured quantum groupoids on a finite basis
[ "math.OA" ]
In this article, we generalize to the case of measured quantum groupoids on a finite basis some important results concerning actions of locally compact quantum groups on C*-algebras [S. Baaj, G. Skandalis and S. Vaes, 2003]. Let $\cal G$ be a measured quantum groupoid on a finite basis. We prove that if $\cal G$ is regular, then any weakly continuous action of $\cal G$ on a C*-algebra is necessarily strongly continuous. Following [S. Baaj and G. Skandalis, 1989], we introduce and investigate a notion of $\cal G$-equivariant Hilbert C$^*$-modules. By applying the previous results and a version of the Takesaki-Takai duality theorem obtained in [S. Baaj and J. C., 2015] for actions of $\cal G$, we obtain a canonical equivariant Morita equivalence between a given $\cal G$-C$^*$-algebra $A$ and the double crossed product $(A\rtimes{\cal G})\rtimes\widehat{\cal G}$.
math.OA
math
A C T I O N S O F M E A S U R E D Q U A N T U M G R O U P O I D S O N A F I N I T E B A S I S by jonathan crespo Abstract In this article, we generalize to the case of measured quantum groupoids on a finite basis some important results concerning actions of locally compact quantum groups on C*-algebras [5]. Let G be a measured quantum groupoid on a finite basis. We prove that if G is regular, then any weakly continuous action of G on a C*-algebra is necessarily strongly continuous. Following [3], we introduce and investigate a notion of G-equivariant Hilbert C∗-modules. By applying the previous results and a version of the Takesaki-Takai duality theorem obtained in [2] for actions of G, we obtain a canonical product (A (cid:111) G) (cid:111) (cid:98)G. equivariant Morita equivalence between a given G-C∗-algebra A and the double crossed Keywords Locally compact quantum groups, measured quantum groupoids, monoidal equivalence, (semi-)regularity, Hilbert C*-modules, Morita equivalence. AMS classification 46L55, 16T99, 46L89. contents Introduction 1 Preliminary notations 2 Locally compact quantum groups 2.1 Hopf C*-algebras associated with a quantum group . . . . . . . . . . . . . . . 2.2 Continuous actions of locally compact quantum groups . . . . . . . . . . . . 2.3 Equivariant Hilbert C*-modules and bimodules . . . . . . . . . . . . . . . . . 3 Measured quantum groupoids 2 4 5 6 7 7 9 . . . . . . . . . . . . . . . . . . . . 11 . . . . 3.1 Case where the basis is finite-dimensional 3.2 Weak Hopf C*-algebras associated with a measured quantum groupoid on a . . . finite basis . 6 Notion of equivariant Hilbert C*-modules 4 Contributions to the notions of semi-regularity and regularity 5 Measured quantum groupoids on a finite basis in action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13 3.3 Measured quantum groupoid associated with a monoidal equivalence . . . . 14 17 20 5.1 Continuous actions, crossed product and biduality . . . . . . . . . . . . . . . 20 5.2 Case of a colinking measured quantum groupoid . . . . . . . . . . . . . . . . 25 5.3 Actions of (semi-)regular measured quantum groupoids . . . . . . . . . . . . 27 30 6.1 Actions of measured quantum groupoids on Hilbert C*-modules . . . . . . . 30 6.2 Case of a colinking measured quantum groupoid . . . . . . . . . . . . . . . . 40 6.3 Induction of equivariant Hilbert C*-modules . . . . . . . . . . . . . . . . . . . 44 54 62 8.1 Normal linear forms, weights and operator-valued weights . . . . . . . . . . 62 8.2 Relative tensor product and fiber product . . . . . . . . . . . . . . . . . . . . . 64 8.3 Unitary equivalence of Hilbert C*-modules . . . . . . . . . . . . . . . . . . . . 70 71 74 7 Takesaki-Takai duality and equivariant Morita equivalence 8 Appendix References Index of notations and symbols 2 introduction J. CRESPO The notion of monoidal equivalence of compact quantum groups has been introduced by Bichon, De Rijdt and Vaes in [6]. Two compact quantum groups G1 and G2 are said to be monoidally equivalent if their categories of representations are equivalent as monoidal C*-categories. They have proved that G1 and G2 are monoidally equivalent if and only if there exists a unital C*-algebra equipped with commuting continuous ergodic actions of full multiplicity of G1 on the left and of G2 on the right. Many crucial results of the geometric theory of free discrete quantum groups rely on the monoidal equivalence of their dual compact quantum groups. Among the applications of monoidal equivalence to this theory, we mention the contributions to randow walks and their associated boundaries [26, 15], CCAP property and Haagerup property [14], the Baum-Connes conjecture and K-amenability [29, 28]. In his Ph.D. thesis [12], De Commer has extended the notion of monoidal equivalence to the locally compact case. Two locally compact quantum groups G1 and G2 (in the sense of Kustermans and Vaes [19]) are said to be monoidally equivalent if there exists a von Neumann algebra equipped with a left Galois action of G1 and a right Galois action of G2 that commute. He proved that this notion is completely encoded by a measured quantum groupoid (in the sense of Enock and Lesieur [17]) on the basis C2. Such a groupoid is called a colinking measured quantum groupoid. The measured quantum groupoids have been introduced and studied by Lesieur and Enock (see [17, 20]). Roughly speaking, a measured quantum groupoid (in the sense of Enock-Lesieur) is an octuple G = (N, M, α, β, Γ, T, T(cid:48), ν), where N and M are von Neumann algebras (the basis N and M are the algebras of the groupoid corresponding respectively to the space of units and the total space for a classical groupoid), α and β are faithful normal *-homomorphisms from N and No (the opposite algebra) to M (corresponding to the source and target maps for a classical groupoid) with commuting ranges, Γ is a coproduct taking its values in a certain fiber product, ν is a normal semi-finite weight on N and T and T(cid:48) are operator-valued weights satisfying some axioms. In the case of a finite-dimensional basis N, the definition has been greatly simplified by De Commer [11, 12] and we will use this point of view in this article. Broadly speaking, we can 1(cid:54)l(cid:54)k Mnl (C). The relative tensor product of Hilbert spaces (resp. the fiber product of von Neumann algebras) is replaced by the ordinary tensor product of Hilbert spaces (resp. von Neumann algebras). The coproduct takes its values in M ⊗ M but is no longer unital. In the following, these objects will be referred to as 'measured quantum groupoids on a finite basis'. In [2], the authors introduce a notion of (strongly) continuous actions on C*-algebras of measured quantum groupoids on a finite basis. They extend the construction of the crossed product, the dual action and give a version of the Takesaki-Takai duality generalizing the Baaj-Skandalis duality theorem [3] in this setting. If a colinking measured quantum groupoid G, associated with a monoidal equivalence of two locally compact quantum groups G1 and G2, acts (strongly) continuously on a C*-algebra A, then A splits up as a direct sum A = A1 ⊕ A2 of C*-algebras and the action of G on A restricts to an action of G1 (resp. G2) on A1 (resp. A2). They also extend the induction procedure to the case of monoidally equivalent regular locally compact quantum groups. To any continuous action of G1 on a C*-algebra A1, they associate canonically a C*-algebra A2 endowed with a continuous action of G2. As important consequences of this construction, we mention the following: take for ν the non-normalized Markov trace on the C*-algebra N =(cid:76) • a one-to-one functorial correspondence between the continuous actions of the quan- tum groups G1 and G2, which generalizes the compact case [15] and the case of deformations by a 2-cocycle [21]; • a complete description of the continuous actions of colinking measured quantum groupoids; • the equivalence of the categories KKG1 and KKG2, which generalizes to the regular locally compact case a result of Voigt [29]. ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS 3 The proofs of the above results rely crucially on the regularity of the quantum groups G1 and G2. They prove that the regularity of G1 and G2 is equivalent to the regularity in the sense of [16] (see also [23, 24]) of the associated colinking measured quantum groupoid. By passing, this result solves some questions raised in [21] in the case of deformations by a 2-cocycle. In this article, we generalize to the case of (semi-)regular measured quantum groupoid on a finite basis some important properties of (semi-)regular locally compact quantum groups [4, 1]. This work will give us enough formulas to generalize some crucial results of [5] concerning actions of (semi-)regular locally compact quantum groups. More precisely, if G is a semi-regular measured quantum groupoid on a finite basis, then the space consisting of the continuous elements of any action of G is a C*-algebra. Moreover, if G is regular we prove that any weakly continuous action of G is necessarily continuous in the strong sense. We introduce a notion of action of G on Hilbert C*-modules in line with the corresponding notion for quantum groups [3]. A G-equivariant Hilbert C*-module is a Hilbert C*-module endowed with a continuous action (in a sense that will be specified). By using the previous result, if G is regular we prove that any action of G on a Hilbert C*-module is necessarily continuous. We are able to define the notion of G-equivariant Morita equivalence of G-C*- algebras. By applying a version of the Takesaki-Takai duality theorem obtained in [2], we product (A (cid:111) G) (cid:111) (cid:98)G in a canonical way. prove that any G-C*-algebra A is G-equivariantly Morita equivalent to its double crossed This article is organized as follows. • Chapter 1. We recall the general conventions and notations used throughout this paper. • Chapter 2. We make an overview of the theory of locally compact quantum groups (cf. [19] and [4]). We recall the construction of the Hopf C*-algebra associated with a locally compact quantum group and the notion of action of locally compact quantum groups in the C*-algebraic setting. We also recall the notion of equivariant Hilbert C*-modules (cf. [3]).• Chapter 3. We make a very brief survey of the theory of measured quantum groupoid (cf. [20, 17]) and we recall the simplified definition in the case where the basis is finite- dimensional and the associated C*-algebraic structure provided by De Commer in [11, 12]. In the last section, we make an outline of the reflection technique across a Galois object provided by De Commer (cf. [12, 13]), the construction and the structure of the colinking measured quantum groupoid associated with monoidally equivalent locally compact quantum groups. We also recall the precise description of the C*-algebraic structure of colinking measured quantum groupoids (cf. [2]). • Chapter 4. In this chapter, we make a review of the notions of regularity and semi- regularity for measured quantum groupoids on a finite basis (cf. [16, 23, 24, 2]) and we obtain new relations equivalent to the (semi-)regularity generalizing some results of Baaj and Skandalis [4, 1]. Given a (semi-)regular measured quantum groupoid, we derive new relations that will play a crucial role in the subsequent chapters. • Chapter 5. In the first section of this chapter, we recall the definitions and the main results of [2] concerning the notion of (strongly) continuous action of measured quantum groupoids on a finite basis on C*-algebras. We also recall the version of the Takesaki-Takai duality theorem obtained in [2] in this framework. The second section is dedicated to a brief overview of C*-algebras acted upon by a colinking measured quantum groupoid (cf. [2]). In the last section, we generalize to the setting of measured quantum groupoids on a finite basis the results of Baaj, Skandalis and Vaes [5] concerning the notion of weakly/strongly continuous action of (semi-)regular locally compact quantum groups. • Chapter 6. We introduce the notion of action of measured quantum groupoid on a finite basis on Hilbert C*-module and we investigate in detail the case of a colinking measured quantum groupoid. In the last paragraph, we provide a direct approach of the induction procedure for equivariant Hilbert C*-modules equivalent to that obtained in [2]. In particular, if GG1,G2 is a colinking measured quantum groupoid associated with two monoidally equivalent regular locally compact groups G1 and G2 we obtain one-to-one 4 J. CRESPO correspondences between the actions of G1, G2 and GG1,G2 on Hilbert C*-modules. • Chapter 7. In this chapter, we introduce and discuss the notion of equivariant Morita (A (cid:111) G) (cid:111) (cid:98)G are G-equivariantly Morita equivalent in a canonical way. equivalence. Given a G-C*-algebra A, we prove that A and its double crossed product • Chapter 8. In the appendix of this article, we have assembled a very brief review of the main notions and notations of the non-commutative measure and integration theory. We can also find some notations and important results used throughout this paper. In a forthcoming article [10], we use the results of this paper to generalize those of Baaj and Skandalis concerning the equivariant Kasparov theory (cf. §6 [3] and 7.7 b) [4]). Acknowledgements Some of the results of this article were part of the author's Ph.D. thesis and he wishes to thank his advisor Prof. S. Baaj for his supervision. The author is also very grateful to Prof. K. De Commer for fruitful discussions on measured quantum groupoids and for the financial support of the F.W.O. 1 preliminary notations of the multipliers of A (resp. the C*-algebra obtained from A by adjunction of a unit We specify here some elementary notations and conventions used in this article. For more notations, we refer the reader to the appendix and the index of this article. • For all subset X of a normed vector space E, we denote (cid:104)X(cid:105) (resp. [X]) the linear span (resp. closed linear span) of X in E. If X, Y ⊂ E, we denote XY := {xy ; x ∈ X, y ∈ Y}, where xy denotes the product/composition of x and y or the evaluation of x at y (when these operations make sense). If X is a subset of a *-algebra A, we denote by X∗ the subset {x∗ ; x ∈ X} of A. • We denote by ⊗ the tensor product of Hilbert spaces, the tensor product of von Neumann algebras, the minimal tensor product of C*-algebras or the external tensor product of Hilbert C*-modules. We also denote by (cid:12) (resp. (cid:12)A) the algebraic tensor product over the • Let A and B be C*-algebras. We denote by M(A) (resp. (cid:101)A) the C*-algebra consisting field of complex numbers C (resp. an algebra A). element). We denote by (cid:102)M(A ⊗ B) (or (cid:102)MB(A ⊗ B) in case of ambiguity, §1 [3]) the B- relative multiplier algebra, i.e. the C*-algebra consisting of the elements m of M((cid:101)A ⊗ B) such that the relations ((cid:101)A ⊗ B)m ⊂ A ⊗ B and m((cid:101)A ⊗ B) ⊂ A ⊗ B hold. exists a unique strictly continuous *-homomorphism π ⊗ idD : (cid:102)M(A ⊗ D) → M(B ⊗ D) Let π : A → M(B) be a (possibly degenerate) *-homomorphism. For all C*-algebra D, there satisfying the relation (π ⊗ idD)(x)(1B ⊗ d) = (π ⊗ idD)(x(1A ⊗ d)) for all x ∈ (cid:102)M(A ⊗ D) and d ∈ D. Indeed, denote by (cid:101)π the unital extension of π to (cid:101)A. The non-degenerate *-homomorphism (cid:101)π ⊗ idD : (cid:101)A ⊗ D → M(B ⊗ D) uniquely extends to M((cid:101)A ⊗ D). By restricting to (cid:102)M(A ⊗ D), we obtain the desired extension of π ⊗ idD (§1 [3]). • If x and y are two elements of an algebra A, we denote by [x, y] the commutator of x and y, i.e. the element of A defined by [x, y] := xy − yx. Let H and K be Hilbert spaces (all inner products are assumed to be anti-linear in the first variable and linear in the second variable). • We denote by B(H, K) (resp. K(H, K)) the Banach space of bounded (resp. compact) linear operators from H to K. For all ξ ∈ K and η ∈ H, we denote by θξ,η ∈ B(H, K) the rank-one operator defined by θξ,η(ζ) := (cid:104)η, ζ(cid:105)ξ for all ζ ∈ H. We have the relation K(H, K) = [θξ,η ; ξ ∈ K, η ∈ K]. Denote by B(H) := B(H, H) (resp. K(H) := K(H, H)) the C*-algebra of bounded (resp. compact) linear operators on H. Recall that K(H) is a closed two-sided ideal of B(H) and B(H) = M(K(H)). • We denote by ΣK⊗H (or simply Σ) the flip map, that is to say the unitary operator K ⊗ H → H ⊗ K ; ξ ⊗ η (cid:55)→ η ⊗ ξ. • For u ∈ B(H), we denote by Adu the bounded operator on B(H) defined for all x ∈ B(H) by Adu(x) = uxu∗. ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS 5 In this article, we will use the notion of (right) Hilbert C*-module over a C*-algebra and their tensor products (internal and external). All the definitions and conventions are those of [18]. In particular, let E and F be two Hilbert C*-modules over a C*-algebra A. • We denote by L(E, F) the Banach space consisting of all adjointable operators from E to F and L(E) the C*-algebra L(E, E). • For ξ ∈ F and η ∈ E, we denote by θξ,η the elementary operator of L(E, F) defined by θξ,η(ζ) := ξ(cid:104)η, ζ(cid:105)A for all ζ ∈ E. Let K(E, F) := [θξ,η ; ξ ∈ F, η ∈ E] be the Banach space of "compact" adjointable operators. Denote by K(E) the C*-algebra K(E, E) consisting of the compact adjointable operators of L(E). Recall that K(E) is a closed two-sided ideal of L(E) and L(E) = M(K(E)). • Let E∗ := K(E, A). We have E∗ = {T ∈ L(E, A) ; ∃ ξ ∈ E, ∀ η ∈ E, T(η) = (cid:104)ξ, η(cid:105)A}. We will identify E = K(A, E) ⊂ L(A, E). Up to this identification, for ξ ∈ E the operator ξ∗ ∈ E∗ satisfies ξ∗(η) = (cid:104)ξ, η(cid:105)A for all η ∈ E. We recall that E∗ is a Hilbert K(E)-module for the inner product defined by (cid:104)T, T(cid:48)(cid:105)K(E) := T∗ ◦ T(cid:48) for T, T(cid:48) ∈ E∗ and the right action is defined by the composition of maps. In this article, we will also use the leg numbering notation. Let H be a Hilbert space and T ∈ B(H ⊗ H). We define the operators T12, T13, T23 ∈ B(H ⊗ H) by setting T12 := T ⊗ 1, T23 := 1 ⊗ T and T13 := (Σ ⊗ 1)(1 ⊗ T)(Σ ⊗ 1). We can generalize the leg numbering notation for operators acting on any tensor product of Hilbert spaces and for adjointable operators acting on any external tensor product of Hilbert C*-modules over possibly different C*-algebras. 2 locally compact quantum groups For the notations and conventions used in this article concerning the non-commutative integration theory and the canonical objects of the Tomita-Takesaki theory, we refer the reader to the appendix of this article (cf. §8.1). 2.1 Definition. -- [19] A locally compact quantum group is a pair G = (L∞(G), ∆), where L∞(G) is a von Neumann algebra and ∆ : L∞(G) → L∞(G) ⊗ L∞(G) is a unital normal *-homomorphism satisfying the following conditions: 1. (∆ ⊗ id)∆ = (id ⊗ ∆)∆; 2. there exist n.s.f. weights ϕ and ψ on L∞(G) such that: (a) ϕ is left invariant, i.e. ϕ((ω ⊗ id)∆(x)) = ϕ(x)ω(1), for all ω ∈ L∞(G)+∗ and x ∈ M+ ϕ , x ∈ M+ ψ . (b) ψ is right inveriant, i.e. ψ((id ⊗ ω)∆(x)) = ψ(x)ω(1), for all ω ∈ L∞(G)+∗ and A left (resp. right) invariant n.s.f. weight on L∞(G) is called a left (resp. right) Haar weight (cid:78) on G. 2.2. For a locally compact quantum group G, there exists a unique left (resp. right) Haar weight on G up to a positive scalar [19]. Let us fix a locally compact quantum group G := (L∞(G), ∆). Let us fix a left Haar weight ϕ on G. Let (L2(G), π, Λ) be the G.N.S. con- struction for (L∞(G), ϕ). The left regular representation of G is the multiplicative unitary [19, 4] W ∈ B(L2(G) ⊗ L2(G)) defined by W∗(Λ(x) ⊗ Λ(y)) = (Λ ⊗ Λ)(∆(y)(x ⊗ 1)), for all x, y ∈ Nϕ. By identifying L∞(G) with its image by the G.N.S. representation π, we obtain: • L∞(G) is the strong closure of the algebra {(id ⊗ ω)(W) ; ω ∈ B(L2(G))∗}; • ∆(x) = W∗(1 ⊗ x)W, for all x ∈ L∞(G). (cid:78) 6 J. CRESPO 2.3. The Hopf-von Neumann algebra (L∞(G), ∆) admits [19] a unitary antipode map RG : L∞(G) → L∞(G) and we can choose for right Haar weight on G the weight ψ defined by ψ(x) := ϕ(RG(x)), for all x ∈ L∞(G)+. The Connes cocycle derivative [8, 25] of ψ with respect to ϕ is given by (Dψ : Dϕ)t := ν it2/2d it, for all t ∈ R, (cid:78) (cid:78) defined by: V(Λψ(x) ⊗ Λψ(y)) = (Λψ ⊗ Λψ)(∆(x)(1 ⊗ y)), where ν > 0 is the scaling constant of G and the operator dηM is the modular element ϕ := {x ∈ M ; xd1/2 is bounded and its closure xd1/2 belongs to Nϕ}. The of G [19]. Let Nd G.N.S. construction [25] for (L∞(G), ψ) is given by (L2(G), id, Λψ), where Λψ is the closure ϕ → L2(G) ; x (cid:55)→ Λ(xd1/2). We recall that Jψ = ν i/4J, where J denotes the of the map Nd (cid:78) modular conjugation for ϕ. 2.4. The right regular representation of the quantum group G is the multiplicative unitary V ∈ B(L2(G) ⊗ L2(G)) defined by for all x, y ∈ Nψ. 2.5 Definition. -- The quantum group (cid:98)G dual of G is defined by the Hopf-von Neumann algebra (L∞((cid:98)G),(cid:98)∆), where: • L∞((cid:98)G) is the strong closure of the algebra {(id ⊗ ω)(V) ; ω ∈ B(L2(G)}; • the coproduct (cid:98)∆ : L∞((cid:98)G) → L∞((cid:98)G) ⊗ L∞((cid:98)G) is defined by (cid:98)∆(x) := V∗(1 ⊗ x)V for all x ∈ L∞((cid:98)G). The quantum group (cid:98)G admits left and right Haar weights [19] and we can take the Hilbert space L2(G) for G.N.S. space. We denote by(cid:98)J the modular conjugation of the left Haar weight on (cid:98)G. We associate [4, 19] with the quantum group G two Hopf C*-algebras (S, δ) and ((cid:98)S,(cid:98)δ) • S (resp. (cid:98)S) is the norm closure of the algebra {(ω ⊗ id)(V) ; ω ∈ B(L2(G))∗} (resp. {(id ⊗ ω)(V) ; ω ∈ B(L2(G))∗}); • the coproduct δ : S → M(S ⊗ S) (resp.(cid:98)δ : (cid:98)S → M((cid:98)S ⊗(cid:98)S)) is given by: for all x ∈ (cid:98)S). We call (S, δ) (resp. ((cid:98)S,(cid:98)δ)) the Hopf C*-algebra (resp. dual Hopf C*-algebra) associated with G. We can also denote by C0(G) := S the Hopf C*-algebra of G. Note that C0((cid:98)G) = (cid:98)S. • Consider the unitary operator U := (cid:98)JJ ∈ B(L2(G)). Since for all x ∈ S (resp.(cid:98)δ(x) := V∗(1 ⊗ x)V, 2.1 Hopf C*-algebras associated with a quantum group δ(x) := V(x ⊗ 1)V∗, 2.1.1 Notations. -- U = ν i/4J(cid:98)J, we have U∗ = ν−i/4U. In particular, AdU = AdU∗ on B(L2(G)). (resp. (cid:98)S): • We have the following non-degenerate faithful representation of the C*-algebra S L : S → B(L2(G)) ; y (cid:55)→ y; R : S → B(L2(G)) ; y (cid:55)→ UL(y)U∗ (resp. ρ : (cid:98)S → B(L2(G)) ; x (cid:55)→ x; λ : (cid:98)S → B(L2(G)) ; x (cid:55)→ Uρ(x)U∗). (cid:78) regular representation of (cid:98)G is the multiplicative unitary (cid:101)V := Σ(1 ⊗ U)V(1 ⊗ U∗)Σ. It follows from 2.15 [19] that W = Σ(U ⊗ 1)V(U∗ ⊗ 1)Σ and [W12, V23] = 0. The right 2.1.2 Notation. -- Let H be a Hilbert space and X ∈ B(H ⊗ H). We denote by C(X) the norm closure of the subspace {(id⊗ ω)(ΣX) ; ω ∈ B(H)∗} of B(H). If X is a multiplicative unitary, then {(id ⊗ ω)(ΣX) ; ω ∈ B(H)∗} is a subalgebra of B(H) [4]. (cid:78) 2.1.3 Definition. -- [4, 1] The quantum group G is said to be regular (resp. semi-regular), if K(L2(G)) = C(V) (resp. K(L2(G)) ⊂ C(V)). (cid:78) Note that G is regular (resp. semi-regular) if, and only if, K(L2(G)) = C(W) (resp. K(L2(G)) ⊂ C(W)). ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS 7 2.2 Continuous actions of locally compact quantum groups We use the notations introduced in the previous paragraph. Let A be a C*-algebra. 2.2.1 Definition. -- 1. An action of the quantum group G on A is a non-degenerate *-homomorphism δA : A → M(A ⊗ S) satisfying (δA ⊗ idS)δA = (idA ⊗ δ)δA. 2. An action δA of G on A is said to be strongly (resp. weakly) continuous if [δA(A)(1A ⊗ S)] = A ⊗ S (resp. A = [(idA ⊗ ω)δA(A) ; ω ∈ B(L2(G))∗]). 3. A G-C*-algebra is a pair (A, δA), where A is a C*-algebra and δA : A → M(A ⊗ S) is (cid:78) a strongly continuous action of G on A. If G is regular, any weakly continuous action of G is necessarily continuous in the strong sense, cf. 5.8 [5]. 2.2.2 Notations. -- Let δA : A → M(A ⊗ S) (resp. δA : A → M(A ⊗(cid:98)S)) be a strongly continuous action of G (resp. (cid:98)G) on the C*-algebra A. Denote by πL (resp. (cid:98)πλ) the (resp. (cid:98)πλ := (idA ⊗ λ)δA). representation of A on the Hilbert A-module A ⊗ L2(G) defined by πL := (idA ⊗ L)δA (cid:78) 2.2.3 Definition. -- (cf. 7.1 [4]) Let (A, δA) be a G-C*-algebra (resp. (cid:98)G-C*-algebra). We call (reduced) crossed product of A by G (resp. (cid:98)G), the C*-subalgebra A (cid:111) G (resp. A (cid:111)(cid:98)G) of L(A ⊗ L2(G)) generated by the products πL(a)(1A ⊗ ρ(x)) (resp. (cid:98)πλ(a)(1A ⊗ L(x))) for a ∈ A and x ∈ (cid:98)S (resp. x ∈ S). The crossed product A (cid:111) G (resp. A (cid:111)(cid:98)G) is endowed with a strongly continuous action of (cid:98)G (resp. G), cf. 7.3 [4]. If G is regular, then the Takesaki-Takai duality extends to this setting, cf. 7.5 [4]. 2.2.4 Definition. -- Let A and B be two C*-algebras. Let δA : A → M(A ⊗ S) and δB : B → M(B ⊗ S) be two actions of G on A and B respectively. A non-degenerate *-homomorphism f : A → M(B) is said to be G-equivariant if ( f ⊗ idS)δA = δB ◦ f . We denote by AlgG the category whose objects are the G-C*-algebras and the morphisms are (cid:78) the G-equivariant non-degenerate *-homomorphisms. (cid:78) 2.3 Equivariant Hilbert C*-modules and bimodules preliminaries. In this paragraph, we briefly recall some classical notations and el- ementary facts concerning Hilbert C*-modules. Let A be a C*-algebra and E a Hilbert A-module. 2.3.1 Notations. -- Let us consider the following maps: • ιA : A → K(E ⊕ A), the *-homomorphism given by ιA(a)(ξ ⊕ b) = 0 ⊕ ab for all a, b ∈ A and ξ ∈ E; • ιE : E → K(E ⊕ A), the bounded linear map given by ιE(ξ)(η ⊕ a) = ξa ⊕ 0 for all a ∈ A and ξ, η ∈ E; • ιE∗ : E∗ → K(E ⊕ A), the bounded linear map given by ιE∗ (ξ∗)(η ⊕ a) = 0 ⊕ ξ∗η for all ξ, η ∈ E and a ∈ A; • ιK(E) : K(E) → K(E ⊕ A), the *-homomorphism given by ιK(E)(k)(η ⊕ a) = kη ⊕ 0 for all k ∈ K(E), η ∈ E and a ∈ A. (cid:78) The result below follows from straightforward computations. 2.3.2 Proposition. -- We have the following statements: 1. ιE(ξa) = ιE(ξ)ιA(a) and ιA(a)ιE∗ (ξ∗) = ιE∗ (aξ∗) for all ξ ∈ E and a ∈ A; 8 J. CRESPO 2. ιE∗ (ξ∗) = ιE(ξ)∗ and ιK(E)(θξ,η) = ιE(ξ)ιE(η)∗ for all ξ, η ∈ E; 3. ιE(ξ)∗ιE(η) = ιA((cid:104)ξ, η(cid:105)) for all ξ, η ∈ E; 4. K(E ⊕ A) is the C*-algebra generated by the set ιA(A) ∪ ιE(E). (cid:19) (cid:78) 2.3.3 Remarks. -- 1. For a ∈ A, ξ ∈ E and k ∈ K(E), the operators ιA(a), ιE(ξ), ιE∗ (ξ∗) and ιK(E)(k) can be represented by 2-by-2 matrices acting on the Hilbert A-module (cid:19) (cid:19) E ⊕ A as follows: 0 ιA(a) = 0 Moreover, any operator x ∈ K(E ⊕ A) can be written in a unique way as follows: (cid:18)0 0 ιK(E)(k) = (cid:18) 0 (cid:18)k ιE(ξ) = ∗) = (cid:19) ιE∗ (ξ 0 a ξ∗ ξ 0 ; 0 0 ; 0 . (cid:18)0 (cid:19) 0 , with k ∈ K(E), ξ, η ∈ E and a ∈ A. ; x = (cid:18) k η∗ ξ a 2. Note that ιA and ιK(E) extend uniquely to strictly/*-strongly continuous unital *- homomorphisms ιA : M(A) → L(E ⊕ A) and ιK(E) : L(E) → L(E ⊕ A). Besides, we have ιA(m)(ξ ⊕ a) = 0 ⊕ ma and ιK(E)(T)(ξ ⊕ a) = Tξ ⊕ 0 for all m ∈ M(A), T ∈ L(E), ξ ∈ E and a ∈ A. 3. ιE∗ admits an extension to a bounded linear map ιE∗ : L(E, A) → L(E ⊕ A) in a straightforward way. Similarly, up to the identification E = K(A, E), we can also extend ιE to a bounded linear map ιE : L(A, E) → L(E ⊕ A). 4. As in 1, we can represent the operators ιA(m), ιK(E)(T), ιE∗ (S) and ιE(S∗), for m ∈ M(A), T ∈ L(E) and S ∈ L(A, E), by 2-by-2 matrices. Moreover, any operator x ∈ L(E ⊕ A) can be written in a unique way as follows: , with T ∈ L(E), S, S(cid:48) ∈ L(A, E) and m ∈ M(A). (cid:78) (cid:19) (cid:18) T S(cid:48) S∗ m x = By using the matrix notations described above, we derive easily the following useful technical lemma: 2.3.4 Lemma. -- Let x ∈ L(E ⊕ A) (resp. x ∈ K(E ⊕ A)). We have: 1. x ∈ ιE(L(A, E)) (resp. ιE(E)) if, and only if, xιE(ξ) = 0 for all ξ ∈ E and ιA(a)x = 0 for all a ∈ A; in that case, we have ιA(m)x = 0 for all m ∈ M(A); 2. x ∈ ιK(E)(L(E)) (resp. ιK(E)(K(E))) if, and only if, xιA(a) = 0 and ιA(a)x = 0 for all (cid:78) a ∈ A; in that case, we have xιA(m) = ιA(m)x = 0 for all m ∈ M(A). inner product is given by: (cid:104)ξ, η(cid:105) := ξ {T ∈ L(A ⊗ B, E ⊗ B) ; ∀x ∈ B, (1E ⊗ x)T ∈ E ⊗ B and T(1A ⊗ x) ∈ E ⊗ B}. Let us recall the notion of relative multiplier module, cf. 2.1 [3]. A. Up to the identification E ⊗ B = K(A ⊗ B, E ⊗ B), we define (cid:102)M(E ⊗ B) (or (cid:102)MB(E ⊗ B) 2.3.5 Definition. -- Let A and B be two C*-algebras and let E be a Hilbert C*-module over in case of ambiguity) to be the following subspace of L(A ⊗ B, E ⊗ B): Note that (cid:102)M(E ⊗ B) is a Hilbert C*-module over (cid:102)M(A ⊗ B), whose (cid:102)M(A ⊗ B)-valued Note also that we have K((cid:102)M(E ⊗ B)) ⊂ (cid:102)M(K(E) ⊗ B). (cid:78) 2.3.6 Proposition-Definition. -- Let B ⊂ B(K) be a C*-algebra of operators on a Hilbert space K. For all T ∈ L(A ⊗ B, E ⊗ B) and ω ∈ B(K)∗, there exists a unique (idE ⊗ ω)(T) ∈ L(A, E) such that is non-degenerate and T ∈ (cid:102)M(E ⊗ B), then we have (idE ⊗ ω)(T) ∈ E. where ιE⊗B : L(A ⊗ B, E ⊗ B) → L((E ⊗ B) ⊕ (A ⊗ B)) = M(K(E ⊕ A) ⊗ B). If B ⊂ B(K) Proof. This is a direct consequence of 2.3.4 1 and the fact that ιE⊗B(T) ∈ (cid:102)M(K(E ⊕ A) ⊗ B) (cid:78) if T ∈ (cid:102)M(E ⊗ B). for all ξ, η ∈ (cid:102)M(E ⊗ B) ⊂ L(A ⊗ B, E ⊗ B). ιE(idE ⊗ ω)(T) = (idK(E⊕A) ⊗ ω)(ιE⊗B(T)) ∈ L(E ⊕ A), ∗ ◦ η, ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS 9 notion of equivariant hilbert c*-module. In this paragraph, we recall the notion of equivariant Hilbert C*-module through the three equivalent pictures developed in §2 [3]. Let us fix a G-C*-algebra (A, δA) and a Hilbert A-module E. 2.3.7 Definition. -- An action of the locally compact quantum group G on E is a linear map δE : E → (cid:102)M(E ⊗ S) such that: 1. δE(ξ)δA(a) = δE(ξa) and δA((cid:104)ξ, η(cid:105)) = (cid:104)δE(ξ), δE(η)(cid:105), for all a ∈ A and ξ, η ∈ E; 2. [δE(E)(A ⊗ S)] = E ⊗ S; 3. the linear maps δE ⊗ idS and idE ⊗ δ extend to linear maps from L(A ⊗ S, E ⊗ S) to L(A ⊗ S ⊗ S, E ⊗ S ⊗ S) and we have (δE ⊗ idS)δE = (idE ⊗ δ)δE. An action δE of G on E is said to be continuous if we have [(1E ⊗ S)δE(E)] = E ⊗ S. A action δE : E → (cid:102)M(E ⊗ S) of G on E. G-equivariant Hilbert A-module is a Hilbert A-module E endowed with a continuous (cid:78) 2.3.8. The datum of a continuous action of G on E is equivalent to that of a continuous action δK(E⊕A) : K(E ⊕ A) → M(K(E ⊕ A) ⊗ S) of G on the linking C*-algebra K(E ⊕ A) such that the *-homomorphism ιA : A → K(E ⊕ A) is G-equivariant, cf. 2.7 [3]. (cid:78) 2.3.9. If δE is an action of G on E, we have the unitary operator V ∈ L(E ⊗δA (A ⊗ S), E ⊗ S) defined by V(ξ ⊗δA x) := δE(ξ)x for all ξ ∈ E and x ∈ A ⊗ S. This unitary satisfies the relation in L(E ⊗δ2 A (A ⊗ S ⊗ S), E ⊗ S ⊗ S), ( V ⊗C idS)( V ⊗δA⊗idS 1) = V ⊗idA⊗ δ 1 A := (δA ⊗ idS)δA = (idA ⊗ δ)δA, cf. 2.3 and 2.4 (a) [3] for the details. Conversely, where δ2 and the fact that VTξ ∈ (cid:102)M(E ⊗ S) for all ξ ∈ E, where Tξ ∈ L(A ⊗ S, E ⊗δA (A ⊗ S)) is if there exists a unitary operator V ∈ L(E ⊗δA (A ⊗ S), E ⊗ S) satisfying the above relation defined by Tξ (x) := ξ ⊗δA x for all x ∈ A ⊗ S, then the map δE : E → (cid:102)M(E ⊗ S) ; ξ (cid:55)→ VTξ 2.3.10. An action of G on E determines an action δK(E) : K(E) → (cid:102)M(K(E) ⊗ S) of G on K(E) defined by δK(E)(k) = V(k ⊗δA 1) V∗ for all k ∈ K(E), where V is the unitary operator associated to the action, cf. 2.8 [3]. Moreover, if E is a G-equivariant Hilbert module, then K(E) turns into a G-C*-algebra. (cid:78) is an action of G on E, cf. 2.4 (b) [3]. (cid:78) 3 measured quantum groupoids For reminders concerning the relative tensor product of Hilbert spaces and the fiber product of von Neumann algebras, we refer the reader to the appendix of this article (cf. 8.2). 3.1 Definition. -- (cf. 3.7 [17], 4.1 [20]) We call a measured quantum groupoid an octuple G = (N, M, α, β, Γ, T, T(cid:48), ν), where: • M and N are von Neumann algebras; • Γ : M → M β(cid:63)α M is a faithful normal unital *-homomorphism, called the coproduct; • α : N → M and β : No → M are faithful normal unital *-homormorphisms, called the range and source maps of G; • T : M+ → α(N)ext • ν is a n.s.f. weight on N; + and T(cid:48) : M+ → β(No)ext + are n.s.f. operator-valued weights; such that the following conditions are satisfied: 1. [α(n(cid:48)), β(no)] = 0, for all n, n(cid:48) ∈ N; 2. Γ(α(n)) = α(n) β⊗α 1 and Γ(β(no)) = 1 β⊗α β(no), for all n ∈ N; 10 J. CRESPO 3. Γ is coassociative, i.e. (Γ 4. the n.s.f. weights ϕ and ψ on M given by ϕ = ν ◦ α−1 ◦ T and ψ = ν ◦ β−1 ◦ T(cid:48) satisfy: β(cid:63)α id)Γ = (id β(cid:63)α Γ)Γ; ∀x ∈ M+ T(cid:48), T(cid:48)(x) = (ψ β(cid:63)α id)Γ(x), • ∀x ∈ M+ • σϕ t and σψ T , T(x) = (id β(cid:63)α ϕ)Γ(x), s commute for all s, t ∈ R. (cid:78) Let G = (N, M, α, β, Γ, T, T(cid:48), ν) be a measured quantum groupoid. We denote by (H, π, Λ) the G.N.S. construction for (M, ϕ) where ϕ := ν ◦ α−1 ◦ T. Let (σt)t∈R, ∇ and J be respectively the modular automorphism group, the modular operator and the modular conjugation for ϕ. In the following, we identify M with its image by π in B(H). • We have a coinvolutive *-antiautomomorphism RG : M → M such that R2G = idM (cf. 3.8 [17]). From now on, we will assume that T(cid:48) = RG ◦ T ◦ RG and then also ψ = ϕ ◦ RG. • There exist self-adjoint positive non-singular operators λ and d respectively affiliated to Z (M) and M such that (Dψ : Dϕ)t = λit2/2d it for all t ∈ R. The operators λ and d are respectively called the scaling operator and the modular operator of G. • The G.N.S. construction for (M, ψ) is given by (H, πψ, Λψ), where: Λψ is the closure of the operator which sends any element x ∈ M such that xd1/2 is closable and its closure xd1/2 ∈ Nϕ to Λϕ(xd1/2); πψ : M → B(H) is given by the formula πψ(a)Λψ(x) = Λψ(ax). G σ(cid:98)βα; WG (x β⊗α 1) W∗ β⊗α H → H α⊗(cid:98)β H the pseudo-multiplicative unitary of • The modular conjugation Jψ for ψ is given by Jψ = λi/4J. • We will denote by WG : H G (cf. 3.3.4 [27], 3.6 [17]). measured quantum groupoid (cid:98)G := (N, (cid:98)M, α,(cid:98)β,(cid:98)Γ,(cid:98)T,(cid:98)R ◦(cid:98)T ◦(cid:98)R, ν), where: 3.2 Proposition-Definition. -- (cf. 3.10 [17]) We define the (Pontryagin) dual of G to be the • (cid:98)M is the von Neumann algebra generated by {(ω (cid:63) id)( WG ) ; ω ∈ B(H)∗} ⊂ B(H); • (cid:98)β : No → (cid:98)M is given by(cid:98)β(no) := Jα(n∗)J for all n ∈ N; (cid:63)α (cid:98)M is given for all x ∈ (cid:98)M by(cid:98)Γ(x) := σα(cid:98)β • (cid:98)Γ : (cid:98)M → (cid:98)M(cid:98)β • there exists a unique n.s.f. weight (cid:98)ϕ on (cid:98)M whose G.N.S. construction is (H, id, Λ(cid:98)ϕ), where Λ(cid:98)ϕ is the closure of the operator (ω (cid:63) id)( WG ) (cid:55)→ aϕ(ω) defined for normal linear forms ω in a dense subspace of Iϕ = {ω ∈ B(H)∗ ; ∃k ∈ R+, ∀x ∈ Nϕ, ω(x∗)2 (cid:54) kϕ(x∗x)} and aϕ(ω) ∈ H satisfies ω(x∗) = (cid:104)Λϕ(x), aϕ(ω)(cid:105) for all x ∈ Nϕ; • (cid:98)T is the unique n.s.f. operator-valued weight from (cid:98)M to α(N) such that (cid:98)ϕ = ν ◦ α−1 ◦(cid:98)T and (cid:98)T(cid:48) = R(cid:98)G ◦(cid:98)T ◦ R(cid:98)G, where R(cid:98)G : (cid:98)M → (cid:98)M is given by R(cid:98)G (x) := Jx∗ J for all x ∈ (cid:98)M. The pseudo-multiplicative unitary W(cid:98)G of (cid:98)G is given by W(cid:98)G = σβα We will denote by(cid:98)J the modular conjugation for (cid:98)ϕ. Note that the scaling operator of (cid:98)G is λ−1. In particular, we have λit ∈ Z (M) ∩ Z ((cid:98)M) for all t ∈ R. • Let(cid:98)α(n) := Jβ(no)∗ J =(cid:98)J(cid:98)β(no)∗(cid:98)J for n ∈ N. We recall the following relations (cf. 3.11 (v) [17]): M ∩ (cid:98)M = α(N), M ∩ (cid:98)M(cid:48) = β(No), M(cid:48) ∩ (cid:98)M = (cid:98)β(No) and M(cid:48) ∩ (cid:98)M(cid:48) =(cid:98)α(N). • Let U := (cid:98)JJ ∈ B(H). Then, U∗ = λ−i/4U and U2 = λi/4 (cf. 3.11 (iv) [17]). In particular, U is unitary. We have(cid:98)α(n) = Uα(n)U∗ and (cid:98)β(no) = Uβ(no)U∗ for all n ∈ N. Since λ−i/4 ∈ Z (M), we also have(cid:98)α(n) = U∗α(n)U and (cid:98)β(no) = U∗β(no)U for all n ∈ N. G σ(cid:98)βα. W∗ (cid:78) 3.3 Proposition-Definition. -- (cf. 3.12 [17]) ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS 11 • The octuple (No, M, β, α, ςβα ◦ Γ, RG ◦ T ◦ RG, T, νo) is a measured quantum groupoid by WGo = (β(cid:98)J α⊗(cid:98)α(cid:98)J(cid:98)β denoted by Go and called the opposite of G. The pseudo-multiplicative unitary of Go is given ) WG (β(cid:98)J α⊗α(cid:98)Jβ). • Let CM : M → M(cid:48) be the canonical *-antihomomorphism given by CM(x) := Jx∗ J for all x ∈ M. Let us define: Γc := (CM β(cid:63)α CM) ◦ Γ ◦ C−1 M ; RcG := CM ◦ RG ◦ C−1 M ; Tc = CM ◦ T ◦ C−1 M . ) WG (β J(cid:98)α⊗α J(cid:98)β Then, the octuple (No, M(cid:48),(cid:98)β,(cid:98)α, Γc, Tc, RcG TcRcG, νo) is a measured quantum groupoid de- noted by Gc and called the commutant of G. The pseudo-multiplicative unitary WGc of Gc is given by WGc = ((cid:98)β J α⊗α J(cid:98)β (cid:78) 3.4 Notations. -- For a given measured quantum groupoid G, we will need the following (cid:98)V := WG; pseudo-multiplicative unitaries: 3.5 Convention. -- Henceforth, we will refer to ((cid:98)G)c instead of (cid:98)G as the dual of G since this groupoid is better suited for right actions of G. We have (cid:101)V := W V := W(cid:91)(Go) ((cid:98)G)c; (Go)c. ). = W (cid:78) ((cid:98)G)c = (No, (cid:98)M(cid:48), β,(cid:98)α,(cid:98)Γc,(cid:98)Tc,(cid:98)Tc(cid:48), νo), ((cid:98)G)c )∗(1 β⊗α x) W ((cid:98)G)c, for all x ∈ (cid:98)M(cid:48); where the coproduct and the operator-valued weights are given by: • (cid:98)Γc(x) = ( W • (cid:98)Tc = C(cid:98)M ◦(cid:98)T ◦ C−1(cid:98)M , where C(cid:98)M : (cid:98)M → (cid:98)M(cid:48) ; x (cid:55)→(cid:98)Jx∗(cid:98)J; • (cid:98)Tc(cid:48) = R((cid:98)G)c ◦(cid:98)Tc ◦ R((cid:98)G)c. Note also that the commutant weight (cid:98)ϕc := νo ◦ β−1 ◦(cid:98)Tc derived from the weight (cid:98)ϕ is left invariant for the coproduct(cid:98)Γc. In the following, we will simply denote by (cid:98)G the dual groupoid of G (since no ambiguity will arise with the Pontryagin dual). Note that the bidual groupoid is (Go)c = (Gc)o. (cid:78) 3.1 Case where the basis is finite-dimensional In [12], De Commer provides an equivalent definition of a measured quantum groupoid on a finite basis. This definition is far more tractable since it allows us to circumvent the use of relative tensor products and fiber products. In the following, we fix a finite-dimensional C*-algebra N := (cid:76) with the non-normalized Markov trace  := (cid:76) 1(cid:54)l(cid:54)k Mnl (C) endowed 1(cid:54)l(cid:54)k nl · Trl, where Trl denotes the non- normalized trace on Mnl (C). We refer to §8.2 of the appendix for the definitions of vβα and qβα. Let us a fix a measured quantum groupoid G = (N, M, α, β, Γ, T, T(cid:48), ). We have a unital normal *-isomorphism M β(cid:63)α M → qβα(M ⊗ M)qβα ; x (cid:55)→ v∗ βαxvβα (cf. 8.2.14). Let ∆ : M → M ⊗ M be the (non necessarily unital) faithful normal *-homomorphism given by ∆(x) = v∗ Γ(x)vβα for all x ∈ M. We have ∆(1) = qβα. This has led De Commer to the following equivalent definition of a measured quantum groupoid on a finite basis. 3.1.1 Definition. -- (cf. 11.1.2 [12]) A measured quantum groupoid on the finite-dimen- sional basis N is an octuple G = (N, M, α, β, ∆, T, T(cid:48), ), where: βα • M is a von Neumann algebra, α : N → M and β : No → M are unital faithful normal *-homomorphisms; • ∆ : M → M ⊗ M is a faithful normal *-homomorphism; 12 J. CRESPO • T : M+ → α(N)ext + and T(cid:48) : M+ → β(No)ext such that the following conditions are satisfied: + are n.s.f. operator-valued weights; 1. [α(n(cid:48)), β(no)] = 0, for all n, n(cid:48) ∈ N; 2. ∆(1) = qβα; 3. (∆ ⊗ id)∆ = (id ⊗ ∆)∆; 4. ∆(α(n)) = ∆(1)(α(n) ⊗ 1) and ∆(β(no)) = ∆(1)(1 ⊗ β(no)), for all n ∈ N; 5. the n.s.f. weights ϕ and ψ on M given by ϕ :=  ◦ α−1 ◦ T and ψ :=  ◦ β−1 ◦ T(cid:48) satisfy: T(x) = (id ⊗ ϕ)∆(x) t ◦ β = β and σT(cid:48) for all x ∈ M+ t ◦ α = α, for all t ∈ R. 6. σT T , T(cid:48)(x) = (ψ ⊗ id)∆(x), for all x ∈ M+ T(cid:48); (cid:78) H, H . Let ((cid:98)V) Similarly, we also define ια β⊗α H) → B(H ⊗ H) β(cid:98)β V := ιβ(cid:98)αα( V), W := ια ιβ(cid:98)αα : B(H(cid:98)α⊗β β(cid:98)β Let us fix a measured quantum groupoid G = (N, M, α, β, ∆, T, T(cid:48), ). 3.1.2 Notations. -- Let us consider the injective bounded linear map βαXv(cid:98)αβ. ; X (cid:55)→ v∗ ((cid:101)V), and ι(cid:98)α(cid:98)ββ where V = W(cid:98)G, (cid:98)V = WG and (cid:101)V = W In what follows, we recall the main properties satisfied by V, W and (cid:101)V. The proof of unitaries V, (cid:98)V and (cid:101)V (cf. [17], §11 [12] and §2 [2]). 3.1.3 Proposition. -- (cf. 3.11 (iii), 3.12 (v), (vi) [17], 2.2 [2]) The operators V, W and (cid:101)V are multiplicative partial isometries acting on H ⊗ H such that: (cid:101)V = Σ(1 ⊗ U)V(1 ⊗ U∗)Σ = (U ⊗ U)W(U∗ ⊗ U∗); 1. W = Σ(U ⊗ 1)V(U∗ ⊗ 1)Σ, 2. V∗ = (J ⊗(cid:98)J)V(J ⊗(cid:98)J), W∗ = ((cid:98)J ⊗ J)W((cid:98)J ⊗ J); the results below are derived from the properties satisfied by the pseudo-multiplicative and (cid:101)V := ι(cid:98)α(cid:98)ββ (Go)c (cf. 3.4). (cid:78) and 3. the initial and final projections are given by 3.1.4 Proposition. -- (cf. 3.8, 3.12 [17]) V∗V = q(cid:98)αβ = (cid:101)V(cid:101)V∗, W∗W = qβα = VV∗, WW∗ = qα(cid:98)β (cid:101)V∗(cid:101)V = q(cid:98)β(cid:98)α. (cid:78) 1. The von Neumann algebra M (resp. (cid:98)M) is the weak closure of {(id ⊗ ω)(W) ; ω ∈ B(H)∗} (resp. {(ω ⊗ id)(W) ; ω ∈ B(H)∗}). 2. We have W ∈ M ⊗ (cid:98)M, V ∈ (cid:98)M(cid:48) ⊗ M, and (cid:101)V ∈ M(cid:48) ⊗ (cid:98)M(cid:48). In particular, we have the commutation relations [W12, V23] = 0 and [V12, (cid:101)V23] = 0. 3. The coproduct ∆ : M → M ⊗ M of G (resp.(cid:98)∆ : (cid:98)M(cid:48) → (cid:98)M(cid:48) ⊗ (cid:98)M(cid:48) of (cid:98)G) satisfies ∆(x) = V(x ⊗ 1)V∗ = W∗(1 ⊗ x)W, (resp.(cid:98)∆(x) = V∗(1 ⊗ x)V = (cid:101)V(x ⊗ 1)(cid:101)V∗, for all x ∈ (cid:98)M(cid:48)). for all x ∈ M 3.1.5 Proposition. -- (cf. 3.2. (i), 3.6. (ii) [17] and 11.1.2 [12]) For all n ∈ N, we have: [V, 1 ⊗(cid:98)β(no)] = 0; [V, 1 ⊗(cid:98)α(n)] = 0, 1. [V, α(n) ⊗ 1] = 0, 2. V(1 ⊗ α(n)) = ((cid:98)α(n) ⊗ 1)V, V(β(no) ⊗ 1) = (1 ⊗ β(no))V; [V,(cid:98)β(no) ⊗ 1] = 0, (cid:78) ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS [W,(cid:98)α(n) ⊗ 1] = 0, [(cid:101)V, β(no) ⊗ 1] = 0, 3. [W,(cid:98)β(no) ⊗ 1] = 0, [W, 1 ⊗ β(no)] = 0, 4. W(1 ⊗(cid:98)β(no)) = (β(no) ⊗ 1)W, W(α(n) ⊗ 1) = (1 ⊗ α(n))W; [(cid:101)V, 1 ⊗ α(n)] = 0, 5. [(cid:101)V, α(n) ⊗ 1] = 0, 6. (cid:101)V(1 ⊗ β(no)) = ((cid:98)β(no) ⊗ 1)(cid:101)V, (cid:101)V((cid:98)α(n) ⊗ 1) = (1 ⊗(cid:98)α(n))(cid:101)V. 3.1.6 Proposition. -- (cf. 11.1.4 [12]) For all n ∈ N, we have: (1 ⊗(cid:98)β(no))W = (α(n) ⊗ 1)W; 1. W(β(no) ⊗ 1) = W(1 ⊗ α(n)), 2. V(1 ⊗ β(no)) = V((cid:98)α(n) ⊗ 1), (1 ⊗ α(n))V = (β(no) ⊗ 1)V; 3. (cid:101)V((cid:98)β(no) ⊗ 1) = (cid:101)V(1 ⊗(cid:98)α(n)), (1 ⊗ β(no))(cid:101)V = ((cid:98)α(n) ⊗ 1)(cid:101)V. [W, 1 ⊗(cid:98)α(n)] = 0; [(cid:101)V, 1 ⊗(cid:98)β(no)] = 0; 13 (cid:78) (cid:78) 3.2 Weak Hopf C*-algebras associated with a measured quantum groupoid on a finite basis We recall the definitions and the main results concerning the weak Hopf C*-algebras associated with a measured quantum groupoid on a finite basis, cf. §11.2 [12] (with different notations and conventions, cf. §2.3 [2]). Let us fix a measured quantum groupoid G = (N, M, α, β, ∆, T, T(cid:48), ) on the finite-dimensional basis N =(cid:76) 3.2.1 Notations. -- With the notations of §3.1, we denote by S (resp. (cid:98)S) the norm closure 1(cid:54)l(cid:54)k Mnl (C). of the subalgebra {(ω ⊗ id)(V) ; ω ∈ B(H)∗} (resp.{(id ⊗ ω)(V) ; ω ∈ B(H)∗}). According to §11.2 [12], we have the following statements: • the Banach space S (resp. (cid:98)S) is a non-degenerate C*-subalgebra of B(H), weakly dense in M (resp. (cid:98)M(cid:48)); • the C*-algebra S (resp. (cid:98)S) is endowed with the faithful non-degenerate *-representa- tions: L : S → B(H) ; x (cid:55)→ x; R : S → B(H) ; x (cid:55)→ UL(x)U∗ (resp. ρ : (cid:98)S → B(H) ; x (cid:55)→ x; λ : (cid:98)S → B(H) ; x (cid:55)→ Uρ(x)U∗); • α(N) ⊂ M(S), β(No) ⊂ M(S), β(No) ⊂ M((cid:98)S) and(cid:98)α(N) ⊂ M((cid:98)S); • V ∈ M((cid:98)S ⊗ S), W ∈ M(S ⊗ λ((cid:98)S)) and (cid:101)V ∈ M(R(S) ⊗(cid:98)S); • ∆ (resp.(cid:98)∆) restricts to a strictly continuous *-homomorphism δ : S → M(S ⊗ S) (resp. (cid:98)δ : (cid:98)S → M((cid:98)S ⊗(cid:98)S)), which uniquely extends to a strictly continuous *-homomorphism δ : M(S) → M(S ⊗ S) (resp. (cid:98)δ : M((cid:98)S) → M((cid:98)S ⊗(cid:98)S)) satisfying δ(1S) = qβα (resp. (cid:98)δ(1(cid:98)S) = q(cid:98)αβ); • δ (resp.(cid:98)δ) is coassociative and satisfies [δ(S)(1S ⊗ S)] = δ(1S)(S⊗ S) = [δ(S)(S⊗ 1S)] (resp. [(cid:98)δ((cid:98)S)(1(cid:98)S ⊗(cid:98)S)] =(cid:98)δ(1(cid:98)S)((cid:98)S ⊗(cid:98)S) = [(cid:98)δ((cid:98)S)((cid:98)S ⊗ 1(cid:98)S)]); • the unital faithful *-homomorphisms α : N → M(S) and β : No → M(S) satisfy for all n ∈ N; δ(α(n)) = δ(1S)(α(n) ⊗ 1S) and δ(β(no)) = δ(1S)(1S ⊗ β(no)), • the unital faithful *-homomorphisms β : No → M((cid:98)S) and(cid:98)α : N → M((cid:98)S) satisfy (cid:98)δ(β(no)) =(cid:98)δ(1(cid:98)S)(β(no) ⊗ 1(cid:98)S) and (cid:98)δ((cid:98)α(n)) =(cid:98)δ(1(cid:98)S)(1(cid:98)S ⊗(cid:98)α(n)), for all n ∈ N. (cid:78) 3.2.2 Definition. -- With the above notations, we call the pair (S, δ) (resp. ((cid:98)S,(cid:98)δ)) the weak Hopf C*-algebra (resp. dual weak Hopf C*-algebra) associated with the measured quantum groupoid G. 3.2.3 Remark. -- With the notations of the above definition, the pair ((cid:98)S,(cid:98)δ) is the weak (cid:78) Hopf C*-algebra of (cid:98)G while its dual weak Hopf C*-algebra is the pair (R(S), δR), where R(S) = USU∗ and the coproduct δR is given by δR(y) := (cid:101)V∗(1 ⊗ y)(cid:101)V for all y ∈ R(S). (cid:78) 14 J. CRESPO 3.3 Measured quantum groupoid associated with a monoidal equivalence We will recall the construction of the measured quantum groupoid associated with a monoidal equivalence between two locally compact quantum groups provided by De Commer [12, 13]. First of all, we will need to recall the definitions and the crucial results of De Commer [12, 13]. 3.3.1 Definition. -- Let G be a locally compact quantum group. A right (resp. left) Galois action of G on a von Neumann algebra N is an ergodic integrable right (resp. left) action αN : N → N ⊗ L∞(G) (resp. γN : N → L∞(G) ⊗ N) such that the crossed product N (cid:111)αN G (cid:110) N) is a type I factor. Then, the pair (N, αN) (resp. (N, γN)) is called a right (resp. G γN (cid:78) (resp. left) Galois object for G. Let G be a locally compact quantum group and let us fix a right Galois object (N, αN) for G. In his thesis, De Commer was able to build a locally compact quantum group H equipped with a left Galois action γN on N commuting with αN, i.e. (id ⊗ αN)γN = (γN ⊗ id)αN. This construction is called the reflection technique and H is called the reflected locally compact quantum group across (N, αN). In a canonical way, he was also able to associate a right Galois object (O, αO) for H and a left Galois action γO : O → L∞(G) ⊗ O of G on O commuting with αO. Finally, De Commer has built a measured quantum groupoid GH,G = (C2, M, α, β, ∆, T, T(cid:48), ) where: M = L∞(H) ⊕ N ⊕ O ⊕ L∞(G); ∆ : M → M ⊗ M is made up of the coactions and coproducts of the constituents of M; the operator-valued weights T and T(cid:48) are given by the invariants weights; the non-normalized Markov trace  on C2 is simply given by (a, b) = a + b for all (a, b) ∈ C2. Moreover, the source and target maps α and β have range in Z (M) and generate a copy of C4. Conversely, if G = (C2, M, α, β, ∆, T, T(cid:48), ) is a measured quantum groupoid whose source and target maps have range in Z (M) and generate a copy of C4, then G is of the form GH,G in a unique way, where H and G are locally compact quantum groups canonically associated with G. In what follows, we fix a measured quantum groupoid G = (C2, M, α, β, ∆, T, T(cid:48), ) whose source and target maps have range in Z (M) and generate a copy of C4. It is worth noticing that for such a groupoid we have: 3.3.2 Lemma. -- (cf. 2.21 [2])(cid:98)α = β and(cid:98)β = α. (cid:78) Following the notations introduced in [12], we recall the precise description of the left and right regular representations W and V of G introduced in the previous section. We identify M with its image by π in B(H), where (H, π, Λ) is the G.N.S. construction for M endowed with the n.s.f. weight ϕ =  ◦ α−1 ◦ T. We also consider the n.s.f. weight ψ =  ◦ β−1 ◦ T(cid:48). Denote by (ε1, ε2) the standard basis of the vector space C2. 3.3.3 Notations. -- Let us introduce some useful notations and make some remarks concerning them. • For i, j = 1, 2, we define the following nonzero central self-adjoint projection of M: pij := α(εi)β(εj). It follows from β(ε1) + β(ε2) = 1M and α(ε1) + α(ε2) = 1M that α(εi) = pi1 + pi2 and β(εj) = p1j + p2j, for all i, j = 1, 2. • We have ∆(1) = α(ε1) ⊗ β(ε1) + α(ε2) ⊗ β(ε2) and (cid:98)∆(1) = β(ε1) ⊗ β(ε1) + β(ε2) ⊗ β(ε2) since(cid:98)α = β. ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS 15 • Let Mij := pijM, for i, j = 1, 2. Then, Mij is a nonzero von Neumann subalgebra of M. • Let Hij := pijH, for i, j = 1, 2. Then, Hij is a nonzero Hilbert subspace of H for all i, j = 1, 2. • Let ϕij := ϕ (cid:22) n.s.f. weights on Mij. (Mij)+ and ψij := ψ (cid:22) (Mij)+, for i, j = 1, 2. Then, ϕij and ψij are • For all i, j, k = 1, 2, we denote by ∆k ij homomorphism given by : Mij → Mik ⊗ Mkj the unital normal *- ij(x) := (pik ⊗ pkj)∆(x), ∆k for all x ∈ Mij. • We have Jpkl = pkl J, (cid:98)Jpkl = plk(cid:98)J and Upkl = plkU for all k, l = 1, 2. We define the anti-unitaries Jkl : Hkl → Hkl,(cid:98)Jkl : Hkl → Hlk and the unitary Ukl : Hkl → Hlk by setting Jkl = pkl Jpkl,(cid:98)Jkl = plk(cid:98)Jpkl and Ukl = plkUpkl =(cid:98)Jkl Jkl. • For i, j, k, l = 1, 2, let Σij⊗kl := ΣHij⊗Hkl : Hij ⊗ Hkl → Hkl ⊗ Hij. (cid:78) We readily obtain: (cid:77) (cid:77) M = Mij; H = Hij; ∆(pij) = pi1 ⊗ p1j + pi2 ⊗ p2j, for all i, j = 1, 2. i,j=1,2 i,j=1,2 Note that in terms of the parts ∆k ik ⊗ idMkj )∆k ij of ∆, the coassociativity condition reads as follows: ij = (idMil ⊗ ∆k for all i, j, k, l = 1, 2. lj)∆l ij, (∆l The G.N.S. representation for (Mij, ϕij) is obtained by restriction of the G.N.S. representa- tion of (M, ϕ) to Mij. In particular, the G.N.S. space Hϕij is identified with Hij. 3.3.4 Proposition. -- For all i, j, k, l = 1, 2, we have: (pij ⊗ 1)V(pkl ⊗ 1) = δi (1 ⊗ pij)W(1 ⊗ pkl) = δl (1 ⊗ pji)(cid:101)V(1 ⊗ plk) = δl k(pij ⊗ pjl)V(pil ⊗ pjl); j(pik ⊗ pij)W(pik ⊗ pkj); j(pki ⊗ pji)(cid:101)V(pki ⊗ pjk). ik : Hik ⊗ Hkj → Hik ⊗ Hij and (cid:101)Vj jl = (pij ⊗ pjl)V(pil ⊗ pjl), Wj 3.3.5 Notations. -- The operators V, W and (cid:101)V each splits up into eight unitaries jl : Hil ⊗ Hjl → Hij ⊗ Hjl, Wj Vi (cid:101)Vj ki = (pki ⊗ pji)(cid:101)V(pki ⊗ pjk). for i, j, k, l = 1, 2, given by Vi ki : Hki ⊗ Hjk → Hki ⊗ Hji ik = (pik ⊗ pij)W(pik ⊗ pkj) and (cid:78) Let i, j, k, l, l(cid:48) = 1, 2. These unitaries are related to each other by the following relations (cf. 3.1.3): (cid:78) jk ⊗ 1)Σik⊗kj; (cid:101)Vj ik(U∗ ki = Σji⊗ki(1 ⊗ Uik)Vj ik = Σij⊗ik(Uji ⊗ 1)Vj Wj (cid:101)Vj ik(U∗ ik ⊗ U∗ ki = (Uik ⊗ Uij)Wj kj). ik)∗ = ((cid:98)Jki ⊗ Jkj)Wj lj(Jij ⊗(cid:98)Jjl) jl)∗ = (Jil ⊗(cid:98)Jlj)Vi and (Wl (Vi Furthermore, we also have: ik(⊗U∗ ik)Σki⊗jk; ki((cid:98)Jik ⊗ Jij). Moreover, these unitaries satisfy the following pentagonal equations: (Vi jk)12(Vi kl)13(Vj kl)23(Vi ((cid:101)Vk ji )12((cid:101)Vl kl)23 = (Vj jl)12; ji)13((cid:101)Vl (Wk ij)12(Wl kj)23 = ((cid:101)Vl ij)13(Wl ki)23((cid:101)Vk jk)23 = (Wl ji )12. ik)23(Wk ij)12; ki)23 = ((cid:101)Vj ki)12((cid:101)Vj kj)∗, kj(x ⊗ 1)(Vi ki)23(Vl ki)12. for all x ∈ Mij. (pikωpik ⊗ id)(Wj ki) = pji(id ⊗ ω)((cid:101)V)pjk. ik) = pij(ω ⊗ id)(W)pkj; 16 J. CRESPO We also have the following commutation relations: (Vl ll(cid:48) )12(Vl(cid:48) ll(cid:48) )12 = (Wk kj)23; kj)23(Wj Furthermore, we have (Vl ij(x) = (Wj ∆k ik)∗(1 ⊗ x)Wj Note that for all ω ∈ B(H)∗ we have: jl) = pij(id ⊗ ω)(V)pil; (pkiωpki ⊗ id)((cid:101)Vj (id ⊗ pjlωpjl)(Vi ik = Vi 3.3.6 Proposition. -- Let i, j = 1, 2 such that i (cid:54)= j. With the notations of 3.3.3, we have: 1. Gi := (Mii, ∆i ii, ϕii, ψii) is a locally compact quantum group whose left (resp. right ) regular representation is Wi ii (resp. Vi ii); ij) is a right Galois object for Gj whose canonical implementation is Vi jj; ij) is a left Galois object for Gi whose canonical implementation is Wj ii; 2. (Mij, ∆j 3. (Mij, ∆i 4. the actions ∆j 5. the Galois isometry associated with the right Galois object (Mij, ∆j ij on Mij commute; ij and ∆i ij)∗Σij⊗ij. [12]) is the unitary Σij⊗jj(Wj ij) for Gj (cf. 6.4.1, 6.4.2 (cid:78) 3.3.7 Definition. -- A measured quantum groupoid (C2, M, α, β, ∆, T, T(cid:48), ) such that the source and target maps have range in Z (M) and generate a copy of C4 will be denoted by GG1,G2, where Gi = (Mii, ∆i ii, ϕii, ψii) (cf. 3.3.6) and will be called a colinking measured (cid:78) quantum groupoid. 3.3.8 Definition. -- Let G and H be two locally compact quantum groups. We say that G and H are monoidally equivalent if there exists a colinking measured quantum groupoid GG1,G2 between two locally compact quantum groups G1 and G2 such that H (resp. G) is (cid:78) isomorphic to G1 (resp. G2). Let (S, δ) be the weak Hopf C*-algebra associated with G. Note that for all i, j = 1, 2. pij = α(εi)β(εj) ∈ Z (M(S)), 3.3.9 Notations. -- Let us recall the notations below (cf. 2.26 [2]). 1. Let Sij := pijS, for i, j = 1, 2. Then, Sij is a C*-algebra (actually a closed two-sided ideal) of S weakly dense in Mij. 2. For i, j, k = 1, 2, let ιk extension of the inclusion map Sik ⊗ Skj ⊂ S ⊗ S satisfying ιk ij : M(Sik ⊗ Skj) → M(S ⊗ S) be the unique strictly continuous ij(1Sik⊗Skj ) = pik ⊗ pkj. 3. Let δk ij : Sij → M(Sik ⊗ Skj) be the unique *-homomorphism such that ij ◦ δk ιk With these notations, we have: 3.3.10 Proposition. -- (cf. 7.4.13, 7.4.14 [12], 2.27 [2]) Let i, j, k, l = 1, 2. ij(x) = (pik ⊗ pkj)δ(x), for all x ∈ Sij. 1. (δl 2. δk 3. [δk ik ⊗ idSkj )δk ij = (idSil ⊗ δk lj)δl ij. ik)∗(1Hik ⊗ x)Wj ij(x) = (Wj ik = Vi ij(Sij)(1Sik ⊗ Skj)] = Sik ⊗ Skj = [δk Skj = [(idSik ⊗ ω)δk kj)∗, for all x ∈ Sij. kj(x ⊗ 1Hkj )(Vi ij(Sij)(Sik ⊗ 1Skj )]. In particular, we have ij(x) ; x ∈ Sij, ω ∈ B(Hkj)∗]. 4. The pair (Sjj, δ j jj) is the Hopf C*-algebra associated with Gj. (cid:78) (cid:78) ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS 17 4 contributions to the notions of semi-regularity and regularity The notion of regular measured quantum groupoid has been introduced in [16] and stu- died in the compact case. Note that this notion has been generalized in the setting of pseudo-multiplicative unitaries, cf. [23, 24]. The notion of semi-regular measured quantum groupoid has been introduced in [2, 9], where the notions of regularity and semi-regularity have been studied in the case of a finite-dimensional basis. In this chapter, we fix a measured quantum groupoid G = (N, M, α, β, ∆, T, T(cid:48), ) on the 1(cid:54)l(cid:54)k Mnl (C) and we use all the notations introduced in §§3.1, 3.2. In the appendix (cf. 8.2.1), for any ξ ∈ H we have given the definition of the operator finite-dimensional basis N =(cid:76) ξ ∈ B(H, H) Rα (resp. Lβ ξ ∈ B(H, H)) and the definition of the weakly dense ideal of α(N)(cid:48) (resp. β(No)(cid:48)) Kα := [Rα η)∗ ; ξ, η ∈ H] ξ (Rα (resp. Kβ := [Lβ ξ (Lβ η )∗ ; ξ, η ∈ H]). Note that Kα and Kβ are C*-subalgebras of K := K(H). We first recall the following important consequence of the irreducibility (cf. 2.13 [2]) of G. 4.1 Proposition. -- (cf. 2.15 [2]) The Banach spaces [S(cid:98)S] and C(V) (cf. 2.1.2) are C*-algebras and we have [S(cid:98)S] = UC(V)U∗. (cid:78) 4.2 Definition. -- (cf. 4.7 [16], 2.37 [2]) The groupoid G is said to be semi-regular (resp. regular) if we have Kβ ⊂ C(V) (resp. Kβ = C(V)). (cid:78) 4.3 Proposition. -- (cf. 2.8 [2], 3.2.8 [9]) The following statements are equivalent: (i) G is semi-regular (resp. regular), i.e. Kα ⊂ C(W) (resp. Kα = C(W)); (ii) (cid:98)G is semi-regular (resp. regular), i.e. Kβ ⊂ C(V) (resp. Kβ = C(V)); (iii) (Go)c is semi-regular (resp. regular), i.e. K(cid:98)α ⊂ C((cid:101)V) (resp. K(cid:98)α = C((cid:101)V)). 4.4 Proposition. -- (cf. 2.8 [2], 3.2.9 [9]) The following statements are equivalent: (cid:78) ⊂ [(cid:98)SS] (resp. K(cid:98)β (i) G is semi-regular (resp. regular); = [(cid:98)SS]); (ii) K(cid:98)β (iii) Kα ⊂ [R(S)(cid:98)S] (resp. Kα = [R(S)(cid:98)S]); (iv) K(cid:98)α ⊂ [Sλ((cid:98)S)] (resp. K(cid:98)α = [Sλ((cid:98)S)]). In particular, if G is regular we have [(cid:98)SS] ⊂ K, [R(S)(cid:98)S] ⊂ K and [Sλ((cid:98)S)] ⊂ K (and also C(V) ⊂ K, C(W) ⊂ K and C((cid:101)V) ⊂ K). (cid:78) The semi-regularity and the regularity of colinking measured quantum groupoids have been treated in detail in §2.5 [2]. 4.5 Theorem. -- (cf. 2.45 [2]) Let GG1,G2 be a colinking measured quantum groupoid associated with two monoidally equivalent locally compact quantum groups G1 and G2. The groupoid GG1,G2 (cid:78) is semi-regular (resp. regular) if, and only if, G1 and G2 are semi-regular (resp. regular). In the following, we use the multi-index notation introduced in the appendix §8 of this article (cf. 8.2.21, 8.2.22 and 8.2.23) with γ := α and π := β. 4.6 Lemma. -- For all ξ, η ∈ H, we have · eI θξ,ηeI · fI θξ,η fI. and Lβ Rα ξ (Rα n−1 I n−1 I (cid:78) η)∗ = ∑ I∈I ξ (Lβ η )∗ = ∑ I∈I 18 J. CRESPO Proof. For η, ζ ∈ H, let Xη,ζ ∈ N be defined by Xη,ζ := ∑I∈I n−1 x ∈ N and η ∈ H, we have Rα (cid:104)(Rα I, J ∈ I. On the other hand, we have I (cid:104)eI η, ζ(cid:105) · ε I. For all Λ(x) = α(x)η = ∑I∈I xI · eI η. Let η, ζ ∈ H and I ∈ I. J I nI for all η)∗ζ, Λ(ε I)(cid:105) = (cid:104)ζ, eI η(cid:105). By disjunction elimination, we prove that (ε JI) = δ η (cid:33) (cid:32) (cid:104)Λ(Xη,ζ ), Λ(ε I)(cid:105) = (X∗ η,ζ ε I) n−1 J (cid:104)eJη, ζ(cid:105) · ε JI ∑ J∈I n−1 J (cid:104)ζ, eJη(cid:105)(ε JI) = ∑ J∈I = (cid:104)ζ, eI η(cid:105). =  Hence, (cid:104)(Rα Let ξ, η ∈ H. For all ζ ∈ H, we have η)∗ζ, Λ(ε I)(cid:105) = (cid:104)Λ(Xη,ζ ), Λ(ε I)(cid:105). Hence, (Rα η)∗ζ = Λ(Xη,ζ ) for all η, ζ ∈ H. η)∗ ξ (Rα Rα ζ = Rα ξ (Xη,ζ )I · eI ξ = ∑ I∈I Λ(Xη,ζ ) = ∑ I∈I · eI θξ,ηeI. The second formula is proved in a similar way. n−1 I (cid:104)η, eI ζ(cid:105) · eI ξ = ∑ I∈I n−1 I θeI ξ, eI η(ζ). η)∗ = ∑I∈I n−1 I Hence, Rα ξ (Rα We refer to 8.2.15 and 8.2.18 for the definition of the operators qα, qβ and q(cid:98)α. The proposi- tions 4.7, 4.8 and 4.9 below have to be compared with their corresponding statements in the quantum group case, cf. 3.2 b), 3.6 b) and 3.6 d) [4]. 4.7 Proposition. -- The following statements are equivalent: (i) G is regular (resp. semi-regular); (ii) [(K ⊗ 1)W(1 ⊗ K)] = [(K ⊗ 1)qα(1 ⊗ K)] (resp. ⊃ [(K ⊗ 1)qα(1 ⊗ K)]); (iii) [(K ⊗ 1)V(1 ⊗ K)] = [(K ⊗ 1)qβ(1 ⊗ K)] (resp. ⊃ [(K ⊗ 1)qβ(1 ⊗ K)]); (iv) [(K ⊗ 1)(cid:101)V(1 ⊗ K)] = [(K ⊗ 1)q(cid:98)α(1 ⊗ K)] (resp. ⊃ [(K ⊗ 1)q(cid:98)α(1 ⊗ K)]). Proof. It is known that G is regular (resp. semi-regular) if, and only if, (cid:98)G is regular (resp. (cid:78) semi-regular). Therefore, it suffices to prove that (i) is equivalent to (ii). We have [(K ⊗ 1)W(1 ⊗ K)] = Σ(C(W) ⊗ K) (cf. 3.1 [4]) (C(W) is a C∗-algebra regardless of the regularity of G). Note that G is regular (resp. semi- regular) if, and only if, Σ(C(W) ⊗ K) = Σ(Kα ⊗ K) (resp. Σ(C(W) ⊗ K) ⊃ Σ(Kα ⊗ K)). Let ξ, η, ζ, χ ∈ H. We have eI θξ, ηeI ⊗ θζ, χ = θeI ξ, eI η ⊗ θζ, χ = θeI ξ⊗ζ, eI η⊗χ, for all I ∈ I. Hence, Σ(eI θξ, ηeI ⊗ θζ, χ) = θζ⊗eI ξ, eI η⊗χ = θζ, eI η ⊗ θeI ξ, χ = θζ, ηeI ⊗ eI θξ, χ for all I ∈ I. By Lemma 4.6, we obtain Σ(Rα ξ (Rα η)∗ ⊗ θζ,χ) = (θζ,η ⊗ 1)qα(1 ⊗ θξ,χ). Hence, Σ(Kα ⊗ K) = [(K ⊗ 1)qα(1 ⊗ K)] and the equivalence ((i) ⇔ (ii)) is proved. 4.8 Proposition. -- If G is regular (resp. semi-regular), we have: 1. [(S ⊗ 1)W(1 ⊗ K)] = [(S ⊗ 1)qα(1 ⊗ K)] (resp. ⊃ [(S ⊗ 1)qα(1 ⊗ K)]); 2. [(K ⊗ 1)V(1 ⊗ S)] = [(K ⊗ 1)qβ(1 ⊗ S)] (resp. ⊃ [(K ⊗ 1)qβ(1 ⊗ S)]); 3. [(R(S) ⊗ 1)(cid:101)V(1 ⊗ K)] = [(R(S) ⊗ 1)q(cid:98)α(1 ⊗ K)] (resp. ⊃ [(R(S) ⊗ 1)q(cid:98)α(1 ⊗ K)]). (cid:78) ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS 19 Proof. Assume that G is regular (resp. semi-regular). Let us prove the first statement. The others will be obtained by using similar arguments. Let a, b ∈ K, ω ∈ B(H)∗ and y = (id ⊗ aω)(W). We have (y ⊗ 1)W(1 ⊗ b) = (id ⊗ ω ⊗ id)(W12W13(1 ⊗ a ⊗ b)) = (id ⊗ ω ⊗ id)(W23W12(1 ⊗ W∗(a ⊗ b))). However, W∗(K ⊗ K) = qβα(K ⊗ K). Moreover, since [W, 1 ⊗ β(no)] = 0 for all n ∈ N, we have [W12, qβα,23] = 0. Hence, W23W12qβα,23 = W23W12. We obtain (cf. 4.7) [(S ⊗ 1)W(1 ⊗ K)] = [(id ⊗ aω ⊗ id)(W23W12(1 ⊗ 1 ⊗ b)) ; ω ∈ B(H)∗, a, , b ∈ K] = [(id ⊗ ωa ⊗ id)(W23W12(1 ⊗ 1 ⊗ b)) ; ω ∈ B(H)∗, a, b ∈ K] = [(id ⊗ ω ⊗ id)(((a ⊗ 1)W(1 ⊗ b))23W12) ; ω ∈ B(H)∗, a, b ∈ K] = [(id ⊗ ω ⊗ id)(((a ⊗ 1)qα(1 ⊗ b))23W12) ; ω ∈ B(H)∗, a, b ∈ K] (resp. ⊃ [(id ⊗ ω ⊗ id)(((a ⊗ 1)qα(1 ⊗ b))23W12) ; ω ∈ B(H)∗, a, b ∈ K]). However, for all ω ∈ B(H)∗ and a, b ∈ K we have (id ⊗ ω ⊗ id)(((a ⊗ 1)qα(1 ⊗ b))23W12) = (id ⊗ ωa ⊗ id)(qα,23W12(1 ⊗ 1 ⊗ b)). Since (1 ⊗ α(n))W = W(α(n) ⊗ 1) for all n ∈ N, we have qα,23W12 = W12qα,13. Hence, (id ⊗ ω ⊗ id)(((a ⊗ 1)qα(1 ⊗ b))23W12) = ((id ⊗ ωa)(W) ⊗ 1)qα(1 ⊗ b) and the result is proved. 4.9 Proposition. -- If G is regular (resp. semi-regular), then we have: 1. [(S ⊗ 1)W(1 ⊗ λ((cid:98)S))] = [(S ⊗ 1)qα(1 ⊗ λ((cid:98)S))] (resp. ⊃ [(S ⊗ 1)qα(1 ⊗ λ((cid:98)S))]); 2. [((cid:98)S ⊗ 1)V(1 ⊗ S)] = [((cid:98)S ⊗ 1)qβ(1 ⊗ S)] (resp. ⊃ [((cid:98)S ⊗ 1)qβ(1 ⊗ S)]); 3. [(R(S) ⊗ 1)(cid:101)V(1 ⊗(cid:98)S)] = [(R(S) ⊗ 1)q(cid:98)α(1 ⊗(cid:98)S)] (resp. ⊃ [(R(S) ⊗ 1)q(cid:98)α(1 ⊗(cid:98)S)]). In particular, we have [(S ⊗ 1)W(1 ⊗ λ((cid:98)S))] ⊂ S ⊗ λ((cid:98)S), [((cid:98)S ⊗ 1)V(1 ⊗ S)] ⊂ (cid:98)S ⊗ S and [(R(S) ⊗ 1)(cid:101)V(1 ⊗(cid:98)S)] ⊂ R(S) ⊗(cid:98)S. 23. Since V ∈ M((cid:98)S ⊗ S) is a partial isometry, we have V∗((cid:98)S ⊗ S) = q(cid:98)αβ((cid:98)S ⊗ S). Since [V, 1 ⊗(cid:98)α(n)] = 0 for all n ∈ N, Proof. We have the pentagonal equation V12V13 = V23V12V∗ we have [V12, q(cid:98)αβ,23] = 0. Hence, V23V12q(cid:98)αβ,23 = V23V12. Hence, [((cid:98)S ⊗ 1)V(1 ⊗ S)] = [((id ⊗ ω)(V) ⊗ 1)V(1 ⊗ y) ; ω ∈ B(H)∗, y ∈ S] 23(1 ⊗ x ⊗ y)) ; ω ∈ B(H)∗, y ∈ S, x ∈ (cid:98)S] = [(id ⊗ ω ⊗ id)(V12V13(1 ⊗ 1 ⊗ y)) ; ω ∈ B(H)∗, y ∈ S] = [(id ⊗ ω ⊗ id)(V23V12V∗ = [(id ⊗ ω ⊗ id)(V23V12(1 ⊗ x ⊗ y)) ; ω ∈ B(H)∗, x ∈ (cid:98)S, y ∈ S] = [(id ⊗ ω ⊗ id)(V23(1 ⊗ 1 ⊗ y)V12) ; ω ∈ B(H)∗, y ∈ S] = [(id ⊗ ω ⊗ id)(((a ⊗ 1)V(1 ⊗ y))23V12) ; ω ∈ B(H)∗, a ∈ K, y ∈ S]. Let X := [(id ⊗ ω ⊗ id)(((a ⊗ 1)qβ(1 ⊗ y))23V12) ; ω ∈ B(H)∗, a ∈ K, y ∈ S]. Since G is [((cid:98)S ⊗ 1)V(1 ⊗ S)] = X (resp. [((cid:98)S ⊗ 1)V(1 ⊗ S)] ⊃ X). regular (resp. semi-regular), it follows from 4.8 that (cid:78) However, since (1 ⊗ β(no))V = V(β(no) ⊗ 1) for all n ∈ N, we have X = [(id ⊗ ωa ⊗ id)(qβ,23V12(1 ⊗ 1 ⊗ y)) ; ω ∈ B(H)∗, a ∈ K, y ∈ S] = [(id ⊗ ω ⊗ id)(V12qβ,13(1 ⊗ 1 ⊗ y)) ; ω ∈ B(H)∗, y ∈ S] = [((id ⊗ ω)(V) ⊗ 1)qβ(1 ⊗ y) ; ω ∈ B(H)∗, y ∈ S] = [((cid:98)S ⊗ 1)qβ(1 ⊗ S)]. 20 J. CRESPO The second statement is proved and the third one follows by applying it to (cid:98)G. We obtain the first statement by combining the third one with the formulas W = (U∗ ⊗ U∗)(cid:101)V(U ⊗ U) and(cid:98)α = AdU ◦ α. Finally, the last statement follows from the inclusions β(No) ⊂ M(S), (cid:98)β(No) ⊂ M((cid:98)S), α(N) ⊂ M(S) and(cid:98)α(N) ⊂ M((cid:98)S). In the result below, we refer again to 8.2.15 and 8.2.18 for the definition of the operators qβ(cid:98)β, q(cid:98)αα and q(cid:98)ββ. 4.10 Corollary. -- If G is regular (resp. semi-regular), then we have: (S ⊗ 1)] (resp. ⊃ [(1 ⊗ λ((cid:98)S))qβ(cid:98)β 1. [(1 ⊗ λ((cid:98)S))W(S ⊗ 1)] = [(1 ⊗ λ((cid:98)S))qβ(cid:98)β 2. [(1 ⊗ S)V((cid:98)S ⊗ 1)] = [(1 ⊗ S)q(cid:98)αα((cid:98)S ⊗ 1)] (resp. ⊃ [(1 ⊗ S)q(cid:98)αα((cid:98)S ⊗ 1)]); (R(S) ⊗ 1)] (resp. ⊃ [(1 ⊗(cid:98)S)q(cid:98)ββ 3. [(1 ⊗(cid:98)S)(cid:101)V(R(S) ⊗ 1)] = [(1 ⊗(cid:98)S)q(cid:98)ββ If G is regular, then we have [(1 ⊗ λ((cid:98)S))W(S ⊗ 1)] ⊂ S ⊗ λ((cid:98)S), [(1 ⊗ S)V((cid:98)S ⊗ 1)] ⊂ (cid:98)S ⊗ S and [(1 ⊗(cid:98)S)(cid:101)V(R(S) ⊗ 1)] ⊂ R(S) ⊗(cid:98)S. Proof. This is a direct consequence of Proposition 4.9 and the formulas (cid:98)β = AdU ◦ β, (cid:98)α = AdU ◦ α, W = Σ(U ⊗ 1)V(U∗ ⊗ 1)Σ and (cid:101)V = Σ(1 ⊗ U)V(1 ⊗ U∗)Σ. The second statement follows from the inclusions β(No) ⊂ M(S),(cid:98)β(No) ⊂ M((cid:98)S), α(N) ⊂ M(S) and (cid:98)α(N) ⊂ M((cid:98)S). (R(S) ⊗ 1)]). (S ⊗ 1)]); (cid:78) 5 measured quantum groupoids on a finite basis in action finite-dimensional basis N =(cid:76) 5.1 Continuous actions, crossed product and biduality In this section, we fix a measured quantum groupoid G = (N, M, α, β, ∆, T, T(cid:48), ) on the 1(cid:54)l(cid:54)k Mnl (C) and we use all the notations introduced in §§ 3.1, 3.2. In the following, we recall the definitions, notations and results of §§ 3.1, 3.2.1, 3.2.2 and 3.3.1 [2] (see also [9] chapter 4). 5.1.1 Notion of actions of measured quantum groupoids on a finite basis 5.1.1 Lemma. -- Let A and B be two C*-algebras, f e ∈ M(B). The following statements are equivalent: : A → M(B) a *-homomorphism and (i) there exists an approximate unit (uλ)λ of A such that f (uλ) → e with respect to the strict topology; (ii) f extends to a strictly continuous *-homomorphism f : M(A) → M(B), necessarily unique, such that f (1A) = e; (iii) [ f (A)B] = eB. In that case, e is a self-adjoint projection, for all approximate unit (vµ)µ of A we have f (vµ) → e (cid:78) with respect to the strict topology and [B f (A)] = Be. 5.1.2 Definition. -- An action of G on a C*-algebra A is a pair (βA, δA) consisting of a non-degenerate *-homomorphism βA : No → M(A) and a faithful *-homomorphism δA : A → M(A ⊗ S) such that: 1. δA extends to a strictly continuous *-homomorphism δA : M(A) → M(A ⊗ S) such that δA(1A) = qβAα (cf. 8.2.19); 2. (δA ⊗ idS)δA = (idA ⊗ δ)δA; 3. δA(βA(no)) = qβAα(1A ⊗ β(no)), for all n ∈ N. ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS 21 We say that the action (βA, δA) is strongly continuous if we have [δA(A)(1A ⊗ S)] = qβAα(A ⊗ S). If that case, we say that the triple (A, βA, δA) is a G-C*-algebra. 5.1.3 Remarks. -- (cid:78) • By 5.1.1, the condition 1 is equivalent to requiring that for some (and then any) approximate unit (uλ) of A, we have δA(uλ) → qβAα with respect to the strict topology of M(A ⊗ S). It is also equivalent to [δA(A)(A ⊗ S)] = qβAα(A ⊗ S). • Condition 1 implies that the *-homomorphisms δA ⊗ idS and idA ⊗ δ extend uniquely to strictly continuous *-homomorphisms from M(A ⊗ S) to M(A ⊗ S ⊗ S) such that (δA ⊗ idS)(1A⊗S) = qβAα,12 and (idA ⊗ δ)(1A⊗S) = qβα,23. In particular, condition 2 A := (δA ⊗ idS)δA : A → M(A ⊗ S ⊗ S) the does make sense and we denote by δ2 (cid:78) iterated coaction map. 5.1.4 Examples. -- Let us give two basic examples. • (S, β, δ) is a G-C*-algebra. • Let βNo := idNo. Let δNo : No → M(No ⊗ S) be the faithful unital *-homomorphism given by δNo (no) := qβNo α(1No ⊗ β(no)) for all n ∈ N. Then, the pair (βNo, δNo ) is an action of G on No called the trivial action. (cid:78) 5.1.5 Proposition. -- Let (δA, βA) be an action of G on A. We have the following statements: A extends uniquely to a strictly continuous *-homomorphism A(1A) = qβAα,12qβα,23; moreover, we have A : M(A) → M(A ⊗ S ⊗ S) such that δ2 δ2 A(m) = (δA ⊗ idS)δA(m) = (idA ⊗ δ)δA(m) for all m ∈ M(A); δ2 1. the iterated coaction map δ2 (cid:78) Let us provide a more explicit definition of what an action of the dual measured quantum 2. for all n ∈ N, we have δA(βA(no)) = (1A ⊗ β(no))qβAα; 3. if (βA, δA) is strongly continuous, then we have [(1A ⊗ S)δA(A)] = (A ⊗ S)qβAα. groupoid (cid:98)G on a C*-algebra B is. 5.1.6 Definition. -- An action of (cid:98)G on a C*-algebra B is a pair (αB, δB) consisting of δB : B → M(B ⊗(cid:98)S) such that: a non-degenerate *-homomorphism αB : N → M(B) and a faithful *-homomorphism 1. δB extends to a strictly continuous *-homomorphism δB : M(B) → M(B ⊗(cid:98)S) such 2. (δB ⊗ id(cid:98)S)δB = (idB ⊗(cid:98)δ)δB; 3. δB(αB(n)) = qαB β(1B ⊗(cid:98)α(n)), for all n ∈ N. that δB(1B) = qαB β (cf. 8.2.19); We say that the action (αB, δB) is strongly continuous if we have [δB(B)(1B ⊗(cid:98)S)] = qαB β(B ⊗(cid:98)S). If (δB, αB) is a strongly continuous action of (cid:98)G on B, we say that the triple (B, αB, δB) is a (cid:98)G-C*-algebra. 5.1.7 Remarks. -- As for actions of G, we have: (cid:78) • the condition 1 is equivalent to requiring that for some (and then any) approximate also equivalent to the relation [δB(B)(B ⊗(cid:98)S)] = qαB β(B ⊗(cid:98)S); unit (uλ)λ of B we have δB(uλ) → qαB β with respect to the strict topology, which is • the *-homomorphisms idB ⊗(cid:98)δ and δB ⊗ id(cid:98)S extend uniquely to strictly continuous *- homomorphisms from M(B⊗(cid:98)S) to M(B⊗(cid:98)S⊗(cid:98)S) such that (idB ⊗(cid:98)δ)(1B⊗(cid:98)S) = q(cid:98)αβ,23 and (δB ⊗ id(cid:98)S)(1B⊗(cid:98)S) = qαB β,12. In particular, condition 2 does make sense and we (cid:78) B := (δB ⊗ id(cid:98)S)δB : B → M(B ⊗(cid:98)S ⊗(cid:98)S) the iterated coaction map. denote by δ2 22 J. CRESPO (cid:78) δ2 δ2 1. the iterated coaction map δ2 5.1.8 Examples. -- Let us give two basic examples: • ((cid:98)S,(cid:98)α,(cid:98)δ) is a (cid:98)G-C*-algebra; • Let αN := idN and δN : N → M(N ⊗(cid:98)S) ; n (cid:55)→ qαN β(1N ⊗(cid:98)α(n)); then, the pair (αN, δN) is an action of (cid:98)G on N called the trivial action. 5.1.9 Proposition. -- Let (αB, δB) be an action of (cid:98)G on B. We have the following statements: B : M(B) → M(B ⊗(cid:98)S ⊗(cid:98)S) such that δ2 B extends uniquely to a strictly continuous *-homomorphism B(1B) = qαB β,12q(cid:98)αβ,23; moreover, we have B(m) = (δB ⊗ id(cid:98)S)δB(m) = (idB ⊗(cid:98)δ)δB(m) for all m ∈ M(B); 2. for all n ∈ N, we have δB(αB(n)) = (1B ⊗(cid:98)α(n))qαB β; 3. if (αB, δB) is strongly continuous, then we have [(1B ⊗(cid:98)S)δB(B)] = (B ⊗(cid:98)S)qαB β. (cid:78) (βAi, δAi ) (resp. (αBi, δBi )) be an action of G (resp. (cid:98)G) on Ai (resp. Bi). A non-degenerate *- 5.1.10 Definition. -- For i = 1, 2, let Ai (resp. Bi) be a C*-algebra. For i = 1, 2, let (cid:98)G-equivariant) if ( f ⊗ idS)δA1 = δA2 ◦ f and f ◦ βA1 = βA2 (resp. ( f ⊗ id(cid:98)S)δB1 = δB2 ◦ f homomorphism f : A1 → M(A2) (resp. f : B1 → M(B2)) is said to be G-equivariant (resp. and f ◦ αB1 = αB2). (cid:78) 5.1.11 Remark. -- With the notations and hypotheses of 5.1.10, if f satisfies the relation ( f ⊗ idS)δA1 = δA2 ◦ f (resp. ( f ⊗ id(cid:98)S)δB1 = δB2 ◦ f ), then f satisfies necessarily the relation f ◦ βA1 = βA2 (resp. f ◦ αB1 = αB2), i.e. f is G-equivariant (resp. (cid:98)G-equivariant). Indeed, let n ∈ N. For all a ∈ A1 and x ∈ A2, we have δA2 ( f (βA1 (no)) f (a)x) = ( f ⊗ idS)δA1 (βA1 (no)a)δA2 (x) = (1A2 ⊗ β(no))( f ⊗ idS)δA1 (a)δA2 (x) = (1A2 ⊗ β(no))δA2 ( f (a)x) = δA2 (βA2 (no) f (a)x). Hence, f (βA1 (no)) f (a)x = βA2 (no) f (a)x for all a ∈ A1 and x ∈ A2 since δA2 is faithful. (cid:78) Hence, we have f (βA1 (no)) = βA2 (no) since f is non-degenerate. 5.1.12 Notation. -- We denote by AlgG the category whose objects are the G-C*-algebras and whose set of arrows between G-C*-algebras is the set of G-equivariant non-degenerate (cid:78) *-homomorphisms. 5.1.2 Crossed product and dual action Let us fix a strongly continuous action (βA, δA) of G on a C*-algebra A. 5.1.13 Notations. -- The *-representation πL := (idA ⊗ L) ◦ δA : A → L(A ⊗ H) of A on the Hilbert A-module A ⊗ H extends uniquely to a strictly/*-strongly continuous faithful *-representation πL : M(A) → L(A ⊗ H) such that πL(1A) = qβAα. Moreover, we have πL(m) = πL(m)qβAα = qβAαπL(m) for all m ∈ M(A). Consider the Hilbert A-module EA,L := qβAα(A ⊗ H). By restricting πL, we obtain a strictly/*-strongly continuous faithful unital *-representation continuous unital *-representation We have [1A ⊗ T, qβAα] = 0 for all T ∈ M((cid:98)S). We then obtain a strictly/*-strongly Note that if βA is faithful, then so is(cid:98)θ. π : M(A) → L(EA,L) ; m (cid:55)→ πL(m)(cid:22)EA,L. (cid:98)θ : M((cid:98)S) → L(EA,L) ; T (cid:55)→ (1A ⊗ T)(cid:22)EA,L. (cid:78) ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS 23 products of the form π(a)(cid:98)θ(x) for a ∈ A and x ∈ (cid:98)S is a C*-subalgebra called the (reduced) crossed 5.1.14 Proposition-Definition. -- The norm closed subspace of L(EA,L) spanned by the product of A by the strongly continuous action (βA, δA) of G and denoted by A (cid:111) G. (cid:78) In particular, the morphism π (resp.(cid:98)θ) defines a faithful unital *-homomorphism (resp. unital *-homomorphism) π : M(A) → M(A (cid:111) G) (resp.(cid:98)θ : (cid:98)S → M(A (cid:111) G)). Since [(cid:101)V, α(n) ⊗ 1] = 0, we have [(cid:101)V23, qβAα,12] = 0. The operator (cid:101)V23 ∈ L(A ⊗ H ⊗ H) restricts to a partial isometry X := (cid:101)V23(cid:22)EA,L⊗H ∈ L(EA,L ⊗ H), whose initial and final projections are X∗X = q(cid:98)βα,23 5.1.15 Proposition-Definition. -- Let (cid:22)EA,L⊗H and XX∗ = q(cid:98)αβ,23(cid:22)EA,L⊗H. δA(cid:111)G : A (cid:111) G → L(EA,L ⊗ H) and αA(cid:111)G : N → M(A (cid:111) G) be the linear maps defined by: Then, δA(cid:111)G is a faithful *-homomorphism and αA(cid:111)G is a non-degenerate *-homomorphism. Moreover, we have the following statements: • δA(cid:111)G (b) := X(b ⊗ 1)X∗, for all b ∈ A (cid:111) G; • αA(cid:111)G (n) :=(cid:98)θ((cid:98)α(n)) = (1A ⊗(cid:98)α(n))(cid:22)EA,L, for all n ∈ N. 1. δA(cid:111)G (π(a)(cid:98)θ(x)) = (π(a) ⊗ 1(cid:98)S)((cid:98)θ ⊗ id(cid:98)S)(cid:98)δ(x), for all a ∈ A and x ∈ (cid:98)S; in particular, δA(cid:111)G takes its values in M((A (cid:111) G) ⊗(cid:98)S); 2. αA(cid:111)G (n)π(a)(cid:98)θ(x) = π(a)(cid:98)θ((cid:98)α(n)x) and π(a)(cid:98)θ(x)αA(cid:111)G (n) = π(a)(cid:98)θ(x(cid:98)α(n)) for all n ∈ N, a ∈ A and x ∈ (cid:98)S. strongly continuous action of (cid:98)G on A (cid:111) G called the dual action of (βA, δA). action (αB, δB) of the dual measured quantum groupoid (cid:98)G. (cid:78) 5.1.16 Proposition-Definition. -- With the notations of 5.1.15, the pair (αA(cid:111)G, δA(cid:111)G ) is a (cid:78) In a similar way, we define the crossed product of a C*-algebra B by a strongly continuous 5.1.17 Notations. -- The *-representation of B on the Hilbert B-module B ⊗ H extends uniquely to a strictly/*-strongly continuous (cid:98)πλ := (idB ⊗ λ) ◦ δB : B → L(B ⊗ H) faithful *-representation (cid:98)πλ : M(B) → L(B ⊗ H) such that (cid:98)πλ(1B) = qαB(cid:98)β. Moreover, we have (cid:98)πλ(m) = (cid:98)πλ(m)qαB(cid:98)β = qαB(cid:98)β(cid:98)πλ(m), for all m ∈ M(B). Consider the Hilbert B-module By restricting (cid:98)πλ, we obtain a strictly/*-strongly continuous faithful unital *-representation We have [1B ⊗ T, qαB(cid:98)β (cid:98)π : M(B) → L(EB,λ) ; m (cid:55)→ (cid:98)πλ(m)(cid:22)EB,λ . ] = 0 for all T ∈ M(S). We then obtain a strictly/*-strongly EB,λ := qαB(cid:98)β (B ⊗ H). continuous unital *-representation θ : M(S) → L(EB,λ) ; T (cid:55)→ (1B ⊗ T)(cid:22)EB,λ . (cid:78) Note that if αB is faithful, then so is θ. products of the form (cid:98)π(b)θ(x) for b ∈ B and x ∈ S is a C*-subalgebra called the (reduced) 5.1.18 Proposition-Definition. -- The norm closed subspace of L(EB,λ) spanned by the crossed product of B by the strongly continuous action (αB, δB) of (cid:98)G and denoted by B (cid:111) (cid:98)G. (cid:78) In particular, the morphism (cid:98)π (resp. θ) defines a faithful unital *-homomorphism (resp. unital *-homomorphism) (cid:98)π : M(B) → M(B (cid:111) (cid:98)G) (resp. θ : S → M(B (cid:111) (cid:98)G)). 24 J. CRESPO Since [V, β(no) ⊗ 1] = 0, we have [V23, qαB β,12] = 0. The operator V23 ∈ L(B ⊗ H ⊗ H) restricts to a partial isometry Y := V23(cid:22)EB,λ⊗H ∈ L(EB,λ ⊗ H), be the linear maps defined by: and βB(cid:111)(cid:98)G : No → L(EB,λ) whose initial and final projections are Y∗Y = q(cid:98)αβ,23(cid:22)EB,λ⊗H and YY∗ = qβα,23(cid:22)EB,λ⊗H. 5.1.19 Proposition-Definition. -- Let δB(cid:111)(cid:98)G : B (cid:111) (cid:98)G → L(EB,λ ⊗ H) • δB(cid:111)(cid:98)G (c) := Y(c ⊗ 1H )Y∗, for all c ∈ B (cid:111) (cid:98)G; • βB(cid:111)(cid:98)G (no) := θ(β(no)) = (1B ⊗ β(no))(cid:22)EB,λ⊗H, for all n ∈ N. Then, δB(cid:111)(cid:98)G is a faithful *-homomorphism and βB(cid:111)(cid:98)G is a non-degenerate *-homomorphism. Moreover, 1. δB(cid:111)(cid:98)G ((cid:98)π(b)θ(x)) = ((cid:98)π(b) ⊗ 1S)(θ ⊗ idS)δ(x), for all b ∈ B and x ∈ S; in particular, δB(cid:111)(cid:98)G takes its values in M((B (cid:111) (cid:98)G) ⊗ S); 2. βB(cid:111)(cid:98)G (no)(cid:98)π(b)θ(x) = (cid:98)π(b)θ(β(no)x) and (cid:98)π(b)θ(x)βB(cid:111)(cid:98)G (no) = (cid:98)π(b)θ(xβ(no)) for all n ∈ N, b ∈ B and x ∈ S. (cid:78) 5.1.20 Proposition-Definition. -- With the notations of 5.1.19, the pair (βB(cid:111)(cid:98)G, δB(cid:111)(cid:98)G ) is a strongly continuous action of G on B (cid:111) (cid:98)G called the dual action of (αB, δB). (cid:78) we have the following statements: 5.1.3 Takesaki-Takai duality Let (βA, δA) be a strongly continuous action of the groupoid G on a C*-algebra A. 5.1.21 Notations. -- The *-representation πR := (idA ⊗ R) ◦ δA : A → L(A ⊗ H) We recall that the Banach space of A on the Hilbert A-module A ⊗ H extends uniquely to a strictly/*-strongly continuous faithful *-representation πR : M(A) → L(A ⊗ H) satisfying πR(m) = (idA ⊗ R)δA(m) for all m ∈ M(A) and πR(1A) = qβA(cid:98)α. Consider the Hilbert A-module D := [πR(a)(1A ⊗ λ(x)L(y)) ; a ∈ A, x ∈ (cid:98)S, y ∈ S] EA,R := qβA(cid:98)α(A ⊗ H). is a C*-subalgebra of L(A ⊗ H) such that dqβA(cid:98)α = d = dqβA(cid:98)α for all d ∈ D. Moreover, we have D(A ⊗ H) = EA,R. We also recall that there exists a unique strictly/*-strongly continuous faithful *-representation jD : M(D) → L(A ⊗ H) extending the inclusion map D ⊂ L(A ⊗ H) such that jD(1D) = qβA(cid:98)α. (cid:78) 5.1.22 Proposition. -- There exists a unique *-isomorphism φ : (A (cid:111) G) (cid:111) (cid:98)G → D such that φ((cid:98)π(π(a)(cid:98)θ(x))θ(y)) = πR(a)(1A ⊗ λ(x)L(y)) for all a ∈ A, x ∈ (cid:98)S and y ∈ S. (cid:78) 5.1.23 Notations. -- We denote K := K(H) for short. Let δ0 : A ⊗ K → M(A ⊗ K ⊗ S) be the *-homomorphism defined by δ0(a ⊗ k) = δA(a)13(1A ⊗ k ⊗ 1S) for all a ∈ A and k ∈ K. The morphism δ0 extends uniquely to a strictly continuous *-homomorphism δ0 : M(A ⊗ K) → M(A ⊗ K ⊗ S) such that δ0(1A⊗K) = qβAα,13. Let V ∈ L(H ⊗ S) be the unique partial isometry such that (idK ⊗ L)(V ) = V. (cid:78) C*-algebra D = [πR(a)(1A ⊗ λ(x)L(y)) ; a ∈ A, x ∈ (cid:98)S, y ∈ S] defined by the relations: 5.1.24 Theorem. -- There exists a unique strongly continuous action (βD, δD) of G on the (jD ⊗ idS)δD(d) = V23δ0(d)V∗ n ∈ N. Moreover, the canonical *-isomorphism φ : (A (cid:111) G) (cid:111) (cid:98)G → D (cf. 5.1.22) is G-equivariant. If the 23, groupoid G is regular, then we have D = qβA(cid:98)α(A ⊗ K)qβA(cid:98)α. The G-C*-algebra D defined above will be referred to as the bidual G-C*-algebra of A. jD(βD(no)) = qβA(cid:98)α(1A ⊗ β(no)), d ∈ D; (cid:78) ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS 25 5.2 Case of a colinking measured quantum groupoid In this section, we fix a colinking measured quantum groupoid G := GG1,G2 associated with two monoidally equivalent locally compact quantum groups G1 and G2. We follow all the notations recalled in §3.3 concerning the objects associated with G. In the following, we recall the notations and the main results of §3.2.3 [2] concerning the equivalent description of the G-C*-algebras in terms of G1-C*-algebras and G2-C*-algebras. Let us fix a G-C*-algebra (A, βA, δA). 5.2.1 Notations. -- • The morphism βA : C2 → M(A) is central. Let qj := βA(εj) for j = 1, 2. Then, qj is a central self-adjoint projection of M(A) and q1 + q2 = 1A. Let Aj := qjA for j = 1, 2. For j = 1, 2, Aj is a C*-subalgebra (actually a closed two-sided ideal) of A and we have A = A1 ⊕ A2. • For j, k = 1, 2, let πk extension of the inclusion Ak ⊗ Skj ⊂ A ⊗ S such that πk A,j and qA,j instead of πk j : M(Ak ⊗ Skj) → M(A ⊗ S) be the unique strictly continuous (cid:78) In case of ambiguity, we will denote πk 5.2.2 Proposition. -- For all j, k = 1, 2, there exists a unique faithful non-degenerate *-homo- morphism j (1Ak⊗Skj ) = qk ⊗ pkj. j and qj. Aj : Aj → M(Ak ⊗ Skj) δk such that for all x ∈ Aj, we have j ◦ δk πk Aj (x) = (qk ⊗ pkj)δA(x) = (qk ⊗ 1S)δA(x) = (1A ⊗ α(εk))δA(x) = (1A ⊗ pkj)δA(x). Moreover, we have: 1. δA(a) = ∑ k,j=1,2 j ◦ δk πk Aj (qja), for all a ∈ A; ⊗ idSkj )δk 2. (δl Ak Aj (Aj)(1Ak ⊗ Skj)] = Ak ⊗ Skj, for all j, k = 1, 2; in particular, we have 3. [δk Aj = (idAl ⊗ δk , for all j, k, l = 1, 2; lj)δl Aj 4. δ j Aj Ak = [(idAk ⊗ ω)δk Aj (a) ; a ∈ Aj, ω ∈ B(Hkj)∗]; : Aj → M(Aj ⊗ Sjj) is a strongly continuous action of Gj on Aj. (cid:78) From this concrete description of G-C*-algebras we can also give a convenient description of the G-equivariant *-homomorphisms. With the above notations, we have the result below. 5.2.3 Proposition. -- Let A and B be two G-C*-algebras. For k = 1, 2, let ιk : M(Bk) → M(B) be the unique strictly continuous extension of the inclusion map Bk ⊂ B such that ιk(1Bk ) = qB,k. : A → M(B) be a non-degenerate G-equivariant *-homomorphism. Then, for all j = 1, 2, there exists a unique non-degenerate *-homomorphism fj : Aj → M(Bj) such that for k = 1, 2 we have (5.1) 1. Let f ( fk ⊗ idSkj ) ◦ δk Aj = δk Bj Moreover, we have f (a) = ι1 ◦ f1(aqA,1) + ι2 ◦ f2(aqA,2) for all a ∈ A. 2. Conversely, for j = 1, 2 let fj : Aj → M(Bj) be a non-degenerate *-homomorphism such that (5.1 ) holds for all j, k = 1, 2. Then, the map f : A → M(B), defined for all a ∈ A by f (a) := ι1 ◦ f1(aqA,1) + ι2 ◦ f2(aqA,2), is a non-degenerate G-equivariant *-homomorphism. The above results show that for j = 1, 2 we have a functor (cid:78) ◦ fj. 26 J. CRESPO AlgG → AlgGj ; (A, βA, δA) (cid:55)→ (Aj, δ j Aj ). In §4 [2], it has been proved that if G is regular (cf. 4.5), then (A, δA, βA) → (A1, δ1 ) is A1 an equivalence of categories. Moreover, the authors build explicitly the inverse functor (A1, δA1 ) → (A, βA, δA). More precisely, to any G1-C*-algebra (A1, δA1 ) they associate a G2-C*-algebra (A2, δA2 ) in a canonical way. Then, the C*-algebra A := A1 ⊕ A2 can be equipped with a strongly continuous action (βA, δA) of the groupoid G. This allowed them to build the inverse functor (A1, δA1 ) → (A, βA, δA). The equivalence of categories (A1, δA1 ) → (A2, δA2 ) generalizes the correspondence of actions for monoidally equivalent compact quantum groups of De Rijdt and Vander Vennet [15]. We bring to the reader's attention that an induction procedure has been developed by De Commer in the von Neumann algebraic setting (cf. §8 [12]). In the following, we recall the notations and the main results of §4 [2]. We assume that the quantum groups G1 and G2 are regular. 5.2.4 Notations. -- Let δA1 : A1 → M(A1 ⊗ S11) be a continuous action of G1 on a C*-algebra A1. Let us denote: δ1 A1 := δA1, δ (2) A1 := (idA1 ⊗ δ2 11)δA1 : A1 → M(A1 ⊗ S12 ⊗ S21). (2) A1 Then, δ S21 with a C*-subalgebra of B(H21). Let is a faithful non-degenerate *-homomorphism. In the following, we will identify IndG2 G1 (A1) := [(idA1⊗S12 ⊗ ω)δ (2) A1 (a) ; a ∈ A1, ω ∈ B(H21)∗] ⊂ M(A1 ⊗ S12). (cid:78) 5.2.5 Proposition. -- The Banach space A2 := IndG2 G1 Moreover, we have: (A1) ⊂ M(A1 ⊗ S12) is a C*-algebra. 1. [A2(1A1 ⊗ S12)] = A1 ⊗ S12 = [(1A1 ⊗ S12)A2]; in particular, A2 ⊂ M(A1 ⊗ S12) defines a faithful non-degenerate *-homomorphism and M(A2) ⊂ M(A1 ⊗ S12); 2. let δA2 := (idA1 ⊗ δ2 12)(cid:22)A2, we have δA2 (A2) ⊂ M(A2 ⊗ S22) and δA2 is a continuous action of G2 on A2; 3. the correspondence IndG2 G1 : AlgG1 → AlgG2 is functorial. (cid:78) : AlgG2 → AlgG1. By exchanging the roles of the quantum groups G1 and G2, we obtain mutatis mutandis a functor IndG1 G2 5.2.6 Proposition. -- Let j, k = 1, 2 with j (cid:54)= k. Let (Aj, δAj ) be a Gj-C*-algebra. Let (Ak) ⊂ M(Ak ⊗ Skj) (Aj) ⊂ M(Aj ⊗ Sjk) and C := Ind Ak := IndGk Gj Gj Gk endowed with the continuous actions δAk := (idAj ⊗ δk Then, we have: jk)(cid:22)Ak and δC := (idAk ⊗ δ j kj)(cid:22)C respectively. 1. C ⊂ M(Ak ⊗ Skj) ⊂ M(Aj ⊗ Sjk ⊗ Skj) and C = δ (k) Aj (Aj); 2. πj : Aj → C ; a (cid:55)→ δ (k) Aj (a) := (idAj ⊗ δk jj)δAj (a) is a Gj-equivariant *-isomorphism; : Aj → M(Ak ⊗ Skj) ; a (cid:55)→ δ 3. δk Aj *-homomorphism. (k) Aj (a) := (idAj ⊗ δk jj)δAj (a) is a faithful non-degenerate (cid:78) The above result shows that the functors IndG2 G1 and IndG1 G2 are inverse of each other. ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS 27 5.2.7 Notations. -- Let (B1, δB1 ) be a G1-C*-algebra. Let (B2, δB2 ) be the induced G2-C*- 12)(cid:22)B2. In virtue of 5.2.6, we have algebra, that is to say B2 = IndG2 G1 four *-homomorphisms: (B1) and δB2 = (idB1 ⊗ δ2 Bj : Bj → M(Bk ⊗ Skj), δk j, k = 1, 2. Let us give a precise description of them. We denote δ1 B1 *-homomorphism δ2 B1 b ∈ B1 (cid:55)→ δ2 B1 whereas the *-homomorphism δ1 : B1 → M(B2 ⊗ S21) is given by (b) ∈ M(B2 ⊗ S21) (with δ (b) := δ (2) B1 (2) B1 (b) := (idB1 ⊗ δ2 11)δ1 B1 (b), for b ∈ B1) B2 : B2 → M(B1 ⊗ S12) is defined by the relation := δB1 and δ2 B2 := δB2. The ((cid:101)Aj, δ(cid:101)Aj (b) = δ (π1 ⊗ idS12 )δ1 B2 B2 and π1 : B1 → IndG1 (1) B2 for b ∈ B2, (b) (B2) ; b (cid:55)→ δ (1) B2 := (idB2 ⊗ δ1 (cid:78) where δ 5.2.8 Proposition. -- Let (A, βA, δA) be a G-C*-algebra. Let j, k = 1, 2 with j (cid:54)= k. With the notations of 5.2.2, let (b) (cf. 5.2.6 2). 22)δ2 (2) B1 G2 Gj Gk (Ak, δk Ak ). ) := Ind (x) ∈ (cid:101)Aj ⊂ M(Ak ⊗ Skj) and the map (cid:101)πj : Aj → (cid:101)Aj ; x (cid:55)→ δk If x ∈ Aj, then we have δk Aj a Gj-equivariant *-isomorphism. 5.2.9 Proposition. -- Let (B1, δB1 ) be a G1-C*-algebra. Let B2 = IndG2 (B1) be the induced G1 G2-C*-algebra. Let B := B1 ⊕ B2. For j, k = 1, 2 with j (cid:54)= k, let πk j : M(Bk ⊗ Skj) → M(B ⊗ S) be the strictly continuous *-homomorphism extending the canonical injection Bk ⊗ Skj → B ⊗ S Bj : Bj → M(Bk ⊗ Skj) the *-homomorphisms defined in 5.2.7. Let βB : C2 → M(B) and and δk δB : B → M(B ⊗ S) be the *-homomorphisms defined by: δB(b) := ∑ k,j=1,2 b = (b1, b2) ∈ B. (cid:18)λ 0 (λ, µ) ∈ C2; j ◦ δk πk Bj βB(λ, µ) := (x) is (cid:78) (cid:19) 0 µ (bj), Aj , Therefore, we have: 1. (βB, δB) is a strongly continuous action of G on B; 2. the correspondence AlgG1 → AlgG ; (B1, δB1 ) (cid:55)→ (B, βB, δB) is functorial; 3. the functors AlgG1 → AlgG and AlgG → AlgG1 are inverse of each other. (cid:78) 5.3 Actions of (semi-)regular measured quantum groupoids In this section, we fix a measured quantum groupoid G = (N, M, α, β, ∆, T, T(cid:48), ) on a 1(cid:54)l(cid:54)k Mnl (C) endowed with the non-normalized Markov finite-dimensional basis N :=(cid:76) trace  =(cid:76) 1(cid:54)l(cid:54)k nl · Trl and we use all the notations introduced in §3.1. We begin this section by a characterization of the regularity (resp. semi-regularity) of G in terms of the action of G on itself (cf. 5.1.4), which generalizes 2.6 [5] to the setting of measured quantum groupoids on a finite basis. (β, δ) of G. Then, we have a canonical *-isomorphism S (cid:111) G (cid:39) [S(cid:98)S]. In particular, G is regular 5.3.1 Proposition. -- Let S (cid:111) G be the crossed product of S by the strongly continuous action (resp. semi-regular) if, and only if, we have K(cid:98)β (cid:78) Proof. Let us identify L(ES,L) = {T ∈ L(S ⊗ H) ; Tqβα = T = qβαT}. Let us denote by for all u ∈ S (cid:111)G ⊂ L(ES,L). Let π : S → M(S (cid:111)G) and(cid:98)θ : (cid:98)S → M(S (cid:111)G) be the canonical jS(cid:111)G : S (cid:111) G → B(H ⊗ H), the faithful *-representation defined by jS(cid:111)G (u) = (L ⊗ idK)(u) = S (cid:111) G (resp. K(cid:98)β ⊂ S (cid:111) G). J. CRESPO 28 such that map for all s ∈ S and x ∈ (cid:98)S. morphisms (cf. 5.1.13). We claim that there exists a unique *-isomorphism φ : S (cid:111) G → [S(cid:98)S] φ(π(s)(cid:98)θ(x)) = L(s)ρ(x), Let s ∈ S and x ∈ (cid:98)S. Since W ∈ M ⊗ (cid:98)M and ρ((cid:98)S) ⊂ (cid:98)M(cid:48), we have jS(cid:111)G (π(s)(cid:98)θ(x)) = (L ⊗ L)δ(s)(1 ⊗ ρ(x)) = W∗(1 ⊗ L(s))W(1 ⊗ ρ(x)) = W∗(1 ⊗ L(s)ρ(x))W. Let C := im(jS(cid:111)G ) = {W∗(1 ⊗ z)W ; z ∈ [S(cid:98)S]}. The representation jS(cid:111)G induces a *- ] = 0 for all z ∈ [S(cid:98)S], the isomorphism ψ : S (cid:111) G → C. Since WW∗ = qα(cid:98)β and [1 ⊗ z, qα(cid:98)β χ : [S(cid:98)S] → C ; z → W∗(1 ⊗ z)W (1 ⊗ z) for all z ∈ [S(cid:98)S]. Let ω ∈ B(H)∗ for all z ∈ [S(cid:98)S]. is a *-homomorphism satisfying Wχ(z)W∗ = qα(cid:98)β such that ω ◦ α = . We have Hence, χ is a *-isomorphism. Hence, φ := χ−1 ◦ ψ : S (cid:111) G → [S(cid:98)S] ; π(s)(cid:98)θ(x) (cid:55)→ L(s)(cid:98)θ(x) is (ω ⊗ id)(Wχ(z)W∗) = z, a *-isomorphism. The second statement of the proposition follows from 4.4. Proposition 5.3.4 and Theorem 5.3.6 are the generalizations of 5.7 and 5.8 of [5] to measured quantum groupoids on a finite basis. 5.3.2 Notations. -- Let (βA, δA) be an action of G on a C*-algebra A. With the notations of 8.2.21 and 8.2.22, let eI := α(ε I) and qI := βA(ε I) for all I ∈ I. (cid:78) 5.3.3 Lemma. -- Let (βA, δA) be an action of G on a C*-algebra A. With the notations of 5.3.2, we have: 1. qβAα = ∑ I∈I n−1 I qI ⊗ eI; 2. (qI ⊗ 1S)δA(a) = (1A ⊗ eI)δA(a), for all a ∈ A and I ∈ I; 3. δA(a)(qI ⊗ 1S) = δA(a)(1A ⊗ eI), for all a ∈ A and I ∈ I. (cid:78) Proof. Statement 1 is just restatement of 8.2.18. By a straightforward computation, we verify that (qI ⊗ 1S)qβAα = (1A ⊗ eI)qβAα for all I ∈ I. Statement 2 then follows from the fact that δA(1A) = qβAα. The last statement follows from the second one by taking the adjoint. 5.3.4 Proposition. -- Let (βA, δA) be an action of G on a C∗-algebra A. If G is semi-regular, the Banach space [(idA ⊗ ω)δA(a) ; a ∈ A, ω ∈ B(H)∗] ⊂ M(A) is a C∗-algebra. (cid:78) Proof. Let us denote T := [(idA ⊗ ω)δA(a) ; a ∈ A, ω ∈ B(H)∗]. For all a ∈ A and ω ∈ B(H)∗, we have (idA ⊗ ω)(δA(a))∗ = (idA ⊗ ω)δA(a∗). Hence, T∗ ⊂ T. Let us prove that TT ⊂ T. Let ω, φ ∈ B(H)∗, a, b ∈ A and x, y ∈ K. We have (idA ⊗ yω)δA(a)(idA ⊗ φx)δA(b) = (idA ⊗ φ ⊗ ω)(δA(a)13(1A ⊗ x ⊗ y)δA(b)12). By 5.3.3 1,2, we have δA(a)13(1A ⊗ x ⊗ y)δA(b)12 = ∑ I∈I = ∑ I∈I = δA(a)13((x ⊗ 1)qα(1 ⊗ y))23δA(b)12. n−1 I δA(a)13(1A ⊗ x ⊗ eIy)(qI ⊗ 1 ⊗ 1)δA(b)12 n−1 I δA(a)13(1A ⊗ xeI ⊗ eIy)δA(b)12 ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS 29 It follows from 4.7 that (idA ⊗ yω)δA(a)(idA ⊗ φx)δA(b) is the norm limit of finite sums of elements of the form where x(cid:48), y(cid:48) ∈ K. By combining the following formulas; c := (idA ⊗ φ ⊗ ω)(δA(a)13((x(cid:48) ⊗ 1)W(1 ⊗ y(cid:48)))23δA(b)12) = (idA ⊗ φx(cid:48) ⊗ y(cid:48) 23δA(a)13W23; WW∗ = qα(cid:98)β; [δA(a)13, qα(cid:98)β,23] = 0 (since (cid:98)β(No) ⊂ M(cid:48)); ω)(δA(a)13W23δA(b)12), A(a) = W∗ δ2 A(a) = δ2 A(a). Hence, we have c = (idA ⊗ ψ)(δ2 we obtain δA(a)13W23 = W23δ2 A(a)13δA(b)12), where ψ := (φx(cid:48) ⊗ yω)W ∈ B(H ⊗ H)∗. Therefore, c is the norm limit of finite sums of ele- ments of the form (idA ⊗ φ(cid:48) ⊗ ω(cid:48))(δ2 A(a)13δA(b)12) = (id⊗ φ(cid:48))δA((idA ⊗ ω(cid:48))δA(a)b), where φ(cid:48), ω(cid:48) ∈ B(H)∗. Hence, (idA ⊗ yω)δA(a)(idA ⊗ φx)δA(b) ∈ T. 5.3.5 Definition. -- Let (βA, δA) be an action of G on a C∗-algebra A. We say that (βA, δA) is weakly continuous if we have A = [(idA ⊗ ω)δA(a) ; a ∈ A, ω ∈ B(H)∗]. (cid:78) Note that any strongly continuous action (βA, δA) of G on a C∗-algebra A is necessarily continuous in the weak sense. Indeed, if (βA, δA) is strongly continuous we have the inclusion δA(A)(1A ⊗ S) ⊂ A ⊗ S. Hence, [(idA ⊗ ω)δA(a) ; a ∈ A, ω ∈ B(H)∗] ⊂ A since S ⊂ B(H) is non-degenerate. Conversely, let ω ∈ B(H)∗ such that ω ◦ α = . We have (idA ⊗ ω)(qβAα) = 1A. By writing ω = yω(cid:48) for some ω(cid:48) ∈ B(H)∗ and y ∈ S, we obtain (idA ⊗ ω(cid:48))(qβAα(a ⊗ y)) = a for all a ∈ A. 5.3.6 Theorem. -- If the groupoid G is regular, then any weakly continuous action of G is strongly (cid:78) continuous. Proof. Let us fix an action (βA, δA) of G on a C∗-algebra A. Let us assume that (βA, δA) is weakly continuous. Since W ∈ M(S ⊗ K) is a partial isometry such that W∗W = qβα, we have (S ⊗ K)W = (S ⊗ K)qβα. We recall that δA(a)13W23 = W23δ2 A(a) for all a ∈ A A(a) for all a ∈ A. By combining the (cf. proof of 5.3.4). By 5.1.5 1, we have qβα,23δ2 assertions of the above discussion with 5.3.3 3 and 4.8 1, we have (A ⊗ S)qβAα = [((idA ⊗ ω)δA(a) ⊗ y)qβAα ; a ∈ A, y ∈ S, ω ∈ B(H)∗] A(a)) ; a ∈ A, x ∈ K, y ∈ S, ω ∈ B(H)∗] = [((idA ⊗ idS ⊗ xω)(δA(a)13(1A ⊗ y ⊗ 1S)qβAα,12) ; a ∈ A, x ∈ K, y ∈ S, ω ∈ B(H)∗] = [(idA ⊗ idS ⊗ ω)(δA(a)13((y ⊗ 1K)qα(1S ⊗ x))23) ; a ∈ A, x ∈ K, y ∈ S, ω ∈ B(H)∗] = [(idA ⊗ idS ⊗ ω)(δA(a)13((y ⊗ 1K)W(1S ⊗ x))23) ; a ∈ A, x ∈ K, y ∈ S, ω ∈ B(H)∗] = [(idA ⊗ idS ⊗ ω)((1A ⊗ y ⊗ 1)δA(a)13W23) ; a ∈ A, x ∈ K, y ∈ S, ω ∈ B(H)∗] = [(idA ⊗ idS ⊗ ω)((1A ⊗ (y ⊗ x)W)δ2 A(a)) ; a ∈ A, x ∈ K, y ∈ S, ω ∈ B(H)∗] = [(idA ⊗ idS ⊗ ω)((1A ⊗ y ⊗ x)δ2 = [(1A ⊗ y)δA((idA ⊗ ωx)δA(a)) ; a ∈ A, x ∈ K, y ∈ S, ω ∈ B(H)∗] = [(1A ⊗ S)δA(A)]. As a first application, we have the result below. strongly continuous and there exists a unique (cid:98)G-equivariant *-isomorphism τ : No (cid:111) G → (cid:98)S such 5.3.7 Proposition. -- If the groupoid G is regular, then the trivial action of G on No (cf. 5.1.4) is that τ(π(no)(cid:98)θ(x)) = β(no)x for all n ∈ N and x ∈ (cid:98)S. (cid:78) Proof. In this proof, we use the notations of 5.3.2 with A := No. In this case, we have I for all I ∈ I. According to Theorem 5.3.6, it suffices to show that the trivial action qI = εo is weakly continuous. Since the C*-algebra N is finite-dimensional, it amounts to proving I ∈ (cid:104)(idNo ⊗ ω)δNo (no) ; n ∈ N, ω ∈ B(H)∗(cid:105) for all I ∈ I. Let I ∈ I. For all n(cid:48) ∈ N, that εo there exists ω ∈ B(H)∗ such that ω(α(n)) = (n(cid:48)n) for all n ∈ N (extension of normal linear forms). In particular, there exists ω ∈ B(H)∗ such that ω(α(ε J)) = nI δI J for all J ∈ I. By a straightforward computation, we have εo I = (idNo ⊗ ω)δNo (1No ) and the weak (cid:98)θ(x) = π(1No )(cid:98)θ(x) ∈ No (cid:111) G for all x ∈ (cid:98)S. Moreover, we have π(no)(cid:98)θ(x) =(cid:98)θ(β(no)x) for continuity of the trivial action is then proved since No is unital. Since No is unital, we have all n ∈ N and x ∈ (cid:98)S. Hence, the morphism(cid:98)θ induces a *-isomorphism (cid:98)S (cid:39) No (cid:111) G. The equivariance is easily obtained from the definitions. 30 J. CRESPO 6 notion of equivariant hilbert c*-modules 6.1 Actions of measured quantum groupoids on Hilbert C*-modules In this section, we introduce a notion of G-equivariant Hilbert C*-module for a measured quantum groupoid G on a finite basis in the spirit of [3]. We fix a measured quantum 1(cid:54)l(cid:54)k Mnl (C) endowed with the non- 1(cid:54)l(cid:54)k nl · Trl. We use all the notations introduced in §§3.1, groupoid G on a finite-dimensional basis N = (cid:76) normalized Markov trace  =(cid:76) 3.2. Let us fix a G-C*-algebra A. the three pictures. be defined by three equivalent data: Following §2 [3], an action of G on a Hilbert A-module E will δE : E → (cid:102)M(E ⊗ S), cf. 6.1.1; • a pair (βE, δE) consisting of a *-homomorphism βE : No → L(E) and a linear map • a pair (βE, VE) consisting of a *-homomorphism βE : No → L(E) and an isometry V ∈ L(E ⊗δA (A ⊗ S), E ⊗ S), cf. 6.1.4; • an action (βJ, δJ) of G on J := K(E ⊕ A), cf. 6.1.8; satisfying some conditions. We have the following unitary equivalences of Hilbert modules: A ⊗δA (A ⊗ S) → qβAα(A ⊗ S) ; a ⊗δA x (cid:55)→ δA(a)x; (6.1) (A ⊗ S) ⊗δA⊗ idS (A ⊗ S ⊗ S) → qβAα,12(A ⊗ S ⊗ S) ; x ⊗δA⊗ idS y (cid:55)→ (δA ⊗ idS)(x)y; (6.2) (A ⊗ S) ⊗idA⊗ δ (A ⊗ S ⊗ S) → qβα,23(A ⊗ S ⊗ S) ; x ⊗idA⊗ δ y (cid:55)→ (idA ⊗ δ)(x)y. (6.3) In the following, we fix a Hilbert A-module E. We will apply the usual identifications M(A ⊗ S) = L(A ⊗ S), K(E) ⊗ S = K(E ⊗ S) and M(K(E) ⊗ S) = L(E ⊗ S). βE : No → L(E) is a non-degenerate *-homomorphism and δE : E → (cid:102)M(E ⊗ S) is a linear 6.1.1 Definition. -- An action of G on the Hilbert A-module E is a pair (βE, δE), where 1. for all a ∈ A and ξ, η ∈ E, we have δE(ξa) = δE(ξ)δA(a) and (cid:104)δE(ξ), δE(η)(cid:105) = δA((cid:104)ξ, η(cid:105)); map such that: 2. [δE(E)(A ⊗ S)] = qβEα(E ⊗ S); 3. for all ξ ∈ E and n ∈ N, we have δE(βE(no)ξ) = (1E ⊗ β(no))δE(ξ); 4. the linear maps δE ⊗ idS and idE ⊗ δ extend to linear maps from L(A ⊗ S, E ⊗ S) to L(A ⊗ S ⊗ S, E ⊗ S ⊗ S) and we have (δE ⊗ idS)δE(ξ) = (idE ⊗ δ)δE(ξ) ∈ L(A ⊗ S ⊗ S, E ⊗ S ⊗ S), for all ξ ∈ E. (cid:78) • If the second formula of the condition 1 holds, then δE is isometric 6.1.2 Remarks. -- (cf. [4], 8.3.2 1). • If the condition 1 holds, then the condition 2 is equivalent to: [δE(E)(1A ⊗ S)] = qβEα(E ⊗ S). Indeed, if (uλ)λ is an approximate unit of A we have δE(ξ) = lim λ δE(ξuλ) = lim λ δE(ξ)δA(uλ) = δE(ξ)qβAα, for all ξ ∈ E. By strong continuity of the action (βA, δA), the condition 1 of Definition 6.1.1 and the equality EA = E, we then have [δE(E)(A ⊗ S)] = [δE(E)(1A ⊗ S)] and the equivalence follows. ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS 31 • Note that we have qβEαδE(ξ) = δE(ξ) = δE(ξ)qβAα for all ξ ∈ E. • We will prove (cf. 6.1.7) that if δE satisfies the conditions 1 and 2 of 6.1.1, then the extensions of δE ⊗ idS and idE ⊗ δ always exist and satisfy the formulas: (idE ⊗ δ)(T)(idA ⊗ δ)(x) = (idE ⊗ δ)(Tx); (δE ⊗ idS)(T)(δA ⊗ idS)(x) = (δE ⊗ idS)(Tx); for all x ∈ A ⊗ S and T ∈ L(A ⊗ S, E ⊗ S). (cid:78) 6.1.3 Notation. -- For ξ ∈ E, let us denote by Tξ ∈ L(A ⊗ S, E ⊗δA (A ⊗ S)) the operator defined by Tξ (x) := ξ ⊗δA x, for all x ∈ A ⊗ S. ξ (η ⊗δA y) = δA((cid:104)ξ, η(cid:105))y for all η ∈ E and y ∈ A ⊗ S. In particular, we have We have T∗ ξ Tη = δA((cid:104)ξ, η(cid:105)) for all ξ, η ∈ E. T∗ (cid:78) 6.1.4 Definition. -- Let V ∈ L(E ⊗δA (A ⊗ S), E ⊗ S) be an isometry and βE : No → L(E) a non-degenerate *-homomorphism such that: Then, V is said to be admissible if we further have: 1. V V∗ = qβEα; 2. V(βE(no) ⊗δA 1) = (1E ⊗ β(no)) V, for all n ∈ N. 3. VTξ ∈ (cid:102)M(E ⊗ S), for all ξ ∈ E; 4. ( V ⊗C idS)( V ⊗δA⊗ idS 1) = V ⊗idA⊗ δ 1 ∈ L(E ⊗δ2 A (A ⊗ S ⊗ S), E ⊗ S ⊗ S). (cid:78) The fourth statement in the previous definition makes sense since we have used the canonical identifications thereafter. By combining the associativity of the internal tensor product with the unitary equivalences (6.2) and (6.3), we obtain the following unitary equivalences of Hilbert A ⊗ S-modules: (E ⊗δA (A ⊗ S)) ⊗δA⊗ idS (A ⊗ S ⊗ S) → E ⊗δ2 (ξ ⊗δA x) ⊗δA⊗ idS y (cid:55)→ ξ ⊗δ2 (E ⊗δA (A ⊗ S)) ⊗idA⊗ δ (A ⊗ S ⊗ S) → E ⊗δ2 (ξ ⊗δA x) ⊗idA⊗ δ y (cid:55)→ ξ ⊗δ2 A A A A (A ⊗ S ⊗ S) (δA ⊗ idS)(x)y; (A ⊗ S ⊗ S) (idA ⊗ δ)(x)y. We also have the following: (E ⊗ S) ⊗δA⊗ idS (A ⊗ S ⊗ S) → (E ⊗δA (A ⊗ S)) ⊗ S (ξ ⊗ s) ⊗δA⊗ idS (x ⊗ t) (cid:55)→ (ξ ⊗δA x) ⊗ st; (6.4) (6.5) (6.6) (E ⊗ S) ⊗idA⊗ δ (A ⊗ S ⊗ S) → qβα,23(E ⊗ S ⊗ S) ⊂ E ⊗ S ⊗ S (6.7) A ξ ⊗idA⊗ δ y (cid:55)→ (idE ⊗ δ)(ξ)y. (A ⊗ S ⊗ S), (E ⊗ S) ⊗δA⊗ idS (A ⊗ S ⊗ S)) (6.4) and In particular, V ⊗δA⊗ idS 1 ∈ L(E ⊗δ2 V ⊗C idS ∈ L((E ⊗ S) ⊗δA⊗ idS (A ⊗ S ⊗ S), E ⊗ S ⊗ S) (6.6). The next result provides an equivalence of the definitions 6.1.1 and 6.1.4. 6.1.5 Proposition. -- a) Let δE : E → (cid:102)M(E ⊗ S) be a linear map and βE : No → L(E) a non-degenerate *-homomorphism which satisfy the conditions 1, 2, and 3 of Definition 6.1.1. Then, there exists a unique isometry V ∈ L(E ⊗δA (A ⊗ S), E ⊗ S) such that δE(ξ) = VTξ for all ξ ∈ E. Moreover, the pair (βE, V) satisfies the conditions 1, 2, and 3 of Definition 6.1.4. 32 J. CRESPO b) Conversely, let V ∈ L(E ⊗δA (A ⊗ S), E ⊗ S) be an isometry and βE : No → L(E) a non-degenerate *-homomorphism, which satisfy the conditions 1, 2, and 3 of Definition 6.1.4. Let us consider the map δE : E → L(A ⊗ S, E ⊗ S) given by δE(ξ) := VTξ for all ξ ∈ E. Then, the pair (βE, δE) satisfies the conditions 1, 2 and 3 of Definition 6.1.1. c) Let us assume that the above statements hold. Then, the pair (βE, δE) is an action of G on E (cid:78) if, and only if, V is admissible. In the proof, we will use the following notation. projection and T ∈ L(qE,F ). Let (cid:101)T : E → F be the map defined by (cid:101)Tξ := Tqξ, for all 6.1.6 Notation. -- Let E and F be Hilbert C*-modules. Let q ∈ L(E ) be a self-adjoint ξ ∈ E. Therefore, (cid:101)T ∈ L(E,F ) and (cid:101)T∗ = qT∗. By abuse of notation, we will still denote by T the adjointable operator (cid:101)T. Proof of Proposition 6.1.5. a) By definition of the internal tensor product and 6.1.1 1, there exists a unique isometric (A ⊗ S)-linear map V : E ⊗δA (A ⊗ S) → E ⊗ S such that (cid:78) V(ξ ⊗δA x) = δE(ξ)x, for all ξ ∈ E and x ∈ A ⊗ S. In other words, we have VTξ = δE(ξ) for all ξ ∈ E. Now, it follows from 6.1.1 2 that the ranges of V and qβEα are equal. Then, denote by v the range restriction of V. Hence, the map v−1qβEα is an adjoint for V. Indeed, for all x ∈ E ⊗ S and y ∈ E ⊗δA (A ⊗ S) we have (cid:104)v−1qβEαx, y(cid:105) = (cid:104) Vv−1(qβEαx), Vy(cid:105) = (cid:104)qβEαx, Vy(cid:105) = (cid:104)x, Vy(cid:105) − (cid:104)(1 − qβEα)(x), Vy(cid:105) = (cid:104)x, Vy(cid:105). ( V is isometric) ( Vy ∈ im(qβEα)) Hence, V ∈ L(E ⊗δA (A ⊗ S), E ⊗ S) and then V∗ V = 1 and V V∗ = Vv−1qβEα = qβEα. The conditions 1 and 3 of Definition 6.1.4 are then fulfilled. Now, we have V(βE(no) ⊗δA 1)(ξ ⊗δA x) = δE(βE(no)ξ)x = (1E ⊗ β(no))δE(ξ)x = (1E ⊗ β(no)) V(ξ ⊗δA x), for all ξ ∈ E, x ∈ A ⊗ S and n ∈ N. Hence, the condition 2 of Definition 6.1.4 holds. b) is straightforward. c) Let T ∈ L(A ⊗ S, E ⊗ S). By using 6.1.6 and the identifications (6.3, 6.7), we have T ⊗idA⊗ δ 1 ∈ L(A ⊗ S ⊗ S, E ⊗ S ⊗ S). Now, we can define the extension idE ⊗ δ : L(A ⊗ S, E ⊗ S) → L(A ⊗ S ⊗ S, E ⊗ S ⊗ S) by setting (idE ⊗ δ)(T) := T ⊗idA⊗ δ 1, for all T ∈ L(A ⊗ S, E ⊗ S). We also have T ⊗δA⊗ idS 1 ∈ L(A ⊗ S ⊗ S, (E ⊗δA (A ⊗ S)) ⊗ S) by using again 6.1.6 and the identifications (6.2, 6.6). Let us define the extension δE ⊗ idS : L(A ⊗ S, E ⊗ S) → L(A ⊗ S ⊗ S, E ⊗ S ⊗ S) by setting (δE ⊗ idS)(T) := ( V ⊗C 1S)(T ⊗δA⊗ idS 1), for all T ∈ L(A ⊗ S, E ⊗ S). Therefore, for all ξ ∈ E we have: (δE ⊗ idS)δE(ξ) = ( V ⊗C 1S)( V ⊗δA⊗ idS 1)(Tξ ⊗δA⊗ idS 1) ∈ L(A ⊗ S ⊗ S, E ⊗ S ⊗ S); (idE ⊗ δ)δE(ξ) = ( V ⊗idA⊗ δ 1)(Tξ ⊗idA⊗ δ 1) ∈ L(A ⊗ S ⊗ S, E ⊗ S ⊗ S); ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS 33 where: Tξ ⊗δA⊗ idS 1 ∈ L(A ⊗ S ⊗ S, E ⊗δ2 Tξ ⊗idA⊗ δ 1 ∈ L(A ⊗ S ⊗ S, E ⊗δ2 A A (A ⊗ S ⊗ S)); (A ⊗ S ⊗ S)); by using the identifications (6.2, 6.4) and (6.3, 6.5) respectively and 6.1.6. In particular, if V is admissible, then the condition 4 of Definition 6.1.1 holds. Conversely, let us assume that the above condition is satisfied. In order to show that V is admissible, we only have to prove that the restrictions of the operators Tξ ⊗δA⊗ idS 1 and Tξ ⊗idA⊗ δ 1 to the Hilbert submodule qβAα,12qβα,23(A ⊗ S ⊗ S) are surjective. Let a ∈ A, x ∈ A ⊗ S and y ∈ A ⊗ S ⊗ S. Let us set z = (δA ⊗ idS)(δA(a)x)y. It is clear that z ∈ qβAα,12qβα,23(A ⊗ S ⊗ S). By a straightforward computation, we have (Tξ ⊗δA⊗ idS 1)(z) = ξa ⊗δ2 A (δA ⊗ idS)(x)y. Hence, the restriction of Tξ ⊗δA⊗ idS 1 to qβAα,12qβα,23(A ⊗ S ⊗ S) is surjective in virtue of (6.4) and the fact that EA = E. The same statement is also true for Tξ ⊗idA⊗ δ 1. 6.1.7 Remarks. -- In the proof of Proposition 6.1.5, we have proved the statements below. • By applying 6.1.6 and the identifications (6.3, 6.7), we have obtained a linear map idE ⊗ δ : L(A ⊗ S, E ⊗ S) → L(A ⊗ S ⊗ S, E ⊗ S ⊗ S) given by (idE ⊗ δ)(T) := T ⊗idA⊗ δ 1, for all T ∈ L(A ⊗ S, E ⊗ S); • If δE satisfies the conditions 1 and 2 of Definition 6.1.1, let V be the isometry associated with δE (cf. 6.1.5 a)). By applying 6.1.6 and the identifications (6.2, 6.6), the linear map δE ⊗ idS : L(A ⊗ S, E ⊗ S) → L(A ⊗ S ⊗ S, E ⊗ S ⊗ S) is defined by (δE ⊗ idS)(T) := ( V ⊗C 1S)(T ⊗δA⊗ idS 1), for all T ∈ L(A ⊗ S, E ⊗ S). Note that the extensions idE ⊗ δ and δE ⊗ idS satisfy the following formulas: (idE ⊗ δ)(T)(idA ⊗ δ)(x) = (idE ⊗ δ)(Tx); (δE ⊗ idS)(T)(δA ⊗ idS)(x) = (δE ⊗ idS)(Tx); (6.8) for all x ∈ A ⊗ S and T ∈ L(A ⊗ S, E ⊗ S). (cid:78) Let us denote by J := K(E ⊕ A) the linking C*-algebra associated with the Hilbert A- module E. In the following, we apply the usual identifications M(J) = L(E ⊕ A) and M(J ⊗ S) = L((E ⊗ S) ⊕ (A ⊗ S)). 6.1.8 Definition. -- An action (βJ, δJ) of G on J is said to be compatible with the action (βA, δA) if: 1. δJ : J → M(J ⊗ S) is compatible with δA, i.e. ιA⊗S ◦ δA = δJ ◦ ιA; 2. βJ : No → M(J) is compatible with βA, i.e. ιA(βA(no)a) = βJ(no)ιA(a), for all n ∈ N (cid:78) 6.1.9 Proposition. -- Let (βJ, δJ) be a compatible action of G on J. There exists a unique non-degenerate *-homomorphism βE : No → L(E) such that and a ∈ A. (cid:19) 0 βA(no) (cid:18)βE(no) (cid:19) 0 0 qβAα . βJ(no) = (cid:18)qβEα 0 Moreover, we have qβJ α = for all n ∈ N. , (cid:78) 34 J. CRESPO Proof. Note that since ιA, βA and βJ are *-homomorphisms, the condition 2 of Definition 6.1.8 is equivalent to: ιA(aβA(no)) = ιA(a)βJ(no), for all a ∈ A, n ∈ N. Therefore, there exists a map βE : No → L(E) necessarily unique such that (cid:18)βE(no) 0 (cid:19) , 0 βA(no) βJ(no) = for all n ∈ N. Then, it is clear that βE is a non-degenerate *-homomorphism and the last statement is then an immediate consequence. 6.1.10 Remarks. -- Note that if βA is injective, then so is βJ. For all n ∈ N, ξ ∈ E and k ∈ K(E), we have ιK(E)(βE(no)k) = βJ(no)ιK(E)(k) and ιE(βE(no)ξ) = βJ(no)ιE(ξ). In particular, we have βE(no)θξ,η = θβE(no)ξ,η for all n ∈ N and ξ, η ∈ E (cf. 2.3.2 2). (cid:78) (βJ, δJ) of G such that δJ(J) ⊂ (cid:102)M(J ⊗ S). Then, we have the following statements: 6.1.11 Proposition. -- a) Let us assume that the C*-algebra J is endowed with a compatible action • there exists a unique linear map δE : E → (cid:102)M(E ⊗ S) such that ιE⊗S ◦ δE = δJ ◦ ιE; furthermore, the pair (βE, δE) is an action of G on E, where βE : No → L(E) is the • there exists a unique faithful *-homomorphism δK(E) : K(E) → (cid:102)M(K(E) ⊗ S) such that *-homomorphism defined in 6.1.9; ιK(E⊗S) ◦ δK(E) = δJ ◦ ιK(E); moreover, the pair (βE, δK(E)) is an action of G on K(E). *-homomorphism δJ : J → (cid:102)M(J ⊗ S) such that ιE⊗S ◦ δE = δJ ◦ ιE. Moreover, we define a unique b) Conversely, let (βE, δE) be an action of G on the Hilbert A-module E. Then, there exists a faithful (cid:19) action (βJ, δJ) of G on J compatible with (βA, δA) by setting (cid:18)βE(no) 0 for all n ∈ N. , (cid:78) βJ(no) = 0 βA(no) Proof. a) Let us assume that the C*-algebra J is endowed with a compatible action (βJ, δJ) of G. Let βE : No → L(E) be the *-homomorphism defined in Proposition 6.1.9. By strict continuity and 6.1.8 1, we have δJ(ιA(m)) = ιA⊗S(δA(m)) for all m ∈ M(A). It then follows from 6.1.9 that δJ(ιK(E)(1E)) = δJ(1J) − δJ(ιA(1A)) = qβJ α − ιA⊗S(qβAα) = ιK(E⊗S)(qβEα). Let ξ ∈ E. We have ιK(E)(1E)ιE(ξ) = ιE(ξ) and ιE(ξ)ιK(E)(1E) = 0. Hence, ιK(E⊗S)(qβEα)δJ(ιE(ξ)) = δJ(ιE(ξ)) and δJ(ιE(ξ))ιK(E⊗S)(qβEα) = 0. We have ιA⊗S(x)δJ(ιE(ξ)) = ιA⊗S(x)ιL(E⊗S)(qβEα)δJ(ιE(ξ)) = 0, for all x ∈ A ⊗ S. Now, let (uλ)λ be an approximate unit of A. We have δJ(ιE(ξ)) = lim λ δJ(ιE(ξuλ)) = lim λ δJ(ιE(ξ))ιA⊗S(δA(uλ)). Hence, δJ(ιE(ξ))ιE⊗S(η) = 0 for all η ∈ E ⊗ S. Hence, there exists a unique linear map takes its values in the subspace (cid:102)M(E ⊗ S) of L(A ⊗ S, E ⊗ S). Indeed, let us fix ξ ∈ E and δE : E → L(A ⊗ S, E ⊗ S) such that ιE⊗S ◦ δE = δJ ◦ ιE (cf. 2.3.4 1). Moreover, δE actually s ∈ S. By assumption, we have that ιE⊗S((1E ⊗ s)δE(ξ)) = (1J ⊗ s)δJ(ιE(ξ)) and ιE⊗S(δE(ξ)(1A ⊗ s)) = δJ(ιE(ξ))(1J ⊗ s) belong to J ⊗ S = K((E ⊗ S) ⊕ (A ⊗ S)). It then follows that (1E ⊗ s)δE(ξ) ∈ E ⊗ S and δE(ξ)(1A ⊗ s) ∈ E ⊗ S. The first condition of 6.1.1 is easily derived from the compatibility of δJ. The vector subspace of δJ(1J)((E ⊕ A) ⊗ S) spanned by {δJ(θξ ⊕ a, η ⊕ b)(ζ) ; ξ, η ∈ E, a, b ∈ A, ζ ∈ (E ⊕ A) ⊗ S} ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS 35 is dense. However, we have δJ(θξ ⊕ a, η ⊕ b)(ζ) = (δE(ξ) ⊕ δA(a))(δE(η) ⊕ δA(b))∗(ζ), where δE(ξ)⊕ δA(a), δE(η)⊕ δA(b) ∈ L(A⊗ S, E⊗ S)⊕L(A⊗ S) ⊂ L(A⊗ S, (E⊕ A)⊗ S). In particular, the vector subspace of δJ(1J)((E ⊕ A) ⊗ S) spanned by {δE(ξ)x ⊕ δA(a)x ; ξ ∈ E, a ∈ A, x ∈ A ⊗ S} is dense. Therefore, the relation [δE(E)(A ⊗ S)] = qβEα(E ⊗ S) follows since we also have δJ(1J)((E ⊕ A) ⊗ S) = qβEα(E ⊗ S) ⊕ qβAα(A ⊗ S). Let us fix ξ ∈ E and n ∈ N. We have ιE⊗S(δE(βE(no)ξ)) = δJ(ιE(βE(no)ξ)) = δJ(βJ(no)ιE(ξ)) = (1J ⊗ β(no))δJ(ιE(ξ)) = ιE⊗S((1E ⊗ β(no)δE(ξ)). Hence, δE(βE(no)ξ) = (1E ⊗ β(no))δE(ξ). Let us consider the linear maps idE ⊗ δ : L(A ⊗ S, E ⊗ S) → L(A ⊗ S ⊗ S, E ⊗ S ⊗ S) and δE ⊗ idS : L(A ⊗ S, E ⊗ S) → L(A ⊗ S ⊗ S, E ⊗ S ⊗ S) (cf. 6.1.7). By using (6.8) and the compatibility of δJ with δA and δE, it follows from a straightforward computation that ιE⊗S⊗S((idE ⊗ δ)(T)) = (idJ ⊗ δ)(ιE⊗S(T)); ιE⊗S⊗S((δE ⊗ idS)(T)) = (δJ ⊗ idS)(ιE⊗S(T)); for all T ∈ L(A ⊗ S, E ⊗ S). In particular, ιE⊗S⊗S((idE ⊗ δ)δE(ξ)) = (idJ ⊗ δ)δJ(ιE(ξ)) and ιE⊗S⊗S((δE ⊗ idS)δE(ξ)) = (δJ ⊗ idS)δJ(ιE(ξ)) for all ξ ∈ E. Hence, for all ξ ∈ E we have (δE ⊗ idS)δE(ξ) = (idE ⊗ δ)δE(ξ). Therefore, the pair (βE, δE) is an action of G on E. We claim that there exists a unique *-homomorphism δK(E) : K(E) → M(K(E) ⊗ S) such that ιK(E⊗S) ◦ δK(E) = δJ ◦ ιK(E). We recall that δJ(ιK(E)(1E)) = ιK(E⊗S)(qβEα). We also have ιA⊗S(x)ιK(E⊗S)(qβEα) = 0 and ιK(E⊗S)(qβEα)ιA⊗S(x) = 0 for all x ∈ A ⊗ S. It follows that ιA⊗S(x)δJ(ιK(E)(k)) = 0 and δJ(ιK(E)(k))ιA⊗S(x) = 0 for all k ∈ K(E) and x ∈ A ⊗ S. Hence, δJ(ιK(E)(k)) ∈ ιK(E⊗S)(L(E ⊗ S)) (cf. 2.3.4) and the claim is proved since ιK(E⊗S) is faithful. Since ιK(E⊗S) is isometric and δJ ◦ ιK(E) is strictly continuous, the *-homomorphism δK(E) is strictly continuous and extend uniquely to a strictly continuous *-homomorphism δK(E) : L(E) → M(K(E) ⊗ S) such that δK(E)(1E) = qβEα. Moreover, for all ξ, η ∈ E we have (cf. 2.3.2 2) δK(E)(θξ,η) = δE(ξ) ◦ δE(η)∗ = θδE(ξ), δE(η) ∈ K((cid:102)M(E ⊗ S)) ⊂ (cid:102)M(K(E) ⊗ S). (6.9) Hence, δK(E)(K(E)) ⊂ (cid:102)M(K(E) ⊗ S). We have δK(E)(βE(no)) = (1E ⊗ β(no))qβEα, for all n ∈ N (cf. 6.9, 6.1.10). By strict continuity, we have the formulas: ιK(E⊗S⊗S)(idK(E) ⊗ δ)(T) = (idJ ⊗ δ)(ιK(E⊗S)(T)); ιK(E⊗S⊗S)(δK(E) ⊗ idS)(m)) = (δJ ⊗ idS)(ιK(E⊗S)(T)); for all T ∈ M(K(E) ⊗ S) = L(E ⊗ S). By applying the above formulas to T := δK(E)(k) for k ∈ K(E), we show that (δK(E) ⊗ idS)δK(E)(k) = (idK(E) ⊗ δ)δK(E)(k). b) First, it is clear that βJ is a non-degenerate *-homomorphism. It is also clear that βJ is compatible with the fibration map βA, i.e. βJ(no)ιA(a) = ιA(βA(no)a), for all a ∈ A and n ∈ N. Let V ∈ L(E ⊗δA (A ⊗ S), E ⊗ S) be the isometry associated with the action δE. Let i : qβAα(A ⊗ S) → A ⊗ S be the inclusion map. We verify that i is an (A ⊗ S)-linear adjointable map and i∗ = qβAα. In particular, the map i is an isometry since we have i∗i(x) = qβAαx = x for all x ∈ qβAα(A ⊗ S). Let W := V ⊕ i ∈ L((E ⊗δA (A ⊗ S)) ⊕ qβAα(A ⊗ S), (E ⊗ S) ⊕ (A ⊗ S)). We have W∗ W = 1, then W is an isometry. Henceforth, we will use the following identification (see (6.1)): (E ⊗δA (A ⊗ S)) ⊕ qβAα(A ⊗ S) = (E ⊗δA (A ⊗ S)) ⊕ (A ⊗δA (A ⊗ S)) = (E ⊕ A) ⊗δA (A ⊗ S). 36 J. CRESPO Hence, W ∈ L((E ⊕ A) ⊗δA (A ⊗ S), (E ⊕ A) ⊗ S). Let us define δJ(x) := W(x ⊗δA 1) W∗ ∈ M(J ⊗ S), for all x ∈ J. In that way, we define a strictly continuous *-homomorphism δJ : J → M(J ⊗ S) satisfying δJ(1J) = W W∗ = qβEα ⊕ qβAα = qβJ α. Let a ∈ A. Let us prove that ιA⊗S(δA(a)) = δJ(ιA(a)). Since ιA⊗S(δA(a)) W W∗ = ιA⊗S(δA(a)) and δJ(ιA(a)) W W∗ = δJ(ιA(a)), it amounts to proving that ιA⊗S(δA(a)) W = δJ(ιA(a)) W, for all a ∈ A. Therefore, it is enough to prove that ιA⊗S(δA(a)) W = W(ιA(a) ⊗δA 1) since W∗ W = 1. However, for all η ∈ E, b ∈ A and x ∈ A ⊗ S we have W((η ⊕ b) ⊗δA x) = V(η ⊗δA x) ⊕ δA(b)x = δE(η)x ⊕ δA(b)x. We finally obtain W(ιA(a) ⊗δA 1)((η ⊕ b) ⊗δA x) = W((0 ⊕ ab) ⊗δA x) = ( V ⊕ i)(0 ⊕ δA(ab)x) = 0 ⊕ δA(a)δA(b)x = ιA⊗S(δA(a))(δE(η)x ⊕ δA(b)x) = ιA⊗S(δA(a)) W((η ⊕ b) ⊗δA x), for all η ∈ E, b ∈ A and x ∈ A ⊗ S. By using similar arguments, we also prove that ιE⊗S(δE(ξ)) = δJ(ιE(ξ)) for all ξ ∈ E. By strict continuity, we obtain the formulas: (δJ ⊗ idS)ιA⊗S(m) = ιA⊗S⊗S(δA ⊗ idS)(m); (idJ ⊗ δ)ιA⊗S(m) = ιA⊗S⊗S(idA ⊗ δ)(m); for all m ∈ M(A ⊗ S). By applying the above formulas to m := δA(a) for a ∈ A and by using again the compatibility of δJ with δA, we obtain the formulas: (δJ ⊗ idS)δJ(ιA(a)) = ιA⊗S⊗S(δA ⊗ idS)δA(a); (idJ ⊗ δ)δJ(ιA(a)) = ιA⊗S⊗S(idA ⊗ δ)δA(a). Hence, (δJ ⊗ idS)δJ(ιA(a)) = (idJ ⊗ δ)δJ(ιA(a)) for all a ∈ A. In a similar way, we have (δJ ⊗ idS)δJ(ιE(ξ)) = (idJ ⊗ δ)δJ(ιE(ξ)) for all ξ ∈ E. However, since J is generated by ιE(E) ∪ ιA(A) as a C*-algebra, the coassociativity of δJ is then proved. For all η ∈ E, b ∈ A, x ∈ A ⊗ S and n ∈ N, we have δJ(βJ(no)) W((η ⊕ b) ⊗δA x) = W(βJ(no)(η ⊕ b) ⊗δA x) = W((βE(no)η ⊕ βA(no)b) ⊗δA x) = δE(βE(no)η)x ⊕ δA(βA(no)b)x = (1J ⊗ β(no))(δE(η)x ⊕ δA(b)x) = (1J ⊗ β(no)) W((η ⊕ b) ⊗δA x). Hence, δJ(βJ(no)) = δJ(βJ(no)) W W∗ = (1J ⊗ β(no)) W W∗ = (1J ⊗ β(no))δJ(1J), for all n ∈ N. Therefore, (βJ, δJ) is an action of G on J, compatible with (βA, δA). equivariant unitary equivalence. In this paragraph, we define a notion of equivariance for unitary equivalences of Hilbert modules acted upon by G. We refer the reader to §8.3 for the definitions and notations used below. 6.1.12 Definition. -- Let A and B be two G-C*-algebras and φ : A → B a G-equivariant *-isomorphism. Let E and F be two Hilbert modules over respectively A and B acted upon by G. A φ-compatible unitary operator Φ : E → F is said to be G-equivariant if we have (cid:78) We recall that the linear map Φ ⊗ idS : L(A ⊗ S, E ⊗ S) → L(B ⊗ S, F ⊗ S) (8.3.6) is the extension of the φ ⊗ idS-compatible unitary operator Φ ⊗ idS : E ⊗ S → F ⊗ S (8.3.4). 6.1.13 Proposition. -- With the notations and hypotheses of 6.1.12, for all n ∈ N we have βF(no) ◦ Φ = Φ ◦ βE(no). (cid:78) δF(Φξ) = (Φ ⊗ idS)δE(ξ), for all ξ ∈ E. ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS 37 Proof. It is clear that (Φ ⊗ idS)((1E ⊗ s)T) = (1F ⊗ s)(Φ ⊗ idS)(T) for all s ∈ S and T ∈ L(A ⊗ S, E ⊗ S). Let n ∈ N and ξ ∈ E. We have δF(Φ(βE(no)ξ)) = δF(βF(no)Φξ) by 6.1.1 3. Hence, Φ(βE(no)ξ) = βF(no)Φξ by 6.1.2 1. 6.1.14 Definition. -- Two Hilbert C*-modules E and F acted upon by G are said to be G-equivariantly unitarily equivalent if there exists a G-equivariant unitary operator from E (cid:78) onto F. It is clear that the G-equivariant unitary equivalence defines an equivalence relation on the class consisting of the Hilbert C*-modules acted upon by G. In the following, we provide equivalent definitions of the G-equivariant unitary equivalence in the two other pictures. Let A and B be two G-C*-algebras and φ : A → B a G-equivariant *-isomomorphism. Let E and F be two Hilbert C*-modules over A and B respectively and Φ : E → F a φ-compatible unitary operator. 6.1.15 Lemma. -- The linear map E ⊗δA (A ⊗ S) → F ⊗δB (B ⊗ S) ; ξ ⊗δA x (cid:55)→ Φξ ⊗δB (φ ⊗ idS)(x) is a φ ⊗ idS-compatible unitary operator. Proof. For all ξ ∈ E, a ∈ A and x ∈ A ⊗ S, we have (cid:78) Φ(ξa) ⊗δB (φ ⊗ idS)(x) = Φ(ξ)φ(a) ⊗δB (φ ⊗ idS)(x) = Φξ ⊗δB δB(φ(a))(φ ⊗ idS)(x) = Φξ ⊗δB (φ ⊗ idS)(δA(a)x). Therefore, we have a well-defined linear map Ψ : E (cid:12)A (A ⊗ S) → F ⊗δB (B ⊗ S) ; ξ (cid:12)A x (cid:55)→ Φξ ⊗δB (φ ⊗ idS)(x). For all ξ, η ∈ E, we have δB((cid:104)Φξ, Φη(cid:105)) = δB(φ((cid:104)ξ, η(cid:105))) = (φ ⊗ idS)δA((cid:104)ξ, η(cid:105)). Therefore, for all ξ, η ∈ E and x, y ∈ A ⊗ S, we have (cid:104)Φξ ⊗δB (φ ⊗ idS)(x), Φη ⊗δB (φ ⊗ idS)(y)(cid:105) = (φ ⊗ idS)((cid:104)ξ ⊗δA x, η ⊗δA y(cid:105)). Hence, (cid:104)Ψ(t), Ψ(t(cid:48))(cid:105) = (φ ⊗ idS)((cid:104)t, t(cid:48)(cid:105)) for all t, t(cid:48) ∈ E (cid:12)A (A ⊗ S). In particular, we have (cid:107)Ψ(t)(cid:107) = (cid:107)t(cid:107) for all t ∈ E (cid:12)A (A ⊗ S) (φ ⊗ idS is isometric). Therefore, Ψ extends uniquely to a bounded operator from E ⊗δA (A ⊗ S) to F ⊗δB (B ⊗ S) still denoted by Ψ. We have (cid:104)Ψ(t), Ψ(t(cid:48))(cid:105) = (φ ⊗ idS)((cid:104)t, t(cid:48)(cid:105)) for all t, t(cid:48) ∈ E ⊗δA (A ⊗ S). Since Ψ is isometric and has a dense range, we conclude that Ψ is surjective. A staightforward computation shows that Ψ(tx) = Ψ(t)(φ ⊗ idS)(x) for all t ∈ E ⊗δA (A ⊗ S) and x ∈ A ⊗ S. 6.1.16 Proposition. -- Let (βE, δE) (resp. (βF, δF)) be an action of G on E (resp. F). Denote by VE ∈ L(E ⊗δA (A ⊗ S), E ⊗ S) (resp. VF ∈ L(F ⊗δB (B ⊗ S), F ⊗ S)) the isometry associated with (βE, δE) (resp. (βF, δF)). Assume that Φ ◦ βE(no) = βF(no) ◦ Φ for all n ∈ N. Then, Φ is G-equivariant if, and only if, we have V∗ F(Φ ⊗ idS) VE(ξ ⊗δA x) = Φξ ⊗δB (φ ⊗ idS)(x), for all ξ ∈ E and x ∈ A ⊗ S. (cid:78) Proof. Let Ψ : E ⊗δA (A ⊗ S) → F ⊗δB (B ⊗ S) be the φ ⊗ idS-compatible unitary operator defined in the above lemma. For all ξ ∈ E and x ∈ A ⊗ S, we have δF(Φξ)(φ ⊗ idS)(x) = VF(Φξ ⊗δB x) = VF ◦ Ψ(ξ ⊗δA x) and (Φ⊗ idS)(δE(ξ))(φ⊗ idS)(x) = (Φ⊗ idS)(δE(ξ)x) = (Φ⊗ idS) VE(ξ ⊗δA x). Therefore, δF ◦ Φ = (Φ ⊗ idS) ◦ δE if, and only if, VF ◦ Ψ = (Φ ⊗ idS) VE. In order for the formula F(Φ ⊗ idS) VE. VF ◦ Ψ = (Φ ⊗ idS) VE to hold true, it is necessary and sufficient that Ψ = V∗ Indeed, it is necessary since V∗ F = qβFα, qβFα(Φ ⊗ idS) = (Φ ⊗ idS)qβEα (by assumption) and qβEα F VF = 1. It is also sufficient since we have VF V∗ VE = VE. 38 J. CRESPO 6.1.17 Remark. -- Let A be a G-C*-algebra. Let E and F be two Hilbert A-modules acted upon by G. Let Φ ∈ L(E, F) be a unitary. The following statements are equivalent: F(Φ ⊗ 1S) VE = Φ ⊗δA 1A⊗S; (i) Φ is G-equivariant; (ii) Φ ◦ βE(no) = βF(no) ◦ Φ for all n ∈ N and V∗ (iii) Φ ◦ βE(no) = βF(no) ◦ Φ for all n ∈ N and VF(Φ ⊗δA 1A⊗S) V∗ E = qβEα(Φ ⊗ 1S). (cid:78) 6.1.18 Proposition. -- Let A and B be two G-C*-algebras and φ : A → B a G-equivariant *-isomorphism. Let E and F be two Hilbert modules over respectively A and B acted upon by G. Let Φ : E → F be a φ-compatible unitary operator. Denote by f : K(E ⊕ A) → K(F ⊕ B) the unique *-homomorphism such that f ◦ ιE = ιF ◦ T and f ◦ ιA = ιB ◦ φ (cf. 8.3.5). Then, Φ is G-equivariant if, and only if, f is G-equivariant. (cid:78) Proof. Let J := K(E ⊕ A) and L := K(F⊕ B). Assume that Φ is equivariant. The following formulas are immediate consequences of the definitions: ιB⊗S ◦ (φ ⊗ idS)(m) = ( f ⊗ idS) ◦ ιA⊗S(m), m ∈ M(A ⊗ S); ιF⊗S ◦ (Φ ⊗ idS)(T) = ( f ⊗ idS) ◦ ιE⊗S(T), T ∈ L(A ⊗ S, E ⊗ S). By combining the first (resp. second) formula with the G-equivariance of φ (resp. Φ) and the fact that f ◦ ιA = ιB ◦ φ (resp. f ◦ ιE = ιF ◦ Φ), we obtain δL ◦ f (ιA(a)) = ( f ⊗ idS)δJ(ιA(a)), (resp. δL ◦ f (ιE(ξ)) = ( f ⊗ idS)δJ(ιE(ξ)), for all a ∈ A for all ξ ∈ E). (cid:19) 0 φ(βA(no)) = βL(no), Therefore, we have δL( f (x)) = ( f ⊗ idS)δJ(x) for all x ∈ J. Moreover, by definition of the fibration map on a linking C*-algebra (cf. 6.1.11) and the G-equivariance of Φ, we have (cid:18)βE(no) 0 f (βJ(no)) = (cid:19) (cid:18)Φ ◦ βE(no) ◦ Φ−1 0 βA(no) = 0 for all n ∈ N. The converse is proved in a similar way. continuous actions. In this paragraph, we introduce the notion of continuity for actions of the quantum groupoid G on Hilbert A-modules. If G is regular, we prove that any action of G is necessarily continuous. 6.1.19 Definition. -- An action (βE, δE) of G on a Hilbert A-module E is said to be continuous if we have [(1E ⊗ S)δE(E)] = (E ⊗ S)qβAα. A G-equivariant Hilbert A-module is a Hilbert A-module E endowed with a continuous action of G. (cid:78) 6.1.20 Proposition. -- Let E be a G-equivariant Hilbert A-module. Let B := K(E). We have the following statements: 1. the action (βB, δB) of G on B defined in 6.1.11 is strongly continuous; 2. we define a continuous action of G on the Hilbert B-module E∗ by setting: • βE∗ (no)T := βA(no) ◦ T, for all n ∈ N and T ∈ E∗, • δE∗ (T)x := δE(T∗)∗ ◦ x, for all T ∈ E∗ and x ∈ B ⊗ S; where we have applied the usual identifications B ⊗ S = K(E ⊗ S) and E = K(A, E). (cid:78) Proof. 1. We have [δB(B)(1B ⊗ S)] = [δB(θξ,η)(1B ⊗ y) ; ξ, η ∈ E, y ∈ S]. However, we have δB(θξ,η)(1B ⊗ y) = δE(ξ)δE(η)∗(1B ⊗ y) = δE(ξ)((1B ⊗ y∗)δE(η))∗ for all y ∈ S and ξ, η ∈ E. It then follows from the continuity of the action (βE, δE) and 6.1.2 that [δB(B)(1B ⊗ S)] = [δE(E)qβAα(E∗ ⊗ S)] = [δE(E)(E∗ ⊗ S)]. Now, by combining the formulas [δE(E)(1E ⊗ S)] = qβEα(E ⊗ S) and B = [EE∗] with the fact that any element of S can be written as a product of two elements of S, we obtain [δB(B)(1B ⊗ S)] = [δE(E)(1E ⊗ S)(E∗ ⊗ S)] = qβEα(B ⊗ S). 2. Straightforward. ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS 39 6.1.21 Proposition. -- Let E be a Hilbert A-module endowed with an action (βE, δE) of G on E. Let J := K(E ⊕ A) be the associated linking C*-algebra. Let (βJ, δJ) be the action defined in 6.1.11. Then, the action (βE, δE) is continuous if, and only if, the action (βJ, δJ) is strongly (cid:78) continuous. Proof. Assume that the action (βE, δE) is continuous. Let B := K(E). Let us prove that (βJ, δJ) is strongly continuous. Let x ∈ J and s ∈ S. Let us write (cid:18) b η∗ x = Then, we have (cid:19) ξ a , where a ∈ A, b ∈ B and ξ, η ∈ E. δJ(x)(1J ⊗ s) = ιB⊗S(δB(b)(1B ⊗ s)) + ιE⊗S(δE(ξ)(1A ⊗ s)) + ιE∗⊗S(δE∗ (η ∗)(1B ⊗ s)) + ιA⊗S(δA(a)(1A ⊗ s)). Then, the continuity of (βJ, δJ) follows from the continuity of (βA, δA), (βE, δE) and the continuity of (βB, δB) and (βE∗, δE∗ ) (6.1.20). Conversely, assume that (βJ, δJ) is continuous. We have ιE⊗S((E ⊗ S)qβAα) = (ιE(E) ⊗ S)qβJ α. Let ζ ∈ E and y ∈ S. As in the above computation, we prove that ιE⊗S((ζ ⊗ y)qβAα) is the norm limit of finite sums of elements of the following forms: ιB⊗S((1B ⊗ s)δB(b)), ιE⊗S((1E ⊗ s)δE(ξ)), ιE∗⊗S((1E∗ ⊗ s)δE∗ (η∗)) and ιA⊗S((1A ⊗ s)δA(a)). By multiplying by ιB⊗S(1B⊗S) on the left and by ιA⊗S(1A⊗S) on the right, we have that ιE⊗S((ζ ⊗ y)qβAα) is the norm limit of finite sums of elements of the form ιE⊗S((1E ⊗ s)δE(ξ)). The continuity of (βE, δE) follows from the fact that ιE⊗S is isometric. 6.1.22 Definition. -- A linking G-C*-algebra is a quintuple (J, βJ, δJ, e1, e2) consisting of a C*-algebra J endowed with a continuous action (βJ, δJ) and nonzero self-adjoint projections e1, e2 ∈ M(J) satisfying the following conditions: 1. e1 + e2 = 1J; 2. [Jej J] = J, for all j = 1, 2; 3. δJ(ej) = qβJ α(ej ⊗ 1S), for all j = 1, 2. (cid:78) 6.1.23 Remarks. -- • Let (A, βA, δA) be a G-C*-algebra and m ∈ M(A) such that δA(m) = qβAα(m ⊗ 1S). Let n ∈ N, we have [m, βA(no)] = 0. Indeed, since α and β commute pointwise we have [qβAα(1A ⊗ β(no)), qβAα(m ⊗ 1S)] = 0. It then follows that δA([m, βA(no)]) = [δA(mβA(no)), δA(βA(no)m)] = 0. Hence, [m, βA(no)] = 0 by faithfulness of δA. In particular, we have [qβAα, m ⊗ 1S] = 0. • Let (J, βJ, δJ, e1, e2) be a linking G-C*-algebra. By restriction of the action (βJ, δJ), the corner e2Je2 (resp. e1Je2) turns into a G-C*-algebra (resp. G-equivariant Hilbert C*-module over e2Je2). Furthermore, we also have the identification of G-C*-algebras K(e1Je2) = e1Je1. • Conversely, if (E, βE, δE) is a G-equivariant Hilbert A-module, then the C*-algebra J := K(E ⊕ A) endowed with the continuous action (βJ, δJ) (cf. 6.1.11, 6.1.21) and the projections e1 := ιE(1E) and e2 := ιA(1A) is a linking G-C*-algebra. (cid:78) 6.1.24 Lemma. -- Let E be a Hilbert A-module endowed with an action (βE, δE) of G. We have E = [(idE ⊗ ω)δE(ξ) ; ξ ∈ E, ω ∈ B(H)∗] (cf. 2.3.6). (cid:78) Proof. We have E ⊃ [(idE ⊗ ω)δE(ξ) ; ξ ∈ E, ω ∈ B(H)∗] (cf. 2.3.6). To obtain the converse inclusion, we combine the fact that there exists ω ∈ B(H)∗ such that (idE ⊗ ω)(qβEα) = 1E with the formula [δE(E)(1B ⊗ S)] = qβEα(E ⊗ S). Now, we can state the main results of this chapter. 6.1.25 Theorem. -- Let A be a G-C∗-algebra and E a Hilbert A-module acted upon by G. Let J := K(E ⊕ A) be the associated linking C*-algebra endowed with the action (βJ, δJ) defined in 6.1.11. If G is semi-regular (resp. regular), then the action (βJ, δJ) is weakly (resp. strongly) (cid:78) continuous. 40 J. CRESPO Proof. Assume that G is semi-regular. Denote by T := [(idJ ⊗ ω)δJ(x) ; x ∈ J, ω ∈ B(H)∗] the C∗-algebra of continuous elements (cf.5.3.4). By combining the compatibility of δJ with δA (resp. δE) with the fact that (βA, δA) is (weakly) continuous (resp. Lemma 6.1.24), we δJ(J) ⊂ (cid:102)M(J ⊗ S). Hence, (βJ, δJ) is weakly continuous. It follows from 5.3.6 that the obtain ιA(A) ⊂ T (resp. ιE(E) ⊂ T). Hence, J ⊂ T. The converse inclusion also holds since action (βJ, δJ) is strongly continuous if G is regular. 6.1.26 Corollary. -- Let E be a Hilbert A-module. If the quantum groupoid G is regular, then any action of G on E is continuous. (cid:78) Proof. This is an immediate consequence of 6.1.21 and 6.1.25. 6.2 Case of a colinking measured quantum groupoid Let us fix a colinking measured quantum groupoid G := GG1,G2 between two monoidally equivalent locally compact quantum groups G1 and G2. Let (A, βA, δA) be a G-C*-algebra. We follow all the notations of §3.3 (resp. 5.2.1 and 5.2.2) concerning the objects associated with G (resp. (A, βA, δA)). In the following, we provide a description of Hilbert modules acted upon by G in terms of Hilbert modules acted upon by G1 and G2. Let us fix a Hilbert A-module E endowed with an action (βE, δE) of G. 6.2.1 Notations. -- We introduce some useful notations to describe the action (βE, δE). • Let qE,j := βE(εj) for j = 1, 2. Note that qE,1 and qE,2 are orthogonal self-adjoint projections of L(E) and qE,1 + qE,2 = 1E. • Let J := K(E ⊕ A) be the linking C*-algebra associated with E endowed with the action (βJ, δJ) of G (cf. 6.1.11 b)). Since βJ(C2) ⊂ Z (M(J)) (cf. 3.2.3 [2]), we have βE(n)ξ = ξβA(n) in L(A, E) for all n ∈ C2 and ξ ∈ E, i.e. (βE(n)ξ)a = ξ(βA(n)a) for all n ∈ C2, ξ ∈ E and a ∈ A. Hence, (qE,jξ)a = ξ(qA,ja), for all ξ ∈ E, a ∈ A, j = 1, 2. (6.10) In particular, we have (cid:104)qE,jξ, qE,jη(cid:105) = qA,j(cid:104)ξ, η(cid:105), for all ξ, η ∈ E. Indeed, fix ξ, η ∈ E and write ξ = ξ(cid:48)a and η = η(cid:48)b with ξ(cid:48), η(cid:48) ∈ E and a, b ∈ A. Since the projection qA,j is central in A, we have (cid:104)qE,jξ, qE,jη(cid:105) = (cid:104)(qE,jξ(cid:48))a, (qE,jη(cid:48))b(cid:105) = (cid:104)ξ(cid:48)(qA,ja), η(cid:48)(qA,jb)(cid:105) = qA,ja∗(cid:104)ξ(cid:48), η(cid:48)(cid:105)b = qA,j(cid:104)ξ, η(cid:105). For j = 1, 2, we then define the following Hilbert Aj-module Ej := qE,jE. Note that E = E1 ⊕ E2. • For j, k = 1, 2, let Πk j It is clear that Πk j -compatible operator (cf. 8.3.1). We can consider its canonical linear j j : L(Ak ⊗ Skj, Ek ⊗ Skj) → L(A ⊗ S, E ⊗ S), up to the canonical injective extension Πk maps Ek ⊗ Skj → L(Ak ⊗ Skj, Ek ⊗ Skj) and E ⊗ S → L(A ⊗ S, E ⊗ S), defined by j ◦ T((qA,k ⊗ pkj)x) for all T ∈ L(Ak ⊗ Skj, Ek ⊗ Skj) and x ∈ A⊗ S. (cid:78) j (T)(x) := Πk Πk 6.2.2 Lemma. -- With the above notations and hypotheses, we have a canonical unitary equivalence (cid:78) of Hilbert A ⊗ S-modules E ⊗δA (A ⊗ S) =(cid:76) : Ek ⊗ Skj → E ⊗ S. be the inclusion map. (Ak ⊗ Skj). j,k=1,2 Ej ⊗δk is a πk Aj Proof. This is a straightforward verification to see that we define a unitary adjointable operator by the following formula: Ej ⊗δk (Ak ⊗ Skj) ; ξ ⊗δA x (cid:55)→ (cid:77) E ⊗δA (A ⊗ S) → (cid:77) (qA,k ⊗ pkj)x. qE,jξ ⊗δk j,k=1,2 Aj j,k=1,2 Aj ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS 41 6.2.3 Proposition-Definition. -- Let V ∈ L(E ⊗δA (A ⊗ S), E ⊗ S) be the isometry associated with the action (βE, δE) (cf. 6.1.5 a). For all j, k = 1, 2, there exists a unique unitary j ∈ L(Ej ⊗δk Vk Aj (Ak ⊗ Skj), Ek ⊗ Skj) such that V(ξ ⊗δA x) = ∑ j,k=1,2 j (qE,jξ ⊗δk Vk Aj (qA,k ⊗ pkj)x), for all ξ ∈ E and x ∈ A ⊗ S. (cid:78) Proof. Let j, k = 1, 2. Fix ξ ∈ E, x ∈ A ⊗ S and write x = x(cid:48)x(cid:48)(cid:48) with x(cid:48), x(cid:48)(cid:48) ∈ A ⊗ S. We have V(qE,jξ ⊗δA (qA,k ⊗ pkj)x) = (1E ⊗ β(εj)) V(ξ ⊗δA (qA,k ⊗ pkj)x) (6.1.4 2) (qA,k, pkj are central). ( V is A ⊗ S-linear). = (1E ⊗ β(εj)) V(ξ ⊗δA x(cid:48)(qA,k ⊗ pkj)x(cid:48)(cid:48)) = (1E ⊗ β(εj)) V(ξ ⊗δA x(cid:48))(qA,k ⊗ pkj)x(cid:48)(cid:48) Now if η ∈ E, y, s ∈ S and a ∈ A, we have (1E ⊗ β(εj))(η ⊗ y)(qA,k ⊗ pkj)(a ⊗ s) = η(qA,ka) ⊗ β(εj)ypkjs = η(qA,ka) ⊗ pkjys = qE,kηa ⊗ pkjys ∈ Ek ⊗ Skj by using (6.10) and the fact that pkj is central in S. In particular, for all ξ ∈ E and x ∈ A ⊗ S we have V(qE,jξ ⊗δA (qA,k ⊗ pkj)x) ∈ Ek ⊗ Skj. By combining the fact that V is isometric (a) = xδA(a) for all a ∈ Aj and x ∈ Ak ⊗ Skj, we obtain a well-defined with the fact that xδk Aj isometric Ak ⊗ Skj-linear map Aj Vk j : Ej ⊗δk (Ak ⊗ Skj) → Ek ⊗ Skj ; ξ ⊗δk It follows from im( V) = qβEα(E ⊗ S) (6.1.4 1) that Vk j j ∈ L(Ej ⊗δk Vk is unitary. For j, k, l = 1, 2 we have the following set of unitary equivalences of Hilbert modules: (Ak ⊗ Skj), Ek ⊗ Skj) and Vk x (cid:55)→ V(ξ ⊗δA x). is surjective. As a result, we have Aj Aj j Aj ⊗δk Aj (Ak ⊗ Skj) → Ak ⊗ Skj (a)x; x (cid:55)→ δk Aj a ⊗δk Aj (Ak ⊗ Skj) ⊗δl Ak ⊗ idSkj (Al ⊗ Slj) ⊗idAl ⊗ δk lj (Al ⊗ Slk ⊗ Skj) → Al ⊗ Slk ⊗ Skj x ⊗δl (Al ⊗ Slk ⊗ Skj) → Al ⊗ Slk ⊗ Skj y (cid:55)→ (δl Ak ⊗ idSkj ⊗ idSkj )(x)y; Ak x ⊗idAl ⊗ δk lj y (cid:55)→ (idAl ⊗ δk lj)(x)y; (Ej ⊗δk Aj (Ak ⊗ Skj)) ⊗δl (Ej ⊗δl Aj (Al ⊗ Slj)) ⊗idAl (Ek ⊗ Skj) ⊗δl Ak ⊗ idSkj (ξ ⊗ s) ⊗δl (El ⊗ Slj) ⊗idAl ⊗ δk lj Ak ⊗ idSkj (ξ ⊗δk Aj ⊗ δk lj (ξ ⊗δl Ak Ak )δk Aj ⊗ idSkj ⊗ idSkj Ak ⊗ idSkj )δk Aj (Al ⊗ Slk ⊗ Skj) → Ej ⊗(δl (Al ⊗ Slk ⊗ Skj) y (cid:55)→ ξ ⊗(δl x) ⊗δl ⊗ idSkj )(x)y; (δl Ak (Al ⊗ Slk ⊗ Skj) → Ej ⊗(idAl (Al ⊗ Slk ⊗ Skj) ⊗ δk lj)δl Aj y (cid:55)→ ξ ⊗(idAl x) ⊗idAl (idAl ⊗ δk lj)(x)y; ⊗ δk lj)δl Aj (Al ⊗ Slk)) ⊗ Skj (Al ⊗ Slk ⊗ Skj) → (Ek ⊗δl Ak x) ⊗ st; (x ⊗ t) (cid:55)→ (ξ ⊗δl (Al ⊗ Slk ⊗ Skj) → El ⊗ Slk ⊗ Skj ⊗ idSkj ⊗ δk lj Ak Ak Aj ξ ⊗idAl ⊗ δk lj y (cid:55)→ (idEl ⊗ δk lj)(ξ)y. (6.11) (6.12) (6.13) (6.14) (6.15) (6.16) (6.17) 42 J. CRESPO 6.2.4 Proposition. -- For all j, k, l = 1, 2, we have Ak Ak Ak Ak (cid:78) 1. ⊗δk lj ⊗ idSkj j ⊗δl j ⊗δl 1) = Vl k ⊗C idSkj )( Vk ( Vl (Al ⊗ Slk ⊗ Skj), (Ek ⊗ Skj) ⊗δl ⊗idSkj k ⊗C idSkj ∈ L((Ek ⊗ Skj) ⊗δl 1 ∈ L(Ej ⊗(δl ⊗ δk lj j ⊗idAl (Al ⊗ Slk ⊗ Skj), El ⊗ Slk ⊗ Skj) (Al ⊗ ⊗ idSkj 1 ∈ L(Ej ⊗(idAl (Al ⊗ Slk ⊗ Skj), El ⊗ Slk ⊗ Skj) ⊗ δk lj)δl Aj k ⊗C idSkj )( Vk j ⊗δl 1) does make sense since ⊗idSkj For j, k, l = 1, 2, Vl (6.16), Vk ⊗ idSkj Slk ⊗ Skj)) (6.14) and Vl (6.17). Moreover, the composition ( Vl (δl Ak Proof. Straightforward consequence of ( V ⊗C idS)( V ⊗δA⊗idS 1) = V ⊗idA⊗ δ 1. 6.2.5 Proposition-Definition. -- For j, k = 1, 2, let δk Ej linear map defined by : Ej → L(Ak ⊗ Skj, Ek ⊗ Skj) be the = (idAl ⊗ δk ⊗ idSkj )δk Aj j ⊗idAl ⊗ idSkj lj)δl Aj )δk Aj Ak . Ak δk Ej (ξ)x := Vk j (ξ ⊗δk Aj x), for all ξ ∈ Ej and x ∈ Ak ⊗ Skj. (i) δE(ξ) = ∑ k,j=1,2 For all j, k, l = 1, 2, we have: j ◦ δk Πk Ej (Ej) ⊂ (cid:102)M(Ek ⊗ Skj); (ii) δk Ej (qE,jξ), for all ξ ∈ E; (iii) δk Ej (ξa) = δk Ej (ξ)δk Aj (a) and (cid:104)δk Ej (ξ), δk Ej (η)(cid:105) = δk Aj ((cid:104)ξ, η(cid:105)), for all ξ, η ∈ Ej and a ∈ Aj; (iv) [δk Ej (Ej)(1Ak ⊗ Skj)] = Ek ⊗ Skj; in particular, we have Ek = [(idEk ⊗ ω)δk Ej (ξ) ; ω ∈ B(Hkj)∗, ξ ∈ Ej] (cf. 2.3.6). ⊗ idSkj (resp. idEl ⊗ δk lj) extends to a linear map from L(Ak ⊗ Skj, Ek ⊗ Skj) (resp. (v) δl Ek L(Al ⊗ Slj, El ⊗ Slj)) to L(Al ⊗ Slk ⊗ Skj, El ⊗ Slk ⊗ Skj) and for all ξ ∈ Ej we have (ξ) ∈ L(Al ⊗ Slk ⊗ Skj, El ⊗ Slk ⊗ Skj); (ξ) = (idEl ⊗ δk ⊗ idSkj )δk Ej lj)δl Ej (δl Ek j Ej (Ej)] = Ek ⊗ Skj. (vi) if E is a G-equivariant Hilbert A-module, then we have [(1Ek ⊗ Skj)δk Ej If E is a G-equivariant Hilbert module, then (Ej, δ is the associated unitary. ) is a Gj-equivariant Hilbert Aj-module and Vj j (cid:78) : Ej → L(Ak ⊗ Skj, Ek ⊗ Skj) is a well-defined linear map. Moreover, Proof. It is clear that δk Ej statement (i) follows straightforwardly from 6.2.3 and the fact that δE(ξ)x = V(ξ ⊗δA x) for all ξ ∈ E and x ∈ A ⊗ S. Let ξ ∈ Ej and s ∈ Skj. We have It then follows from δE(E) ⊂ (cid:102)M(E ⊗ S) that Πk belong to E ⊗ S. Moreover, (qE,k ⊗ pkj)Πk hence, Πk then follows that δk Ej j ((1Ek ⊗ s)δk Ej (ξ)(1Ak ⊗ s)) and Πk j (δk (ξ)) Ej j (T) for all T ∈ L(Ak ⊗ Skj, Ek ⊗ Skj); j (T) = Πk j ((1Ek ⊗ s)δk j (Ek ⊗ Skj) and Πk It Ej (ξ)(1Ak ⊗ s) and (1Ek ⊗ s)δk (ξ) belong to Ek ⊗ Skj by injectivity of Ej (ξ)) = (1E ⊗ s)δE(ξ) j ((1Ek ⊗ s)δk Ej j (Ek ⊗ Skj). (ξ)(1Ak ⊗ s)) = δE(ξ)(1A ⊗ s) (ξ)(1Ak ⊗ s)) ∈ Πk (ξ)) ∈ Πk and Πk j (δk Ej j (δk Ej Πk ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS 43 j : L(Ak ⊗ Skj, Ek ⊗ Skj) → L(A ⊗ S, E ⊗ S). Hence, statement (ii) is proved. Πk Let ξ, η ∈ Ej. We have j ((cid:104)δk πk Ej (η)(cid:105)) = (cid:104)Πk (ξ), δk Ej (ξ)), Πk (η))(cid:105) j (δk Ej j (δk Ej = (cid:104)(qE,k ⊗ pkj)δE(ξ), (qE,k ⊗ pkj)δE(η)(cid:105) = (qA,k ⊗ pkj)δA((cid:104)ξ, η(cid:105)) = πk j (δk Aj ((cid:104)ξ, η(cid:105))). ((cid:104)ξ, η(cid:105)) by injectivity of πk (ξ), δk Ej (η)(cid:105) = δk Aj j . The first formula of statement (iii) Hence (cid:104)δk Ej is derived immediately from the definition of δk . Ej (Ej)(Ak ⊗ Skj)] = Ek ⊗ Skj. The identity is just a restatement of [δk The surjectivity of Vk Ej j (Ej)(1Ak ⊗ Skj)] = Ek ⊗ Skj follows by combining the previous formula with the first [δk Ej (Aj)(1Ak ⊗ Skj)] = Ak ⊗ Skj. Let us prove the formula relation of (iii) and the relation [δk Aj Ek = [(idEk ⊗ ω)δk (ξ) ; ω ∈ B(Hkj)∗, ξ ∈ Ej]. By statement (ii) and 2.3.6, we already Ej have the relation Ek ⊃ [(idEk ⊗ ω)δk (ξ) ; ω ∈ B(Hkj)∗, ξ ∈ Ej]. Conversely, let us fix Ej η ∈ Ek. Let ω ∈ B(Hkj)∗ and s ∈ Skj such that ω(s) = 1. It then follows from the formula (Ej)(1Ak ⊗ Skj)] = Ek ⊗ Skj that η = (idEk ⊗ ω)(η ⊗ s) is the norm limit of finite sums [δk Ej of elements of the form (idEk ⊗ ω)(δk (ξ), where ξ ∈ Ej Ej and y ∈ S. Therefore, statement (iv) is proved. By using the identifications (6.12) and (6.16) (resp. (6.13) and (6.17)), the linear map (ξ)(1Ak ⊗ y)) = (idEk ⊗ yω)δk Ej δl Ek ⊗ idSkj : L(Ak ⊗ Skj, Ek ⊗ Skj) → L(Al ⊗ Slk ⊗ Skj, El ⊗ Slk ⊗ Skj) lj : L(Al ⊗ Slj, El ⊗ Slj) → L(Al ⊗ Slk ⊗ Skj, El ⊗ Slk ⊗ Skj)) (resp. idEl ⊗ δk is defined for all T ∈ L(Ak ⊗ Skj, Ek ⊗ Skj) (resp. T ∈ L(Al ⊗ Slj, El ⊗ Slj)) by (δl Ek k ⊗ 1Skj )(T ⊗δl ⊗ idSkj )(T) := ( Vj ⊗ idSkj )δk Ej (resp. (idEl ⊗ δk (ξ) for ξ ∈ Ej is then derived from 6.2.4 as The relation (δl Ek in the proof of 6.1.5. Assume that (βE, δE) is continuous. Since pkj is central in S, we have ⊗ idSkj (ξ) = (idEl ⊗ δk lj)(T) := T ⊗idAl lj)δl Ej ⊗ δk lj 1). 1) Ak j (Ek ⊗ Skj) = (qE,j ⊗ pkj)(E ⊗ S)qβAα Πk = (qE,j ⊗ pkj)[(1E ⊗ S)δE(E)] = Πk j [(1Ek ⊗ Skj)δk Ej (Ej)] and statement (vi) is proved. From this concrete description of G-equivariant Hilbert C*-modules, we can also provide a corresponding description of the G-equivariant unitary equivalences between them. 6.2.6 Lemma. -- Let A and B be G-C*-algebras. Let E and F be Hilbert C*-modules over A and B respectively acted upon by G. 1. Let Φ : E → F be a G-equivariant unitary equivalence over a G-equivariant *-isomorphism φ : A → B. For j = 1, 2, there exists a unique map Φj : Ej → Fj satisfying the formula Φ(ξ) = Φ1(qE,1ξ) + Φ2(qE,2ξ) for all ξ ∈ E. Moreover, we have: (i) for j = 1, 2, the map Φj is a unitary equivalence over the *-isomorphism φj : Aj → Bj (cf. 5.2.3 1); (ii) for all j, k = 1, 2, we have (Φk ⊗ idSkj ) ◦ δk Ej = δk Fj ◦ Φj. (6.18) 44 J. CRESPO In particular, Φj is a Gj-equivariant φj-compatible unitary operator. 2. Conversely, for j = 1, 2 let Φj : Ej → Fj be a Gj-equivariant unitary equivalence over a Gj-equivariant *-isomorphism φj : Aj → Bj such that (5.1) and (6.18) hold for all j, k = 1, 2. Then, the map Φ : E → F, defined by Φ(ξ) := Φ1(qE,1ξ) + Φ2(qE,2ξ) for all ξ ∈ E, is a G-equivariant unitary equivalence over the G-equivariant *-isomorphism φ : A → B (cf. (cid:78) 5.2.3 2). Proof. 1. Let j = 1, 2. Since Φ is G-equivariant, we have Φ ◦ qE,j = qF,j ◦ Φ. It then follows that Φ(Ej) ⊂ Fj. Let us denote Φj := Φ(cid:22)Ej: Ej → Fj. For ξ ∈ E, we have ξ = qE,1ξ + qE,2ξ; hence, Φ(ξ) = Φ1(qE,1ξ) + Φ2(qE,2ξ). Moreover, such a decomposition of Φ is unique since F1 and F2 are orthogonal in F. Statement (i) is straightforward. Let j, k = 1, 2 and x ∈ Ak ⊗ Skj. For all T ∈ L(Ak ⊗ Skj, Ek ⊗ Skj) we have j (T))(φ ⊗ idS)(x) = (Φk ⊗ idSkj )(T)(φk ⊗ idSkj )(x). j (δk Ej (Φ ⊗ idS)(Πk In particular, (Φ ⊗ idS)(Πk (ξ)))(φ ⊗ idS)(x) = (Φk ⊗ idSkj )(δk (ξ))(φk ⊗ idSkj )(x) for Ej all ξ ∈ Ej; hence, (Φ ⊗ idS)(δE(ξ))(φ ⊗ idS)(x) = (Φk ⊗ idSkj )(δk (ξ))(φk ⊗ idSkj )(x) (6.2.5 Ej (i)) for all ξ ∈ Ej. We also have δF(Φ(ξ))(qA,k ⊗ pkj) = Πk (Φj(ξ))) for all ξ ∈ Ej. j (δk Fj (Φj(ξ))(φk ⊗ idSkj )(x) for all ξ ∈ Ej and statement (ii) Hence, δF(Φ(ξ))(φ ⊗ idS)(x) = δk Fj is proved. 2. Straightforward. 6.2.7 Example. -- Let (βN, δN) be the trivial action (cf. 5.1.4). Let i = 1, 2. Consider the Hilbert N-module E := Hi1 ⊕ Hi2. Let V ∈ L(E ⊗δN (N ⊗ S), E ⊗ S) and βE : N → L(E) be the maps defined by the formulas: kj(ξ ⊗ 1), Vi βE(εj) = pij, ξ ∈ Hij j = 1, 2. ; V(ξ ⊗ 1) = ∑ k=1,2 Then, the pair ( V, βE) is an action of G on E. 6.3 Induction of equivariant Hilbert C*-modules (cid:78) Let G1 and G2 be two monoidally equivalent regular locally compact quantum groups. Fix a G1-C*-algebra (A1, δA1 ) and a G1-equivariant Hilbert A1-module (E1, δE1 ). We denote by J1 := K(E1 ⊕ A1) the associated linking C*-algebra endowed with the continuous action δJ1 of G1. 6.3.1 Notations. -- Let us fix some notations. 11 : L(A1 ⊗ S11, E1 ⊗ S11) → L(A1 ⊗ S12 ⊗ S21, E1 ⊗ S12 ⊗ S21) be the 11 : E1 ⊗ S11 → L(A1 ⊗ S12 ⊗ S21, E1 ⊗ S12 ⊗ S21) 11)(Tx) for all x ∈ M(A1 ⊗ S11) 11)(x) = (idE1 ⊗ δ2 • Let idE1 ⊗ δ2 unique linear extension of idE1 ⊗ δ2 such that (idE1 ⊗ δ2 11)(T)(idA1 ⊗ δ2 and T ∈ L(A1 ⊗ S11, E1 ⊗ S11). • Let δ (2) E1 (2) E1 (ξ) := (idE1 ⊗ δ2 δ 11)δE1 (ξ) for all ξ ∈ E1. : E1 → L(A1 ⊗ S12 ⊗ S21, E1 ⊗ S12 ⊗ S21) be the linear map defined by • Consider the Banach subspace of L(A1 ⊗ S12, E1 ⊗ S12) defined by (cf. 2.3.6): IndG2 G1 (E1) := [(idE1⊗S12 ⊗ ω)δ (2) E1 (ξ) ; ξ ∈ E1, ω ∈ B(H21)∗]. 6.3.2 Proposition. -- We have [IndG2 G1 In particular, IndG2 G1 (E1) ⊂ (cid:102)M(E1 ⊗ S12). (E1)(1A1 ⊗ S12)] = E1 ⊗ S12 = [(1E1 ⊗ S11)IndG2 G1 (cid:78) (E1)]. (cid:78) ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS 45 (E1)(1A1 ⊗ S12)] = E1 ⊗ S12. Fix ξ ∈ E1, s ∈ S12 Proof. Let us prove the formula [IndG2 G1 and ω ∈ B(H21)∗. Write ω = s(cid:48)ω(cid:48) with s(cid:48) ∈ S21 and ω(cid:48) ∈ B(H21)∗. It follows from 11(S11)(1S12 ⊗ S21)] that S12 ⊗ S21 = [δ2 (idE1⊗S12 ⊗ ω)(δ (ξ))(1A1 ⊗ s) = (idE1⊗S12 ⊗ ω is the norm limit of finite sums of elements of the form (2) E1 (cid:48))(δ (2) E1 (ξ)(1E1 ⊗ s ⊗ s(cid:48))) η = (idE1 ⊗ idS12 ⊗ ω (cid:48))(δ (2) E1 (ξ)(1E1 ⊗ δ2 11(t(cid:48))(1S12 ⊗ t))), with t(cid:48) ∈ S11 and t ∈ S21. It follows from [δE1 (E1)(1A1 ⊗ S11)] = E1 ⊗ S11 that η = (idE1⊗S12 ⊗ ω (cid:48))((idE1 ⊗ δ2 11)(δE1 (ξ)(1E1 ⊗ t(cid:48)))(1E1⊗S12 ⊗ t)) is the norm limit of finite sums of elements of the form η (cid:48) = (idE1⊗S12 ⊗ ω By using S12 = [(idS12 ⊗ ω)(δ2 11(t(cid:48)(cid:48))(1S12 ⊗ t)), with ζ ∈ E1 and t(cid:48)(cid:48) ∈ S21. (cid:48))(ζ ⊗ δ2 11(y)) ; ω ∈ B(H21)∗, y ∈ S11], we obtain Hence, (idE1⊗S12 ⊗ ω)(δ Therefore, the inclusion (2) E1 η (cid:48) = ζ ⊗ (idS12 ⊗ tω (ξ))(1A1 ⊗ s) ∈ E1 ⊗ S12 for all ξ ∈ E1, ω ∈ B(H21)∗ and s ∈ S12. 11(t(cid:48)(cid:48))) ∈ E1 ⊗ S12. (cid:48))(δ2 [IndG2 G1 (E1)(1A1 ⊗ S12)] ⊂ E1 ⊗ S12 is proved. The converse inclusion is obtained by following backwards the above argument. By a similar argument, we prove by using the relation (1E1 ⊗ S12)δE1 (E1) ⊂ E1 ⊗ S12 (E1)] ⊂ E1 ⊗ S12. For the that (1E1 ⊗ S12)IndG2 converse inclusion, it suffices to follow backwards the proof as above and to use the continuity of the action δE1. 6.3.3 Lemma. -- For all a ∈ A1, ξ ∈ E1, k ∈ K(E1) and ω ∈ B(H21)∗, we have: (E1) ⊂ E1 ⊗ S12. Hence, [(1E1 ⊗ S12)IndG2 G1 G1 1. ιA1⊗S11 (idA1⊗S12 ⊗ ω)δ 2. ιE1⊗S12 (idE1⊗S12 ⊗ ω)δ 3. ιK(E1⊗S12)(idK(E1)⊗S12 (2) E1 ⊗ ω)δ (2) A1 (a) = (idJ1⊗S12 ⊗ ω)δ (ξ) = (idJ1⊗S12 ⊗ ω)δ (2) J1 (ιA1 (a)); (2) J1 (ιE1 (ξ)); (2) J1 (2)K(E1)(k) = (idJ1⊗S12 ⊗ ω)δ (ιK(E1)(k)). (cid:78) Proof. These formulas are straightforward consequences of definitions and the compatibility of δJ1 with δA1 and δE1 and δK(E1) (2.7 (b), 2.8 (a) [3], 2.3.6). 6.3.4 Proposition. -- Let IndG2 G1 IndG2 G1 given by (cid:104)ξ, η(cid:105) := ξ∗ ◦ η for ξ, η ∈ IndG2 G1 Proof. Let ω, ω(cid:48) ∈ B(H21)∗, a ∈ A1 and ξ ∈ E1. Let η := (idE1⊗S12 ⊗ ω)δ x := (idA1⊗S12 ⊗ ω(cid:48))δ (A1)-module for the right action by composition and the IndG2 G1 (E1) is a Hilbert (A1)-valued inner product (cid:78) (A1) be the induced C*-algebra. Then IndG2 G1 (a). We have (ξ) and (E1). (2) E1 (2) A1 ιE1⊗S12 (ηx) = ιE1⊗S12 (η)ιA1⊗S12 (a) = (idJ1⊗S12 ⊗ ω)δ (2) J1 = (idJ1⊗S12 ⊗ ω ⊗ ω (ιE1 (ξ))(idJ1⊗S12 ⊗ ω (cid:48))(δ (ιE1 (ξ))123δ (2) J1 (2) J1 (2) J1 (cid:48))δ (ιA(a)) (ιA(a))124). (2.3.2 1) (6.3.3) In virtue of 4.2 a) [2], we have ιE1⊗S12 (ηx) ∈ IndG2 limit of finite sums of elements of the form y = (idJ1⊗S12 ⊗ φ)δ (cid:18) k (cid:19) G1 , with k ∈ K(E1), ζ ∈ E1, χ (2) J1 χ∗ ζ b (J1). Therefore, ιE1⊗S12 (ηx) is the norm ∗ ∈ E∗ 1 , b ∈ A1, φ ∈ B(H21)∗. 46 We have (6.3.3) J. CRESPO y = ιK(E1⊗S12)(idK(E1)⊗S12 ⊗ φ)δ (2)K(E1)(k) + ιE1⊗S12 (idE1⊗S12 ⊗ φ)δ (2) E1 (ζ) + ιE1⊗S12 (idE1⊗S12 ⊗ φ)δ (2) E1 (χ)∗ + ιA1⊗S12 (idA1⊗S12 ⊗ φ)δ (2) A1 (b). By multiplying on the left (resp. right) by ιK(E1⊗S12)(1E1⊗S12 ) (resp. ιA1⊗S12 (1A1⊗S12 )), we obtain (2.3.4) that ιE1⊗S12 (ηx) is the norm limit of finite sums of elements of the form ιE1⊗S12 (idE1 ⊗ idS12 ⊗ φ)δ (2) E1 (ζ), with ζ ∈ E1 and φ ∈ B(H21)∗. Since ιE1⊗S12 is isometric, we have proved that ηx ∈ IndG2 G1 Let us prove that ζ∗ ◦ χ ∈ IndG2 (A1) for all ζ, χ ∈ IndG2 G1 G1 Let us fix ξ, η ∈ E1 and ω, ψ ∈ B(H21)∗. Let us denote ζ := (idE1⊗S12 ⊗ ω)δ χ := (idE1⊗S12 ⊗ ψ)δ (E1). (E1) ⊂ L(A1 ⊗ S12, E1 ⊗ S12). (ξ) and (η). We have (2) E1 (2) E1 ιA1⊗S12 (ζ ∗ ◦ χ) = ιE1⊗S12 (ζ)∗ ιE1⊗S12 (χ) = (idJ1⊗S12 ⊗ ω)δ = (idJ1⊗S12 ⊗ ω ⊗ ψ)(δ (2) J1 (ιE1 (ξ)∗)(idJ1⊗S12 ⊗ ψ)δ (ιE1 (ξ)∗)123δ (2) J1 (2) J1 (2) J1 (ιE1 (η)) (ιE1 (η))124). (2.3.2 3) (6.3.3) G1 (2) A1 (A1) since ιA1⊗S12 is isometric. (A1, δA1 ) and (J2, δJ2 ) := IndG2 G1 (J1, δJ1 ) the induced G2-C*- (E1) the induced (J1) (4.2 a) [2]). As above, we prove that ιA1⊗S12 (ζ∗ ◦ χ) is the (a) with a ∈ A1 Hence, ιA1⊗S12 (ζ∗ ◦ χ) ∈ IndG2 norm limit of finite sums of elements of the form ιA1⊗S12 (idA1⊗S12 ⊗ φ)δ and φ ∈ B(H21)∗. We have proved that ζ∗ ◦ χ ∈ IndG2 G1 Let us denote (A2, δA2 ) := IndG2 G1 algebra of (A1, δA1 ) and (J1, δJ1 ) respectively. We also denote E2 := IndG2 G1 Hilbert A2-module as defined above. In the technical lemma below, we make the identification M(A) = L(A). We first recall a well-known corollary of Cohen-Hewitt factorization theorem. 6.3.5 Lemma. -- Let A be a C*-algebra and E a Hilbert A-module. If T : A → E is a map such that T(ab) = T(a)b for all a, b ∈ A, then T is linear and continuous. (cid:78) 6.3.6 Lemma. -- Let A be a C*-algebra, B ⊂ M(A) a non-degenerate C*-subalgebra and E a Hilbert A-module. Let F ⊂ L(A,E ) be a Hilbert B-module (where B is acting on the right by 1 ◦ η2, for all η1, η2 ∈ F) composition and the B-valued inner product is given by (cid:104)η1, η2(cid:105) := η∗ such that [F A] = E. (i) There exists a unique map i : L(B,F ) → L(A,E ) such that i(T)(ba) = (Tb)a for all T ∈ L(B,F ), b ∈ B and a ∈ A. Moreover, i is an injective linear map whose image is im(i) = {S ∈ L(A,E ) ; SB ⊂ F, S∗F ⊂ B}. (ii) There exists a unique map j : L(F ⊕ B) → L(E ⊕ A) such that j(x)(ηa) = (xη)a for all x ∈ L(F ⊕ B), η ∈ F ⊕ B and a ∈ A. Moreover, j is a unital faithful *-homomorphism. (cid:78) Proof. (i) We have A = BA. Let T ∈ L(B,F ). Let (uλ) be an approximate unit of B, we have (Tb)a = limλ [T(uλb)]a = limλ [T(uλ)b]a = limλ T(uλ)ba, for all b ∈ B and a ∈ A. In particular, we have (Tb)a = (Tb(cid:48))a(cid:48) for all b, b(cid:48) ∈ B and a, a(cid:48) ∈ A such that ba = b(cid:48)a(cid:48). Therefore, i(T) is well defined. Moreover, we have i(T)(aa(cid:48)) = (i(T)a)a(cid:48) for all a, a(cid:48) ∈ A. Indeed, let us fix a, a(cid:48) ∈ A. Let us write a = ba(cid:48)(cid:48) with b ∈ B and a(cid:48)(cid:48) ∈ A. We have i(T)(aa(cid:48)) = i(T)(b(a(cid:48)(cid:48)a(cid:48))) = (Tb)a(cid:48)(cid:48)a(cid:48) = i(T)(ba(cid:48)(cid:48))a(cid:48) = (i(T)a)a(cid:48). By Lemma 6.3.5, it then follows that i(T) is a bounded linear map. By a straightforward computation, we have (cid:104)i(T)(ba(cid:48)), ηa(cid:105) = (cid:104)ba(cid:48), T∗(η)a(cid:105), for all b ∈ B, a, a(cid:48) ∈ A and η ∈ F. Hence, (cid:104)i(T)x, ηa(cid:105) = (cid:104)x, T∗(η)a(cid:105) for all x, a ∈ A and η ∈ F. Let S ∈ L(F, B). We have (cid:104) x, for all a1,· · · an ∈ A, η1,· · · , ηn ∈ F and x ∈ A. S(ηl)al (cid:105) = (cid:104) i(S∗)x, ηlal (cid:105), n∑ n∑ l=1 l=1 ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS 47 As a consequence, the following map i(cid:48)(S) : (cid:104)F A(cid:105) → A ; n∑ ηlal (cid:55)→ n∑ l=1 S(ηl)al l=1 is well-defined and we have (cid:104)x, i(cid:48)(S)(ξ)(cid:105) = (cid:104)i(S∗)x, ξ(cid:105) for all ξ ∈ (cid:104)F A(cid:105) and x ∈ A. It follows from the boundedness of the linear operator i(S∗) and the Cauchy-Schwarz in- equality that (cid:107)i(cid:48)(S)ξ(cid:107)2 = (cid:107)(cid:104)i(cid:48)(S)ξ, i(cid:48)(S)ξ(cid:105)(cid:107) = (cid:107)(cid:104)ξ, i(S∗)(i(cid:48)(S)ξ)(cid:105)(cid:107) (cid:54) (cid:107)ξ(cid:107)(cid:107)i(S∗)(i(cid:48)(S)ξ)(cid:107) (cid:54) (cid:107)ξ(cid:107)(cid:107)i(S∗)(cid:107)(cid:107)i(cid:48)(S)ξ(cid:107) for all ξ ∈ E. Hence, (cid:107)i(cid:48)(S)ξ(cid:107) (cid:54) (cid:107)i(S∗)(cid:107)(cid:107)ξ(cid:107) for all ξ ∈ E, which proves the continuity of i(cid:48)(S) since i(cid:48)(S) is linear by definition. In particular, i(cid:48)(S) extends uniquely to a bounded linear map i(cid:48)(S) : E → A. By continuity of the inner product, we have proved that i(cid:48)(S) ∈ L(E, A) and i(cid:48)(S)∗ = i(S∗). As a result, we have well-defined maps i : L(A,E ) → L(B,F ) and i(cid:48) : L(F, B) → L(E, A) such that i(T)∗ = i(cid:48)(T∗) for all T ∈ L(A,E ). It is clear that i is linear and injective. It remains to prove that im(i) = {S ∈ L(A,E ) ; SB ⊂ F, S∗F ⊂ B}. Let T ∈ L(B,F ) and b ∈ B ⊂ L(A). For all a ∈ A, we have [i(T) ◦ b]a = i(T)(ba) = (Tb)a. Hence i(T) ◦ b = T(b) ∈ F. Fix η ∈ F. Write η = ζb with ζ ∈ F and b ∈ B. For all a ∈ A, we have [i(T)∗ ◦ η]a = [i(cid:48)(T∗)η]a = i(cid:48)(T∗)(ηa) = i(cid:48)(T∗)(ζ(ba)) = (T∗ζ)ba = T∗(ζb)a = T∗(η)a. Hence, i(T)∗ ◦ η = T∗(η) ∈ B. Conversely, let us fix S ∈ L(A,E ) such that SB ⊂ F and S∗F ⊂ B. Let T : A → E and T(cid:48) : F → B be the maps defined by: T(b) := S ◦ b, b ∈ B; T(cid:48)(η) := S∗ ◦ η, η ∈ F. For all b ∈ B and η ∈ F, we have (cid:104)T(b), η(cid:105) = (S ◦ b)∗ ◦ η = b∗(S∗ ◦ η) = (cid:104)b, T(cid:48)(η)(cid:105). Hence T ∈ L(A,E ) and T∗ = T(cid:48). Moreover, we have i(T)(ba) = T(b)a = S(ba) for all a ∈ A and b ∈ B. Thus, we have S = i(T). (ii) Since [F A] = E, we have [(F ⊕ B)A] = E ⊕ A, which proves the uniqueness of j. Let i12 := i and i21 := i(cid:48). By a similar argument as in statement (i), we prove that there exists a unique map i11 : L(F ) → L(E ) such that i11(T)(ηa) = (Tη)a for all η ∈ F and a ∈ A. The non-degenerate inclusion of C*-algebras B ⊂ M(A) extends to a unital *-homomorphism i22 : M(B) → M(A). Then we consider the map j : L(F ⊕ B) → L(E ⊕ A) defined by j(x) := (ikl(xkl))k,l=1,2 for all x = (xkl)k,l=1,2 ∈ L(F ⊕ B). It is clear that j(x)(ηa) = (xη)a for all x ∈ L(F ⊕ B), η ∈ F ⊕ B and a ∈ A. The fact that j is a unital faithful *- homomorphism is then straightforward. 6.3.7 Remarks. -- With the notations and hypotheses of the previous proposition, we have: (cid:78) (i) for all T ∈ L(B,F ), j(ιF (T)) = ιE (i(T)); (ii) for all m ∈ M(B), j(ιB(m)) = ιA(m), where we identify M(B) ⊂ M(A). L(A2 ⊗ S2j, E2 ⊗ S2j) → L(A1 ⊗ S12 ⊗ S2j, E1 ⊗ S12 ⊗ S2j) ; T (cid:55)→ (cid:101)T, 6.3.8 Lemma. -- Let j = 1, 2. We have a canonical embedding where for T ∈ L(A2 ⊗ S2j, E2 ⊗ S2j) the operator (cid:101)T ∈ L(A1 ⊗ S12 ⊗ S2j, E1 ⊗ S12 ⊗ S2j) is defined by (cid:101)T(xa) = T(x)a for all x ∈ A2 ⊗ S2j and a ∈ A1 ⊗ S12 ⊗ S2j. Moreover, the image of L(A2 ⊗ S2j, E2 ⊗ S2j) → L(A1 ⊗ S12 ⊗ S2j, E1 ⊗ S12 ⊗ S2j) is {X ∈ L(A1 ⊗ S12 ⊗ S2j, E1 ⊗ S12 ⊗ S2j) ; X(A2 ⊗ S2j) ⊂ E2⊗S2j and X∗(E2 ⊗ S2j) ⊂ A2 ⊗ S2j}. (cid:78) Proof. This follows from 6.3.6 with A := A1 ⊗ S12 ⊗ S2j, B := A2 ⊗ S2j, E := E1 ⊗ S12 ⊗ S2j and F := E2 ⊗ S2j ⊂ L(A1 ⊗ S12 ⊗ S2j, E1 ⊗ S12 ⊗ S2j). The assumptions of 6.3.6 are satisfied in this case in virtue of 5.2.5 1 and 6.3.2. 6.3.9 Notation. -- Let idE1 ⊗ δ2 be the unique linear extension of idE1 ⊗ δ2 such that (idE1 ⊗ δ2 T ∈ L(A1 ⊗ S12, E1 ⊗ S12). 12 : L(A1 ⊗ S12, E1 ⊗ S12) → L(A1 ⊗ S12 ⊗ S22, E1 ⊗ S12 ⊗ S22) 12)(T)(idA1 ⊗ δ2 12 : E1 ⊗ S12 → L(A1 ⊗ S12 ⊗ S22, E1 ⊗ S12 ⊗ S22) 12)(Tx) for all x ∈ M(A1 ⊗ S12) and (cid:78) 12)(x) = (idE1 ⊗ δ2 48 J. CRESPO 6.3.10 Proposition-Definition. -- There exists a unique linear map δE2 : E2 → L(A2 ⊗ S22, E2 ⊗ S22) satisfying the relation [δE2 (ξ)a]b = (idE1 ⊗ δ2 b ∈ A1 ⊗ S12 ⊗ S22. Proof. Let us prove the inclusion (idE1 ⊗ δ2 6.3.3 2 that ιE1⊗S12 (E2) ⊂ J2. Fix ξ ∈ E2 and x ∈ A2 ⊗ S22. We have 12)(ξ)(ab) for all ξ ∈ E2, a ∈ A2 ⊗ S22 and (cid:78) 12)(E2)(A2 ⊗ S22) ⊂ E2 ⊗ S22. It follows from ιE1⊗S12⊗S22 ((idE1 ⊗ δ2 12)(ξ)x) = (idJ1 ⊗ δ2 12)(ιE1⊗S12 (ξ))ιA1⊗S12⊗S22 (x) = δJ2 (ιE1⊗S12 (ξ))ιA2⊗S22 (x) ∈ J2 ⊗ S22. (2) E1 12)(E2)(A2 ⊗ S22) ⊂ E2 ⊗ S22. The inclusion (idE1 ⊗ δ2 As in the proof of 6.3.4, ιE1⊗S12⊗S22 ((idE1 ⊗ δ2 12)(ξ)x) is the norm limit of finite sums of (η) ⊗ s) with η ∈ E1, ω ∈ B(H21)∗ and elements of the form ιE1⊗S12⊗S22 ((idE1⊗S12 ⊗ ω)δ 12)(ξ)x ∈ E2 ⊗ S22 since ιE1⊗S12⊗S22 is isometric. Therefore, we have s ∈ S22. Hence, (idE1 ⊗ δ2 12)(E2)∗(E2 ⊗ S22) ⊂ A2 ⊗ S22 (idE1 ⊗ δ2 is obtained by a similar argument. Then, the existence and uniqueness of the operator δE2 (ξ) ∈ L(A2 ⊗ S22, E2 ⊗ S22) follows as an application of 6.3.8 with j = 2. It is clear that the map δE2 : E2 → L(A2 ⊗ S22, E2 ⊗ S22) is linear. In the following, we prove that δE2 is a continuous action of G2 on E2. We also show that the induction procedure for equivariant Hilbert modules is equivalent to that of §4.3 [2]. 6.3.11 Notations. -- Let e1,1 := ιK(E1)(1E1 ) ∈ M(J1) and e2,1 := ιA1 (1A1 ) ∈ M(J1), where we identify M(J1) = L(E1 ⊕ A1). Let (J2, δJ2, e1,2, e2,2) be the induced linking G2- C*-algebra, with el,2 := el,1 ⊗ 1S12 ∈ M(J2) for l = 1, 2 (cf. 4.14 [2]). Consider e2,2J2e2,2 and e1,2J2e2,2 endowed with their structure of G2-C*-algebra and G2-equivariant Hilbert ιA1 : A2 → J2 ; x (cid:55)→ (ιA1 ⊗ idS12 )(x) e2,2J2e2,2-module [3]. Recall that the morphism IndG2 induces a G2-equivariant *-isomormorphism A2 → e2,2J2e2,2 (cf. 4.17, 4.18 [2]). G1 (cid:78) 6.3.12 Proposition. -- We use the above notations. (i) The map δE2 : E2 → L(A2 ⊗ S22, E2 ⊗ S22) is a continuous action of G2 on E2. (ii) There exists a unique bounded linear map IndG2 G1 ιE1 : E2 → J2 such that IndG2 G1 ιE1 ((idE1⊗S12 ⊗ ω)δ (ξ)) = (idJ1⊗S12 ⊗ ω)δ for all ξ ∈ E1 and ω ∈ B(H21)∗. Moreover, we have IndG2 G1 IndG2 G1 over the G2-equivariant *-isomorphism A2 → e2,2J2e2,2 ; a (cid:55)→ IndG2 G1 ιE1 induces a G2-equivariant unitary equivalence E2 → e1,2J2e2,2 ; ξ (cid:55)→ IndG2 (2) J1 ιE1 (E2) = e1,2J2e2,2 and ιE1 (ξ) (ιE1 (ξ)), ιA1 (a). (2) E1 G1 (iii) There exists a unique *-homomorphism τ : K(E2 ⊕ A2) → J2 such that τ ◦ ιE2 = IndG2 ιE1 G1 and τ ◦ ιA2 = IndG2 G1 ιA1. Moreover, τ is an isomorphism of linking G2-C*-algebras. (iv) If T ∈ IndG2 G1 have T ◦ η ∈ E2. Moreover, for all T ∈ IndG2 G1 More precisely, the map IndG2 G1 *-isomorphism. (K(E1)) ⊂ L(E1 ⊗ S12) and η ∈ E2 ⊂ L(A1 ⊗ S12, E1 ⊗ S12), then we (K(E1)), we have [η (cid:55)→ T ◦ η] ∈ K(E2). (K(E1)) → K(E2) ; T (cid:55)→ [η (cid:55)→ T ◦ η] is a G2-equivariant (cid:78) Proof. Let us denote B := e2,2J2e2,2 and F := e1,2J2e2,2 for short. (i)-(ii) We have ιE1⊗S12 (E2) ⊂ J2 (cf. 6.3.3). Let IndG2 follows from the formulas δ (el,1) = el,2 ⊗ 1S21 for l = 1, 2 (4.14 [2]) that ιE1 := ιE1⊗S12 G1 (2) J1 (cid:22)E2: E2 → J2. It also e1,2J2e2,2 = [(idJ1⊗S12 ⊗ ω)δ = [(idJ1⊗S12 ⊗ ω)δ = IndG2 G1 ιE1 (E2). (e1,1xe2,1) ; x ∈ J1, ω ∈ B(H21)∗] (ιE1 (ξ)) ; ξ ∈ E1, ω ∈ B(H21)∗] (2) J1 (2) J1 ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS 49 Let ξ, η ∈ E2. Since (cid:104)ξ, η(cid:105) ∈ A2, we have (cf. 2.3.2 3) (cid:104)IndG2 G1 ιE1 (η)(cid:105) = ιE1⊗S12 (ξ)∗ ιE1 (ξ), IndG2 G1 ιE1⊗S12 (η) = ιA1⊗S12 ((cid:104)ξ, η(cid:105)) = IndG2 ιA1 ((cid:104)ξ, η(cid:105)). ιA1 (a) for all a ∈ A2 and ξ ∈ E2 (cf. 2.3.2 1). ιE1 (ξ)IndG2 ιE1 (ξ) = IndG2 We also have IndG2 G1 G1 G1 The map Φ : E2 → F ; ξ (cid:55)→ IndG2 ιE1 (ξ) is a unitary equivalence of Hilbert modules over G1 the *-isomorphism φ : A2 → B ; a (cid:55)→ IndG2 G1 Let us prove that (Φ ⊗ idS22 ) ◦ δE2 = δF ◦ Φ. It is immediately verified that for ξ ∈ E1 ⊗ S12, the formula (ιE1⊗S12 ⊗ idS22 )(idE1 ⊗ δ2 12)(ιE1⊗S12 (ξ)) holds true. Let us fix ξ ∈ L(A1 ⊗ S12, E1 ⊗ S12). For all a ∈ A1 ⊗ S12 and x ∈ A1 ⊗ S12 ⊗ S22, we have ιE1⊗S12⊗S22 ((idE1 ⊗ δ2 12)(ξ))ιA1⊗S12⊗S22 ((idA1 ⊗ δ2 12)(ξ) = (idJ1 ⊗ δ2 ιA1 (a). G1 12)(a)x) = (idJ1 ⊗ δ2 12)(ιE1⊗S12 (ξa))ιA1⊗S12⊗S22 (x) 12)(ιE1⊗S12 (ξa)) = (idJ1 ⊗ δ2 12)(ιE1⊗S12 (ξ))ιA1⊗S12⊗S22 ((idA1 ⊗ δ2 12)(a)). Hence, and (idJ1 ⊗ δ2 ιE1⊗S12⊗S22 ((idE1 ⊗ δ2 12)(ξ))ιA1⊗S12⊗S22 ((idA1 ⊗ δ2 12)(a)x) = (idJ1 ⊗ δ2 12)(ιE1⊗S12 (ξ))ιA1⊗S12⊗S22 ((idA1 ⊗ δ2 12)(a)x). 12)(ξ))ιA1⊗S12⊗S22 (x) = (idJ1 ⊗ δ2 12. Let us fix ξ ∈ E2. For all x ∈ A2 ⊗ S22 and y ∈ A1 ⊗ S12 ⊗ S21 we have Thus, ιE1⊗S12⊗S22 ((idE1 ⊗ δ2 12)(ιE1⊗S12 (ξ))ιA1⊗S12⊗S22 (x) for all ξ ∈ L(A1 ⊗ S12, E1 ⊗ S12) and a ∈ A1 ⊗ S12 ⊗ S22 in virtue of the non-degeneracy of idA1 ⊗ δ2 [(Φ ⊗ idS22 )δE2 (ξ)(φ ⊗ idS22 )(x)]ιA1⊗S12⊗S22 (y) = ιE1⊗S12⊗S22 ((δE2 (ξ)x)y) = ιE1⊗S12⊗S22 ((idE1 ⊗ δ2 = (idJ1 ⊗ δ2 12)(ιE1⊗S12 (ξ))ιA1⊗S12⊗S22 (xy) = [δF(Φ(ξ))(φ ⊗ idS22 )(x)]ιA1⊗S12⊗S22 (y), In particular, by which also holds for all y ∈ M(A1 ⊗ S12 ⊗ S22) by strict continuity. applying this formula for y ∈ A2 ⊗ S22, we have then proved that 12)(ξ)(xy)) (Φ ⊗ idS22 )(δE2 (ξ))(φ ⊗ idS22 )(x) = δF(Φ(ξ))(φ ⊗ idS22 )(x) for all x ∈ A2 ⊗ S22. Hence, (Φ ⊗ idS22 )(δE2 (ξ)) = δF(Φ(ξ)) for all ξ ∈ E2. This proves that δE2 is a continuous action of G2 on E2 and Φ is G2-equivariant. (iii) There exists a unique unital faithful *-homomorphism j : L(E2 ⊕ A2) → L((E1 ⊗ S12) ⊕ (A1 ⊗ S12)) such that j(x)(ηa) = (xη)a for all x ∈ L(E2 ⊕ A2), η ∈ E2 ⊕ A2 and a ∈ A1 ⊗ S12 (6.3.8 (ii), with A := A1 ⊗ S12, B := A2, E := E1 ⊗ S12 and F := E2). Now, it should be noted that we have the following canonical identifications J2 ⊂ M(J1 ⊗ S12) = L((E1 ⊕ A1) ⊗ S12) = L((E1 ⊗ S12) ⊕ (A1 ⊗ S12)). ιE1 (ξ) for all ξ ∈ E2 and j(ιA2 (b)) = IndG2 ιA1 (b) for all b ∈ A2 We have j(ιE2 (ξ)) = IndG2 G1 (cf. 6.3.7). In particular, we have j(K(E2 ⊕ A2)) ⊂ J2. Let τ := j(cid:22)K(E2⊕A2): K(E2 ⊕ A2) → J2. Since J2 is generated by e1,2J2e2,2 and e2,2J2e2,2 as a C*-algebra, τ has dense range (cf. (ii)); moreover, τ is also isometric (faithful), therefore τ is surjective. Thus, we have proved that τ is a *-isomorphism. The G2-equivariance of τ is derived from straightforward computations. (iv) Consider the G2-equivariant *-isomorphism G1 ϕ : IndG2 G1 (K(E1)) → e1,2J2e1,2 ; k (cid:55)→ IndG2 G1 ιK(E1)(k) 50 J. CRESPO (cf. 4.18 [2], note that K(F) = e1,2J2e1,2). By statement (ii), τ induces by restriction a G2- equivariant *-isomorphism τ : f1,2K(E2 ⊕ A2) f1,2 → e1,2J2e1,2, where f1,2 := ιE2 (1E2 ) and f2,2 := ιA2 (1A2 ). We have an isomorphism ψ : K(E2) → f1,2K(E2 ⊕ A2) f1,2 ; k (cid:55)→ ιK(E2)(k) of G2-C*-algebras. Hence, χ := ψ−1 ◦ τ−1 ◦ ϕ : IndG2 (K(E1)) → K(E2) is an isomorphism G1 of G2-C*-algebras. It is clear that χ(T)ξ = T ◦ ξ for all T ∈ IndG2 (K(E1)) ⊂ L(E1 ⊗ S12) G1 and ξ ∈ E2 ⊂ L(A1 ⊗ S12, E1 ⊗ S12). 6.3.13 Proposition-Definition. -- Let us fix some notations. Consider: • two G1-C*-algebras A1 and B1; • two G1-equivariant Hilbert modules E1 and F1 over A1 and B1 respectively; • a G1-equivariant unitary equivalence Φ1 : E1 → F1 over a G1-equivariant *-isomorphism φ1 : A1 → B1. Denote by: (A1) and B2 := IndG2 G1 (B1) the induced G2-C*-algebras; (φ1) : A2 → B2 the induced G2-equivariant *-isomorphism; • A2 := IndG2 G1 • IndG2 G1 • E2 := IndG2 G1 A2 and B2 respectively; (E1) and F2 := IndG2 G1 (F1) the induced G2-equivariant Hilbert modules over • Φ1 ⊗ idS12 : L(A1 ⊗ S12, E1 ⊗ S12) → L(B1 ⊗ S12, F1 ⊗ S12) the unique linear map such that (Φ1 ⊗ idS12 )(T)(φ1 ⊗ idS12 )(x) = (Φ1 ⊗ idS12 )(Tx) for all L(A1 ⊗ S12, E1 ⊗ S12) and x ∈ A1 ⊗ S12 (cf. 8.3.6). (Φ1)((idE1⊗S12 ⊗ ω)δ (Φ1) := (Φ1 ⊗ idS12 ) (cid:22)E2: E2 → F2 is a Then, (Φ1 ⊗ idS12 )(E2) ⊂ F2 and the map IndG2 G1 (φ1) : A2 → B2. Moreover, for all ξ ∈ E1 and G2-equivariant unitary equivalence over IndG2 G1 ω ∈ B(H21)∗ we have IndG2 (cid:78) G1 Proof. Denote by J1 := K(E1 ⊕ A1) and K1 := K(F1 ⊕ B1) the linking G1-C*-algebras, whose linking structures are respectively defined by: e1,1 := ιE1 (1E1 ), e2,1 := ιA1 (1A1 ); f1,1 := ιF1 (1F1 ), f2,1 := ιB1 (1B1 ). We also denote by (J2, δJ2, e1,2, e2,2) and (K2, δK2, f1,2, f2,2) the induced linking G2-C*-algebras, where el,2 := el,1 ⊗ 1S12 and fl,2 := fl,1 ⊗ 1S12 for l = 1, 2 (cf. 4.14 [2]). There exists a unique *-isomorphism τ1 : J1 → K1 such that τ1 ◦ ιE1 = ιF1 ◦ Φ1 and τ1 ◦ ιA1 = ιB1 ◦ φ1 (cf. 8.3.5 and 6.1.18). We then denote by (ξ)) = (idF1⊗S12 ⊗ ω)δ (2) F1 (Φ1ξ). (2) E1 τ2 := IndG2 G1 τ1 : J2 → K2 the induced morphism. Since τ2 is an isomorphism of linking G2-C*-algebras, it induces a G2-equivariant unitary equivalence Ψ : e1,2J2e2,2 → f1,2K2 f2,2 over the isomorphism of G2-C*-algebras ψ : e2,2J2e2,2 → f2,2K2 f2,2. Since τ1 ◦ ιA1 = ιB1 ◦ φ1, we have τ2 ◦ IndG2 G1 ιA1 = IndG2 G1 ιB1 ◦ φ2. Therefore, by composition of G2-equivariant unitary equivalences (cf. 8.3.2 2) and by applying 6.3.12, we obtain a G2-equivariant φ2-compatible unitary operator Φ2 : E2 → F2. By a straightforward computation, we show that Φ2 = (Φ1 ⊗ idS12 )(cid:22)E2. By exchanging the roles of G1 and G2, we define as above an induction procedure for G2-equivariant Hilbert modules. In the following, we investigate the composition of IndG2 G1 algebra and E1 a G1-equivariant Hilbert A1-module. Denote by: . Let A1 be a G1-C*- and IndG1 G2 • A2 := IndG2 G1 (A1) and E2 = IndG2 G1 (E1) ⊂ L(A1 ⊗ S12, E1 ⊗ S12) the induced G2-C*- algebra and the induced G2-equivariant Hilbert A2-module; ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS 51 • C = IndG1 G2 (A2) and F := IndG1 G2 (E2) ⊂ L(A2 ⊗ S21, E2 ⊗ S21) the induced G1-C*- algebra and the induced G1-equivariant Hilbert C-module. 6.3.14 Proposition. -- With the above notations and hypotheses, we have the following statements: 1. there exists a unique map Π1 : E1 → F such that (Π1(ξ)x)a = δ (2) E1 (ξ)(xa), for all ξ ∈ E1, x ∈ A2 ⊗ S21 and a ∈ A1 ⊗ S12 ⊗ S21; (ξ), δ2 E1 (a) and (cid:104)δ2 E1 (η)(cid:105) = δ2 (ξa) = δ2 E1 A1 (E1)(1A2 ⊗ S21)] = E2 ⊗ S21 = [(1E2 ⊗ S21)δ2 E1 moreover, Π1 is a G1-equivariant unitary equivalence over the G1-equivariant *-isomorphism π1 : A1 → C ; a (cid:55)→ δ : E1 → (cid:102)M(E2 ⊗ S21) ; ξ (cid:55)→ Π1(ξ) is a well-defined linear map such that: 2. δ2 E1 (i) δ2 E1 a ∈ A1, (ii) [δ2 E1 ((cid:104)ξ, η(cid:105)) for all ξ, η ∈ E1 and (cid:78) Proof. 1. The existence and uniqueness of Π1 is an immediate application of 6.3.8 with j = 1 and the proof is very similar to that of 6.3.10. The fact that Π1 is a G1-equivariant takes its values in (cid:102)M(E2 ⊗ S21) are proved by unitary equivalence over π1 is a straightforward consequence of 6.3.12 (ii), (iii) and 5.2.6 2. 2. Statement (ii) and the fact that δ2 E1 combining the formulas [F(1A2 ⊗ S21)] = E2 ⊗ S21 = [(1E2 ⊗ S21)F] (cf. 6.3.2) with the fact that Π1 is bijective. Statement (i) follows from the compatibility of Π1 with π1. We have proved the following result: 6.3.15 Theorem. -- Let G1 and G2 be two monoidally equivalent regular locally compact quantum groups. The map IndG2 G1 : (E1, δE1 ) (cid:55)→ (E2 := IndG2 (E1), δE2 : ξ ∈ E2 (cid:55)→ [x ∈ A2 ⊗ S22 (cid:55)→ (idE1 ⊗ δ2 12)(ξ)x]), (E1)]. (2) A1 (a); (ξ)δ2 A1 G1 where E1 is a Hilbert module over the G1-C*-algebra A1 and A2 = IndG2 (A1) denotes the induced G1 G2-C*-algebra, is a one-to-one correspondence up to unitary equivalence. The inverse map, up to unitary equivalence, is IndG1 G2 : (F2, δF2 ) (cid:55)→ (F1 := IndG1 G2 (F2), δF1 : ξ ∈ F1 (cid:55)→ [x ∈ B1 ⊗ S11 (cid:55)→ (idF2 ⊗ δ1 21)(ξ)x]), where F2 is a Hilbert module over the G2-C*-algebra B2 and B1 = IndG2 (B2) denotes the induced G1 (cid:78) G1-C*-algebra. Proof. This is a consequence of Propositions 6.3.14, 6.3.13 and the corresponding results obtained by exchanging the roles of G1 and G2. Let B1 be a G1-C*-algebra. Let us denote by B2 := IndG2 G1 Let δk 6.3.16 Notations. -- Let E1 be a G1-equivariant Hilbert B1-module. Let us denote by F2 = IndG2 (F1) the induced G2-equivariant Hilbert B2-module. We have four linear maps G1 Bj : Bj → M(Bk ⊗ Skj) for j, k = 1, 2 be the *-homomorphisms defined in 5.2.7. (B1) the induced G2-C*-algebra. Fj : Fj → L(Bk ⊗ Skj, Fk ⊗ Skj), δk for j, k = 1, 2, defined as follows: • δ1 F1 • δ2 F1 := δF1 and δ2 F2 : F1 → L(B2 ⊗ S21, F2 ⊗ S21) is the unique linear map such that := δF2; (δ2 F1 (ξ)x)b = δ (2) F1 (ξ)(xb) for all ξ ∈ F1, x ∈ B2 ⊗ S21 and b ∈ B1 ⊗ S12 ⊗ S22, where δ (2) F1 (cf. 6.3.14); (ξ) := (idE1 ⊗ δ2 11)δF1 (ξ) 52 J. CRESPO • δ1 F2 x ∈ IndG1 G2 : F2 → L(B1 ⊗ S12, F1 ⊗ S12) is the unique linear map such that for all ξ ∈ F2, (B2) ⊗ S12 and y ∈ B2 ⊗ S21 ⊗ S12, we have (ξ))x]y = δ [(Π1 ⊗ idS12 )(δ1 F2 (1) F2 (ξ)(xy), 6.3.17 Lemma. -- For all j, k, l = 1, 2, we have the following statements: 22)δF2 (ξ) and Π1 : F1 → IndG1 G2 (F2) (cf. 6.3.14 1). (cid:78) where δ (1) F2 (ξ) := (idF1 ⊗ δ1 (Fj) ⊂ (cid:102)M(Fk ⊗ Skj); (ξ)δk Bj (b) and (cid:104)δk Fj (η)(cid:105) = δk ((cid:104)ξ, η(cid:105)) for all ξ, η ∈ Fj and b ∈ Bj; (ξb) = δk Fj Bj (Fj)(1Bk ⊗ Skj)] = Fk ⊗ Skj = [(1Fk ⊗ Skj)δk (Fj)]; Fj ⊗ idSkj (resp. idFl ⊗ δk lj) extends uniquely to a linear map from L(Bk ⊗ Skj, Ek ⊗ Skj) (ξ), δk Fj 4. δl Fk to L(Bl ⊗ Slk ⊗ Skj, El ⊗ Slk ⊗ Skj) such that 1. δk Fj 2. δk Fj 3. [δk Fj (δl Fk ⊗ idSkj )(T)(δl Bk ⊗ idSkj )(x) = (δl Fk (resp. (idFl ⊗ δk lj)(T)(idBl ⊗ δk lj)(x) = idFl ⊗ δk ⊗ idSkj )(Tx) lj)(Tx)) for all T ∈ L(Bk ⊗ Skj, Ek ⊗ Skj) and x ∈ Bk ⊗ Skj; . (cid:78) 5. (δl Fk (C), K = IndG1 G2 = (idFl ⊗ δk lj)δl Fj (B2), D := IndG2 G1 ⊗ idSkj )δk Fj Proof. Let C := IndG1 (F2) and L = IndG2 (K). There exists a unique G2-equivariant unitary equivalence Π2 : F2 → L (6.3.14 1, after exchanging G1 G2 the roles of G1 and G2) over the G2-equivariant *-isomorphism π2 : B2 → D. 1. This statement will follow straightforwardly from the third one. 2. This statement has already been proved for (j, k) = (1, 1) (by definition), (j, k) = (2, 2) (cf. 6.3.12 (i)) and for (j, k) = (1, 2) (cf. 6.3.14). Moreover, the case (j, k) = (2, 1) follows from the formulas δ1 F2 3. This statement is true by assumption for (j, k) = (1, 1), for (j, k) = (2, 2) (cf. 6.3.12 (i)) and for (j, k) = (1, 2) (cf. 6.3.14 2 (ii)). By 6.3.2 and 6.3.14, we have [L(C ⊗ S12)] = K ⊗ S12, L = Π2(F2) and K = Π1(F1). Therefore, we have 1 ⊗ idS12 )Π2 and δ1 −1 1 ⊗ idS12 )π2. = (Π−1 = (π B2 [δ1 F2 (F2)(B1 ⊗ S12)] = [(Π−1 = [(Π−1 = F1 ⊗ S12. 1 ⊗ idS12 )(L)(π 1 ⊗ idS12 )(L(D ⊗ S12))] −1 1 ⊗ idS12 )(D ⊗ S12)] It then follows from the second statement and the fact that [δ1 B2 (F2)(1B1 ⊗ S12)] = F1 ⊗ S12, which is statement 3 for (j, k) = (2, 1). that [δ1 F2 4. Let j, k, l = 1, 2. The uniqueness of the extensions is obvious by the non-degeneracy of δk lj lj : L(Bk ⊗ Skj, Ek ⊗ Skj) → L(Bl ⊗ Slk ⊗ Skj, El ⊗ Slk ⊗ Skj) . The linear map idFl ⊗ δk and δl Bk is defined by (B2)(1B1 ⊗ S12)] = B1 ⊗ S12 (idFl ⊗ δk lj)(T) := T ⊗idBl (Bl ⊗ Slk), Fl ⊗ Slk) such that ⊗δk lj 1, where we use the identifications (6.13) and (6.17). As in 2.4 (a) [3], there exists a unique unitary Vl k ∈ L(Fk ⊗δl for all T ∈ L(Bk ⊗ Skj, Ek ⊗ Skj), Bk k (ξ ⊗δl Vl Bk (ξ)x, for all ξ ∈ Fk and x ∈ Bl ⊗ Slk. x) = δl Fk ⊗ idSkj : L(Bk ⊗ Skj, Ek ⊗ Skj) → L(Bl ⊗ Slk ⊗ Skj, El ⊗ Slk ⊗ Skj) is 1) for all T ∈ L(Bk ⊗ Skj, Ek ⊗ Skj), k ⊗C 1)(T ⊗δl ⊗ idSkj Bk = (idFl ⊗ δk lj)δl Fj is derived from 6.3.12 after long but ⊗ idSkj )(T) := ( Vl The linear extension δl Fk defined by (δl Fk up to the identifications (6.12) and (6.16). ⊗ idSkj )δk 5. The formula (δl Fj Fk straightforward computations. ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS 53 Let us consider the C*-algebra B := B1 ⊕ B2 endowed with the continuous action (βB, δB) (cf. 5.2.9). 6.3.18 Proposition. -- Let F1 be a G1-equivariant Hilbert B1-module. Let F2 := IndG2 (F1) be the induced G2-equivariant Hilbert B2-module. Consider the Hilbert B-module F := F1 ⊕ F2. G1 j : L(Bk ⊗ Skj, Fk ⊗ Skj) → L(B ⊗ S, F ⊗ S) the linear extension of the canonical Denote by Πk injection Fk ⊗ Skj → F ⊗ S. Let us consider the linear maps δF : F → L(B ⊗ S, F ⊗ S) and βF : C2 → L(F) defined by: (cid:18)λ 0 (cid:19) 0 µ δF(ξ) := ∑ k,j=1,2 j ◦ δk Πk Fj (ξj), ξ = (ξ1, ξ2) ∈ F; βF(λ, µ) := (λ, µ) ∈ C2. , Then, the triple (F, βF, δF) is a G-equivariant Hilbert B-module. (cid:78) Proof. Let us consider J1 := K(F1 ⊕ B1) (resp. K(F2 ⊕ B2)) the linking G1-C*-algebra (resp. linking G2-C*-algebra) associated with F1 (resp. F2). Let J2 := IndG2 (J1) be the induced G1 G2-C*-algebra. Let us consider J := J1 ⊕ J2 endowed with the continuous action (βJ, δJ) of G (see above). We denote L := K(F ⊕ B) the linking C*-algebra associated with F and we identify L = J1 ⊕ K(F2 ⊕ B2). We have an isomorphism of linking C*-algebras f := idJ1 ⊕ τ : L → J (6.3.12 (ii)). Let (βL, δL) be the continuous action of G on L obtained by transport of structure, i.e.: δL(x) := ( f −1 ⊗ idS)δJ( f (x)), x ∈ L; βL(n) := f −1(βJ(n)), n ∈ C2. By straightforward computations, we show that (βL, δL) is compatible with (βB, δB) (cf. 6.1.8) and we prove that δL(ιF(ξ)) = ιF⊗S(δF(ξ)), for all ξ ∈ F. Therefore, the result follows from 6.1.11 a) and 6.1.21. 6.3.19 Proposition. -- Let (E, βE, δE) be a G-equivariant Hilbert A-module. In the following, we use the notations of 6.2.5. Let j, k = 1, 2 with j (cid:54)= k. Let and (cid:101)Ej := Ind Gj-equivariant unitary equivalence over (cid:101)πj : Aj → (cid:101)Aj (cf. 5.2.8). If ξ ∈ Ej, then we have δk Ej Proof. We have Ej = [(idEj ⊗ ω)δ ξ ∈ Ej and ω ∈ B(Hjk)∗ we have (cid:101)Aj := Ind (ξ) ∈(cid:101)Ej ⊂ (cid:102)M(Ek ⊗ Skj) and the map (cid:101)Πj : Ej →(cid:101)Ej ; ξ (cid:55)→ δk (ξ) is a (cid:78) (ξ) ; ω ∈ B(Hjk)∗, ξ ∈ Ek] (cf. 6.2.5 (iv)) and for all (Ak, δk Ak (Ek, δk Ek Gj Gk Gj Gk j Ek ). Ej ) δk Ej (idEj ⊗ ω)δ j Ek (ξ) = (idEk⊗Skj ⊗ ω)(δk Ej (ξ) := (idEj ⊗ δk j Ek (ξ) = (idEk⊗Skj ⊗ ω)δ ⊗ idSjk )δ (ξ). As a consequence, statement 1 is proved (k) Ej (ξ) j Ej (k) Ej jj)δ (cf. 6.2.5 (v)), where δ as well as the surjectivity of (cid:101)Πj. The fact that (cid:101)Πj is a Gj-equivariant (cid:101)πj-compatible unitary operator is just a restatement of 6.2.5 (iii) and (idEk ⊗ δ (v)). 6.3.20 Theorem. -- Let GG1,G2 be a colinking measured quantum groupoid between two regular monoidally equivalent locally compact quantum groups G1 and G2. Let j = 1, 2. The map (E, βE, δE) (cid:55)→ (Ej, δ ) is a one-to-one correspondence up to unitary equivalence (cf. 6.2.5 and 6.2.6 1). The inverse map, up to unitary equivalence, is (Fj, δFj ) (cid:55)→ (F, βF, δF) (cf. 6.3.18, 6.3.13 and (cid:78) 6.2.6 2). Proof. Let A be a G-C*-algebra and E a G-equivariant Hilbert A-module. Let us use all the notations introduced in §6.2. Let us denote: ⊗ idSjj )δ = (δk Ej j kj)δk Ej (6.2.5 j Ej j Ej ), (B2, δB2 ) := IndG2 (B1, δB1 ) := (A1, δ1 A1 G1 ), (F2, δF2 ) := IndG2 (F1, δF1 ) := (E1, δ1 E1 G1 (B1, δB1 ); (F1, δF1 ). 54 J. CRESPO Let us endow the C*-algebra B := B1 ⊕ B2 with the continuous action (βB, δB) of G and F := F1 ⊕ F2 with the structure of G-equivariant Hilbert B-module (βF, δF) (cf. 5.2.9, ψA(a) := (qA,1a,(cid:101)π2(qA,2a)) (cf. 4.10 [2]). Then, we consider the map Ψ : E → F given by 6.3.18). Let ψA : A → B the canonical G-equivariant *-isomorphism defined for all a ∈ A by Ψ(ξ) := (qE,1ξ,(cid:101)Π2(qE,2ξ)), for all ξ ∈ E. It is clear from 6.3.19 that Ψ is a ψA-compatible unitary operator. Let us consider the G-C*-algebras K := K(E ⊕ A) and L := K(F ⊕ B). Let f : K → L be the associated isomorphism of linking C*-algebras (cf. 8.3.5). In virtue of 6.1.18, it only remains to prove that f is G-equivariant. We also consider the G1-C*-algebra J1 := K(F1 ⊕ B1) and the induced G2-C*-algebra J2 := IndG2 (J1). We recall that we have a canonical isomorphism G1 τ : K(F2 ⊕ B2) → J2 (cf. 6.3.12 (ii)). Let us endow the C*-algebra J := J1 ⊕ J2 with the continuous action (βJ, δJ) of G. Therefore, it amounts to proving that the *-isomorphism (idJ1 ⊕ τ) f : K → J is G-equivariant (we identify L = J1 ⊕ K(F2 ⊕ B2)). We apply the notations of §6.2 to the G-C*-algebra K and identify Kj := qK,jK = K(Ej ⊕ Aj) for j = 1, 2. Let us consider as above (by exchanging the roles of A and K) the G-equivariant *-isomorphism ψK : K → J. By evaluating on elements of the form ιE(ξ) for ξ ∈ E and ιA(a) for a ∈ A, it is staightforward to see that (idJ1 ⊕ τ) f = ψK. 7 takesaki-takai duality and equivariant morita equivalence In this section, we fix a measured quantum groupoid G = (N, M, α, β, ∆, T, T(cid:48), ) on the 1(cid:54)l(cid:54)k Mnl (C) and we use all the notations introduced in §§ finite-dimensional basis N =(cid:76) 3.1 and 3.2. We will use the notations and results of §§5.1, 5.3 and 6.1. equivariant hilbert bimodules and morita equivalence. In this paragraph, we introduce the notion of equivariant representation of a G-C*-algebra on a Hilbert module acted upon by G. We then introduce the notion of equivariant Morita equivalence. 7.1 Notation. -- Let A and B be C*-algebras. Let E be a Hilbert B-module. If γ : A → L(E) extend γ ⊗ idS to a *-homomorphism γ ⊗ idS : (cid:102)M(A ⊗ S) → L(E ⊗ S) (cf. §1). is a *-homomorphism then, up to the identification M(K(E) ⊗ S) = L(E ⊗ S), we can (cid:78) As in 2.9 [3], we have: 7.2 Definition. -- Let A and B be two G-C*-algebras, E a Hilbert B-module, (βE, δE) an action of G on E and γ : A → L(E) a *-representation. We say that γ is G-equivariant if we have: 1. δE(γ(a)ξ) = (γ ⊗ idS)(δA(a)) ◦ δE(ξ), for all a ∈ A and ξ ∈ E; 2. βE(no) ◦ γ(a) = γ(βA(no)a), for all n ∈ N and a ∈ A. (cid:78) 7.3 Remarks. -- 1. Provided that the second condition in the above definition is verified, the first condition is equivalent to: V(γ(a) ⊗δB 1) V∗ = (γ ⊗ idS)δA(a), (7.1) where V ∈ L(E ⊗δB (B ⊗ S), E ⊗ S) denotes the isometry defined in 6.1.5 a). Indeed, we can interpret it as follows: V(γ(a) ⊗δB 1) = (γ ⊗ idS)(δA(a)) V, for all a ∈ A. Moreover, for all a ∈ A we have for all a ∈ A, (γ ⊗ idS)(δA(a)) V V∗ = (γ ⊗ idS)(δA(a))qβEα = (γ ⊗ idS)(δA(a)qβAα) = (γ ⊗ idS)δA(a). Hence, ( V(γ(a) ⊗δB 1) = (γ ⊗ idS)(δA(a)) V ⇔ V(γ(a) ⊗δB 1) V∗ = (γ ⊗ idS)δA(a)), for all a ∈ A. ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS 55 2. We recall that the action δK(E) of G on K(E) is defined by δK(E)(k) := V(k ⊗δB 1) V∗ for all k ∈ K(E). Hence, (7.1) can be restated as follows: δK(E)(γ(a)) = (γ⊗ idS)δA(a) for all a ∈ A. In particular, if γ is non-degenerate, then Definition 7.2 simply means that the *-homomorphism γ : A → M(K(E)) is G-equivariant (cf. 5.1.10). 3. If γ : A → L(E) is a non-degenerate *-representation such that δE(γ(a)ξ) = (γ ⊗ idS)(δA(a)) ◦ δE(ξ), for all a ∈ A and ξ ∈ E, then we have βE(no) ◦ γ(a) = γ(βA(no)a) for all n ∈ N and a ∈ A. Indeed, this will (cid:78) be inferred from 5.1.11 and the previous remark. 7.4 Definition. -- (cf. §6 [22]) Let A and B be two C*-algebras. An imprimitivity A-B- bimodule is an A-B-bimodule E, which is a full left Hilbert A-module for an A-valued inner product A(cid:104)·, ·(cid:105) and a full right Hilbert B-module for a B-valued inner product (cid:104)·, ·(cid:105)B such that A(cid:104)ξ, η(cid:105)ζ = ξ(cid:104)η, ζ(cid:105)B for all ξ, η, ζ ∈ E. (cid:78) 7.5 Remarks. -- Let A and B be two C*-algebras and E an imprimitivity A-B-bimodule. We recall that the norms defined by the inner products A(cid:104)·, ·(cid:105) on AE and (cid:104)·, ·(cid:105)B on EB coincide. We also recall that the left (resp. right) action of A (resp. B) on E defines a non-degenerate *-homomorphism γ : A → L(EB) (resp. ρ : B → L(AE)). (cid:78) 7.6 Definition. -- Let A and B be two G-C*-algebras. A G-equivariant imprimitivity A-B-bimodule is an imprimitivity A-B-bimodule E endowed with a continuous action of G on EB such that the left action γ : A → L(EB) is G-equivariant. (cid:78) 7.7 Examples. -- Let A and B be two G-C*-algebras. (i) B is a G-equivariant imprimitivity B-B-bimodule for the inner products given by B(cid:104)x, y(cid:105) := xy∗ and (cid:104)x, y(cid:105)B := x∗y for all x, y ∈ B. (ii) Let E be a G-equivariant Hilbert B-module. If E is full, then E is a G-equivariant imprimitivity K(E)-B-bimodule for the natural left action and the inner product given by K(E)(cid:104)ξ, η(cid:105) := θξ,η for all ξ, η ∈ E. Conversely, if E is a G-equivariant imprimitivity A-B-bimodule, then the the left action γ : A → L(EB) induces an isomorphism of G-C*-algebras A (cid:39) K(EB). (iii) Let (J, βJ, δJ, e1, e2) be a linking G-C*-algebra (cf. 6.1.22). Let A := e1Je1 and B := e2Je2 be the corner C*-algebras endowed with the continuous actions of G induced by (βJ, δJ). Let us endow E := e1Je2 with its structure of G-equivariant Hilbert B-module (cf. 6.1.23). Then, E is a G-equivariant imprimitivity A-B-module whose actions and inner products are defined as in (i). (iv) If E is a G-equivariant imprimitivity A-B-bimodule, then E∗ turns into a G-equivariant imprimitivity B-A-bimodule for the actions and inner products given by the following formulas: bξ∗a := (a∗ξb∗)∗, for ξ∗ ∈ E∗, a ∈ A and b ∈ B; B(cid:104)ξ∗, η∗(cid:105) := (cid:104)ξ, η(cid:105)B and (cid:104)ξ∗, η∗(cid:105)A := A(cid:104)ξ, η(cid:105), for ξ∗, η∗ ∈ E∗. (cid:78) 7.8 Proposition. -- Let A and B be G-C*-algebras. The following statements are equivalent: (i) there exists a G-equivariant imprimitivity A-B-bimodule; (ii) there exists a full G-equivariant Hilbert B-module E such that we have an isomorphism A (cid:39) K(E) of G-C*-algebras; (iii) there exists a linking G-C*-algebra (J, βJ, δJ, e1, e2) such that we have G-equivariant *- (cid:78) isomorphisms A (cid:39) e1Je1 and B (cid:39) e2Je2. Proof. This is a straightforward consequence of 7.7 (ii), (iii), 6.1.11 b) and 6.1.23. Now, we investigate the tensor product construction (cf. 2.10 [3] for the quantum group case). 56 J. CRESPO 7.9 Proposition. -- Let C (resp. B) be a G-C*-algebra. Let E1 (resp. E2) be a Hilbert module over C (resp. B) endowed with an action (βE1, δE1 ) (resp. (βE2, δE2 )) of G. Let γ2 : C → L(E2) be a G-equivariant *-representation. Consider the Hilbert B-module E := E1 ⊗γ2 We have ∆(ξ1, ξ2) ∈ (cid:102)M(E ⊗ S) for all ξ1 ∈ E1 and ξ2 ∈ E2. Let βE : No → L(E) be the ∆(ξ1, ξ2) := (δE1 (ξ1) ⊗(cid:101)γ2⊗idS 1) ◦ δE2 (ξ2), for ξ1 ∈ E1 and ξ2 ∈ E2. E2. Denote by *-homomorphism defined by βE(no) := βE1 (no) ⊗γ2 1, for all n ∈ N. There exists a unique map δE : E → (cid:102)M(E ⊗ S) defined by the formula δE(ξ1 ⊗γ2 ξ2) := ∆(ξ1, ξ2) for ξ1 ∈ E1 and ξ2 ∈ E2 such that the pair (βE, δE) is an action of G on E. (cid:78) The operator δE1 (ξ1) is considered here as an element of L((cid:101)C ⊗ S, E1 ⊗ S) ⊃ (cid:102)M(E1 ⊗ S). In particular, we have δE1 (ξ1) ⊗(cid:101)γ2⊗idS 1 ∈ L(E2 ⊗ S, E ⊗ S) since we use the identifications: ((cid:101)C ⊗ S) ⊗(cid:101)γ2⊗idS (E2 ⊗ S) = E2 ⊗ S, x ⊗(cid:101)γ2⊗idS η (cid:55)→ ((cid:101)γ2 ⊗ idS)(x)η; (7.2) (E1 ⊗ S) ⊗(cid:101)γ2⊗idS (E2 ⊗ S) = E ⊗ S, (ξ1 ⊗ s) ⊗(cid:101)γ2⊗idS (ξ2 ⊗ t) (cid:55)→ (ξ1 ⊗γ2 ξ2) ⊗ st. (7.3) reader to it for the proof of the fact that ∆(ξ1, ξ2) ∈ (cid:102)M(E ⊗ S) for all ξ1 ∈ E1 and Proof. The proof is basically the same as that of 2.10 [3]. For example, we refer the ξ2 ∈ E. Let V1 and V2 be the isometries associated with δE1 and δE2. Since V2 intertwines (cid:101)V2 ∈ L(E ⊗δB (B ⊗ S), E1 ⊗(γ2⊗idS)δC (E2 ⊗ S)) such that the left actions c (cid:55)→ γ2(c) ⊗δB 1 and (γ2 ⊗ idS)δC of C, there exists a unique isometry (cid:101)V2((ξ1 ⊗γ2 ξ2) ⊗δB x) = ξ1 ⊗(γ2⊗idS)δC V2(ξ2 ⊗δB x), for all ξ1 ∈ E1, ξ2 ∈ E2 and x ∈ B ⊗ S. Let us prove that (cid:101)V2 is a unitary. It amounts to proving that (cid:101)V2 is surjective. Since im( V2) = im(qβE2 α), we have im((cid:101)V2) = [ξ ⊗(γ2⊗idS)δC qβE2 αη ; ξ ∈ E1, η ∈ E2 ⊗ S]. Let ξ ∈ E1 and η ∈ E2 ⊗ S. Write ξ = ξ(cid:48)c with ξ(cid:48) ∈ E1 and c ∈ C. Since V (γ2 ⊗ idS)δC(c)qβE2 α = (γ2 ⊗ idS)δC(c) (cf. 7.3). Hence, V∗ 2 = qE2α, we have 2 ξ ⊗(γ2⊗idS)δC qβE2 αη = ξ (cid:48) ⊗(γ2⊗idS)δC (γ2 ⊗ idS)δC(c)qβE2 αη (cid:48) ⊗(γ2⊗idS)δC (γ2 ⊗ idS)δC(c)η = ξ = ξ ⊗(γ2⊗idS)δC η. Therefore we have shown that im((cid:101)V2) = E1 ⊗(γ2⊗idS)δC (E2 ⊗ S), which proves that (cid:101)V2 is unitary. Let us identify (E1 ⊗δC (C ⊗ S)) ⊗γ2⊗idS (E2 ⊗ S) → E1 ⊗(γ2⊗idS)δC (E2 ⊗ S) (ξ1 ⊗δC x) ⊗γ2⊗idS η (cid:55)→ ξ1 ⊗(γ2⊗idS)δC (γ2 ⊗ idS)(x)η and (E1 ⊗ S) ⊗γ2⊗idS (E2 ⊗ S) = E ⊗ S (cf. (7.3)). Let V := ( V1 ⊗γ2⊗idS 1)(cid:101)V2 ∈ L(E ⊗δB (B ⊗ S), E ⊗ S). 2 = 1, (cid:101)V 2 (cid:101)V It follows from the formulas (cid:101)V∗ 2(cid:101)V∗ 2 = 1, V∗ V∗ V = 1 and V V∗ = qβE1 α ⊗γ2⊗idS 1 = qβEα (by definition of βE). Let n ∈ N. On one hand, we have (cid:101)V2(βE(no) ⊗δB 1) = (βE1 (no) ⊗(γ2⊗idS)δC 1)(cid:101)V2 (by definition of βE and (cid:101)V2). On the other, we have 1 1 = 1 and V V 1 V∗ 1 = qβE1 α that ( V1 ⊗γ2⊗idS 1)(βE1 (no) ⊗(γ2⊗idS)δC 1) = ((1E1 ⊗ β(no)) ⊗γ2⊗idS 1)( V1 ⊗γ2⊗idS 1). ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS 57 Hence, we have proved that V(βE(no) ⊗δB 1) = (1 ⊗ β(no)) V for all n ∈ N. Exactly as in for the definition of Tξ). In particular, VTξ ∈ (cid:102)M(E ⊗ S) for all ξ ∈ E. It then follows the proof of 2.10 [3], we have VTξ1⊗γ2 ξ2 = ∆(ξ1, ξ2) for all ξ1 ∈ E1 and ξ2 ∈ E2 (cf. 6.1.3 from 6.1.5 b) that the pair (βE, δE), where δE : E → (cid:102)M(E ⊗ S) is defined for all ξ ∈ E by δE(ξ) := VTξ, satisfies the conditions 1, 2, and 3 of Definition 6.1.4. The coassociativity condition of δE is derived from those of δE1 and δE2 exactly as in the proof of 2.10 [3]. 7.10 Proposition. -- We use all the notations and hypotheses of 7.9. If A is a G-C*-algebra and γ1 : A → L(E1) is a G-equivariant *-representation, then γ : A → L(E1 ⊗γ2 E2) the *-representation defined by γ(a) := γ1(a) ⊗γ2 1 for all a ∈ A is G-equivariant. (cid:78) Proof. Through the identification (7.3), for all x ∈ A⊗ S the operator (γ1 ⊗ idS)(x)⊗(cid:101)γ2⊗idS 1 is identified to (γ ⊗ idS)(x). This identification also holds for x ∈ (cid:102)M(A ⊗ S) (by using the fact that any element of S can be written as a product of two elements of S). In particular, for all a ∈ A the operator (γ1 ⊗ idS)δA(a) ⊗(cid:101)γ2⊗idS 1 is identified to (γ ⊗ idS)δA(a). Hence, δE(γ(a)ξ) = (γ ⊗ idS)δA(a) ◦ δE(ξ) for all ξ ∈ E and a ∈ A by definition of δE. The relation βE(no) ◦ γ(a) = γ(βA(no)a) for n ∈ N and a ∈ A is straightforward. From now on, we assume the quantum groupoid G to be regular. We recall that any action of the quantum groupoid G on a Hilbert module is necessarily continuous (cf. 6.1.26). 7.11 Proposition-Definition. -- Let A, C and B be G-C*-algebras. Let E1 (resp. E2) be a G-equivariant imprimitivity A-C-bimodule (resp. C-B-bimodule). Denote by E1 ⊗C E2 the internal tensor product E1 ⊗γ2 E2, where γ2 : C → L(E2) is the G-equivariant *-representation defined by the left action of C on E2. The Hilbert B-module E1 ⊗C E2 endowed with the action of G defined in 7.9 is a G-equivariant imprimitivity A-B-bimodule for the left action of A and the A-valued inner product defined by the formulas: • a(ξ1 ⊗C ξ2) := aξ1 ⊗C ξ2, for all a ∈ A, ξ1 ∈ E1 and ξ2 ∈ E2; • A(cid:104)ξ1 ⊗C ξ2, η1 ⊗C η2(cid:105) := A(cid:104)ξ1, η1 C(cid:104)ξ2, η2(cid:105)(cid:105), for all ξ1, η1 ∈ E1 and ξ2, η2 ∈ E2. (cid:78) Proof. It is known that E1 ⊗C E2 is an imprimitivity A-B-bimodule. The rest of the proof is contained in 7.9 and 7.10. 7.12 Proposition. -- Let A and B be G-C*-algebras. Let E be a G-equivariant imprimitivity A-B-bimodule. Then, the map E∗ ⊗A E → B ; ξ∗ ⊗A η (cid:55)→ (cid:104)ξ, η(cid:105)B defines an isomorphism of G-equivariant imprimitivity B-B-bimodules. (cid:78) Proof. It is known that the map Φ : E∗ ⊗A E → B ; ξ∗ ⊗A η (cid:55)→ (cid:104)ξ, η(cid:105)B is an isomorphism of imprimitivity B-B-bimodules. The G-equivariance of Φ is a restatement of the formula δB((cid:104)ξ, η(cid:105)B) = δE(ξ)∗ ◦ δE(η) for ξ, η ∈ E. 7.13 Definition. -- Let A and B be G-C*-algebras. We say that A and B are G-equivariantly Morita equivalent if there exists a G-equivariant imprimitivity A-B-bimodule. The G- equivariant Morita equivalence is a reflexive (7.7 (i)), symmetric (7.7 (iv)) and transitive (7.11) relation on the class of G-C*-algebras. (cid:78) biduality and equivariant morita equivalence. In this paragraph, the mea- sured quantum groupoid G is assumed to be regular. Let us fix a G-C*-algebras A. We crossed product (A (cid:111) G) (cid:111) (cid:98)G. show that there is a canonical G-equivariant Morita equivalence between A and the double 7.14 Notations. -- Denote by K := K(H) for short. Consider the Hilbert A-modules E0 := A ⊗ H and EA,R := qβA(cid:98)α(A ⊗ H). Let V ∈ L(H ⊗ S) be the unique partial isometry such that (idK ⊗ L)(V ) = V. (cid:78) 7.15 Proposition. -- There exists a unique bounded linear map δE0 : E0 → L(A ⊗ S,E0 ⊗ S) such that δE0 (a ⊗ ζ) = V23δA(a)13(1A ⊗ ζ ⊗ 1S), for all a ∈ A and ζ ∈ H. (cid:78) 58 J. CRESPO Proof. If B is a C*-algebra and K a Hilbert space, we identify M(B) ⊗ K with a closed vector subspace of L(B, B ⊗ K). We have (δA ⊗ idH )(ξ) ∈ L(A ⊗ S, A ⊗ S ⊗ H) and (δA ⊗ idH )(ξ)∗ = (δA ⊗ idH∗ )(ξ∗) for ξ ∈ E0. Let σ ∈ L(S ⊗ H, H ⊗ S) be the flip map. Denote by δE0 : E0 → L(A ⊗ S,E0 ⊗ S) the map defined by δE0 (ξ) := V23σ23(δA ⊗ idH )(ξ) for ξ ∈ E0. It is clear that δE0 : E0 → L(A ⊗ S,E0 ⊗ S) is linear map satsifying the formula δE0 (a ⊗ ξ) = V23δA(a)13(1A ⊗ ξ ⊗ 1S) for all a ∈ A and ζ ∈ H. 7.16 Proposition. -- We have the following statements: 1. δE0 (ξ)∗δE0 (η) = δA((cid:104)qβA(cid:98)αξ, qβA(cid:98)αη(cid:105)), for all ξ, η ∈ E0; 2. δE0 (ξa) = δE0 (ξ)δA(a), for all ξ ∈ E0 and a ∈ A; 3. δE0 (qβA(cid:98)αξ) = δE0 (ξ), for all ξ ∈ E0; 4. δE0 (E0)(A ⊗ S) ⊂ EA,R ⊗ S. Proof. 1. Let ξ, η ∈ E0, we have δE0 (ξ)∗δE0 (η) = (δA ⊗ id)(ξ∗)σ∗ 23V∗ have σ∗V∗V σ = qβ(cid:98)α. Let n, n(cid:48) ∈ N. For all a ∈ A and ζ ∈ H, we have (1A ⊗ β(no) ⊗(cid:98)α(n(cid:48)))(δA ⊗ idH )(a ⊗ ζ) = (1A ⊗ β(no))δA(a) ⊗(cid:98)α(n(cid:48))ζ = δA(βA(no)a) ⊗(cid:98)α(n(cid:48))ζ = (δA ⊗ idH )((βA(no) ⊗(cid:98)α(n(cid:48)))(a ⊗ ζ)). Hence, (1A ⊗ β(no) ⊗(cid:98)α(n(cid:48)))(δA ⊗ idH )(η) = (δA ⊗ idH )((βA(no) ⊗(cid:98)α(n(cid:48)))η). It then fol- (cid:78) 23V23σ23(δA ⊗ id)(η). We 23V23σ23(δA ⊗ idH )(η) = (δA ⊗ idH )(qβA(cid:98)αη). We finally have lows that σ∗ 23V∗ δE0 (ξ)∗ δE0 (η) = (δA ⊗ idH∗ )(ξ = δA((cid:104)ξ, qβA(cid:98)αη(cid:105)) = δA((cid:104)qβA(cid:98)αξ, qβA(cid:98)αη(cid:105)), ∗)(δA ⊗ idH )(qβA(cid:98)αη) where the last equality follows from the fact that qβA(cid:98)α ∈ L(E0) is a self-adjoint projection. 2. Let us fix a, b ∈ A and ζ ∈ H. We have δA(a)13(1A ⊗ ζ ⊗ 1S) = (1A ⊗ ζ ⊗ 1S)δA(a) in L(A ⊗ S,E0 ⊗ S). Hence, δE0 ((b ⊗ ζ)a) = δE0 (b ⊗ ζ)δA(a). 3. Let a ∈ A and ζ ∈ H. For all n, n(cid:48) ∈ N, we have δE0 (βA(no)a ⊗(cid:98)α(n(cid:48))ζ) = V23(1A ⊗(cid:98)α(n(cid:48)) ⊗ 1S)δA(βA(no)a)13(1A ⊗ ζ ⊗ 1S) = V23(1A ⊗(cid:98)α(n(cid:48)) ⊗ β(no))δA(a)13(1A ⊗ ζ ⊗ 1S). Hence, δE0 (qβA(cid:98)α(a ⊗ ζ)) = V23q(cid:98)αβ,23δA(a)13(1A ⊗ ζ ⊗ 1S) = δE0 (a ⊗ ζ). 4. It suffices to show that qβA(cid:98)α,12δE0 (ξ) = δE0 (ξ) for all ξ ∈ E0. We recall (cf. 3.1.5) that ((cid:98)α(n) ⊗ 1S)V = V (1K ⊗ α(n)) for all n ∈ N. Hence, qβA(cid:98)α,12V23 = V23qβAα,13. follows from qβAα = δA(1A) that qβA(cid:98)α,12V23δA(a)13 = V23δA(a)13 for all a ∈ A. Hence, qβA(cid:98)α,12δE0 (a ⊗ ζ) = δE0 (a ⊗ ζ) for all a ∈ A and ζ ∈ H. 7.17 Notations. -- According to the previous proposition, δE0 restricts to a linear map It then δEA,R : EA,R → L(A ⊗ S,EA,R ⊗ S), which satisfies the following statements: then have a non-degenerate *-homomorphism • δEA,R (ξ)∗δEA,R (η) = δA((cid:104)ξ, η(cid:105)), for all ξ, η ∈ EA,R; • δEA,R (ξa) = δEA,R (ξ)δA(a), for all ξ ∈ EA,R and a ∈ A. Since [(cid:98)α(n(cid:48)), β(no)] = 0 for all n, n(cid:48) ∈ N, we have [1A ⊗ β(no), qβA(cid:98)α] = 0 for all n ∈ N. We βEA,R : No → L(EA,R) ; n (cid:55)→ (1A ⊗ β(no))(cid:22)EA,R . Since β and(cid:98)α commute pointwise and VV∗ = qβα, we have [V23V∗ qβEA,R α = V23V∗ 23, qβA(cid:98)α,12] = 0. Hence, (cid:22)EA,R⊗S ∈ L(EA,R ⊗ S). (cid:78) 23 ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS 59 (cid:78) 7.18 Proposition. -- We have the following statements: 1. δEA,R (EA,R) ⊂ (cid:102)M(EA,R ⊗ S); 2. [δEA,R (EA,R)(A ⊗ S)] = qβEA,R α(EA,R ⊗ S); 3. δEA,R (βEA,R (no)ξ) = (1EA,R ⊗ β(no))δEA,R (ξ), for all ξ ∈ EA,R and n ∈ N. Proof. 1. Let us prove that δEA,R (ξ)(1A ⊗ s) ∈ EA,R ⊗ S for all ξ ∈ EA,R and s ∈ S. It amounts to proving that δE0 (ξ)(1A ⊗ s) ∈ E0 ⊗ S for all ξ ∈ E0 and s ∈ S (cf. 7.16 3, 4). Let a ∈ A and ζ ∈ H. It follows from the relation δA(A)(1A ⊗ S) ⊂ A ⊗ S that δE0 (a ⊗ ζ)(1A ⊗ s) = (1A ⊗ V (ζ ⊗ 1S))δA(a)(1A ⊗ s) is the norm limit of finite sums of the form ∑i ai ⊗ V (ζ ⊗ si), where ai ∈ A and si ∈ S. Hence, δE0 (a ⊗ ζ)(1A ⊗ s) ∈ E0 ⊗ S. Now, let us prove that (1EA,R ⊗ y)δEA,R (ξ) ∈ EA,R ⊗ S for all ξ ∈ EA,R and y ∈ S. This also amounts to proving that (1E0 ⊗ y)δE0 (ξ) ∈ E0 ⊗ S for all ξ ∈ E0 and y ∈ S. Let a ∈ A, ζ ∈ H with x ∈ (cid:98)S and η ∈ H. We have (1K ⊗ y)V (ζ ⊗ 1S) = (ρ ⊗ idS)((1(cid:98)S ⊗ y)V(x ⊗ 1S))(η ⊗ 1S). and y ∈ S, we have (1E0 ⊗ y)δE0 (a ⊗ ζ) = (1A ⊗ (1H ⊗ y)V (ζ ⊗ 1S))δA(a). Write ζ = ρ(x)η Since G is regular, we have (1(cid:98)S ⊗ y)V(x ⊗ 1S) ∈ (cid:98)S ⊗ S (cf. 4.10 2). Hence, (1E0 ⊗ y)δE0 (a ⊗ ζ) is the norm limit of finite sums of elements of the form (1A ⊗ ρ(x(cid:48))η ⊗ y(cid:48))δA(a) with x(cid:48) ∈ (cid:98)S and y(cid:48) ∈ S. Hence, (1E0 ⊗ y)δE0 (a ⊗ ζ) ∈ E0 ⊗ S since (1A ⊗ S)δA(A) ⊂ A ⊗ S. 23δE0 (ξ) = δE0 (ξ) for all ξ ∈ E0. It then follows that 2. Since VV∗V = V, we have V23V∗ qβEA,R αδEA,R (ξ) = δEA,R (ξ), for all ξ ∈ EA,R. By the first statement, we then obtain δEA,R (EA,R)(A ⊗ S) ⊂ qβEA,R α(EA,R ⊗ S). Conversely, let a ∈ A, ζ ∈ H and s ∈ S. Since V23qβAα,13 = qβA(cid:98)α,12V23, we have qβEA,R α(qβA(cid:98)α(a ⊗ ζ) ⊗ s) = V23V∗ 23qβA(cid:98)α,12(a ⊗ ζ ⊗ s) = V23qβAα,13(a ⊗ V∗(ζ ⊗ s)). Hence, qβEA,R α(qβA(cid:98)α(a ⊗ ζ) ⊗ s) is the norm limit of finite sums of elements of the form: (cid:48) ∈ H, s(cid:48) ∈ S. (cid:48) ⊗ s(cid:48)) = V23(qβAα(a ⊗ s(cid:48)))23(1A ⊗ ζ (cid:48) ⊗ 1S), where ζ V23qβAα,13(a ⊗ ζ By continuity of the action (δA, βA), V23qβAα,13(a ⊗ ζ(cid:48) ⊗ s(cid:48)) is the norm limit of finite sums of the form ∑i V23δA(ai)13(1A ⊗ ζ(cid:48) ⊗ si) = ∑i δEA,R (qβA(cid:98)α(ai ⊗ ζ(cid:48)))(1A ⊗ si), where ai ∈ A and si ∈ S. As a result, we have qβEA,R α(qβA(cid:98)α(a ⊗ ζ) ⊗ s) ∈ [δEA,R (EA,R)(A ⊗ S)] for all a ∈ A, ζ ∈ H and s ∈ S. Hence, qβEA,R α(EA,R ⊗ S) ⊂ [δEA,R (EA,R)(A ⊗ S)]. 3. Let ξ = qβA(cid:98)α(a ⊗ ζ), with a ∈ A and ζ ∈ H. We have βEA,R (no)ξ = (1A ⊗ β(no))qβA(cid:98)α(a ⊗ ζ) = qβA(cid:98)α(a ⊗ β(no)ζ). Moreover, we have V (β(no) ⊗ 1S) = (1K ⊗ β(no))V for all n ∈ N (cf. 3.1.5). It then follows that δEA,R (βEA,R (no)ξ) = δE0 (a ⊗ β(no)ζ) = (1E0⊗ β(no))δE0 (a ⊗ ζ) = (1EA,R⊗ β(no))δEA,R (ξ). Consequently, δEA,R ⊗ idS and idEA,R ⊗ δ extend to linear maps from L(A ⊗ S,EA,R ⊗ S) to L(A ⊗ S ⊗ S,EA,R ⊗ S ⊗ S) (cf. 6.1.7) and we have: (δEA,R ⊗ idS)(T)(δA ⊗ idS)(x) = (δEA,R ⊗ idS)(Tx); (idEA,R ⊗ δ)(T)(idA ⊗ δ)(x) = (idEA,R ⊗ δ)(Tx); for all x ∈ A ⊗ S and T ∈ L(A ⊗ S,EA,R ⊗ S). 7.19 Proposition. -- For all ξ ∈ EA,R, (δEA,R ⊗ idS)δEA,R (ξ) = (idEA,R ⊗ δ)δEA,R (ξ). (cid:78) 60 J. CRESPO Proof. Let a ∈ A, ζ ∈ H and x ∈ A ⊗ S. Let ξ := qβA(cid:98)α(a ⊗ ζ). We have (δEA,R ⊗ idS)δEA,R (ξ)(δA ⊗ idS)(x) = (δE0 ⊗ idS)(V23(δA(a)x)13(1A ⊗ ζ ⊗ 1S)). For all b ∈ A, ζ(cid:48) ∈ H and s(cid:48) ∈ S, we have (δE0 ⊗ idS)(b ⊗ ζ (cid:48) ⊗ s(cid:48)) = V23δA(b)13(1A ⊗ ζ (cid:48) ⊗ 1S ⊗ s(cid:48)). Hence, (δE0 ⊗ idS)(b ⊗ X) = V23δA(b)13X24 ∈ L(A ⊗ S ⊗ S, A ⊗ H ⊗ S ⊗ S) for all b ∈ A and X ∈ H ⊗ S. In particular, we have (δE0 ⊗ idS)(V23(b ⊗ ζ ⊗ s)) = (δE0 ⊗ idS)(b ⊗ V (ζ ⊗ s)) = V23δA(b)13V24(1A ⊗ ζ ⊗ 1S ⊗ s) = V23V24δA(b)13(1A ⊗ ζ ⊗ 1S ⊗ s). However, we have (idK ⊗ δ)(V ) = V12V13. Hence, V23V24 = (idA⊗K ⊗ δ)(V23). Moreover, we have δA(b)13(1A ⊗ ζ ⊗ 1S ⊗ s) = (δA,13 ⊗ idS)(b ⊗ s)(1A ⊗ ζ ⊗ 1S ⊗ 1S), for all b ∈ A and s ∈ S, where δA,13 : A → M(A ⊗ K ⊗ S) is the strictly continuous *-homomorphism defined by δA,13(a) := δA(a)13 for all a ∈ A. As a result, for all Y ∈ A ⊗ S we have (δE0 ⊗ idS)(V23Y13(1A ⊗ ζ ⊗ 1S)) = (idA⊗K ⊗ δ)(V23)(δA,13 ⊗ idS)(Y)(1A ⊗ ζ ⊗ 1S ⊗ 1S). In particular, we have (δE0 ⊗ idS)(δE0 (a ⊗ ζ)x) = (idA⊗K ⊗ δ)(V23)(δA,13 ⊗ idS)(δA(a)x)(1A ⊗ ζ ⊗ 1S ⊗ 1S) = (idA⊗K ⊗ δ)(V23)(δA,13 ⊗ idS)δA(a)(1A ⊗ ζ ⊗ 1S ⊗ 1S)(δA ⊗ idS)(x). Moreover, we have (δA,13 ⊗ idS)δA = (idA⊗K ⊗ δ)δA,13. Hence, (δEA,R ⊗ idS)δEA,R (ξ)x = (idA⊗K ⊗ δ)(V23δA(a)13)(1A ⊗ ζ ⊗ 1S ⊗ 1S)x, for all x ∈ qβAα,12(A ⊗ S ⊗ S). In particular, if x ∈ qβAα,12qβα,23(A ⊗ S ⊗ S) we have (δEA,R ⊗ idS)δEA,R (ξ)x = (idA⊗K ⊗ δ)(V23δA(a)13)qβα,34(1A ⊗ ζ ⊗ 1S ⊗ 1S)x = (idA⊗K ⊗ δ)(V23δA(a)13)(idE0 ⊗ δ)(1A ⊗ ζ ⊗ 1S)x = (idE0 ⊗ δ)δE0 (a ⊗ ζ)x = (idEA,R ⊗ δ)δEA,R (ξ)x. unique faithful continuous *-homomorphism for the strict/*-strong topologies such that Hence, (δEA,R ⊗ idS)δEA,R (ξ) = (idEA,R ⊗ δ)δEA,R (ξ). Now, we can assemble the previous results (see also 6.1.26) in the statement below. 7.20 Proposition. -- The triple (EA,R, βEA,R, δEA,R ) is a G-equivariant Hilbert A-module. (cid:78) φ : (A (cid:111) G) (cid:111) (cid:98)G → D of G-C*-algebras (cf. 5.1.22). Let jD : M(D) → L(E0) be the Let D be the bidual G-C*-algebra of A. We have a canonical G-equivariant *-isomorphism jD(1D) = qβA(cid:98)α. 7.21 Proposition. -- The *-representation (A (cid:111) G) (cid:111) (cid:98)G → L(EA,R) ; x (cid:55)→ φ(x)(cid:22)EA,R is G- (cid:78) equivariant. Proof. We have to prove that δE0 (dξ) = (jD ⊗ idS)(δD(d)) ◦ δE0 (ξ) for all d ∈ D and ξ ∈ E0 • Let b ∈ A, x ∈ (cid:98)S and ζ ∈ H. We have (cf. 7.3 3 and 7.16 3). Let us prove it in three steps: However, [V, λ(x) ⊗ 1S] = 0 (as λ((cid:98)S) ⊂ (cid:98)M and V ∈ (cid:98)M(cid:48) ⊗ M). Hence, δE0 (b ⊗ λ(x)ζ) = (1A ⊗ λ(x) ⊗ 1S)δE0 (b ⊗ ζ) δE0 (b ⊗ λ(x)ζ) = (1A ⊗ V (λ(x) ⊗ 1S))δA(b)13(1A ⊗ ζ ⊗ 1S). ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS and then δE0 ((1A ⊗ λ(x))ξ) = (1A ⊗ λ(x) ⊗ 1S)δE0 (ξ), for all x ∈ (cid:98)S and ξ ∈ E0. • Let y ∈ S. Since L(y) ∈ M ⊂(cid:98)α(N)(cid:48), we have V (L(y) ⊗ 1S) =Vq(cid:98)αβ(L(y) ⊗ 1S) =V (L(y) ⊗ 1S)q(cid:98)αβ =V (L(y) ⊗ 1S)V∗V = (L ⊗ idS)δ(y)V. For all b ∈ A and ζ ∈ H, we have 61 δE0 ((1A ⊗ L(y))(b ⊗ ζ)) = (1A ⊗ V (L(y) ⊗ 1S))δA(b)13(1A ⊗ ζ ⊗ 1S) = (1A ⊗ (L ⊗ idS)δ(y))δE0 (b ⊗ ζ). Hence, δE0 ((1A ⊗ L(y))ξ) = (1A ⊗ (L ⊗ idS)δ(y))δE0 (ξ) for all y ∈ S and ξ ∈ E0. In virtue of the first two steps, for all ξ ∈ EA,R we have δEA,R ((1A ⊗ λ(x)L(y))ξ) = (1A ⊗ λ(x) ⊗ 1S)(1A ⊗ (L ⊗ idS)δ(y))δEA,R (ξ). • Let s ∈ S. We have (cf. 3.1.3) (R(s) ⊗ 1)V = (U ⊗ 1)Σ(1 ⊗ L(s))Σ(U∗ ⊗ 1)V = (U ⊗ 1)Σ(1 ⊗ L(s))WΣ(U∗ ⊗ 1). WW∗ = qα(cid:98)β and L(s) ∈ M ⊂ (cid:98)β(No)(cid:48). Therefore, since (U ⊗ 1)ΣW = V(U ⊗ 1)Σ we have Besides, (1 ⊗ L(s))W = (1 ⊗ L(s))WW∗W = WW∗(1 ⊗ L(s))W = Wδ(s) since we have (R(s) ⊗ 1)V = VΣ(1 ⊗ U)δ(s)(1 ⊗ U∗)Σ. Hence, (R(s) ⊗ 1S)V = V σ(idS ⊗ R)(δ(s))σ∗ for all s ∈ S. We then have ((idA ⊗ R)(x) ⊗ 1S)V23 = V23σ23(idA⊗S ⊗ R)((idA ⊗ δ)(x))σ∗ 23 for all x ∈ A ⊗ S. But, since R and δ are strictly continuous this equality also holds for all x ∈ M(A ⊗ S). In particular, we have πR(a)12V23 = V23σ23(idA⊗S ⊗ R)(δ2 23 for all a ∈ A. By coassociativity of δA, we have A(a))σ∗ ∗ πR(a)12V23 = V23σ23(δA ⊗ idK)(πR(a))σ 23, for all a ∈ A. It then follows that πR(a)12δE0 (ξ) = V23σ23(δA ⊗ idK)(πR(a))(δA ⊗ idH )(ξ) = V23σ23(δA ⊗ idH )(πR(a)ξ) = δE0 (πR(a)ξ), for all a ∈ A and ξ ∈ E0. In particular, πR(a)12δE0 (ξ0) = δE0 (πR(a)ξ0) for all a ∈ A and ξ ∈ E0. We have proved that for all a ∈ A, x ∈ (cid:98)S, y ∈ S and ξ ∈ E0, we have δE0 (πR(a)(1A ⊗ λ(x)L(y))ξ) = πR(a)12(1A ⊗ λ(x) ⊗ 1S)(1A ⊗ (L ⊗ idS)δ(y))δE0 (ξ). However, for all a ∈ A, x ∈ (cid:98)S and y ∈ S we have (cf. 3.37 d) [2]) (jD ⊗ idS)δD(πR(a)(1A ⊗ λ(x)L(y))) = πR(a)12(1A ⊗ λ(x) ⊗ 1S)(1A ⊗ (L ⊗ idS)δ(y)). If d = πR(a)(1A ⊗ λ(x)L(y)) ∈ D, where a ∈ A, x ∈ (cid:98)S and y ∈ S, we have proved that D = [πR(a)(1A ⊗ λ(x)L(y)) ; a ∈ A, x ∈ (cid:98)S, y ∈ S]. δE0 (dξ) = (jD ⊗ idS)(δD(d)) ◦ δE0 (ξ) for all ξ ∈ E0. Thus, the statement is proved since 7.22 Theorem. -- The G-C*-algebras (A (cid:111) G) (cid:111) (cid:98)G and A are Morita equivalent via the G- equivariant imprimitivity (A (cid:111) G) (cid:111) (cid:98)G-A-bimodule EA,R. (cid:78) Proof. Let us prove that the Hilbert A-module EA,R is full. Fix x ∈ A and write x = a∗b for a, b ∈ A. There exists ω ∈ B(H)∗ such that (idA ⊗ ω)(qβA(cid:98)α) = 1A. Hence, x is the norm limit of finite sums of elements of the form a∗(idA ⊗ ωξ,η)(qβA(cid:98)α)b, where ξ, η ∈ H. However, for all ξ, η ∈ H we have a∗(idA ⊗ ωξ,η)(qβA(cid:98)α)b = (a ⊗ ξ)∗qβA(cid:98)α(b ⊗ η) = (cid:104)qβA(cid:98)α(a ⊗ ξ), qβA(cid:98)α(b ⊗ η)(cid:105). Hence, A = [(cid:104)ξ, η(cid:105) ; ξ, η ∈ EA,R]. Now, we recall that D = qβA(cid:98)α(A ⊗ K)qβA(cid:98)α (cf. 5.1.24). It is easily seen that the left action of (A (cid:111) G) (cid:111) (cid:98)G (cf. 7.21) induces a G-equivariant *-isomorphism (A (cid:111) G) (cid:111) (cid:98)G (cid:39) K(EA,R). 62 8 appendix J. CRESPO 8.1 Normal linear forms, weights and operator-valued weights on von Neumann algebras [8] Let M be a von Neumann algebra. Denote by M∗ (resp. M+∗ ) the Banach space (resp. positive cone) of the normal linear forms (resp. positive normal linear forms) on M. Let ω ∈ M∗ and a, b ∈ M. Denote by aω ∈ M∗ and ωb ∈ M∗ the normal linear functionals on M given for all x ∈ M by: We have a(cid:48)(aω) = (a(cid:48)a)ω and (ωb)b(cid:48) = ω(bb(cid:48)), for all a, a, b, b(cid:48) ∈ M. We also denote (aω)(x) := ω(xa); (ωb)(x) := ω(bx). aωb := a(ωb) = (aω)b; ωa := a∗ ωa. If ω ∈ M+∗ , then ωa ∈ M+∗ . Note that (ωa)b = ωab for all a, b ∈ M. If ω ∈ M∗ we define ω ∈ M∗ by setting ω(x) := ω(x∗), for all x ∈ M. Let H be a Hilbert space and let us fix ξ, η ∈ H. Denote by ωξ,η ∈ B(H)∗ the normal linear form defined by ωξ,η(x) := (cid:104)ξ, xη(cid:105), for all x ∈ B(H). Note that we have ωξ,η = ωη,ξ, aωξ,η = ωξ,aη and ωξ,ηa = ωa∗ξ,η for all a ∈ B(H). 8.1.1. Tensor product of normal linear forms. Let M and N be von Neumann algebras, φ ∈ M∗ and ψ ∈ N∗. There exists a unique φ ⊗ ψ ∈ (M ⊗ N)∗ such that (φ ⊗ ψ)(x ⊗ y) = φ(x)ψ(y) have an (completely) isometric identification M∗(cid:98)⊗π N∗ = (M ⊗ N)∗, where (cid:98)⊗π denotes the for all x ∈ M and y ∈ N. Moreover, (cid:107)φ ⊗ ψ(cid:107) (cid:54) (cid:107)φ(cid:107) · (cid:107)ψ(cid:107). Actually, it is known that we projective tensor product of Banach spaces. In particular, any ω ∈ (M ⊗ N)∗ is the norm limit of finite sums of the form ∑i φi ⊗ ψi, where φi ∈ M∗ and ψi ∈ N∗. (cid:78) 8.1.2. Slicing with normal linear forms. We will also need to slice maps with normal linear forms. Let M1 and M2 be von Neumann algebras, ω1 ∈ (M1)∗ and ω2 ∈ (M2)∗. Therefore, the maps ω1 (cid:12) id : M1 (cid:12) M2 → M1 and id(cid:12) ω2 : M1 (cid:12) M2 → M2 extend uniquely to norm continuous normal linear maps ω1 ⊗ id : M1 ⊗ M2 → M2 and id ⊗ ω2 : M1 ⊗ M2 → M1. Let H and K be Hilbert spaces, for ξ ∈ H and η ∈ K we define θξ ∈ B(K, H ⊗ K) and η ∈ B(H, H ⊗ K) by setting: θ(cid:48) θξ (ζ) := ξ ⊗ ζ, for all ζ ∈ K; η(ζ) := ζ ⊗ η, θ(cid:48) for all ζ ∈ H. If T ∈ B(H ⊗ K), φ ∈ B(K)∗ and ω ∈ B(H)∗, then the operators (id ⊗ φ)(T) ∈ B(H) and (ω ⊗ id)(T) ∈ B(K) are determined by the formulas: (cid:104)ξ1, (id ⊗ φ)(T)ξ2(cid:105) = φ(θ (cid:104)η1, (ω ⊗ id)(T)η2(cid:105) = ω(θ ∗ ξ1 Tθξ2 ), (cid:48)∗ η1 Tθ (cid:48) η2 ), ξ1, ξ2 ∈ H; η1, η2 ∈ K. In particular, we have: (id ⊗ ωη1,η2 )(T) = θ (cid:48)∗ η1 Tθ (cid:48) η2, η1, η2 ∈ K; (ωξ1,ξ2 ⊗ id)(T) = θ ∗ ξ1 Tθξ2, ξ1, ξ2 ∈ H. Let us recall some formulas that will be used several times. For all φ ∈ B(K)∗, ω ∈ B(H)∗ and T ∈ B(H ⊗ K), we have: x(id ⊗ φ)(T)y = (id ⊗ φ)((x ⊗ 1)T(y ⊗ 1)), (yωx ⊗ id)(T) = (ω ⊗ id)((x ⊗ 1)T(y ⊗ 1)) for all x, y ∈ B(H); a(ω ⊗ id)(T)b = (ω ⊗ id)((1 ⊗ a)T(1 ⊗ b)), (id ⊗ bφa)(T) = (id ⊗ φ)((1 ⊗ a)T(1 ⊗ b)) for all a, b ∈ B(K). We also have (id ⊗ φ)(T)∗ = (id ⊗ φ)(T∗), (φ ⊗ id)(ΣH⊗KTΣK⊗H ) = (id ⊗ φ)(T), (ω ⊗ id)(T)∗ = (ω ⊗ id)(T∗), (id ⊗ ω)(ΣH⊗KTΣK⊗H ) = (ω ⊗ id)(T), for all T ∈ B(H ⊗ K), φ ∈ B(K)∗ and ω ∈ B(H)∗. (cid:78) ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS 63 8.1.3 Definition. -- A weight ϕ on M is a map ϕ : M+ → [0, ∞] such that: • for all x, y ∈ M+, ϕ(x + y) = ϕ(x) + ϕ(y); • for all x ∈ M+ and λ ∈ R+, ϕ(λx) = λϕ(x). We denote by Nϕ := {x ∈ M ; ϕ(x∗x) < ∞} the left ideal of square ϕ-integrable elements ϕ := {x ∈ M+ ; ϕ(x) < ∞} the cone of positive ϕ-integrable elements of M and of M, M+ Mϕ := (cid:104)M+ ϕ (cid:105) the space of ϕ-integrable elements of M. (cid:78) 8.1.4 Definition. -- Let ϕ be a weight on M. The opposite weight of ϕ is the weight ϕo on Mo given by ϕo(xo) := ϕ(x) for all x ∈ M+. Then, we have Nϕo = (N∗ ϕ )o (cid:78) and Mϕo = (Mϕ)o. 8.1.5 Definition. -- A weight ϕ on M is called: • semi-finite, if Nϕ is σ-weakly dense in M; • faithful, if for x ∈ M+ the condition ϕ(x) = 0 implies x = 0; • normal, if ϕ(supi∈I xi) = supi∈I ϕ(xi) for all increasing bounded net (xi)i∈I of (cid:78) ϕo = (M+ ϕ)o, M+ M+. From now on, we will mainly use normal semi-finite faithful (n.s.f.) weights. Fix a n.s.f. weight ϕ on M. 8.1.6 Definition. -- We define an inner product on Nϕ by setting (cid:104)x, y(cid:105)ϕ := ϕ(x∗y), for all x, y ∈ Nϕ. We denote by (Hϕ, Λϕ) the Hilbert space completion of Nϕ with respect to this inner product, where Λϕ : Nϕ → Hϕ is the canonical map. There exists a unique unital normal *-representation πϕ : M → B(Hϕ) such that πϕ(x)Λϕ(y) = Λϕ(xy), for all x ∈ M and y ∈ Nϕ. (cid:78) The triple (Hϕ, πϕ, Λϕ) is called the G.N.S. construction for (M, ϕ). 8.1.7 Remarks. -- The linear map Λϕ is called the G.N.S. map. We have that Λϕ(Nϕ) is dense in Hϕ and (cid:104)Λϕ(x), Λϕ(y)(cid:105)ϕ = ϕ(x∗y) for all x, y ∈ Nϕ. In particular, Λϕ is injective. (cid:78) Moreover, we also call πϕ the G.N.S. representation. We recall below the main objects of the Tomita-Takesaki modular theory. 8.1.8 Proposition-Definition. -- Let M be a von Neumann algebra and ϕ a n.s.f. weight ϕ ∩ Nϕ) → Λϕ(N∗ ϕ ∩ Nϕ) ; Λϕ(x) (cid:55)→ Λϕ(x∗) is closable on M. The anti-linear map Λϕ(N∗ and its closure is a possibly unbounded anti-linear map Tϕ : D(Tϕ) ⊂ Hϕ → Hϕ such that D(Tϕ) = imTϕ and Tϕ ◦ Tϕ(x) = x for all x ∈ D(Tϕ). be the polar decomposition of Tϕ. The anti-unitary Jϕ : Hϕ → Hϕ is called the Let Tϕ = Jϕ∇1/2 modular conjugation for ϕ and the injective positive self-adjoint operator ∇ϕ is called the modular (cid:78) operator for ϕ. 8.1.9 Proposition-Definition. -- There exists a unique one-parameter group (σϕ morphisms on M, called the modular automorphism group of ϕ, such that t )t∈R of auto- ϕ πϕ(σϕ t (x)) = ∇it ϕπϕ(x)∇−it ϕ , for all t ∈ R and x ∈ M. t (x)) = ∇it ϕ t (x) ∈ Nϕ and Λϕ(σϕ Then, for all t ∈ R and x ∈ M we have σϕ (cid:78) 8.1.10 Proposition-Definition. -- The map CM : M → M(cid:48) ; x (cid:55)→ Jϕπϕ(x)∗ Jϕ is a normal (cid:78) unital *-antihomomorphism. 8.1.11 Definition. -- Let N be a von Neumann algebra. The extended positive cone of + consisting of the maps m : N+∗ → [0, ∞], which satisfy the following N is the set Next conditions: Λϕ(x). 64 J. CRESPO • for all ω1, ω2 ∈ N+∗ , m(ω1 + ω2) = m(ω1) + m(ω2); • for all ω ∈ N+∗ and λ ∈ R+, m(λω) = λm(ω); • m is lower semicontinuous with respect to the norm topology on N∗. (cid:78) 8.1.12 Notations. -- Let N be a von Neumann algebra. 1. From now on, we will identify N+ with its part inside Next and ω ∈ N+∗ we will denote by ω(m) the evaluation of m at ω. + . Accordingly, if m ∈ Next + + , we define a∗ma ∈ Next + and λ ∈ R+, we also define m + n ∈ Next 2. Let a ∈ N and m ∈ Next all ω ∈ N+∗ . If m, n ∈ Next by setting ω(m + n) := ω(m) + ω(n) and ω(λm) := λω(m) for all ω ∈ N+∗ . + by setting ω(a∗ma) := aωa∗(m) for + and λm ∈ Next + (cid:78) 8.1.13 Definition. -- Let N ⊂ M be a unital normal inclusion of von Neumann algebras. An operator-valued weight from M to N is a map T : M+ → Next + such that: • for all x, y ∈ M+, T(x + y) = T(x) + T(y); • for all x ∈ M+, ∀λ ∈ R+, T(λx) = λT(x); • for all x ∈ M+ and a ∈ N, T(a∗xa) = a∗T(x)a. Let NT := {x ∈ M ; T(x∗x) ∈ N+}, M+ T (cid:105). (cid:78) 8.1.14 Definition. -- Let N ⊂ M be a unital normal inclusion of von Neumann algebras. An operator-valued weight T from M to N is said to be: T := {x ∈ M+ ; T(x) ∈ N+} and MT := (cid:104)M+ • semi-finite, if NT is σ-weakly dense in M; • faithful, if for x ∈ M+ the condition T(x) = 0 implies x = 0; • normal, if for every increasing bounded net (xi)i∈I of elements of M+ and ω ∈ N+∗ , (cid:78) we have ω(T(supi∈I xi)) = limi∈I ω(T(xi)). Note that if T : M+ → Next + is an operator-valued weight, it extends uniquely to a semi- + → Next linear map T : Mext + . This will allow us to compose n.s.f. operator-valued weights. Indeed, let P ⊂ N ⊂ M be unital normal inclusions of von Neumann algebras. Let S (resp. T) be an operator-valued weight from N (resp. M) to P (resp. N). We define an operator-valued weight from M to P by setting (S ◦ T)(x) := S(T(x)) for all x ∈ N+. 8.2 Relative tensor product of Hilbert spaces and fiber product of von Neumann algebras In this paragraph, we will recall the definitions, notations and important results concerning the relative tensor product and the fiber product which are the main technical tools of the theory of measured quantum groupoids. For more information, we refer the reader to [7]. In the whole section, N is a von Neumann algebra endowed with a n.s.f. weight ϕ. Let π (resp. γ) be a normal unital *-representation of N (resp. No) on a Hilbert space H (resp. K). The Hilbert space H (resp. K) may be considered as a relative tensor product. left (resp. right) N-module. Moreover, Hϕ is an N-bimodule whose actions are given by xξ := πϕ(x)ξ and ξy := Jϕπϕ(y∗)Jϕξ, for all ξ ∈ Hϕ and x, y ∈ N. 8.2.1 Definition. -- We define the set of right (resp. left) bounded vectors with respect to ϕ and π (resp. γ) to be: ϕ(π,H) := {ξ ∈ H ; ∃ C ∈ R+, ∀ x ∈ Nϕ, (cid:107)π(x)ξ(cid:107) (cid:54) C(cid:107)Λϕ(x)(cid:107)}, (resp. (K, γ)ϕ := {ξ ∈ K ; ∃ C ∈ R+, ∀ x ∈ N∗ ϕ, (cid:107)γ(xo)ξ(cid:107) (cid:54) C(cid:107)Λϕo (xo)(cid:107)}). ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS 65 If ξ ∈ ϕ(π,H), we denote by Rπ,ϕ unique bounded operator such that ξ ∈ B(Hϕ,H) (or simply Rπ ξ if ϕ is understood) the Rπ,ϕ ξ Λϕ(x) = π(x)ξ, for all x ∈ Nϕ. Similarly, if ξ ∈ (K, γ)ϕ we denote Lγ,ϕ unique bounded operator such that ξ ∈ B(Hϕ,K) (or simply Lγ ξ if ϕ is understood) the Lγ,ϕ ξ JϕΛϕ(x∗) = γ(xo)ξ, for all x ∈ N∗ ϕ, where we have used the identification Hϕo → Hϕ ; Λϕo (xo) (cid:55)→ JϕΛϕ(x∗). (cid:78) Note that ξ ∈ K is left bounded with respect to ϕ and γ if, and only if, it is right bounded with respect to the n.s.f. weight ϕc := ϕ ◦ C−1 N on N(cid:48) and the normal unital *-representation γc := γ ◦ C−1 N of N(cid:48). It is important to note that (K, γ)ϕ (resp. ϕ(π,H)) is dense in K (resp. H) (cf. Lemma 2 of [7]). If ξ ∈ ϕ(π,H) (resp. ξ ∈ (K, γ)ϕ), we have that Rπ,ϕ Therefore, for all ξ, η ∈ ϕ(π,H) (resp. (K, γ)ϕ) we have ) is left (resp. right) N-linear. (resp. Lγ,ϕ ξ ξ (Rπ,ϕ ξ )∗Rπ,ϕ η ∈ πϕ(N)(cid:48) = CN(N) and Rπ,ϕ ξ (Lγ,ϕ η η ∈ πϕ(N) and Lγ,ϕ )∗Lγ,ϕ )∗ ∈ π(N)(cid:48) (Rπ,ϕ )∗ ∈ γ(No)(cid:48)). η ξ (resp. (Lγ,ϕ ξ 8.2.2 Notations. -- (cf. 2.1 [16]) Let Kπ,ϕ := [Rπ,ϕ ξ (Rπ,ϕ η )∗ ; ξ, η ∈ ϕ(π,H)] (resp. Kγ,ϕ := [Lγ,ϕ ξ (Lγ,ϕ η )∗ ; ξ, η ∈ (H, γ)ϕ]). ξ η )∗Rπ,ϕ )o ∈ No N ((Rπ,ϕ (cid:104)ξ, η(cid:105)No := C−1 Note that Kπ,ϕ (resp. Kγ,ϕ) is a weakly dense ideal of π(N)(cid:48) (resp. γ(No)(cid:48)) (cf. Proposition 3 of [7]). If ϕ is understood, we denote Kπ (resp. Kγ) instead of Kπ,ϕ (resp. Kγ,ϕ). (cid:78) 8.2.3 Notations. -- Let ξ, η ∈ ϕ(π,H) (resp. (K, γ)ϕ), we denote −1 ϕ ((Lγ,ϕ ) ∈ N). (cid:78) 8.2.4 Proposition. -- For all ξ, η ∈ ϕ(π,H) (resp. ξ, η ∈ (K, γ)ϕ) and y ∈ N analytic for (σϕ t )t∈R, we have: 1. (cid:104)ξ, η(cid:105)∗ 2. (cid:104)ξ, ηyo(cid:105)No = (cid:104)ξ, η(cid:105)No σϕ No = (cid:104)η, ξ(cid:105)No (resp. (cid:104)ξ, η(cid:105)∗ i/2(y)o (resp. (cid:104)ξ, ηy(cid:105)N = (cid:104)ξ, η(cid:105)Nσϕ−i/2(y)). 8.2.5 Lemma. -- For all ξ1, ξ2 ∈ ϕ(π,H) and η1, η2 ∈ (K, γ)ϕ, we have (cid:104)η1, γ((cid:104)ξ1, ξ2(cid:105)No )η2(cid:105)K = (cid:104)ξ1, π((cid:104)η1, η2(cid:105)N)ξ2(cid:105)H. (resp. (cid:104)ξ, η(cid:105)N := π N = (cid:104)η, ξ(cid:105)N); )∗Lγ,ϕ η (cid:78) (cid:78) ξ 8.2.6 Definition. -- The relative tensor product K γ⊗π H (or simply denoted by K γ⊗π H) ϕ is the Hausdorff completion of the pre-Hilbert space (K, γ)ϕ (cid:12) ϕ(π,H), whose inner product is given by (cid:104)η1 ⊗ ξ1, η2 ⊗ ξ2(cid:105) := (cid:104)η1, γ((cid:104)ξ1, ξ2(cid:105)No )η2(cid:105)K = (cid:104)ξ1, π((cid:104)η1, η2(cid:105)N)ξ2(cid:105)H, for all η1, η2 ∈ (K, γ)ϕ and ξ1, ξ2 ∈ ϕ(π,H). If η ∈ (K, γ)ϕ and ξ ∈ ϕ(π,H), we will denote by η γ⊗π ξ (or simply η γ⊗π ξ) the image of η ⊗ ξ by the canonical map (K, γ)ϕ (cid:12) ϕ(π,H) → K γ⊗π H (isometric dense (cid:78) range). ϕ 66 J. CRESPO 8.2.7 Remarks. -- 1. By applying this construction to (No, ϕo) instead of (N, ϕ) we obtain the relative tensor product H π⊗γ ϕo K. 2. The relative tensor product K γ⊗π H is also the Hausdorff completion of the pre- Hilbert space (K, γ)ϕ (cid:12) H (resp. K (cid:12) ϕ(π,H)), whose inner product is given by: (cid:104)η1 ⊗ ξ1, η2 ⊗ ξ2(cid:105) := (cid:104)ξ1, π((cid:104)η1, η2(cid:105)N)ξ2(cid:105)H (resp. (cid:104)η1 ⊗ ξ1, η2 ⊗ ξ2(cid:105) := (cid:104)η1, γ((cid:104)ξ1, ξ2(cid:105)No )η2(cid:105)K). 3. Moreover, for all η ∈ K, ξ ∈ ϕ(π,H) and y ∈ N analytic for (σϕ t )t∈R we have γ(yo)η γ⊗π ξ = η γ⊗π π(σϕ−i/2(y))ξ. (cid:78) 8.2.8. The relative flip map is the isomorphism σγπ ϕ from K γ⊗π ϕ H onto H π⊗γ ϕo K given by: ϕ (η γ⊗π σγπ ϕ ξ) := ξ π⊗γ ϕo η, for all ξ ∈ (K, γ)ϕ and η ∈ ϕ(π,H) (or simply σγπ). Note that σγπ ϕ homomorphism is unitary and (σγπ ϕ )∗ = σπγ ϕo . Then, we can define a relative flip *- ϕ : B(K γ⊗π ςγπ ϕ H) → B(H π⊗γ ϕo K) (or simply denoted by ςγπ) by setting ςγπ ϕ (X) := σγπ ϕ X(σγπ ϕ )∗ for all X ∈ B(K γ⊗π H). (cid:78) ϕ fiber product of von neumann algebras. We continue to use the notations of the previous paragraph. 8.2.9 Proposition-Definition. -- Let Ki and Hi be Hilbert spaces, and γi : No → B(Ki) and πi : N → B(Hi) be unital normal *-homomorphisms for i = 1, 2. Let T ∈ B(K1,K2) and S ∈ B(H1,H2) such that T ◦ γ1(no) = γ2(no) ◦ T and S ◦ π1(n) = π2(n) ◦ S for all n ∈ N. Then, the linear map (K1, γ1)ϕ (cid:12) ϕ(π1,H1) → K2 γ2⊗π2 H2 ; ξ (cid:12) η (cid:55)→ Tξ γ2⊗π2 Sη extends uniquely to a bounded operator γ2 T γ1⊗π2 Sπ1 ∈ B(K1 γ1⊗π1 H1,K2 γ2⊗π2 H2) (or simply denoted by T γ1⊗π2 S), whose adjoint operator is γ1 T∗ γ2⊗π1 S∗). In particular, if x ∈ γ(No)(cid:48) and y ∈ π(N)(cid:48), then the linear map π2 (or simply T∗ γ2⊗π1 S∗ (K, γ)ϕ (cid:12) ϕ(π,H) → K γ⊗π H ; ξ (cid:12) η (cid:55)→ xξ γ⊗π yη extends uniquely to a bounded operator on K γ⊗π H denoted by x γ⊗π y ∈ B(K γ⊗π H). (cid:78) 8.2.10 Remark. -- With the notations of 8.2.9, let T : K1 → H2 and S : H1 → K2 be bounded antilinear maps such that T ◦ γ1(no)∗ = π2(n) ◦ T and S ◦ π1(n) = γ2(no)∗ ◦ S for all n ∈ N. In a similar way, we define π2 T γ1⊗γ2 Sπ1 ∈ B(K1 γ1⊗π1 H1,H2 π2⊗γ2 K2) (or simply T γ1⊗γ2 S). Note that these notations are different from those used in [17, 20]. (cid:78) Let M ⊂ B(K) and P ⊂ B(H) be two von Neumann algebras. Let us assume that π(N) ⊂ P and γ(No) ⊂ M. 8.2.11 Definition. -- The fiber product M γ(cid:63)π P of M and P over N is the commutant of {x γ⊗π y ; x ∈ M(cid:48), y ∈ P(cid:48)} ⊂ B(K γ⊗π H). Then, M γ(cid:63)π P is a von Neumann algebra. (cid:78) Note that we have ςγπ(M γ(cid:63)π P) = P π(cid:63)γ M. We still denote by ςγπ : M γ(cid:63)π P → P π(cid:63)γ M the restriction of ςγπ to M γ(cid:63)π P. ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS 67 8.2.12. Slicing with normal linear forms. Now, let us recall how to slice with normal linear forms. For ξ ∈ (K, γ)ϕ and η ∈ ϕ(π,H), we consider the following bounded linear maps: : H → K γ⊗π H, ζ (cid:55)→ ξ γ⊗π ζ; ργπ η : K → K γ⊗π H, ζ (cid:55)→ ζ γ⊗π η. λγπ ξ Let T ∈ B(K γ⊗π H) and ω ∈ B(H)∗ (resp. ω ∈ B(K)∗). By using the fact that (K, γ)ϕ (resp. ϕ(π,H)) is dense in H (resp. K), there exists a unique (id γ(cid:63)π ω)(T) ∈ B(K) (resp. (ω γ(cid:63)π id)(T) ∈ B(H)) such that (cid:104)ξ1, (id γ(cid:63)π ω)(T)ξ2(cid:105) = ω((λγπ (resp. (cid:104)η1, (ω γ(cid:63)π id)(T)η2(cid:105) = ω((ργπ )∗Tλγπ ξ2 η1 )∗Tργπ η2 ), ξ1 ), for all ξ1, ξ2 ∈ (K, γ)ϕ for all η1, η2 ∈ ϕ(π,H)). In particular, we have: (id γ(cid:63)π ωη1,η2 )(T) = (ργπ (ωξ1,ξ2 γ(cid:63)π id)(T) = (λγπ ξ1 η1 )∗Tργπ )∗Tλγπ η2 ∈ B(K), ∈ B(H), ξ2 for all η1, η2 ∈ ϕ(π,H); for all ξ1, ξ2 ∈ (K, γ)ϕ. If x ∈ M γ(cid:63)π P, then for all ω ∈ B(H)∗ (resp. ω ∈ B(K)∗) we have (id γ(cid:63)π ω)(x) ∈ M (resp. (ω γ(cid:63)π id)(x) ∈ P). We refrain from writing the details but we can easily define the slice maps if T takes its values in a different relative tensor product. Note that we can extend the notion of slice maps for normal linear forms to normal semi-finite weights. (cid:78) fiber product over a finite-dimensional von neumann algebra. Now, let us assume that N := Mnl (C) and ϕ := (cid:77) 1(cid:54)l(cid:54)k (cid:77) 1(cid:54)l(cid:54)k Trl(Fl−), where Fl is a positive invertible matrix of Mnl (C) and Trl is the non-normalized trace on Mnl (C). Denote by (Fl,i)1(cid:54)i(cid:54)nl the eigenvalues of Fl. 8.2.13 Proposition-Definition. -- (§7 [11]) The bounded linear map ϕ : K ⊗ H → K γ⊗π vγπ ϕ H ; ξ ⊗ η (cid:55)→ ξ γ⊗π η ϕ (or simply denoted by vγπ) is a coisometry if, and only if, we have ∑1(cid:54)i(cid:54)nl F−1 In the following, we assume the above condition to be satisfied. 8.2.14 Proposition-Definition. -- (§7 [11]) Let us denote l,i = 1 for all 1 (cid:54) l (cid:54) k. (cid:78) qγπ ϕ := (vγπ ϕ )∗vγπ ϕ (or simply qγπ). Then, qγπ ϕ is a self-adjoint projection of B(K ⊗ H) such that ϕ = ∑ qγπ 1(cid:54)l(cid:54)k ∑ 1(cid:54)i,j(cid:54)nl F−1/2 l,i F−1/2 l,j γ(e(l) o ij ) ⊗ π(e(l) ji ), ; x (cid:55)→ (vγπ ϕ )∗xvγπ ϕ (M ⊗ P)qγπ . Moreover, M γ(cid:63)π P → qγπ where, for all 1 (cid:54) l (cid:54) k, (e(l) ij )1(cid:54)i,j(cid:54)nl is a system of matrix units (s.m.u.) diagonalizing Fl, i.e. Fl = ∑1(cid:54)i(cid:54)nl Fl,ie(l) is a unital ii (cid:78) normal *-isomorphism. Since N is finite-dimensional, the inner product given by (cid:104)x, y(cid:105) := ϕ(x∗y) for all x, y ∈ N defines a structure of finite-dimensional Hilbert space on N. We have a (bounded) linear map µϕ : N ⊗ N → N defined for all x, y ∈ N by µϕ(x ⊗ y) = xy, where N ⊗ N is endowed with its canonical structure of finite-dimensional Hilbert space. 8.2.15 Proposition-Definition. -- For i = 1, 2, let πi *-representation of N on a Hilbert space Hi. Let us denote : N → B(Hi) be a unital normal ϕ ϕ qπ1π2 ϕ := (π1 ⊗ π2)(µ ∗ ϕ(1N)) ∈ B(H1 ⊗ H2) (or simply qπ1π2). 68 J. CRESPO We denote qπ1 ϕ := qπ1π1 ϕ (or simply qπ1) for short. Then, we have ϕ = ∑ qπ1π2 1(cid:54)l(cid:54)k F−1 l,j π1(e(l) 1(cid:54)i,j(cid:54)nl ∑ ij ) ⊗ π2(e(l) ji ), where, for all 1 (cid:54) l (cid:54) k, (e(l) Proof. For 1 (cid:54) l (cid:54) k, fix a s.u.m. (e(l) prove that ij )1(cid:54)i,j(cid:54)nl is a s.m.u. diagonalizing Fl. µ ∗ ϕ(1N) = ∑ 1(cid:54)l(cid:54)k ∑ 1(cid:54)i,j(cid:54)nl F−1 l,j e(l) ij ⊗ e(l) ji . ij )1(cid:54)i,j(cid:54)nl of Mnl (C) diagonalizing Fl. It suffices to (cid:78) Since 1N = ∑1(cid:54)l(cid:54)k ∑1(cid:54)i(cid:54)nl e(l) ii , it is enough to prove that µ ∗ ϕ(e(l) rs ) = ∑ 1(cid:54)j(cid:54)nl F−1 l,j e(l) rj ⊗ e(l) js , for all 1 (cid:54) r, s (cid:54) nl. ij ) be the family of N given by f (l) ij := F−1/2 l,j e(l) ij for all 1 (cid:54) l (cid:54) k and 1 (cid:54) i, j (cid:54) nl. It Let ( f (l) is clear that ( f (l) ij ) is an orthonormal basis of N. We have ϕ(e(l) sq ) = Trl(Fl e(l) sq ) = nl∑ i = 1 Fl,i Trl(e(l) ii e(l) sq ) = Fl,s Trl(e(l) sq ) = Fl,s δs q. We have ∗ ϕ(e(l) rs ) = µ = = = = nl(cid:48)(cid:48) ∑ p, q = 1 nl(cid:48)(cid:48) ∑ p, q = 1 k∑ l(cid:48), l(cid:48)(cid:48)= 1 k∑ l(cid:48), l(cid:48)(cid:48)= 1 k∑ l(cid:48)= 1 nl∑ nl(cid:48) ∑ i, j = 1 nl(cid:48) ∑ i, j = 1 nl(cid:48) ∑ i,j, q = 1 F−1 l, j F−1 j, q = 1 nl∑ j = 1 F−1 l, j e(l) rj ⊗ e(l) js . (cid:104)µ ∗ ϕ(e(l) rs ), f (l(cid:48)) ij ⊗ f (l(cid:48)(cid:48)) pq (cid:105) f (l(cid:48)) ij ⊗ f (l(cid:48)(cid:48)) pq δl(cid:48) l(cid:48)(cid:48) δ j p F−1 l(cid:48), j F−1 l(cid:48)(cid:48), q (cid:104)e(l) rs , e(l(cid:48)) iq (cid:105) e(l(cid:48)) ij ⊗ e(l(cid:48)(cid:48)) pq i F−1 l(cid:48), j F−1 l(cid:48), q ϕ(e(l) sq ) e(l(cid:48)) ij ⊗ e(l(cid:48)) jq δl l(cid:48) δr l, q Fl, s δs rj ⊗ e(l) q e(l) jq 8.2.16 Remarks. -- 1. For i = 1, 2, let γi : No → B(Ki) be a unital normal *-representa- ϕo ∈ B(K1 ⊗ K2) (or tion of No on a Hilbert space Ki. In a similar way, we define qγ1γ2 simply qγ1γ2) such that ϕo = ∑ qγ1γ2 1(cid:54)l(cid:54)k F−1 l,j γ1(e(l)o ij ) ⊗ γ2(e(l)o ∑ ), ji 1(cid:54)i,j(cid:54)nl where, for all 1 (cid:54) l (cid:54) k, (e(l) 2. It should be noted that qπ1π2 ϕ ij )1(cid:54)i,j(cid:54)nl is a s.m.u. diagonalizing Fl. If N is commutative (i.e. N = Ck), then qπ1π2 ϕ and qγ1γ2 ϕo are self-adjoint but not idempotent in general. (cid:78) are projections. and qγ1γ2 ϕo , qπγ case of the non-normalized markov trace. In this paragraph, we take for ϕ the non-normalized Markov trace on N = ⊕1(cid:54)l(cid:54)k Mnl (C), i.e.  = ⊕1(cid:54)l(cid:54)k nl · Trl. From now on, the operators qγπ o will be simply denoted by qγπ, qπγ, qπ1π2  and qγ1γ2. As a corollary of 8.2.14, we have: 8.2.17 Proposition. -- For all s.u.m. (e(l) ) ⊗ π(e(l) qγπ = ∑ 1(cid:54)l(cid:54)k ij )1(cid:54)l(cid:54)k, 1(cid:54)i,j(cid:54)nl of N, we have n−1 l ∑ ji ) and qπγ = ∑ 1(cid:54)i,j(cid:54)nl 1(cid:54)l(cid:54)k n−1 l ∑ 1(cid:54)i,j(cid:54)nl ij ) ⊗ γ(e(l)o o qπ1π2 and qγ1γ2 γ(e(l)o π(e(l) ). (cid:78) ij ji  ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS As a corollary of 8.2.15, we have: 8.2.18 Proposition. -- For all s.u.m. (e(l) ij )1(cid:54)l(cid:54)k, 1(cid:54)i,j(cid:54)nl of N, we have n−1 l ∑ 1(cid:54)i,j(cid:54)nl n−1 l ∑ 1(cid:54)i,j(cid:54)nl ij ) ⊗ π2(e(l) ) ⊗ γ2(e(l)o γ1(e(l)o π1(e(l) ji ) and ). ij ji qπ1π2 = ∑ 1(cid:54)l(cid:54)k qγ1γ2 = ∑ 1(cid:54)l(cid:54)k 69 (cid:78) The following result is a slight generalization of 8.2.17 to the setting of C*-algebras. 8.2.19 Proposition-Definition. -- (2.6 [2]) Let A, B be two C*-algebras. We consider two non-degenerate *-homomorphisms γA : No → M(A) and πB : N → M(B). There exists a unique self-adjoint projection qγAπB ∈ M(A ⊗ B) (resp. qπBγA ∈ M(B ⊗ A)) such that qγAπB = ∑ 1(cid:54)l(cid:54)k (resp. qπBγA = ∑ 1(cid:54)l(cid:54)k n−1 l ∑ 1(cid:54)i,j(cid:54)nl n−1 l ∑ 1(cid:54)i,j(cid:54)nl γA(e(l)o ij ) ⊗ πB(e(l) ji ) ij ) ⊗ γA(e(l)o ji )), πB(e(l) for all s.u.m. (e(l) ij )1(cid:54)l(cid:54)k, 1(cid:54)i,j(cid:54)nl of N. (cid:78) Proof. The uniqueness of such a self-adjoint projection is straightforward. In virtue of the Gelfand-Naimark theorem, we can consider faithful non-degenerate *-homomorphisms θA : A → B(K) and θB : B → B(H). Let us denote γ := θA ◦ γA and π := θB ◦ πB. Then, γ : No → B(K) and π : N → B(H) are normal unital *-representations. Let us fix ij )1(cid:54)i,j(cid:54)nl for Mnl (C) for each 1 (cid:54) l (cid:54) k. We define a self-adjoint an arbitrary s.u.m. (e(l) projection qγAπB ∈ M(A ⊗ B) by setting: n−1 l ∑ 1(cid:54)i,j(cid:54)nl qγAπB := ∑ 1(cid:54)l(cid:54)k ) ⊗ πB(e(l) ji ). γA(e(l)o ij By 8.2.17, we have qγπ = (θA ⊗ θB)(qγAπB ). By using again 8.2.17 and the fact that θA ⊗ θB is faithful, we obtain that qγAπB is independent of the chosen systems of matrix units. Moreover, the definition of qγAπB shows that qγAπB is also independent of the chosen faithful non-degenerate *-homomorphisms θA and θB. In a similar way, we have the following generalization of 8.2.18 to the setting of C*-algebras. 8.2.20 Proposition. -- For i = 1, 2, let Bi (resp. Ai) be a C*-algebra and πi : N → M(Bi) : No → M(Ai)) a non-degenerate *-homomorphism. Then, there exists a unique (resp. γi qπ1π2 ∈ M(B1 ⊗ B2) (resp. qγ1γ2 ∈ M(A1 ⊗ A2)) such that π1(e(l) qπ1π2 = ∑ 1(cid:54)l(cid:54)k (resp. qγ1γ2 = ∑ 1(cid:54)l(cid:54)k n−1 l ∑ 1(cid:54)i,j(cid:54)nl n−1 l ∑ 1(cid:54)i,j(cid:54)nl ij ) ⊗ π2(e(l) ji ) ) ⊗ γ2(e(l)o ji γ1(e(l)o ij )), (cid:78) for all s.u.m. (e(l) ij )1(cid:54)l(cid:54)k, 1(cid:54)i,j(cid:54)nl of N. In the following, we adopt a multi-index notation to simplify formulas and computations. 8.2.21 Notations. -- 1. Consider the index sets I := {(l, i, j) ; 1 (cid:54) l (cid:54) k, 1 (cid:54) i, j (cid:54) nl} and I0 := I (cid:116) {∅}. 2. For I = (l, i, j) ∈ I, we denote I := (l, j, i) ∈ I. Denote also ∅ := ∅. The map I0 → I0 ; I (cid:55)→ I is involutive. 70 J. CRESPO 3. A pair of indices (I, J) ∈ I × I is said to be composable if we have I = (l, i, m) and J = (l, m, j) for some indices 1 (cid:54) l (cid:54) k and 1 (cid:54) i, m, j (cid:54) nl. In this case, we denote I J := (l, i, j) ∈ I. We also denote I J = ∅ if I and J are not composable, I = ∅ or J = ∅. This defines a map I0 × I0 → I0 ; (I, J) (cid:55)→ I J. It is clear that I J = J I for all I, J ∈ I0. (cid:78) Let us fix a s.u.m. (e(l) ij )1(cid:54)l(cid:54)k, 1(cid:54)i,j(cid:54)nl of N. 8.2.22 Notations. -- 1. Denote by ε I := e(l) ij by eI := π(ε I) and fI := γ(εo n∅ := 1. Notice that we have nI = nI for all I ∈ I0. for I = (l, i, j) ∈ I and ε∅ := 0. Denote I ) for I ∈ I0. Denote by nI := nl for I = (l, i, j) ∈ I and I ∈ I. Note that x∗ = ∑I∈I xI · ε I. 2. Since (ε I)I∈I is a basis of N, for x ∈ N we denote x = ∑I∈I xI · ε I, with xI ∈ C for (cid:78) I = ε I and ε I ε J = ε I J. For all I, J ∈ I0, we have: 8.2.23 Remarks. -- 1. For all I, J ∈ I0, we have ε∗ f ∗ I = fI, fI ⊗ eI, qπγ = ∑I∈I n−1 e∗ I = eI, 2. We have qγπ = ∑I∈I n−1 eIeJ = eI J; qγ = ∑I∈I fI ⊗ fI. I fI fJ = fJI. I eI ⊗ fI, qπ = ∑I∈I eI ⊗ eI and (cid:78) 8.3 Unitary equivalence of Hilbert C*-modules In the following, we recall the notion of morphism between Hilbert modules over possibly different C*-algebras. 8.3.1 Definition. -- Let A and B be two C*-algebras and φ : A → B a *-homomorphism. Let E and F be two Hilbert C*-modules over A and B respectively. A φ-compatible operator from E to F is a linear map Φ : E → F such that: (i) for all ξ ∈ E and a ∈ A, Φ(ξa) = Φ(ξ)φ(a); (ii) for all ξ, η ∈ E, (cid:104)Φξ, Φη(cid:105) = φ((cid:104)ξ, η(cid:105)). Furthermore, if φ is a *-isomorphism and Φ is surjective, we say that Φ is φ-compatible (cid:78) unitary operator (or a unitary equivalence over φ) from E onto F. 8.3.2 Remarks. -- 1. It follows from (ii) that Φ : E → F is bounded and even isometric if φ is faithful. Indeed, we have (cid:107)(cid:104)Φξ, Φη(cid:105)(cid:107) = (cid:107)φ((cid:104)ξ, η(cid:105))(cid:107) = (cid:107)(cid:104)ξ, η(cid:105)(cid:107) for all ξ, η ∈ E. Then, for all ξ ∈ E we have (cid:107)Φξ(cid:107)2 = (cid:107)(cid:104)Φξ, Φξ(cid:105)(cid:107) = (cid:107)(cid:104)ξ, ξ(cid:105)(cid:107) = (cid:107)ξ(cid:107)2. In particular, if φ is a *-isomorphism and Φ is a φ-compatible unitary operator, then Φ is bijective and the inverse map Φ−1 : F → E is a φ−1-compatible unitary operator. 2. It is clear that idE is a idA-compatible unitary operator. Let A, B and C be C*-algebras and E, F and G be Hilbert modules over A, B and C respectively. Let φ : A → B and ψ : B → C be *-homomorphisms (resp. *-isomorphisms). If Φ : E → F is a φ-compatible operator (resp. unitary operator) and Ψ : F → G a ψ-compatible operator (resp. unitary operator), then Ψ ◦ Φ : E → G is a ψ ◦ φ-compatible operator (resp. unitary operator). 3. Let Φ : E → F be a unitary equivalence over a given *-isomorphism φ. If T ∈ L(E), then the map Φ ◦ T ◦ Φ−1 : F → F is an adjointable operator whose adjoint operator is Φ−1 ◦ T∗ ◦ Φ. We define a *-isomorphism L(E) → L(F) ; T (cid:55)→ Φ ◦ T ◦ Φ−1. Note that Φ ◦ θξ,η ◦ Φ−1 = θΦξ,Φη for all ξ, η ∈ E. In particular, for all k ∈ K(E) we have Φ ◦ k ◦ Φ−1 ∈ K(F). More precisely, the map K(E) → K(F) ; k (cid:55)→ Φ ◦ k ◦ Φ−1 is a (cid:78) *-isomorphism. The notion of unitary equivalence defines an equivalence relation on the class consisting of all Hilbert C*-modules (cf. 8.3.2 1, 2). Actually, this notion of morphism between Hilbert modules over possibly different C*-algebra can be understood in terms of unitary adjointable operator between two Hilbert modules over the same C*-algebra. ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS 71 8.3.3 Proposition. -- Let A and B be two C*-algebras and φ : A → B a *-isomorphism. Let E and F be two Hilbert C*-modules over A and B respectively. 1. If Φ : E → F is a surjective φ-compatible unitary operator, then there exists a unique unitary adjointable operator U ∈ L(E ⊗φ B, F) such that U(ξ ⊗φ b) = Φ(ξ)b, for all ξ ∈ E and b ∈ B. 2. Conversely, if U ∈ L(E ⊗φ B, F) is a unitary, then there exists a unique φ-compatible (cid:78) unitary operator Φ : E → F such that Φ(ξ)b = U(ξ ⊗φ b) for all ξ ∈ E and b ∈ B. As an application of the above proposition, we can the state the following result. 8.3.4 Proposition-Definition. -- Let A1, B1, A2 and B2 be C*-algebras, φ1 : A1 → B1 and φ2 : A2 → B2 *-isomorphisms. Let E1, F1, E2 and F2 be Hilbert C*-modules over A1, B1, A2 and B2 respectively. Let Φ1 : E1 → F1 and Φ2 : E2 → F2 be unitary equivalences over φ1 and φ2 respectively. Then, the linear map E1 (cid:12) E2 → F1 ⊗ F2 ; ξ1 ⊗ ξ2 (cid:55)→ Φ1(ξ1) ⊗ Φ2(ξ2) extends to a bounded linear map Φ1 ⊗ Φ2 : E1 ⊗ E2 → F1 ⊗ F2. Moreover, Φ1 ⊗ Φ2 is a φ1 ⊗ φ2-compatible (cid:78) unitary operator. The notion of unitary equivalence can also be understood in terms of isomorphism between the associated linking C*-algebras. 8.3.5 Proposition. -- Let A and B be two C*-algebras and φ : A → B a *-isomorphism. Let E and F be two Hilbert C*-modules over A and B respectively. 1. If Φ : E → F is a φ-compatible unitary operator, then there exists a unique *-homomorphism f : K(E ⊕ A) → K(F ⊕ B) such that f ◦ ιE = ιF ◦ Φ and f ◦ ιA = ιB ◦ φ. Moreover, f is a *-isomorphism. 2. Conversely, let f : K(E ⊕ A) → K(F ⊕ B) be a *-isomorphism such that f ◦ ιA = ιB ◦ φ. Then, there exists a unique map Φ : E → F such that f ◦ ιE = ιF ◦ Φ. Moreover, Φ is a (cid:78) φ-compatible unitary operator. Proof. 1. The *-homomorphism f : K(E ⊕ A) → K(F ⊕ B) is defined by (cf. 8.3.2 3): for all k ∈ K(E), ξ, η ∈ E and a ∈ A. (cid:18)Φ ◦ k ◦ Φ−1 Φξ (cid:19) := (cid:19) , (cid:18) k η∗ f ξ a (Φη)∗ φ(a) 2. This is a straightforward consequence of 2.3.4 1. 8.3.6 Notation. -- Let A, B be C*-algebras and E and F be two Hilbert C*-modules over unitary operator. If T ∈ L(A, E), we define the map (cid:101)Φ(T) := Φ ◦ T ◦ φ−1 : B → F. By a A and B respectively. Let φ : A → B be a *-isomorphism and Φ : E → F a φ-compatible straightforward computation, we show that (cid:101)Φ(T) ∈ L(B, F) whose adjoint operator is (cid:101)Φ(T)∗ = φ ◦ T∗ ◦ Φ−1. We have a bounded linear map (cid:101)Φ : L(A, E) → L(B, F), which is an extension of Φ up to the canonical injections E → L(A, E) and F → L(B, F). (cid:78) references [1] S. Baaj, Repésentation régulière du groupe quantique des déplacements de Worono- wicz, Astérisque 232 (1995), 11 -- 49. [2] S. Baaj et J. Crespo, Équivalence monoïdale de groupes quantiques et K-théorie bivariante, Bull. Soc. Math. France 145 (4) (2017), 711-802. [3] S. Baaj et G. Skandalis, C∗-algèbres de Hopf et théorie de Kasparov équivariante, K-theory 2 (1989), 683-721. [4] S. Baaj et G. Skandalis, Unitaires multiplicatifs et dualité pour les produits croisés de C∗-algèbres, Ann. Sci. Éc. Norm. Supér. 4e série, 26 (4) (1993), 425-488. [5] S. Baaj, G. Skandalis and S. Vaes, Non-semi-regular quantum groups coming from number theory, Comm. Math. Phys. 235 (1) (2003), 139-167. 72 J. CRESPO [6] J. Bichon, A. De Rijdt and S. Vaes, Ergodic coactions with large multiplicity and monoidal equivalence of quantum groups, Comm. Math. Phys. 262 (2006), 703-728. [7] A. Connes, On the spatial theory of von Neumann algebras, J. Funct. Anal. 35 (1980), 153-164. [8] A. Connes, Noncommutative Geometry, Academic Press, San Diego, CA, 1994. [9] J. Crespo, Monoidal equivalence of locally compact quantum groups and application to bivariant K-theory, Ph.D. thesis, Université Blaise Pascal (2015). [10] J. Crespo, Measured quantum groupoid on a finite basis and equivariant Kasparov theory, preprint: arXiv:1706.08516 [math.OA]. [11] K. De Commer, Monoidal equivalence for locally compact quantum groups, preprint: arXiv:math.OA/0804.2405v2. [12] K. De Commer, Galois coactions for algebraic and locally compact quantum groups, Ph.D. thesis, Leuven, Katholieke Universiteit Leuven (2009). [13] K. De Commer, Galois coactions and cocycle twisting for locally compact quantum groups, J. Operator theory 66 (1) (2011), 59-106. [14] K. De Commer, A. Freslon and M. Yamashita, CCAP for Universal Discrete Quan- tum Groups (with an appendix by S. Vaes), Comm. Math. Phys. 331 (2) (2014), 677-701. [15] A. De Rijdt and N. Vander Vennet, Actions of monoidally equivalent compact quantum groups and applications to probabilistic boundaries, Ann. Inst. Fourier 60 (1) (2010), 169-216. [16] M. Enock, Quantum groupoids of compact type, J. Inst. Math. Jussieu 4 (2005), 29-133. [17] M. Enock, Measured Quantum Groupoids in Action, Mém. Soc. Math. Fr. 114 (2008), 1-150. [18] G. G. Kasparov, Hilbert C∗-modules: theorems of Stinespring and Voiculescu, J. Operator Theory 4 (1) (1980), 133-150. [19] J. Kustermans and S. Vaes, Locally compact quantum groups in the von Neumann algebraic setting, Math. Scand. 92 (1) (2003), 68-92. [20] F. Lesieur, Measured Quantum Groupoids, Mém. Soc. Math. Fr. 109 (2007), 1-117. [21] S. Neshveyev and L. Tuset, Deformation of C*-algebras by cocycles on locally compact quantum groups, Adv. Math. 254 (2014), 454-496. [22] M. A. Rieffel, Induced representations of C*-algebras, Adv. Math. 13 (2) (1974), 176- 257. [23] T. Timmermann, Pseudo-multiplicative unitaries and pseudo-Kac systems on C∗- modules, 05/2005, Dissertation, Preprint des SFB 478 Münster (394). [24] T. Timmermann, Coactions of Hopf C∗-bimodules, J. Operator Theory 68 (1) (2012), 19-66. [25] S. Vaes, The unitary implementation of a locally compact quantum group action, J. Funct. Anal. 180 (2001), 426-480. [26] S. Vaes and N. Vander Vennet, Identification of the Poisson and Martin boundaries of orthogonal discrete quantum groups, J. Inst. Math. Jussieu 7 (2008), 391-412. [27] J.-M. Vallin, Unitaire pseudo-multiplicatif associé à un groupoïde. Applications à la moyennabilité, J. Operator Theory 44 (2) (2000), 347-368. [28] R. Vergnioux and C. Voigt, The K-theory of free quantum groups, Math. Ann. 357 (1) (2013), 355-400. ACTIONS OF MEASURED QUANTUM GROUPOIDS ON A FINITE BASIS 73 [29] C. Voigt, The Baum-Connes conjecture for free orthogonal quantum groups, Adv. Math. 227 (5) (2011), 1873-1913. Vrije Universiteit Brussel, Vakgroep Wiskunde, Pleinlaan 2, B-1050 Brussels (Belgium). E-mail addresses: [email protected], [email protected] Supported by the FWO grant G.0251.15N index of notations and symbols AlgG, category of G-C*-algebras, 22 A (cid:111) G, crossed product, 23 B (cid:111) (cid:98)G, crossed product, 23 C(−), 6 CM, 63 ∆k ij, 15 ij, 16 δk A, iterated coaction map, 21 δ2 D, bidual G-C*-algebra, 24 , 25 δk Aj (2) , 26 δ A1 , 42 δk Ej (2) , 44 E1 (ε1, ε2), standard basis of C2, 14 EA,L, 22 EB,λ, 23 EA,R, 24 ε I, 70 eI, fI, 70 GG1,G2, colinking measured quantum δ groupoid, 16 (A1), induced C*-algebra, 26 (E1), induced Hilbert module, 44 Hij, 15 ιA, ιE, ιE∗, ιK(E), canonical morphisms, 7 ij, 16 ιk IndG2 G1 IndG2 G1 I, multi-index set, 69 jD, 24 K, C*-algebra of compact operators on Kπ, Kγ, 65 L, R, ρ, λ, canonical representations of S the G.N.S. space L2(G), 17 and (cid:98)S, 13 Mnl (C), square matrices of order nl with (cid:102)M(E ⊗ B), relative multiplier module, 8 entries in C, 2 Mij, 15 ϕ , 63 Nϕ, M+ + , extended positive cone, 63 Next T , MT, 64 NT, M+ ω, 62 ωξ,η, 62 pij, 14 ϕij, ψij, 15 πL, 22 π,(cid:98)θ, 22 (cid:98)πλ, 23 (cid:98)π, θ, 23 (cid:101)πj, 27 (cid:101)Πj, 53 πR, 24 πk j , πk πj, 26 Πk Πj, 51 j , 40, 53 A,j, 25 ϕo, opposite weight, 63 ϕc, commutant weight, 65 qj, qA,j, 25 qI, 28 qE,j, 40 qγπ, qπγ, 67 qπ1π2, qπ1, 67 RG, unitary coinverse, 10 ξ , Lγ Rπ η , 65 S, (cid:98)S, weak Hopf C*-algebras, 13 Sij, 16 σγπ/ςγπ, relative flip map/*-homomorphism, 66 Trl, non-normalized Markov trace on Mnl (C), 11 Tξ, 31 12 ki, 15 U, 10(cid:98)V, V, (cid:101)V, 11 V, W, (cid:101)V, multiplicative partial isometries, ik, (cid:101)Vj jl, Wj Vi V, 24 j , 41, 52 Vk vγπ, canonical coisometry, 67 WG, pseudo-multiplicative unitary, 10 Z (−), center, 10
1510.08581
4
1510
2016-10-19T10:59:34
Composition of topological correspondences
[ "math.OA" ]
In the previous article, we proved that a topological correspondence $(X,\lambda)$ from a locally compact groupoid with a Haar system $(G,\alpha)$ to another one, $(H,\beta)$, produces a $C^*$-correspondence $\mathcal{H}(X)$ from $C^*(G,\alpha)$ to $C^*(H,\beta)$. In the present article, we describe how to form a composite of two topological correspondences when the bispaces are Hausdorff and second countable in addition to being locally compact.
math.OA
math
COMPOSITION OF TOPOLOGICAL CORRESPONDENCES ROHIT DILIP HOLKAR Abstract. In [5], we define a topological correspondence from a locally com- pact groupoid equipped with a Haar system to another one. In [5], we show that a topological correspondence, (X, λ), from a locally compact groupoid with a Haar system (G, α) to another one, (H, β), produces a C∗-correspondence H(X) from C∗(G, α) to C∗(H, β). In the present article, we describe how to form a composite of two topological correspondences when the bispaces are Hausdorff and second countable in addition to being locally compact. Contents Introduction 1. Preliminaries 1.1. Revision 1.2. Cohomology of proper groupoids 2. Composition of correspondences 2.1. Preparation for composition 2.2. Composition of topological correspondences 3. Examples References 1 5 5 6 8 8 10 19 21 Introduction Let (G, α) and (H, β) be locally compact groupoids with Haar systems. A topo- logical correspondence from (G, α) to (H, β) is a G-H-bispace X which is equipped with a continuous family of measures λ along the momentum map sX : X → H (0), and the action of H and the family of measures satisfy certain conditions (See [5, Definition 2.1]). We need that the action of H is proper, and the condition on λ is that it is H-invariant and each measure in λ is (G, α)-quasi-invariant. The groupoids G and H, and the space X are locally compact. Recall the definitions from [5]: we call a subset A ⊆ X of a topological space X quasi-compact if every open cover of A has a finite subcover, and A is called compact if it is quasi-compact and Hausdorff. The space X is called locally compact if every point x ∈ X has a locally compact neighbourhood. We call a topological groupoid G locally compact if G is a locally compact topological space and G(0) ⊆ G is Hausdorff. The main result in [5] says that a topological correspondence (X, λ) from (G, α) to (H, β) produces a C∗-correspondence H(X) from C∗(G, α) to C∗(H, β). Section 3 of [5] discusses many examples of topological correspondences. Two C∗-correspondences, K : A → B and F : B → C, may be composed to get a correspondence K ⊗B F : A → C. On the similar lines, consider two topological correspondences (X, α) and (Y, β) from (G1, χ1) to (G2, χ2) and (G2, χ2) to (G3, χ3), respectively. We describe the composite (Y, β) ◦ (X, α) : (G1, χ1) → (G3, χ3) when X and Y are Hausdorff and second countable in addition to being locally compact. In fact, our construction works when X and and Y are Hausdorff and the space 1 2 ROHIT DILIP HOLKAR (X ×sX ,G2 (0),rY Y )/G2 is paracompact; here sX and rX denote the momentum maps for the actions of G2 on X and Y , respectively. And the quotient is taken for the diagonal action of G2 on X ×sX ,G2 (0),rY Y . The composite (Y, β)◦ (X, α) should be a pair (Ω, µ) where Ω is a G1-G3-bispace, (0) and µ is a continuous family of measures along the momentum map sΩ : Ω → G3 the conditions in [5, Definition 2.1] are satisfied. Furthermore, we must have an isomorphism H(Ω) ≃ H(X) ⊗C∗(G2,χ2)H(Y ) of C∗-correspondences. The construction of Ω is well-known -- it is the quotient space (X ×sX ,G2 (0),rY Y )/G2 for diagonal action of G2 on X ×sX ,G2 (0),rY Y . The diagonal action is proper, since the action of G2 on X is proper. Thus the quotient space inherits all the nice properties of the fibre product such as Hausdorffness. The harder task is to get the continuous family of measures µ satisfying the required conditions. We need that µ := {µu}u∈G3 (0) is G3-invariant and each µu is (G1, χ1)-quasi- invariant. We explain how to get one such family of measures. The reason to write 'one such family of measures' is that the family is not unique; it depends on the choice of a certain continuous function on X ×sX ,G2 (0),rY Y . However, for any two such families of measures the corresponding C∗-correspondences are naturally isomorphic to H(X) ⊗C∗(G2,χ2) H(Y ). The construction of µ is one of the most technical part of this article. To ex- plain the problem, motivation and idea of constructing the composite of families of measures, we have to do a computation and discuss some technical ideas. Denote (0),rY Y by Z. Then Z carries a G3-invariant continuous family the space X ×sX ,G2 of measures m := {mu}u∈G(0) 3 which is given by (0.1) ZZ f dmu = ZY ZX f (x, y) dαrY (y)(x) dβu(y) for f ∈ Cc(Z). Let π : Z → Ω be the quotient map and λ be the continuous family of measures along it defined as (0.2) Zπ−1([x,y]) f dλ[x,y] := ZG rY (y) 2 f (xγ, γ−1y) dχrY (y) 2 (γ). for f ∈ Cc(Z), and [x, y] ∈ Ω which is the equivalence class of (x, y) ∈ Z. A very natural choice for µ is that it is the family of measures on Ω which gives the disintegration m = µ ◦ λ. Furthermore, one may expect that the isomorph- ism H(X) ⊗C∗(G2,χ2)H(Y ) ≃ H(Ω) is induced by the map Ψ : Cc(Z) → Cc(Ω) where Ψ(F )([x, y]) = RG2 (γ). To be more explicit, we view Cc(Z) and Cc(Ω) as pre-Hilbert C∗(G3, χ3)-modules which complete to the Hilbert C∗(G3, χ3)-modules H(X) ⊗H(Y ) and H(Ω). And the map Ψ : Cc(Z) → C(Ω) is expected to induce the required isomorphism of Hilbert C∗(G3, χ3)-modules which also gives the desired isomorphism of C∗-correspondences. F (xγ, γ−1y) dχsX (x) 2 However, that is not exactly the case. Consider the following example: let G be a group and H a closed proper subgroups of G. Let α and κ be the Haar measures on G and H, respectively. Then (G, α−1) is a topological correspondence from (G, α) to (H, κ) which is called the induction correspondence in [5, Example 3.13]. The constant function 1 is the adjoining function for this correspondence. Let X be a left H-space carrying an (H, κ)-quasi-invariant measure β. Let ∆X denote the 1-cocycle on the transformation groupoid H ⋉ X that gives the quasi-invariance. Assume that ∆X is continuous. Then (X, β) is a topological correspondence from (H, κ) to the trivial group Pt, see [5, Example 3.6]. The adjoining function of this correspondence is ∆X . Furthermore, H(X) = L2(X, β) and the action of H induces the representation of C∗(H) on L2(X, β). Thus we COMPOSITION OF TOPOLOGICAL CORRESPONDENCES 3 have (G, α) (G,α−1) −−−−−→ (K, κ) (X,β) −−−→ Pt. Let Z, Ω, π, m, λ and µ have the similar meaning as in the above discussion. Then in this situation, Z = G×X, Ω = (G×X)/K, m = α−1×β and π : G×X → (G×X)/K the quotient map. For f ∈ Cc(Z) Equation (0.2) now reads Zπ−1([γ,x]) f dλ[γ,x] = ZH f (γη, η−1x) dκ(η). What are the necessary and sufficient conditions get a measure µ on (G × X)/K satisfying α−1 × β = m = µ ◦ λ? We may draw a square as in Figure 3 comprising of the spaces, maps and meas- ures discussed above. And then (i) of Proposition 2.2 implies that if there is such a measures µ, then the equality m ◦ κ = m ◦ κ−1 must hold -- this is the necessary condition. Recall from the discussion above that m = α−1 × β. Thus we must have (α−1 × β) ◦ κ = (α−1 × β) ◦ κ−1 on Z × K = G × X × K. Let f ∈ Cc(G × X × K), then a direct computation gives that (α−1 × β) ◦ κ(f ) = ZGZX ZK f (γ, x, η) dκ(η) dβ(x) dα−1(γ). On the other hand, (α−1 × β) ◦ κ−1(f ) = ZGZX ZK f (γ, x, η−1) dκ(η) dβ(x) dα−1(γ). Now (i) first apply Fubini's theorem to dκ dβ, (ii) then change the variable (γ, x, η−1) to (γη, η−1x, η), (iii) then use the (H, κ)-quasi-invariance of β and the right invari- ance of α−1 and (iv) finally apply Fubini's theorem to dβ dκ to see that the last term equals ZGZX ZK f (γ, η−1x, η) ∆X (η, η−1x) dκ(η) dβ(x) dα−1(γ). Thus the 1-cocycle ∆X on the transformation groupoid H ⋉X is the obstruction for the measures (α−1 × β) ◦ κ and (α−1 × β) ◦ κ−1 to be equal. (ii) of Proposition 2.2 says that equality of these measures, m ◦ κ = m ◦ κ−1, is also a sufficient condition in our situation for the measure µ to exist. A similar problem appears in the general setting; there the cocycle ∆X is replace by the adjoining function of the second correspondence involved in the composition. How to overcome this obstruction? Let (X, α) : (G1, χ1) → (G2, χ2) and (Y, β) : (G2, χ2) → (G3, χ3) be topological correspondences and let ∆2 be the adjoining function of (Y, β). Then we realise ∆2 as a 1-cocycle on the proper groupoid (Z ⋊ G2) and decompose it into a quotient ∆2 = b ◦ sZ⋊G2 /b ◦ rZ⋊G2 for a 0-cochain b on Z ⋊ G2. Here sZ⋊G2 and rZ⋊G2 denote the source and the range maps of Z ⋊G2, respectively. Using Proposition 2.2 we show that there is a unique measure µ which gives the disintegration bm = µ ◦ λ. We modify the map Ψ : Cc(Z) → Cc(Ω) discussed above (the discussion following Equation (0.2) on page 2) to consider the 0-cochain b. Then this µ and modified Ψ produce the desired isomorphism of C∗-correspondences. In this construction, the required 0-cochain b, which is a function of Z, is not unique. However, as mentioned earlier, for any two such 0-cochains the C∗-correspondences associated with the composites are isomorphic to H(X) ⊗C∗(G2,χ2) H(Y ). Given two 0-cochains b and b′ which decompose ∆2 as above, with some slight work, one may show that there is a positive continuous function c on Z with b = cb′. This explains the isomorphism of C∗-correspondences associated with the two composites. 4 ROHIT DILIP HOLKAR The theory of groupoid cohomology which we need is developed in [5, Section 1] and we prove the other necessary results in this article. The main result we need is Proposition 1.10 which asserts that the first R-valued cohomology of a proper groupoid is trivial. Lemma 1.7 is the main tool to prove this proposition. This lemma says that a locally compact proper groupoid G equipped with a Haar system and G\G(0) paracompact carries an invariant family of probability measures. Anantharaman Delaroche and Renault introduced the notions of amenability for a measured Borel groupoid ([1, Definition 3.2.8]), measurewise amenability for a Borel groupoid ([1, Definition 3.3.1]) and topological amenability for a locally compact topological groupoid ([1, Definition 2.2.7]). In [1, Proposition 3.3.5], they prove that topological amenability implies measurewise amenability. The definition of measurewise amenability implies that if G is measurewise amenable, then for any Borel Haar system on G and a quasi-invariant measure for the Haar system, G is an amenable measured groupoid. Our Lemma 1.7 implies that a locally compact proper groupoid with a Haar system (G, α) and G/G(0) paracompact is topologically amenable. Anantharaman Delaroche and Renault prove this fact -- [1, Proposition 3.3.5] says that topolo- gical amenability implies measurewise amenability. [1, Proposition 2.2.5] implies Lemma 1.7 also. But the proposition proves a general statement than the lemma, and both proofs are different. Our proof is a simple minded one. As shown in [5, Example 3.10 and 3.7], the generalized morphism defined by Buneci and Stachura in [3], and the topological correspondences for the groupoids with Hausdorff space of units introduced by Tu in [8], respectively, are topological correspondences. Our construction of composite matches the ones described by Tu [8], and Buneci and Stachura [3], see Examples 3.5 and 3.6, respectively. We discuss some examples at the end of the article. A continuous map f : X → Y of spaces gives a topological correspondence (X, δX ) from Y to X ([5, Example 3.1]). Here δX = {δx}x∈X is the familiy of measures where each δx is the point mass at x ∈ X. A continuous group homomorphism φ : G → H gives a topological corres- pondences (H, β−1) from G to H ([5, Example 3.4]) where β is the Haar measure on H. If g : Y → Z is a function, then Example 3.1 shows that the composite of the topological correspondences (X, δX ) and (Y, δZ ) is the correspondence obtained from the function g ◦ f : X → Z. Example 3.3 shows that a similar result holds for group homomorphisms. Both these examples agree with the well-known behaviour of the C∗-functor for spaces and groups. Let G be a locally compact group, and let H and K be closed subgroups of G. Example 3.7 shows how to use topological correspondences to induce a topological representation of K to H. Most of our terminology, definitions, hypotheses and notation are defined in [5]. Now we describe the structure of the article briefly. In the first section, we revise few definitions, notation and results in [5]. We prove that every locally compact proper groupoid equipped with a Haar system carries an invariant continuous family of probability measures. Then using this result we prove that the first cohomology of a proper groupoid is trivial. In the second section, we describe the composition of topological correspondences and prove the main result Theorem 2.18 which says that the C∗-correspondence as- sociated with a composite is isomorphic to the composite of the C∗-correspondences. The last section contains examples. Most of the examples are related to the ones in [5]. COMPOSITION OF TOPOLOGICAL CORRESPONDENCES 5 1. Preliminaries 1.1. Revision. The symbols ≃, ≈, R+ and R+ ∗ stand for isomorphic, homeomorphic, the set of positive real numbers and the multiplicative group of positive real num- bers, respectively. The symbol ⊗ and ⊗ indicate the algebraic tensor product modules and the interior tensor product Hilbert modules, respectively. We work with continuous families of measures and all the measures are assumed to be positive, Radon and σ-finite. The families of measures are denoted by small Greek letters and the corresponding integration function that appears in the con- tinuity condition is denoted by the Greek upper case letter used to denote the family of measures. For example, if λ is a family of measures along a map f : X → Y , then Λ : Cc(X) → Cc(Y ) is the function Λ(f )(y) = RX f dy. The capitalisations of α, β, χ and µ are A, B, χ and M , respectively. However, for a single measure on a space, which is a family of measures along the constant map onto a point, we follow the traditional convention, that is, the same letter is used to denote the measure and the corresponding integration functional. For example, if α is a measure on X, then α(f ) = RX f dα for f ∈ Cc(X). Let G be a groupoid, then rG, sG and invG denote the source, range and the in- version maps for G. Given a left G-space X, we tacitly assume that the momentum map for the action is rX . If X is a right G-space, then sX is the momentum map for the action. For A, B ⊆ G(0) we define GA = r−1 We denote G×sG,G(0),rX X, the fibre product for G and X over G(0) along sG and rX , by G ×G(0) X. If X is a right G-space, then X ×G(0) G has a similar meaning. For a left G-space X, G ⋉ X is the transformation groupoid and its set of arrows is the fibre product G ×G(0) X. Similar is the meaning of X ⋊ G a right G-space X. B = GA ∩ GB. When A = {u} and B = {v} are singletons, we simply write Gu, Gv and Gu v instead of G{u}, G{v} and G{u} {v} , respectively. Let X and Y be left and right G-spaces, respectively, and let A ⊆ G(0) and u ∈ G(0). Then X A, X u, YB and Yu have the similar obvious meanings. G (A), GB = s−1 G (B) and GA We denote a C∗-correspondence only by the Hilbert module involved in it; we do not write the representation of the left C∗-algebra. Thus we say 'H is a C∗-correspondence from a C∗-algebra A to B', and not '(H, φ) is a C∗-correspondence from a C∗-algebra A to B' where φ : A → BB(H) is the nondegenerate *-representation involved in the definition of the correspondence. We also write 'H : A → B is a C∗-correspondence'. Now we sketch the process of composing the C∗-correspondences briefly and explain a few notation along the way. Let A, B and C be C∗-algebras, and let H : A → B and F : B → C be C∗-correspondences. Endow H ⊗C F with the inner product hζ ⊗ ξ , ζ′ ⊗ ξ′i = hξ , hζ , ζ′i ξ′i. Let N ⊆ H ⊗C F be the closed vector subspace of the vectors of zero norm, that is, N = {z ∈ H ⊗C F : hz , zi = 0}. The proof of Proposition 4.5 in [6] shows that the subspace N is same as the the subspace spanned by the elements of the form ζb ⊗ ξ − ζ ⊗ bξ where ζ ∈ H, ξ ∈ F and b ∈ B. The Hilbert C-module H ⊗BF is the completion of (H ⊗C F )/N in the norm induced by h,i. We denote the equivalence class of ζ ⊗ ξ ∈ H ⊗C F in H ⊗BF by ζ ⊗ξ. The action of A on H ⊗BF is a(ζ ⊗ξ) = aζ ⊗ξ where a ∈ A and ζ ⊗ξ ∈ H ⊗F. We call the map ξ ⊗C ζ 7→ ξ ⊗ζ, H ⊗C F → H ⊗F, the obvious map of Hilbert C-modules which, clearly, has a dense image. Definition 1.1 (Topological correspondence). A topological correspondence from a locally compact groupoid G with a Haar system α to a locally compact groupoid H equipped with a Haar system β is a pair (X, λ), where: i) X is a locally compact G-H-bispace, 6 ROHIT DILIP HOLKAR ii) the action of H is proper, iii) λ = {λu}u∈H(0) is an H-invariant proper continuous family of measures along the momentum map sX : X → H (0), iv) there exists a continuous function ∆ : G⋉X → R+ such that for each u ∈ H (0) and F ∈ Cc(G ×G(0) X), ZXu ZGrX (x) F (γ−1, x) dαrX (x)(γ) dλu(x) = ZXu ZGrX (x) F (γ, γ−1x) ∆(γ, γ−1x) dαrX (x)(γ) dλu(x). The function ∆ is unique and is called the adjoining function of the correspond- ence. For φ ∈ Cc(G), f ∈ Cc(X) and ψ ∈ Cc(H) define the functions φ · f and f · ψ on X as follows: (1.2)   (φ · f )(x) := ZGrX (x) (f · ψ)(x) := ZHsX (x) φ(γ)f (γ−1x) ∆1/2(γ, γ−1x) dαrX (x)(γ), f (xη)ψ(η−1) dβsX (x)(η). For f, g ∈ Cc(X) define the function hf, gi on H by (1.3) hf, gi(η) := ZXrH (η) f (x)g(xη) dλrH (η)(x). Very often we write φf and f ψ instead of φ · f and f · ψ, respectively. [5, Lemma 2.10] proves that φf, f ψ ∈ Cc(X) and hf , gi ∈ Cc(H). Theorem 1.4 ([5, Theorem 2.39]). Let (G, α) and (H, β) be locally compact group- oids with Haar systems. Then a topological correspondence (X, λ) from (G, α) to (H, β) produces a C∗-correspondence H(X) from C∗(G, α) to C∗(H, β). 1.2. Cohomology of proper groupoids. In this subsection, we show that the first continuous cohomology group with real coefficients is trivial. It can be readily checked that the result is valid for the groupoid equivariant continuous cohomology introduced in [5, Section 1], and also for the (equivariant) Borel cohomology of a proper (topological) groupoid. Lemma 1.5 (Lemma 1, Appendix I in [2]). Let X be a locally compact Hausdorff space, R an open equivalence relation in X, such that the quotient space X/R is paracompact; let π be the canonical mapping of X onto X/R. There is a continuous real-valued function F ≥ 0 on X such that: i) F is not identically zero on any equivalence class with respect to R; ii) for every compact subset K of X/R, the intersection of π−1(K) with supp(F ) is compact. A continuous map f : X → Y is proper, if for each y ∈ Y , f −1(y) ⊆ X is quasi- compact. We call a locally compact groupoid proper if the map (rG, sG) : G → G(0) × G(0), (rG, sG)(γ) = (rG(γ), sG(γ)), is proper. Proposition 1.6 ([8, Proposition 2.10]). Let G be a locally compact groupoid. Then the following assertions are equivalent. i) G is proper; ii) the map (rG, sG) : G → G(0) × G(0) is closed and for each u ∈ G(0), Gu u ⊆ G is quasi-compact; iii) for all quasi-compact subsets K, L ⊆ G(0), GL iv) for all compact subsets K, L ⊆ G(0), GL K is compact; K is quasi-compact; COMPOSITION OF TOPOLOGICAL CORRESPONDENCES 7 v) for every quasi-compact subsets K ⊆ G(0), GK vi) for all x, y ∈ G(0), there are compact neighbourhoods Kx and Ly of x and K is quasi-compact; y, respectively,such that GKx Ly is quasi-compact. [8, Proposition 2.10] is stated for groupoids which do not, necessarily, have Haus- dorff space of units in which case (i)-(v) are equivalent and (v) =⇒ (vi). Lemma 1.7. Let (G, α) is a locally compact proper groupoid with a Haar system. If G\G(0) is paracompact then there is a continuous invariant family of probability measures on G. Furthermore, each measure in this family has a compact support. Proof. Since G has a Haar system, the range map of G is open, hence the quotient map π : G(0) → G\G(0) is open. Since G is proper, G\G(0) is locally compact and Hausdorff. By hypothesis G\G(0) is paracompact. Now we apply Lemma 1.5 to get a function F on G(0) such that F is not identically zero on any G-orbit in G(0) and for every compact K ⊆ G\G(0) the intersection supp(F ) ∩ π−1(K) is compact. Define h : G(0) → R+ by h(u) = ZGu F ◦ sG(γ) dαu(γ). Property (ii) of F from Lemma 1.5 and the full support condition of αu imply that h(u) > 0. To see that h(u) < ∞, notice that supp(F ◦ sG) ∩ Gu ⊆ G is compact: γ ∈ supp(F ◦ sG) ∩ Gu =⇒ sG(γ) ∈ supp(F ) and rG(γ) = u. Thus if u ⊆ G(0) denotes the orbit of u, then supp(F ◦ sG) ∩ Gu ⊆ (rG, sG)−1({u} × supp(F u)). Property (ii) of F from Lemma 1.5 says that supp(F u) is compact. As G is a proper groupoid, the set (rG × sG)−1({u} × supp(F u)) is compact which implies that supp(F ◦ sG) ∩ Gu is compact. Using the invariance of α, it is not hard to see that the function h is constant on the orbits of G(0). Put F ′ = (F/h) ◦ sG, then (1.8) ZGu F ′(γ) dαu(γ) = 1. Denote F ′ αu by pu, then p := {pu}u∈G(0) is a family of probability measures on G. Explicitly, p is given by ZGu f dpu = ZGu f (γ) F ′(γ) dαu(γ) for f ∈ Cc(G). It follows from the definition of measure pu, where u ∈ G(0), that the the compact set supp(F ◦ sG) ∩ Gu is the support of pu. To check that p is invariant, let f ∈ Cc(G) and η ∈ G, by the definition of p we have ZGsG(η) f (ηγ) dpsG(η)(γ) = ZGsG(η) f (ηγ) F ′(γ) dαsG(η)(γ). Now change the variable ηγ 7→ γ to that the previous term equals ZGrG(η) f (γ) F ′(η−1γ) dαrG(η)(γ). Use the invariance of α and the fact that F ′(η−1γ) := F ◦ sG(η−1γ)/h ◦ sG(η−1γ) = F ◦ sG(γ)/h ◦ sG(γ) = F ′(γ) and compute further: ZGrG(η) f (γ) F ′(η−1γ) dαrG(η)(γ) = ZGsG(η) = ZGsG(η) f (γ) F ′(γ) dαrG(η)(γ) f (γ) dprG(η)(γ). (cid:3) 8 ROHIT DILIP HOLKAR Remark 1.9. Anantharaman Delaroche and Renault define topological amenability for a locally compact topological groupoid ([1, Definition 2.2.7]). Lemma 1.7 says that every proper groupoid G with a Haar system and G\G(0) paracompact is topologically amenable. Proposition 1.10. Let G be a locally compact proper groupoid and α a Haar system on G. Then every R-valued 1-cocycle is a coboundary, that is, H1(G; R) = 0. Proof. Let p = {pu}u∈G(0) be an invariant family of probability measures on G which is obtained using Lemma 1.7. We claim that for a 1-cocycle c : G → R, the function b(u) = ZG c(γ) dpu(γ) for u ∈ G(0) satisfies c = b ◦ s − b ◦ r. Lemma 1.7 says that the support of each measure in p is compact, hence the above integral is well-defined. To see that b is the desired cochain, let η ∈ G and compute: c(γ) dps(η)(γ) −ZGrG(η) c(γ) dps(η)(γ) −ZGsG(η) (b ◦ s − b ◦ r)(η) = ZGsG(η) = ZGsG(η) = ZGsG(η) = c(η)Z dps(η)(γ) = c(η). (c(γ) − c(η) + c(γ)) dps(η)(γ) c(γ) dpr(η)(γ) c(ηγ) dps(η)(γ) We used the invariance of p to get the second equality above. (cid:3) 2. Composition of correspondences 2.1. Preparation for composition. Let G be a locally compact proper groupoid with G/G(0) paracompact. Since G is proper the action of G on G(0) given by γ · sG(γ) = rG(γ) for γ ∈ G is proper. Let λ be a Haar system on G. Then λ induces a family of measures [λ] along π : G(0) → G/G(0). And [λ] is defined as (2.1) ZG(0) f d[λ][u] = ZG f (γ−1 · u)) dλu(λ) = ZG f ◦ sG(γ) dλu(γ) for f ∈ Cc(G(0)). Note that in Equation 2.1, γ−1·u does not stand for the composite of γ−1 and u but for the action of G on G(0), that is, γ−1 · u = rG(γ−1) = sG(γ). We draw Figure 1 which contains all this data. G rGλ G(0) λ−1 sG [λ] π G(0) π [λ] G/G(0) Figure 1 A measure m on G(0) induces measures m ◦ λ and m ◦ λ−1 on G. For f ∈ Cc(G) ZG f d(m ◦ λ−1) := ZG(0) ZG f (γ−1) dλu(γ) dm(u). The measure m ◦ λ is defined similarly. We call the measure m on G(0) invariant with respect to (G, λ) if m ◦ λ = m ◦ λ−1 and the measure m ◦ λ on G is called symmetric. COMPOSITION OF TOPOLOGICAL CORRESPONDENCES 9 Proposition 2.2. Let G be a proper groupoid with G/G(0) paracompact, let λ be a Haar system for G and π : G(0) → G/G(0) the quotient map. i) Let µ be a measure on G/G(0) and let m denote the measure µ ◦ [λ] on G(0). Then m is an invariant measure. ii) Let m be a measure on G(0). If m is invariant, then there is a measure µ on G/G(0) with µ ◦ [λ] = m. iii) The measure µ in (ii), with µ ◦ [λ] = m, is unique. Proof. (i): Let f ∈ Cc(G), then ZG f d(m ◦ λ) = ZG f d(µ ◦ [λ] ◦ λ) = ZG/G(0) ZG(0) Λ(f )(u) d[λ][u](u) dµ([u]) Λ(f ) ◦ sG(γ) dλu(γ) dµ([u]) = ZG/G(0) ZGu = ZG/G(0) ZGu (cid:18)ZGsG(γ) f (η) dλsG (γ)(η)(cid:19) dλu(γ) dµ[u]. We know that sG(γ) = rG(γ−1). Now change the variable η 7→ γ−1η, and use the left invariance of λ to see that the previous term equals ZG/G(0) ZGZG f (γ−1η) dλrG(γ)(η)dλu(γ) dµ([u]). We have removed the superscripts of G in above equation for simplicity. Now apply Fubini's theorem to dλrG(γ)(η) dλu(γ) which is allowed since rG(γ) = u and f is compactly supported continuous function. Moreover, note that u = rG(γ) = sG(γ−1), use the right invariance of λ−1 and compute further: ZG/G(0) ZGZG = ZG/G(0) ZGZG = ZG/G(0) ZG = ZG/G(0) f (γ−1η) dλsG(γ−1)(γ) dλu(η) dµ([u]) f (γ−1) dλsG(η)(γ) dλu(η) dµ([u]) Λ−1(f ) ◦ sG(η) dλu(η) dµ([u]) f d(µ ◦ [λ] ◦ λ−1) = ZG f d(m ◦ λ−1). To be precise, in the last the first equality is obtained by using the right invariance of λ−1. (ii): Let e be a function on G(0) which is similar to function F/h in Lemma 1.7, and thus Λ(e ◦ sG) = 1. Function e ◦ sG is like function F ′ in Equation 1.8. Let µ be the measure on G/G(0) which is defined as µ(g) = m((g ◦ π) · e) for g ∈ Cc(G/G(0)). Let [Λ] denote the integration function corresponding to [λ]. For f ∈ Cc(G(0)) ZG(0) f d(µ ◦ [λ]) = ZG/G(0) = ZG(0) ZG [Λ](f )([u]) dµ([u]) = ZG(0) f ◦ sG(γ) e(rG(γ)) dλu(γ) dm(u). [Λ](f ) ◦ π(u) e(u) dm(u) We change γ 7→ γ−1, then use the symmetry of the measure m ◦ λ and continue the computation: ZG(0) ZG f ◦rG(γ) e(sG(γ)) dλu(γ) dm(u) = ZG(0) f (u) Λ(e◦sG)(γ) dm(u) = Z f dm. The last equality is due to the property of e that Λ(e ◦ sG) = 1. 10 ROHIT DILIP HOLKAR (iii): Let µ′ be another measure on G/G(0) which satisfies the condition µ′ ◦ [λ] = m. Since the integration map [Λ] : Cc(G(0)) → Cc(G/G(0)) is surjective, µ◦[λ] = µ′ ◦[λ] implies µ = µ′. (cid:3) Now we study the case when the measure m is not invariant, but strongly quasi- invariant; m is called quasi-invariant with respect to (G, λ) if m ◦ λ ∼ m ◦ λ−1. Following Folland, see [4, Chapter 2, Section 6, page 58], we call m strongly quasi- invariant with respect to (G, λ) if there is a continuous homomorphism ∆ : G → R+ ∗ with m ◦ λ = ∆ · (m ◦ λ−1), that is, m is quasi-invariant with respect to (G, λ) and the Radon-Nikodym derivative implementing the equivalence of the measures m ◦ λ and m◦λ−1 is continuous. Very often, when there is no chance of confusion, we drop the phrase 'with respect to (G, λ)' while talking about a (strongly) quasi-invariant measure. The cohomology theory of groupoids tells us that a homomorphism from G to an abelian group R is same as an R-valued 1-cocycle (See [5, Section 1]). Let (G, λ) be as in Proposition 2.2. Let m be a strongly quasi-invariant measure on G(0), and let ∆ be the R+ ∗ -valued continuous 1-cocycle which implements the quasi-invariance. Then ∆ gives an R-valued 1-cocycle log ◦∆ : G → R. Proposi- tion 1.10 says that log ◦ ∆ = b◦sG−b◦rG for some continuous function b : G(0) → R. Thus ∆ = eb◦sG eb◦rG . Write b = eb, then b > 0 and it can be checked that ∆ = eb◦sG eb◦rG = eb ◦ sG eb ◦ rG = b ◦ sG b ◦ rG . Rewriting the definition of (G, λ)-quasi-invariance of m using the above value of b◦rG(cid:17) m ◦ λ−1 which is equivalent to (b ◦ rG)(m ◦ λ) = ∆ gives that m ◦ λ = (cid:16) b◦sG (b◦sG)(m◦λ−1). A straightforward calculation shows that (b◦rG)(m◦λ) = (bm)◦λ and (b ◦ sG)(m ◦ λ−1) = (bm) ◦ λ−1. Thus we get Proposition 2.3. Let (G, λ) be a locally compact proper groupoid with a Haar system. Assume that G/G(0) is paracompact. Let m be a strongly (G, λ)-quasi- invariant measure on G(0). Let ∆ be the R+ ∗ -valued continuous 1-cocycle which implements the quasi-invariance. Then there is a continuous function b : G(0) → R+ with b◦rG(γ) = ∆(γ) for all γ ∈ G; i) b◦sG(γ) ii) the measure bm on G(0) is (G, λ)-invariant, that is, bm ◦ λ = bm ◦ λ−1. 2.2. Composition of topological correspondences. Let (X, α) and (Y, β) be correspondences from (G1, χ1) to (G2, χ2) and from (G2, χ2) to (G3, χ3), respect- ively. Let ∆1 and ∆2 be the adjoining functions of (X, α) and (Y, β), respectively. Additionally, assume that X and Y are Hausdorff and second countable. We draw Figure 2 that comprises of this data. X ∆1 α ∆2 Y β (G1, χ1) (G2, χ2) (G3, χ3) Figure 2 COMPOSITION OF TOPOLOGICAL CORRESPONDENCES 11 We need to create a G1-G3-bispace Ω equipped with a G3-invariant continuous family of measures µ = {µu}u∈H(0) with each µu G1-quasi-invariant. And the C∗(G1, χ1)-C∗(G3, χ3)-Hilbert module H(Ω) should be isomorphic to the Hilbert module H(X) ⊗C∗(G2,χ2)H(Y ). 3 Denote the fibred product X ×G2 (0) Y by Z. Then Z carries the diagonal action of G2. Since the action of G2 on X is proper, its action on Z is proper. Thus the the transformation groupoid Z ⋊ G2 is proper. We define the space Ω = Z/G2 = (Z ⋊ G2)(0)/(Z ⋊ G2). Since Z is locally compact, Hausdorff and second countable, so is Ω. Being a locally compact, Hausdorff and second countable space, Ω is paracompact. The following discussion in this section goes through under a milder hypothesis, namely, X and Y are locally compact Hausdorff and Ω is paracompact. Observation 2.4. The space Z is a G1-G3-bispace. The momentum maps are rZ (x, y) = rX (x) and sZ (x, y) = sY (y). For (γ1, (x, y)) ∈ G1 ×G1 (0) Z and ((x, y), γ3) ∈ Z ×G(0) G3, the actions are γ1 ·(x, y) = (γ1x, y) and (x, y)·γ3 = (x, yγ3), respectively. These actions descend to Ω and make it a G1-G3-bispace. Thus rΩ([x, y]) = rX (x) and sΩ([x, y]) = sY (y) and γ1[x, y]γ2 = [γ1x, yγ2] for appropri- ate γ1 ∈ G1, [x, y] ∈ Ω and γ2. 3 Lemma 2.5. The right action of G3 on Ω defined in Observation 2.4 is proper. Proof. Follows from [8, Lemma 2.33]. (cid:3) For each u ∈ G(0) 3 define a measure mu on the space Z as follows: for f ∈ Cc(Z) ZZ f dmu = ZY ZX f (x, y) dαrY (y)(x) dβu(y). Lemma 2.6. The family of measures {mu}u∈G(0) family of measures on Z. 3 is a G3-invariant continuous Proof. It is a routine computation to check that the G3-invariance of the family of measures β makes {mu}u∈G(0) 3 G3-invariant. The computation is similar to that in Proposition 2.13. To check the continuity let f ∈ Cc(X) and g ∈ Cc(Y ), then ZZ f ⊗ g dmu = B((A(f ) ◦ rY ) g)(u). (0)). Now use the theorem of Stone-Weierstrass to see that the set which is in Cc(G3 {f ⊗ g : f ∈ Cc(X), g ∈ Cc(Y )} ⊆ Cc(Z) is dense which concludes the lemma. (cid:3) The Haar system χ2 of G2 induces a Haar system χ on Z ⋊G2; for f ∈ Cc(Z ⋊G2) and (x, y) ∈ Z Z(Z⋊G2)(x,y) f dχ(x,y) := ZG sX (x) 2 f ((x, y), γ) dχsX (x) 2 (γ). The quotient map π : Z → Ω carries the family of measures [χ]; the definition of which is similar to the one in Equation 2.1. We write λ instead of [χ], and λω instead of [χ]ω for all ω ∈ Ω. Recall from Subsection 2.1 that for f ∈ Cc(Z) and ω = [x, y] ∈ Ω Zπ−1(ω) f dλω := ZG rY (y) 2 f (xγ, γ−1y) dχrY (y) 2 (γ). We wish to prove that, up to equivalence, {mu}u∈G3 (0) can be pushed down from Z to Ω to a G3-invariant family of measures {µu}u∈G3 (0) . We use λ to achieve this. To be precise, we find a continuous function b : Z → R+ and a family of measures µ 12 ROHIT DILIP HOLKAR on Ω which gives a disintegration bm = µ ◦ λ. Before we proceed we prove a small lemma. Lemma 2.7. Let (X, α) and (Y, β) be correspondences from (G1, χ1) to (G2, χ2) and from (G2, χ2) to (G3, χ3), respectively, with ∆1 and ∆2 as their adjoining functions. Then for each u ∈ G(0) 3 there is a function bu on Z such that the measure bumu on on Z = (Z ⋊ G2)(0) is an invariant measure with respect to (Z ⋊ G2, χ). Furthermore, bu satisfies the relation b(xγ, γ−1y)b(x, y)−1 = ∆2(γ−1, y). We work with a single µu at a time, so we prefer to drop the suffix u of bu and (0) G2 → R+ as simply write b. Using Lemma 2.7 we define the function ∆ : Z ×G2 ∆((x, y), γ) = ∆2(γ−1, y). Proof. The proof follows the steps below: i) Firstly, we show that for each u ∈ G(0) 3, mu is strongly quasi-invariant with respect to (Z ⋊ G2, χ) and ∆((x, y), γ) := ∆2(γ−1, y) is the cocycle which implements the quasi-invariance. ii) Since Z ⋊ G2 is proper, we appeal to Proposition 2.3 to get a function b : Z = (Z ⋊ G2)(0) → R+ having the desired properties. (i): We draw Figure 3 which is similar to Figure 1. Let f ∈ Cc(Z ⋊ G2), then Z ⋊ G2 χ−1 sZ⋊G2 Z χ rZ⋊G2 Z λ π λπ Ω Figure 3 ZZ⋊G2 f d(mu ◦ χ) = ZZ ZG2 f ((x, y), γ) dχsX (x) 2 (γ) dmu(x, y) = ZY ZX ZG2 f ((x, y), γ) dχsX (x) 2 (γ) dαrY (y)(x) dβu(y). Change variable ((x, y), γ) 7→ ((xγ, γ−1y), γ−1). Then use the fact that the family measures α is G2-invariant and each measures in β is G2-quasi-invariant to see that the previous term equals ZY ZX ZG2 The function f ((xγ, γ−1y), γ−1) ∆2(γ, γ−1y) dχsX (x) 2 (γ) dαrY (y)(x) dβu(y). ∆ : Z ⋊ G2 → R+, ∆((x, y), γ) = ∆2(γ−1, y), is clearly continuous. Furthermore, ∆ is an R+ pair ((x, y), γ), ((xγ, γ−1y), η) ∈ Z ⋊ G2 a small routine computation shows that ∗ -valued 1-cocycle; for a composable ∆((x, y), γ)∆((xγ, γ−1y), η) = ∆((x, y), γη). Thus ∆ implements the (Z ⋊ G2, χ)-quasi-invariance of the measure mu. (ii): Since Z ⋊ G2 is a proper groupoid, we apply Proposition 2.3 which gives a function b : Z → R+ ∗ such that bmu is an invariant measure on Z (with respect to (Z ⋊ G2, χ)). The function b also satisfies the relation b ◦ sZ⋊G2/b ◦ rZ⋊G2 = ∆, that is, b(xγ, γ−1y)b(x, y)−1 = ∆((x, y), γ) for all ((x, y), γ) ∈ Z ⋊ G2. (cid:3) Remark 2.8. For the cocycle ∆ : Z ⋊ G2 → R+ observe: ∗ , ∆((x, y), γ)) = ∆2(γ−1, y), we COMPOSITION OF TOPOLOGICAL CORRESPONDENCES 13 i) since ∆ does not depend on x, ∆ is G1-invariant; ii) ∆2 is G3-invariant ([5, Remark 2.5]). Hence ∆(((x, y), γ)γ3) = ∆2(γ−1, yγ3) = ∆2(γ−1, y) = ∆((x, y), γ) for all ((x, y), γ) ∈ Z ⋊ G2 and appropriate γ3 ∈ G3. Thus ∆ depends only on γ and [y] ∈ Y /G3. The function b appearing in Lemma 2.7 can be computed explicitly. Let p = {pz}z∈Z be a family of probability measures on Z ⋊ G2 as in Lemma 1.7. Then Propositions 1.10 and 2.3 give (2.9) b(x, y) = (exp ◦ b)(x, y) = exp(cid:18)ZZ⋊G2 log ◦∆((x, y), γ) dp(x,y)((x, y), γ)(cid:19) . This implies that b is a continuous positive function on Z. Remark 2.10. i) The G1-invariance of ∆ from Remark 2.8 along with Equa- tion 2.9 imply that b is G1-invariant. ii) The G3-invariance of ∆ (Remark 2.8 and Equation 2.9) implies that b is G3-invariant. Indeed, for composable ((x, y), γ3) ∈ Z × G3 b(x, yγ3) = b((x, y)γ3) = exp(cid:18)Z log ◦∆((x, yγ3), γ) dp(x,yγ3)((x, yγ3), γ)(cid:19) = exp(cid:18)Z log ◦∆((x, yγ3), γ) F ′((x, y), γ) dχrY (yγ3) (γ)(cid:19) where F ′ is a function as in Equation 1.8 for groupoid Z ⋊ G2 used to get the family of probability measures p. The G3 invariance of ∆ and the fact that rY (yγ3) = rY (y) give that the previous term equals 2 exp(cid:18)Z log ◦∆((x, y), γ) F ′((x, y), γ) dχrY (y) 2 (γ)(cid:19) = b(x, y). The last equality is obtained from a computation similar to the one we started with, but in reverse order. Remark 2.11. Once we have bmu ◦ χ = bmu ◦ χ−1, (ii) of Proposition 2.2 gives a measure µu on Ω with bmu = µu ◦ λ. And, as we shall see, {µu}u∈G3 is the required family of measures. For f ∈ Cc(Ω) (2.12) ZΩ f dµu = ZZ (f ◦ π) · e · b dmu = ZY ZX f ◦ π(x, y)e(x, y) b(x, y) dαrY (y)(x) dβu(y) where π : Z → Ω is the quotient map, and e is the function on Z with R e ◦ sZ⋊G2 dχz = 1 for all z ∈ Z. In the discussion that follows, the letter e will always stand for such a function. Due to (iii) of Proposition 2.2 the measure µu is independent of the choice of the function e. Recall that Ω is a G1-G3-bispace (Observation 2.4) and the action of G3 is proper (Lemma 2.5). Proposition 2.13. The family of measures {µu}u∈G(0) tinuous family of measures on Ω along the momentum map sΩ. 3 is a G3-invariant con- Proof. We check the invariance first and then check the continuity. Let f ∈ Cc(Ω) and γ ∈ G3, then ZΩ f ([x, yγ]) dµrG3 (γ)[x, y] = ZY ZX f ([x, yγ])e(x, yγ)b(x, y) dαrY (y)(x) dβrG3 (γ ′)(y). 14 ROHIT DILIP HOLKAR Change yγ → y, then use the G3-invariance of the family β and that of the function b to see that the last term in the above computation equals ZY ZX = ZZ f ([x, y])e(x, y)b(x, y) dαrY (y)(x) dβsG3 (γ)(y) f · e · b dmsG3 (γ) = ZΩ f [x, y] dµsG3 (γ)[x, y]. (0) is G3-invariant. Thus {µu}u∈G3 Now we check that µ is a continuous family of measures. Let M, µ and Λ denote the integration maps which the families of measures m, µ and λ induce between the corresponding spaces of continuous compactly supported functions. Remark 2.11 says that M : Cc(Z) → Cc(G(0) 3), that is, Figure 4 commutes: Lemma 2.6 shows that M is continuous, [5, Example 3) is the composite of Cc(Z) Λ−→ Cc(Ω) µ −→ Cc(G(0) Cc(Ω) Λ µ Cc(Z) M Cc(G(0) 3). Figure 4 1.8] shows that Λ is continuous and surjective. Hence µ is continuous. (cid:3) The family of measures µ on Ω is the required family of measures for the com- posite correspondence. We still need to show that each µu is G1-quasi-invariant. The following computation shows this quasi-invariance and also yields the adjoining function. Let f ∈ Cc(G1 ×G1 (0) Ω) and u ∈ G(0) 3, then f (η−1, [x, y]) dχrΩ([x,y]) 1 (η) dµu[x, y] ZΩZG1 = ZY ZX ZG1 f (η−1, [x, y]) e(x, y) b(x, y) dχrX (x) 1 (η) dαrY (y)(x) dβu(y). Now we change variable (η−1, [x, y]) 7→ (η, [η−1x, y]). Then the (G1, χ1)-quasi- invariance of α changes dχrX (x) 1 (η) dαrY (y)(x) 7→ ∆1(η, η−1x) dχrX (x) 1 (η) dαrY (y)(x). We incorporate this change and continue the computation further: R. H. S. = ZY ZX ZG1 f (η, [η−1x, y]) e(η−1x, y) b(η−1x, y) ∆1(η, η−1x) dχrX (x) 1 (η) dαrY (y)(x) dβu(y) = ZY ZX ZG1 f (η, [η−1x, y]) b(η−1x, y) b(x, y) ∆1(η, η−1x) e(η−1x, y)b(x, y) dχrX (x) 1 (η) dαrY (y)(x) dβu(y). We transfer the integration on Ω where the previous term equals ZΩZG1 f (η, [η−1x, y]) b(η−1x, y) b(x, y) ∆1(η, η−1x) dχrΩ([x,y]) 1 (η) dµu[x, y]. Define ∆1,2 : G1 ⋉ Ω → R+ ∗ by (2.14) ∆1,2(η, [x, y]) = b(ηx, y)−1∆1(η, x)b(x, y), COMPOSITION OF TOPOLOGICAL CORRESPONDENCES 15 then the above computation gives ZΩZG1 f (η−1, [x, y]) dχ1(η) dµu[x, y] = ZΩZG1 f (η, [η−1x, y]) ∆1,2(η, η−1[x, y]) dχ1(η) dµu[x, y] for u ∈ G(0) 3. To announce that µu is G1-quasi-invariant and ∆1,2 is the adjoining function, we must check that the function ∆1,2 is well-defined which the next lemma does. Lemma 2.15. The function ∆1,2 defined in Equation (2.14) is a well-defined R+ ∗ -valued continuous 1-cocycle on the groupoid G1 ⋉ Ω. Proof. Let (xγ, γ−1y) ∈ [x, y], then ∆1,2(η−1, [xγ, γ−1y]) = b(η−1xγ, γ−1y)−1∆1(η−1, xγ)b(xγ, γ−1y). We multiply and divide the last term by b(η−1x, y)−1b(x, y), and use the G2-invariance of ∆1, then re-write the term as b(η−1x, y)−1∆1(η−1, x)b(x, y) (cid:18) b(η−1x, y) b(η−1xγ, γ−1y) b(xγ, γ−1y) b(x, y) (cid:19) Now use the last claim in Lemma 2.7 which relates b and ∆2, use the definition of ∆1,2, and compute the above term further: ∆1,2(η−1, [x, y]) (cid:0)∆2(γ−1, y)∆2(γ, γ−1y)(cid:1) = ∆1,2(η−1, [x, y]). To get the equality above, observe that (γ−1, y)−1 = (γ, γ−1y) and use the fact that ∆2 is a homomorphism. Due to the continuity of b and ∆1, the cocycle ∆1,2 is continuous. Using a computation as above, it can be checked that ∆1,2 is a groupoid homomorphism. (cid:3) Proposition 2.16. The family of measures {µu}u∈G(0) adjoining function for the quasi-invariance is given by Equation (2.14). 3 is G1-quasi-invariant. The Proof. Clear from the discussion above. Definition 2.17 (Composition). Let (cid:3) (X, α) : (G1, χ1) → (G2, χ2) and (Y, β) : (G2, χ2) → (G3, χ3) be topological correspondences with ∆1 and ∆2 as the adjoining functions, respect- ively. A composite of these correspondence (Ω, µ) : (G1, χ1) → (G3, χ3) is defined by: i) the space Ω := (X ×G2 ii) a family of measures µ = {µu}u∈G(0) (0) Y )/G2; 3 such that (a) let ∆ ∈ C1 ∆2(γ−1, y), G3((X ×G2 (0) Y ) ⋊ G2, R+ ∗ ) be the 1-cocycle ∆((x, y), γ) = (b) let b ∈ C0 (c) then µ disintegrate the family of measures {b(α × βu)}u∈G3 ∗ ) be a cochain with d0(b) = ∆; (0) Y ) ⋊ G2, R+ G3((X ×G2 (0) on X ×G2 (0) (0) Y → Ω using λ, that is, b(α×βu) = Y along the quotient map π : X ×G2 µu ◦ λ for each u ∈ G3 (0). G3((X ×G2 In Definition 2.17, Cn consisting of G3-invariant R+ Y ) ⋊ G2. For the composite (Ω, µ) as above, the adjoining function ∆1,2 is given by Equation (2.14). ∗ -valued continuous cochains on groupoid (X ×G2 ∗ ) denotes the n-th cochain group (0) Y ) ⋊ G2, R+ (0) 16 ROHIT DILIP HOLKAR Theorem 2.18. Let (X, α) : (G1, χ1) → (G2, χ2) and (Y, β) : (G2, χ2) → (G3, χ3) be topological correspondences of locally compact groupoids with Haar systems. In addition, assume that X and Y are Hausdorff and second countable. Let (Ω, µ) : (G1, χ1) → (G3, χ3) be a composite of the correspondences. Then H(Ω) and H(X) ⊗C∗(G2,χ2)H(Y ) are isomorphic C∗-correspondences from C∗(G1, χ1) to C∗(G3, χ3). Proof. The symbols Z, Ω and the families of measures m, λ and µ continue to have the same meaning as in the earlier discussion. Let b be a fixed zeroth cocycle on Z ⋊ G2 with ∆ = d0(b) as in Definition 2.17. In the calculations below, the subscripts to h,i indicate the Hilbert module on which the inner product is defined. We write H(X) ⊗H(Y ) instead of H(X) ⊗C∗(G2,χ2)H(Y ) in this proof to reduce the complexity in writing. Recall the process of composing two C∗-correspondences in Section 1.1 on page 5. We know that Cc(X) ⊆ H(X) and Cc(Y ) ⊆ H(Y ) are, respectively, pre-Hilbert C∗(G2, χ2) andC∗(G3, χ3)-modules. Due to this density, the image of Cc(X) ⊗C Cc(Y ) → H(X) ⊗ H(Y ) under the obvious map is dense. Recall from the same discussion that the obvious map sends f ⊗ g ∈ Cc(X) ⊗C Cc(Y ) to its equivalence class in (Cc(X) ⊗C Cc(Y ))/(Cc(X) ⊗C Cc(Y )) ∩ N where N ⊆ H(X) ⊗ H(Y ) is the subspace of vectors of zero norm. The norm on Cc(X) ⊗C Cc(Y ) induced by the inner product (2.19) hf ⊗ g , f ′ ⊗ g′iH(X) ⊗H(Y ) := Dg , hf , f ′iH(X) g′EH(Y ) for f ⊗ g, f ′ ⊗ g′ ∈∈ Cc(X) ⊗C Cc(Y ). Thus when equipped with the inner product in Equation 2.19, the pre-Hilbert C∗(G3, χ3)-module Cc(X) ⊗C Cc(Y ) completes to the Hilbert C∗(G3, χ3)-module H(X) ⊗ H(Y ). On the other hand, Cc(Ω) ⊆ H(Ω) is a pre-Hilbert C∗(G3, χ3)-module. We define an inner product preserving Cc(G3)-module map Λ′ : Cc(X)⊗Cc(Y ) → Cc(Ω) which has a dense image. Then Λ′ is an inner product preserving map of pre-Hilbert C∗(G3, χ3)-modules and image of Λ′ is dense. Thus Λ′ extends to an isomorphism of Hilbert C∗(G3, χ3)-modules H(X) ⊗H(Y ) → H(Ω). After showing that Λ′ is an isometry of Cc(G3)-modules and has a dense image, we show that Λ′ intertwines the representations of C∗(G1, χ1) on H(X) ⊗H(Y ) and H(Ω) which completes the proof that Λ′ induces an isomorphism H(X) ⊗H(Y ) → H(Ω) of C∗-correspondences. Again, due to the density arguments as above, it is enough to show that Λ′ intertwines the representations of Cc(G1) on Cc(X) ⊗C Cc(Y ) and Cc(Ω); and this is what we show. Thus the proof is divided into two parts, the first part proves the isomorphism of Hilbert modules and the other the isomorphism of the representations. The strategy of the proof is explained and we start the proof by defining Λ′. Map f ⊗ g ∈ Cc(X) ⊗C Cc(Y ) to (f ⊗ g)Z ∈ Cc(Z) where (f ⊗ g)Z (x, y) = f (x)g(y) for (x, y) ∈ Z. Then the Stone-Weierstrass theorem gives that the set {(f ⊗ g)Z: f ⊗ g ∈ Cc(X) ⊗C Cc(Y )} ⊆ Cc(Z) is dense. Define Λ′ : Cc(X) ⊗C Cc(Y ) → Cc(Ω) by Λ′(f ⊗ g)[x, y] = Λ((f ⊗ g)Z b−1/2)[x, y] = ZG2 (f ⊗ g)Z(xγ, γ−1y) b−1/2(xγ, γ−1y) dχsX (x) 2 (γ) where f ⊗ g ∈ Cc(X) ⊗C Cc(Y ). Since b is a positive function, the multiplication by b−1/2 is an isomorphism from Cc(Z) to itself. As λ is a continuous family of measure with full support, Λ : Cc(Z) → Cc(Ω) is surjection. Thus the composite COMPOSITION OF TOPOLOGICAL CORRESPONDENCES 17 Λ′ : Cc(X) ⊗C Cc(Y ) a continuous and has dense image. f ⊗g7→(f ⊗g)Z −−−−−−−−−→ Cc(Z) multiplication by b−1/2 −−−−−−−−−−−−−−−→ Cc(Z) Λ−→ Cc(Ω) is Let z ∈ C, f, f ′ ∈ Cc(X) and g, g′ ∈ Cc(Y ). Then it is straightforward com- putation to check that Λ′(zf ⊗ g + f ′ ⊗ g′) = zΛ′(f ⊗ g) + Λ′(f ′ ⊗ g′). Fur- thermore, if ψ ∈ Cc(G3), then a computation using Fubini's theorem shows that Λ′((f ⊗ g)ψ) = Λ′(f ⊗ g)ψ. Thus Λ′ is a homomorphism of Cc(G3)-modules. The isomorphism of the Hilbert modules: In this part, we show that Λ′ pre- serves C∗(G3, χ3)-valued inner products. Let f ⊗ g ∈ Cc(X) ⊗C Cc(Y ) and γ ∈ G3, then (γ) g(y) (hf , f iH(X) g)(yγ) dβrG3 (γ)(y) hf ⊗ g , f ⊗ giH(X) ⊗H(Y ) (γ) := Dg , hf , f iH(X) gEH(Y ) = ZY = ZY ZG2 = ZY ZG2 g(y) hf, f iH(X)(γ) g(γ−1yγ)∆1/2 f (x)f (xγ) dαrG2 (γ)(x)(cid:19) g(y)(cid:18)ZX (γ, γ−1yγ) dχrY (y) g(γ−1yγ) ∆1/2 2 (γ) dβrG3 (γ)(y) 2 2 (γ, γ−1yγ) dχrY (y) 2 (γ)dβrG3 (γ)(y) We rearrange the functions, note that rG2(γ) = rY (y) and write the last term as (2.20) ZY ZG2 ZX f (x) g(y) f (xγ)g(γ−1yγ) ∆1/2 2 (γ, γ−1yγ) dαrY (y)(x) dχrY (y) 2 (γ) dβrG3 (γ)(y). Now we calculate the norm of Λ′(f ⊗ g) ∈ Cc(Ω): hΛ′(f ⊗ g) , Λ′(f ⊗ g)iH(Ω) (γ) := ZΩ Λ′(f ⊗ g)[x, y] Λ′(f ⊗ g)[x, yγ] dµrG3 (γ)[x, y]. We plug the value of the first Λ′(f ⊗ g) and continue computing further: f (xγ∗)g(γ−1 ∗ y)b−1/2(xγ∗, γ−1 ∗ y) dχrY (y) 2 f (xγ∗)g(γ−1 ∗ y)b−1/2(xγ∗, γ−1 ∗ y)Λ′(f ⊗ g)[x, yγ] dχrY (y) (γ∗)(cid:19) Λ′(f ⊗ g)[x, yγ] dµrG3 (γ)[x, y] (γ∗) dµrG3 (γ)[x, y] 2 ZΩ(cid:18)ZG2 = ZΩZG2 = ZY ZX f (x) g(y)b−1/2(x, y)Λ′(f ⊗ g)[x, yγ]b(x, y) dαrY (y)(x) dβrG3 (γ)(y). The last equality above is due to Remark 2.11, which says that dχrY (y) 2 (γ∗) dµrG3 (γ)[x, y] = b(x, y) dαrY (y)(x) dβrG3 (γ)(y). 18 ROHIT DILIP HOLKAR We process the function b in the previous term, plug in the value of Λ′(f ⊗ g) and compute further, L. H. S. = ZY ZX f (x) g(y)(cid:18)ZG2 f (xγ)g(γ−1yγ)b−1/2(xγ, γ−1yγ) dχrY (y) 2 (γ)(cid:19) b1/2(x, y) dαrY (y)(x) dβrG3 (γ)(y) = ZY ZX ZG2 f (x) g(y)f (xγ)g(γ−1yγ) (cid:18) b(x, y) b(xγ, γ−1yγ)(cid:19)1/2 dχrY (y) 2 (γ) dαrY (y)(x) dβrG3 (γ)(y). First we use the G3-invariance of b (Remark 2.10) to write b(x, y) = b(x, yγ). Then we use Lemma 2.3 to relate the factors of b and get a factor of ∆ which can be written in terms of ∆2 using Remark 2.8. At the end of these computations, the last term of the previous becomes ZY ZX ZG2 (cid:16)f (x) g(y)f (xγ)g(γ−1yγ)(cid:17) ∆2 1/2(γ, γ−1yγ) dχrY (y) 2 (γ) dαrY (y)(x) dβrG3 (γ)(y). Finally, we apply Fubini's Theorem to χrY (y) 2 and αrY (y) to get (2.21) hΛ′(f ⊗ g) , Λ′(f × g)iH(Ω) (γ) = ZY ZG2 ZX (cid:16)f (x) g(y)f (xγ)g(γ−1yγ)(cid:17) 1/2(γ, γ−1yγ) dαrY (y)(x) dχrY (y) ∆2 2 (γ) dβrG3 (γ)(y). Comparing the values of both inner products, that is, Equation 2.20 and 2.21, we conclude that hf ⊗ g , f ⊗ giH(X) ⊗H(Y ) = hΛ′(f ⊗ g) , Λ′(f ⊗ g)iH(Ω) . The isomorphism of representations: Denote the actions of C∗(G1, χ1) on H(X) ⊗H(Y ) and H(Ω) by ρ1 and ρ2, respectively, that is, ρ1 : C∗(G1, χ1) → B(H(X) ⊗C∗(G2)H(Y )) and ρ2 : C∗(G1, χ1) → B(H(Ω)) are the nondegenerate *-representations that give the C∗-correspondences from C∗(G1, χ1) to C∗(G3, χ3). Now we show that Λ′ intertwines ρ1 and ρ2. Let ∆1,2 be the adjoining function of (Ω, µ) which is given by Equation 2.14. Let φ ∈ Cc(G1) and f ⊗ g ∈ Cc(X) ⊗C Cc(Y ), then (ρ2(φ)Λ′)(f ⊗ g)[x, y] = (φ ∗ Λ′(f ⊗ g))[x, y] φ(η)Λ′(f ⊗ g))[η−1x, y] ∆1/2 1,2 (η, [η−1x, y]) dχrX (x) 1 (η) = ZG1 = ZG1 ZG2 ∆1/2 (2.22) φ(η)f (η−1xγ)g(γ−1y) b−1/2(η−1xγ, γ−1y) 1,2 (η, [η−1x, y]) dχsX (x) 2 (γ) dχrX (x) 1 (η). Lemma 2.15 and Equation (2.14) allows us to write ∆1,2(η, [η−1x, y]) = ∆1,2(η, [η−1xγ, γ−1y]) = ∆1(η, η−1xγ) b(η−1xγ, γ−1y) b(xγ, γ−1y) . COMPOSITION OF TOPOLOGICAL CORRESPONDENCES 19 Substitute this value of ∆1,2(η, [η−1x, y]) in Equation 2.22. Then apply Fubini's theorem and continuing computing further: ZG2 (cid:18)ZG1 φ(η)f (η−1xγ) ∆1/2 1 (η, η−1xγ) dχrX (x) 1 (η)(cid:19) g(γ−1y) b−1/2(xγ, γ−1y) dχsX (x) 2 (γ) = ZG2 (φ ∗ f )(xγ)g(γ−1y) b−1/2(xγ, γ−1y) dχsX (x) 2 (γ) = Λ′((φ ∗ f ) ⊗ g)[x, y] = Λ′(ρ1(φ)(f ⊗ g))[x, y]. (cid:3) 3. Examples Example 3.1. Let X, Y and Z be locally compact Hausdorff spaces and let f : X → Y and g : Y → Z be a continuous functions. Then [5, Example 3.1] shows that (X, δX ) is a topological correspondence from Y to X and (Y, δY ) is the one from Y to Z. Here δX = {δx}x∈X is the family of measures consisting of point masses along the identity map X → X. Similar is the meaning of δY . The constant function 1 is the adjoining function for both correspondences. The space involved the composite of (Y, δY ) and (X, δX ) is (Y ×IdY ,Y,f X) ≈ X, and the homeomorphism (Y ×IdY ,Y,f X) → X is implemented by the function (f (x), x) 7→ x. The inverse of this function is x 7→ (f (x), x). The left momentum map Y ×IdY ,Y,f X → Z is (f (x), x) 7→ g(f (x)) which we identify with g ◦ f : X → Z. Thus the composite of the topological correspondences related to continuous maps is same as the topological correspondence related to the composite of the maps. Reader may check the C∗-algebraic counterpart of this example agree with Theorem 2.18. Example 3.2. Let V, W, X, Y and Z be locally compact Hausdorff spaces and let f : X → Z, g : X → Y, k : V → Y and l : V → W be continuous maps. Let λ1 and λ2 be continuous families of measures along g and l, respectively (See Figure 5 on page 19). Then (X, λ1) is a topological correspondence from Z to Y and (V, λ2) is one from Y to W ([5, Example 3.3]). The composite correspondence is (X ×g,Y,k X f g k Z λ1 Y V l λ2 W Figure 5 V, λ1 ◦ λ2) where (λ1 ◦ λ2)w is defined by ZX×g,Y,kV f d(λ1 ◦ λ2)w = ZV ZX f (x, v) dλ1k(v)(x) dλ2w(v) for w ∈ W and f ∈ Cc(X ×g,Y,k V ). Note that in this example λ1 ◦ λ2 is the family of measures m in Lemma 2.6 and, since there are only the trivial actions, it is same as the family of measures µ in Proposition 2.13. Example 3.3. Let G, H and K be locally compact groups, ψ : K → H and φ : H → G continuous homomorphisms. Let α, β and λ be the Haar measures on G, H and K, respectively. Then (G, α−1) is a correspondence from (H, β) to (G, α), and (H, β−1) is one from (K, λ) to (H, β) ([5, Example 3.4]). Let δG, δH and δK denote 20 ROHIT DILIP HOLKAR the modular functions of G, H and K, respectively. Then δG◦φ δH adjoining function for these correspondences, respectively. and δH ◦φ δK are the The K-G-bispace in composite of these correspondences is (H × G)/H ≈ G. The map a : γ 7→ [eH , γ], G → (H × G)/H, gives the homeomorphism where eH is the unit in H. The inverse of this map a−1 is a−1 : [η, γ] 7→ φ(η)γ, (H × G)/H → G. We figure out the action of K on this K-G-bispace: if κ ∈ K then κγ = a−1(κ[eH , γ]) := a−1([ψ(κ), γ]) = φ(ψ(κ))γ. Thus K acts on G via the homo- morphism φ ◦ ψ : K → G. Similarly, the right action of G on the composite space (H × G)/H ≈ G is identified with the right multiplication action of G on itself. A computation as in [5, Example 3.4] gives that δG◦φ◦ψ is the adjoining function for the composite correspondence. δK This shows that the composite of (H, β−1) and (G, α−1) is same as the corres- pondence associated with the homomorphism φ ◦ ψ : K → G. Example 3.4. Let (G, α), (H, β) and (K, λ) be locally compact groups with Haar measures, and let φ : H → G and ψ : K → G be continuous homomorphisms. As- sume the ψ is a proper map. Then φ gives a correspondences (G, α−1) from (H, β) to (G, α) as in [5, Example 3.4] and ψ gives a correspondence (G, α−1) from (G, α) to (K, λ) as in [5, Example 3.5]. The adjoining function of the topological corres- pondence associated with ψ is the constant function 1. The composite of these correspondences is a correspondence (H, β) → (K, λ). One the similar lines of Example 3.3, one may show that the space involved in the composite is homeomorphic to G, the actions of H and K are identified with the left and right multiplication via φ and ψ, the K-invariant family of measures on G is α−1. From [5, Example 3.4] we know that the α−1 is (H, β)-quasi-invariant and the function δG◦φ is the δH adjoining function for the composite. is the cocycle involved in the quasi-invariance. Hence δG◦φ δH An interesting situation is when H, K ⊆ G are closed subgroups, and φ and ψ are the inclusion maps. Then the composite correspondence from (H, β) to (K, λ) is (G, α−1) where G is made into an H-K bispace using the left and right multiplication actions, respectively. The adjoining function in this case is δG δH . Example 3.5. Example 3.7 in [5] shows that the correspondences defined by Macho Stadler and O'uchi in [7] are topological correspondences. The same example shows that the topological correspondences defined by Tu in [8, Proposition 7.5] are also topological correspondences provided that the spaces of the units of the groupoids are Hausdorff. The composition of correspondences of Macho Stadler and O'uchi defined by Tu ([8]) is same as the composition we define. Recall from [7, Definition 1.2] or [8, Definition 7.3] that a correspondence (G1, χ1) → (G2, χ2) is a G1-G2-bispace, and the actions and the quotient G1\X satisfy certain conditions. Since the correspondences of Macho Stadler and O'uchi or Tu do not involve explicit families of measures, the construction of the composite in this is purely topological. If Y a correspondence in there sense from (G2, χ2) to (G3, χ3), then Tu shows [8] the space Ω in Definition 2.17 the composite. Example 3.6. Example 3.10 in [5] shows that the generalized morphisms defined by Buneci and Stachura are topological correspondences in our sense. Though it is not as straightforward as in Example 3.5 above, but it may be checked that the composition of the generalized morphisms of Buneci and Stachura defined in [5, Section 2.2] match our definition of composition. Example 3.7. Let G be a locally compact group, let H and K be closed subgroups of G, and let α, β and λ be the Haar measures on G, H and K, respectively. Let δG and δH be the modular functions of G and H, respectively. Then (cid:0)G, α−1(cid:1) is a as the adjoining function, see in Example 3.4. correspondence from H to K with δG δH COMPOSITION OF TOPOLOGICAL CORRESPONDENCES 21 Let X be a left K-space carrying a strongly (K, λ)-quasi-invariant measure κ, that is, κ is a (K, λ)-quasi-invariant measure on X and the Radon-Nikodym deriv- ative for the quasi-invariance, say ∆ : K ⋉ X → R+ ∗ , is a continuous function. Then (X, κ) is a correspondence from K to Pt with ∆ as adjoining function. Here Pt stands for the trivial group(oid) which consists of the unit only. We discuss the composite of these two correspondences. The space in the com- posite is (G × X)/K, which we denote by Z. In this example, writing the measure ν on Z concretely is not always possible. However, when (X, κ) = (K, λ), we get Z ≈ G and ν = α−1. The correspondence (X, κ) gives a representation of K on L2(X, κ) and the composite correspondence is the representation of H induced by this representation of K. Acknowledgement: I am grateful to Jean Renault and Ralf Meyer for their guid- ance and many fruitful discussions. References [1] C. Anantharaman-Delaroche and J. Renault, Amenable groupoids, Monographies de L'Enseignement Mathématique [Monographs of L'Enseignement Mathématique], vol. 36, L'Enseignement Mathématique, Geneva, 2000. With a foreword by Georges Skandalis and Appendix B by E. Germain. MR1799683 (2001m:22005) [2] Nicolas Bourbaki, Integration. II. Chapters 7 -- 9, Elements of Mathematics (Berlin), Springer- Verlag, Berlin, 2004. Translated from the 1963 and 1969 French originals by Sterling K. Ber- berian. MR2098271 (2005f:28001) [3] Mădălina Roxana Buneci and Piotr Stachura, Morphisms of locally compact groupoids endowed with Haar systems (2005), eprint, available at arxiv:0511613. [4] Gerald B. Folland, A course in abstract harmonic analysis, Studies in Advanced Mathematics, CRC Press, Boca Raton, FL, 1995. MR1397028 (98c:43001) [5] Rohit Dilip Holkar, Topological construction of C ∗-correspondences for groupoid C ∗-algebras (2015), available at arxiv:1510.07534. [6] E. C. Lance, Hilbert C ∗-modules, London Mathematical Society Lecture Note Series, vol. 210, Cambridge University Press, Cambridge, 1995. A toolkit for operator algebraists. MR1325694 (96k:46100) [7] Marta Macho Stadler J. Operator algebras, www.mathjournals.org/jot/1999-042-001/1999-042-001-005.pdf. no. 42 and Moto O'uchi, Correspondence Theory (1999), of 103 -- 119, 1, groupoid C ∗- at available [8] Jean-Louis Tu, Non-Hausdorff groupoids, proper actions and K-theory, Doc. Math. 9 (2004), 565 -- 597 (electronic). MR2117427 (2005h:22004) E-mail address: [email protected] Department of Mathematics, Federal University of Santa Catarina, 88. 040-900, Flori- anopólis, SC, Brazil
1701.01610
1
1701
2017-01-06T12:13:06
Vietoris topology on hyperspaces associated to a noncommutative compact space
[ "math.OA" ]
We study some topological spaces that can be considered as hyperspaces associated to noncommutative spaces. More precisely, for a NC compact space associated to a unital C*-algebra, we consider the set of closed projections of the second dual of the C*-algebra as the hyperspace of closed subsets of the NC space. We endow this hyperspace with an analog of Vietoris topology. In the case that the NC space has a quantum metric space structure in the sense of Rieffel we study the analogs of Hausdorff and infimum distances on the hyperspace. We also formulate some interesting problems about distances between sub-circles of a quantum torus.
math.OA
math
VIETORIS TOPOLOGY ON HYPERSPACES ASSOCIATED TO A NONCOMMUTATIVE COMPACT SPACE MAYSAM MAYSAMI SADR 7 1 0 2 n a J 6 ] Abstract. We study some topological spaces that can be considered as hyperspaces asso- ciated to noncommutative spaces. More precisely, for a NC compact space associated to a unital C*-algebra, we consider the set of closed projections of the second dual of the C*- algebra as the hyperspace of closed subsets of the NC space. We endow this hyperspace with an analog of Vietoris topology. In the case that the NC space has a quantum metric space structure in the sense of Rieffel we study the analogs of Hausdorff and infimum distances on the hyperspace. We also formulate some interesting problems about distances between sub-circles of a quantum torus. . A O h t a m [ 1 v 0 1 6 1 0 . 1 0 7 1 : v i X r a 1. Introduction This note is a contribution to Noncommutative Topology. We introduce and study some topological spaces that can be considered as the hyperspaces associated to noncommutative spaces. More precisely, let qA denote the (imaginary) NC compact Hausdorff space asso- ciated to a unital C*-algebra A. We consider the set of nonzero closed projections of the second dual A∗∗ as the hyperspace SclqA of nonempty closed subsets of qA. In the case that A is commutative these closed projections are canonically identified with closed subsets of the Gelfand space of A. (To the best of our knowledge the study of closed projections as closed subsets of NC spaces goes back to Akemann [1, 2, 3]. Closed projections have been considered also in some recent papers, see [10, 11] and references therein.) There is a canonical bijection between closed projections in A∗∗ and weak*-closed faces of the state space SA of A (see Section 4). Thus we can identify the hyperspace SclcSA of such subsets of SA with SclqA. We have a canonical Vietoris topology on SclcSA induced from the weak*-topology of SA. In the case that A is commutative it is proved in Section 2 that this Vietoris topology coincides with the vietoris topology on the hyperspace of closed subsets of the Gelfand space of A. Suppose that qA has a quantum metric space structure in the sense of Rieffel [23, 24, 25]. This induces Hausdorff and infimum distances on SclqA. Again if A is commutative it is proved that these distances coincide with the usual Hausdorff and infimum distances (see the last paragraph of Section 4). The notion of quantum (or NC) metric space have been considered by many authors, see [25, 15, 14, 12, 16, 17, 29, 27] and references therein. The main subjects studied in most of the mentioned papers are variations of the quantum Gromov-Hausdorff distance and quantum metric spaces defined by Rieffel [25]. The notions introduced by Rieffel [25] are based on order unit spaces. Since our attention here is to NC Topology we are more interested in order unit spaces arising from C*-algebras. 2010 Mathematics Subject Classification. 46L05; 46L85; 54B20. Key words and phrases. C*-algebra, state space, closed projection, hyperspace, Vietoris topology, Haus- dorff distance, infimum distance. 1 2 M. M. SADR The plan of the paper is as follows. In Section 2 we consider some properties of hyperspaces associated to ordinary topological spaces. Also we consider Hausdorff and infimum distances. In Section 3 we review the notion of quantum metric space. In Section 4 we introduce our main object SclqA, the hyperspace of closed subsets of a compact NC space. In Section 5 we study the Vietoris topology on SclqA. In Section 6 using the infimum distance we define an analog of Lipschitz seminorm for quantum metric spaces. At last in Section 7 we consider some questions and problems on finite NC spaces and quantum tori. 2. Hyperspace of closed subsets of an ordinary topological space Let X be a compact Hausdorff space. We denote by SclX the set of all nonempty closed subsets of X. For every open U ⊆ X, let U − := {K ∈ SclX : K ∩ U 6= ∅} and U + := {K ∈ SclX : K ⊆ U}. The smallest topology on SclX containing all U ±'s is called Vietoris topology. The space SclX together with the Vietoris topology is called the hyperspace of closed sets in X. It is easy to see ([13, Exercise 3.12]) that the hyperspace is compact and Hausdorff. Also the subspace topology of X, where X is considered as a subspace of SclX via the canonical embedding x 7→ {x}, coincides with the original topology of X. Let CX denote the C*-algebra of complex valued continuous functions on X. We always endow the state space SCX of CX with weak* topology. We also identify SCX with the space of Borel regular probability measures on X. Then the map δ : x 7→ δx is a homeomorphism from X onto the space of pure states of CX where δx denote the point mass measure concentrated at x. For a nonempty closed subset K of X let FK denote the set of those measures µ in SCX with Spt(µ) ⊆ K. Then FK is a weak*-closed face of SCX. Also note that FK is the weak* closed convex hull of {δx : x ∈ K}. Proposition 1. The map F : K 7→ FK is a homeomorphism from the hyperspace SclX onto a closed subspace of the hyperspace SclSCX. Proof. Since both of the hyperspaces are compact Hausdorff spaces and F is injective it is enough to show that F is continuous. Let U, V be arbitrary open subsets of SCX. We must show that F −1(U +) and F −1(V −) are open in SclX. Suppose that K ∈ F −1(U +). Thus FK ⊆ U. Since FK is convex it follows from [26, Theorem 1.10] that there is a convex open subset U0 of SCX with FK ⊆ U0 ⊆ U0 ⊆ U. We have K ∈ (δ−1(U0))+ ⊆ F −1(U +). Thus F −1(U +) is open. Now suppose that K ∈ F −1(V −). Thus there are µ ∈ FK \ V c and open subset W of SCX such that µ ∈ W and W ∩ V C = ∅. It follows that there exist x1, . . . , xn ∈ Sptµ ⊆ K, t1, . . . , tn > 0 with Pn i=1 ti = 1, and an open subset W0 of SCX, such that Pn i=1 tiδxi ∈ W0 ⊆ W . Thus there are open subsets O1, . . . , On of X with xi ∈ Oi, and with the property that if yi ∈ Oi then Pn i ⊆ F −1(V −). Thus F −1(V −) is open. This completes the proof. (cid:3) i=1 tiδyi ∈ W0. We have K ∈ ∩n i=1O− Now suppose that X is metrizable and let d be a compatible metric on X. The Hausdorff distance Hd (associated to d) on SclX is defined by Hd(K, K ′) = inf{r > 0 : K ⊆ Ball(K ′, r), K ′ ⊆ Ball(K, r)} (K, K ′ ∈ SclX), where Ball(K, r) = {y ∈ X : d(x, y) < r, ∃x ∈ K}. It is well known that Hd is a metric and the topology induced by Hd coincides with the Vietoris topology ([13, Theorem 3.1]). Also the mapping x 7→ {x} is an isometric embedding of X into SclX. The Lipschitz seminorm HYPERSPACES ASSOCIATED TO A NONCOMMUTATIVE SPACE 3 Ld for (self-adjoint) elements of CX is defined by (1) Ld(f ) := sup{ f (x) − f (y) d(x, y) : x, y ∈ X, x 6= y} (f ∈ CXsa). This seminorm satisfies the Leibniz inequality: Ld(f g) ≤ Ld(f )kgk∞ + kf k∞Ld(g). The Lipschitz algebra of (X, d) is defined by LipdX := {f ∈ CXsa : Ld(f ) < ∞}. This is a real uniformly-dense subalgebra of CXsa. (For an extensive account on Lipschitz algebras see [28].) The Monge-Kantorovich distance is defined by (2) ρd(µ, ν) := sup{µ(f ) − ν(f ) : Ld(f ) ≤ 1} (µ, ν ∈ SCX) It is well known that the topology of ρd coincides with weak* topology and also the restriction of ρd to the space of pure states of CX is equal to d where the pure state space is canonically identified with X. The metric version of Proposition 1 is as follows. Proposition 2. K 7→ FK is an isometric embedding from (SclX, Hd) into (SclSCX, Hρd ). Proof. Let h denote the Hausdorff distance of FK and FK ′. Suppose that Hd(K, K ′) < r. Then for every x ∈ K there is y ∈ K ′ such that d(x, y) < r. Let t1, . . . , tn ≥ 0 with Pn i=1 ti = 1 and let x1, . . . , xn ∈ K. Then it is easily verified that ρd(Pn i=1 tiδyi) < r where yi ∈ K ′ is such that d(xi, yi) < r. This shows that FK ⊆ Ball(FK ′, r + ǫ) for every ǫ > 0. Similarly, we have FK ′ ⊆ Ball(FK, r + ǫ). Thus h ≤ r, and hence h ≤ Hd(K, K ′). Now suppose that h < s. Let x ∈ K. Then there are z1, . . . , zn ∈ K ′ and t1, . . . , tn ≥ 0 with Pn i=1 tiδzi) < s. Let f be the function on X defined by y 7→ d(x, y). Then Ld(f ) = 1 (if X at least has two points). We have i=1 ti = 1 and ρd(δx, Pn i=1 tiδxi, Pn n X i=1 tid(x, zi) = δx(f ) − ( n X i=1 tiδzi)(f ) < s. Thus d(x, zi0) < s for some i0. This shows that K ⊆ Ball(K ′, s). Similarly we have K ′ ⊆ Ball(K, s). Thus Hd(K, K ′) ≤ s, and hence Hd(K, K ′) ≤ h. The proof is complete. (cid:3) For two subsets K, K ′ of X their infimum distance is defined by Id(K, K ′) := inf{d(x, y) : x ∈ K, y ∈ K ′}. In the case that K or K ′ is empty we let Id(K, K ′) = ∞. Note also that in general Id is not a metric on SclX. Proposition 3. Let K, K ′ ∈ SclX. Then Id(K, K ′) = Iρd(FK, FK ′). Proof. Let I denote the infimum distance of FK and FK ′. For x ∈ K, y ∈ K ′ we have δx ∈ FK, δy ∈ FK ′ and d(x, y) = ρd(δx, δy). Thus I ≤ Id(K, K ′). Let I < r. There are µ := Pn j=1 sjδyj ∈ FK ′ such that ρd(µ, ν) < r. Let the function f on X be defined by x 7→ Id({x}, {y1, . . . , yn}). Then Ld(f ) = 1, and we have i=1 tiδxi ∈ FK and ν := Pm n X i=1 tif (xi) = µ(f ) − ν(f ) < r. Thus f (xi0) < r for some i0, and hence there is j0 such that d(xi0, yj0) < r. This shows that Id(K, K ′) < r. Since r > I is arbitrary we conclude that Id(K, K ′) ≤ I. (cid:3) 4 M. M. SADR It is well known that (3) Ld(f ) = sup{ µ(f ) − ν(f ) ρd(µ, ν) : µ, ν ∈ SCX, µ 6= ν} (f ∈ CXsa). Also the following formula follows from (1) and Proposition 3. λ′ − λ λ′ − λ (4) Ld(f ) = sup λ<λ′∈R Id(f −1λ′, f −1λ) Iρd(Ff −1λ′, Ff −1λ) = sup λ<λ′∈R (f ∈ CXsa). 3. Compact quantum metric spaces For the theory of order unit spaces we refer the reader to [5]. We denote the state space of an order unit space B by SB. This space is always considered with the weak* topology. Let A be a unital C*-algebra with the self-adjoint part Asa and state space SA. Suppose that B is any real linear subspace of Asa that contains 1A. Then B together with the usual partial ordering between self-adjoint elements and 1A as order unit becomes an order unit space. Moreover if B is dense in Asa (with the norm topology) then the mapping µ 7→ µB defines an affine homeomorphism from SA with the weak* topology onto SB. It is clear that the state space of an order unit space with weak* topology is a compact convex subset of a locally convex Hausdorff space. The converse of this fact is also well known (see the details after Corollary II.2.3 of [5]). Indeed, let E be a compact convex subset of a locally convex Hausdorff space; if AE denotes the order unit space of all continuous affine real valued functions on E with the constant function 1E as order unit, then E and SAE are affinely homeomorphic via the map x 7→ (f 7→ f (x)) (x ∈ E). Thus there is no difference that we formulate our results in terms of order unit spaces or else using compact convex sets. Let B be an order unit space and L be a seminorm on B. By analogy with Formula (2) we define a pseudo-metric on SB as follows. (5) ρL(µ, ν) := sup{µ(b) − ν(b) : L(b) ≤ 1} (µ, ν ∈ SB) Note that in general ρL does not separate the points and may take value +∞. Definition 4. ([23, 24, 25]) Let B be an order unit space with order unit e and let L be a seminorm on B satisfying the following two conditions: i) L(b) = 0 if and only if b = λe for some λ ∈ R. ii) The topology induced by ρL, given by Formula (5), coincides with the weak*-topology on SB. Then the pair (B, L) is called a compact quantum metric space. Also if a unital C*-algebra A is given such that B as an order unit space is a subspace of Asa containing 1A, and B is dense in Asa w.r.t. the C*-norm, then (A, B, L) is called a C*-algebraic compact quantum metric space. In the case that (B, L) is understood we say that qA is a C*-algebraic quantum metric space. Let (X, d) be an ordinary compact metric space. Then (CX, LipdX, Ld) is a C*-algebraic compact quantum metric space. Also by (2) and (5) we have ρLd = ρd. Thus the structure of (X, d) is completely recovered by (CX, LipdX, Ld). We remark that there are examples of C*-algebraic compact quantum metric spaces (A, B, L) with A = CX for a compact space X such that L does not arise from any ordinary metric d on X i.e. L 6= Ld, see [23, Example 7.1 and Theorem 8.1]. For other examples of nonclassical quantum metric spaces we refer the reader to the list of papers in Introduction. HYPERSPACES ASSOCIATED TO A NONCOMMUTATIVE SPACE 5 4. Hyperspace of closed sets in a NC space Let A be a unital C*-algebra with the state space SA. Let A′′ denote the second com- mutant of A in the universal representation of A. By the Sherman Theorem the second dual A∗∗ is canonically isomorphic to the von Neumann algebra A′′ where A∗∗ is consid- ered as a C*-algebra with the Arens product. A projection p ∈ A∗∗ is called closed [1] if there is a decreasing net of positive elements of A that converges to p in the weak* topol- ogy. A projection q is called open if 1 − q is closed. For every projection p ∈ A∗∗ we let Fp := {µ ∈ SA : hµ, pi = 1}. In the case that A = CX for a compact Hausdorff space X, there is a bijection K 7→ pK between closed subsets K of X and closed projections of A∗∗ such that FK = FpK with the notations of Section 2. Proposition 5. The assignment p 7→ Fp is a bijection between closed projections of A∗∗ and weak*-closed faces of SA. Proof. It follows from [21, Theorems 3.6.11 and 3.10.7] or [4, Theorem 2.5]. (cid:3) Let E be a compact convex subset of a locally convex Hausdorff space. We denote by SclcE the set of nonempty closed convex subsets of E, and by Sclf E the set of all closed faces of E. Thus we have the chain Sclf E ⊂ SclcE ⊂ SclE of Hyperspaces. Throughout the paper these hyperspaces are endowed with Vietoris topology. Let B be an order unit space. In [25] (imaginary) closed subsets of the quantum space qB are identified with elements of SclcSB. As we saw above it is more natural to consider the closed subsets as elements of SclfSB. So by analogy with the notations of Section 2 we would use the symbol SclqB instead of SclfSB. Analogously, for a unital C*-algebra A we let SclqA := SclfSA. In the case that A = CX for a compact Hausdorff space X it follows from Proposition 1 that SclqA is homeomorphic to SclX. Suppose that X has a compatible metric d and consider the C*-algebraic quantum metric space (CX, LipdX, Ld). Let ρ := ρLd = ρd. It follows from Proposition 2 that the metric spaces (SclqA, Hρ) and (SclX, Hd) are isometrically isomorphic. Also it follows from Proposition 3 that the distance functions Iρ on SclqA and Id on SclX coincide when the two spaces are considered canonically identical. 5. Vietoris topology Throughout this section E denotes a compact convex subset of a locally convex Hausdorff space. The following result stated as Theorem 7 is very well known, at least in the case that E is metrizable; but we did not find in literatures any proof for the general case; (Let Λ, Λ′ be directed sets however its proof is easy and based on the following lemma. and (xλ)λ∈Λ be a net in X. Let f : Λ′ → Λ be an order preserving function such that ∀λ ∈ Λ, ∃λ′ ∈ Λ′ : f (λ′) ≥ λ. Then the net (xf (λ′))λ′∈Λ′ is called a subnet of (xλ)λ∈Λ.) Lemma 6. Let X be a compact Hausdorff space and (Kλ)λ a net in SclX converging to K. (i) If (xλ)λ is a net in X such that xλ → x and xλ ∈ Kλ, then x ∈ K. (ii) If x ∈ K, then there are a subnet (Kλ′)λ′ of (Kλ)λ and a net (xλ′)λ′ such that xλ′ ∈ Kλ′ and xλ′ → x. Proof. Straightforward. Theorem 7. The hyperspace SclcE is a compact Hausdorff space. (cid:3) 6 M. M. SADR Proof. It is enough to show that SclcE is a closed subset of SclE. Suppose that (Kλ)λ is a net in SclcE converging to K ∈ SclE. We must show that K is convex. Suppose that x, y ∈ K and 0 ≤ t ≤ 1. By Lemma 6(ii) there exist a subnet (Kλ′)λ′ of (Kλ)λ and nets (xλ′)λ′, (yλ′)λ′ such that xλ′, yλ′ ∈ Kλ′ and xλ′ → x, yλ′ → y. Thus (txλ′ + (1 − t)yλ′)λ′ is a net in Kλ′ converging to tx + (1 − t)y. Now it follows from Lemma 6(i) that tx + (1 − t)y ∈ K. The proof is complete. (cid:3) For metrizable E we have the following strong result of Nadler-Quinn-Stavrakos: Theorem 8. Suppose that E is metrizable and the real dimension of the smallest real hy- perplane containing E is ≥ 2. Then SclcE is homeomorphic to Hilbert cube. Proof. This is a restatement of [18, Theorem 2.2]. (Note that in the proof of [18, Theorem 2.2] it is enough that K be metrizable.) (cid:3) For some results similar to Theorem 8 in the case that E is not metrizable see [9]. A direct consequence of Theorem 8 is the following. Corollary 9. Let (B, L) be a compact quantum metric space such that the (real vector space) dimension of B is ≥ 2. Then SclcSB is homeomorphic to Hilbert cube. In particular, if qA is a C*-algebraic quantum metric space such that A 6= 0, C then SclcSA is homeomorphic to Hilbert cube. Let ∂eE denote the subspace of extreme points of E. Theorem 10. If the hyperspace SclfE is compact then ∂eE is compact. Proof. Suppose that SclfE is compact. We must show that ∂eE is a closed subset of E. Let (eλ)λ be a net in ∂eE converging to x ∈ E. Since eλ is an extreme point {eλ} is a closed face of E. Thus there is a subnet (eλ′)λ′ such that {eλ′} → K in Sclf E. Now it follows from Lemma 6 that K = {x} which means that x ∈ ∂eE. Thus ∂eE is a closed subset of E. (cid:3) In general the converse of Theorem 10 is not satisfied even if E is finite dimensional, see [22]. We say that E is stable [20] if for every 0 ≤ t ≤ 1 the map (x, y) 7→ tx + (1 − t)y from E × E into E is open. Among examples of stable compact convex sets are Bauer simplices [19, Theorem 1]. For a complete account on Bauer simplices see [5]. It is well known that a unital C*-algebra is commutative if and only if its state space with weak*-topology is a Bauer simplex [6, Remark in page 296],[8]. To our knowledge the following result has not been mentioned before in the literatures. Theorem 11. If E is stable then SclfE is compact. Proof. Suppose that E is stable. We must show that SclfE is a closed subset of SclcE. Let (Kλ)λ be a net in SclfE converging to K ∈ SclcE. We show that K is a face of E. Suppose that for some 0 ≤ t ≤ 1 and x, y ∈ E we have z := tx + (1 − t)y ∈ K. By Lemma 6(ii) there exist a subnet (Kλ′)λ′ of (Kλ)λ and a net (zλ′)λ′ such that zλ′ ∈ Kλ′ and zλ′ → z. Since the map φ : (x′, y′) 7→ tx′ + (1 − t)y′ is open and continuous, for every open U in E containing z there are opens V, W respectively containing x, y such that φ(V × W ) is an open in U. This property enables us to find a subnet (zλ′′)λ′′ of (zλ′)λ′ and nets (xλ′′)λ′′, (yλ′′)λ′′ such that xλ′′ → x, yλ′′ → y and zλ′′ = txλ′′ + (1 − t)yλ′′. Since Kλ′′ is a face we have xλ′′ , yλ′′ ∈ Kλ′′. Now it follows from Lemma 6(i) that x, y ∈ K. The proof is complete. (cid:3) HYPERSPACES ASSOCIATED TO A NONCOMMUTATIVE SPACE 7 A direct consequence of Theorems 10 and 11 is the following result. Corollary 12. Let A be a unital C*-algebra. If SclqA is compact then the space of pure states of A is weak*-compact. If SA is stable then SclqA is compact. 6. Infimum distance and an analog of Lipschitz seminorm Let A be a unital C*-algebra and let ρ be a compatible metric on SA. Let L1 : Asa → [0, ∞] be a seminorm given by the analog of Formula (3): L1(a) = sup{ µ(a) − ν(a) ρ(µ, ν) : µ, ν ∈ SA, µ 6= ν}. Let H denote the Hilbert space of the universal representation of A, and BH be the algebra of bounded operators on H. Then by definition we have A ⊂ A′′ ⊆ BH. For a ∈ Asa let Ea denotes the spectral measure of a defined on the Borel subsets of R where a is considered as an element of BH. It is well known that for every closed subset S of R the projection Ea(S) is a closed projection in A′′. (The converse is also true [2, Theorem A1],[3], that is if a ∈ A′′ sa and Ea(S) is a closed projection for every closed subset S ⊆ R then a ∈ A.) For a ∈ Asa and λ ∈ R let Fa,λ denote the weak*-closed face of SA corresponding to the closed projection Ea({λ}) as in Proposition 5. (In the case that Ea({λ}) = 0 we let Fa,λ = ∅.) In Section 2 we restated the definition of Lipschitz seminorm for an ordinary metric space as Formula (4). Now analogously we define a function L2 : Asa → [0, ∞] by L2(a) = sup λ<λ′∈R λ′ − λ Iρ(Fa,λ′, Fa,λ) . We have L2 ≤ L1 but in general L2 is not a seminorm. Question 13. Under which conditions is L2 a seminorm (with Leibniz property) on any commutative subalgebra of A? Suppose ρ is induced by a C*-algebraic quantum metric structure (A, B, L) i.e. ρ = ρL. Then we have L2(a) ≤ L1(a) ≤ L(a) for a ∈ B. As we saw in Section 2 in the classical case (A, B, L) = (CX, LipdX, Ld) we have L2 = L1 = L. By Theorem 4.1 of [23] we know that if L is lower semicontinuous (which means {a ∈ B : L(a) ≤ 1} is closed in B w.r.t. the C*-norm) then L1(a) = L(a) for every a ∈ B. 7. Some questions and problems We saw that for a unital commutative C*-algebra A, SclqA is compact. Problem 14. Characterize those unital C*-algebras A such that SclqA is compact. Question 15. For which C*-algebras A, is SclqA (path or locally path) connected? (See [7] in the classical case.) Let Mn denote the C*-algebra of n× n matrixes. In NC Geometry qMn is usually consid- ered as the finite NC space with n points. Since Mn = M∗∗ any projection in Mn is closed and open ([1, Proposition II.18]), and hence SclqMn as a set is canonically identified with ∪n i=1Gr(i, n) where Gr(i, n) denote the Grassmannian manifold of i-dimensional subspaces of Cn. 8 M. M. SADR Question 16. Is the subspace of SclqMn containing projections of rank i homeomorphic to Gr(i, n)? Let X be a compact Hausdorff space and C be a C*-subalgebra of CX containing 1X. Let Z denote the pure state space of C with weak*-topology. We have a canonical continuous surjective map Γ : X → Z defined by Γ(x)(c) = c(x) (c ∈ C). It is easily checked that the topology of Z is the quotient topology under Γ. Also Γ induces the family {Kz}z∈Z of nonempty disjoint closed subsets of X parameterized by Z where Kz := Γ−1(z). A generalization of this notion is as follows. Definition 17. Let A be a unital C*-algebra and C be a C*-subalgebra of A containing the unit. Let Z denote the pure state space of C. For every z ∈ Z let Fz := {µ ∈ SA : µ(c) = z(c), c ∈ C}. Then Fz is a weak*-closed face of SA. We say that {Fz}z∈Z is the family of closed subsets of qA parameterized by qC. θ := qCT2 Let 0 ≤ θ < 1. The quantum torus T2 θ is the NC space associated to the universal C*-algebra CT2 θ generated by two unitary elements u, v satisfying uv = e2πiθvu. Let T := {z ∈ C : z = 1} denote the unit circle. We identify the C*-subalgebra generated by v with CT via the *-isomorphism given by the assignment v 7→ idT where idT ∈ CT is the identity function. For every z ∈ T let Tθ,z := {µ ∈ SCT2 θ : µ(f ) = f (z), f ∈ CT}. Then we call {Tθ,z}z∈T the family of v-sub-circles in T2 θ. The name is justified as follows. It is clear that CT2 0 can be identified with CT2 via the *-isomorphism given by the assignments u 7→ id1, v 7→ id2 where id1, id2 ∈ CT2 are respectively the projection functions on the first and second components of T2 = T × T. Then T0,z is identified with the set of Borel probability measures µ on T2 such that the support of µ is contained in the sub-circle {(w, z) ∈ T2 : w ∈ T}. It is not hard to see that the map z 7→ {(w, z) ∈ T2 : w ∈ T} from T into SclT2 is continuous with Vietoris topology. So it is natural to ask the following questions. Question 18. Is the map z 7→ Tθ,z (θ 6= 0) from T into SclT2 v-sub-circles in T2 θ compact or (path) connected? θ continuous? Is the family of If we have a (Riemannian) metric on T2 we can ask about the Hausdorff and infimum distances of sub-circles. Analogously we have the following problem. Problem 19. Consider T2 find the Hausdorff and infimum distances between two arbitrary sub-circles Tθ,z and Tθ,z ′. θ as a C*-algebraic quantum metric space described in [24, 25] and References 1. C.A. Akemann, The general Stone-Weierstrass problem, J. Func. Anal. 4, no. 2 (1969): 277–294. 2. C.A. Akemann, Left ideal structure of C*-algebras, J. Func. Anal. 6, no. 2 (1970): 305–317. 3. C.A. Akemann, A Gelfand representation theory for C*-algebras, Pacif. J. Math. 39, no. 1 (1971): 1–11. 4. C.A. Akemann, G.K Pedersen, Facial structure in operator algebra theory, Proc. Lon. Math. Soc. 3, no. 2 (1992): 418–448. 5. E.M. Alfsen, Compact convex sets and boundary integrals, Springer-Verlag, New York, 1971, Ergebnisse der Mathematik und ihrer Grenzgebiete, Band 57. 6. E.M. Alfsen, F.W. Shultz, Geometry of state spaces of operator algebras, Springer Science & Business Media, 2012. 7. T. Banakh, R. Voytsitskyy, Characterizing metric spaces whose hyperspaces are absolute neighborhood retracts, Top. & App. 154, no. 10 (2007): 2009–2025. (arXiv:math/0509395 [math.GT]) HYPERSPACES ASSOCIATED TO A NONCOMMUTATIVE SPACE 9 8. C.J.K. Batty, Simplexes of states of C*-algebras, J. Operator Theory 4 (1980): 3–23. 9. L. Bazylevych, D. Repovs, M. Zarichnyi, Hyperspace of convex compacta of nonmetrizable compact convex subspaces of locally convex spaces, Top. & App. 155, no. 8 (2008): 764–772. (arXiv:0803.4243 [math.GN]) 10. D.P. Blecher, M. Neal, Open projections in operator algebras II: compact projections, Studia Math, 209 (2012), 203–224. (arXiv:1109.5347 [math.OA]) 11. H. Comman, Capacities on C*-algebras, Infinite Dimensional Analysis, Quantum Probability & Related Topics 6, no. 03 (2003): 373–388. 12. D. Guido, T. Isola, The problem of completeness for GromovHausdorff metrics on C*-algebras, J. Func. Anal. 233, no. 1 (2006): 173–205. (arXiv:math/0502013 [math.OA]) 13. A. Illanes, S.B. Nadler, Hyperspaces, fundamentals and recent advances, Vol. 216. CRC Press, 1999. 14. D. Kerr, H. Li, On Gromov-Hausdorff convergence for operator metric spaces, (2004). (arXiv:math/0411157 [math.OA]) 15. G. Kuperberg, N. Weaver, A von Neumann algebra approach to quantum metrics/Quantum relations, Vol. 215, no. 1010. American Mathematical Society, 2012. (arXiv:1005.0353 [math.OA]) 16. F. Latr´emoli`ere, Curved noncommutative tori as Leibniz quantum compact metric spaces, J. Math. Phys. 56 (2015), no. 12, 123503, 16 pp. (arXiv:1507.08771 [math.OA]) 17. P. Martinetti, F. Mercati, L. Tomassini, Minimal length in quantum space and integrations of the line element in noncommutative geometry, Rev. Math. Phy. 24, no. 05 (2012): 1250010. (arXiv:1106.0261 [math-ph]) 18. S.B. Nadler Jr., J. Quinn, N.M. Stavrakos, Hyperspace of compact convex sets, Pacif. J. Math. 83(1979), 441-462. 19. R.C. O'Brien, On the openness of the barycentre map, Math. Ann. 223, no. 3 (1976): 207–212. 20. S. Papadopoulou, On the geometry of stable compact convex sets, Math. Ann. 229, no. 3 (1977): 193–200. 21. G.K. Pedersen, C*-algebras and their automorphism groups, London Mathematical Society Monographs 14, Academic Press, London, 1979. 22. H.B. Reiter, N.M. Stavraks, On the compactness of the hyperspace of faces Pacif. J. Math. 73, no. 1 (1977): 193–196. 23. M.A. Rieffel, Metrics on state spaces, Doc. Math. 4 (1999): 559-600. (arXiv:math/9906151 [math.OA]) 24. M.A. Rieffel, Group C*-algebras as compact quantum metric spaces, Doc. Math 7 (2002): 605–651. (arXiv:math/0205195 [math.OA]) 25. M.A. Rieffel, Gromov-Hausdorff distance for quantum metric spaces/Matrix algebras converge to the sphere for quantum Gromov-Hausdorff distance, Vol. 168, no. 796. American Mathematical Soc., 2004. (arXiv:math/0011063 [math.OA]) (arXiv:math/0108005 [math.OA]) 26. W. Rudin, Functional analysis, International series in pure and applied mathematics, 1991. 27. M.M. Sadr, Quantum metrics on noncommutative spaces, (2016). (arXiv:1606.00661 [math.OA]) 28. N. Weaver, Lipschitz algebras, PhD diss., University of California, Berkeley, 1994. 29. W. Wu, Quantized Gromov-Hausdorff distance, J. Func. Anal. 238, no. 1 (2006): 58–98. (arXiv:math/0503344 [math.OA]) Department of Mathematics, Institute for Advanced Studies in Basic Sciences, P.O. Box 45195-1159, Zanjan 45137-66731, Iran E-mail address: [email protected]
1001.4202
1
1001
2010-01-23T20:02:30
An index theorem to solve the gap-labeling conjecture for the pinwheel tiling
[ "math.OA", "math.KT" ]
In this paper, we study the K0-group of the C?-algebra associated to a pinwheel tiling. We prove that it is given by the sum of Z + Z^6 with a cohomological group. The C?-algebra is endowed with a trace that induces a linear map on its K0-group. We then compute explicitly the image, under this map, of the summand Z+Z^6, showing that the image of Z is zero and the image of Z^6 is included in the module of patch frequencies of the pinwheel tiling. We finally prove that we can apply the measured index theorem due to A. Connes to relate the image of the last summand of the K0-group to a cohomological formula which is more computable. This is the first step in the proof of the gap-labeling conjecture for the pinwheel tiling.
math.OA
math
An index theorem to solve the gap-labeling conjecture for the pinwheel tiling Haıja MOUSTAFA Abstract In this paper, we study the K0-group of the C ∗-algebra associated to a pinwheel tiling. We prove that it is given by the sum of Z ⊕ Z6 with a cohomological group. The C ∗-algebra is endowed with a trace that induces a linear map on its K0-group. We then compute explicitly the image, under this map, of the summand Z⊕ Z6, showing that the image of Z is zero and the image of Z6 is included in the module of patch frequencies of the pinwheel tiling (see [Mou]). We finally prove that we can apply the measured index theorem due to A. Connes ([Con79]) to relate the image of the last summand of the K0-group to a cohomological formula which is more computable. This is the first step in the proof of the gap-labeling conjecture for the pinwheel tiling, the second step is done in [Mou] where we study the cohomological formula obtained by the index theorem. Contents 1 Introduction 2 Reminders 2.1 Pinwheel tiling and continuous hull . . . . . . . . . . . . . . . . . 2.2 The canonical transversal . . . . . . . . . . . . . . . . . . . . . . 2.3 C∗-algebra associated to a tiling . . . . . . . . . . . . . . . . . . 3 Gap-labeling 3.1 Quasicrystals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2 Integrated density of states - IDS . . . . . . . . . . . . . . . . . . 3.3 Shubin's formula . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.4 Gap-labeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4 Index theorem for the gap-labeling of the pinwheel tiling 4.1 Study of K0(cid:0)C(Ω) ⋊ R2 ⋊ S1) . . . . . . . . . . . . . . . . . . . . ∗ of the summand Z . . . . . c (cid:0)(Ω \ F )/S1 ; Z(cid:1) in K0(cid:0)C(Ω) ⋊ R2 ⋊ S1(cid:1) Index theorem to compute the trace . . . . . . . . . . . . . . . . ∗ of the summands Z.qi . . . . 4.2 Computation of the image under τ µ 4.3 Study of the summand H 2 4.4 4.5 Computation of the image under τ µ 5 Conclusion A Proof of lemma 4.5 1 2 9 9 12 15 16 16 17 18 20 21 22 26 30 33 45 47 50 1 Introduction In 1982, D. Shechtman discovered, in a rapidly solidified aluminium alloy, a phase similar to the one obtained from crystals ([SBGC84]). To study the atomic distribution of such a solid, he realized its diffraction diagram that is shown in Figure 1. Figure 1: Diffraction diagram. This diagram is discrete and is similar to the one obtained from crystals. From its study, he could deduce that the atoms are distributed in such a way that the distance between two atoms is bounded below and such that there is only a finite number of local configurations upto translations. However, the diffraction diagram of this solid has a 10-fold symmetry that is forbidden by crystallographic classification. The structure of this solid is thus different from the one of crystals. In particular, the atomic distribution isn't stable by any translation. Neverthe- less, each local configuration is repeted uniformly in the alloy. For these reasons, such a solid was called quasicrystal. This discovery gave rise to a big interest in solid physics and in mathematics. In mathematics, this atomic distribution is naturally modeled by aperiodic tilings encoding its geometric properties into combinatoric properties. A tiling of Rn is a countable family P = {t0, t1, . . .} of non empty compact sets ti, called tiles, each tile being homeomorphic to the unit ball, such that the ti's cover Rn, with a fixed origin, and the tiles meet each other only on their border. In the sequel, we consider the particular case in which the ti's are G-copies of elements of a finite family {p0, . . . , pn}, where G is some subgroup of the group of rigid motions of Rn. The pk's are called prototiles. A patch is a finite collection of tiles in a tiling. A tiling is of finite G-type if, for all R > 0, there are only finitely many patches of diameter less than or equal to R upto the G-action. A tiling T ′ is G-repetitive if, for each patch A in T ′, there exists R(A) > 0 such that every ball of radius R(A) in Rn meets T ′ on a patch containing a G-copy of A. In this paper, a G-aperiodic tiling is a nonperiodic tiling (for translations of Rn) of finite G-type and G-repetitive. 2 The atoms of a quasicrystal are located in the tiles of the aperiodic tiling mod- eling it. Motivated by electromagnetic and macroscopic properties of such solids, J. Bellissard studied the electonic motion in a quasicrystal (see [Bel82], [Bel86] and [Bel92]). This motion is closely related to the spectral gaps of the Shrodinger operator associated to the solid. This operator is defined as follows : H = − ℏ 2m ∆ + V (x) 2m is some constant, ∆ is the Laplacian on Rn and V is some potential where − ℏ depending only on the structure of the studied quasicrystal. V is given by V (x) = Xy∈L v(x − y) where L is the point set of equilibrium atomic positions in the quasicrystal and v the effective potential for valence electrons near an atom. In other words, v is the function governing the interaction be- tween an electron and an atom (see [BHZ00]). Bellissard looked for a mathematical way to label the gaps of the spectrum of this Schrodinger operator and this is the aim of its gap-labeling conjecture. Tilings were studied before the discovery of quasicrystals but the interest generated by this new material greatly increased progresses made in their study, looking tiling properties from a new point of view, the noncommutative one. In 2000, in [KP00], J. Kellendonk and I. Putnam associated a C∗-algebra to a G-aperiodic tiling to study its properties. It is the crossed product of the continuous functions on a topological space Ω by G. The space Ω encodes the combinatorial properties of the tiling into topological and dynamical properties. We recall, in section 2, how this space is obtained together with other defi- nitions and the construction of the C∗-algebra associated to tilings. To construct Ω, let's fix an orthonormal basis of Rn and a G-aperiodic tiling T . Letting G act on the right on T , we have a family of tilings T .G and one can take its closure under some distance (adapted to such tilings) to obtain the continuous hull Ω of T endowed with a natural action of G and containing tilings with the same properties as T (this construction is due to Kellendonk [Kel95]) . The combinatorial properties of T ensure that Ω is a compact topological space and that each orbit under the action of G is dense in Ω. The canonical transversal Ξ is the set of all the tilings of Ω satisfying some rigid conditions. This is a Cantor space and allows us to see Ω as a foliated space, i.e Ω is covered by open sets given by K × U , where K is a clopen (closed and open) subset of Ξ and U is an open subset of G, with some smooth conditions on transition maps. The space Ξ is then a transversal of Ω and the leaves of Ω are homeomorphic to G or to a quotient of G by a discrete subgroup. We can associate to the dynamical system (Ω, G) the C∗-crossed product C(Ω) ⋊ G which is defined in the end of section 2. 3 Since G is amenable, Ω is endowed with a G-invariant, ergodic probability mea- sure µ. This measure induces a densely defined trace τ µ on C(Ω) ⋊ G and a linear map τ µ ∗ : K0(cid:0)C(Ω) ⋊ G(cid:1) → R on the K-theory of C(Ω) ⋊ G. In section 3, we present the works made by J. Bellissard. For G = Rn, J. Bellissard linked the spectral gaps of the Schrodinger operator H to this K-theory, noting that each spectral gap gives rise to an element of K0(cid:0)C(Ω) ⋊ Rn(cid:1) (see [Bel92] and [BHZ00]). He also showed that the image under τ µ ∗ of K0(cid:0)C(Ω) ⋊ Rn(cid:1) is related to some physical function N : R → R+, called the integrated density of states, abrevi- ated IDS (see p.17). For E ∈ R, N (E) can be seen as the number of eigenvalues of H per unit volume up to E. Thus, this is a nonnegative, nondecreasing function on R and thereby, it defines a Stieljes-Lebesgue measure dN on R given by dN(cid:0)[E; E′](cid:1) := N (E′) − N (E). dN is absolutely continuous with respect to the Lebesgue measure dE (i.e RA dE = 0 ⇒RA dN = 0) and, by the Radon-Nikodym theorem, we can define the density of states to be the derivative dN /dE. This density of states is a well known quantity in solid state physics which is in fact accessible by scattering experiments (see [Bel92]). This quantity is important in physics to deduce conductivity properties of qua- sicrystals and it is crucial to predict its values. For this, J. Bellissard proved that this function is related to a mathematical object thanks to Shubin's formula. Shubin's formula expresses that, if E is in a spectral gap, there is a projection pE in C(Ω) ⋊ Rn such that N (E) is equal to τ µ(pE) and that this does not depend on E chosen in the spectral gap. Thereby, the integrated density of states N takes values in the image, under τ µ ∗ , of K0(cid:0)C(Ω) ⋊ Rn(cid:1). To study conductivity properties of quasicrystals, it is thus important to know this image. The gap-labeling conjecture established by Bellissard predicts this image : µ induces a measure µt on the canonical transversal Ξ (given locally by the quotient of µ(K × U ) by the volume of U ) and the conjecture expresses the link between the image of K0(cid:0)C(Ω) ⋊ Rn(cid:1) under τ µ ∗ and the image, under µt, of continuous functions on Ξ with integer values: Conjecture : ([Bel92], [BHZ00]) τ µ ∗(cid:16)K0(cid:0)C(Ω) ⋊ Rn(cid:1)(cid:17) = µt(cid:0)C(Ξ, Z)(cid:1) In this conjecture, µt(cid:0)C(Ξ, Z)(cid:1) is the space (cid:26)ZΞ f dµt; f ∈ C(Ξ, Z)(cid:27). Since then, several works have been done to prove this conjecture. The Pimsner-Voiculescu exact sequence was used by J. Bellissard in [Bel92] to prove the conjecture for n = 1 and A. van Elst iterated this sequence twice in [vE94] to prove the case n = 2. J. Bellissard, J. Kellendonk and A. Legrand have proved the conjecture for n = 3 in [BKL01] using a spectral sequence due to Pimsner. 4 Finally, in 2002, the conjecture was proved in full generality for Rn-aperiodic tilings in three independent papers : by J. Bellissard, R. Benedetti and J.-M. Gambaudo in [BBG06], by M.-T. Benameur and H. Oyono-Oyono in [BOO02] and by J. Kaminker and I. Putnam in [KP03]. A natural question is to know if this conjecture remains true for more general tilings. This paper gives a first step to solve the gap-labeling conjecture in the particular case of the (1 , 2 )-pinwheel tiling of the plane introduced by Conway and Radin (see [Rad94] and [Rad95]) which is not a R2-aperiodic tiling but a R2 ⋊ S1- aperiodic one. It is a tiling built from two right triangles of side 1, 2 and √5, one being the mirror image of the other. The substitution method gives a pinwheel tiling as follows : we can cover the stretched image, by a √5 factor, of the two triangles by the union of copies by a rigid motion of these triangles, these copies meeting only on their border, as follows : Figure 2: Substitution of the pinwheel tiling. Iterating this process, we build a union of triangles that covers bigger and bigger regions of the plane and that, in the limit, covers the whole plane and, thus, gives a tiling of the plane called a pinwheel tiling. A patch of this tiling is shown in Figure 3. This tiling is nonperiodic for translations of the plane, of finite R2 ⋊ S1-type and R2 ⋊ S1-repetitive, where R2 ⋊ S1 is the group of rigid motion of the plane. It is not of finite R2-type since the triangles appear in infinitly many orientations. The continuous hull Ω of this tiling contains six circles fixed by a rotation of angle π around the origin. The union of these 6 circles is denoted F and these are the only fixed points for rotations. The constrution explained above can be applied to obtain a dynamical sys- tem (Ω, R2 ⋊ S1), where Ω is a compact topological space and such that each R2 ⋊ S1-orbit is dense in Ω. 5 Figure 3: Patch of a pinwheel tiling. The canonical transversal Ξ can still be built and Ω is a foliated space with leaves homeomorphic to R2 ⋊ S1 except six that are homeomorphic to the quotient of R2 ⋊ S1 by the subgroup generated by (0, Rπ). In this paper, we give a first step of the proof of the following theorem, the second step is given in [Mou] : Theorem : If T is a pinwheel tiling, Ω = Ω(T ) its hull provided with an invariant ergodic probability measure µ and Ξ its canonical transversal provided with the induced measure µt, then τ µ ∗(cid:16)K0(cid:0)C(Ω × R2 ⋊ S1)(cid:1)(cid:17) = µt(cid:0)C(Ξ, Z)(cid:1). This theorem aims to link the image of K0(cid:0)C(Ω × R2 ⋊ S1)(cid:1) under τ µ Z-module of patch frequencies µt(cid:0)C(Ξ, Z)(cid:1). For this, we follow the guideline of the works made by M.-T. Benameur and ∗ to the H. Oyono-Oyono in [BOO02]. In their paper, they used the measured index theorem proved by A. Connes (see [Con79] or [MS06]) to link the analytical part τ µ ∗(cid:16)K0(cid:0)C(Ω × Rn)(cid:1)(cid:17) to a topological part, easier to compute, Chτ(cid:0)Kn(C(Ω))(cid:1) which is in H∗τ(cid:0)Ω(cid:1), the tangential cohomology group of Ω (Chτ is the tangential Chern character, see [MS06]). In section 4, we prove that this theorem can still be applied for pinwheel tilings. For this, we want to prove that any element of K0(cid:0)C(Ω) ⋊ R2 ⋊ S1(cid:1) can be seen as the analytical index of a Dirac operator "twisted by a unitary" of K1(cid:0)C(Ω)(cid:1), i.e for any b ∈ K0(cid:0)C(Ω) ⋊ R2 ⋊ S1(cid:1), we want to find a [u] ∈ K1(cid:0)C(Ω)(cid:1) such that b = [u]⊗C(Ω) [D3], where [D3] ∈ KK1(cid:0)C(Ω), C(Ω) ⋊ R2 ⋊ S1(cid:1) is the class of the tangential Dirac operator D3 along the leaves of Ω and ⊗C(Ω) is the Kasparov product over C(Ω). 6 ∗ (b) = τ µ apply the index theorem to obtain : a subgroup H (see below for more details on H). The image under τ µ In fact, we prove that K0(cid:0)C(Ω) ⋊ R2 ⋊ S1(cid:1) is isomorphic to the sum of Z7 and ∗ of Z7 is explicitly computable and for any b ∈ H, there is some [u] ∈ K1(cid:0)C(Ω)(cid:1) such that τ µ ∗(cid:0)[u] ⊗C(Ω) [D3](cid:1), which is enough to Theorem 4.13 : ∀b ∈ H, ∃[u] ∈ K1(cid:0)C(Ω)(cid:1) such that τ µ ∗ (b) = τ µ where [Cµt ] ∈ H τ 3 (Ω) is the Ruelle-Sullivan current associated to µt (locally, it is given by integration on Ξ × R2 × S1, see [MS06]) and h , i is the pairing between the tangential cohomology and homology groups. ∗ ([u] ⊗C(Ω) [D3]) =(cid:10)Chτ ([u]), [Cµt ](cid:11), To obtain this theorem, we see Ω as a double foliated space. First, it is foliated by the action of S1 ⊂ R2 ⋊ S1 and we can build the map ψ1 : K1(cid:0)C(Ω)(cid:1) → K0(cid:0)C(Ω) ⋊ S1(cid:1) given by the Kasparov product, over C(Ω), with the class of the cycle induced by the tangential Dirac operator d1 along the leaves S1. Then, Ω is also foliated by the action of R2 ⋊ S1 and thus, the Kasparov cy- cle defined by the tangential Dirac operator D2 transverse to the inclusion of the foliation by S1 into the foliation by R2 ⋊ S1 is used to construct the map ψ2 : K0(cid:0)C(Ω) ⋊ S1(cid:1) → K0(cid:0)C(Ω) ⋊ R2 ⋊ S1(cid:1) given as the Kasparov product, over C(Ω) ⋊ S1, with the class of this cycle in KK0(cid:0)C(Ω) ⋊ S1, C(Ω) ⋊ R2 ⋊ S1(cid:1). Using the Dirac-dual Dirac construction (see [Kas88], [Ska91]), one can show that ψ2 is an isomorphism. We then prove that ψ2 ◦ ψ1 is given by the Kasparov product with the class of the cycle induced by the tangential Dirac operator of dimension 3 (see [HS87] where, in the case of foliations, they studied the product of Gysin maps associ- ated to double foliations). To prove these results, we first use a six term exact sequence in K-theory to show ( H 2(·; Z) is the Cech cohomology group with integer coefficients and H 2 c (·; Z) is the one with compact support) : Proposition 4.6 : Then, we prove that τ µ K0(C(Ω) ⋊ S1) ≃ Z ⊕ Z6 ⊕ H 2 c(cid:16)(Ω \ F )/S1 ; Z(cid:17) ∗(cid:0)ψ2(Z)(cid:1) = 0 and to study the image, under τ µ ∗ , of c(cid:16)(Ω\ F )/S1 ; Z(cid:17) is isomorphic H := ψ2(cid:16) H 2 to H 3(cid:0)Ω ; Z(cid:1) and that this isomorphism is ψ1, modulo the Chern character. Thus, ψ2 ◦ ψ1 is surjective on the H summand of K0(cid:0)C(Ω) ⋊ R2 ⋊ S1(cid:1). Using a result proved by Douglas, Hurder and Kaminker in [DHK91] on the odd index theorem for foliated spaces, we obtain the main theorem of this paper : c(cid:0)(Ω\ F )/S1 ; Z(cid:1)(cid:17), we prove that H 2 Theorem 4.13 : ∀b ∈ H, ∃[u] ∈ K1(cid:0)C(Ω)(cid:1) such that : τ µ ∗ (b) = τ µ ∗ ([u] ⊗C(Ω) [D3]) =(cid:10)Chτ ([u]), [Cµt ](cid:11). 7 Finally, an explicit computation, using the index theorem for foliated spaces on Ω seen as a foliated space for the R2-action, gives the inclusion In this paper, we thus obtain the following result : τ µ ∗(cid:0)ψ2(Z6)(cid:1) ⊂ µt(C(Ξ, Z)). Theorem : If T is a pinwheel tiling, Ω = Ω(T ) its hull provided with an invariant ergodic probability measure µ and Ξ its canonical transversal provided with the induced measure µt, then K0(C(Ω) ⋊ R2 ⋊ S1) ψ2≃ Z ⊕ Z6 ⊕ H 2 c(cid:16)(Ω \ F )/S1 ; Z(cid:17). And • τ µ ∗(cid:0)ψ2(Z)(cid:1) = 0. • τ µ ∗(cid:0)ψ2(Z6)(cid:1) ⊂ µt(C(Ξ, Z)). • ∀b ∈ H, ∃[u] ∈ K1(cid:0)C(Ω)(cid:1) such that : τ µ ∗ (b) = τ µ ∗ ([u] ⊗C(Ω) [D3]) =(cid:10)Chτ ([u]), [Cµt ](cid:11). To prove the gap-labeling conjecture for pinwheel tilings, it thus remains to study the topological part of the last point of the theorem to prove that τ µ ∗ (cid:0)K0(cid:0)C(Ω) ⋊ R2 ⋊ S1(cid:1)(cid:1) ⊂ µt(cid:0)C(Ξ, Z)(cid:1). This is done in [Mou] and the inclusion in the other direction is easily obtained in this paper too. Aknowledgements. Oyono-Oyono who always supported and advised me during this work. I also want to thank Jean Bellissard for useful discussions on the gap-labeling conjecture. It is a pleasure for me to thank my advisor Herv´e 8 2 Reminders 2.1 Pinwheel tiling and continuous hull A tiling of the plane is a countable family P = {t1, t2 . . .} of non empty compact subsets ti of R2, called tiles (each tile being homeomorphic to the unit ball), such that: ti = E2 where E2 is the euclidean plane with a fixed origin O; • [i∈N • Tiles meet each other only on their border ; • Tiles's interiors are pairwise disjoint. We are interested in the special case where there exists a finite family of tiles {p1, . . . , pn}, called prototiles, such that each tile ti is the image of one of these prototiles under a rigid motion (i.e. a direct isometry of the plane). In fact this paper will focus on the particular tiling called pinwheel tiling or (1,2)-pinwheel tiling which is obtained by a substitution explained below. Our construction of a pinwheel tiling is based on the construction made by Charles Radin in [Rad94]. It's a tiling of the plane obtained by the substitution described in Figure 4. Figure 4: Substitution of the pinwheel tiling. This tiling is constructed from two prototiles, the right triangle in Figure 4.(a) with legs 1, 2 and √5 and its mirror image. To obtain this tiling, we begin from the right triangle with the following vertices in the plane : (0, 0) , (2, 0) and (2, 1). This tile and its reflection are called supertiles of level 0 or 0-supertiles. We will next define 1-supertiles as follows : take the right triangle with vertices 9 (−2, 1), (2,−1) and (3, 1) and take the decomposition of Figure 4.(b). This 1-supertile is thus decomposed in five 0-supertiles, which are isometric copies of the first tile, with the beginning tile in its center (see Figure 5.(b)). Figure 5: Construction of a pinwheel tiling. We next repeat this process by mapping this 1-supertile in a 2-supertile with vertices (−5, 5), (1,−3) and (5, 0) (see Figure 5.(c)). Including this 2-supertile in a 3-supertile with correct orientation and so on, this process leads to the desired pinwheel tiling T . We will now attach to this tiling a topological space reflecting the combinatorial properties of the tiling into topological and dynamical properties of this space. For this, we observe that the direct isometries of the plane are acting on the euclidean plane E2 where we have fixed the origin O. Direct isometries E2 = R2 ⋊ S1 thus act naturally on our tiling T on the right . If Rθ denotes the rotation about the origin with angle θ and s ∈ R2, T .(s, Rθ) := R−θ(T − s). We will also denote (s, Rθ) by (s, θ). Definition 2.1 A patch is a finite union of tiles of a tiling. A tiling T ′ is of finite E2-type or of Finite Local Complexity (FLC) if for any R > 0, there is only a finite number of patches in T ′ of diameter less than R up to direct isometries. A tiling T ′ of finite E2-type is E2-repetitive if, for any patch A in T ′, there is R(A) > 0 such that any ball of radius R(A) intersects T ′ in a patch containing a E2-copy of A. 10 The tiling T is of finite E2-type, E2-repetitive and nonperiodic for translations (see [Pet05]). To attach a topological space to T , we define a metric on T .E2 : If T1 and T2 are two tilings in T .E2, we define A =nε ∈h0, 1√2i / ∃s, s′ ∈ B2 ε (0) , θ, θ′ ∈ B1 ε (0) s.t. (O) = T2.(s′, θ′) ∩ B 1 (O) is the euclidean ball centered in O with radius 1 T1.(s, θ) ∩ B 1 ε ε ǫ where B 1 ǫ(0) are the euclidean balls in Ri centered in 0 and with radius ǫ (i.e. we consider direct isometries near Id). Then, define : ǫ and Bi (O)o d(T1,T2) =(cid:26) Inf A if A 6= ∅ 1√2 else . d is a bounded metric on T .E2. For this topology, a base of neighborhoods is defined by: two tilings T1 and T2 are close if, up to a small direct isometry, they coincide on a large ball around the origin. The topology defined here works because the tilings considered are of finite E2- type. There exist other topologies (equivalent) that one can put on this space. The topology thus obtained is metrizable but none of the metrics that can be defined to produce the topology is canonical. A more canonical way to define the topology was given in [BHZ00]. The problem of non-uniqueness of the metric has been investigated in [PB09]. Definition 2.2 The continuous hull of T is then the completion of (T .E2, d) and will be denoted Ω(T ). Let's enumerate some well known properties of this continuous hull: Property 2.3 ([BBG06], [BHZ00], [BG03], [KP00], [LP03], [Rad94]) • Ω(T ) is formed by finite E2-type, E2-repetitive and nonperiodic (for trans- lations) tilings and each tiling of Ω(T ) has the same patches as T . • Ω(T ) is a compact space since T is of finite E2-type. • Each tiling in Ω(T ) are uniquely tiled by n-supertiles, for all n ∈ N. • The dynamical system (Ω(T ), E2) is minimal since T is repetitive, i.e each orbit under direct isometries is dense in Ω(T ). The last property of Ω(T ) allows us to write Ω without mentioning the tiling T (in fact, if T ′ ∈ Ω(T ), Ω(T ′) = Ω(T )). Definition 2.4 Any tiling in Ω is called a pinwheel tiling. Remark : we can easily see that our continuous hull is the compact space Xφ defined by Radin and Sadun in [RS98]. 11 2.2 The canonical transversal In this section, we will construct a Cantor transversal for the action of E2 and we show that this transversal gives the local structure of a foliated space. For this, we fix a point in the interior of the two prototiles of the pinwheel tiling. This, in fact, gives for any tiling T1 in Ω (i.e constructed by these two prototiles), a punctuation of the plane denoted T punct Define then Ω0 to be the set of every tilings T1 of Ω such that O ∈ T punct The canonical transversal is the space Ω0/S1. 1 . 1 . We can identify this space with a subspace of Ω by constructing a continuous section s : Ω0/S1 −→ Ω. To obtain such a section, we fix an orientation of the two prototiles of our tilings once for all. Hence when we consider a patch of a tiling in the transversal Ω0, there is only one orientation of this patch where the tile containing the origin have the orientation chosen for the prototiles. Let's then [ω] be in Ω0/S1, there is only one θ ∈ [0; 2π[ such that the tile in Rθ(ω) containing the origin has the good orientation. We define s([ω]) := Rθ(ω). s is well defined because θ depends on the representative ω chosen but not Rθ(ω). s : Ω0/S1 −→ s(Ω0/S1) is then a bijection. We easily see that s is continuous and thus it is a homeomorphism from the canonical transversal onto a compact subspace Ξ of Ω. We also call this space the canonical transversal. We can see Ξ as the set of all the tilings T1 in Ω with the origin on the punctu- ation of T1 and with the tile containing the origin in the orientation chosen for the prototiles. We then have : Proposition 2.5 ([BG03]) The canonical transversal is a Cantor space. A base of neighborhoods is obtained as follows : consider T ′ ∈ Ξ and A a patch around the origin in T ′ then U (T ′,A) = {T1 ∈ Ξ T1 = T ′ on A} is a closed and open set in Ξ, called a clopen set. Before defining the foliated stucture on Ω, we must study the rotations which can fix tilings in Ω. In pinwheel tilings, we can sometimes find regions tiled by supertiles of any level and so we introduce the following definition: Definition 2.6 A region of a tiling which is tiled by n-supertiles for all n ∈ N is called an infinite supertile or supertile of infinite level. 12 If a ball in a tiling T1 fails to lie in any supertile of any level n, then T1 is tiled by two or more supertiles of infinite level, with the offending ball straddling a boundary. We can, in fact, construct a pinwheel tiling with two half-planes as infinite supertiles as follows: Consider the rectangle consisting of two (n − 1)-supertiles in the middle of a n-supertile. For each n > 1, orient this rectangle with its center at the origin and its diagonal on the x-axis, and fill out the rest of a (non-pinwheel) tiling Tn by periodic extension. By compactness this sequence has a convergent subsequence, which will be a pinwheel tiling and which will consist of two infinite supertiles (this example comes from [RS98]). Note that the boundary of an infinite supertile must be either a line, or have a single vertex, since it is tiled by supertiles of all levels. We call such a line a fault line. Lemma 2.7 If (s, θ) fixes a pinwheel tiling T ′ then θ ∈ {0, π}mod(2π). Moreover, if θ = 0 then s = 0. In other terms, translations can't fix a pinwheel tiling. Proof : Let's consider the different cases: 1. First, if the tiling T ′ which is fixed by (s, θ) have no fault line (i.e have no infinite supertile), then s = 0 and θ = 0(mod2π). Indeed, let's x be in E2 be such that −→Ox = s then O and x is in the interior of a m-supertile since there isn't infinite supertiles (see p.29 in [RS98]). Since no direct isometry fixes our prototiles, s and θ must be zero. 2. Let's see the case in which T ′ have some infinite supertiles. By [RS98] p.30, the number of infinite supertiles in T ′ is bounded by a constant K (in fact for pinwheel tilings, we can take K = 2π α where α is the smallest angle in the prototiles). Thus, T ′ doesn't contain more than K infinite supertiles and in fact, it has only a finite number of fault lines. This will give us the result. Indeed, since (s, θ) fixes T ′, if F is a fault line, (s, θ) sends it on another fault line F1 in T ′ and thus, Rθ sends F on a line parallel to F1. As there is only a finite number of fault lines in T ′, there is M ∈ N∗, m ∈ Z∗ such that M θ = 2πm. If we now use results obtained in [RS98] p.32, θ must be in the group of relative orientations GRO(P in) of pinwheel tilings which is the subgroup of S1 generated by π Hence, if θ = 2kα + l π 2 with k ∈ Z∗ and l ∈ Z, it would mean that α is rationnal with respect to π, which is impossible (see [Rad94] p.664) 2 and 2α. hence k = 0 and θ ∈(cid:26)0, π 2 , π, 3π 2 (cid:27) mod(2π). Now, if we study the first vertex coronas (i.e the minimal patches around a vertex or around the middle point of the hypothenuses), there is only patches with a 2-fold symmetry ([Sad] or see Figure 6 and Figure 7 p.48 and p.49). Thus, θ ∈ {0, π} mod(2π) and if θ = 0, s = 0 since pinwheel tilings are not fixed by translations. (cid:3) 13 We note that, in fact, there exists only 6 pinwheel tilings with a 2-fold symmetry up to rotations ([Sad]). Hence, there is only 6 orbits with fixed points for the R2 ⋊ S1 action on Ω. Moreover, there is exactly 6 circles F1, . . . , F6 containing fixed points for the S1 action on Ω (of course, therefore, the 6 orbits of these circles contain all the fixed points of the R2 ⋊ S1-action). We thus obtain the following important result on the dynamic of our tiling space: Theorem 2.8 ([BG03]) The continuous hull is a minimal foliated space. Proof : The proof follows the one in [BG03] except that, locally, Ω looks like an open subset of S1 × an open subset of R2 × a Cantor set instead of S1 × an open subset of R2 × a Cantor set, like in [BG03]. Ω is covered by a finite number of open sets Ui = φi(Vi × Ti) where : • Ti is a clopen set in Ξ; • Vi is an open subset of R2 ⋊ S1 which read Vi = Γi × Wi where Wi is an open subset of R2 and Γi an open subset of S1 of the form ]lπ/2 − π/3; lπ/2 + π/3[, l ∈ {0, 1, 2, 3}; • φi : Vi × Ti −→ Ω is defined by φi(v, ω0) = ω0.v. As we can find finite partitions of Ξ in clopen sets with arbitrarily small diameter, it is possible to choose this diameter small enough so that: • the maps φi are homeomorphisms on their images; • whenever T1 ∈ Ui ∩ Uj, T1 = φi(v, ω0) = φj(v′, ω′0), the element v′.v−1 is independent of the choice of T1 in Ui ∩ Uj, we denote it by gij. The transition maps read : (v′, ω′0) = (gij.v, ω0.g−1 ij ). It follows that the boxes Ui and charts hi = φ−1 : Ui −→ Vi × Ti define a foliated structure on Ω. By construction, the leaves of Ω are the orbits of Ω under the E2-action. i (cid:3) We must do several remarks on the actions. E2 isn't acting freely on Ω, even if the translations are, but we could adapt results of Benedetti and Gambaudo obtained in their paper [BG03] studying the possible symmetries in our pinwheel tilings. The E2-action is not free on Ω0 too but the S1-action is. Using the group of relative orientations GRO(P in), we can see that each R2-orbit of Ω is in fact a dense subset of Ω (see [HRS05]). 14 2.3 C ∗-algebra associated to a tiling As in the paper of Kellendonk and Putnam, one can naturally associate a C∗- algebra to pinwheel tilings. It is the crossed product C∗-algebra A = C(Ω) ⋊ (R2 ⋊ S1) obtained from the dynamical system (Ω, R2 ⋊ S1). For the sake of the next session, let's consider a topological space Ω endowed with a right action of a locally compact group G (for pinwheel tilings, G = R2 ⋊ S1). For simplicity, let's suppose that G is unimodular and let's fix a Haar measure λ on G. In order to construct C(Ω) ⋊ G, let's first consider the vector space Cc (Ω × G) of continuous functions with compact support on Ω× G. One endows this space with a convolution product and an involution : f ∗ g(ω, g) :=ZG f (ω, h)g(ω.h, h−1g)dλ(h) f∗(ω, g) := f (ω.g, g−1) with f, g ∈ Cc(Ω × G) and ω ∈ Ω, g ∈ G. One then defines a norm on it : k f k∞,1= M ax(cid:26)sup ω∈ΩZG f (ω, g) dλ(g), sup ω∈ΩZG f∗(ω, g) dλ(g)(cid:27) Define the involutive algebra L∞,1(Ω× G) as the completion of Cc(Ω× G) under this norm. This algebra is represented on L2(G) by the family {πω, ω ∈ Ω} of representa- tions given by : πω(f )ψ(g) :=ZG f (ω.g, g−1.h)ψ(h)dλ(h), ψ ∈ L2(G). Namely, πω is linear, πω(f g) = πω(f )πω(g), πω(f )∗ = πω(f∗) and πω is boun- ded, k πω(f )k6k f k∞,1. Set k f k= supω∈Ω k πω(f ) k which defines a C∗-norm on L∞,1(Ω × G) and permits to define C(Ω) ⋊ G as the completion of Cc(Ω × G) or of L∞,1(Ω × G) under this norm. For pinwheel tilings, G = R2 ⋊ S1 and this C∗-algebra C(Ω) ⋊ (R2 ⋊ S1) is isomorphic to (C(Ω) ⋊ R2) ⋊ S1 ([Cha99] p.15). S1 acts on Cc(Ω× R2) by : θ.f (w, s) := f (R−θ(w), R−θ(s)) for f ∈ Cc(Ω× R2). This action extends to the crossed product C(Ω) ⋊ R2. This algebra is interesting for two reasons. First, it contains dynamical in- formations related to the combinatorial properties of the tiling. Secondly, the gap-labeling conjecture made by J. Bellissard in [Bel92] links the electronic motion in a quasicrystal to the K0-group of the C∗-algebra of the tiling. This is the subject of the next session. A natural question is to know if, in the case of pinwheel tilings, we can relate the K-theory of C(Ω) ⋊ R2 ⋊ S1 to the Z-module of patch frequencies of the pinwheel tiling (see the end of next session for motivations). 15 3 Gap-labeling 3.1 Quasicrystals Before beginning the presentation of the works of Bellissard on the gap-labeling conjecture, we introduce the definition of aperiodic tilings which model qua- sicrystals. It is a tiling of the space Rn with the same combinatorial properties as pinwheel tilings. Definition 3.1 A tiling of Rn is a countable family P = {t0, t1, . . .} of non empty compact sets ti, called tiles (each tile is suppoed to be homeomorphic to the unit ball), such that : ti = En where En is the Euclidean space Rn endowed with a fixed origin O; • [i∈N • Tiles meet only on their border; • The interior of two tiles are disjoint. Let G be a subgroup of the group of rigid motions of En. We consider in the sequel the particular case where there exists a finite family {p0, . . . , pn} such that each compact set ti is a G-copy of some pk. The pk's are then called prototiles. A tiling is of finite G-type if, for any R > 0, there are only finitely many patches of diameter less than or equal to R upto the G-action. A tiling T ′ is G-repetitive if, for each patch A in T ′, there exists R(A) > 0 such that every ball of radius R(A) in Rn meets T ′ on a patch containing a G-copy of A. A G-aperiodic tiling is a tiling of Rn of finite G-type, G-repetitive and non- periodic with respect to translations of Rn. For example, pinwheel tilings are aperiodic tilings with n = 2 and G = R2 ⋊ S1, the group of direct isometries of the plane. Many more examples are known, the most famous being the one constructed by Penrose (n = 2, G = R2) (see [Pet05]). In this section on the gap-labeling, we fix a Rn-aperiodic tiling T . One can then associate to such a tiling a topological space Ω as we have done for pinwheel tilings. Let T .Rn be the space of all the translations of T . Ω is then the completion of this space for the metric d defined as follows : A =nε ∈h0, 1√2i / ∃s, s′ ∈ B2 ε (0) , θ, θ′ ∈ B1 ε (0) s.t. T1.(s, θ) ∩ B 1 ε (O) = T2.(s′, θ′) ∩ B 1 ε (O)o (O) is the ball of En centered at O with radius 1 ε (0) is the where B 1 ball of Rn centered at 0 with radius ε (i.e one only considers translations near Id) . ε and Bn ε 16 d is then defined by : else . 1√2 d(T1,T2) =(cid:26) Inf A if A 6= ∅ Then Ω(T ) := T .Rn is the completion of T .Rn under d. One can prove that Ω(T ) is a compact metric space, that every Rn-orbits are dense in this space and thus, all the tilings of Ω(T ) have the same completion. One can then define, as in the first section, the canonical transversal by fixing a point in each prototile and letting the canonical transversal Ξ be the set of all T ′ ∈ Ω with one point on the origin O. This space is still a Cantor set (see [BG03]) and Ω is a foliated space with leaves homeomorphic to Rn. 3.2 Integrated density of states - IDS This section summarizes the works of Bellissard [Bel92] (see also [Ypm]). The gap-labeling conjecture describes qualitatively the spectrum of the opera- tor associated to the electronic motion in a quasicrystal (which is a solid with a particular atomic distribution, modeled by aperiodic tilings). This motion is described by the Schrodinger operator on the Hilbert space L2(Rn) : H = − ℏ 2m ∆ + V (x) h 2π where h is the Planck constant, m where ℏ is the Dirac constant, equals to is the electron mass, ∆ the Laplacian on Rn and V is a potential in L∞(Rn) depending on the atomic distribution of the quasicrystal. The domain of H is D(H) = {ψ ∈ L2(Rn) ∆ψ ∈ L2(Rn)}. Let fix a Rn-aperiodic tiling T which is a model for the quasicrystal. One can then consider the two spaces, introduced in the last section, Ω and Ξ together with the C∗-algebra, associated to the dynamical system, C(Ω) ⋊ Rn called the C∗-algebra of observables. Definition 3.2 A covariant family of selfadjoint operators {Hω; ω ∈ Ω} sat- isfies Ua.Hω.U∗a = Hω−a for all x ∈ Rn and Ua : L2(Rn) → L2(Rn) defined by Ua(f )(x) := f (x + a). Such a family is affiliated to the C∗-algebra A = C(Ω) ⋊ Rn if, for every f ∈ C0(R), the bounded operator f (Hω) can be represented as πω(hf ) for some hf ∈ A such that the map h : C0(R) → A : f 7→ hf is a bounded *- homomorphism. In the sequel, we fix such a covariant family of selfadjoint operators {Hω; ω ∈ Ω} affiliated to C(Ω) ⋊ Rn (see Bellissard's papers [Bel92] and [BHZ00] for the existence of such family). We are then interested in the spectrum of Hω. 17 For this, Bellissard linked a physical function defined on R (the Integrated Den- sity of States IDS) to a map obtained from C(Ω) ⋊ Rn. Let's define the IDS N (E) which can be seen as the number of eigenvalues per unit volume up to E. Definition 3.3 Let G be a locally compact group. A F∅lner sequence is a sequence (Λn) of open subsets of G, each with finite Haar measure Λn , such that G = ∪Λn and such that for all x ∈ G, lim n→∞ Λn∆x.Λn Λn = 0 where V ∆W = (V ∪ W ) \ (V ∩ W ). It can be shown that there exists a F∅lner sequence in G if and only if G is amenable (see [Gre69]). One can now define Nω,Λn (E) := #{E′ ∈ Sp(Hω,Λn) E′ 6 E}, where Hω,Λn is the Hamiltonian Hω restricted to Λn, acting on the Hilbert space L2(Λn), subject to certain boundary conditions (see [Bel92]). The IDS Nω : R −→ R+ is then defined by 1 Nω(E) := lim n→∞ Λn Nω,Λn (E). Jean Bellissard proved that this limit exists and is independent of the chosen boundary conditions ([Bel92]). 3.3 Shubin's formula The link between the IDS and the C∗-algebra C(Ω) ⋊ Rn is given by the Shu- bin's formula. On a Hilbert space H with an orthonormal basis {ei}, one can define the oper- ator trace of a bounded operator A by : T r(A) := hei, Aeii. ∞ Xi=1 This trace is independent of the choice of a basis in H if A is trace-class, i.e if T r( A ) < ∞ where A := √A∗A. One then has : Nω(E) = lim n→∞ 1 Λn T rΛn(cid:0)χ]−∞;E](Hω,Λn )(cid:1) where T rΛn is the restriction to L2(Λn) of the operator trace on L2(Rn) and χ]−∞;E] is the characteristic function of ] − ∞; E]. To link this formula to the C∗-algebra of observables, one defines an ergodic invariant measure on Ω. This is given by the next proposition which is a conse- quence of the amenability of Rn and of the Krein-Milman theorem : 18 Proposition 3.4 There exists a translation invariant, ergodic probability mea- sure µ on Ω. A trace is then densely defined on C(Ω) ⋊ Rn by : ∀ f ∈ Cc(Ω × Rn), τ µ(f ) :=Z f (ω, 0)dµ(ω). Since µ is translation invariant, τ µ has the properties of a positive trace i.e τ µ(f ∗ f∗) > 0 and τ µ(f ∗ g) = τ µ(g ∗ f ). This functional can be extended in a faithful trace (which follows from the ergodicity of µ and the minimality of the action of Rn on Ω) and semi-finite (see [MS06] p150 and following) on the Von Neumann algebra of C(Ω) ⋊ Rn. Moreover, this trace is finite on projections of the Von Neumann algebra (see [MS06] p154 and [Ped79] 5.6.7). The link between the IDS and this trace is highlighted by the following Birkhoff theorem : Theorem 3.5 Let Ω be a compact space with a probability measure µ that is ergodic and invariant under the action T of an amenable group G on Ω. Then the left Haar measure λ on G can be normalized in such a way that for all f ∈ C(Ω), for µ-almost all ω ∈ Ω, we have : Λn ZΛn f (ω′)dµ(ω′) = lim n→∞ where (Λn) is a F∅lner sequence in G. ZΩ 1 f (Tgω)dλ(g), Applying this theorem, we get for all f ∈ Cc(Ω × Rn) and µ-almost all ω ∈ Ω : τ µ(f ) = lim n→∞ 1 Λn ZΛn f (ω.g, 0)dλ(g). By definition of πω and T r, the formula of the theorem becomes : τ µ(f ) = lim n→∞ 1 Λn T rΛn (πω(f )), for µ − almost all ω. (1) Note that, if E ∈ R, then χ]−∞;E](Hω) = χH6E for some χH6E in the Von Neumann algebra of the C∗-algebra of observables. Moreover, if E ∈ g for some gap g in the spectrum of Hω (i.e a connected component of the complement of the spectrum) then χ]−∞;E] is a bounded con- tinuous map on Sp(Hω) and thus χ]−∞;E](Hω) can be represented as πω(χH6E) for some element χH6E ∈ C(Ω) ⋊ Rn, since Hω is affiliated to this C∗-algebra. There is then a link between a physically measurable function by experiments and our C∗-algebra : Definition 3.6 A covariant family (Hω) affiliated to C(Ω)⋊Rn is said to satisfy the Shubin's formula if for µ-almost all ω ∈ Ω, we have: The common value is denoted by N (E). Nω(E) = τ µ(χH6E). 19 3.4 Gap-labeling In this section, we fix a covariant family of selfadjoint operators (Hω) affiliated to C(Ω) ⋊ Rn and satisfying the Shubin's formula (see [Bel92] for the existence of such a family). As we said, if E /∈ Sp(Hω), χ]−∞;E](Hω) is a projection of the C∗-algebra. Moreover, if g is a gap in Sp(Hω) then for any two values E, E′ ∈ g, we have χ]−∞;E](Hω) = χ]−∞;E′](Hω). Thus, there is a labeling of the gaps g by projec- tions P (g) of C(Ω) ⋊ Rn : P (g) := χ]−∞;E](Hω) for any E ∈ g. Now, an important property of the trace τ µ is its invariance under unitary transformations which implies that it induces a trace on the set of equivalence classes of projections under unitary transformations. Thus, τ µ induces a linear map τ µ We then have another form for the Shubin's formula on gaps of Sp(Hω) : ∗ on K0(C(Ω) ⋊ Rn). Using the fact that C(Ω) ⋊ Rn is separable, we can state : N (g) = τ µ ∗ [P (g)]. Proposition 3.7 On gaps of Sp(Hω), the Integrated Density of States takes its values in τ µ ∗ (K0(C(Ω) ⋊ Rn)), which is a countable subset of R. The goal is then to compute the image under τ µ ∗ of the K-theory of C(Ω) ⋊ Rn. For Rn-aperiodic tilings, this image was conjectured in [BHZ00] in 2000 and was proven independently by Bellissard, Benedetti and Gambaudo [BBG06] on one hand, Benameur and Oyono-Oyono [BOO02] on the other and finally by Kaminker and Putnam [KP03]: Theorem 3.8 Let Ω be the continuous hull of a Rn-aperiodic tiling with a to- tally disconnected canonical transversal Ξ. Let µ be a translation invariant, ergodic probability measure and µt the induced measure on Ξ. We then have : where C(Ξ, Z) is the set of continuous functions on Ξ with integer values and τ µ ∗(cid:16)K0(cid:0)C(Ω) ⋊ Rn(cid:1)(cid:17) = µt(cid:0)C(Ξ, Z)(cid:1) µt(cid:0)C(Ξ, Z)(cid:1) :=(cid:26)ZΞ f dµt, f ∈ C(Ξ, Z)(cid:27) . A natural question is whether this theorem remains true for pinwheel tilings i.e : Theorem 3.9 Let Ω be the continuous hull of a pinwheel tiling, µ a R2 ⋊ S1- invariant, ergodic probability measure and µt the induced measure on Ξ. We have : τ µ ∗(cid:16)K0(cid:0)C(Ω) ⋊ R2 ⋊ S1(cid:1)(cid:17) = µt(cid:0)C(Ξ, Z)(cid:1). 20 In other words, is there still this link between the image of the K-theory of the C∗-algebra of pinwheel tilings under the linear map induced by the trace and the Z-module of patch frequencies? In this paper, we make a first step in the direction of this theorem. More precisely, we link the image of the K-theory to the tangential cohomology group of the continuous hull (see [MS06]). This cohomological part is more computable as could be seen in [Mou] and the gap-labeling is proved in this paper, showing moreover that the module of patch frequencies is given explicitly by 1 To make this link between K-theory and cohomology, we follow the ideas of [BOO02]. Let's remind the important steps of their proof. In their article, Benameur and Oyono-Oyono used the index theorem for foliations established by Alain Connes in [Con79] and more precisely, the version for foliated spaces proved in [MS06]. They then proceeded in several steps. 264 Z(cid:2) 1 5(cid:3). Ξ,Rn ) where ∂e First, the inclusion µt(C(Ξ, Z)) ⊂ τ µt ∗ (K0(C(Ξ) ⋊ Zn)) is easy and thus they only have to prove the inclusion in the other side. Then, they notice that it is enough to prove it for even integers n. Next, they prove that the K-theory of C(Ω) is isomorphic to the one of C(Ξ)⋊Zn and that this isomorphism is given by the map e 7→ IndΩ(∂e Ξ,Rn is a Dirac operator with coefficients in the fiber bundle associated to e and IndΩ is the analytical index. The second crucial point of the proof is the fact that the top dimensional tan- gential cohomology group is isomorphic, by a map ΨZn , to the integer group of coinvariants C(Ξ, R)Zn of C(Ξ, R) under the action of Zn. The last step is to link τ µt ∗ (cid:16)K0(cid:0)C(Ξ) ⋊ Zn(cid:1)(cid:17) to this cohomology group using : Ξ)) = hchn l ([e]), [CZn,µ]i where CZn,µ is some current on Ω and chn l ([e]) is an element of the top co- homology group (it is the image of [e] under the component of degree n of the tangential Chern character). Theorem 3.10 τ µ ∗ (IndΩ(∂e Finally, they prove that the image of chn ΨZn(chn l ([e])) ⊂ C(Ξ, Z)Zn . l ([e]) under ΨZn is integer valued i.e 4 Index theorem for the gap-labeling of the pin- wheel tiling In a first step, we study the K-theory of the C∗-algebra associated to the dy- namical system C(Ω) ⋊ R2 ⋊ S1 and then we compute its image under the linear map τ µ ∗ . In the sequel, one identifies the class of an unbounded triple defined in [BJ83] with the class that it defines in KK-theory. The reader can refer to [Ska91] for definitions and properties of equivariant KK-theory groups of Kasparov that will be used in this section. 21 4.1 Study of K0(cid:0)C(Ω) ⋊ R2 ⋊ S 1) To compute this K-theory, we will proceed in two steps. The first one consists in using the Dirac-Dual Dirac construction. For this, let's consider the Dirac operator ∂2 on R2. Then, F = ∂2(1 + ∂2 an elliptic pseudodifferential operator of order 0 on H = L2(R2, C ⊕ C). Letting C0(R2) act on H by multiplication f 7→ M (f ), one has a Kasparov cycle (H, M, F ). Moreover, since G = R2 ⋊ S1 acts naturally on the left of C0(R2) and of H, and since F is then G-invariant, the class of (H, M, F ) defines an element αG of KK G 0 (C0(R2), C), called the fundamental element. 2 )− 1 2 is There exists an element σG ∈ KK G 0 (C, C0(R2)) such that the Kasparov product of αG with σG over C is αG ⊗C σG = 1C0(R2) i.e σG is a right inverse for αG (see [Kas95], [Kas88], [Ska91]). Furthermore, since G = R2 ⋊ S1 is amenable, σG ⊗C0(R2) αG = 1C. αG is thus an invertible element in KK G 0 (C0(R2), C) and we can prove : Proposition 4.1 K0 (cid:0)C(Ω) ⋊ R2 ⋊ S1(cid:1) is isomorphic to K0 (cid:0)C(Ω) ⋊ S1(cid:1). Proof : JG : KK G on this element, we obtain the element Denote [∂2] the class of the above cycle (H, M, F ). We have τΩ[∂2] ∈ KK G Using the descent homomorphism 0 (cid:0)C0(Ω × R2), C(Ω)(cid:1) where τΩ[∂2] := 1Ω ⊗C [∂2]. 0 (cid:0)C0(Ω × R2), C(Ω)(cid:1) −→ KK0(cid:0)C0(Ω × R2) ⋊ G, C(Ω) ⋊ G(cid:1) , JG (τΩ ([∂2])) ∈ KK0(cid:0)C0(Ω × R2) ⋊ G, C(Ω) ⋊ G(cid:1) . Since the homomorphisms τΩ and JG are compatible with Kasparov prod- uct and unit elements, JG (τΩ ([∂2])) is an invertible element and thus defines an isomorphism β := ⊗C0(Ω×R2)⋊G JG (τΩ ([∂2])) : β : KK0(cid:0)C, C0(Ω × R2) ⋊ G(cid:1) −→ KK0 (C, C(Ω) ⋊ G) . Moreover, since C(Ω) ⋊ S1 is Morita equivalent to C0(Ω × R2) ⋊ R2 ⋊ S1 (see lemma 4.2), we have an isomorphism δ : KK0(cid:0)C, C(Ω) ⋊ S1(cid:1) −→ KK0(cid:0)C, C0(Ω × R2) ⋊ G(cid:1) . The desired isomorphism of the proposition is then obtained as the com- position βoδ. (cid:3) We recall a result that we used in the proof of the proposition and that we will often use in the sequel : 22 Lemma 4.2 Let G be a locally compact group and X a locally compact space with a right G-action. Let K(cid:0)L2(G)(cid:1) be the compact operators on L2(G). Then the C∗-algebras K(cid:0)L2(G)(cid:1) ⊗ C0(X) and C0(X × G) ⋊ G are isomorphic. Moreover, for G = R2, this isomorphism is S1-equivariant and C0(X) ⋊ S1 is Morita equivalent to C0(X × R2) ⋊ R2 ⋊ S1. Thus, we proved that, to study K0(cid:0)C(Ω) ⋊ R2 ⋊ S1(cid:1), of angle π around the origin and F := S Fi, to obtain the following six term K0(C(Ω) ⋊ S1). To investigate this group, we use the 6 circles F1, ... , F6 stable under a rotation it suffices to study exact sequence : K S1 0 (C0(Ω \ F )) / K S1 0 (C(Ω)) / K S1 0 (C(F )) K S1 1 (C(F )) K S1 1 (C(Ω)) ∂ K S1 1 (C0(Ω \ F )) i (C(F )) ≃ Ki(cid:0)C(F ) ⋊ S1(cid:1) ≃ Ki(cid:16)(C ⊕ C)6(cid:17), the first isomorphism is But, K S1 proved in [Jul81] and the second comes from the isomorphism of C(Fi)⋊S1 with IndS1 {−1,1}(C) ⋊ S1 and from the fact that this C∗-algebra is Morita equivalent to C∗(Z/2Z) ≃ C ⊕ C, see [Seg68] (the S1-action on Fi is obtained as the S1-action on S1 defined by eiθ.z := e2iθz and IndS1 {−1,1}(C) is the induced C∗- algebra defined as the space of continuous functions f on S1 with values in C such that f (−z) = −f (z)). Thus, K S1 Note that K S1 0 (C(Fi)) ≃ Z2 is generated by two S1-equivariant vector bundles. The first one is the trivial fiber bundle Fi × C with the diagonal action of S1 where S1 acts trivially on C and the second one is the fiber bundle Fi × C with the diagonal action of S1 where S1 acts by multiplication on C. Furthermore, S1 acts freely and properly on Ω\ F since all the fixed points have been removed. Thus, 0 (C(F )) ≃ Z12 (= Z6 ⊕ Z6) and K S1 1 (C(F )) ≃ 0. K S1 i (C0(Ω \ F )) ≃ Ki(C0(Ω \ F ) ⋊ S1) ≃ Ki(cid:0)C0(cid:0)(Ω \ F )/S1(cid:1)(cid:1) , where the last isomorphism is obtained by Morita equivalence (see [Rie82]). The six term exact sequence becomes : K0(C0(Ω \ F )/S1)) / K S1 0 (C(Ω)) Z12 ∂ 0 K S1 1 (C(Ω)) K1(C0((Ω \ F )/S1)) To study the connecting map ∂, we need the following lemma. It links the K-theory group of a topological space of low dimension with its cohomology group (see [Mat], section 3.4) : 23 / /   O O o o o o / / /   O O o o o o Lemma 4.3 Let X be a connected finite CW -complex of dimension 6 3. Then, there exist canonical isomorphisms chZ ev := chZ chZ odd := chZ 0 ⊕ chZ 1 ⊕ chZ 2 : K0(cid:0)C(X)(cid:1) −→ H 0(X; Z) ⊕ H 2(X; Z) 3 : K1(cid:0)C(X)(cid:1) −→ H 1(X; Z) ⊕ H 3(X; Z) that are natural for such complexes and compatible with the usual Chern char- acter i.e such that the following diagram commutes Kj(cid:0)C(X)(cid:1) chZ n H n(X; Z) chn )RRRRRRRRRRRRRR / H n(X; Q) where j = n mod 2, H n(X; Z) → H n(X; Q) is the canonical homomorphism induced by the inclusion Z ֒→ Q and chn is the component of degree n of the Chern character. Since Ω/S1 is the inverse limit of connected finite CW -complexes of dimension 2 (see [ORS02] or [Mou]), we obtain the following lemma: Lemma 4.4 The Cech cohomology groups with compact support and with inte- ger coefficients H k Moreover, c ((Ω \ F )/S1; Z) vanish for k > 3. K0(cid:16)C0(cid:16)(Ω \ F )/S1(cid:17)(cid:17) ≃ H 2 K1(cid:16)C0(cid:16)(Ω \ F )/S1(cid:17)(cid:17) ≃ H 1 c(cid:16)(Ω \ F )/S1 ; Z(cid:17) c(cid:16)(Ω \ F )/S1 ; Z(cid:17). and Proof : In fact, according to [ORS02] or [Mou], Ω/S1 is the inverse limit of CW -complexes of dimension 2. Thus, H k(Ω/S1 ; Z) = 0 for k > 3. The long exact sequence of relative cohomology groups associated to the pair (Ω/S1, F/S1) then gives : . . . / H 2(Ω/S1, F/S1; Z) / H 2(Ω/S1; Z) / H 2(F/S1; Z) / H 3(Ω/S1, F/S1; Z) / H 3(Ω/S1; Z) / H 3(F/S1; Z) / . . . Since F/S1 is composed by 6 points, H k(F/S1; Z) = 0 for k > 1 and thus, H k+1(Ω/S1, F/S1; Z) ≃ H k+1(Ω/S1; Z) for k > 1. From the result reminded above on cohomology groups of Ω/S1, we proved that the relative cohomology groups associated to (Ω/S1, F/S1) with in- teger coefficients vanish for degrees greater than 3. To conclude, we use lemma 11 p.321 from [Spa66] to state that, for any 24   ) / / / / / / / / k, H k(Ω/S1, F/S1; Z) ≃ H k Moreover, since F is stable for the S1-action, (Ω/S1\ F/S1) = (Ω\ F )/S1, which completes the proof of the first point of the lemma. c(cid:16)(cid:0)Ω/S1(cid:1) \(cid:0)F/S1(cid:1); Z(cid:17). Furthermore, we have : and K0(cid:16)C0(cid:16)(Ω \ F )/S1(cid:17)(cid:17) = K0(cid:16)C(cid:16)(cid:0)(Ω \ F )/S1(cid:1)+(cid:17)(cid:17) K1(cid:16)C0(cid:16)(Ω \ F )/S1(cid:17)(cid:17) = K1(cid:16)C(cid:16)(cid:0)(Ω \ F )/S1(cid:1)+(cid:17)(cid:17) and K0 is the reduced K-theory. is the Alexandroff compactification of (Ω \ F )/S1 where (cid:0)(Ω \ F )/S1(cid:1)+ Using results from [Mou], we can easily prove that (cid:0)(Ω \ F )/S1(cid:1)+ inverse limit of CW -complexes of dimension 2. Thus, applying the results of lemma 4.3, we have proved the second point of the lemma. is the We can then compute the kernel of ∂ : (cid:3) Lemma 4.5 Ker ∂ = Z.qi where 7 Li=1 q1 = (1, 1, 1, 1, 1, 1, 0, 0, 0, 0, 0, 0) q2 = (0, 1, 1, 1, 1, 1, 1, 0, 0, 0, 0, 0) q3 = (1, 0, 1, 1, 1, 1, 0, 1, 0, 0, 0, 0) q4 = (1, 1, 0, 1, 1, 1, 0, 0, 1, 0, 0, 0) q5 = (1, 1, 1, 0, 1, 1, 0, 0, 0, 1, 0, 0) q6 = (1, 1, 1, 1, 0, 1, 0, 0, 0, 0, 1, 0) q7 = (1, 1, 1, 1, 1, 0, 0, 0, 0, 0, 0, 1). The proof is given in Appendix A. Since q1 lifts on the constant projection equal to 1 on Ω, we thus have proved : Proposition 4.6 We have K0(C(Ω) ⋊ R2 ⋊ S1) ≃ K0(C(Ω) ⋊ S1) ≃ Z⊕ 7 Mi=2 Z.qi!⊕ H 2 c(cid:16)(Ω\ F )/S1 ; Z(cid:17), where Z is generated by the constant projection equal to 1. 25 4.2 Computation of the image under τ µ ∗ of the summand Z ∗ of β◦δ(Z) in K0(C(Ω)⋊R2 ⋊S1) In this section, we compute the image under τ µ where β ◦ δ is the isomorphism contructed in the proof of proposition 4.1. To compute it, we will consider the maps φ : C → C(Ω) and φA : A → C(Ω)⊗ A (A a C∗-algebra with a R2 ⋊ S1-action) given by φ(z)(ω) := z for any ω ∈ Ω and φA := φ⊗Id. We will also use the induced map φR2 A : A⋊R2 →(cid:0)C(Ω)⊗A(cid:1)⋊R2. The maps φ and φA are R2 ⋊ S1-equivariant and thus define, by functoriality of KK-theory, the following maps : 0 (C, C(Ω)), φ∗ : KK S1 (φA)∗ : KK S1 0 (C, C) −→ KK S1 0 (C, A) −→ KK S1 0 (C, A ⋊ R2) −→ KK S1 0 (C(Ω) ⊗ A, C(Ω)) −→ KK S1 A )∗ : KK S1 φ∗A : KK S1 0 (C, C(Ω) ⊗ A), 0 (cid:16)C,(cid:0)C(Ω) ⊗ A(cid:1) ⋊ R2(cid:17), 0 (A, C(Ω)). (φR2 We will denote φ∗ and φ∗ these homomorphisms when no confusion would be possible or if A = C. Proposition 4.7 We have the following commutative diagram : δ1 K S1 K S1 0 (C) η1 K S1 Ψ∗ 0 (cid:0)K(L2(R2))(cid:1) 0 (cid:0)C0(R2) ⋊ R2(cid:1) 0 (cid:0)C∗(R2)(cid:1) K S1 β1 φ∗ (cid:0)φK(cid:1)∗ (cid:0)φR2 C0 (R2 )(cid:1)∗ (cid:0)φR2 C (cid:1)∗ K S1 0 (C(Ω)) η δ K S1 K S1 Ψ∗ 0 (cid:0)C(Ω) ⊗ K(L2(R2))(cid:1) 0 (cid:0)C0(Ω × R2) ⋊ R2(cid:1) 0 (cid:0)C(Ω) ⋊ R2(cid:1) / K S1 β where the isomorphisms δ1 and β1 are constructed in a similar way as the maps used in the proof of proposition 4.1, taking C instead C(Ω), and by using the partial descent homomorphism J S1 R2 defined in [Cha99]. Moreover, Ψ∗ is the isomorphism induced by the S1-equivariant isomorphism of lemma 4.2, between C(X) ⊗ K(cid:0)L2(R2)(cid:1) and C0(X × R2) ⋊ R2, with X = Ω and X = one point. η and η1 are the isomorphisms given by Morita equivalence. Proof : The proof is essentially based on naturality properties of the Kasparov product and the descent homomorphism. 26 / /   ( (   u u / /     / /     / • The middle diagram commutes trivially. • The isomorphism η−1 equivalence between K(cid:0)L2(R2)(cid:1) and C i.e : K S1 1 0 (C) is the Kasparov product by the class of the S1-equivariant bimodule giving the Morita 0 (cid:16)K(cid:0)L2(R2)(cid:1)(cid:17) → K S1 η−1 0 (cid:16)K(cid:0)L2(R2)(cid:1), C(cid:17). 1 = ⊗K(L2(R2))h(cid:0)L2(R2), i, 0(cid:1)i. We denote y1 the class of (cid:0)L2(R2), i, 0(cid:1) in KK S1 0 (cid:16)K(cid:0)L2(R2)(cid:1) ⊗ C(Ω)(cid:17) → K S1 Similarly, η−1 : K S1 ⊗K(L2(R2))⊗C(Ω)y2 where y2 :=h(cid:0)L2(R2) ⊗ C(Ω), θ, 0(cid:1)i ∈ KK S1 0 (cid:16)K(cid:0)L2(R2)(cid:1) ⊗ C(Ω), C(Ω)(cid:17) where θ is the action of K(cid:0)L2(R2)(cid:1)⊗C(Ω) on the S1-´equivariant bimodule L2(R2)⊗ C(Ω) (this is the natural action of compact operators on L2(R2) and multiplication operator on C(Ω)). We then have 2 commutative diagrams (see [Ska91]) : 0 (C(Ω)) is defined by K S1 0 (C) NK(L2 (R2)) y1 φ∗ / K S1 0 (C(Ω)) 7nnnnnnnnnnnnnnnnnnnnnnnnn NK(L2 (R2)) φ∗(y1) K S1 0 (cid:0)K(cid:0)L2(R2)(cid:1)(cid:1) NK(L2 (R2 )) φ∗ 5llllllllllllllllllllllllllllll K(y2) K S1 0 (C(Ω)) NK(L2(R2 ))⊗C(Ω) y2 K S1 0 (cid:16)K(cid:0)L2(R2)(cid:1)(cid:17) But, (cid:0)φK(cid:1)∗ / K S1 0 (cid:16)K(cid:0)L2(R2)(cid:1) ⊗ C(Ω)(cid:17) φ∗ and K(y2) =(cid:2)(L2(R2) ⊗ C(Ω), θ ◦ φK, 0)(cid:3) φ∗(y1) =(cid:2)(L2(R2) ⊗ C(Ω), i ⊗ 1, 0)(cid:3). Since θ ◦ φK acts as i ⊗ 1 on L2(R2) ⊗ C(Ω), we have φ∗ (y2) = φ∗(y1) K and thus, the first upper diagram in the proposition is commutative (the vertical arrows are isomorphisms). 27 / O O 7 5 / O O • As above, we get the following commutative diagrams : K S1 0 (cid:0)C0(R2) ⋊ R2(cid:1) K S1 0 (cid:0)C0(Ω × R2) ⋊ R2(cid:1) (cid:0)φR2 C0 (R2 )(cid:1)∗ (RRRRRRRRRRRRRRRRRRRRRRRRRRRRR R2 (cid:0)τΩ([∂2])(cid:1)(cid:17) N(cid:0)φR2 C0(R2 )(cid:1)∗(cid:16)J S1 N J S1 R2 (cid:0)τΩ([∂2])(cid:1) K S1 0 (cid:0)C(Ω) ⋊ R2(cid:1) K S1 0 (cid:0)C0(R2) ⋊ R2(cid:1) C (cid:1)∗ We have N J S1 R2 ([∂2]) / K S1 0 (cid:0)C(Ω) ⋊ R2(cid:1) is R2-equivariant. N(cid:0)φR2 C (cid:1)∗(cid:0)J S1 R2 ([∂2])(cid:1) K S1 0 (cid:0)C ∗(R2)(cid:1) C0(R2)(cid:1)∗(cid:16)J S1 C (cid:1)∗(cid:0)J S1 (QQQQQQQQQQQQQQQQQQQQQQQQQQQ (cid:0)φR2 R2(cid:16)φ′∗(cid:0)τΩ([∂2])(cid:1)(cid:17) where φ′ = φC0(R2) R2(cid:0)τΩ([∂2])(cid:1)(cid:17) = J S1 R2 ([∂2])(cid:1) = J S1 R2(cid:0)φ∗([∂2])(cid:1). But,(cid:0)φR2 Similarly, (cid:0)φR2 φ∗([∂2]) =h(cid:0)H ⊗ C(Ω), M ⊗ 1, F ⊗ 1(cid:1)i ∈ KK R2⋊S1 with the notation of the class [∂2] and φ′∗(cid:0)τΩ([∂2])(cid:1) = h(cid:0)H ⊗ C(Ω), (M ⊗ iC(Ω)) ◦ φ′, F ⊗ 1(cid:1)i ∈ KK G with iC(Ω) : C(Ω) → L(C(Ω)) the multiplication operator on C(Ω). Since (M⊗iC(Ω))◦φ′ is the representation of C0(R2) on H⊗C(Ω) obtained by multiplication of C0(R2) on H and M ⊗ 1 is the same representation, we thus have equality of these two classes and of the two diagonal homo- morphisms in the above diagrams. The bottom diagram of the proposition is thus commutative. (cid:0)C0(R2), C(Ω)(cid:1) 0 (cid:0)C0(R2), C(Ω)(cid:1) 0 Let's consider the forgetful homomorphism r∗ : KK S1 (A, B) −→ KK(A, B) for A and B two C∗-algebras endowed with S1-actions. This homomorphism commutes trivially with the homomorphism η1 obtained by Morita equivalence (equivariant or not) : (cid:3) K0(C) r∗ K S1 0 (C) η1 η1 K0(K) K S1 0 (K) r∗ 28 / / (     ( /   o o   o o It also commutes trivially with the homomorphism Ψ∗ of proposition 4.7. Finally, the definition of the partial descent homomorphism in [Cha99] and the one of the usual descent homomorphism gives the equality r∗ ◦ J S1 R2 = JR2 ◦ r′∗ where r′∗ : KK R2⋊S1 (A, B) is the restriction homomorphism of R2 ⋊ S1 on R2 (see [Ska91]). In conclusion, we obtain the following commutative diagram : (A, B) −→ KK R2 r∗ wnnnnnnnnnnnn K S1 0 (C) φ∗ / K S1 0 (C(Ω)) ≃ β1◦δ1 ≃ β◦δ K S1 0 (C∗(R2)) φR2 ∗ / K S1 0 (C(Ω) ⋊ R2) K0(C) β′◦δ′ ≃ woooooooooooo r∗ K0(C∗(R2)) r∗ / K0(C(Ω) ⋊ R2) φR2 ∗ F∗ ≃ K0(C0(R2)) 0 (C(Ω)) comes from the 0 (C) that is sent on the projection 1 in where β′ and δ′ are obtained in a similar way as β1 and δ1 but without the S1-equivariance. Thus, the projection generating the summand Z in K S1 constant projection equal to 1 in K S1 K0(C). Thus, r∗◦β◦δ◦φ∗([1]) = φR2 ∗ ◦F∗(±βott) where βott is the element [Bott]−1 of K0(C0(R2)) obtained from the Bott projection and F is the Fourier transform giving the isomorphism C∗(R2) ≃ C0(R2)(cid:16)r∗([1]) = [1] spans K0(C) and since β′◦δ′ is an isomorphism, the image of [1] by this homomorphism has to be ± a generator of K0(C0(R2))(cid:17). ∗ ◦β′◦δ′◦r∗([1]) = φR2 It then suffices to follow how τ µ To begin, the map τ µ 0 (C(Ω) ⋊ R2) and on K0(C(Ω) ⋊ R2) where τ′µ is given on the dense subalgebra Cc(Ω × R2) by (see [Jul81]) ∗ is changed under all these homomorphisms. ∗ on K0(C(Ω) ⋊ R2 ⋊ S1)) becomes τ′µ ∗ on K S1 τ′µ(f ) =ZΩ f (ω, 0)dµ(ω). ∗ ◦ r∗ = τ′µ ∗ . Moreover, τ′µ τ′µ ∗ then reads on K0(C∗(R2)) as τ∗ where τ is the trace on C∗(R2) defined by τ (f ) := f (0) for any f ∈ Cc(R2). We then have τ ◦ F (f ) = F (f )(0) =RR2 f (x)dx and thus, τ∗ ◦ F∗(βott) =ZR2(cid:16)T r(Bott(x)) − 1(cid:17) dx = 0. 29 w   /     w /   / O O Thereby, and so τ′µ(cid:16)β ◦ δ ◦ φ∗([1])(cid:17) = τ′µ(cid:16)r∗ ◦ β ◦ δ ◦ φ∗([1])(cid:17) τ′µ(cid:16)β ◦ δ ◦ φ∗([1])(cid:17) = τ′µ(cid:16)φR2 ∗ ◦ F∗(±βott)(cid:17) = τ∗ ◦ F∗(±βott) = 0. We just proved : Theorem 4.8 The summand Z.q1 in K0(cid:16)C(Ω) ⋊ R2 ⋊ S1(cid:17) is traceless. 4.3 Study of the summand H 2 We will now focus on the image of the summand H 2 c (cid:0)(Ω \ F )/S1 ; Z(cid:1) in K0(cid:0)C(Ω)⋊R2 ⋊S1(cid:1) c(cid:16)(Ω\ F )/S1 ; Z(cid:17) under τ µ ∗ . For this, we construct a map K1(cid:0)C(Ω)(cid:1) −→ K0(cid:0)C(Ω) ⋊ S1(cid:1) which is onto on c(cid:16)(Ω \ F )/S1 ; Z(cid:17). This map, denoted βS1 ◦ δS1 , is defined in a similar way H 2 as the one in proposition 4.1. We consider the Dirac operator ∂1 which is S1-equivariant on the circle S1. It defines a class [∂1] ∈ KK S1 The desired map is given by the composition of the isomorphism δS1 (coming from the Morita equivalence of C(Ω) and C(Ω × S1) ⋊ S1) and the Kasparov product βS1 :=N JS1(cid:0)τΩ[∂1](cid:1). The same reasoning as the one used at the beginning of lemma 4.4 allow us to state : 1 (cid:0)C(S1), C). Lemma 4.9 H k(Ω) = 0 for every k > 4. Thus K0(C(Ω)) ≃ H 0(Ω; Z)⊕ H 2(Ω; Z) and K1(C(Ω)) ≃ H 1(Ω; Z)⊕ H 3(Ω; Z). We then prove that the H 3(cid:0)Ω ; Z(cid:1) summand of K1(cid:0)C(Ω)(cid:1) is isomorphic to the c(cid:16)(Ω \ F )/S1 ; Z(cid:17) summand of K0(cid:0)C(Ω) ⋊ S1(cid:1) and then we show that this H 2 isomorphism can be read in K-theory as the map constructed above. First, we have the following commutative diagram : K1(cid:0)C0(Ω \ F )(cid:1) β′ S1 ◦δ′ S1 Ψ tiiiiiiiiiiiiiiiii M orita eq. K0(cid:16)C0(cid:0)(Ω \ F )/S1(cid:1)(cid:17) K0(cid:0)C0(Ω \ F ) ⋊ S1(cid:1) i∗ K1(cid:0)C(Ω)(cid:1) β S1 ◦δ S1 / K0(cid:0)C(Ω) ⋊ S1(cid:1) i∗ where ψ is the composition of β′S1 ◦ δ′S1 and of the isomorphism induced by Morita equivalence of C0(Ω \ F ) ⋊ S1 and C0(cid:0)(Ω \ F )/S1(cid:1). Writing the long exact sequence of the relative cohomology groups of the pair (Ω, F ), one can prove that i∗ : H 3(Ω; Z) → H 3 c (Ω \ F ; Z) is an isomorphism 30 / /   t   o o / since H 2(F ; Z) = H 3(F ; Z) = 0. It then suffices to show that the H 3 isomorphically on H 2 with β′S1 ◦ δ′S1. For this, we have the following proposition : c (Ω \ F ; Z) part of K1(cid:0)C0(Ω \ F )(cid:1) is sent c(cid:16)(Ω \ F )/S1 ; Z(cid:17) and then to compare this isomorphism Proposition 4.10 ([Bre72]) The projection Ω \ F −→ (Ω \ F )/S1 is a S1- principal bundle. Thanks to this proposition, to the Leray-Serre spectral sequence and to the resulting Gysin sequence, we then get : Corollary 4.11 H 3 c(cid:0)Ω \ F ; Z(cid:1) = H 2 ≃ H 2 c(cid:16)(cid:0)Ω \ F(cid:1)/S1 ; Z(cid:17) ⊗ H 1(S1; Z) c(cid:16)(cid:0)Ω \ F(cid:1)/S1 ; Z(cid:17) The isomorphism of this corollary is given by "integration along the fiber". It remains to see that this isomorphism is the same map as Ψ, under the Chern character. We want to prove that the following diagram is commutative : H 3 c(cid:0)Ω \ F ; Z(cid:1)  chZ 3 K1(cid:0)C0(Ω \ F )(cid:1) ψ chZ / H 2 c(cid:16)(cid:0)Ω \ F(cid:1)/S1 ; Z(cid:17) / K0(cid:16)C0(cid:0)(Ω \ F )/S1(cid:1)(cid:17) c(cid:16)(cid:0)Ω\ F(cid:1)/S1 ; Z(cid:17) is spanned by H 3 c (D2 × S1; Z) resp. H 2 resp. But H 3 c(cid:0)Ω\ F ; Z(cid:1) resp. H 2 H 2 c (C × D2; Z) i.e by H 3 c (C × D2× S1; Z) c (D2; Z) (where C is a Cantor set and D2 is an open subset of R2) since Ω \ F resp. (cid:0)Ω \ F(cid:1)/S1 is covered by the closure of disjoint union of open sets of the form C × D2 × S1 resp. C × D2 which intersections are of dimension 2 resp. 1. We then have the following diagram (the cohomology groups are with integer coefficients) : H3 c(cid:0)Ω \ F(cid:1) hPPPPPPPPPPPPP ≃ (cid:9) H3 c (D2 × S1) H2 c (D2) c(cid:16)(cid:0)Ω \ F(cid:1)/S1(cid:17) H2 6mmmmmmmmmmmm (cid:9) (cid:9) ≃ K1(cid:0)C0(D2 × S1)(cid:1) K0(cid:0)C0(D2)(cid:1) vnnnnnnnnnnnnn K1(cid:0)C0(Ω \ F )(cid:1) (QQQQQQQQQQQQ / K0(cid:16)C0(cid:0)(Ω \ F )/S1(cid:1)(cid:17) ψ′ (cid:9) ψ 31  / / / O O O O / ?  O O O O / / h 6 O O O O / / v ( O O O O / O O where ψ′ : K1(cid:0)C0(D2 × S1)(cid:1) → K0(cid:0)C0(D2 × S1) ⋊ S1(cid:1) → K0(cid:0)C0(D2)(cid:1) is the composition of β′′S1 ◦ δ′′S1 (obtained as above with D2 × S1 instead of Ω) and the isomorphism induced by Morita equivalence. It is enough to show that the middle diagram is commutative to prove that the integration along the fiber at the level of cohomology groups is Ψ at the level of K-groups. Integration along the fiber sends a generator of H 3 tor of H 2 c (D2 × S1; Z) on a genera- c (D2; Z), and so, we must show that ψ′ maps the Bott generator of K1(cid:0)C0(D2 × S1)(cid:1) on the one of K0(cid:0)C0(D2)(cid:1). We can reduce the problem a little bit more thanks to the following diagram : K1(cid:0)C(S1)(cid:1) β′′◦δ′′=⊗C(S1 )y′ 1 ⊗Cβott K0(C) ⊗CBott (2) β′◦δ′=⊗C0(D2 ×S1)y′ 2 K1(cid:0)C0(D2 × S1)(cid:1) / K0(cid:0)C0(D2)(cid:1) the map ⊗Cβott (βott ∈ KK0(C, C0(D2))) is given by Bott periodicity in KK- theory ([Ska91] p.214). We denote α ∈ KK0(C0(D2), C) its inverse. Let's ε be the class, in KK0(cid:0)C(S1), C(S1 × S1) ⋊ S1(cid:1), of the equivalence bi- module of the Morita equivalence of C(S1) with C(S1 × S1) ⋊ S1. We then have and y′1 = ε ⊗C(S1×S1)⋊S1 JS1(cid:0)τC(S1)[∂1](cid:1) y′2 = τC0(D2)(ε) ⊗C0(D2)⊗C(S1×S1)⋊S1 JS1(cid:0)τC0(D2)⊗C(S1)[∂1](cid:1). Since S1 is not acting on D2 (the S1-action on Ω \ F is given by action on S1 in C × D2 × S1), we obtain JS1(cid:0)τC0(D2×S1)[∂1](cid:1) = τC0(D2)(cid:16)JS1(cid:0)τC(S1)[∂1](cid:1)(cid:17). Thus (see [Ska91]) : y′2 = τC0(D2)(cid:16)ε ⊗C(S1×S1)⋊S1 JS1(cid:0)τC(S1)[∂1](cid:1)(cid:17) = τC0(D2)(y′1) have (thanks to commutativity of the Kasparov product over C) : The above diagram is thus commutative since, if X ∈ K1(cid:0)C0(D2 × S1)(cid:1), we (cid:0)(X ⊗C0(D2) α) ⊗C(S1) y′1(cid:1) ⊗C βott = βott ⊗C(cid:0)(X ⊗C0(D2) α) ⊗C(S1) y′1(cid:1) = (cid:0)(βott ⊗C X) ⊗C0(D2) α(cid:1) ⊗C(S1) y′1 = (cid:0)(X ⊗C βott) ⊗C0(D2) α(cid:1) ⊗C(S1) y′1 = (cid:0)X ⊗C (βott ⊗C0(D2) α)(cid:1) ⊗C(S1) y′1 = (X ⊗C 1C) ⊗C(S1) y′1 = X ⊗C(S1) y′1 = X ⊗C0(D2×S1) y′2 It then remains to prove that the generator [u] of K1(C(S1)) is sent on the generator [1] of K0(C). 32 / /     / But the map β′′ ◦ δ′′ : K1(C(S1)) −→ K0(C), in the diagram (2), is given by ⊗C(S1)[∂1] i.e the Kasparov product by the class of the Dirac operator on the circle [∂1] ∈ KK1(C(S1), C) or, in other words, by the odd index of the Dirac operator twisted by a unitary in K1(C(S1)). This map sends [u] on [1]. Thus, we just have proved : Theorem 4.12 The map K1(C(Ω)) −→ K0(cid:0)C(Ω) ⋊ R2 ⋊ S1(cid:1) is surjective from H 3(Ω; Z) onto H 2(cid:16)(Ω \ F )/S1) ; Z(cid:17). 4.4 Index theorem to compute the trace We now want to identify the map β◦ δ◦ βS1 ◦ δS1 constructed from K1 (C(Ω)) to K0(cid:0)C(Ω) ⋊ R2 ⋊ S1(cid:1) with the Kasparov product by the class of the unbounded Kasparov cycle (see [BJ83] for definitions) D3 := (C(Ω) ⋊ R2 ⋊ S1 ⊕ C(Ω) ⋊ R2 ⋊ S1, MC(Ω), D3) where D3 is the Dirac operator along the leaves of Ω and MC(Ω) is the natural representation of C(Ω) on C(Ω) ⋊ R2 ⋊ S1 ⊕ C(Ω) ⋊ R2 ⋊ S1. Using the measured index theorem for foliation due to Alain Connes [Con79] extended to foliated spaces in [MS06], we will have (see also [DHK91]) : Theorem 4.13 ∀ b ∈ Im(β ◦ δ ◦ βS1 ◦ δS1), τ µ Sullivan current associated to µt, Chτ is the tangential Chern character (see [MS06]) and [u] ∈ K1(C(Ω)) is a lift of b for β ◦ δ ◦ βS1 ◦ δS1 . ∗ (b) =DChτ ([u]), [Cµt ]E where Cµt is the Ruelle- Cµt defines a class in the tangential cohomology group H τ an element of the longitudinal cohomology group H 3 The representative of a class in H 3 τ (Ω) locally looks like 3 (Ω) and Chτ ([u]) is τ (Ω) (see [MS06]). σ = a(x1, x2, θ, ω)dx1 ∧ dx2 ∧ dθ. The Ruelle-Sullivan current is then locally defined as the integral : [Cµt ](cid:0)[σ](cid:1) =ZΞZR2×S1 a(x1, x2, θ, ω)dx1dx2dθdµt(ω). To obtain theorem 4.13, we proceed in several steps. First, we identify βS1◦δS1 with the Kasparov product by the class of the element in KK1(C(Ω), C(Ω)⋊S1) induced by the Dirac operator d1 of dimension 1 along the leaves of Ω foliated by S1. Next, we identify β ◦ δ with the Kasparov product by the class of the element in KK0(C(Ω) ⋊ S1, C(Ω) ⋊ R2 ⋊ S1) induced by the tangential Dirac operator 33 D2 of dimension 2 transverse to the inclusion of the foliation of Ω by S1 in the foliation by R2 ⋊ S1 (see [HS87]). Finally, we prove that the composition of these two homomorphisms is given by ⊗C(Ω)(cid:2) D3(cid:3). For these computations, we will use the characterisation of the product of un- bounded cycles made by Kucerovsky in [Kuc97] (see also [BJ83] for the defi- nitions). Definition 4.14 Let A and B be two graded C∗-algebras. An unbounded Kasparov module is a triple (H, φ, D) where H is a graded Hilbert B-module, φ : A → L(H) is a ∗-homomorphism, and D is a densely defined, selfadjoint and regular unbounded operator of order 1, such that : 1. φ(a)(1 + D2)−1 ∈ K(H), 2. for all a in some dense subalgebra of A, the domain of D is stable under φ(a) and [D, φ(a)] = Dφ(a) − (−1)∂aφ(a)D is defined on Dom D and can be extended to an element of L(H). We denote Ψ(A, B) the set of unbounded Kasparov modules. If H is ungraded, such a module is an ungraded unbounded Kasparov mod- ule and we denote ψ1(A, B) the set of such modules. Julg and Baaj have proved the following result (see [BJ83] and [Kuc97]): Proposition 4.15 If (H, φ, D) ∈ Ψ(A, B) then (H, φ, D(1 + D2)−1/2) defines a Kasparov cycle in KK0(A, B). Similarly, if (H, φ, D) ∈ Ψ1(A, B) then (H, φ, D(1+D2)−1/2) defines a Kasparov cycle in KK1(A, B). There is then a map βBJ : Ψ(A, B) → KK0(A, B) (resp. Ψ1(A, B) → KK1(A, B)) defined by βBJ (H, φ, D) =(cid:2)(H, φ, D(1 + D2)−1/2)(cid:3). Kucerovsky then proved a criterion to compute the Kasparov product of two unbounded modules: Theorem 4.16 Fix (E1 ⊗φ2E2, φ1 ⊗1, D) ∈ Ψ(A, C), (E1, φ1, D1) ∈ Ψ(A, B) and (E2, φ2, D2) ∈ Ψ(B, C) such that : (i) for any x in a dense subset of φ1(A)E1, (cid:20)(cid:18) D 0 0 D2 (cid:19) ,(cid:18) 0 T ∗x Tx 0 (cid:19)(cid:21) is bounded on Dom (D⊕ D2) (where Dom F is the domain of the unbounded operator F ); (ii) Dom D ⊂ Dom (D1 ⊗1) (or vice versa); (iii) and (cid:10)(D1 ⊗1)x, Dx(cid:11) +(cid:10)Dx, (D1 ⊗1)x(cid:11) > κhx, xi for any x in the domain; 34 where x ∈ E1 is homogeneous and Tx : E2 → E maps e 7→ x ⊗e. Then (E1 ⊗φ2 E2, φ1 ⊗1, D) represents the Kasparov product of (E1, φ1, D1) and (E2, φ2, D2). This theorem applies on unbounded Kasparov modules of Ψ(A, B). In the sequel, we use ungraded unbounded Kasparov module of Ψ1(A, B). To apply the theorem of Kucerovsky, we thus need to use the isomorphism KK1(A, B) ≃ KK0(A, B ⊗ C1) (where C1 is the complex Clifford algebra of dimension 2 spanned by an element α of degree 1 satisfying α∗ = α and α2 = 1) given by : KK1(A, B) −→ (cid:2)(H, φ, F )(cid:3) KK0(A, B ⊗ C1) 7−→ (cid:2)(H ⊗C1, φ ⊗Id, F ⊗α)(cid:3) We also have : if (H, φ, D) ∈ Ψ1(A, B) then (cid:0)H ⊗C1, φ ⊗Id, D ⊗α(cid:1) ∈ Ψ(A, B ⊗ C1). To compute the Kasparov product of an element in KK1 and an element in KK0 or conversely, we use the two following lemmas (the proof can be found in [Mou09] and is easily deduced from the theorem of Kucerovsky): Lemma 4.17 Let A, B, C be three ungraded C∗-algebras. Let (E1 ⊗φ2E2, φ1 ⊗1, F ) and (E2, φ2, F2) be two unbounded Kasparov modules in Ψ1(A, C) and (E1, φ1, 0) ∈ Ψ(A, B) such that : (a) E1 is trivially graded, (b) for any x in a dense subset D of φ1(A)E1, 0 F2 (cid:19) ,(cid:18) 0 0 (cid:19)(cid:21) (cid:20)(cid:18) F T ∗x Tx 0 is bounded on Dom (F ⊕ F2), where Tx : E2 → E1 ⊗φ2 E2 maps e2 7→ x ⊗e2. Then the module (E1 ⊗φ2E2, φ1 ⊗1, F ) represents the Kasparov product of the two modules (E1, φ1, 0) and (E2, φ2, F2), i.e βBJ (E1 ⊗φ2 E2, φ1 ⊗1, F ) = βBJ (E1, φ1, 0) ⊗B βBJ (E2, φ2, F2). Lemma 4.18 Let A, B, C be three ungraded C∗-algebras. Let (B ⊗φ2 E2, φ1 ⊗1, F ) and (B, φ1, F1) be two unbounded Kasparov modules in Ψ1(A, B) and (E2, φ2, F2) ∈ Ψ(B, C) such that φ2 is nondegenerated and : (a) F is the closure of the sum P1 + P2 where P1 is an unbounded selfadjoint operator of degree 0 on B ⊗φ2E2 and P2 an unbounded selfadjoint operator of degree 1 on B ⊗φ2 E2 with Dom F = Dom P1 ∩ Dom P2, (b) for any x in a dense subset B of φ1(A)B (such that for all b ∈ B and e2 ∈ Dom F2, b ⊗e2 ∈ Dom P1 ∩ Dom P2), F2 (cid:19) ,(cid:18) 0 (cid:20)(cid:18) P2 0 0 T ∗x Tx 0 (cid:19)(cid:21) is bounded on Dom (F ⊕ F2), where Tx : E2 → B ⊗φ2E2 maps e2 7→ x ⊗e2; 35 (c) For any b ∈ B, is bounded, and Dom F2 −→ B ⊗E2 7−→ P1(b ⊗e2) e2 Dom P1 ∩ Dom P2 −→ E2 b′ ⊗e2 7−→ T ∗b P1(b′ ⊗e2) can be extended in a bounded operator on Dom F . (d) Dom F ⊂ Dom (F1 ⊗1) (or vice versa); (e) h(F1 ⊗1)x, P2yi + hP2x, (F1 ⊗1)yi = 0 for all x, y in Dom F . (f ) (−1)∂x(cid:2)h(F1 ⊗1)x, P1xi + h(P1 ⊗1)x, (F1 ⊗1)xi(cid:3) > κhx, xi for all homoge- Then (B ⊗φ2E2, φ1 ⊗1, F ) represents the Kasparov product of (B, φ1, F1) and (E2, φ2, F2), i.e neous x in the domain Dom F . βBJ (B ⊗φ2 E2, φ1 ⊗1, F ) = βBJ (B, φ1, F1) ⊗ βBJ (E2, φ2, F2). First Kasparov product We first prove that the homomorphism K1(cid:0)C(Ω)(cid:1) → K0(cid:0)C(Ω) ⋊ S1(cid:1) is the Kasparov product by the class of the Dirac operator of dimension 1 along the leaves of Ω foliated by S1. The decomposition of this map is given by : K1(cid:0)C(Ω)(cid:1)⊗[(L2(S1)op⊗C(Ω),MC(Ω),0)] ⊗[Ψ] K1(cid:0)C(Ω) ⊗ C(S1) ⋊ S1(cid:1) K1(cid:0)C(Ω × S1) ⋊ S1(cid:1) K0(cid:0)C(Ω) ⋊ S1(cid:1) ⊗JS1 (τΩ[∂1]) where Ψ : C(Ω) ⊗ C(S1) ⋊ S1 → C(Ω × S1) ⋊ S1 is the densely defined S1- equivariant map: Ψ(f )(ω, θ) = f (ω.(0, θ), θ), for f ∈ C(Ω × S1), MC(Ω) is the multiplication by an element of C(Ω) and L2(S1)op is the space L2(S1) endowed with the multiplication : λ.ξ := λξ for λ ∈ C and ξ ∈ L2(S1). We can then simplify this map : 36       Proposition 4.19 [E] ⊗ [Ψ] ⊗ JS1(τΩ[∂1]) =(cid:2)(C(Ω) ⋊ S1, MC(Ω), d1)(cid:3), where E = (L2(S1)op ⊗ C(Ω), MC(Ω), 0). Proof : (a) In KK1(cid:0)C(Ω) ⊗ C(S1) ⋊ S1, C(Ω) ⋊ S1(cid:1), [Ψ] ⊗ JS1(τΩ[∂1]) = Ψ∗(JS1 (τΩ[∂1])) and thus: [Ψ] ⊗ JS1(τΩ[∂1]) =(cid:2)(L2(S1) ⊗ C(Ω) ⋊ S1, πα ◦ Ψ, ∂1 ⊗ 1)(cid:3), with JS1 (τΩ[∂1]) =(cid:2)(L2(S1) ⊗ C(Ω) ⋊ S1, πα, ∂1 ⊗ 1)(cid:3) (see [Ska91]) where, for ζ ∈ C(S1, C(Ω)), a ∈ C(S1, C(S1)), ξ ∈ L2(S1) and f ∈ C(S1 × Ω), πα(ζ ⊗ a)(ξ ⊗ f )(k, h, ω) :=ZS1 a(g)(k)ξ(g−1k)ζ(g)(ω)f (g−1h, ω.g)dg. Define U : L2(S1) ⊗ C(S1 × Ω) → L2(S1) ⊗ C(S1 × Ω) by : ∀ f ∈ C(S1 × Ω), ζ ∈ L2(S1) U (ζ ⊗ f )(g, h, ω) := ζ(g)f (gh, ω.g−1). This map extends to a unitary operator of L2(S1) ⊗ C(Ω) ⋊ S1. Moreover, the representation of C(Ω) ⊗ C(S1) ⋊ S1 on the module defining [Ψ] ⊗ JS1 (τΩ[∂1]) is given by the natural representation obtained from the action of C(Ω) on C(Ω) ⋊ S1 and of C(S1) ⋊ S1 by compact operator action on L2(S1). The operator ∂1 ⊗ 1 is transformed in the operator ∂1 ⊗ 1 + 1 ⊗ d1. Thus, [Ψ] ⊗ JS1 (τΩ[∂1]) q . (cid:2)(L2(S1) ⊗ C(Ω) ⋊ S1, πC(S1)⋊S1 ⊗ MC(Ω), ∂1 ⊗ 1 + 1 ⊗ d1)(cid:3) (b) Since L2(S1)op ⊗C(S1)⋊S1 L2(S1) ≃ C, it remains to prove that (C(Ω) ⋊ S1, MC(Ω), d1) , E and (L2(S1) ⊗ C(Ω) ⋊ S1, πC(S1)⋊S1 ⊗ MC(Ω), ∂1 ⊗ 1 + 1 ⊗ d1) satisfy the conditions of lemma 4.17 to obtain the equality of the proposi- tion. We must verify that : (cid:20)(cid:18) d1 0 0 ∂1 ⊗ 1 + 1 ⊗ d1 (cid:19) ,(cid:18) 0 T ∗x⊗f 0 (cid:19)(cid:21) Tx⊗f 37 is a bounded operator on Dom (d1⊕ (∂1⊗1+1⊗d1)) for any x⊗f in a dense subset of L2(S1) ⊗ C(Ω) (see [BJ83] ou [Vas01] for the definition of the domain of the tensor product of an unbounded operator with the identity). Thanks to the identification L2(S1)op ⊗C(S1)⋊S1 L2(S1) ≃ C : for all x⊗ f ∈ Dom ∂1⊗ C∞τ (Ω) (where C∞τ (Ω) is the algebra of tangentially smooth maps on Ω, see [MS06]), Tx⊗f : L2(S1) ⊗ C(Ω) ⋊ S1 −→ C(Ω) ⋊ S1 ζ ⊗ b 7−→ hx, ζiL2 MC(Ω)(f )b and T ∗x⊗f : C(Ω) ⋊ S1 −→ L2(S1) ⊗ C(Ω) ⋊ S1 b 7−→ (hx, ζiC(S1)⋊S1 . ζ) ⊗ MC(Ω)(f )b q where ζ ∈ L2(S1) is chosen such that hζ, ζiL2 = 1 (for example the constant function equal to 1). One can easily obtain x ⊗ MC(Ω)(f )b 0 (cid:20)(cid:18) d1 =(cid:18) Tx⊗Grad1(f ) − T∂1x⊗f ∂1 ⊗ 1 + 1 ⊗ d1 (cid:19) ,(cid:18) 0 T ∗x⊗f T ∗∂1x⊗f + T ∗x⊗Grad1(f ) (cid:19) , 0 (cid:19)(cid:21) Tx⊗f 0 0 0 and thus this operator is bounded for any x ⊗ f in Dom ∂1 ⊗ C∞τ (Ω) which is dense in L2(S1) ⊗ C(Ω) and this ends the proof. Grad1 is the gradient on C∞τ (Ω) along S1 and is defined as follows : Let f be in C∞τ (Ω), the gradient of f along S1 is given by : Grad1(f )(ω) := lim θ→0 f (ω.θ) − f (ω) θ . Let's remark that we have the following relation : for any f ∈ C∞τ (Ω) and ζ ∈ Dom d1: d1(cid:0)MC(Ω)(f )ζ(cid:1) = MC(Ω)(f )d1ζ + MC(Ω)(cid:0)Grad1(f )(cid:1)ζ. (3) (cid:3) Second Kasparov product We now prove that the second map K0(cid:0)C(Ω) ⋊ S1(cid:1) → K0(cid:0)C(Ω) ⋊ R2 ⋊ S1(cid:1) is the Kasparov product by the class of the tangential Dirac operator D2 of dimension 2 transverse to the inclusion of the foliation by S1 in the foliation by R2 ⋊ S1 of Ω. 38 Let's remind its definition : K0(cid:0)C(Ω) ⋊ S1(cid:1) ⊗JS1 [(L2(R2,C⊕C)op⊗C(Ω),MC(Ω),0)] K0(cid:16)(cid:0)C(Ω) ⊗ C0(R2) ⋊ R2(cid:1) ⋊ S1(cid:17) K0(cid:16)(cid:0)C0(Ω × R2) ⋊ R2(cid:1) ⋊ S1(cid:17) ⊗JS1hΨR2i ⊗JR2⋊S1 (τΩ[∂2]) K0(cid:0)C(Ω) ⋊ R2 ⋊ S1(cid:1) where ΨR2 R2-equivariant map: : C(Ω) ⊗ C(S1) ⋊ S1 → C(Ω × S1) ⋊ S1 is the densely defined ΨR2 (f )(ω, x) = f (ω − x, x), for f ∈ C(Ω × R2). We define E(Ω, G; C ⊕ C) (where G represents R2 or R2 ⋊ S1) as the Hilbert C∗-module on C(Ω) ⋊ G obtained as the completion of Cc(Ω× G; C⊕ C) for the following scalar product with values in Cc(Ω × G) hf, f′i(ω, g) :=ZGhf (ω.h, h−1), f′(ω.h, h−1g)iHdg where h,iH is given by : h(λ, ν), (λ′, ν′)iH := λλ′ + νν′. This Hilbert module can be endowed with a S1-action by rotation on the second summand of C ⊕ C. As for the first Kasparov product, we can show : Proposition 4.20 JS1[E2] ⊗ JS1(cid:2)ΨR2(cid:3) ⊗ JR2⋊S1(τΩ[∂2]) =(cid:2)E(Ω, R2 ⋊ S1; C ⊕ C), πC(Ω)⋊S1 , D2)(cid:3), where E2 =(cid:0)L2(R2, C ⊕ C)op ⊗ C(Ω), MC(Ω), 0(cid:1) and πC(Ω)⋊S1 (f )(g)(ω, x, θ) =ZS1 f (ω, α) α · g((ω, 0, α)−1(ω, x, θ))dα for f ∈ C(Ω × S1), g ∈ Cc(Ω × R2 × S1, C ⊕ C) and where α · represents the action of S1 on C ⊕ C given by trivial action on the first summand and the rotation of angle α on the second. 39       Proof : (a) We have JS1 [E2] ⊗ JS1(cid:2)ΨR2(cid:3) ⊗ JR2⋊S1 (τΩ[∂2]) = JS1(cid:16)[E2] ⊗(cid:2)ΨR2(cid:3) ⊗ J S1 R2 (τΩ[∂2])(cid:17) R2 is the partial descent homomorphism constructed in [Cha99]. where J S1 (b) Using the same proof as for proposition 4.19, we can show that the homo- morphism : K S1 0 (cid:0)C(Ω)(cid:1) ⊗[E2] K S1 0 (cid:16)C(Ω) ⊗ C0(R2) ⋊ R2(cid:17) ⊗hΨR2i K S1 0 (cid:16)C0(Ω × R2) ⋊ R2(cid:17) R2 (τΩ[∂2]) K S1 ⊗J S1 0 (cid:0)C(Ω) ⋊ R2(cid:1) is the Kasparov product by h(cid:0)E(Ω, R2; C ⊕ C), MC(Ω), d2(cid:1)i where d2 is the Dirac operator along the leaves of Ω foliated by R2 that is S1-invariant. (c) Now, applying the descent homomorphism JS1, we obtain a class [(H ′, π ′, F ′)] given by : • H′ = E(Ω, R2; C ⊕ C) ⊗C(Ω)⋊R2 C(Ω) ⋊ R2 ⋊ S1 where the left action of C(Ω) ⋊ R2 on C(Ω) ⋊ R2 ⋊ S1 is given by : b.ZS1 ahuhdh :=ZS1 b ∗C(Ω)⋊R2 ahuhdh for any b ∈ C(Ω) ⋊ R2 and RS1 ahuhdh ∈ C(Ω) ⋊ R2 ⋊ S1 . • If we use the identification of E(Ω, R2; C⊕ C)⊗C(Ω)⋊R2 C(Ω) ⋊ R2 ⋊ S1 with E(Ω, R2⋊S1; C⊕C), F ′ is the tangential Dirac operator transverse to the inclusion of S1 in R2 ⋊ S1 since D2(b ∗ ah) = (d2b) ∗ ah for any h ∈ S1 and any b ∈ C(Ω) ⋊ R2 ⊕ C(Ω) ⋊ R2. • π′ is the action of C(Ω) ⋊ S1 on E(Ω, R2 ⋊ S1; C ⊕ C) described in the proposition (see [Ska91]). (cid:3) 40       Kasparov product by the tangential Dirac operator along the leaves of Ω To end this section, it remains to show that the Kasparov product of h(cid:0)C(Ω) ⋊ S1, MC(Ω), d1(cid:1)i with over C(Ω) ⋊ S1 is given by the class of the unbounded module h(cid:0)C(Ω) ⋊ R2 ⋊ S1 ⊕ C(Ω) ⋊ R2 ⋊ S1, πC(Ω)⋊S1 , D2(cid:1)i (cid:0)C(Ω) ⋊ R2 ⋊ S1 ⊕ C(Ω) ⋊ R2 ⋊ S1, MC(Ω), D3(cid:1) where D3 is the tangential Dirac operator of dimension 3 along the leaves of Ω foliated by R2 ⋊ S1. For this, we apply lemma 4.18 to prove : Theorem 4.21 h(cid:0)C(Ω) ⋊ S1, MC(Ω), d1(cid:1)i ⊗C(Ω)⋊S1 h(cid:0)ER2⋊S1, πC(Ω)⋊S1 , D2(cid:1)i =h(cid:0)ER2⋊S1 , MC(Ω), D3(cid:1)i where ER2⋊S1 = E(Ω, R2 ⋊ S1; C ⊕ C). Proof : We prove that the three unbounded Kasparov modules h(cid:0)C(Ω) ⋊ S1, MC(Ω), d1 − 1 2 Id(cid:1)i , h(cid:0)ER2⋊S1, πC(Ω)⋊S1 , D2(cid:1)i and h(cid:0)ER2⋊S1, MC(Ω), D3 − 1 2 Id(cid:1)i satisfy the hypotheses of lemma 4.18 with B = C(S1, C∞τ (Ω)). Since 2 Id(cid:1)i =h(cid:0)ER2⋊S1, MC(Ω), D3(cid:1)i 2 Id(cid:1)i =h(cid:0)C(Ω) ⋊ S1, MC(Ω), d1(cid:1)i, h(cid:0)ER2⋊S1 , MC(Ω), D3 − 1 h(cid:0)C(Ω) ⋊ S1, MC(Ω), d1 − 1 and this would prove the theorem. (a) The operator D3 is given, on the Hilbert field (cid:16)L2({ω} × R2 × S1, C ⊕ C)(cid:17)ω∈Ω 41 d3 =  0 − ∂ ∂θ 0 ∂ ∂ ∂θ ∂ − ∂ ∂x1 − ∂ − ∂ ∂x2 ∂x1 ∂x2 ∂ ∂x1 ∂x2 ∂ ∂x2 − ∂ ∂x1 d′′1   associated to the Hilbert module E(Ω, R2 ⋊ S1; C ⊕ C) (see [Con82]), by the matrix where, if f ∈ C∞c (R2 ⋊ S1, C) and Rθ is the rotation of angle θ on C, d′′1 (f )(x1, x2, θ) = iRθ ∂g ∂θ (x1, x2, θ) and g(x1, x2, θ) = R−θ(f (x1, x2, θ)). For any θ ∈ S1, define Uθ : C ⊕ C −→ C ⊕ C as the identity action on the first summand C and the action by rotation of angle −θ on the second summand. Thus, d3 = ∂′1 + ∂2 ⊗ 1 where ∂2 ⊗ 1 is the tensor product of the Dirac operator on R2 and the identity, and if f ∈ C∞c (R2 ⋊ S1, C ⊕ C), ∂′1(f )(x1, x2, θ) = iU∗θ (x1, x2, θ) ∂g ∂θ with g(x1, x2, θ) = Uθ(f (x1, x2, θ)). Thus, D3 = D′1 + D2 where D′1 (resp. D2) is the operator given by ∂′1 (resp. ∂2 ⊗ 1) on the Hilbert field associated to E(Ω, R2 ⋊ S1; C ⊕ C). Identifying C(Ω) ⋊ S1 ⊗C(Ω)⋊S1 ER2⋊S1 with πC(Ω)⋊S1 (C(Ω) ⋊ S1).ER2⋊S1 ≃ ER2⋊S1 , for any f ⊗g ∈ Dom (d1 ⊗ 1) ∩ Dom D2, (cid:0)D3 − 1 = = (−1)∂gπC(Ω)⋊S1 (d1f )g − 1 = 2 Id(cid:1)(cid:0)f ⊗g(cid:1) = (cid:0)D3 − 1 (cid:0)D′1 − 1 (−1)∂g(d1f ) ⊗g − 1 2 Id(cid:1)(cid:0)πC(Ω)⋊S1(f )g(cid:1) 2 Id(cid:1)(cid:0)πC(Ω)⋊S1 (f )g(cid:1) + D2(cid:0)πC(Ω)⋊S1 (f )g(cid:1) 2 (f ⊗g) + D2(cid:0)πC(Ω)⋊S1 (f )g(cid:1) 2 (f ⊗g) + D2(cid:0)πC(Ω)⋊S1(f )g(cid:1). The sign is due to the fact that C is acting on C ⊕ C by multiplication on the first summand and by conjugation on the second. We thus need to verify the hypotheses of the lemma for P1(f ⊗g) = (−1)∂f d1(f ) ⊗g − 1 2 f ⊗g and P2 = D2. (b) If x ∈ B, one can easily compute : Tx 0 (cid:20)(cid:18) D2 0 D2 (cid:19) ,(cid:18) 0 T ∗x 0 (cid:19)(cid:21) = T ′Grad2(x) 0 0 T ′Grad2(x∗)! where 42 (1) D2(πC(Ω)⋊S1(f )g) = πC(Ω)⋊S1 (f )D2g + πC(Ω)⋊S1(cid:0)Grad2(f )(cid:1) · g since d2 commutes with the S1-action on C(Ω) ⋊ R2 ⊕ C(Ω) ⋊ R2. Grad2 is the gradient on C(S1, C∞τ (Ω)) along R2 and is defined as follows: If f ∈ C∞τ (Ω) and e1, e2 are the two vectors of the canonical basis of R2, then, if ∂1f (ω) := lim y→0 ∂2f (ω) := lim y→0 f (ω + ye1) − f (ω) f (ω + ye2) − f (ω) y y , the gradient along R2 is given by : Grad2(f )(ω) := ∂1f (ω)e1 + ∂2f (ω)e2 = (cid:18) 0 ∂1f (ω) + i∂2f (ω) −∂1f (ω) + i∂2f (ω) 0 (cid:19) . For any function x in C(S1, C∞τ (Ω)), we still denote Grad2(x) the func- tion of C(S1, C∞τ (Ω) ⊕ C∞τ (Ω)) : θ 7−→ Grad2(x(θ, .)). Moreover, we denoted πC(Ω)⋊S1(cid:0)Grad2(f )(cid:1) · g the product : (cid:19) · g πC(Ω)⋊S1(−f1 + if2) πC(Ω)⋊S1(f1 + if2) (cid:18) 0 0 if Grad2(f ) = f1e1 + f2e2. (2) for any x ∈ B, Tx : ER2⋊S1 −→ C(Ω) ⋊ S1 ⊗C(Ω)⋊S1 ER2⋊S1 ≃ b 7−→ x ⊗ b ER2⋊S1 ≃ πC(Ω)⋊S1(x)b (3) for any x ∈ B, T ∗x : C(Ω) ⋊ S1 ⊗C(Ω)⋊S1 ER2⋊S1 −→ ER2⋊S1 7−→ hx, fi.b f ⊗ b where hx, fi.b := πC(Ω)⋊S1 (hx, fi)b. (4) for any x ∈ B, T ′Grad2(x) : ER2⋊S1 −→ ER2⋊S1 is given by multiplication by 0 πC(Ω)⋊S1 (G1 + iG2) (cid:18) πC(Ω)⋊S1(−G1 + iG2) 0 (cid:19) , if Grad2(x) = G1e1 + G2e2. (c) For any b ∈ B and e2 ∈ Dom D2 , P1(b ⊗e2) = (−1)∂e2 d1(b) ⊗e2 − 1 2 b ⊗e2 is bounded for fixed b thus Dom D2 −→ B ⊗E2 7−→ P1(b ⊗e2) e2 43 is bounded. Moreover, for any b ∈ B and b′ ⊗e2 ∈ Dom P1 ∩ Dom P2, T ∗b P1(b′ ⊗e2) = (−1)∂e2hb, d1(b′)i.e2 − 1 2hb, b′i.e2. Thereby, since d1 is selfadjoint, and T ∗b P1(b′ ⊗e2) =(cid:2)T ∗d1(b) − 1 Dom P1 ∩ Dom P2 −→ 2 T ∗b(cid:3)(b′ ⊗e2) 7−→ T ∗b P1(b′ ⊗e2) b′ ⊗e2 E2 (d) Using the ellipticity of D3, we get extends to a bounded operator on Dom (cid:0)D3 − 1 2 Id(cid:1). Dom D3 ⊂ Dom (d1 ⊗1). where Dom (cid:0)d1 ⊗1(cid:1) := Im(1 + d2 Thus, Dom (cid:0)D3 − 1 2 ⊗1. 2 Id(cid:1) ⊂ Dom (cid:0)(cid:0)d1 − 1 1)− 1 2 Id(cid:1) ⊗1(cid:1). 2 Id and D2 are selfadjoint and (e) Since d1 ⊗1 − 1 for any x in the domain, (cid:0)d1 ⊗1 − 1 2 Id(cid:1)D2 + D2(cid:0)d1 ⊗ 1 − 1 2 Id(cid:1)x, D2x(cid:11) +(cid:10)D2x,(cid:0)d1 ⊗1 − 1 2 Id(cid:1) = 0, 2 Id(cid:1)x(cid:11) = 0. (cid:10)(cid:0)d1 ⊗1 − 1 (f) One can easily prove : • for any x in the domain such that x is of degree 0, D(cid:16)(cid:0)d1 − 1 2 Id(cid:1) ⊗1(cid:17)x, P1xE = D(cid:16)(cid:0)d1 − 1 2 Id(cid:1) ⊗1(cid:17)x,(cid:16)(cid:0)d1 − 1 2 Id(cid:1) ⊗1(cid:17)xE > 0, • for any x in the domain such that x is of degree 1, 2 Id(cid:1) ⊗1(cid:17)x, P1xE = D(cid:0)d1 ⊗1(cid:1)x,(cid:0)d1 ⊗1(cid:1)xE − 1 −D(cid:16)(cid:0)d1 − 1 4 hx, xi > − 1 4 hx, xi, Thus, 2 Id(cid:1) ⊗1(cid:17)x, P1xE + DP1x,(cid:16)(cid:0)d1 − 1 (−1)∂x hD(cid:16)(cid:0)d1 − 1 for any x in the domain. The last hypothesis of lemma 4.18 is thus satisfied and the theorem is thus proved. 2 Id(cid:1) ⊗1(cid:17)xEi > − 1 4 hx, xi, (cid:3) 44 4.5 Computation of the image under τ µ ∗ of the summands Z.qi In this section, we prove that τ µ tation, where qi is the lift of qi in K S1 ∗(cid:0)β ◦ δ( qi)(cid:1) ∈ µt(C(Ξ, Z)) by an explicit compu- 0 (C(Ω)) built in Appendix A. It suffices to make this computation for one of the qi's, for example q2. To compute this image, we use the index theorem for Ω foliated by R2. In fact, the following diagram is commutative K0(C(Ω)) ≃ β2◦δ2 K0(C(Ω) ⋊ R2) &LLLLLLLLLLL τ ′µ ∗ K S1 0 (C(Ω)) ≃ β◦δ K S1 0 (C(Ω) ⋊ R2) xqqqqqqqqqqq τ µ ∗ r∗ r∗ R where β2 and δ2 are obtained similarly to β and δ but forgetting the S1- equivariance and τ′µ Thus, to compute τ µ ∗ was defined in section 4.2. ∗ (β ◦ δ( q2)), it suffices to compute τ′µ ∗ (cid:0)β2 ◦ δ2 ◦ r∗( q2)(cid:1). A nonequivariant version of proposition 4.20 prove that β2 ◦ δ2 is given by the Kasparov product by the class of the unbounded Kasparov module given by the tangential Dirac operator along the leaves of Ω foliated by R2 and the index theorem for foliated spaces gives : τ′µ ∗ (cid:16)β2 ◦ δ2 ◦ r∗( q2)(cid:17) =(cid:10)Ch2 τ (r∗( q2)) [CνZ ](cid:11) where Z is a transversal of the foliated space, νZ an invariant transverse mea- sure, [CνZ ] its associated Ruelle-Sullivan current and Ch2 τ ( q2) is the Chern char- acter component that is in H 2 τ (Ω), the tangential cohomology of Ω foliated by R2. Let's remind some notations from Appendix A. Let ω0 be the tiling in F1 fixed by the rotation Rπ of angle π around the origin and used to build q2. Ξ00 is the set of all the tilings with the same 1-corona as ω0. Fix r0 < r < r1 with r = (r0 + r1)/2 small enough and r0, r1 close enough to r. Denote Ωr0,r1 := {eiθw + v; w ∈ Ξ00, θ ∈ R, v ∈ R2, r0 ≤ v ≤ r1}. Since outside Ωr0,r1, the bundle defining q2 is trivial (see Appendix A), it suffices to compute the Chern character on this corona. Let's recall that X is formed by all the tilings with the same 1-corona as ω0 but that are not fixed by Rπ. Define X 0 := X ∪ {ω0}. In the sequel, we will consider the transversal Z = X × S1, endowed with the invariant transverse measure µt X is the measure induced on X by µ and dS1 is the usual measure on S1. X ⊗ dS1 where µt 45     o o & o o x We denote lk (k = 1, . . . , 4) the four quadrants of the circle of radius r in C i.e lk =(cid:8)rekiθ, θ ∈(cid:2)0; π 2(cid:3)(cid:9). Define also Vk Yk Yk := {eiθ(w + v); w ∈ X 0, v ∈ lk, θ ∈ R}, := {w + v; w ∈ X 0, v ∈ lk} := Yk/S1. and Let's now fix a smooth function ε : S1 → C satisfying ε(z) = 1, ε(1) = 1, ε(i) = −1, ε′(1) = 0 = ε′(i) and ε(ei(θ+π)) = ε(eiθ). We then define the following functions : f1 : Y1 → C is given by f1(y) = ε(eiθ), if y ≃ (w, reiθ ) under the identifications Y1 ≃ Y1 ≃ X 0 × l1. f2 : Y2 → C is defined by f2(y) = −1. f3 : Y3 → C is given by f3(y) = −ε(eiθ), if y ≃ (w, reiθ) under the identifications Y3 ≃ Y3 ≃ X 0 × l3. f4 : Y4 → C is defined by f4(y) = 1. In the appendix, we didn't need a smooth function and we took ε(z) = z2 but the construction of q2 obtained in the appendix is still available with any map ε satisfying the above conditions. Since {eiθw0 + v; θ ∈ R, v ∈ R2, r0 ≤ v ≤ r1} is of µ-measure zero ([RS98]), τ ( q2) on Ωr0,r1 \{eiθw0 + v; θ ∈ R, v ∈ R2, r0 ≤ v ≤ r1}. it suffices to study Ch2 But this set admits a trivialization : s : X × ([0; π] × S1) × S1 ≃ Ωr0,r1 \ {eiθw0 + v; θ ∈ R, v ∈ R2, r0 ≤ v ≤ r1} given by (w, x, eiθ, eiα) 7→ eiα(cid:0)w + ( r1−r0 On X × ([0; π] × S1) × S1 (that is a chart of Ω foliated by R2), the projection associated to q2 is given, modulo s, on C⊕ C (with S1-action by rotation on the first summand and trivial action on the second) by the matrix : π x + r0)eiθ(cid:1). e(w, x, eiθ, eiα) := sin(x)e−iαf(cid:0)[w + reiθ](cid:1) (cid:19) 1 − cos(x) 1 1 + cos(x) 2(cid:18) sin(x)eiαf(cid:0)[w + reiθ](cid:1) where f : S4 y ∈ Y ′k ⊂ Yk with Y ′k := {w + v; w ∈ X, v ∈ lk} and Y ′k := Y ′k/S1. f ([w + reiθ]) doesn't depend on the chosen w in X but only on eiθ. Y ′k −→ S1 ⊂ C is the continuous map given by f (y) = fk(y) if k=1 Thus, Ch2 τ ( q2) = 1 2iπ T r(edede) is the tangential form defined by (w, x, z, eiα) 7→ 1 4iπ sin(x) f ∂f ∂z dx dz. Thereby (cid:10)Ch2 τ ( q2) [CνZ ](cid:11) =ZXZS1Z[0;π]×S1 1 4iπ sin(x) f ∂f ∂z dx dz dS1 (eiα) dµt X (ω). 46 Moreover, 1 ∂f ∂z sin(x) f 4iπ Z[0;π]×S1 where Ind0(f ) is the winding number of f around 0 (we recall that f f = 1). Since Ind0(f ) ∈ Z, sin(x)dx ZS1 4iπ Z[0;π] dz = Ind0(f ) dx dz = ∂f ∂z 1 f X(X), with l ∈ Z. (cid:10)Ch2 τ ( q2) [CνZ ](cid:11) = lµt 3(X − v) = µt 3 on X3. X (X) = µt As X3 := X ∪ Ξ is a transversal in Ω foliated by R2 ⋊ S1, µ induces an invariant transverse measure µt 3(U \ {ω0 − v}), where v ∈ R2 is such Then µt that ω0 − v ∈ Ξ and U is the set formed by the tilings of Ω coinciding on the 1-corona of ω0 translated by −v. But µt Since U is a clopen subset of Ξ, we deduce 3({ω0 − v}) = 0 thus µt 3(U ) = µt(U ). X (X) = µt 3(X) = µt (cid:10)Ch2 τ ( q2) [CνZ ](cid:11) = lµt(U ) ∈ µt(cid:0)C(Ξ, Z)(cid:1). (cid:3) We thus obtained : Theorem 4.22 5 Conclusion τ µ ∗ 7 Mi=2 Z.qi! ⊂ µt(cid:0)C(Ξ, Z)(cid:1). From theorem 4.8, 4.13 and 4.22, we thus have obtained : Theorem : If T is a pinwheel tiling, Ω = Ω(T ) its hull provided with an invariant ergodic probability measure µ and Ξ its canonical transversal provided with the induced measure µt, then K0(C(Ω) ⋊ R2 ⋊ S1) ψ2≃ Z ⊕ Z6 ⊕ H 2 c(cid:16)(Ω \ F )/S1 ; Z(cid:17). And • τ µ ∗(cid:0)ψ2(Z)(cid:1) = 0. • τ µ ∗(cid:0)ψ2(Z6)(cid:1) ⊂ µt(C(Ξ, Z)). • ∀b ∈ H, ∃[u] ∈ K1(cid:0)C(Ω)(cid:1) such that : ∗ (b) = τ µ τ µ ∗ ([u] ⊗C(Ω) [D3]) =(cid:10)Chτ ([u]), [Cµt ](cid:11). To prove the gap-labeling conjecture for pinwheel tilings, it then remains to study the cohomological part of the index formula proved in the last point of this theorem. This is done in [Mou], proving that τ µ ∗(cid:16)K0(cid:0)C(Ω) ⋊ R2 ⋊ S1(cid:1)(cid:17) = µt(cid:0)C(Ξ, Z)(cid:1) = 1 264 5(cid:21) . Z(cid:20) 1 47 4 8 Figure 6: 28 collared prototiles. 4 9 Figure 7: 26 other collared prototiles. A Proof of lemma 4.5 We prove in this appendix that the kernel of the connecting map in lemma 4.5 is isomorphic to Z7 with explicit generators. First inclusion We have already seen that K S1 section that 7 Li=1 0 (C(F )) ≃ Z12 in section 4.1 and we prove in this Z.qi ⊂ Ker ∂ where q1 = (1, 1, 1, 1, 1, 1, 0, 0, 0, 0, 0, 0) q2 = (0, 1, 1, 1, 1, 1, 1, 0, 0, 0, 0, 0) q3 = (1, 0, 1, 1, 1, 1, 0, 1, 0, 0, 0, 0) q4 = (1, 1, 0, 1, 1, 1, 0, 0, 1, 0, 0, 0) q5 = (1, 1, 1, 0, 1, 1, 0, 0, 0, 1, 0, 0) q6 = (1, 1, 1, 1, 0, 1, 0, 0, 0, 0, 1, 0) q7 = (1, 1, 1, 1, 1, 0, 0, 0, 0, 0, 0, 1). For this, we show the inclusion Z.q1 ⊂ Ker ∂. It suffices to show that (1, . . . , 1, 0, . . . , 0) ∈ Ker ∂. But (1, . . . , 1, 0, . . . , 0) is the projection in K S1 S1-invariant projection f equal to 1 in C(F ). The constant map Ψ := 1 in C(Ω) is then a S1-invariant and selfadjoint map which lifts f (i.e π(Ψ) = f where π : C(Ω) −→ C(F )). Thus, the definition of ∂ gives : 0 (cid:0)C(F )(cid:1) associated to the constant ∂([f ]) = [exp(−2iπΨ)] = [1]. Thereby q1 ∈ Ker ∂. The next step is to show that Z.q2 = Z.(0, 1, 1, 1, 1, 1, 1, 0, 0, 0, 0, 0) ⊂ Ker ∂. The other inclusions are shown in a similar way. Let ω0 ∈ F1 be a fixed tiling for the rotation Rπ of angle π around the origin and let Ξ00 be the set of the tilings having the same 1-corona as ω0 (the 1-corona of ω0 consists of all the tiles intersecting the tiles, in ω0, containing the origin). Set : Ωr :=(cid:8)eiθω + v; θ ∈ R, w ∈ Ξ00, v ∈ R2,v = r(cid:9) with r small enough. The S1-action on Ωr is free, thus we obtain a S1-principal bundle Ωr → Ωr/S1. It is clear, from the study of Ω/S1 made in [Mou], that Ωr/S1 has no coho- mology groups in degree greater than 2 and thus the bundle Ωr → Ωr/S1 is trivializable. We then exhibit a trivialization of this bundle. 50 We first exhibit a fundamental domain of Ωr for its S1-action. The nonsymmetric 2-coronas that can surround the 1-corona around the origin in ω0 can be partitionned in two sets such that Rπ sends one on the other (a 2-corona consists of all the tiles that intersect the 1-corona). Let X2 be the union of the clopen sets of Ξ00 associated to one of these sets. Next, we consider the nonsymmetric 3-coronas of the 1-corona around the origin in ω0 such that the 2-corona is symmetric. We can partition as above and we obtain a clopen set X3. Then we consider nonsymmetric 4-coronas of the 1-corona around the origin in ω0 such that the 3-corona is symmetric and so on. Xn, We also set We have the clopen sets X2, X3, X4, . . . of Ξ00 and if we denote X = Sn>2 then X is an open set of Ξ00 and X, {ω0}, −X is a partition of Ξ00. We denote X 0 := X ∪ {ω0}. In the sequel, we call lk (k = 1, . . . , 4) the four quadrants of the circle of radius r around the origin of C i.e lk =(cid:8)rekiθ, θ ∈(cid:2)0; π 2(cid:3)(cid:9). and Yk := {w + v; w ∈ X 0, v ∈ lk}. We first prove that Vk → Vk/S1 is a trivial principal bundle homeomorphic to Yk × S1. Vk := {eiθ(w + v); w ∈ X 0, v ∈ lk, θ ∈ R} Let Ψk : Yk × S1 −→ Vk be defined by ψk(w + v, eiθ) := eiθ(w + v). We show that this is a homeomorphism. ψk is easily seen to be surjective from the definition of Yk and Vk. If ψk(w + v, eiθ) = ψk(w′ + v′, eiθ′ ) then ei(θ−θ′)(w + v) − v′ = w′. If r is small enough, ei(θ−θ′)(v) − v′ can send ei(θ−θ′)w on another tiling of X 0 only if it vanishes. Thus, ei(θ−θ′)(v) = v′ and ei(θ−θ′)(w) = w′. The only possibility is θ′ = θ modulo π. But, if θ′ = θ + π + 2lπ (l ∈ Z), then v′ = −v, which is not possible since v′, v ∈ lk. Thereby, θ′ = θ + 2lπ and (w + v, eiθ) = (w′ + v′, eiθ′ And ψk is injective. ). Since ψk is a continuous bijection from the compact set Yk × S1 on the compact set Vk, it is a homeomorphism. Thereby, Yk ≃ Vk/S1 and, denoting Yk := Yk/S1, Yk is homeomorphic to Yk. We will call sk this homeomorphism. As Yk ≃ X 0 × lk, we also note that we can identify Yk with X 0 × lk. We next study the intersection Yk ∩ Yk′ (k 6= k′) in Ωr/S1. 51 If [w + v] = [w′ + v′] with [w + v] ∈ Yk and [w′ + v′] ∈ Yk′ , there exists θ ∈ R such that eiθ(w + v) = w′ + v′. As above, v′ must be equal to eiθv and thus w′ = eiθw. There are then only two possibilities: • if w 6= w0, then θ = 2lπ (l ∈ Z) and v′ = v, w′ = w. • if w = w0, then θ = lπ (l ∈ Z) and thus w′ = w0, v′ = (−1)lv. Thereby, Yk ∩ Yk′ ≃ X 0 × (lk ∩ lk′ )[{w0} × (−lk ∩ lk′ ). We then define 4 maps allowing us to glue the ( Yk × S1)'s to identify the union of these sets with Ωr. We can identify each Yk with X 0 × lk which allow us to set pk the projection on the second coordinate. In the sequel, we consider the lk's as imbedded in C which allow us to use the square of elements of lk. f1 : Y1 → C is then defined by f1(y) = p1(s1(y))2 Y1 ≃ Y1 and Y1 ≃ X 0 × l1, f1 is obtained by : if y ≃ (w, reiθ), f1(y) = e2iθ. f2 : Y2 → C is defined by f2(y) = −1. f3 : Y3 → C is defined by f3(y) = − p3(s3(y))2 identifications Y3 ≃ Y3 ≃ X 0 × l3, in the following way : f3(y) = −e2iθ. f4 : Y4 → C is defined by f4(y) = 1. . f3 is obtained, under the if y ≃ (w, reiθ), . Under the identifications r2 r2 We introduce these maps because the transition maps on the intersection of Yk with Yk′ are given by fkfk′ . Let's prove that we can then define a map φ : (cid:18) 4 Sk=1 φ(y, eiθ) := fk(y)eiθsk(y) and that this is a homeomorphism. Yk(cid:19) × S1 −→ Ωr by If y = [w + v] ∈ Yk ∩ Yk′ , then there are two possibilities : • if w 6= w0, then k = k′ ± 1 mod 4. We can assume, even if it means interchanging the roles of k and k′, that k = k′ + 1 mod 4. We then have sk([w + v]) = w + rik−1 and sk′ ([w + v]) = w + rik−1. One can easily prove that fk(y) = fk′ (y) and thus fk(y)eiθsk(y) = fk′ (y)eiθsk′ (y). • if w = w0, then ∗ either k = k′ ± 1 mod 4 and we can suppose k = k′ + 1 mod 4. Then sk([w + v]) = w0 + rik−1 or sk([w + v]) = w0 + rik. In the first case, we obtain sk′ ([w + v]) = w0 + rik−1 and in the second, we have sk′ ([w + v]) = w0 + rik−2. 52 In the first case, one can easily verify that fk(y) = fk′ (y) and, since sk′ ([w + v]) = sk([w + v]), we thus obtain that fk(y)eiθsk(y) = fk′ (y)eiθsk′ (y). In the second case, fk(y) = −fk′(y) and we also obtain that fk(y)eiθsk(y) = fk′ (y)eiθsk′ (y), since sk′ ([w+v]) = −sk([w+v]). ∗ or k = k′ ± 2 mod 4 and if sk([w + v]) = w0 + v, then we have sk′ ([w + v]) = w0 − v. Again, one can verify that fk(y)eiθsk(y) = fk′ (y)eiθsk′ (y) since, if k = k′ ± 2 mod 4, fk = −fk′. Thus, φ is well defined. Now, assume y ∈ Yk and y′ ∈ Yk′ satisfy φ(y, eiθ) = φ(y′, eiθ′ Note sk(y) = w + v and sk′ (y′) = w′ + v′. One has fk(y)eiθ(w + v) = fk′ (y′)eiθ′ Thus, projecting in Ωr/S1, y = y′ is an element of the intersection Yk∩ Yk′ . It is enough then to prove that θ = θ′ + 2lπ, l ∈ Z. As above, we must have v′ = fk′ (y)−1fk(y)ei(θ−θ′)(v) and thereby (w′ + v′). ). w′ = fk′ (y)−1fk(y)ei(θ−θ′)(w). There are several cases : • either fk′ (y)−1fk(y)ei(θ−θ′) = 1 and w′ = w, v′ = v. Then k is equal to k′, k′ + 1 or k′ − 1 mod 4. In the three cases, fk(y) = fk′ (y) thus ei(θ−θ′) = 1. So eiθ = eiθ′ . • or fk′ (y)−1fk(y)ei(θ−θ′) = −1 then w = w0 = −w′ and v′ = −v. Since, for two opposite points, we have fk(y) = −fk′(y′), we also obtain that eiθ = eiθ′ . In every cases, we have (y, eiθ) = (y′, eiθ′ ) thus φ is injective. One can easily see that φ is also surjective and continuous thus φ is a homeomorphism between (cid:18) 4 Sk=1 Yk(cid:19) × S1 and Ωr. Thereby, we have constructed a trivialization of the projection Ωr −→ Ωr/S1. We can now construct the fiber bundle on Ω lifting (0, 1, 1, 1, 1, 1, 1, 0, 0, 0, 0, 0) ∈ K S1 0 (C(F )). To do this, let's take a tiling w0 ∈ F1 fixed by Rπ and any r small enough. The above construction gives Ωr, Ω≤r and H = c (Ω<r) where Ω≤r :=(cid:8)eiθω + v; θ ∈ R, w ∈ Ξ00, v ∈ R2,v ≤ r(cid:9) 53 and Ω<r :=(cid:8)eiθω + v; θ ∈ R, w ∈ Ξ00, v ∈ R2,v < r(cid:9) . On Ω≤r, we take the line bundle Ωs≤r × C with the diagonal action of S1 where the S1-action on C is given by rotation. On H, we take the line bundle H × C with the diagonal action of S1 but, this time, the S1-action on C is trivial. The gluing map is then given on the intersection Ω≤r ∩ H = Ωr ≃ Ωr/S1 × S1 by f (w, eiθ, z) = (w, eiθ, eiθz). Thus, since (0, 1, 1, 1, 1, 1, 1, 0, 0, 0, 0, 0) ∈ K S1 K S1 0 (C(Ω)), it is in the kernel of ∂. 0 (C(F )) lifts on an element of Thereby 7 Li=1 Z.qi ⊂ Ker ∂. Inclusion in the other direction 7 Z.qi. We next prove that Ker ∂ ⊂ Let (n1, . . . , n6, n′1, . . . , n′6) be in Z12. We can assume that the n′i's are zero thanks to the last section, since, for example, as (1, 0, 0, 0, 0, 0,−1, 0, 0, 0, 0, 0) ∈ Ker ∂, Li=1 ∂(n1, . . . , n6, n′1, . . . , n′6) = ∂(n1 + n′1, n2, . . . , n6, 0, n′2, . . . , n′6). Moreover, we can also assume that every ni + n′i are in N thanks to the last section, since, if n is the smallest integer among the (ni + n′i)'s, we have ∂(n1 + n′1, . . . , n6 + n′6, 0, . . . , 0) = ∂(n1 + n′1 − n, . . . , n6 + n′6 − n, 0, . . . , 0), with ni + n′i − n ∈ N. Let's fix (n1, . . . , n6, 0, . . . , 0) ∈ Z12 with ni ∈ N. (n1, . . . , n6, 0, . . . , 0) is associated with the S1-invariant projection : where f (x) =  Pni :=  1 Pn1 ... Pn6 . . . 1 if x ∈ F1 if x ∈ F6   ∈ Mni(C). We will exhibit a S1-invariant projection in M∞(C(Ω)) which lifts this projec- tion. For this, consider Ki :=nω ∈ Ω ; d(ω, Fi) 6 54 1 no and 1 Vi :=nω ∈ Ω ; d(ω, Fi) < n − 1o with n big enough so that the Vi's are disjoint. Set Ki := p( Ki) and Vi := p( Vi) where p : Ω −→ Ω/S1. Ki is then a compact subset of the open set Vi and it contains [xi] := p(xi) where xi is some element of Fi. Thus, there exist maps φi ∈ C(Ω/S1), i ∈ {1, . . . , 6}, such that • Supp(φi) ⊂ Vi; • 0 6 φi 6 1 and φi(x) = 1 if x ∈ Ki ; • φ1 + . . . + φ6 = 1 on S Ki. Set Ψ := φ1.P ′n1 + . . . + φ6.P ′n6 where P ′ni := ∈ MMax(ni)(C). 1 . . . 1 0 . . . 0     Thereby, Ψ ∈ MMax(ni)(cid:0)C(cid:0)Ω/S1(cid:1)(cid:1) and Ψ is selfadjoint. We can then see Ψ as a function of C(Ω), constant on the S1-orbits and Ψ can thus be seen as a S1-invariant function. One easily see that π∗(Ψ) = f in M∞(C(F )), thereby ∂([f ]) = [exp(−2iπΨ)] ∈ K S1 1 (C0(cid:0)Ω \ F )(cid:1). By lemma 4.4, we have an isomorphism between K S1 H 1 c(cid:16)(Ω \ F )/S1, Z(cid:17) =(cid:26)[h] ; h :(cid:16)(Ω \ F )/S1(cid:17)+ 1 (C0(Ω \ F )) and −→ S1 continuous (cid:27) −→ S1 homotopic to is the Alexandroff compactification of (Ω \ F )/S1 (see [Hat02] for the equality of the cohomology group and the set of classes of where [h] is the class of continuous maps (cid:16)(Ω \ F )/S1(cid:17)+ h and (cid:16)(Ω \ F )/S1(cid:17)+ continuous maps from (cid:16)(Ω \ F )/S1(cid:17)+ This isomorphism is given by det∗. Thus : ∂([f ]) = [exp(−2iπ(n1.φ1 + . . . + n6.φ6))] ∈ H 1 Note that a continuous map on(cid:16)(Ω\ F )/S1(cid:17)+ c(cid:16)(Ω \ F )/S1 ; Z(cid:17). on Ω/S1, constant on F/S1. can be seen as a continuous map to S1(cid:17). It remains to find the (n1, . . . , n6, 0, . . . , 0)'s for which there exists a continu- and with values in S1 between ous homotopy Ht defined on (cid:16)(Ω \ F )/S1(cid:17)+ 55 exp(−2iπ(n1.φ1 + . . . + n6.φ6)) and 1. This is equivalent to know if there exists a continuous homotopy Ht : Ω/S1 → S1 between exp(−2iπ(n1.φ1 + . . . + n6.φ6)) and 1 which is constant on F/S1 for any t. Or, in other words, we want to know if we can find a continuous homotopy ht : Ω/S1 −→ R between h0 := n1.φ1 + . . . + n6.φ6 and a continuous map h1 := g defined on Ω/S1 with values in Z and such that ht([xi]) − ht([xj]) ∈ Z for any t ∈ [0, 1], i, j ∈ {1, . . . , 6}. Since Ω/S1 is connected, ht would be a homotopy between n1.φ1 + . . . + n6.φ6 and a constant map g on Ω/S1 equal to an integer and satisfying ht([xi]) − ht([xj ]) ∈ Z pour tout t, i, j. We prove that this is possible only if ni = nj for all i, j. Let evi(h) : [0; 1] −→ R be the continuous map defined by evi(h)(t) := ht([xi]). evi(h)−evj(h) is then a continuous map with integer values and thus is constant equal to ni − nj. Thereby, ∀t ∈ [0; 1], ht([xi]) − ht([xj]) = ni − nj. But, for t = 0, h0 = g = k ∈ Z and so, if there exists such a homotopy ht, then ni = nj for all i, j. Thus, (n1, . . . , n6, 0, . . . , 0) ∈ Ker ∂ ⇒ (n1, . . . , n6, 0, . . . , 0) ∈ Z.(1, . . . , 1, 0, . . . , 0). If now (n1, . . . , n6, n′1, . . . , n′6) ∈ Ker∂, ∂(n1, . . . , n6, n′1, . . . , n′6) = ∂(n1 + n′1, . . . , n6 + n′6, 0, . . . , 0) = ∂(n1 + n′1 − n, . . . , n6 + n′6 − n, 0, . . . , 0) where n = M in(ni + n′i). From the above result, we have ni + n′i − n = nj + n′j − n, i.e ni + n′i = nj + n′j for any i, j. Let k be this common integer, we have obtained that ni = −n′i + k, thus (n1, . . . , n6, n′1, . . . , n′6) = (−n′1 + k, . . . ,−n′6 + k, n′1, . . . , n′6) = k.q1 + n′1(q2 − q1) + . . . + n′6(q7 − q1) ∈ Thereby, we have proved that Ker ∂ ⊂ This completes the proof of lemma 4.5. Z.qi. 7 Li=1 56 Z.qi. 7 Li=1 References [BBG06] J. Bellissard, R. Benedetti, and J.-M. Gambaudo. Spaces of tilings, fi- nite telescopic approximation and gap labelings. Comm. Math. Phys., 261:1 -- 41, 2006. [Bel82] [Bel86] [Bel92] [BG03] J. Bellissard. Schrodinger's operator with an almost periodic potential : an overview. Lecture Notes in Physics, 153, 1982. J. Bellissard. K-theory of C∗-algebras in Solid State Physics. Lecture Notes in Physics, 257:99 -- 156, 1986. J. Bellissard. Gap labelling theorems for Schrodinger operators. In From number theory to physics (Les Houches, 1989), pages 538 -- 630, 1992. Springer, Berlin. R. Benedetti and J-M Gambaudo. On the dynamics of G-Solenoids. Applications to Delone sets. Ergod. Th. & Dynam. Sys., 29:673 -- 691, 2003. [BHZ00] J. Bellissard, D.J.L. Herrmann, and M. Zarrouati. Hulls of aperiodic solids and gap labeling theorems. CRM Monogr. Ser., 13:207 -- 258, 2000. A.M.S., Providence. [BJ83] S. Baaj and P. Julg. Th´eorie bivariante de Kasparov et op´erateurs non born´es dans les C∗-modules hilbertiens. C. R. Acad. Sci. Paris Sr. I Math., 296 no. 21:875 -- 878, 1983. [BKL01] J. Bellissard, J. Kellendonk, and A. Legrand. Gap-labelling for three- dimensional aperiodic solids. C.R.A.S, serie I, 332:521 -- 525, 2001. [BOO02] M.-T. Benameur and H. Oyono-Oyono. Index theory for quasi- crystals. I. Computation of the gap-label group. 252:137 -- 170, 2002. J. Funct. Anal. [Bre72] G.E. Bredon. Introduction to compact transformation groups. Pure and applied mathematics 46, 1972. Academic Press. [Cha99] J. Chabert. Stabilit´e de la conjecture de Baum-Connes pour certains produit semi-directs de groupes. PhD thesis, Univ. de la M´edit´erran´ee Aix-Marseille II, 1999. [Con79] A. Connes. Sur la th´eorie non commutative de l'int´egration. Lecture Notes in Math., 725:19 -- 143, 1979. Springer, New York. [Con82] A. Connes. A survey of foliations and operator algebras. Proc. Sym- pos. Pure Math., 38 part. 1, 1982. A.M.S. , Providence. [DHK91] R.G. Douglas, S. Hurder, and J. Kaminker. The Longitudinal Cocycle and the Index of Toeplitz Operators. J. Funct. Anal., 101:120 -- 144, 1991. [Gre69] F. P. Greenleaf. applications. Van Nostrand mathematical studies, 16, 1969. Invariant means on topological groups and their 57 [Hat02] A. Hatcher. Algebraic topology. 1st ed., Cambridge University Press, 2002. [HRS05] C. Holton, C. Radin, and L. Sadun. Conjugacies for Tiling Dynam- ical Systems. Comm. Math. Phys., 254:343 -- 359, 2005. [HS87] [Jul81] [Kas88] [Kas95] [Kel95] [KP00] M. Hilsum and G. Skandalis. Morphismes K-orient´es d'espaces de feuilles et fonctorialit´e en th´eorie de Kasparov (d'apr`es une conjec- ture d'A. Connes). Ann. Sci. ´Ecole Norm. Sup. (4), 20 no. 3:325 -- 390, 1987. P. Julg. K-th´eorie ´equivariante et produits crois´es. Note C.R.A.S. Paris, 292:629 -- 632, 1981. G.G. Kasparov. Equivariant KK-theory and the novikov conjecture. Inv. Math., 91:147 -- 201, 1988. G.G. Kasparov. K-theory, group C∗-algebras and higher signatures. London Math. Soc. Lecture Note Ser., 226:101 -- 146, 1995. J. Kellendonk. Noncommutative geometry of tilings and gap labeling. Rev. Math. Phys., 7:1133 -- 1180, 1995. J. Kellendonk and I.F. Putnam. Tilings, C∗-algebras and K-theory. CRM monograph Series 13, 177-206, 2000. M.P. Baake and R.V. Moody Eds., A.M.S., Providence. [KP03] J. Kaminker and I. Putnam. A proof of the gap labeling conjecture. Michigan Math. J., 51:537 -- 546, 2003. [Kuc97] D. Kucerovsky. The KK-product of unbounded modules. K-theory, 11 no. 1:17 -- 34, 1997. [LP03] [Mat] [Mou] J.C. Lagarias and P.A.B. Pleasant. Repetitive Delone sets and qua- sicrystals. Ergod. Th. & Dynam. Sys., 23:831 -- 867, 2003. M. Matthey. K-theories, C∗-algebras and assembly maps. PhD thesis, Universit´e de Neuchatel. H. Moustafa. PV cohomology of pinwheel tilings, their integer group of coinvariants and gap-labelling. http://arxiv.org/abs/0906.2107, to appear in Comm. Math. Phys. [Mou09] H. Moustafa. Gap-labeling des pavages de type pinwheel. PhD thesis, Univ. Blaise Pascal, Clermont-Ferrand, 2009. http://math.univ-bpclermont.fr/∼moustafa/These/these-Moustafa.pdf. [MS06] C. C. Moore and C. Schochet. Global analysis on foliated spaces. MSRI Publications, 9, 2006. [ORS02] N. Ormes, C. Radin, and L. Sadun. A homeomorphism invariant for substitution tiling spaces. Geometriae Dedicata, 90:153 -- 182, 2002. [PB09] J. Pearson and J. Bellissard. Noncommutative Riemannian Geometry and Diffusion on Ultrametric Cantor Sets. Journal of Noncommuta- tive Geometry, 3:847 -- 865, 2009. 58 [Ped79] [Pet05] G. K. Pederson. C∗-algebras and their automorphism groups. London Math. Society Monographs, 14, 1979. Academic Press, London. S. Petite. Pavages du demi-plan hyperbolique et laminations. PhD thesis, Univ. de Bourgogne I, 2005. [Rad94] C. Radin. The pinwheel tilings of the plane. Ann. of Math., 139:661 -- 702, 1994. [Rad95] C. Radin. Space tilings and substitutions. Geom. Dedicata, 55:257 -- 264, 1995. [Rie82] M.A. Rieffel. Morita equivalence for operator algebras. Proc. of Sym- posia in Pure Math., 38:285 -- 298, 1982. [RS98] C. Radin and L. Sadun. An algebraic invariant for substitution tiling systems. Geom. Dedicata, 73:21 -- 37, 1998. [Sad] L. Sadun. Private conversation in september 2007. [SBGC84] D. Shechtman, I. Blech, D. Gratias, and J.V. Cahn. Metallic phase with long range orientational order and no translational symmetry. Phys. Rev. Lett., 53:1951 -- 1953, 1984. [Seg68] [Ska91] [Spa66] [Vas01] [vE94] [Ypm] G. Segal. Equivariant K-theory. Math., 34:129 -- 151, 1968. Inst. Hautes ´Etudes Sci. Publ. G. Skandalis. Kasparov's bivariant K-theory and applications. Ex- position. Math., 9:193 -- 250, 1991. E.H. Spanier. Algebraic topology. McGraw-Hill series in higher math- ematics, 1966. S. Vassout. Feuilletages et r´esidu non commutatif longitudinal. PhD thesis, Univ. Pierre et Marie Curie - Paris VI, 2001. A. van Elst. Gap labelling theorems for Schrodınger operators on the square and cubic lattices. Rev. Math. Phys., 6:319 -- 342, 1994. F. Ypma. Quasicrystals, C∗-algebras and K-theory. http://remote.science.uva.nl/ npl/fonger.pdf (2004). Univ. Blaise Pascal, Clermont-Ferrand, FRANCE E-mail address : [email protected] URL : http://math.univ-bpclermont.fr/∼moustafa 59
1308.5084
2
1308
2015-04-07T11:28:52
Classifying crossed product C*-algebras
[ "math.OA" ]
I combine recent results in the structure theory of nuclear C*-algebras and in topological dynamics to classify certain types of crossed products in terms of their Elliott invariants. In particular, transformation group C*-algebras associated to free minimal Z^d-actions on the Cantor set with compact space of ergodic measures are classified by their ordered K-theory. In fact, the respective statement holds for finite dimensional compact metrizable spaces, provided that projections of the crossed products separate tracial states. Moreover, C*-algebras associated to certain minimal homeomorphisms of odd dimensional spheres are only determined by their spaces of invariant Borel probability measures (without a condition on the space of ergodic measures). Finally, I show that for a large collection of classifiable C*-algebras, crossed products by Z^d-actions are generically again classifiable.
math.OA
math
CLASSIFYING CROSSED PRODUCT C∗-ALGEBRAS WILHELM WINTER Abstract. I combine recent results in the structure theory of nuclear C∗- algebras and in topological dynamics to classify certain types of crossed prod- ucts in terms of their Elliott invariants. In particular, transformation group C∗-algebras associated to free minimal Zd-actions on the Cantor set with com- pact space of ergodic measures are classified by their ordered K-theory. In fact, the respective statement holds for finite dimensional compact metriz- able spaces, provided that projections of the crossed products separate tracial states. Moreover, C∗-algebras associated to certain minimal homeomorphisms of spheres S 2n+1 are only determined by their spaces of invariant Borel proba- bility measures (without a condition on the space of ergodic measures). Finally, I show that for a large collection of classifiable C∗-algebras, crossed products by Zd-actions are generically again classifiable. 0. Introduction The structure and classification theory of nuclear C∗-algebras has seen substan- tial progress in recent years, largely spurred by the interplay of certain topological, algebraic and homological regularity properties. These allow for some amount of in- terpretation, but generally arise as finite topological dimension, tensorial absorption of certain touchstone C∗-algebras and order completeness of homological invariants, cf. [5] and [35]. When Andrew Toms and I began to study the tight connections between finite decomposition rank (cf. [14]), Z-stability (cf. [11, 24]) and strict comparison (cf. [23]), this was for very specific (and, to some extent artificial) classes of simple inductive limits (see [31]). However, the three properties soon turned out to occur (or to fail) simultaneously in much broader generality. We then conjectured that they are equivalent for all separable, simple, unital, nonelementary, nuclear C∗- algebras which are finite (hence admit a tracial state). With the introduction of nuclear dimension in [40], this point of view -- and the regularity conjecture -- became available also in a not necessarily finite setting. Of the implications in the regularity conjecture, the problem of when Z-stability implies finite nuclear dimension or even finite decomposition rank arguably remains the most intriguing. While by now there are several results establishing such impli- cations, most of these factorize through classification theorems of some sort. In [29], a more direct argument (based on [13]) was given; this does not use a classification result in any way and in fact it even works in a not necessarily simple situation, Date: September 25, 2018. 2010 Mathematics Subject Classification. 46L35, 47L65, 46L55, 37B05. Key words and phrases. C∗-algebra, crossed product, classification, dynamical system. Supported by EPSRC Grant EP/I019227/1, DFG SFB 878, GIF Grant 1137-30.6/2011. 1 2 WILHELM WINTER but it requires the algebra to be locally homogeneous -- a rather strong structural hypothesis. In [21], Matui and Sato obtained a very convincing result in the simple, nuclear, monotracial case: In this situation, Z-stability implies finite decomposition rank provided the algebra is also quasidiagonal. They moreover showed that in the trace- less case, Z-stability implies finite nuclear dimension. These results are in line with Kirchberg -- Phillips classification of purely infinite C∗-algebras (the traceless, finite nuclear dimension case), and with Lin's classification of TAF algebras (these have traces, and are always quasidiagonal). Of course, it remains an important open question whether tracial C∗-algebras with finite nuclear dimension are automati- cally quasidiagonal. In this paper, I am interested in applying these ideas to the classification prob- lem for simple, nuclear C∗-algebras, following Elliott's program to classify nuclear C∗-algebras by K-theoretic invariants. The problem is particularly relevant for so- called transformation group C∗-algebras, i.e., C∗-algebras associated to topological dynamical systems via the crossed product construction. These are the perhaps most natural source of examples, and they are important invariants of dynami- cal systems as has been impressively demonstrated in [9] (for Cantor minimal Z- actions). Also, for such crossed products there are highly useful tools for computing the invariant, in particular the Pimsner -- Voiculescu exact sequence. For minimal actions on finite dimensional compact spaces, their C∗-algebras were successfully classified by ordered K-theory in [32] (see also [30]), at least if projec- tions separate tracial states (e.g., in the uniquely ergodic situation). While this is already quite satisfactory, it begs to be generalized in two directions: First, what if projections do not separate tracial states? This is interesting even when there are no nontrivial projections and only finitely many extremal traces -- and such examples indeed do exist, like C∗-algebras of minimal homeomorphisms of odd dimensional spheres, as considered in [4] and [33] (those examples rely on the fast approximation method of [6], first introduced by Anosov and Katok). Second, what about more general group actions? Even Zd-actions on Cantor sets are notoriously difficult to handle in this respect (but cf. [7, 8]). One of the reasons is that, on the C∗-side, analogues of Putnam's orbit breaking subalgebras are not easy to find, let alone to use. Below I combine recent results on the structure of dynamical systems and crossed products with a new technique (see Theorem 2.2) to make progress on both of these questions. This new method reduces the problem of showing that a C∗- algebra is classifiable to the problem of embedding it into a classifiable C∗-algebra in a sufficiently nice fashion. (Here, by 'classifiable' I mean TAF, or TAI, in the sense of Lin -- for the crossed products in question we do not have to worry about the Universal Coefficient Theorem.) I am confident that this method will prove to be as useful for classification as the one introduced in [38] (where classification up to Z-stability was reduced to classification up to UHF-stability), see also [19, 17]. Apart from the aforementioned Theorem 2.2, the main ingredients are a recent result of Szab´o, where he establishes finite Rokhlin dimension (in the sense of [10]) of free Zd-actions of finite dimensional spaces, a recent result of Strung which ensures that certain transformation group C∗-algebras nicely embed into classifiable models, CLASSIFYING CROSSED PRODUCTS 3 and a result of Lin which establishes a strong form of AF-embeddability of certain types of crossed product C∗-algebras. As byproducts I obtain a simple proof of a special case of [21, Theorem 6.1] and show that, for certain classifiable C∗-algebras, generic sets of Zd-actions give rise to crossed products which are again classifiable. Section 1 recalls a characterization of TAS algebras (with S a suitable class of building blocks) which will be useful for our purposes. It also summarizes some of the relevant results on the classification of rationally TAF (or TAI) C∗-algebras. Section 2 contains the main technical result, along with a Corollary which illustrates the 'classification by embeddings' method in a somewhat more concise (and less technical) manner. We also obtain a short proof of a special case of a recent result of Matui and Sato along these lines. In Section 3 Theorem 2.2 is applied to transformation group C∗-algebras of free and minimal Zd-actions. The result in particular covers free and minimal Cantor Zd-actions with compact space of ergodic measures; the C∗-algebras in this case are classified by their ordered K-theory. In Section 4 it is shown that C∗-algebras associated to minimal homeomorphisms of spheres (of odd dimension at least 3) obtained by the fast approximation method of [6] are classified by their spaces of invariant Borel probability measures. Finally, Section 5 shows that (in a suitable context) classifiability generically passes to crossed products by Zd-actions. I would like to thank the referee for carefully proofreading the paper and for a number of helpful comments and suggestions. 1. TAS algebras and classification 1.1 For convenience, let us recall the following characterization of rationally TAS algebras from [27, Lemma 1.2] (see also [34, Lemma 3.2]). By Q we denote the universal UHF C∗-algebra. Let S be a class of separable, unital C∗-algebras which can be Proposition: finitely presented with weakly stable relations, which is closed under taking direct sums and which contains all finite dimensional C∗-algebras. Let A be a separable, simple, unital, stably finite, exact C∗-algebra. Then, A ⊗ Q is TAS if and only if the following holds: There is η > 0 such that, for any ǫ > 0 and any finite subset F ⊂ A ⊗ Q, there are a projection p ∈ A ⊗ Q and a unital C∗-subalgebra B ⊂ p(A ⊗ Q)p, B ∈ S, such that (i) kpb − bpk < ǫ for all b ∈ F , (ii) dist(pbp, B) < ǫ for all b ∈ F , (iii) τ (p) > η for all τ ∈ T (A ⊗ Q). 1.2 We will mainly be interested in the cases where S is the class of finite dimen- sional C∗-algebras or the class of interval algebras, i.e., in TAF and TAI algebras, respectively. For such algebras, if they are in addition nuclear and satisfy the Uni- versal Coefficient Theorem (UCT), the Elliott invariant is complete. Moreover, we know the range of the invariant in these situations. We summarize the relevant results from [15, 17] for the reader's convenience. 4 WILHELM WINTER Theorem: Let Ai, i = 0, 1, be separable, simple, unital, nuclear C∗-algebras which satisfy the UCT. Suppose A1 and A2 are TAI. Then, A1 ∼= A2 if and only if their Elliott invariants (K0(Ai), K0(Ai)+, [1Ai], K1(Ai), T (Ai), rAi : T (Ai) → S(K0(Ai))) are isomorphic, and every isomorphism of invariants lifts to a ∗-isomorphism of algebras. Moreover, the Ai are approximately subhomogeneous (ASH) algebras of topolog- ical dimension (and hence decomposition rank) at most 2, and are approximately homogeneous (AH) of topological dimension at most 3. Finally, they are in fact TAF if and only if they have real rank zero, and in this case the classifying invariant degenerates to ordered K-theory, (K0(Ai), K0(Ai)+, [1Ai], K1(Ai)). 1.3 The previous theorem illustrates that classification in term of K-theory works best in the case of real rank zero, i.e., with an abundance of projections around. In [38] it was shown that classification up to UHF-stability will yield classification up to Z-stability, where Z denotes the Jiang -- Su algebra, cf. [11, 24]. This is useful when there are at least enough projections to distinguish traces, since then tensoring with UHF algebras will enforce real rank zero. In this case it only remains to confirm Z-stability, for which we have all sorts of highly useful criteria (for example finite decomposition rank). We summarize the situation as follows. Theorem: Let A be a separable, simple, unital C∗-algebra with finite decomposi- tion rank. Then, conditions (i) -- (iv) below are equivalent: (i) The canonical map T (A) → S(K0(A)) is a homeomorphism. (ii) K0(A) separates the tracial states of A. (iii) A is rationally TAF, i.e., A ⊗ Q is TAF. (iv) A is rationally of real rank zero, i.e., A ⊗ Q has real rank zero. If, moreover, A satisfies the UCT, under any of these conditions A is ASH of topological dimension at most 2, and such C∗-algebras are classified by their ordered K-theory. 2. From approximate tracial embeddings to tracial approximations In this section we introduce and illustrate the 'classification by embedding' method. We start with a technical result that allows us to compare order zero maps in terms of traces. Recall that a completely positive contractive (c.p.c.) map is order zero if it preserves orthogonality. There is a structure theorem for such maps which in particular yields a notion of functional calculus, see [39] for a detailed exposition. In the sequel, we will encounter matrix algebras of different sizes. We will usually write matrix units for these in the form emn, without distinguishing between the matrix sizes; this should cause no confusion. CLASSIFYING CROSSED PRODUCTS 5 2.1 Proposition: comparison. Let F be a finite dimensional C∗-algebra and let Let A be a separable, simple, unital C∗-algebra with strict ϕ : F → A, ϕi : F → A, i ∈ N, be c.p.c. order zero maps such that, for each x ∈ F+ and f ∈ C0((0, 1])+, sup{τ (f (ϕ)(x) − f (ϕi)(x)) τ ∈ T (A)} i→∞−→ 0 and lim sup kf (ϕi)(x)k ≤ kf (ϕ)(x)k. i Then, there are such that for each y ∈ F+ si ∈ (M4 ⊗ A)1, i ∈ N, (1) and (2) and that (3) and (4) ksi(14 ⊗ ϕ(y)) − (e11 ⊗ ϕi(y))sik i→∞−→ 0 k(e11 ⊗ ϕi(y))sis∗ i − (e11 ⊗ ϕi(y))k i→∞−→ 0, sis∗ i ∈ (e11 ⊗ ϕi(1F ))M4 ⊗ A(e11 ⊗ ϕi(1F )) s∗ i si ∈ (14 ⊗ ϕ(1F ))M4 ⊗ A(14 ⊗ ϕ(1F )). Proof: We first show the statement in the case F = C. When checking (1) and (2) it will clearly suffice to consider y = 1C; let us write h and hi for ϕ(1C) and ϕi(1C), respectively. For the moment let us fix L ∈ N and ǫ > 0. Define functions d(c) l , d(c) l , d(c) l ∈ C0((0, 1])1 + for c ∈ {0, 1}, l ∈ {1, . . . , L}, with the following properties: L , c ∈ {0, 1}, l ∈ {1, . . . , L} (a) k d(c) l (b) d(c) (c) d(c) (d) d(c) d(c) l k < 1 id(0,1] − l L d(c) l = d(c) , c ∈ {0, 1}, l ∈ {1, . . . , L} d(c) l′ = 0, c ∈ {0, 1}, l 6= l′ ∈ {1, . . . , L} l − ǫ)+, c ∈ {0, 1}, l ∈ {1, . . . , L} l = (d(c) l l l (e) kid(0,1] −P1 c=0PL For c ∈ {0, 1}, set l=1 l L · d(c) l k ≤ 1 L . N (c) := {l ∈ {1, . . . , L} d(c) l (h) 6= 0} and × := {1, . . . , L} \ N (c). One checks that there is ¯ı ∈ N such that for each i ≥ ¯ı N (c) τ ( d(c) l (hi)) ≤ 3 2 τ ( d(c) l (h)) 6 WILHELM WINTER for all τ ∈ T (A), c ∈ {0, 1}, l ∈ N (c) and k d(c) l (hi)k < ǫ for c ∈ {0, 1}, l ∈ N (c) × . But then dτ (d(c) l (hi)) ≤ τ ( d(c) l (hi)) ≤ 3 2 τ ( d(c) l (h)) ≤ 3 2 dτ ( d(c) l (h)) for i ≥ ¯ı, τ ∈ T (A), c ∈ {0, 1}, l ∈ N (c). Now by comparison, hd(c) l (hi)i ≤ 2 · h d(c) l (h)i in Cu(A) for i ≥ ¯ı, c ∈ {0, 1}, l ∈ N (c). By the Kirchberg -- Rørdam Lemma [12, Lemma 2.2] there are s(c) i,l ∈ (M2 ⊗ A)1 such that and s(c) i,l s(c)∗ i,l = e11 ⊗ d(c) l (hi) s(c)∗ i,l s(c) i,l ∈ her(12 ⊗ d(c) l (h)) ⊂ M2 ⊗ A for i ≥ ¯ı, c ∈ {0, 1}, l ∈ N (c). Set s(l) i,l := 0 for i ≥ ¯ı, c ∈ {0, 1}, l ∈ N (c) × . For i ≥ ¯ı, c ∈ {0, 1} define l=1 s(c) i,l ∈ M2 ⊗ A, ks(c) i s(c)∗ l=1 e11 ⊗ d(c) l (hi)k ≤ ǫ s(c) i :=PL i −PL i −PL l=1 c=0 s(c) i s(c)∗ i k(e11 ⊗ hi)s(c) i s(c)∗ l L · e11 ⊗ d(c) l (hi)k ≤ 2ǫ + 1/L, k(e11 ⊗ hi)(P1 ) − (e11 ⊗ hi)k ≤ 3ǫ + 3/L then and whence for i ≥ ¯ı. Moreover, ks(c) i (12 ⊗ h) − (e11 ⊗ hi)s(c) i k l=1 s(c) = kPL ≤ kPL ≤ 1/L + 2ǫ l i,l (12 ⊗ h) −PL i,l −PL L · s(c) l=1 l=1(e11 ⊗ hi)s(c) i,l k l=1(e11 ⊗ hi)s(c) i,l k + 1/L for c ∈ {0, 1}, i ≥ ¯ı. Now define then c=0 e1,c+1 ⊗ s(c) i ∈ M2 ⊗ M2 ⊗ A, c=0 s(c) i s(c)∗ i ≤ 1M2⊗M2⊗A si :=P1 i = e11 ⊗P1 sis∗ CLASSIFYING CROSSED PRODUCTS 7 and k(e11 ⊗ e11 ⊗ hi)sis∗ i − (e11 ⊗ e11 ⊗ hi)k ≤ 3ǫ + 3/L for i ≥ ¯ı. Next, we check ksi(12 ⊗ 12 ⊗ h) − (e11 ⊗ e11 ⊗ hi)sik c=0 e1,c+1 ⊗ (s(c) i (12 ⊗ h)) c=0 e1,c+1 ⊗ ((e11 ⊗ hi)s(c) i )k = kP1 −P1 ≤ 2/L + 4ǫ for i ≥ ¯ı. Now if we let L go to infinity and ǫ to zero, a diagonal sequence argument yields satisfying and si ∈ (M2 ⊗ M2 ⊗ A)1, i ∈ N, ksi(12 ⊗ 12 ⊗ h) − (e11 ⊗ e11 ⊗ hi)sik i→∞−→ 0 k(e11 ⊗ e11 ⊗ hi)sis∗ i − e11 ⊗ e11 ⊗ hik i→∞−→ 0. Upon replacing si with (e11 ⊗ h We have thus verified the proposition when F = C. )si(14 ⊗ h i+1 i 1 1 i+1 ) if necessary, we get (3) and (4). Next suppose F = Mk for some k ∈ N. Let π : Mk −→ A∗∗, πi : Mk −→ A∗∗ be supporting ∗-homomorphisms for ϕ and for the ϕi, respectively; cf. [39]. Run the proposition for C ∼= e11Mke11 and for ϕe11Mk e11 and ϕie11Mk e11 ; denote the resulting elements of (M4 ⊗ A)1 by si. The 'amplified' elements m=1(e11 ⊗ πi(em1))si(14 ⊗ π(e1m)) ∈ (M4 ⊗ A)1 si :=Pk will then satisfy (1) through (4) above. This verifies the proposition in the case F = Mk. When F is a sum of N matrix algebras, run the proposition for each matrix ∈ (M4 ⊗ A)1. By (3) and (4), the )∗ are pairwise orthogonal for each fixed i, and . Therefore we can define elements block separately to obtain elements s(1) elements s(1) i i )∗s(1) the same goes for the (s(1) )∗, . . . , s(N ) , . . . , s(N ) )∗s(N ) (s(N ) i (s(1) i i i i i , . . . , (s(N ) i + . . . + s(N ) i i i si := s(1) ∈ (M4 ⊗ A)1. These will again satisfy (1) through (4), thus verifying the proposition for an arbi- trary finite dimensional C∗-algebra F . 2.2 Theorem: Let S be a class of separable, unital C∗-algebras which can be finitely presented with weakly stable relations. Suppose further that S is closed under taking direct sums and under taking tensor products with finite dimensional C∗-algebras, and that S contains all finite dimensional C∗-algebras. 8 WILHELM WINTER Let A be a separable, simple, unital C∗-algebra with dimnuc A < ∞ and T (A) 6= ∅, and let (cid:0)A σi−→ Bi i−→ A(cid:1)i∈N be a system of maps with the following properties: (i) Bi ∈ S, i ∈ N (ii) i is an embedding for each i ∈ N (iii) σi is c.p.c. for each i ∈ N (iv) ¯σ : A →QN Bi/LN Bi induced by the σi is a unital ∗-homomorphism (v) sup{τ (iσi(a) − a) τ ∈ T (A)} i→∞−→ 0 for each a ∈ A. Then, A ⊗ Q is TAS. Proof: Let (Fj = F (0) j ⊕ . . . ⊕ F (m) j , ψj, ϕj )j∈N be a system of m-decomposable c.p. approximations for A with c.p.c. approximately order zero maps ψj and c.p.c. order zero maps ϕ(l) as in [36, Proposi- tion 4.2]; see also [40, Proposition 4.3]. (Here, 'approximately order zero' means that for a, b ∈ A with ab = 0 we have ψj(a)ψj(b) −→ 0 as j goes to infinity.) j = ϕjF (l) j By weak stability of order zero maps, for each j ∈ N there are c.p.c. order zero maps such that j,i : F (l) ϕ(l) j → Bi, i ∈ N, l ∈ {0, . . . , m}, k ϕ(l) j,i(x) − σiϕ(l) j (x)k i→∞−→ 0, x ∈ F (l) j . Now for each x ∈ (F (l) j )+ and f ∈ C0((0, 1])+, we have hence and kf ( ϕ(l) j,i)(x) − σif (ϕ(l) j )(x)k i→∞−→ 0, sup τ ∈T (A) τ (f (i ϕ(l) j,i)(x) − f (ϕ(l) j )(x)) i→∞−→ 0. kif ( ϕ(l) j,i)(x) − iσif (ϕ(l) j )(x)k i→∞−→ 0. Since the i are embeddings, the σi are (eventually) nonzero and approximately multiplicative, we see that lim sup i kiσif (ϕ(l) j )(x)k = kf (ϕ(l) j )(x)k, whence lim sup i kf (i ϕ(l) j,i)(x)k ≤ kf (ϕ(l) j )(x)k for x ∈ (F (l) j )+, f ∈ C0((0, 1])+. By Proposition 2.1, there are s(l) j,i ∈ (M4 ⊗ A)1, i ∈ N, such that and ks(l) j,i(14 ⊗ ϕ(l) j (x)) − (e11 ⊗ i ϕ(l) j,i(x))s(l) j,ik i→∞−→ 0 k(e11 ⊗ i ϕ(l) j,i(x))s(l) j,i s(l)∗ j,i − e11 ⊗ i ϕ(l) j,i(x)k i→∞−→ 0 CLASSIFYING CROSSED PRODUCTS 9 for each x ∈ F (l) j . We obtain contractions s(l) j ∈ (M4 ⊗ A)∞ ∼= M4 ⊗ A∞ with and where j (14 ⊗ ιϕ(l) s(l) j (x)) = (e11 ⊗ ¯¯σϕ(l) j (x))s(l) j (e11 ⊗ ¯¯σϕ(l) j (x))s(l) j s(l)∗ j = e11 ⊗ ¯¯σϕ(l) j (x), is the ∗-homomorphism induced by the i and ¯ :Q Bi/L Bi → A∞ is the canonical embedding. Let ι : A → A∞ ¯ι : A∞ → (A∞)∞ be induced by the canonical embedding i.e., Let ι : A → A∞, [(aj)j∈N] ¯ι7→ [(ι(aj ))j∈N]. ¯γ : A∞ → (A∞)∞ be the ∗-homomorphism induced by ¯¯σ : A → A∞. Let and j → A∞ j /Lj F (l) ¯ϕ(l) :Qj F (l) ¯ψ(l) : A →Qj F (l) j and ψ(l) j /Lj F (l) j j , respectively; these will automatically be be the maps induced by the ϕ(l) c.p.c. order zero. Define then and ¯s(l) := [(s(l) j )j∈N] ∈ (M4 ⊗ A∞)∞ ∼= M4 ⊗ (A∞)∞, ¯s(l)(14 ⊗ ¯ι ¯ϕ(l) ¯ψ(l)(a)) = (e11 ⊗ ¯γ ¯ϕ(l) ¯ψ(l)(a))¯s(l) (e11 ⊗ ¯γ ¯ϕ(l) ¯ψ(l)(a))¯s(l) ¯s(l)∗ = e11 ⊗ ¯γ ¯ϕ(l) ¯ψ(l)(a). Note that for each a ∈ A ¯ϕ(l) ¯ψ(l)(1A)ι(a) = ¯ϕ(l) ¯ψ(l)(a) by [36, Proposition 4.2], and so in particular ( ¯ϕ(l) ¯ψ(l)(1A)) 1 2 ι(a) ∈ C∗( ¯ϕ(l) ¯ψ(l)(b) b ∈ A), 10 WILHELM WINTER which in turn implies that ¯s(l)(14 ⊗ (¯ι ¯ϕ(l) ¯ψ(l)(1A)) 1 2 )(14 ⊗ ¯ιι(a)) = (e11 ⊗ ¯γ(ι(a))(¯γ ¯ϕ(l) ¯ψ(l)(1A)) = (e11 ⊗ ¯γ(ι(a)))(e11 ⊗ (¯γ ¯ϕ(l) ¯ψ(l)(1A)) 2 )¯s(l) 1 1 2 )¯s(l). l=1 e1,l ⊗ ((e11 ⊗ (¯γ ¯ϕ(l) ¯ψ(l)(1A)) l=1 e1,l ⊗ (¯s(l)(14 ⊗ (¯ι ¯ϕ(l) ¯ψ(l)(1A)) 1 2 )¯s(l)) 1 2 )) ∈ Mm+1 ⊗ M4 ⊗ (A∞)∞, Set ¯v then := Pm+1 = Pm+1 ¯v¯v∗ =Pm+1 l=1 e11 ⊗ e11 ⊗ ¯γ ¯ϕ(l) ¯ψ(l)(1A) = e11 ⊗ e11 ⊗ ¯γ(1A∞ ), so in particular ¯v is a partial isometry. Moreover, we check that for a ∈ A ¯v(1m+1 ⊗ 14 ⊗ ¯ιι(a)) l=1 e1,l ⊗ ((e11 ⊗ ¯γ(ι(a)))(e11 ⊗ (¯γ ¯ϕ(l) ¯ψ(l)(1A)) 1 2 )¯s(l)) = Pm+1 = (e11 ⊗ e11 ⊗ ¯γ(ι(a)))¯v, whence for a ∈ A ¯v∗¯v(1m+1 ⊗ 14 ⊗ ¯ιι(a)) = ¯v∗(e11 ⊗ e11 ⊗ ¯γ(ι(a)))¯v = (1m+1 ⊗ 14 ⊗ ¯ιι(a))¯v∗¯v in Mm+1 ⊗ M4 ⊗ (A∞)∞. Now for every finite subset F ⊂ A1 + and ǫ > 0 there are i ∈ N and v ∈ Mm+1 ⊗ M4 ⊗ A such that (a) vv∗ = e11 ⊗ e11 ⊗ i(1Bi ) (b) (trMm+1⊗M4 ⊗ τ )(vv∗) ≥ (c) k[v∗v, 1m+1 ⊗ 14 ⊗ a]k < ǫ for all a ∈ F (d) kv∗v(1m+1 ⊗ 14 ⊗ a) − v∗(e11 ⊗ e11 ⊗ iσi(a))vk < ǫ for all a ∈ F . 2(m+1)4 for all τ ∈ T (A) 1 Define by κ : Bi → Mm+1 ⊗ M4 ⊗ A κ(b) := v∗(e11 ⊗ e11 ⊗ i(b))v, then κ is an embedding such that (e) (trMm+1⊗M4 ⊗ τ )(1κ(Bi)) ≥ (f) k[1κ(Bi), 1m+1 ⊗ 14 ⊗ a]k < ǫ for all a ∈ F (g) 1κ(Bi)(1m+1 ⊗ 14 ⊗ a)1κ(Bi) ∈ǫ κ(Bi) for all a ∈ F . 2(m+1)4 for all τ ∈ T (A) 1 Now by Proposition 1.1, A ⊗ Q is TAS. 2.3 We note a Corollary which nicely illustrates the way in which we are going to use Theorem 2.2 towards classification results in the subsequent sections. Our method reduces the problem of showing that a C∗-algebra A is classifiable to the problem of embedding it into a classifiable (TAF, or TAI) C∗-algebra B such that the embedding induces an isomorphism at the level of invariants. Then, one needs to lift the inverse of this isomorphism to an embedding of the classifiable model CLASSIFYING CROSSED PRODUCTS 11 into the original algebra. Theorem 2.2 now moves the image of the composition of these two embeddings into a position compatible with the TAF (or TAI) condition for A. There are quite a few tools available for finding such embeddings: In our applica- tions, the map from A to B will usually come from properties related to quasidiag- onality. The source of the map from B to A depends on the situation; for example, it might come from the existence theorem of [22]. Although in our applications we are not in the exact situation of the Corollary, it will serve as a blueprint for tracially approximate versions which essentially follow the same pattern, but which do not require keeping track of the entire K-theoretic information of A (see 2.4, 3.1 and 4.2 below). Recall from [22, Section 3.1] that for a unital C∗-algebra D the ordered semigroup Cu∼(D) is given by formal differences x − n · [1D], with x ∈ Cu(D) and n ∈ N. More precisely, Cu∼(D) is the quotient of Cu(D) × N by the equivalence relation given by (x, n) ∼ (y, m) :⇔ x + m · [1D] + k · [1D] = y + n · [1D] + k · [1D] in Cu(D) for some k ∈ N. Corollary: Let A and B be separable, simple, unital C∗-algebras. Suppose that dimnuc A < ∞ and that B is an AI algebra. Suppose there is a unital embedding σ : A → B inducing an isomorphism between the Elliott invariants (K0(A), K0(A)+, [1A], K1(A), T (A), rA : T (A) → S(K0(A))) and (K0(B), K0(B)+, [1B], K1(B), T (B), rB : T (B) → S(K0(B))). Then, A ⊗ Q is TAI. Proof: From the Kunneth Theorem it is clear that σ ⊗ idQ also induces an isomor- phism between the Elliott invariants of A ⊗ Q and of B ⊗ Q. Under the conditions on A ⊗ Q and B ⊗ Q (they are both simple, unital, nuclear, Z-stable, and have nonempty tracial state spaces, hence stable rank one), it follwos from [1] that the Cuntz semigroups are determined in a natural way by the Elliott invariants, so that the isomorphim between the latter induces one between the former. But since also the classes of the units are preserved by this isomorphism, it then also induces an isomorphism Cu∼(σ ⊗ idQ) between Cu∼(A ⊗ Q) and Cu∼(B ⊗ Q). Since B ⊗ Q is AI and A ⊗ Q has stable rank one by [23], it follows from [22, Theorem 1] that the inverse of this isomorphism lifts to a unital embedding : B ⊗ Q → A ⊗ Q. Let be a system of maps such that the Bi are interval algebras, the ψi are approximately multiplicative, the ϕi are embeddings and ϕiψi → idB⊗Q in point norm topology. (cid:0)B ⊗ Q ψi−→ Bi ϕi−→ B ⊗ Q(cid:1)i∈N For i ∈ N define and σi : A ⊗ Q → Bi i : Bi → A ⊗ Q 12 by and WILHELM WINTER σi := ψi ◦ (σ ⊗ idQ) i := ◦ ϕi. It is now straightforward to check that the system (cid:0)A ⊗ Q σi−→ Bi i−→ A ⊗ Q(cid:1)i∈N satisfies the hypotheses of Theorem 2.2, whence A ⊗ Q ⊗ Q ∼= A ⊗ Q is TAI. 2.4 As a first incidence of a tracially approximate version of 2.3, in the monotracial situation we rediscover a special case of a striking recent result of Matui and Sato, [21]. Our version is less general since we need to assume finite nuclear dimension; on the other hand, the proof is substantially simpler since it avoids the von Neumann algebra techniques pivotal for [21] (see also [20]). We will see in the subsequent sections that our approach has the additional advantage that it applies in the situation of more general trace spaces. Corollary: Let A be a separable, simple, unital, monotracial C∗-algebra with dimnuc A < ∞. Suppose that A is quasidiagonal. Then, A ⊗ Q is TAF. Proof: By [3], A ⊗ Q is NF, so there is a unital embedding for a suitable sequence (ni)i∈N ⊂ N. Since A ⊗ Q is nuclear, there is a sequence of c.p.c. maps ¯σ : A ⊗ Q →Qi Mni/Li Mni σi : A ⊗ Q → Mni, i ∈ N, σ : A ⊗ Q →Yi Mni, σ(a) = (σi(a))i∈N, such that given by lifts ¯σ. Let be a unital embedding, then clearly ρi : Mni → 1A ⊗ Q ⊂ A ⊗ Q τ (ρiσi(a) − a) i→∞−→ 0 for each a ∈ A ⊗ Q, where τ denotes the unique trace on A ⊗ Q. Now Theorem 2.2 yields that A ⊗ Q ⊗ Q ∼= A ⊗ Q is TAF. 3. Free, minimal Zd-actions We now combine the method of the previous section with results of Lin and of Szab´o to obtain our classification result for crossed products by Zd-actions. 3.1 Theorem: Let A be a separable, simple, unital C∗-algebra with dimnuc A < ∞ and such that A ⊗ Q has real rank zero; suppose the extreme boundary of the tracial state space of A, ∂eT (A), is nonempty and compact. CLASSIFYING CROSSED PRODUCTS 13 Suppose further that for each τ ∈ ∂eT (A) there are a simple, unital, monotracial AF algebra D with trace δ and a unital embedding with α : A → D δ ◦ α = τ. Then, A ⊗ Q is TAF. Proof: We may clearly replace A by A ⊗ Q. For the moment fix F ⊂ (A ⊗ Q)1 + finite and ǫ > 0. Choose a finite partition of unity (hλ)Λ for ∂eT (A) such that for each λ ∈ Λ there is τλ ∈ supp (hλ) such that τ ⊗ τQ(a) − τλ ⊗ τQ(a) ≤ ǫ for τ ∈ supp (hλ), a ∈ F . Choose 0 < η < ǫ Λ . Since A has real rank zero, the image of the set of projections of A in C(∂eT (A))1 + is dense; using comparison one easily checks that there are pairwise orthogonal projections such that for λ ∈ Λ, τ ∈ ∂eT (A). pλ ∈ A, λ ∈ Λ, τ (pλ) − hλ(τ ) < η For each λ ∈ Λ, find Dλ, δλ, αλ as in the hypotheses, i.e., each Dλ is simple, unital, AF with unique trace δλ, and αλ : A ⊗ Q → Dλ is a unital embedding with δλ ◦ αλ = τλ ⊗ τQ, λ ∈ Λ. Find matrix algebras Mrλ,i, i ∈ N, and approximately multiplicative u.c.p. maps choose unital embeddings Note that for each a ∈ A ⊗ Q βλ,i : Dλ → Mrλ,i; γλ,i : Mrλ,i → Q. Next define τQ ◦ γλ,i ◦ βλ,i ◦ αλ(a) i→∞−→ τλ ⊗ τQ(a). ¯Bi :=Lλ∈Λ Mrλ,i, ¯σi : A ⊗ Q → ¯Bi, ¯σi := ⊕λβλ,i ◦ αλ, ¯ρi : ¯Bi → A ⊗ Q, ¯ρi := ⊕λpλ ⊗ γλ,i. 14 WILHELM WINTER We check for a ∈ F and τ ∈ ∂eT (A) τ ⊗ τQ(¯ρi ¯σi(a) − a) = Pλ τ (pλ) · τQ ◦ γλ,i ◦ βλ,i ◦ αλ(a) − hλ(τ ) · (τ ⊗ τQ)(a) i→∞−→ Pλ τ (pλ) · (τλ ⊗ τQ)(a) − hλ(τ ) · (τ ⊗ τQ)(a) ≤ Pλ hλ(τ ) · (τλ ⊗ τQ)(a) − (τ ⊗ τQ)(a) + Λ · η ≤ Pλ hλ(τ ) · ǫ + Λ · η ≤ 2 · ǫ. Making F bigger and ǫ smaller will now produce (A ⊗ Q σi−→ Bi ρi−→ A ⊗ Q)i∈N as required for Theorem 2.2. 3.2 Corollary: Let X be a compact metrizable space with finite covering dimen- sion and β : Zd → Homeo(X) a free, minimal action with compact space of ergodic measures. Suppose that K0(C(X)⋊Zd) separates the traces of C(X)⋊Zd (this is automatically satisfied if X is a Cantor set). Then, (C(X)⋊ Zd)⊗Q is TAF. As a consequence, the crossed products themselves are classified by their ordered K-theory. Proof: In [28], Szab´o shows that dimnuc(C(X) ⋊ Zd) < ∞; C(X) ⋊ Zd is simple since the action is minimal, hence Z-stable by [36, Corollary 6.3]. Also, traces are separated by projections, and therefore (C(X) ⋊ Zd) ⊗ Q has real rank zero, see [23]. By [16, Theorem 9.3 and its proof], for each τ ∈ T (C(X) ⋊ Zd) there is a unital embedding α : C(X) ⋊ Zd → D into a simple, unital, AF algebra with unique tracial state δ and such that τ = δ ◦ α. Now (C(X) ⋊ Zd) ⊗ Q is TAF by Theorem 3.1. By [19], C(X) ⋊ Zd is classifiable (C(X) ⋊ Zd is well known to satisfy the UCT, see [2]); since it has real rank zero it is TAF. 4. Odd spheres Below we will see that C∗-algebras associated to certain minimal homeomorphisms of spheres (of odd dimension at least 3) are classified by their spaces of invariant Borel probability measures; the crucial point here is that for the crossed products projections do not separate tracial states. The argument combines classification by embeddings with recent results of Strung and of Robert. 4.1 We start with a technical lemma, the crucial step of which relies on [29, Lemma 4.4]. The result is implicitly contained in [25, Sections 2.4 and 2.5]; a full proof is given below for the convenience of the reader. Recall that a purely positive element of a C∗-algebra is one that is not Cuntz equivalent to a projection. Lemma: Let B = C([0, 1]) ⊗ Mr, b ∈ B1 +, ǫ > 0 be given. CLASSIFYING CROSSED PRODUCTS 15 Then, there are q, L ∈ N, purely positive elements b1, . . . , bL ∈ (B ⊗ Mq)1 + and numbers ν1, . . . , νL ∈ [0, 1] such that the bl are pairwise orthogonal and such that (5) for any τ ∈ T (B). (cid:12)(cid:12)τ (b) −PL l=1 νl · dτ ⊗trq (bl)(cid:12)(cid:12) < ǫ Proof: Once the bl are constructed, it will be enough to confirm (5) for extremal traces of B, i.e. for traces of the form τt = evt ⊗ trr, t ∈ [0, 1]. The numbers dτt⊗trq (bl) are then just the ranks, divided by r · q, of the matrices bl(t), t ∈ [0, 1]. Moreover, t 7→ τt(b) is just a positive continuous function of norm at most 1 on [0, 1], and there are 2 ≤ L ∈ N, 0 = t0 < t1 < . . . < tL = 1 and ν0, . . . , νL ∈ [0, 1] such that for t ∈ [tl−1, tl], l = 1, . . . , L, we have τt(b) − νl < ǫ/4 and νl − νl−1 < ǫ/4. Set t′ l := (tl − tl−1)/2, l = 1, . . . , L. Choose q ∈ N such that 1/q < ǫ/4. Let Dq ⊂ Mq denote the subalgebra of diagonal matrices. For each l = 1, . . . , L−1, choose pairwise disjoint nondegenerate closed intervals Il,1, . . . , Il,q ⊂ (t′ + such 2 ,l(t) has rank at most 1 for each t ∈ [t′ l+1] and such that for t ∈ Il,s, the s-th a 1 diagonal entry of a 1 l+1) and a function a 1 2 ,l ∈ C([t′ l, t′ l, t′ l+1], Dq)1 l, t′ 2 ,l(t) is 1. Now by [29, Lemma 4.4], for each l = 1, . . . , L − 1, there are a0,l, a1,l ∈ C([t′ l, t′ l+1], Dq)1 + such that a0,l ⊥ a1,l, a0,l + a 1 l+1], and a0,l(t′ l,t′ l) = a1,l(t′ l+1) = 1q. Note that for each t ∈ [t′ q − 1, and that a0,l(t′ 2 ,l + a1,l = 1[t′ l, t′ l+1) = a1,l(t′ l) = 0. l+1], the ranks of a0,l(t) and a1,l(t) add up to at least We are now ready to define the Mr ⊗ Mq-valued functions bl, l = 1, . . . , l as follows: For l = 2, . . . , L − 1, set 1r ⊗ a1,l−1(t), 1r ⊗ a0,l(t), 0, l−1, t′ t ∈ [t′ l] l, t′ t ∈ [t′ l+1] t ∈ [0, 1] \ [t′ l, t′ l+1]. 1r ⊗ 1q, 1r ⊗ a0,l(t), 0, t ∈ [0, t′ 1] 1, t′ t ∈ [t′ 2] t ∈ [t′ 2, 1] 1r ⊗ 1q, 1r ⊗ a1,L−1(t), 0, t ∈ [t′ t ∈ [t′ t ∈ [0, t′ L, 1] L−1, t′ L] L−1]. For l = 1, L, set and bl(t) :=  b1(t) :=  bL(t) :=  It is clear from our construction that the bl are indeed continuous (thus well- defined) positive contractions; they each take the value 1r ⊗ 1q for some t and are 0 for some other t, hence (the interval is connected) must be purely positive. Each bl is supported on [t′ l+1]. Moreover, for each t ∈ [0, 1], bl(t) can be nonzero for at most two values of l, which then must be consecutive. The bl are pairwise orthogonal since the a0,l and a1,l are orthogonal for each l. l−1, t′ 16 WILHELM WINTER Now let t ∈ [t′ ¯l, t′ ¯l+1] for some ¯l ∈ {1, . . . , L − 1}, so that bl(t) = 0 for all l 6= ¯l, ¯l + 1. We estimate l=1 νl · dτt⊗trq (bl) τt(b) −PL < τt(b) − = τt(b) − ν¯l · dτt⊗trq (b¯l) − ν¯l+1 · dτt⊗trq (b¯l+1) ν¯l r · q · (rank(b¯l(t)) + rank(b¯l+1(t))) + ǫ/4 < τt(b) − ν¯l · 1 + ǫ/4 + 1/(r · q) < ǫ. 1] and t ∈ [t′ L, 1] similar (in fact, easier) estimates hold, so we have For t ∈ [0, t′ confirmed (5) for extremal traces of B. 4.2 Theorem: Let A and B be separable, simple, unital C∗-algebras. Suppose that dimnuc A < ∞. Let B be TAI and suppose there is a unital embedding such that and such that ι : A → B T (ι) : T (B) ≈−→ T (A) τ∗ = τ ′ ∗ ∈ S(K0(B)) for τ, τ ′ ∈ T (B). Then, A ⊗ Q is TAI. Proof: We may assume A = A ⊗ Q. Let Bi ⊂ B, i ∈ N, be a sequence of interval algebras tracially approximating B via u.c.p. maps i.e., the ψi are approximately multiplicative, ψi : B → Bi, and and kψi(b) − 1Bib1Bik i→∞−→ 0 k[1Bi, b]k i→∞−→ 0, supτ ∈T (B){1 − τ (1Bi )} < ǫi for some sequence (ǫi)i∈N ⊂ (0, 1), with ǫi i→∞−→ 0. Each Bi is of the form with each Bi,j being nonzero and either Mri,j or C([0, 1]) ⊗ Mri,j for some ri,j . Set j=1 Bi,j Bi =LMi for some, hence all, τ ∈ T (B) (the λi,j are nonzero by simplicity of B). λi,j := τ (1Bi,j ) ∈ R∗ + Choose µi,j ∈ Q+ such that sup {1 − µi,j λi,j } < ǫi 3 · Mi · ri,j j∈{1,...,Mi} for i ∈ N. Let γ : Cu(B ⊗ Q) ∼=−→ V (B ⊗ Q) ⊔ LAff(T (B ⊗ Q))++ CLASSIFYING CROSSED PRODUCTS 17 be the semigroup isomorphism (in Mor(Cu)) of [1, Theorem 5.27]. (The domain of the isomorphism in [1] is W(B ⊗ Q ⊗ K), which by [1, Corollary 4.31 and Theo- rem 4.33] can be identified with Cu(B ⊗ Q).) Let δ : LAff(T (A ⊗ Q))++ ∼=−→ LAff(T (B ⊗ Q))++ be induced by ι ⊗ idQ, i.e., δf (τ ) = f (τ ◦ (ι ⊗ idQ)) for f ∈ LAff(T (A ⊗ Q))++ and τ ∈ T (B ⊗ Q). Note that δ−1(g)(τ ⊗ τQ) = g(T (ι)−1(τ ) ⊗ τQ) for g ∈ LAff(T (B ⊗ Q))++ and τ ∈ T (A). Let be the canonical map. Define by ζ : Q+ → V (Q) → V (A ⊗ Q) κi : Cu(Bi) → Cu(A ⊗ Q) κi([b]) :=(ζ( µi,j δ−1( µi,j λi,j λi,j · dτ ([b])), · γ([b ⊗ 1Q])), if [b] is the class of a projection in K ⊗ Bi,j, if b is purely positive in K ⊗ Bi,j; here, the dimension function dτ comes from some τ ∈ T (B) (note that by our hypotheses on the pairing between T (B) and S(K0(B)) dτ ([b]) is independent of the particular choice of τ as long as [b] is represented by a projection). Note that µi,j λi,j defined. · dτ ([b]) ∈ Q+ if [b] is the class of a projection, hence κi is well- One checks that κi is a semigroup homomorphism in Mor(Cu). Let κ∼ i : Cu∼(Bi) → Cu∼(A ⊗ Q) be the induced map (cf. [22, Section 3]), i.e., κ∼ i ([x] − n · [1]) = κi([x]) − n · [1]; then κ∼ i ∈ Mor(Cu). By [22, Theorem 1], κ∼ i lifts to a unital ∗-homomorphism βi : Bi → A ⊗ Q. If Bi,j is a matrix algebra, then for a projection p ∈ Bi,j and for traces τ ∈ T (A), τ ′ ∈ T (B), (τ ⊗ τQ)(βi(p)) − (T (ι)−1(τ ))(p) = dτ ⊗τQ(βi(p)) − τ ′(p) = dτ ⊗τQ(κ∼ i ([p])) − dτ ′ ([p]) = dτ ⊗τQ(κi([p])) − dτ ′([p]) = dτ ⊗τQ(ζ( µi,j λi,j = µi,j λi,j ǫi · dτ ′([p]) − dτ ′([p]) · dτ ′([p]))) − dτ ′([p]) ≤ . Mi · ri,j 18 WILHELM WINTER But then for each b ∈ (Bi,j)1 + we have (τ ⊗ τQ)(βi(b)) − (T (ι)−1(τ ))(b) < ǫi Mi for any τ ∈ T (A). Let us now consider the case where Bi,j is of the form C([0, 1]) ⊗ Mri,j . We first check that βBi,j is injective, hence in particular maps purely positive elements to purely positive elements: If 0 6= b ∈ (Bi,j )1 + is purely positive, then for any τ ∈ T (B ⊗ Q), δ([βi(b)])(τ ) = δ(κi([b]))(τ ) · γ([b ⊗ 1Q]))(τ ) = δδ−1( µi,j λi,j = µi,j λi,j = µi,j λi,j 6= 0, · γ([b ⊗ 1Q])(τ ) · dτ (b ⊗ 1Q) from which follows that βBi,j is injective. Now let b ∈ (Bi,j )1 b1, . . . , bL ∈ (Bi,j ⊗ Mq)1 [0, 1] such that for any τ ∈ T (B) + be arbitrary. With the aid of Lemma 4.1, find q, L ∈ N and + pairwise orthogonal and purely positive, and ν1, . . . , νL ∈ l=1 νl · dτ ⊗trq (bl) < ǫi 3 · Mi . But then, for any τ ∈ T (A) we also have τ (b) −PL (τ ⊗ τQ)(βi(b)) −PL l=1 νl · dτ ⊗τQ⊗trq (βi ⊗ idMq (bl)) < ǫi 3 · Mi . We compute for τ ∈ T (A) and l ∈ {1, . . . , L} dτ ⊗τQ⊗trq (βi ⊗ idMq (bl)) = dτ ⊗τQ⊗trq (κi([bl])) = dτ ⊗τQ⊗trq (δ−1( µi,j λi,j = µi,j λi,j = µi,j λi,j = µi,j λi,j · dT (ι⊗idMq )−1(τ ⊗trq)(bl) · dT (ι)−1(τ )⊗trq (bl). · γ([bl ⊗ 1Q]))) · γ([bl ⊗ 1Q])(T (ι ⊗ idMq )−1(τ ⊗ trq) ⊗ τQ) It follows that for τ ∈ ∂eT (A) (T (ι)−1(τ ))(b) − (τ ⊗ τQ)(βi(b)) < PL = PL < 1 − µi,j λi,j l=1 νl · (dT (ι)−1(τ )⊗trq (bl) − dτ ⊗τQ⊗trq (βi ⊗ idMq (bl))) + 2ǫi 3Mi l=1 νl · (1 − µi,j λi,j · 1 + 2ǫi 3Mi ) · dT (ι)−1(τ )⊗trq (bl) + 2ǫi 3Mi < ǫi Mi . As a consequence, we have (T (ι)−1(τ ))(b) − (τ ⊗ τQ)(βi(b)) < ǫi for all b ∈ (Bi)1 + and τ ∈ T (A). CLASSIFYING CROSSED PRODUCTS 19 Set and for i ∈ N. We estimate σi := ψi ◦ ι : A → Bi ρi := Bi βi−→ A ⊗ Q ∼= A sup kτ (ρiσi(a) − a)k τ ∈T (A) = sup (τ ⊗ τQ)(βiσi(a)) − τ (a) τ ∈T (A) ≤ = ≤ sup (T (ι)−1(τ ))(σi(a)) − τ (a) + ǫi τ ∈T (A) sup (T (ι)−1(τ ))(ψi(ι(a)) − ι(a)) + ǫi τ ∈T (A) sup (T (ι)−1(τ ))(ψi(ι(a)) − 1Biι(a)1Bi ) τ ∈T (A) + sup τ ∈T (A) (T (ι)−1(τ ))(1B − 1Bi) +2k[ι(a), 1Bi]k +ǫi ≤ kψi(ι(a)) − 1Biι(a)1Bi k + sup τ ∈T (A) (T (ι)−1(τ ))(1B − 1Bi) +2k[ι(a), 1Bi]k +ǫi i→∞−→ 0. We have now verified the hypotheses of Theorem 2.2, whence A ∼= A ⊗ Q is TAI. 4.3 The following is shown by Strung in [26]; this in particular confirms the hy- potheses of Theorem 4.2 -- see Corollary 4.4 below. The Corollary as it stands only covers the examples of [33], but after this work (and that of Strung) was completed, Lin generalized it to cover arbitrary minimal homeomorphisms of odd dimensional spheres; see [18]. Lin's method still uses our classification by embedding technique and also ideas from [26]. Theorem: Let α : S2n+1 → S2n+1, n ≥ 1, be a minimal homeomorphism which can uniformly be approximated by periodic homeomorphisms as in [33]. Then, there is a uniquely ergodic minimal homeomorphism γ : X → X on the Cantor set X (in fact, γ can be chosen to be an odometer action), such that the product α × γ : S2n+1 × X → S2n+1 × X is minimal and such that C(X × S2n+1) ⋊ Z is TAI with a uniquely determined state on K0. 4.4 Corollary: Let α : S2n+1 → S2n+1, n ≥ 1, be a minimal homeomorphism, as constructed in [33]. Then, (C(S2n+1) ⋊ Z) ⊗ Q is TAI and C(S2n+1) ⋊ Z is classifiable in the sense of [17]. In particular, crossed products of this form are determined by their trace spaces. 20 WILHELM WINTER Proof: By Theorem 4.3, there is a uniquely ergodic Cantor minimal action γ : X → X such that C(S2n+1 × X) ⋊ Z is TAI with only one state on K0. Let ι : C(S2n+1) ⋊ Z → C(S2n+1 × X) ⋊ Z be the canonical embedding. It is straightforward to check that T (ι) is a homeo- morphism. By Theorem 4.2, (C(S2n+1) ⋊ Z) ⊗ Q is TAI, hence classifiable (the algebra is in the bootstrap class and satisfies the UCT, cf. [2]). All such crossed products have the same ordered K-theory and only one state on K0, so their Elliott invariants are just distinguished by their trace spaces. 5. Generic classifiability We now follow up on a theme from [10] to show that, in fair generality, classifiability is generically preserved under crossed products by Zd. We need a notion of Rokhlin dimension for Zd-actions; this is the obvious generalization of [10, Definition 2.3(c)] where {0, . . . , p} is replaced by {0, . . . , p}d, see [37] and [28] for details. 5.1 Proposition: Let A be a separable, simple, unital C∗-algebra with dimnuc A < ∞. Let α : Zd → Aut(A) be an action with finite Rokhlin dimension. Then, A ⋊α Zd is again simple with finite nuclear dimension, and the canonical map T (A ⋊α Zd) → T (A)α is an isomorphism. Proof: This is just a variation of [10, Theorem 4.1], see also [28]. We omit the details. 5.2 Proposition: Let A be a separable, unital, Z-stable C∗-algebra. Then, there is a dense Gδ of Zd-actions of A with finite Rokhlin dimension. Proof: Apply the technique of [10, Theorem 3.4] with Z ⊗d in place of Z to end up with a generic set of Zd-actions with Rokhlin dimension at most 2d − 1. (Alternatively, one could also apply [10, Theorem 3.4] inductively d times.) 5.3 Theorem: Let A be a C∗-algebra with finitely many extremal tracial states and satisfying the conditions of Theorem 1.3 (i.e., A is separable, simple, unital, nonele- mentary, has finite decomposition rank, satisfies the UCT and K0(A) separates the tracial states of A). Then, for a dense Gδ of Zd-actions on A, A⋊ Zd again has finitely many extremal traces and satisfies the conditions of Theorem 1.3. In particular, for generic sets of Zd-actions, being classifiable by ordered K-theory in the sense of Theorem 1.3 and having finitely many extremal tracial states passes to crossed products. Proof: By Proposition 5.2, there is a dense Gδ of Zd-actions of A with finite Rokhlin dimension; by Proposition 5.1, every such action gives rise to a crossed product A ⋊ Zd which is simple, has finite nuclear dimension and with tracial state space being a subspace of T (A). But then T (A ⋊ Zd) is finite dimensional, compact and convex, hence has only finitely many extreme points. These are again separated by projections, whence A ⋊ Zd has real rank zero. Since A is AH (see Theorem 1.3), by [16, Theorem 9.3 and its proof], for each τ ∈ T (A ⋊ Zd) there is a unital embedding α : A ⋊ Zd → D into a simple, unital, AF algebra with unique CLASSIFYING CROSSED PRODUCTS 21 tracial state δ and such that τ = δ ◦ α. Now (A ⋊ Zd) ⊗ Q is TAF by Theorem 3.1. References 1. P. Ara, F. Perera, and A. S. Toms, K-theory for operator algebras. Classification of C∗- algebras, Aspects of operator algebras and applications, Contemp. Math., vol. 534, Amer. Math. Soc, Providence RI., 2011, pp. 1 -- 71. 2. B. Blackadar, K-Theory for Operator Algebras, MSRI Monographs, vol. 5, Springer Verlag, Berlin and New York, 1986. 3. B. Blackadar and E. Kirchberg, Generalized inductive limits of finite-dimensional C∗-algebras, Math. Ann. 307 (1997), 343 -- 380. 4. A. Connes, An analogue of the Thom isomorphism for crossed products of a C∗-algebra by an action of R, Adv. Math. 39 (1981), 31 -- 55. 5. G. A. Elliott and A. S. Toms, Regularity properties in the classification program for separable amenable C∗-algebras, Bull. Amer. Math. Soc. (N.S.) 45 (2008), no. 2, 229 -- 245. 6. A. Fathi and M. R. Herman, Existence de diff´eomorphismes minimaux, Dynamical systems, Vol. I -- Warsaw, Soc. Math. France, Paris, 1977, pp. 37 -- 59. Ast´erisque, No. 49. 7. T. Giordano, H. Matui, I. F. Putnam, and Ch. F. Skau, Orbit equivalence for Cantor minimal Z2-systems, J. Amer. Math. Soc. 21 (2008), no. 3, 863 -- 892. 8. , Orbit equivalence for Cantor minimal Zd-systems, Invent. Math. 179 (2010), no. 1, 119 -- 158. 9. T. Giordiano, I. F. Putnam, and C. F. Skau, Topological orbit equivalence and C∗-crossed products, J. Reine Angew. Math. 469 (1995), 51 -- 111. 10. I. Hirshberg, W. Winter, and J. Zacharias, Rokhlin dimension and C∗-dynamics, Comm. Math. Phys. 335 (2015), 637 -- 670. 11. X. Jiang and H. Su, On a simple unital projectionless C∗-algebra, Amer. J. Math. 121 (1999), no. 2, 359 -- 413. 12. E. Kirchberg and M. Rørdam, Infinite non-simple C∗-algebras: absorbing the Cuntz algebra O∞, Adv. Math. 167 (2002), no. 2, 195 -- 264. 13. , Purely infinite C∗-algebras: Ideal preserving zero homotopies, Geom. Funct. Anal. 15 (2005), no. 2, 377 -- 415. 14. E. Kirchberg and W. Winter, Covering dimension and quasidiagonality, Internat. J. Math. 15 (2004), 63 -- 85. 15. H. Lin, Classification of simple C∗-algebras of tracial topological rank zero, Duke Math. J. 125 (2004), 91 -- 119. 16. 17. 18. , AF-embeddings of the crossed products of AH-algebras by finitely generated abelian groups, Int. Math. Res. Pap. IMRP (2008), no. 3, Art. ID rpn007, 67. , Asymptotic unitary equivalence and classification of simple amenable C∗-algebras, Invent. Math. 183 (2011), no. 2, 385 -- 450. , Minimal dynamical systems on connected odd dimensional spaces, arXiv preprint math.OA/1404.7034, 2014. 19. H. Lin and Z. Niu, Lifting KK-elements, asymptotic unitary equivalence and classification of simple C∗-algebras, Adv. Math. 219 (2008), no. 5, 1729 -- 1769. 20. H. Matui and Y. Sato, Strict comparison and Z-absorption of nuclear C∗-algebras, Acta Math. 209 (2012), no. 1, 179 -- 196. 21. , Decomposition rank of UHF-absorbing C∗-algebras, Duke Math. J. 163 (2014), 2687 -- 2708. 22. L. Robert, Classification of inductive limits of 1-dimensional NCCW complexes, Adv. Math. 231 (2012), no. 5, 2802 -- 2836. 23. M. Rørdam, The stable and the real rank of Z-absorbing C∗-algebras, Internat. J. Math. 15 (2004), no. 10, 1065 -- 1084. 24. M. Rørdam and W. Winter, The Jiang -- Su algebra revisited, J. Reine Angew. Math. 642 (2010), 129 -- 155. 25. K. R. Strung, On classification, UHF-stability, and tracial approximation of simple nuclear C∗-algebras, Ph.D. thesis, Munster, 2013. 26. , C∗-algebras of minimal dynamical systems of the product of a Cantor set and an odd dimensional sphere, to appear in J. Funct. Anal.; arXiv preprint math.OA/1403.3136, 2014. 22 WILHELM WINTER 27. K. R. Strung and W. Winter, UHF-slicing and classifcation of nuclear C∗-algebras, J. Top. Anal. 6 (2014), 465 -- 540. 28. G. Szab´o, The Rokhlin dimension of topological Zm-actions, to appear in Proc. London Math. Soc. (3); arXiv preprint math.OA/1308.5418, 2013. 29. A. Tikuisis and W. Winter, Decomposition rank of Z-stable C∗-algebras, Analysis & PDE 7 (2014), 673 -- 700. 30. A. S. Toms and W. Winter, Minimal dynamics and the classification of C∗-algebras, Proc. Natl. Acad. Sci. USA 106 (2009), no. 40, 16942 -- 16943. 31. 32. , The Elliott conjecture for Villadsen algebras of the first type, J. Funct. Anal. 256 (2009), no. 5, 1311 -- 1340. , Minimal dynamics and K-theoretic rigidity: Elliott's conjecture, Geom. Funct. Anal. 23 (2013), no. 1, 467 -- 481. 33. A. Windsor, Minimal but not uniquely ergodic diffeomorphisms, Smooth ergodic theory and its applications (Seattle, WA, 1999), Proc. Sympos. Pure Math., vol. 69, Amer. Math. Soc., Providence, RI, 2001, pp. 809 -- 824. 34. W. Winter, On the classification of simple Z-stable C∗-algebras with real rank zero and finite decomposition rank, J. London Math. Soc. (2) 74 (2006), no. 1, 167 -- 183. 35. 36. 37. 38. , Decomposition rank and Z-stability, Invent. Math. 179 (2010), no. 2, 229 -- 301. , Nuclear dimension and Z-stability of pure C∗-algebras, Invent. Math. 187 (2012), 259 -- 342. , Dynamic dimension, in preparation, 2013. , Localizing the Elliott conjecture at strongly self-absorbing C∗-algebras. With an ap- pendix by H. Lin., J. Reine Angew. Math. 692 (2014), 193 -- 231. 39. W. Winter and J. Zacharias, Completely positive maps of order zero, Munster J. Math. 2 (2009), 311 -- 324. 40. , The nuclear dimension of C∗-algebras, Adv. Math. 224 (2010), 461 -- 498. Mathematisches Institut, Universitat Munster, Germany E-mail address: [email protected]
1612.07706
3
1612
2017-06-13T17:50:57
Higher l^2-Betti numbers of universal quantum groups
[ "math.OA" ]
We calculate all $\ell^2$-Betti numbers of the universal discrete Kac quantum groups $\hat U_n^+$ as well as their full half-liberated counterparts $\hat U_n^*$.
math.OA
math
Higher ℓ2-Betti numbers of universal quantum groups by Julien Bichon, David Kyed1 and Sven Raum Abstract. We calculate all ℓ2-Betti numbers of the universal discrete Kac quan- tum groups U+ n as well as their half-liberated counterparts U∗ n. 1 Introduction The category of locally compact quantum groups [KV00] is a natural extension of the category of locally compact groups. There are two important reasons to pass from locally compact groups to locally compact quantum groups. First, classical Pontryagin duality for locally compact abelian groups extends to a full duality theory for locally compact quantum groups, in particular establishing a duality between discrete and compact quantum groups. Second, locally compact quantum groups form the correct framework to host a number of important deformations and liberations of classical groups. A convenient operator algebraic setting describing discrete and compact quantum groups was first provided by Woronowicz [Wor87, Wor98], and since his seminal work an abundance of analytical tools have been shown to carry over from discrete groups to discrete quantum groups (cf. [BMT01, Bra12, Ver07, Fim10, MN06, Voi11]). The current paper is concerned with the ℓ2-Betti numbers of discrete quantum groups [Kye08b] and our primary focus is on the duals of the free unitary quantum groups U+ n , which are universal within the class of discrete quantum groups in the same way that the free groups are universal within the class of finitely generated discrete groups; that is, every2 finitely generated discrete quantum group is the quotient of some U+ n , the latter to be understood in the sense of the existence of a surjection at the Hopf algebra level. Due to the lack of a topological interpretation of (co)homology of quantum groups, the computation of their ℓ2-Betti numbers has proven to be a difficult task, and only general structural results [Kye08a, Kye11, Kye12] were available until Vergnioux's paper [Ver12], in which it was proven that the first ℓ2-Betti number vanishes for the duals of the free orthogonal quantum groups O+ n vanish [CHT09]. Concerning the universal quantum groups U+ n , Vergnioux's work also showed that the first ℓ2-Betti number is non-zero. He conjectured that it would equal one, which was recently proven by the second and third author in [KR16] along with the observation that β(2) n) = 0 for p ⩾ 4, thus leaving open the question about the values of the important second and third ℓ2-Betti number of U+ n . n by combining techniques from [KR16] with those from [BNY15, BNY16] and a new free product formula for ℓ2-Betti numbers of discrete quantum groups. In particular, we obtain a different proof of the fact that β(2) n . Subsequently, Collins-Härtel-Thom showed that also the higher ℓ2-Betti numbers of O+ In this article we provide computations of all ℓ2-Betti numbers of U+ p ( U+ 1 ( U+ n) = 1. MSC classification: 16T05, 46L65, 20G42 Keywords: ℓ2-Betti numbers, free unitary quantum groups, half-liberated unitary quantum groups, free product formula, extensions Acknowledgements: The authors would like to thank Étienne Blanchard for pointing out a number of misprints in an earlier version of the article. 1The research leading to these results has received funding from the Villum foundation grant 7423 (D.K). 2We tacitly limit our focus to the case of quantum groups of Kac type here, although the statement remains true of one allows non-trivial matrix-twists in the definition of the unitary quantum groups; cf. [Ban97a]. Higher ℓ2-Betti numbers by Julien Bichon, David Kyed and Sven Raum Theorem A. For any n ⩾ 2 the free unitary quantum group U+ n satisfies β(2) p ( U+ n ) = ⎧⎪⎪ ⎨ ⎪⎪⎩ 1 if p = 1, 0 otherwise. In [CHT09] it was proven that the discrete duals of the free orthogonal quantum groups sat- isomorphism n )-module M. But since this property implies a symmetry isfy a certain Poincaré duality developed in [vdB98], meaning that there is a natural H∗( O+ in the ℓ2-Betti numbers, Theorem A shows the following. n , M) for every Pol(O+ n , M) ≅ H3−∗( O+ Corollary B. The discrete quantum groups U+ n do not satisfy Poincaré duality. The techniques used in the proof of Theorem A are of independent interest. Firstly, we provide a formula for the ℓ2-Betti numbers of arbitrary free product quantum groups [Wan95]. Theorem C. If G and H are non-trivial compact quantum groups of Kac type, then the following holds. β(2) p (G ∗ H) = ⎧⎪⎪⎪⎪⎪ ⎪⎪⎪⎪⎪⎩ ⎨ 0 1 (G) − β(2) β(2) β(2) p (G) + β(2) 0 (G) + β(2) p ( H) 1 ( H) − β(2) if p = 0 0 ( H) + 1 if p = 1 if p ⩾ 2 . Our second ingredient is the fact that ℓ2-Betti numbers enjoy a natural scaling behaviour under cocentral extensions, which is a quantum counterpart of the classical scaling formula for finite index inclusions of groups (cf. [Lüc02, Theorem 6.54 (6)]). Theorem D. Let C → Pol(H) → Pol(G) → CΓ → C be an exact sequence of Hopf ∗-algebras in which G and H are compact quantum groups of Kac type and Γ is a finite abelian group. Then for any p ⩾ 0 we have β(2) p ( H) = Γ β(2) p (G) . The previous theorem allows for other applications. We combine it with results of [Kye08a] in order to obtain the following result (see Section 2 for the definition of the half-liberated quantum groups): Theorem E. For any n ⩾ 2 the half-liberated unitary quantum group U∗ n satisfies β(2) p ( U∗ n) = ⎧⎪⎪ 1 ⎨ ⎪⎪⎩ 0 if p = 1 otherwise. 2 Preliminaries We collect some notation and necessary tools for the sections to follow. Augmented algebras. A ∗-algebra A with a distinguished ∗-homomorphism ε∶ A → C is called an augmented ∗-algebra. We write A+ = ker ε. 2 Higher ℓ2-Betti numbers by Julien Bichon, David Kyed and Sven Raum Compact and discrete quantum groups. Compact quantum groups were introduced in the C∗-algebra setting by Woronowicz [Wor87, Wor98]. If G is a compact quantum group, we denote by Pol(G) the associated Hopf ∗-algebra of polynomial functions, which possesses a unique Haar state denoted by ϕ. In case ϕ is tracial, we call G a compact quantum group of Kac type. We denote the discrete dual of a compact quantum group G by G. Von Neumann algebra completions and measurable operators. If G is a compact quantum group and ϕ is the Haar state on Pol(G), then the von Neumann algebra completion of Pol(G) in the associated GNS-representation is written L∞(G) = πϕ(Pol(G))′′. If G is of Kac type, then L∞(G) is a finite von Neumann algebra and we let M(G) denote the algebra of measurable operators affiliated with L∞(G). ℓ2-Betti numbers. Given a compact quantum group of Kac type G, we denote by dimL∞(G) Lück's dimension function [Lüc02] for modules over the finite von Neumann algebra L∞(G). If G denotes the discrete dual of G, its ℓ2-Betti numbers are defined [Kye08b] as β(2) p (G) = dimL∞(G) Tor Pol(G) p (L∞(G), C) . Work of Thom [Tho08] and Reich [Rei01] allows to calculate the ℓ2-Betti numbers alternatively as β(2) p (G) = dimL∞(K) Extp Pol(G)(C, M(K)), whenever Pol(G) ⊂ Pol(K) as Hopf ∗-algebras for another compact quantum group of Kac type K. This is explained in more detail in Remark 1.8 of [KR16]. Cocentral homomorphisms. Let G be a compact quantum group and Γ a discrete abelian group. A Hopf ∗-algebra homomorphism π ∶ Pol(G) → CΓ is called cocentral if (π ⊗ id)○ ∆ = (π ⊗ id)○ Σ○ ∆ where Σ denotes the map flipping the tensor factors. In this case {a ∈ Pol(G) (id ⊗ π) ○ ∆(a) = a ⊗ 1} = {a ∈ Pol(G) (π ⊗ id) ○ ∆(a) = 1 ⊗ a}, and this subalgebra is denoted Pol(G)e , and it too is the Hopf ∗-algebra of a compact quantum group. Exact sequences of Hopf algebras. Let A, B, L be Hopf ∗-algebras. A sequence of Hopf ∗-algebra maps is called exact if C Ð→ B ιÐ→ A πÐ→ L Ð→ C • ι is injective and π is surjective, • ker π = A ι(B+) = ι(B+) A and • ι(B) = {a ∈ A (id ⊗ π) ○ ∆(a) = a ⊗ 1} = {a ∈ A (π ⊗ id) ○ ∆(a) = 1 ⊗ a}. By Proposition 1.2 of [BNY16], every surjective cocentral Hopf ∗-algebra homomorphism π∶ Pol(G) → CΓ gives rise to an exact sequence of Hopf ∗-algebras πÐ→ CΓ Ð→ C. C Ð→ Pol(G)e ιÐ→ Pol(G) The notation Pol(G)e stems from the fact that π turns Pol(G) into a Γ-graded Hopf algebra. 3 Higher ℓ2-Betti numbers by Julien Bichon, David Kyed and Sven Raum Universal quantum groups. Wang's [Wan98, VW96] universal quantum groups U+ n and O+ n can be described by their associated ∗-algebras of polynomial functions Pol(U+ Pol(O+ n) = ⟨uij , 1 ⩽ i , j ⩽ n uu∗ = u∗u = 1 = ¯u ¯u∗ = ¯u∗ ¯u⟩ , n) = ⟨vij , 1 ⩽ i , j ⩽ n vij = v ∗ ij , v v t = v tv = 1⟩ . Here u, v and ¯u denote the n × n-matrices (vij)ij , (uij)ij and (u∗ ij)ij , respectively. Their comul- tiplications are given by dualising matrix multiplication uij ↦ ∑k uik ⊗ ukj and vij ↦ ∑k vik ⊗ vkj respectively and their counits satisfy ε(uij) = δij = ε(vij) for all i , j ∈ {1, . . . , n}. The matrices u and v are called the fundamental corepresentations of U+ n , respectively. n and O+ Graded twists of universal quantum groups. Examples 2.18 and 3.6 of [BNY15] show that the Hopf ∗-algebra homomorphisms Pol(U+ n) ∗ Pol(O+ n) → CZ2 ∶ uij ↦ u1δij n) → CZ2 ∶ v (k) ij ↦ u1δij , Pol(O+ where v (k) denotes the fundamental corepresentation corresponding to the k-th factor in the free product and u1 denotes the generator of Z2 inside CZ2, are cocentral and induce exact sequences of Hopf ∗-algebras C Ð→ Pol(H) Ð→ Pol(U+ n) Ð→ CZ2 Ð→ C C Ð→ Pol(H) Ð→ Pol(O+ n) ∗ Pol(O+ n) pÐ→ CZ2 Ð→ C (2.1) (2.2) for the same compact quantum group H. To see this, first note that since our ground field is the complex numbers, the Hopf algebra denoted B(In) in [BNY15, Example 2.17] can be equipped with a ∗-structure making the canonical generators self-adjoint and the resulting Hopf ∗-algebra n). Similarly, the Hopf algebra H(In) of [BNY15, Example 2.18] can be identifies with Pol(O+ given a ∗-structure which satisfies u∗ ij = vij on the canonical generators, and the resulting Hopf ∗-algebra naturally identifies with Pol(U+ n) ∗ Pol(O+ n) and note that the cocentral Hopf ∗-algebra morphism p ∗ p∶ A → CZ2 defines a graded twisting At [BNY15, Definition 2.6] of the Hopf ∗-algebra A, which in turn also allows for a cocentral Hopf ∗-algebra morphism (p ∗ p)t ∶ At → CZ2 [BNY15, Proposition 2.7]. We therefore obtain two exact sequences of Hopf ∗-algebras n). For notational convenience, we set A ∶= Pol(O+ C Ð→ A0 Ð→ A C Ð→ (At)0 Ð→ At (p∗p)t p∗pÐ→ CZ2 Ð→ C Ð→ CZ2 Ð→ C. Recall that At is defined as spanC{A0 ⊗ u0, A1 ⊗ u1,} ⊂ A ⋊ Z2, where Z2 acts on A by flipping the factors in the free product and Ag ∶= {a ∈ A (id ⊗ p ∗ p)∆(a) = a ⊗ ug}, g ∈ Z2 . One may now check (see [BNY15, Example 2.18] for details) that the map uij ↦ v (1) ij ⊗ u1 extends to a Hopf ∗-algebra isomorphism Pol(U+ n) ≃ At . A direct calculation verifies that (At)0 = A0 ⊗ u0 ≅ A0, and denoting the compact quantum group underlying this Hopf ∗-algebra by H we obtain the sequences (2.1) and (2.2). 4 Higher ℓ2-Betti numbers by Julien Bichon, David Kyed and Sven Raum Half-liberated quantum groups. The ideal in Pol(O+ n) generated by {abc − cba a, b, c ∈ {vij ∶ i , j = 1, . . . , n}} gives rise to a quotient which is the Hopf ∗-algebra of a compact quantum group of Kac type, known as the half-liberated orthogonal group [BS09] and denoted O∗ n. Similarly, the n) by the ∗-ideal generated by {ab∗c −cb∗a a, b, c ∈ {uij ∶ i , j = 1, . . . n}} is the quotient of Pol(U+ Hopf ∗-algebra of a compact quantum group of Kac type which is known as the half-liberated unitary quantum group [BDD11] and denoted U∗ n. Automorphisms of Hopf ∗-algebras. Whenever α is an automorphism of Pol(G) that preserves its Haar state, then α uniquely extends to an automorphism of L∞(G). In particular, if π ∶ Pol(G) → CΓ is a cocentral Hopf ∗-algebra homomorphism, then the induced action of Γ, defined for χ ∈ Γ by Pol(G) ∆Ð→ Pol(G) ⊗ Pol(G) id⊗πÐ→ Pol(G) ⊗ CΓ Ð→ Pol(G) ⊗ C(Γ) id⊗evχÐ→ Pol(G), preserves the Haar state, thanks to right invariance (ϕ ⊗ id) ○ ∆(a) = ϕ(a)1. 3 A free product formula for discrete quantum groups In this section we combine results from [Bic16] with additional homological calculations to prove Theorem C. We start with a general lemma for inclusions of quantum groups. Lemma 3.1. If G and K are compact quantum groups of Kac type such that Pol(G) ⊂ Pol(K) as Hopf ∗-algebras, then 1 + β(2) 1 (G) − β(2) 0 (G) = dimL∞(K) HomPol(G)(Pol(G)+, M(K)) . Proof. Consider the short exact sequence 0 → Pol(G)+ ι→ Pol(G) ε→ C → 0 of right Pol(G)-modules and the associated long exact sequence of Ext-groups: 0 HomPol(G)(C, M(K)) HomPol(G)(Pol(G), M(K))) HomPol(G)(Pol(G)+, M(K)) Ext1 Pol(G)(C, M(K)) δ0 Pol(G)(Pol(G), M(K)) Ext1 ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ ={0} Ext1 Pol(G)(Pol(G)+, M(K)) ⋯ Splitting this long exact sequence at δ0, we obtain two short exact sequences 0 → HomPol(G)(C, M(K)) Ð→ HomPol(G)(Pol(G), M(K)) Ð→ ker δ0 Ð→ 0 , 0 → ker δ0 Ð→ HomPol(G)(Pol(G)+, M(K)) Ð→ Ext1 Pol(G)(C, M(K)) Ð→ 0 . Since HomPol(G)(Pol(G), M(K)) ≅ M(K) as a right L∞(K)-module, applying the dimension function dimL∞(K) to the first short exact sequence gives 1 = dimL∞(K) HomPol(G)(Pol(G), M(K)) = dimL∞(K) HomPol(G)(C, M(K)) + dimL∞(K) ker(δ0) = β(2) 0 (G) + dimL∞(K) ker(δ0) , 5 Higher ℓ2-Betti numbers by Julien Bichon, David Kyed and Sven Raum where the equality dimL∞(K) HomPol(G)(C, M(K)) = dimL∞(K) Ext0 plained in Section 2. Applying the dimension function to the second exact sequence, we obtain Pol(G)(C, M(K)) = β(2) 0 (G) is ex- dimL∞(K) HomPol(G)(Pol(G)+, M(K)) = dimL∞(K) ker(δ0) + dimL∞(K) Ext1 Pol(G)(C, M(K)) = dimL∞(K) ker(δ0) + β(2) 1 (G), from which the formula follows. Proof of Theorem C. We denote the compact dual of G∗ H [Wan95] by K. Since G and H are assumed non-trivial, the free product Pol(G)∗ Pol(H) is infinite dimensional, so for p = 0 the result follows from [Kye11]. For p ⩾ 2, [Bic16, Theorem 5.1] gives that Extp Pol(K)(C, M(K)) ≃ Extp Pol(G)(C, M(K)) ⊕ Extp Pol(H)(C, M(K)), and since ℓ2-Betti numbers can be calculated by Ext-groups (see Section 2), the result follows from applying dimL∞(K) to both sides. To prove the formula when p = 1, we apply Lemma 3.1 to each of the quantum groups G and H to get 1 (G) + β(2) 1 ( H) − β(2) 2 + β(2) = dimL∞(K) HomPol(G)(Pol(G)+, M(K)) + dimL∞(K) HomPol(H)(Pol(H)+, M(K)) = dimL∞(K) HomPol(K)(Pol(K) ⊗Pol(G) Pol(G)+, M(K)) 0 (G) − β(2) 0 ( H) + dimL∞(K) HomPol(K)(Pol(K) ⊗Pol(H) Pol(H)+, M(K)) = dimL∞(K) HomPol(K) Pol(K) ⊗Pol(G) Pol(G)+ ⊕ Pol(K) ⊗Pol(H) Pol(H)+ , M(K) = dimL∞(K) HomPol(K)(Pol(K)+, M(K)) = 1 + β(2) = 1 + β(2) 1 (G ∗ H) − β(2) 1 (G ∗ H), 0 (G ∗ H) ([Bic16, Lemma 5.8]) (Lemma 3.1) and the formula follows. 4 A scaling formula for cocentral extensions of discrete quantum groups In this section we generalise the considerations of [KR16, Section 2.1], which provides us with a scaling formula for ℓ2-Betti numbers of cocentral extensions of discrete quantum groups by abelian groups. We start by collecting the analogues of [KR16, Lemma 2.1 & 2.2]. Lemma 4.1. Let A the A-module C(Γ) that is induced by the homomorphism A ∋ a ↦ (g ↦ ε ○ αg(a)) ∈ C(Γ). Then ε→ C be an augmented algebra and Γ α ↷ A an action of a finite group. Consider AC(Γ) ≅ ࣷ g∈Γ ε ○ αg C . C denotes C considered as an A-module via the homomorphism ε○αg ∶ A → C. More generally, Here ε ○ αg whenever β ∶ A → B is a ring homomorphism and X is a B-module, we denote by βX the A-module X with module structure defined via β. 6 Higher ℓ2-Betti numbers by Julien Bichon, David Kyed and Sven Raum Proof. The natural direct sum decomposition C(Γ) = ࣷ g∈Γ C1 {g} is compatible with the A-module structure, since C(Γ) is abelian. Because AC1 conclude the lemma. {g} ≅ ε ○ αg C, we can Remark 4.2. If A is Hopf ∗-algebra and Γ is an abelian finite group, one can start out with a Hopf ∗-algebra homomorphism π∶ A → CΓ and consider the induced action Γ ↷ A (cf. Section 2). Then the homomorphism A → C(Γ) constructed in Lemma 4.1 coincides with π after applying the Fourier transform CΓ ≅ C(Γ). Lemma 4.3. Let A state with bounded GNS-representation. Let M = πϕ(A)′′. Then for all p ⩾ 0 ε→ C be an augmented ∗-algebra, α ∈ Aut(A) and ϕ ∈ A∗ an α-invariant tracial dimM TorA p (M, αC) = dimM TorA p (M, C) . Proof. A flat base change [Wei94, Proposition 3.2.9] gives an isomorphism of M-modules TorA p (M, αC) ≅ TorA p (M ⊗A (αA), C) . Note that α extends to an automorphism of M, which provides an isomorphism of left M-modules M ⊗A (αA) ≅ αM . Hence dimM TorA p (M, αC) = dimM TorA p (αM, C) = dimM αTorA p (M, C). The endofunctor X ↦ αX on the category of left M-modules preserves the class of finitely generated projective modules, is dimension preserving on this class and preserves inclusions, hence dimM(X) = dimM(αX) for all M-modules X (see [Lüc02, Section 6.1]). Therefore p (M, C). p (M, αC) = dimM TorA dimM TorA as claimed. Proof of Theorem D. We have p ( H) = dimL∞(H) Tor β(2) Pol(H) p (L∞(H), C) = dimL∞(G) L∞(G) ⊗L∞(H) Tor Pol(H) p (L∞(H), C) , since the functor L∞(G) ⊗L∞(H) − is dimension preserving [Sau02, Theorem 3.18]. This functor is furthermore exact3 [Sau02, Theorem 1.48], and therefore commutes with Tor, so that dimL∞(G) L∞(G) ⊗L∞(H) Tor Pol(H) p (L∞(H), C) = dimL∞(G) Tor Pol(H) p (L∞(G), C) . Since the inclusion Pol(H) ⊂ Pol(G) is flat by [Chi14], we can apply the flat base change formula [Wei94, Proposition 3.2.9], which gives dimL∞(G) Tor Pol(H) p (L∞(G), C) = dimL∞(G) Tor Pol(G) p (L∞(G), Pol(G) ⊗Pol(H) C) . 3For exactness and dimension-preservation in the case of group von Neumann algebras, see [Lüc02, Theorem 6.29] 7 Higher ℓ2-Betti numbers by Julien Bichon, David Kyed and Sven Raum The exactness assumption on our sequence gives in particular Pol(G) ⊗Pol(H) C ≅ Pol(G)~Pol(G)Pol(H)+ ≅ CΓ ≅ C(Γ), as left Pol(G)-modules where the Pol(G)-structure on C(Γ) is defined via the identification with CΓ. The theorem now follows from a combination of Lemmas 4.1, 4.3 and Remark 4.2 together with the fact that the dimension function is additive. 5 Calculations of ℓ2-Betti numbers In this section we apply the formulas obtained in Sections 3 and 4 to the specific examples of cocentral extensions presented in [BNY15, BNY16]. This will provide a complete calculation of the ℓ2-Betti numbers of the universal discrete quantum groups U+ n (all but the second and third were already found in [KR16]) and, furthermore, a complete calculation of the ℓ2-Betti numbers of the duals of the half-liberated unitary quantum groups U∗ n. Proof of Theorem A. As explained in Section 2, we have exact sequences of Hopf ∗-algebras C Ð→ Pol(H) Ð→ Pol(U+ n) Ð→ CZ2 Ð→ C C Ð→ Pol(H) Ð→ Pol(O+ n) ∗ Pol(O+ n) Ð→ CZ2 Ð→ C, for the same compact quantum group H. Applying Theorem C and D this gives p ( U+ β(2) n) = 1 2 p ( H) = β(2) β(2) p ( O+ n ∗ O+ n) = ⎧⎪⎪⎪⎪⎪⎨⎪⎪⎪⎪⎪⎩ 0 2 ⋅ β(2) 2 ⋅ β(2) 1 ( O+ n) − β(2) n) p ( O+ 0 ( O+ p = 0 n) + 1 p = 1 p ⩾ 2 . Since β(2) p ( O+ n) = 0 for all p ⩾ 0 by [Ver12] and [CHT09], Theorem A follows. Proof of Theorem E. By [BNY15, Example 3.7], there exist short exact sequences of Hopf ∗- algebras C Ð→ Pol(H∗) Ð→ Pol(U∗ n) Ð→ CZ2 Ð→ C C Ð→ Pol(H∗) Ð→ Pol(O∗ n) ∗ Pol(O∗ n) Ð→ CZ2 Ð→ C, for the same compact quantum group H∗; the details of this argument are analogous to those sketched in Section 2 for the free unitary quantum groups. By Theorem C and D we therefore obtain p ( U∗ β(2) n) = 1 2 p ( H∗) = β(2) β(2) p ( O∗ n ∗ O∗ n) = 0 ⎧⎪⎪⎪⎪⎪⎨⎪⎪⎪⎪⎪⎩ 2 ⋅ β(2) 2 ⋅ β(2) 1 ( O∗ n) − β(2) n) p ( O∗ 0 ( O∗ p = 0 n) + 1 p = 1 p ⩾ 2 . However, since O∗ [Kye08a] and the result follows. n is infinite and coamenable [BV10, Corollary 9.3], β(2) p ( O∗ n) = 0 for all p ⩾ 0 by 8 Higher ℓ2-Betti numbers References by Julien Bichon, David Kyed and Sven Raum [Ban97a] Teodor Banica. Le groupe quantique compact libre U(n). Comm. Math. Phys., 190(1):143 -- 172, 1997. [BS09] [BV10] Teodor Banica and Roland Speicher. Liberation of orthogonal Lie groups. Adv. Math., 222(4):1461 -- 1501, 2009. Teodor Banica, and Roland Vergnioux. Ann. Inst. Fourier, 60(6):2137 -- 2164, 2010. Invariants of the half-liberated orthogonal group. [BMT01] Erik Bédos, Gerard J. Murphy, and Lars Tuset. Co-amenability of compact quantum groups. J. Geom. Phys., 40(2):130 -- 153, 2001. [vdB98] Michel van den Bergh. A relation between Hochschild homology and cohomology for Goren- stein rings. Proc. Amer. Math. Soc., 126(5): 1345 -- 1348, 1998. [BDD11] Jyotishman Bhowmick, Francesco D'Andrea, and Ludwik D abrowski . Quantum isome- ' tries of the finite noncommutative geometry of the standard model. Comm. Math. Phys., 307(1):101 -- 131, 2011. [Bic16] Julien Bichon. Cohomological dimensions of universal cosoverign Hopf algebras Publicacions Matemàtiques, to appear, arXiv:1611.02069 [BNY15] Julien Bichon, Sergey Neshveyev, and Makoto Yamashita. Graded twisting of categories and quantum groups by group actions Ann. Inst. Fourier, 66 (6):22990 -- 2338, 2016. [BNY16] Julien Bichon, Sergey Neshveyev, and Makoto Yamashita. Graded twisting of comodule algebras and module categories Preprint, arXiv:1604.02078. [Bra12] Michael Brannan. Approximation properties for free orthogonal and free unitary quantum groups. J. Reine Angew. Math., 672:223 -- 251, 2012. [Chi14] Alexandru Chirvasitu. Cosemisimple Hopf algebras are faithfully flat over Hopf subalgebras. Algebra Number Theory, 8(5):1179 -- 1199, 2014. [CHT09] Benoît Collins, Johannes Härtel, and Andreas Thom. Homology of free quantum groups. C. R. Math. Acad. Sci. Paris, 347(5-6):271 -- 276, 2009. [Fim10] Pierre Fima. Kazhdan's property T for discrete quantum groups. Int. J. Math., 21(1):47 -- 65, 2010. [KV00] Johan Kustermans and Stefaan Vaes. Locally compact quantum groups. Ann. Sci. École Norm. Sup. (4), 33(6):837 -- 934, 2000. [Kye08a] David Kyed. L2-Betti numbers of coamenable quantum groups. Münster J. Math., 1(1):143 -- 179, 2008. [Kye08b] David Kyed. L2-homology for compact quantum groups. Math. Scand., 103(1):111 -- 129, 2008. 9 Higher ℓ2-Betti numbers by Julien Bichon, David Kyed and Sven Raum [Kye11] David Kyed. On the zeroth L2-homology of a quantum group. Münster J. Math. 4 119 -- 127, 2011. [Kye12] David Kyed. An L2-Kunneth formula for tracial algebras. J. Oper. Theory, 67(2):317 -- 327, 2012. [KR16] David Kyed and Sven Raum. On the ℓ2-Betti numbers of universal quantum groups. Math. Ann., to appear, arXiv:1610.05474. [Lüc02] Wolfgang Lück. L2-invariants: theory and applications to geometry and K-theory, vol- ume 44 of Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics [Results in Mathematics and Related Areas. 3rd Series. A Series of Modern Surveys in Mathematics]. Springer-Verlag, Berlin, 2002. [MN06] Ralf Meyer and Ryszard Nest. The Baum-Connes conjecture via localisation of categories. Topology, 45(2):209 -- 259, 2006. [Rei01] Holger Reich. On the K- and L-theory of the algebra of operators affiliated to a finite von Neumann algebra. K-Theory, 24(4):303 -- 326, 2001. [Sau02] Roman Sauer. L2-invariants of groups and discrete measured groupoids. PhD Dissertation, University of Münster, 2002. [Tho08] Andreas Thom. L2-cohomology for von Neumann algebras. Geom. Funct. Anal., 18(1):251 -- 270, 2008. [VW96] Alfons Van Daele and Shuzhou Wang. Universal quantum groups. Internat. J. Math., 7(2):255 -- 263, 1996. [Ver07] [Ver12] [Voi11] Roland Vergnioux. The property of rapid decay for discrete quantum groups. J. Oper. Theory, 57(2):303 -- 324, 2007. Roland Vergnioux. Paths in quantum Cayley trees and L2-cohomology. Adv. Math., 229(5):2686 -- 2711, 2012. Christian Voigt. The Baum-Connes conjecture for free orthogonal quantum groups. Adv. Math., 227(5):1873 -- 1913, 2011. [Wan95] Shuzhou Wang. Free products of compact quantum groups. Comm. Math. Phys., 167(3):671 -- 692, 1995. [Wan98] Shuzhou Wang. Quantum symmetry groups of finite spaces. Comm. Math. Phys., 195(1):195 -- 211, 1998. [Wei94] Charles A. Weibel. An introduction to homological algebra, volume 38 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 1994. [Wor87] Stanisław L. Woronowicz. Twisted SU(2) group. An example of a noncommutative differ- ential calculus. Publ. Res. Inst. Math. Sci., 23(1):117 -- 181, 1987. 10 Higher ℓ2-Betti numbers by Julien Bichon, David Kyed and Sven Raum [Wor98] Stanisław L. Woronowicz. Compact quantum groups. In Symétries quantiques (Les Houches, 1995), pages 845 -- 884. North-Holland, Amsterdam, 1998. DAVID KYED, Department of Mathematics and Computer Science, University of Southern Denmark, Campusvej 55, DK-5230 Odense M, Denmark. E-mail address: [email protected] JULIEN BICHON, Campus universitaire des Cézeaux, 3 place Vasarely, 63178 Aubière cedex, France. E-mail address: [email protected] Laboratoire de Mathématiques Blaise Pascal, Université Clermont Auvergne, SVEN RAUM, EPFL SB SMA, Station 8, CH-1015 Lausanne, Switzerland E-mail address: [email protected] 11
1309.3839
1
1309
2013-09-16T07:32:31
Orthogonal forms and orthogonality preservers on real function algebras
[ "math.OA" ]
We initiate the study of orthogonal forms on a real C$^*$-algebra. Motivated by previous contributions, due to Ylinen, Jajte, Paszkiewicz and Goldstein, we prove that for every continuous orthogonal form $V$ on a commutative real C$^*$-algebra, $A$, there exist functionals $\varphi_1$ and $\varphi_2$ in $A^{*}$ satisfying $$V(x,y) = \varphi_1 (x y) + \varphi_2 (x y^*),$$ for every $x,y$ in $A$. We describe the general form of a (not-necessarily continuous) orthogonality preserving linear map between unital commutative real C$^*$-algebras. As a consequence, we show that every orthogonality preserving linear bijection between unital commutative real C$^*$-algebras is continuous.
math.OA
math
ORTHOGONAL FORMS AND ORTHOGONALITY PRESERVERS ON REAL FUNCTION ALGEBRAS JORGE J. GARC´ES AND ANTONIO M. PERALTA Abstract. We initiate the study of orthogonal forms on a real C∗-algebra. Motivated by previous contributions, due to Ylinen, Jajte, Paszkiewicz and Goldstein, we prove that for every continuous orthogonal form V on a commu- tative real C∗-algebra, A, there exist functionals ϕ1 and ϕ2 in A∗ satisfying V (x, y) = ϕ1(xy) + ϕ2(xy∗), for every x, y in A. We describe the general form of a (not-necessarily con- tinuous) orthogonality preserving linear map between unital commutative real C∗-algebras. As a consequence, we show that every orthogonality preserving linear bijection between unital commutative real C∗-algebras is continuous. 1. Introduction and preliminaries Elements a and b in a real or complex C∗-algebra, A, are said to be orthogonal (denoted by a ⊥ b) if ab∗ = b∗a = 0. A bounded bilinear form V : A × A → K is called orthogonal (resp., orthogonal on self-adjoint elements) whenever V (a, b∗) = 0 for every a ⊥ b in A (resp., in the self-adjoint part of A). All the forms considered in this paper are assumed to be continuous. Motivated by the seminal contributions by K Ylinen [51] and R. Jajte and A. Paszkiewicz [29], S. Goldstein proved that every orthogonal form V on a (complex) C∗-algebra, A, is of the form V (x, y) = φ(xy) + ψ(xy) (x, y ∈ A), where φ and ψ are two functionals in A∗ (cf. [21, Theorem 1.10]). A simplified proof of Goldstein's theorem was published by U. Haagerup and N.J. Laustsen in [24]. This characterisation has emerged as a very useful tool in the study of bounded linear operators between C∗-algebras which are orthogonality or disjointness pre- serving (see, for example, [11, 12]). The first aim of this paper is to study orthogonal forms on the wider class of real C∗-algebras. Little or nothing is known about the structure of an orthogonal form V on a real C∗-algebra. At first look, one is tempted to consider the canonical complex bilinear extension of V to a form on the complexification, AC = A ⊕ iA, of A and, when the latter is orthogonal, to apply Goldstein's theorem. However, the complex bilinear extension of V to AC × AC, need not be, in general, orthogonal (see Example 2.7). The study of orthogonal forms on real C∗-algebras requires 2010 Mathematics Subject Classification. Primary 46H40; 4J10, Secondary 47B33; 46L40; 46E15; 47B48. Key words and phrases. Orthogonal form, real C∗-algebra, orthogonality preservers, disjoint- ness preserver, separating map. Authors partially supported by the Spanish Ministry of Economy and Competitiveness, D.G.I. project no. MTM2011-23843, and Junta de Andaluc´ıa grants FQM0199 and FQM3737. 2 J.J. GARC´ES AND A.M. PERALTA a completely independent strategy; surprisingly the resulting forms will enjoy a different structure to that established by S. Goldstein in the complex setting. In section 2 we establish some structure results for orthogonal forms on a general real C∗-algebra, showing, among other properties, that every orthogonal form on a real C∗-algebra extends to an orthogonal form on its multiplier algebra (see Propo- sition 1.3). It is also proved that, for each orthogonal and symmetric form V on a real C∗-algebra, A, there exists a functional φ ∈ A∗ satisfying V (a, b) = φ(ab + ba), for every a, b ∈ A with a = a∗, b∗ = b (cf. Proposition 1.5). In the real setting, the skew-symmetric part of a real C∗-algebra, A, is not determined by the self-adjoint part of A, so the information about the behavior of V on the rest of A is very limited. Section 3 contains one of the main results of the paper: the characterisation of all orthogonal forms on a commutative real C∗-algebra. Concretely, we prove that a form V on a commutative real C∗-algebra A is orthogonal if, and only if, there exist functionals ϕ1 and ϕ2 in A∗ satisfying V (x, y) = ϕ1(xy) + ϕ2(xy∗), for every x, y ∈ A (see Theorem 2.4). Among the consequences, it follows that the complex bilinear extension of V to the complexification of A is orthogonal if, and only if, we can take ϕ2 = 0 in the above representation. We recall that a mapping T : A → B between real or complex C∗-algebras is said to be orthogonality or disjointness preserving (also called separating) whenever a ⊥ b in A implies T (a) ⊥ T (b) in B. The mapping T is bi-orthogonality preserving whenever the equivalence a ⊥ b ⇔ T (a) ⊥ T (b) holds for all a, b in A. As noticed in [13], every bi-orthogonality preserving linear surjection, T : A → B between two C∗-algebras is injective. The study of orthogonality preserving operators between C∗-algebras started with the work of W. Arendt [1] in the setting of unital abelian C∗-algebras. Sub- sequent contributions by K. Jarosz [30] extended the study to the setting of or- thogonality preserving (not necessarily bounded) linear mappings between abelian C∗-algebras. The first study on orthogonality preserving symmetric (bounded) lin- ear operators between general (complex) C∗-algebras is originally due to M. Wolff [49]). Orthogonality preserving bounded linear maps between C∗-algebras, (cf. JB∗-algebras and JB∗-triples were completely described in [11] and [12]. The pioneer works of E. Beckenstein, L. Narici, and A.R. Todd in [8] and [9] (see also [7]) were applied by K. Jarosz to prove that every orthogonality preserving linear bijection between C(K)-spaces is (automatically) continuous (see [30]). More recently, M. Burgos and the authors of this note proved in [13] that every bi- orthogonality preserving linear surjection between two von Neumann algebras (or between two compact C∗-algebras) is automatically continuous (compare [40], [41] for recent additional generalisations). The main goal of section 4 is to describe the orthogonality preserving linear mappings between unital commutative real C∗-algebras (see Theorem 3.2). As a consequence, we shall prove that every orthogonality preserving linear bijection between unital commutative real C∗-algebras is automatically continuous. We shall exhibit some examples illustrating that the results in the real setting are completely independent from those established for complex C∗-algebras. We further give a ORTHOGONAL FORMS AND ORTHOGONALITY PRESERVERS 3 characterisation of those linear mappings between real forms of C(K)-spaces which are bi-orthogonality preserving. 1.1. Preliminary results. Let us now introduce some basic facts and definitions required later. A real C∗-algebra is a real Banach *-algebra A which satisfies the standard C∗-identity, ka∗ak = kak2, and which also has the property that 1 + a∗a is invertible in the unitization of A for every a ∈ A. It is known that a real Banach *-algebra, A, is a real C∗-algebra if, and only if, it is isometrically *-isomorphic to a norm-closed real *-subalgebra of bounded operators on a real Hilbert space (cf. [39, Corollary 5.2.11]). Clearly, every (complex) C∗-algebra is a real C∗-algebra when scalar multiplica- tion is restricted to the real field. If A is a real C∗-algebra whose algebraic com- plexification is denoted by B = A⊕ iA, then there exists a C∗-norm on B extending the norm of A. It is further known that there exists an involutive conjugate-linear ∗-automorphism τ on B such that A = Bτ := {x ∈ B : τ (x) = x} (compare [39, Proposition 5.1.3] or [47, Lemma 4.1.13], and [22, Corollary 15.4]). The dual space the map defined by of a real or complex C∗-algebra A will be denoted by A∗. Leteτ : B∗ → B∗ denote Theneτ is a conjugate-linear isometry of period 2 and the mapping eτ (φ)(b) = φ(τ (b)) (φ ∈ B∗, b ∈ B). (B∗)eτ → A∗ ϕ 7→ ϕA is a surjective linear isometry. We shall identify (B∗)eτ and A∗ without making any explicit mention. When A is a real or complex C∗-algebra, then Asa and Askew will stand for the set of all self-adjoint and skew-symmetric elements in A, respectively. We shall make use of standard notation in C∗-algebra theory. Given Banach spaces X and Y , L(X, Y ) will denote the space of all bounded lin- ear mappings from X to Y . We shall write L(X) for the space L(X, X). Throughout the paper the word "operator" (respectively, multilinear or sesquilinear operator) will always mean bounded linear mapping (respectively bounded multilinear or sesquilinear mapping). The dual space of a Banach space X is always denoted by X ∗. εnxn(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ally Cauchy (w.u.C.) if there exists C > 0 such that for any finite subset F ⊂ N Let us recall that a seriesPn xn in a Banach space is called weakly uncondition- and εn = ±1 we have (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xn∈F conditionally converging if for every w.u.C. seriesPn xn in X, the seriesPn T (xn) is unconditionally convergent in Y , that is, every subseries of Pn T (xn) is norm if, for every w.u.C. seriesPn xn in X, we have kT (xn)k → 0 (compare, for example, converging. It is known that T : X → Y is unconditionally converging if, and only ≤ C. A (linear) operator T : X −→ Y is un- Let us also recall that a Banach space X is said to have Pe lczy´nski's property (V) if, for every Banach space Y , every unconditionally converging operator T : X → Y is weakly compact. [44, page 1257]) The proof of the following elementary lemma is left to the reader. 4 J.J. GARC´ES AND A.M. PERALTA Lemma 1.1. Let X be a complex Banach space, τ : X → X a conjugate-linear period-2 isometry. Then the real Banach space X τ := {x ∈ X : τ (x) = x} satisfies property (V) whenever X does. (cid:3) We shall require, for later use, some results on extensions of multilinear operators. Let X1, . . . , Xn, and X be Banach spaces, T : X1 × · · · × Xn → X a (continuous) n-linear operator, and π : {1, . . . , n} → {1, . . . , n} a permutation. It is known that there exists a unique n-linear extension AB(T )π : X ∗∗ n → X ∗∗ such that for every zi ∈ X ∗∗ αi ) ∈ Xi (1 ≤ i ≤ n), converging to zi in the weak* topology we have and every net (xi 1 × · · · × X ∗∗ i AB(T )π(z1, . . . , zn) = weak*- lim απ(1) · · · weak*- lim απ(n) T (x1 α1, . . . , xn αn ). Moreover, AB(T )π is bounded and has the same norm as T . The extensions AB(T )π coincide with those considered by Arens in [2, 3] and by Aron and Berner for polynomials in [4]. The n-linear operators AB(T )π are usually called the Arens or Aron-Berner extensions of T . Under some additional hypothesis, the Arens extension of a multilinear operator also is separately weak∗ continuous. Indeed, if every operator from Xi to X ∗ j is weakly compact (i 6= j) the Arens extensions of T defined above do not depend on the chosen permutation π and they are all separately weak∗ continuous (see [5], and Theorem 1 in [10]). In particular, the above requirements always hold when every Xi satisfies Pelczynski's property (V ) (in such case X ∗ i contains no copies of c0, therefore every operator from Xi to X ∗ j is unconditionally converging, and hence weakly compact by property (V ), see [43]). When all the Arens extensions of T coincide, the symbol AB(T ) = T ∗∗ will denote any of them. We should note at this point that every C∗-algebra satisfies property (V ) (cf. Corollary 6 in [46]). Since every real C∗-algebra is, in particular, a real form of a (complex) C∗-algebra, it follows from Lemma 1.1 that every real C∗-algebra satisifes property (V ). We therefore have: Lemma 1.2. Let A1, . . . , Ak be real C∗-algebras and let T be a multilinear con- tinuous operator from A1 × . . . × Ak to a real Banach space X. Then T admits k → X ∗∗ which is separately weak∗ a unique Arens extension T ∗∗ : A∗∗ continuous. (cid:3) 1 × . . . × A∗∗ Given a real or complex C∗-algebra, A, the multiplier algebra of A, M (A), is the set of all elements x ∈ A∗∗ such that, for each element a ∈ A, xa and ax both lie in A. We notice that M (A) is a C∗-algebra and contains the unit element of A∗∗. It should be recalled here that A = M (A) whenever A is unital. Proposition 1.3. Let A be a real C∗-algebra. Suppose that V : A × A → R is an orthogonal bounded bilinear form. Then the continuous bilinear form V : M (A) × M (A) → R, V (a, b) := V ∗∗(a, b) is orthogonal. Proof. Let a and b be two orthogonal elements in M (A). Let a[ 1 3 ]) denote the unique element z in M (A) satisfying zz∗z = a (resp., zz∗z = b). We notice that a[ 1 3 ] and b[ 1 3 ] are orthogonal elements in A. Since V is orthogonal, we have 3 ] are orthogonal, so, for each pair x, y in A, a[ 1 3 ] (resp., b[ 1 3 ] and b[ 1 3 ]xa[ 1 3 ]yb[ 1 V (a[ 1 3 ]xa[ 1 3 ], (b[ 1 3 ])∗y(b[ 1 3 ])∗) = 0 ORTHOGONAL FORMS AND ORTHOGONALITY PRESERVERS 5 for every x, y ∈ A. Goldstine's theorem (cf. Theorem V.4.2.5 in [18]) guarantees that the closed unit ball of A is weak*-dense in the closed unit ball of A∗∗. Therefore we can pick two bounded nets (xλ) and (yµ) in A, converging in the weak∗ topology of A∗∗ to (a[ 1 3 ])∗ and b[ 1 We have already mentioned that V ∗∗ : A∗∗ × A∗∗ → R is separately weak∗ 3 ])∗), for every λ and µ, taking continuous. Since 0 = V (a[ 1 limits, first in λ and subsequently in µ, we deduce that 3 ], respectively. 3 ])∗yµ(b[ 1 3 ]xλa[ 1 3 ], (b[ 1 V ∗∗(a[ 1 3 ](a[ 1 3 ])∗a[ 1 3 ], (b[ 1 3 ])∗b[ 1 3 ](b[ 1 3 ])∗)) = V (a, b∗) = 0, which shows that V is orthogonal. (cid:3) Since the multiplier algebra of a real or complex C∗-algebra always has a unit element, Proposition 1.3 allows us to restrict our study on orthogonal bilinear forms on a real C∗-algebra A to the case in which A is unital. A real von Neumann algebra is a real C∗-algebra which is also a dual Banach [28] or [39, §6.1]). Clearly, the self adjoint part of a real von Neumann space (cf. algebra is a JW-algebra in the terminology employed in [25], so every self-adjoint element in a real von Neumann algebra W can be approximated in norm by a [25, finite real linear combination of mutually orthogonal projections in W (cf. Proposition 4.2.3]). We shall explore now the validity in the real setting of some of the results established by S. Goldstein in [21]. Lemma 1.4. Let A be a real von Neumann algebra with unit 1. Suppose that V : A × A → R is a bounded bilinear form. The following are equivalent: (a) V is orthogonal on Asa; (b) V (p, q) = 0, whenever p and q are two orthogonal projections in A; (c) V (a, b) = V (ab, 1) for every a, b ∈ Asa with ab = ba. If any of the above statements holds and V is symmetric, then defining φ1(x) := V (x, 1) (x ∈ A), we have V (a, b) = φ1( ab+ba ), for every a, b ∈ Asa. 2 Proof. Applying the existence of spectral resolutions for self-adjoint elements in a real von Neumann algebra, the argument given by S. Goldstein in [21, Proposition 1.2] remains valid to prove the equivalence of (a), (b) and (c). Suppose now that V is symmetric. Let a =Pm j=1 λjpj be an algebraic element in Asa, where the λj's belong to R and p1, . . . , pm are mutually orthogonal projections in A. Since V is orthogonal, for every projection p ∈ A, we have V (p, 1) = V (p, 1 − p) + V (p, p) = V (p, p). Thus, V (a, a) = mXj=1 λ2 j V (pj, pj) = mXj=1 The (norm) density of algebraic elements in Asa and the continuity of V imply that V (a, a) = V (a2, 1), for every a ∈ Asa. Finally, applying that V is symmetric we have V (a2, 1) + V (b2, 1) + V (ab + ba, 1) = V ((a + b)2, 1) = V (a + b, a + b) = V (a, a) + V (b, b) + 2V (a, b), λ2 j V (pj, 1) = V  mXj=1 λ2 j pj, 1 = V (a2, 1). 6 J.J. GARC´ES AND A.M. PERALTA for every a, b ∈ Asa, and hence V (a, b) = V ( ab+ba 2 , 1), for all a, b ∈ Asa. (cid:3) The above result holds for every monotone σ-complete unital real C∗-algebra A (that is, each upper bounded, monotone increasing sequence of selfadjoint elements of A has a least upper bound). Surprisingly, the final conclusion of the above Lemma can be established for unital real C∗-algebras with independent basic techniques. Proposition 1.5. Let A be a unital real C∗-algebra with unit 1. Suppose that V : A × A → R is an orthogonal, symmetric, bounded, bilinear form. Then defining φ1(x) := V (x, 1) (x ∈ A), we have V (a, b) = φ1( ab+ba ), for every a, b ∈ Asa. 2 Proof. Let a be a selfadjoint element in A. The real C∗-subalgebra, C, of A gener- ated by 1 and a is isometrically isomorphic to the space C(K, R) of all real-valued continuous functions on a compact Hausdorff space K. The restriction of V to C × C is orthogonal, therefore the mapping x 7→ V (x, x) is a 2-homogeneous or- thogonally additive polynomial on C. The main result in [45] implies the existence of a functional ϕa ∈ C ∗ such that V (x, x) = ϕa(x2), for every x ∈ C. It is clear that ϕa(x) = V (x, 1) for every x ∈ C. In particular V (a, a) = ϕa(a2) = V (a2, 1). The argument given at the end of the proof of Lemma 1.4 gives the desired state- ment. (cid:3) The above proposition shows that we can control the form of a symmetric or- thogonal form on the self adjoint part of a (unital) real C∗-algebra. The form on the skew-symmetric part remains out of control for the moment. 2. Orthogonal forms on abelian real C∗-algebras Throughout this section, A will denote a unital, abelian, real C∗-algebra whose complexification will be denoted by B. It is clear that B is a unital, abelian C∗- algebra. It is known that there exists a period-2 conjugate-linear ∗-automorphism τ : B → B such that A = Bτ := {x ∈ B : τ (x) = x} (cf. [47, 4.1.13] and [22, 15.4] or [39, §5.2]). By the commutative Gelfand theory, there exists a compact Hausdorff space K such that B is C∗-isomorphic to the C∗-algebra C(K) of all complex valued continuous functions on K. The Banach-Stone Theorem implies the existence of a homeomorphism σ : K → K such that σ2(t) = t, and τ (a)(t) = a(σ(t)), for all t ∈ K, a ∈ C(K). Real function algebras of the form C(K)τ have been studied by its own right and are interesting in some other settings (cf. [37]). Henceforth, the symbol B will stand for the σ-algebra of all Borel subsets of K, S(K) will denote the space of B-simple scalar functions defined on K, while the Borel algebra over K, B(K), is defined as the completion of S(K) under the supremum norm. It is known that B = C(K) ⊂ B(K) ⊂ C(K)∗∗. The mapping τ ∗∗ : C(K)∗∗ → C(K)∗∗ is a period-2 conjugate-linear ∗-automorphism on B∗∗ = C(K)∗∗. It is easy to see that τ ∗∗(B(K)) = B(K), and hence τ ∗∗B(K) : B(K) → B(K) defines a period-2 conjugate-linear ∗-automorphism on B(K). By an abuse ORTHOGONAL FORMS AND ORTHOGONALITY PRESERVERS 7 of notation, the symbol τ will denote τ , τ ∗∗ and τ ∗∗B(K) indistinctly. It is clear that, for each Borel set B ∈ B, τ (χB ) = χσ(B) . a − a − 1 + a − 1 2 1 ≤ λkχBk < ε. When rXk=1 2 (a + τ (a)), we have Let a be an element in B(K). For each ε > 0, there exist complex numbers a ∈ A is τ -symmetric (i.e. τ (a) = a) then, since a = 1 λkχBk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) λkχσ(Bk )(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) rXk=1 λ1, . . . , λr and disjoint Borel sets B1, . . . , Br such that(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) rXk=1 + λkχσ(Bk )(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) λkχBk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) rXk=1 λkχBk!(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) λkχBk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) τ a − rXk=1 rXk=1 linear combinations of the formPk αkχBk each Borel set B ∈ B and each α ∈ C,(cid:0)αχB + αχσ(B)(cid:1)∗ = 2ℜe(α)(cid:0)2χσ(B)∩B + χσ(B)\B + χB\σ(B)(cid:1) (cid:0)αχB + αχσ(B)(cid:1) +(cid:0)αχB + αχσ(B)(cid:1)∗ Consequently, every element in B(K)τ can be approximated in norm by finite +αkχσ(Bk ) , where α1, . . . , αn are complex numbers and B1, . . . , Bn are mutually disjoint Borel sets. Having in mind that, for = αχB + αχσ(B) , we have 1 2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≤ a − 1 + a − = 2ℜe(α)(cid:0)2χσ(B)∩B + χ(σ(B)\B)∪σ(σ(B)\B)(cid:1) , (cid:0)αχB + αχσ(B)(cid:1) −(cid:0)αχB + αχσ(B)(cid:1)∗ = 2iℑm(α)(cid:0)χB\σ(B) − χσ(B)\B(cid:1) . Suppose now that a ∈ B(K)τ is *-symmetric (i.e. a∗ = a). It follows from the above that a can be approximated in norm by linear combinations of the form < ε. and rXk=1 αkχEk , where αk ∈ R and E1, . . . , Er are mutually disjoint Borel subsets of K with σ(Ei) = Ei. Let b be an element in B(K)τ satisfying b∗ = −b. Similar argu- ments to those given for *-symmetric elements, allow us to show that b can be ap- rXk=1 proximated in norm by finite linear combinations of the form i αk(χEk − χσ(Ek ) ), where αk ∈ R and E1, . . . , Er are mutually disjoint Borel subsets of K with σ(Ei) ∩ Ei = ∅. Lemma 2.1. Let A be a unital, abelian, real C∗-algebra whose complexification is denoted by B = C(K), for a suitable compact Hausdorff space K. Let τ : B → B be a period-2 conjugate-linear ∗-automorphism satisfying A = Bτ and τ (a)(t) = a(σ(t)), for all t ∈ K, a ∈ C(K), where σ : K → K is a period-2 homeomorphism. Then the set N = {t ∈ K : σ(t) 6= t} is an open subset of K, F = {t ∈ K : σ(t) = t} is a closed subset of K and there exists an open subset O ⊂ K maximal with respect to the property O ∩ σ(O) = ∅. Proof. That F is closed follows easily from the continuity of σ, and consequently, N = K/F is open. Let F be the family of all open subsets O ⊆ K such that O ∩ σ(O) = ∅ ordered by inclusion. Let S = {Oλ}λ be a totally ordered subset of F . We shall see that O =Sλ Oλ is an open set which also lies in F , that is, O ∩ σ(O) = ∅. Let us suppose, on the contrary, that there exists t ∈ O ∩ σ(O) 6= ∅. Then there exist λ, β such that t ∈ Oλ and t ∈ σ(Oβ). Since S is totally ordered, Oλ ⊆ Oβ 8 J.J. GARC´ES AND A.M. PERALTA or Oβ ⊆ Oλ. We shall assume that Oλ ⊆ Oβ. Then σ(Oλ) ⊆ σ(Oβ ) and t lies in Oβ ∩σ(Oβ ) = ∅, which is a contradiction. Finally, Zorn's Lemma gives the existence of a maximal element O in F . (cid:3) It should be noticed here that, in Lemma 2.1, O ∪ σ(O) = N , an equality which follows from the maximality of O. Our next lemma analyses the "spectral resolution" of a *-skew-symmetric ele- ment in B(K)τ . Lemma 2.2. In the notation of Lemma 2.1, let B(A) = B(K)τ , let a ∈ B(K)τ and let b be an element in B(A)skew. Then the following statements hold: a) bF = 0; b) For each ε > 0, there exist mutually disjoint Borel sets B1, . . . , Bm ⊂ O and sa, c) For each ε > 0, there exist mutually disjoint Borel sets C1, . . . , Cm ⊂ K and b − real numbers λ1, . . . , λm satisfying (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) real numbers µ1, . . . , µm satisfying σ(Cj ) = Cj, and(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) mXj=1 − χσ(Bj ) )(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) µjχCj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) mXj=1 i λj(χBj < ε; a − < ε. Proof. a) Since b∗ = −b, we have Re(b(t)) = 0, ∀t ∈ K. Now, let t ∈ F , applying σ(t) = t and τ (b) = b we get b(t) = b(σ(t)) = b(t), and hence ℑm(b(t)) = 0. Statements b) and c) follow from the comments prior to Lemma 2.1 and the (cid:3) maximality of O in that Lemma. It is clear that in a commutative real (or complex) C∗-algebra, A, two elements a, b are orthogonal if and only if they have zero-product, that is, ab = 0. Therefore, V (a, b∗) = 0 = V (a, b) whenever V : A × A → R is an orthogonal bilinear form on an abelian real C∗-algebra and a, b are two orthogonal elements in A. We shall make use of this property without an explicit mention. We shall keep the notation of Lemma 2.1 throughout the section. Henceforth, for each C ⊆ O we shall write uC = i (χC − χσ(C) ). The symbol u0 will stand for the element uO . It is easy to check 1 = χF +u0u∗ 0 , where 1 is the unit element in B(K)τ . By Lemma 2.2 a), for each b ∈ B(K)τ skew we have b ⊥ χF , and so b = bu0u∗ 0 . Proposition 2.3. Let K be a compact Hausdorff space, τ a period-2 conjugate- linear isometric ∗-homomorphism on C(K), A = C(K)τ , and V : A × A → R be an orthogonal bounded bilinear form whose Arens extension is denoted by V ∗∗ : A∗∗ × A∗∗ → R. Let σ : K → K be a period-2 homeomorphism satisfying τ (a)(t) = a(σ(t)), for all t ∈ K, a ∈ C(K). Then the following assertions hold for all Borel subsets D, B, C of K with σ(B) ∩ B = σ(C) ∩ C = ∅ and σ(D) = D: a) V (χD , uB ) = V (uB , χD) = 0, whenever D ∩ B = ∅; b) V (uB , uC ) = 0, whenever B ∩ C = ∅; c) V ((u0 u∗ C )uB , uC ) = V (uC , (u0u∗ C )uB ) = 0. 0 − uC u∗ 0 − uC u∗ Proof. By an abuse of notation, we write V for V and V ∗∗. Let K1, K2 be compact subsets of K such that K1, K2 and σ(K2) are mutu- ally disjoint. By regularity and Urysohn's Lemma there exist nets (fλ)λ, (gγ)γ in C(K)+ such that χK1 ≤ gγ ≤ χK\(K1 ∪σ(K1 )∪σ(K2 )) , ≤ fλ ≤ χK\(K2 ∪σ(K2 )) , χK2 ORTHOGONAL FORMS AND ORTHOGONALITY PRESERVERS 9 (fλ)λ (respectively, (gγ)γ ) converges to χK1 of C(K)∗∗. (resp., to χK2 ) in the weak∗ topology 2 (χK1 in the weak∗ topology of C(K)∗∗ to 1 By the separate weak∗ continuity of V ∗∗ ≡ V we have The nets efλ = 1 also clear that fλ ⊥ gγ, τ (fλ) ⊥ gγ, and hence efλ ⊥egγ, for every λ, γ. V (cid:16)ffλ,egγ(cid:17)(cid:19) = 0, 2 (fλ + τ (fλ)) andegγ = i(gγ − τ (gγ)) lie in C(K)τ and converge + χσ(K1 ) ), uK2(cid:19) = w∗ − lim + χσ(K1 ) ) and uK2 V (cid:18) 1 , respectively. It is (χK1 and (1) 2 γ λ (cid:18)w∗ − lim + χσ(K1 ))(cid:19) = 0. , uK2(cid:1) = 0, V (cid:18)uK2 , 1 2 (χK1 V (cid:0)uK1 We can similarly prove that (2) whenever K1 and K2 are two compact subsets of K such that K1, K2, σ(K1) and σ(K2) are pairwise disjoint. B ⊆ O. By inner regularity there exist nets of the form (χ a) Let now D, B be two disjoint Borel subsets of K such that σ(D) = D and )γ such )γ converge in the weak∗ topology of C(K)∗∗ to χD and χB , )λ and (χ )λ and (χ that (χ D λ B γ K K K D λ K B γ D respectively, where each K the assumptions made on D and B we have that K K λ ⊆ D and each K B B γ ⊆ B is compact subset of K. By γ ) = ∅ and λ ∩ σ(K γ = K λ ∩ K D D B B γ ⊆ O for all λ and γ. By (1) and the separate weak∗ continuity of V we have V (χD , uB ) = w∗ − lim λ w∗ − lim γ V χ K D λ + χ 2 σ(K D λ ) , u V (uB , χ D ) = 0. (3) and (4) γ !! = 0, B K A similar argument, but replacing (1) with (2), applies to prove b). To prove the last statement, we observe that (u0u∗ 0 − ucu∗ c )uB = (χO + χσ(O ) − χC − χσ(C) )uB = (χO\C + χσ(O\C) )uB = u(O\C)∩B , and hence the statement c) follows from b). (cid:3) We can now establish the description of all orthogonal forms on a commutative real C∗-algebra. Theorem 2.4. Let V : A × A → R be a continuous orthogonal form on a commu- tative real C∗-algebra, then there exist ϕ1, and ϕ2 in A∗ satisfying V (x, y) = ϕ1(xy) + ϕ2(xy∗), for every x, y ∈ A. 10 J.J. GARC´ES AND A.M. PERALTA Proof. We may assume, without loss of generality, that A is unital (compare Propo- sition 1.3). Let B denote the complexification of A. In this case B identifies with C(K) for a suitable compact Hausdorff space K and A = C(K)τ , where τ is a conjugate-linear period-2 *-homomorphism on C(K). We shall follow the notation employed in the rest of this section. The form V : A×A → R extends to a continuous form V ∗∗ : A∗∗×A∗∗ → R which is separately weak∗ continuous (cf. Lemma 1.2). The restriction V ∗∗B(K)τ ×B(K)τ : B(K)τ × B(K)τ → R also is a continuous extension of V . We shall prove the statement for V ∗∗B(K)τ ×B(K)τ . Henceforth, the symbol V will stand for V , V ∗∗ and V ∗∗B(K)τ ×B(K)τ indistinctly. Let us first take two self-adjoint elements a1, a2 in B(K)τ . By Proposition 1.5, (5) V (a1, a2) = V (a1a2, 1). To deal with the skew-symmetric part, let D, B, C be Borel subsets of K with, D = σ(D) and B, C ⊆ O. From Proposition 2.3 a), we have (6) V (χD , uB ) = V (χD , uB (1 − χD + χD )) = V (χD , uB∩(K\D) ) + V (χD , uB χD ) = V (χD − 1 + 1, uB χD ) = V (−χ(K\D) + 1, u(B∩D) ) = V (1, uB χD ). Similarly, (7) V (uB , χD ) = V (uB χD , 1). Now, Proposition 2.3 b) and c), repeatedly applied give: V (uB , uC ) = V (uB (χF + u0u∗ = V (uB (u0u∗ C − uC u∗ 0 + uC u∗ 0 ), uC ) = V (uB u0u∗ C ), uC ) = V (uB uC u∗ 0 , uC ) C , uC ) = V (uB uC u∗ C , uC − u0 + u0) = V (u(B∩C) , −u(O\C) + u0) = V (u(B∩C) , u0) = V (uB uC (u∗ 0 ), u0) = V (uB uC u∗ C − u∗ 0 + u∗ 0 , u0). Thus, we have (8) and similarly (9) V (uB , uC ) = V (uB uC u∗ 0 , u0), V (uB , uC ) = V (u0, uB uC u∗ 0 ). mlXj=1 plXk=1 Let al = µl,jχ Dl j , bl = λl,ku Bl k (l ∈ {1, 2}) be two simple elements in sa and B(K)τ B(K)τ {Dl of K with σ(Dl 1, . . . , Dl ml} and {Bl skew, respectively, where λl,k, µl,j ∈ R, for each l ∈ {1, 2}, pl} are families of mutually disjoint Borel subsets 1, . . . , Bl j ) = Dl j and Bl i ⊆ O. By (5), (6), (7), and (8), we have V (a1 + b1, a2 + b2) = V (a1a2, 1) + + p1Xk=1 m2Xj=1 µ2,jλ1,kV (cid:18)u , χ B1 k = V (a1a2, 1) + p2Xk=1 p2Xk=1 m1Xj=1 j(cid:19) + p1Xk=1 µ1,jλ2,kV (cid:18)1, χ p2Xk=1 µ1,jλ2,kV (cid:18)χ λ2,kλ1,kV (cid:16)u k(cid:19) D1 j B2 u D2 m1Xj=1 , u D1 j B2 , u B2 B1 k k(cid:19) k(cid:17) ORTHOGONAL FORMS AND ORTHOGONALITY PRESERVERS 11 + p1Xk=1 χ µ2,jλ1,kV (cid:18)u λ2,kλ1,kV (cid:16)u m2Xj=1 = V (a1a2, 1) + V (1, a1b2) + V (b1a2, 1) + V (cid:0)b1b2u∗ = ψ1(a1a2) + ψ2 (a1b2) + ψ1 (b1a2) + ψ4 (b1b2) , , 1(cid:19) + p1Xk=1 p2Xk=1 B1 k B1 k D2 j u∗ 0 , u0(cid:17) u B2 k 0 , u0(cid:1) where ψ1, ψ2, and ψ4 are the functionals in A∗ defined by ψ1(x) = V (x, 1), ψ2(x) = V (1, x), and ψ4(x) = V (xu∗ 0 , u0), respectively. Since, by Proposition 2.2, simple elements of the above form are norm-dense in B(K)τ skew, respectively, and V is continuous, we deduce that sa and B(K)τ V (a1 + b1, a2 + b2) = ψ1(a1a2) + ψ2 (a1b2) + ψ1 (b1a2) + ψ4 (b1b2) , for every a1, a2 ∈ B(K)τ sa, b1, b2 ∈ B(K)τ skew. Now, taking φ1 = 1 4 (2ψ1 + ψ2 + ψ4), φ2 = 1 4 (2ψ1 − ψ2 − ψ4), φ3 = 1 4 (ψ2 − ψ4), and φ4 = 1 4 (ψ4 − ψ2), we get V (a1 + b1, a2 + b2) = φ1((a1 + b1)(a2 + b2)) + φ2 ((a1 + b1)(a2 + b2)∗) +φ3 ((a1 + b1)∗(a2 + b2)) + φ4 ((a1 + b1)∗(a2 + b2)∗) , for every a1, a2 ∈ B(K)τ sa, b1, b2 ∈ B(K)τ skew. Finally, defining ϕ1(x) = φ1(x) + φ4(x∗) and ϕ2(x) = φ2(x) + φ3(x∗) (x ∈ A), (cid:3) we get the desired statement. Remark 2.5. The functionals ϕ1 and ϕ2 appearing in Theorem 2.4 need not be unique. For example, let (ϕ1, ϕ2) and (φ1, φ2) be two couples of elements in the dual of a commutative real C∗-algebra A. It is not hard to check that ϕ1(xy) + ϕ2(xy∗) = φ1(xy) + φ2(xy∗), for every x, y ∈ A if, and only if, ϕ1 + ϕ2 = φ1 + φ2, (ϕ1 − ϕ2)(z) = (φ1 − φ2)(z) and (ϕ1 − ϕ2)(zw) = (φ1 − φ2)(zw), for every z, w ∈ Askew. These conditions are not enough to guarantee that φi = ϕi. Take, for example, A = R ⊕∞ CR, φ1(a, b) = a + ℜe(b) + ℑm(b), φ2(a, b) = 0, ϕ1(a, b) = a 2 + ℜe(b) + ℑm(b), and ϕ2(a, b) = a 2 . Corollary 2.6. Let V : A × A → R be a continuous orthogonal form on a commu- tative real C∗-algebra, then its (unique) Arens extension V ∗∗ : A∗∗ × A∗∗ → R is an orthogonal form. (cid:3) Clearly, the statement of the above Theorem 2.4 doesn't hold for bilinear forms on a commutative (complex) C∗-algebra. The real version established in this paper is completely independent to the result proved by K. Ylinen for commutative com- plex C∗-algebras in [51] and [21]. It seems natural to ask whether the real result follows from the complex one by a mere argument of complexification. Our next ex- ample shows that the (canonical) extension of an orthogonal form on a commutative real C∗-algebra need not be an orthogonal form on the complexification. Example 2.7. Let K = {t1, t2}. We define σ : K → K by σ(t1) = t2. Let A = C(K)τ be the real C∗-algebra whose complexification is C(K) and let V : A × (xy∗) = ℜe(x(t1)y(t1)) = A → R, be the orthogonal form defined by V (x, y) = φt1 ℜe(x(t1)y(t2)), where φt1 = ℜe(δt1 ). In this case, the canonical complex bilinear (xτ (y)∗) = x(t1)y(t2) ) = 1 6= 0, extension eV : C(K) × C(K) → C is given by eV (x, y) = φt1 which implies that eV is not orthogonal. in C(K), however eV (χt1 (x, y ∈ C(K)). It is clear that χt1 ⊥ χt2 , χt2 12 J.J. GARC´ES AND A.M. PERALTA The (complex) bilinear extension of an orthogonal form V on a real C∗-algebra to its complexification is orthogonal precisely when V satisfies the generic form of an orthogonal form on a (complex) C∗-algebra given by the main result in [21]. Corollary 2.8. Let V : A × A → R be a continuous orthogonal form on a commu- tative real C∗-algebra, let B denote the complexification of A and let eV : B ×B → R be the (complex) bilinear extension of V . Then the form eV is orthogonal if, and only if, V writes in the form V (x, y) = ϕ1(xy) (x, y ∈ A), where ϕ1 is a functional in A∗. Proof. Let τ be the period-2 ∗-automorphism on B satisfying that Bτ = B and let V , we get V (a, b) = ℜeφ(ab) = φ(ab), for every a, b ∈ A. In particular, φ(a) ∈ R, eτ : B∗ → B∗ be the involution defined byeτ (φ)(b) = φ(τ (b)). Suppose eV is orthogonal. By the main result in [21] (see also [51]), there exists φ ∈ B∗ satisfying eV (x, y) = φ(xy), for every x, y ∈ B. Since eV is an extension of for every a ∈ A and henceeτ (φ) = φ lies in (B∗)eτ ≡ A∗. eτ (ϕ1) = ϕ1. It is easy to check that eV (x, y) = ϕ1(xy), for every x, y ∈ B. Let us assume that V writes in the form V (x, y) = ϕ1(xy) (x, y ∈ A), where ϕ1 is a functional in A∗. The functional ϕ1 can be regarded as an element in B∗ satisfying (cid:3) 3. Orthogonality preservers between commutative real C∗-algebras Throughout this section, A1 = C(K1)τ1 and A2 = C(K2)τ2 will denote two unital commutative real C∗-algebras, K1 and K2 will be two compact Hausdorff spaces and τi will denote a conjugate-linear period-2 ∗-automorphism on C(Ki) given by τi(f )(t) = f (σi(t)) (t ∈ Ki, f ∈ C(Ki)), where σi : Ki → Ki is a period-2 homeomorphism. We shall write B1 = C(K1) and B2 = C(K2) for the corresponding complexifications of A1 and A2, respectively. By the Banach-Stone theorem, every surjective isometry T : C(K1) → C(K2) is a composition operator, that is, there exist a unitary element u in C(K2) and a homeomorphism σ : K2 → K1 such that T (f )(t) = (uCσ)(f )(t) := u(t) f (σ(t)) (t ∈ K2, f ∈ C(K1)). This result led to the study of the so-called Banach-Stone theorems in different classes of Banach spaces containing C(K)-spaces, in which their algebraic and geometric properties are mutually determined. That is the case of general C∗-algebras (R. Kadison [31] and Paterson and Sinclair [42]), JB- and JB∗-algebras (Wright and M. Youngson [50] and Isidro and A. Rodr´ıguez [27]), JB∗- triples (Kaup [33] and Dang, Friedman and Russo [15]), real C∗-algebras (Grzesiak [23], Kulkarni and Arundhathi [36], Kulkarni and Limaye [37] and Chu, Dang, Russo and Ventura [14]) and real JB∗-triples (Isidro, Kaup and Rodr´ıguez [26], Kaup [34] and Fern´andez-Polo, Mart´ınez and Peralta [19]). In what concerns us, we highlight that any surjective linear isometry T : C(K1)τ1 → C(K2)τ2 is a composition oper- ator given by a homeomorphism φ : K2 → K1 which satisfies σ1 ◦ φ = φ ◦ σ2 (cf. [23] or [36] or [37, Corollary 5.2.4]). The class of orthogonality preserving (continuous) operators between C(K)- spaces is strictly bigger than the class of surjective isometries. Actually, a bounded linear operator T : C(K1) → C(K2) is orthogonality preserving (equivalently, disjointness preserving) if, and only if, there exist u in C(K2) and a mapping ϕ : K2 → K1 which is continuous on {t ∈ K2 : u(t) 6= 0} such that T (f )(t) = (uCϕ)(f )(t) = u(t) f (ϕ(t)) (compare [1, Example 2.2.1]). ORTHOGONAL FORMS AND ORTHOGONALITY PRESERVERS 13 Developing ideas given by E. Beckenstein, L. Narici, and A.R. Todd in [8] and [9] (see also [7]), K. Jarosz showed, in [30], that the above hypothesis of T being continuous can be, in some sense, relaxed. More concretely, for every orthogonality preserving linear mapping T : C(K1) → C(K2), there exists a disjoint decompo- sition K2 = S1 ∪ S2 ∪ S3 (with S2 open, S3 closed), and a continuous mapping ϕ from S1 ∪ S2 into K1 such that T (f )(s) = χ(s)f (ϕ(s)) for all s ∈ S1 (where χ is a continuous, bounded, non-vanishing, scalar-valued function on S1), T (f )(s) = 0 for all s ∈ S3, ϕ(S2) is finite and, for each s ∈ S2, the mapping f 7→ T (f )(s) is not continuous. As a consequence, every orthogonality preserving linear bijec- tion between C(K)-spaces is (automatically) continuous. More recently, M. Burgos and the authors of this note prove, in [13], that every bi-orthogonality preserv- ing linear surjection between two von Neumann algebras (or between two compact C∗-algebras) is automatically continuous (compare [40], [41] for recent additional generalisations). The main goal of this section is to describe the orthogonality preserving lin- ear mappings between C(K)τ -spaces. Among the consequences, we establish that every orthogonality preserving linear bijection between unital commutative real C∗- algebras is automatically continuous. We shall provide an example of an orthogonal- ity preserving linear bijection between C(K)τ -spaces which is not bi-orthogonality preserving and give a characterisation of bi-orthogonality preserving linear maps. We shall borrow and adapt some of the ideas developed in those previously [8, 9] and [30]). In order to have a good balance between mentioned papers (cf. completeness and conciseness, we just give some sketch of the refinements needed in our setting. In any case, the results presented here are independent innovations and extensions of those proved by Beckenstein, Narici, and Todd and Jarosz for C(K)-spaces. Let T : C(K1)τ1 → C(K2)τ2 be an orthogonality preserving linear mapping. Keeping in mind the notation in the previous section, we write Li := Oi ∪ Fi, where Oi and Fi are the subsets of Ki given by Lemma 2.1. The map sending each f in C(Ki)τi to its restriction to Li is a C∗-isomorphism (and hence a surjective linear isometry) from C(Ki)τi onto the real C∗-algebra Cr(Li) of all continuous functions f : Li → C taking real values on Fi. Thus, studying orthogonality preserving linear maps between C(K)τ spaces is equivalent to study orthogonality preserving linear mappings between the corresponding Cr(L)-spaces. Henceforth, we consider an orthogonality preserving (not necessarily continuous) linear map T : Cr(L1) → Cr(L2), where L1 and L2 are two compact Hausdorff spaces and each Fi is a closed subset of Li. Let us consider the sets Z1 = {s ∈ L2 : δsT is a non-zero bounded real-linear mapping}, Z3 = {s ∈ L2 : δsT = 0}, and Z2 = L2\(Z1 ∪ Z3). It is easy to see that Z3 is closed. Following a very usual technique (see, for example, [8, 9, 30, 16] and [17]), we can define a continuous support map ϕ : Z1 ∪ Z2 → L1. More concretely, for each s ∈ Z1 ∪ Z2, we write supp(δsT ) for the set of all t ∈ L1 such that for each open set U ⊆ L1 with t ∈ U there exists f ∈ Cr(L1) with coz(f ) ⊆ U and δs(T (f )) 6= 0. Actually, following a standard argument, it can be shown that, for each s ∈ Z1 ∪ Z2, supp(δsT ) is non-empty and reduces exactly to one point ϕ(s) ∈ L1, and the assignment s 7→ ϕ(s) defines a continuous map from Z1 ∪ Z2 to L1. Furthermore, the value of T (f ) at every s ∈ Z1 depends strictly 14 J.J. GARC´ES AND A.M. PERALTA on the value f (ϕ(s)). More precisely, for each s ∈ Z1 with ϕ(s) /∈ F1, the value T (g)(s) is the same for every function g ∈ Cr(L1) with g ≡ i on a neighborhood of ϕ(s). Thus, defining T (i)(s) := 0 for every s ∈ Z3 ∪ Z2 and for every s ∈ Z1 with ϕ(s) ∈ F1, and T (i)(s) := T (g)(s) for every s ∈ Z1 ∪ Z2 with ϕ(s) /∈ F1, where g is any element in Cr(L1) with g ≡ i on a neighborhood of ϕ(s), we get a (well-defined) mapping T (i) : L2 → C. It should be noticed that "T (i)" is just a symbol to denoted the above mapping and not an element in the image of T . In this setting, the identity T (f )(s) = T (1)(s) ℜef (ϕ(s)) + T (i)(s) ℑmf (ϕ(s)), holds for every s ∈ Z1. Clearly, T (1)(s), T (i)(s) ∈ R, for every s ∈ F2 and T (1)(s)+ T (i)(s) 6= 0, for every s ∈ Z1. The following property also follows from the definition of ϕ by standard argu- ments: Under the above conditions, let s be an element in Z1 ∪ Z2, then (10) δsT (f ) = 0 for every f ∈ Cr(L1) with ϕ(s) /∈ coz(f ). Lemma 3.1. The mapping T (i) is bounded on the set ϕ−1(O1). Furthermore, the inequality holds for all s ∈ Z1 and all f ∈ Cr(L1) with ℜe(f ), ℑm(f ) ≤ 1. T (f )(s) ≤ kT (1)k + sup es∈ϕ−1(O1) T (i)(es) Proof. Arguing by contradiction, we suppose that, for each natural n, there exists sn ∈ ϕ−1(O1) such that T (i)(sn) > n3. The elements s′ ns can be chosen so that ϕ(sn) 6= ϕ(sm) for n 6= m, and consequently we can find a sequence of pairwise disjoint open subsets (Un) of O1 with ϕ(sn) ∈ Un. It is easily seen that we can define a function g = i gn ∈ Cr(L1) with coz(gn) ⊂ Un, 0 ≤ gn ≤ 1 n2 , and gn ≡ 1 n2 on a neighborhood of sn, for all n. By the form of g, and since T is orthogonality preserving, we have T (g)(sn) = n2T (i)(sn) > n for all n, which is absurd. (cid:3) ∞Xn=1 We can easily show now that Z2 is an open subset of L2. With this aim, we consider an element s0 in Z2. We can pick a function f ∈ Cr(L1) such that kf k ≤ 1 and T (f )(s0) > 1 + kT (1)k + sup es∈ϕ−1(O1) T (i)(es). Since T (f )(s) ≤ kT (1)k+ sup es∈ϕ−1(O1) we conclude that there exists an open neighborhood of s0 contained in Z2. The next theorem resumes the above discussion. T (i)(es) < T (f )(s0)−1, for every s ∈ Z1∪Z3, Theorem 3.2. In the notation above, let T : Cr(L1) → Cr(L2) be an orthogonality preserving linear mapping. Then L2 decomposes as the union of three mutually disjoint subsets Z1, Z2, and Z3, where Z2 is open and Z3 is closed, there exist a continuous support map ϕ : Z1 ∪ Z2 → L1, and a bounded mapping T (i) : L2 → C which is continuous on ϕ−1(O1) satisfying: T (i)(s) ∈ R (∀s ∈ F2), T (i)(s) = 0, (∀s ∈ Z3 ∪ Z2 and ∀s ∈ Z1 with ϕ(s) ∈ F1), (11) T (1)(s) + T (i)(s) 6= 0, (∀s ∈ Z1), ∞Xn=1 ∞Xn=1,n6=n0 fn. (cid:3) ORTHOGONAL FORMS AND ORTHOGONALITY PRESERVERS 15 (12) T (f )(s) = T (1)(s) ℜef (ϕ(s)) + T (i)(s) ℑmf (ϕ(s)), (∀s ∈ Z1, f ∈ Cr(L1)), T (f )(s) = 0, (∀s ∈ Z3, f ∈ Cr(L1)), and for each s ∈ L2, the mapping Cr(L1) → C, f 7→ T (f (s)), is unbounded if, and only if, s ∈ Z2. Furthermore, the set ϕ(Z2) is finite. Proof. Everything has been substantiated except perhaps the statement concerning the set ϕ(Z2). Arguing by contradiction, we assume the existence of a sequence (sn) in Z2 such that ϕ(sn) 6= ϕ(sm) for every n 6= m. Find a sequence (Un) of mutually disjoint open subsets of L1 satisfying ϕ(sn) ∈ Un and a sequence (fn) ⊆ Cr(L1) such that kfnk ≤ 1 n , coz(fn) ⊆ Un and δsn T (fn) > n, for every n ∈ N. The element f = fn lies in Cr(L1), and for each natural n0, fn0 ⊥ Thus, δsn0 T (f ) ≥ δsn0 T (fn0) > n0, which is impossible. Remark 3.3. The mapping T (i) : L2 → C has been defined to satisfy T (i)(s) = 0, for all s ∈ Z3 ∪ Z2 and for all s ∈ Z1 with ϕ(s) ∈ F1. It should be noticed here that the value T (i)(s) is uniquely determined only when s ∈ Z1 and ϕ(s) /∈ F1. There are some other choices for the values of T (i)(s) at s ∈ Z3 ∪ Z2 and at s ∈ Z1 with ϕ(s) ∈ F1 under which conditions (11) and (12) are satisfied. Remark 3.4. We shall now explore some of the consequences derived from Theo- rem 3.2. Let T : Cr(L1) → Cr(L2) be an orthogonality preserving linear mapping. (a) The set Z3 is empty whenever T is surjective. (b) Z3 = ∅ implies that Z1 = L2\Z2 is a compact subset of L2. (c) ϕ(Z2) is a finite set of non-isolated points in L1. Indeed, if ϕ(s0) = t0 is isolated for some s0 ∈ Z2, then we can find an open set U ⊆ L1 such that U ∩K1 = {t0}. Therefore, for each f ∈ Cr(L1) with f (t0) = 0 we have δs0 T (f ) = 0. Pick an arbitrary h ∈ Cr(L1). Clearly, χt0 lies in Cr(L1) if, and only if, t0 /∈ F1. Therefore, ∈ Cr(L1), while iχt0 h0 = ℜe(h(t0))χt0 lies in Cr(L1) and (h − h0)(t0) = 0. + ℑm(h(t0)) iχt0 Assume first that t0 /∈ F1. Denoting λ0 = δs0 T (χt0 we have ) and µ0 = δs0 T (iχt0 ), δs0 T (h) = δs0 T (h0) = λ0ℜe(h(t0)) + µ0ℑm(h(t0)) = λ0 − iµ0 2 This shows that δs0T = λ0−iµ0 Cr(L1) to C, which is impossible. 2 λ0 + iµ0 δt0(h) + 2 δt0 + λ0+iµ0 δt0 is a continuous mapping from δt0 (h). 2 When t0 ∈ F1 we have δs0 T = λ0δt0 is a continuous mapping from Cr(L1) to R, which is also impossible. (d) T surjective implies ϕ(Z1 ∩ O2) ⊆ O1. Suppose, on the contrary that there exists s0 ∈ Z1 ∩ O2 with ϕ(s0) ∈ F1. By (12), T (f )(s0) = T (1)(s0)ℜef (ϕ(s0)), for every f ∈ Cr(L1). It follows from the surjectivity of T , together with the condition s0 ∈ O2, that for every complex number ω there exists a real λ satisfying ω = T (1)(s0)λ, which is impossible. 16 J.J. GARC´ES AND A.M. PERALTA (e) Suppose T is surjective and fix s0 ∈ Z1 ∩ O2. The mapping δs0T is a bounded real-linear mapping from Cr(L1) onto C. On the other hand, by (12), δs0 T (f ) = T (1)(s0)ℜef (ϕ(s0)) + T (i)(s0)ℑmf (ϕ(s0)), (∀f ∈ Cr(L1)). Thus, T being surjective implies that the space CR = R × R is linearly spanned by the elements T (1)(s0) and T (i)(s0). Therefore, for each s0 ∈ Z1 ∩ O2, the set {T (1)(s0), T (i)(s0)} is a basis of CR = R × R. Consequently, when T is surjective and s0 ∈ Z1 ∩ O2, the condition T (f )(s0) = 0 implies f (ϕ(s0)) = 0. For any other s1 ∈ Z1 ∩ O2 with ϕ(s0) = ϕ(s1), we have: T (f )(s0) = 0 ⇒ f (ϕ(s0)) = 0 ⇒ T (f )(s1) = 0. The fact that Cr(L2) separates points implies that s1 = s0. Thus, ϕ is injective on Z1 ∩ O2. We can now state the main result of this section which affirms that every orthog- onality preserving linear bijection between unital commutative real C∗-algebras is (automatically) continuous. Theorem 3.5. Every orthogonality preserving linear bijection between unital com- mutative (real) C∗-algebras is (automatically) continuous. Proof. Since T is surjective, Z3 = ∅, and hence Z1 = L2\Z2 is a compact subset of L2. It is also clear that ϕ(L2) is compact. We claim that ϕ(L2) = L1. Otherwise, there would exist a non-zero function f ∈ Cr(L1) with coz(f ) ⊆ L1\ϕ(L2). Thus, by (10), T (f ) = 0, contradicting the injectivity of T . By Remark 3.4(c), ϕ(Z1) = ϕ(Z1) = ϕ(L2) = ϕ(Z1) ∪ ϕ(Z2) = L1. We next see that Z2 = ∅. Otherwise we can take g ∈ Cr(L2) with ∅ 6= coz(g) ⊂ Z2. Let h = T −1(g). Obviously T h(s) = 0 whenever s ∈ Z1. We claim that h(t) = 0, for every t ∈ ϕ(Z1) \ ϕ(Z2). Let us fix t ∈ ϕ(Z1) \ ϕ(Z2). Since ϕ(Z2) is a finite set there are disjoint open sets U1, U2 such that t ∈ U1, ϕ(Z2) ⊂ U2. Let f ∈ C(L1, R) be such that f (t) 6= 0 and coz(f ) ⊂ U1. We see that T (f h) = 0. Indeed, let s ∈ L2 = Z1 ∪ Z2. If s lies in Z1, then the maps f h and f (ϕ(s))h lie in Cr(L1) and coincide at ϕ(s). Since T is linear over R and f takes real values, we deduce, by (12), that T (f h)(s) = f (ϕ(s))T h(s) = 0. If s ∈ Z2 then, since ϕ(s) /∈ coz(f h), then δsT (f h) = T (f h)(s) = 0. We have shown that T (f h) = 0. Thus, since T is injective, f h = 0 and therefore h(t) = 0. We have therefore proved that coz(h) ⊂ ϕ(Z2) which is a finite set. This means that h must be a finite linear combination of characteristic function on points of ϕ(Z2) and these points must be isolated which is impossible, since by c) in Remark 3.4 no point in ϕ(Z2) can be isolated. We have proved that Z2 = ∅. Now the fact that T is continuous follows easily. (cid:3) The above theorem is the first step toward extending, to the real setting, those results proved in [30], [6], [13], [40], [38] and [48] for (complex) C∗-algebras. Orthogonality preserving linear bijections enjoy an interesting additional prop- erty. Proposition 3.6. In the notation of this section, let T : Cr(L1) → Cr(L2) be an orthogonality preserving linear bijection. Then T −1 preserves invertible elements, that is, T −1(g) is invertible whenever g is an invertible element in Cr(L2). ORTHOGONAL FORMS AND ORTHOGONALITY PRESERVERS 17 Proof. Take an invertible element g ∈ Cr(L2). Let f be the unique element in Cr(L1) satisfying T (f ) = g. Theorem 3.2 implies that 0 6= g(s) = T (f )(s) = T (1)(s) ℜef (ϕ(s)) + T (i)(s) ℑmf (ϕ(s)), for every s ∈ Z1. This assures that f (ϕ(s)) 6= 0, for every s ∈ Z1, and since ϕ(Z1) = L1, f = T −1(g) must be invertible in Cr(L1). (cid:3) In the setting of complex Banach algebras, it follows from the Gleason-Kahane- Zelazko theorem that a linear transformation φ from a unital, commutative, com- plex Banach algebra A into C satisfying φ(1) = 1 and φ(a) 6= 0 for every invertible element a in A is multiplicative, that is, φ(ab) = φ(a)φ(b) (see [20, 32]). Although, the Gleason-Kahane- Zelazko theorem fails for real Banach algebras, S.H. Kulka- rni found in [35] the following reformulation: a linear map φ from a real unital Banach algebra A into the complex numbers is multiplicative if ϕ(1) = 1 and φ(a)2 + φ(b)2 6= 0 for every a, b ∈ A with ab = ba and a2 + b2 invertible. It is not clear that statement (b) in the above proposition can be improved to get the hypothesis of Kulkarni's theorem. The structure of orthogonality preserving linear mappings between Cr(L)-spaces described in Theorem 3.2 invites us to affirm that they are not necessarily multiplicative. 3.1. Bi-orthogonality preservers. As a consequence of the description of orthog- onality preserving linear maps given in [30], it can be shown that an orthogonality preserving linear bijection between (complex) C(K)-spaces is bi-orthogonality pre- serving. It is natural to ask wether every orthogonality preserving linear bijection between commutative (unital) real C∗-algebras is bi-orthogonality preserving. This is known to be true in two cases: first, between spaces CR(K) of real (and also complex) valued functions on a compact Hausdorff space K, as it is well-known; second, between spaces of the type CR(K; Rn) (compare [16, Section 3]). Spaces like those we are dealing with in this paper need not satisfy this property, that is, there exists an orthogonality preserving linear bijection T : Cr(L1) → Cr(L2) which is not bi-orthogonality preserving (and even L1 and L2 are not homeomorphic either). Example 3.7. Let L1 = {t1, t2, t3} L2 = {s1, s2, s3, s4} with O1 = {t1, t3}, O2 = {s1}, F1 = {t2} and F2 = {s2, s3, s4}. Define ϕ : L2 → L1 by ϕ(si) = ti, for i = 1, 2, and ϕ(si) = t3, for i = 3, 4. It is easy to check that T (f )(si) = f (ϕ(si)) if i = 1, 2, and T (f )(s3) = ℜef (t3), T (f )(s4) = ℑmf (t3) is an orthogonality preserving linear bijection, but T −1 is not orthogonality preserving. In the above example, ϕ−1(O1) ∩ F2 is non-empty. Our next result shows that a topological condition on F2 assures that an orthogonality preserving linear bijection between unital commutative real C∗-algebras is bi-orthogonality preserving. Proposition 3.8. In the notation of this section, let T : Cr(L1) → Cr(L2) be an orthogonality preserving linear bijection (not assumed to be bounded). The following statements hold: (a) If T is bi-orthogonality preserving then ϕ : L2 → L1 is a (surjective) homeo- morphism, ϕ(F2) = F1, and ϕ(O2) = O1. In particular, ϕ−1(O1) ∩ F2 = ∅. (b) If F2 has empty interior then T is biorthogonality preserving. Proof. (a) If T is bi-orthogonality preserving, it can be easily seen that ϕ : L2 → L1 is a homeomorphism, and for each s ∈ L2, supp(δϕ(s)T −1) = {s}. By Remark 18 J.J. GARC´ES AND A.M. PERALTA 3.4(d), applied to T and T −1, we have ϕ(F2) = F1 and ϕ(O2) = O1. Then ϕ−1(O1) ∩ F2 = ∅. So, a) is clear. (b). Let us assume that F2 has empty interior. Arguing by contradiction we suppose that T −1 is not orthogonality preserving. Then there exist f1, f2 ∈ Cr(L1) with f1f2 6= 0, but T (f1) ⊥ T (f2). Thus U := coz(f1) ∩ coz(f2) is a non-empty open subset of L1. Keeping again the notation of Theorem 3.2 for T , we recall that, by Theorem 3.5 and Remark 3.4, Z3 = ∅, Z2 = ∅, ϕ(L2) = L1, ϕ(O2) ⊂ O1, ϕO2 is injective, and for each s ∈ O2, and {T (1)(s), T (i)(s)} is a basis of CR = R × R. By the form of T , there are no points of ϕ(O2) in U = coz(f1) ∩ coz(f2) (because for each s ∈ O2, T (f )(s) 6= 0 when f (ϕ(s)) 6= 0). Now, let k be a non-zero element in C(L1, R), with coz(k) ⊆ coz(f1) ∩ coz(f2). By Theorem 3.2 (12), it is clear that ϕ(coz(T (k))) ⊆ coz(k), and hence, since ϕ(O2) ⊆ O1, coz(T (k)) is a non-empty subset of F2, against our hypotheses. (cid:3) As we have already seen, an orthogonality preserving linear bijection between Cr(L)-spaces needs not to be biorthogonality preserving. Example 3.7 also shows that, unlike in the complex case, the existence of an orthogonality preserving linear bijection between Cr(L)-spaces does not guarantee that the corresponding compacts spaces are homeomorphic. We next provide a characterisation of those (linear) mappings which are bi-orthogonality preserving. As a consequence, we shall see that if there exists a bi-orthogonality preserving linear map T : Cr(L1) → Cr(L2) then L1 and L2 are homeomorphic. Theorem 3.9. Let T : Cr(L1) → Cr(L2) be a mapping. The following statements are equivalent: (a) T is a bi-orthogonality preserving linear surjection; (b) There exists a (surjective) homeomorphism ϕ : L2 → L1 with ϕ(O2) = O1, a function a1 = γ1 + iγ2 in Cr(L2) with a1(s) 6= 0 for all s ∈ L2, and a function a2 = η1 + iη2 : L2 → C continuous on O2 with the property that s∈O2(cid:12)(cid:12)(cid:12)(cid:12)det(cid:18) γ1(s) γ2(s) η1(s) η2(s) (cid:19)(cid:12)(cid:12)(cid:12)(cid:12) ≤ sup s∈O2(cid:12)(cid:12)(cid:12)(cid:12)det(cid:18) γ1(s) γ2(s) η1(s) η2(s) (cid:19)(cid:12)(cid:12)(cid:12)(cid:12) < +∞, T (f )(s) = a1(s) ℜef (ϕ(s)) + a2(s) ℑmf (ϕ(s)) 0 < inf such that for all s ∈ L2 and f ∈ Cr(L1). Proof. (a) ⇒ (b). Since every bi-orthogonality preserving linear mapping is in- jective, we can assume that T : Cr(L1) → Cr(L2) is a bi-orthogonality preserving linear bijection. We keep the notation given in Theorem 3.2. We have already shown that Z3 = ∅, Z2 = ∅, ϕ : L2 → L1 is a surjective homeomorphism, ϕ(O2) = O1, and for each s ∈ O2, {T (1)(s), T (i)(s)} is a basis of CR = R × R (compare The- orem 3.5, Remark 3.4 and Proposition 3.8). Taking a1 = T (1) = γ1 + iγ2 and a2 = T (i) = η1 + iη2 we only have to show that s∈O2(cid:12)(cid:12)(cid:12)(cid:12)det(cid:18) γ1(s) Let us denote Ms =(cid:18) γ1(s) T (f )(s) = Ms ·(cid:18) ℜef (ϕ(s)) η2(s) (cid:19)(cid:12)(cid:12)(cid:12)(cid:12) ≤ sup η2(s) (cid:19)(cid:12)(cid:12)(cid:12)(cid:12) < +∞. iη2(s) (cid:19) . Clearly det(Ms) 6= 0, for every s ∈ O2 and ℑmf (ϕ(s)) (cid:19) , for every f ∈ Cr(L1), s ∈ L2. By the boundedness s∈O2(cid:12)(cid:12)(cid:12)(cid:12)det(cid:18) γ1(s) 0 < inf iγ2(s) γ2(s) γ2(s) η1(s) η1(s) η1(s) ORTHOGONAL FORMS AND ORTHOGONALITY PRESERVERS 19 of T (1) : L2 → C and T (i)O2 : O2 → C (see Lemma 3.1) there exists M > 0 such that det(Ms) ≤ M for all s ∈ O2. Applying the above arguments to the mapping T −1 we find a surjective home- omorphism ψ = ϕ−1 : L1 → L2, a mapping T −1(i) : L1 → L2 and m > 0, such that ψ(O1) = O2, for each t ∈ O1, {T −1(1)(t), T −1(i)(t)} is a basis of CR = R × R, T −1(g)(t) = Nt ·(cid:18) ℜeg(ψ(t)) t ∈ O1, where Nt = (cid:18) ℜeT −1(1)(t) ℑmg(ψ(t)) (cid:19) (g ∈ Cr(L2), t ∈ L1), det(Nt) ≤ m, for all iℑmT −1(i)(t) (cid:19) . It can be easily seen ℜeT −1(i)(t) iℑmT −1(1)(t) that, for each s ∈ O2, Nϕ(s) = M −1 s ∈ O2. s , which shows that det(Ms) ≥ 1 m , for all (b) ⇒ (a). Let T : Cr(L1) :→ Cr(L2) be a mapping satisfying the hypothesis in (b). Clearly, T is linear, and since ϕ(F2) = F1, T f (s) ∈ R for all s ∈ F2 and f ∈ Cr(L1) (that is, T (f ) ∈ Cr(L2)). We can easily check that, under these hypothesis, T is injective and preserves orthogonality. We shall now prove that T is surjective. Indeed, for each s ∈ O2 T (f )(s) =(cid:18) ℜeg(s) ℑmg(s) (cid:19) =(cid:18) γ1(s) iγ2(s) η1(s) iη2(s) (cid:19) ·(cid:18) ℜef (ϕ(s)) ℑmf (ϕ(s)) (cid:19) thus, = Ms ·(cid:18) ℜef (ϕ(s)) ℑmf (ϕ(s)) (cid:19) , (cid:18) ℜef (ϕ(s)) ℑmf (ϕ(s)) (cid:19) = M −1 ·(cid:18) ℜeg(s) ℑmg(s) (cid:19) . We define b1(t) : L1 → C and b2 : O1 → C by b1(t) = eγ1(t) + ieγ2(t) and ϕ−1(t) = (cid:18) eγ1(t) ieη2(t) (cid:19) , and b1(t) = eη1(t) b2 = eη1(t) + ieη2(t) (t ∈ O1), where M −1 ieγ2(t) γ1(ϕ−1(t)) , for every t ∈ F1. Then S : Cr(L2) → Cr(L1), defined by S(g)(t) = b1(t) ℜeg(ϕ−1(t)) + b2(t) ℑmg(ϕ−1(t)), is linear, preserves orthogonality and it is easy to check that S = T −1. It follows that T is bi-orthogonality preserving. (cid:3) 1 s Let T be a bi-orthogonality preserving linear mapping with associated homeo- morphism ϕ : L2 → L1. Clearly, the operator S : Cr(L1) → Cr(L2), S(f )(s) := f (ϕ(s)) is a ∗-isomorphism. Having in mind that a linear mapping T : A → B between real C∗-algebras is a ∗-isomorphism if, and only if, the complex linear ex- tension eT : A ⊕ iA → B ⊕ iB, eT (a + ib) = T (a) + iT (b) is a ∗-isomorphism, we get the following corollary. Corollary 3.10. The following statements are equivalent: (a) There exists a bi-orthogonality preserving linear bijection T : Cr(L1) → Cr(L2); (b) There exists a C∗-isomorphism S : Cr(L1) → Cr(L2); (c) There exists a C∗-isomorphism eS : C(L1) → C(L2); (d) L1 and L2 are homeomorphic. Acknowledgements: The authors gratefully thank to the Referee for the con- structive comments and detailed recommendations which definitely helped to im- prove the readability and quality of the paper. (cid:3) 20 J.J. GARC´ES AND A.M. PERALTA References [1] W. Arendt, Spectral properties of Lamperti operators, Indiana Univ. Math. J. 32, no. 2, 199-215 (1983). [2] R. Arens, Operations induced in function classes, Monatsh. Math. 55, 1-19 (1951). [3] R. Arens, The adjoint of a bilinear operation, Proc. Amer. Math. Soc. 2, 839-848 (1951). [4] R.M. Aron, P.D. Berner, A Hahn-Banach extension theorem for analytic mappings, Bull. Soc. Math. France. 106, 3-24 (1978). [5] R.M. Aron, B.J. Cole, T.W. Gamelin, Spectra of algebras of analytic functions on a Banach space, J. Reine Angew. Math. 415, 51-93 (1991). [6] J. Araujo, K. Jarosz, Biseparating maps between operator algebras, J. Math. Anal. Appl. 282, no. 1, 48-55 (2003). [7] E. Beckenstein, L. Narici, Automatic continuity of certain linear isomorphisms, Acad. Roy. Belg. Bull. Cl. Sci. (5) 73, no. 5, 191-200 (1987). [8] E. Beckenstein, L. Narici, A.R. Todd, Variants of the Stone-Banach theorem, preprint. [9] E. Beckenstein, L. Narici, A.R. Todd, Automatic continuity of linear maps on spaces of con- tinuous functions, Manuscripta Math. 62, no. 3, 257-275 (1988). [10] F. Bombal, I. Villanueva, Multilinear operators on spaces of continuous functions, Funct. Approx. Comment. Math. 26, 117-126 (1998). [11] M. Burgos, F.J. Fern´andez-Polo, J.J. Garc´es, J. Mart´ınez Moreno, A.M. Peralta, Orthog- onality preservers in C*-algebras, JB*-algebras and JB*-triples, J. Math. Anal. Appl. 348, 220-233 (2008). [12] M. Burgos, F.J. Fern´andez-Polo, J.J. Garc´es, A.M. Peralta, Orthogonality preservers revis- ited, Asian-Eur. J. Math. 2 (3), 387-405 (2009). [13] M. Burgos, J. Garc´es, A.M. Peralta, Automatic continuity of bi-orthogonality preservers between compact C*-algebras and von Neumann algebras, J. Math. Anal. Appl. 376, 221-230 (2011). [14] Ch.-H. Chu, T. Dang, B. Russo, B. Ventura, Surjective isometries of real C*-algebras, J. London Math. Soc. 47, 97-118 (1993). [15] T. Dang, Y. Friedman, B. Russo, Affine geometric proofs of the Banach-Stone theorems of Kadison and Kaup, Proceedings of the Seventh Great Plains Operator Theory Seminar (Lawrence, KS, 1987). Rocky Mountain J. Math. 20, no. 2, 409-428 (1990). [16] L. Dubarbie, Separating maps between spaces of vector-valued absolutely continuous func- tions, Canad. Math. Bull. 53, no. 3, 4466-474 (2010). [17] L. Dubarbie, Maps preserving common zeros between subspaces of vector-valued continuous functions, Positivity, vol. 14, no. 4, 695-703, (2010). [18] N. Dunford, J.T. Schwartz, Linear operators, Part I, Interscience, New York, (1958). [19] F.J. Fern´andez-Polo, J. Mart´ınez, A.M. Peralta, Surjective isometries between real JB∗- triples, Math. Proc. Cambridge Philos. Soc. 137, no. 3, 703-723 (2004). [20] A.M. Gleason, A characterization of maximal ideals, J. Analyse Math. 19, 171-172 (1967). [21] S. Goldstein, Stationarity of operator algebras, J. Funct. Anal. 118, no. 2, 275-308 (1993). [22] K.R. Goodearl, Notes on Real and Complex C∗-algebras, Shiva Publ., 1982. MR 85d:46079. [23] M. Grzesiak, Isometries of a space of continuous functions determined by an involution, Math. Nachr. 145, 217-221 (1990). [24] U. Haagerup, N.J. Laustsen, Weak amenability of C ∗-algebras and a theorem of Goldstein, In Banach algebras '97 (Blaubeuren), 223-243, de Gruyter, Berlin, 1998. [25] H. Hanche-Olsen, E. Størmer, Jordan operator algebras, Monographs and Studies in Mathe- matics, 21. Pitman (Advanced Publishing Program), Boston, MA, 1984. [26] J.M. Isidro, W. Kaup, A. Rodr´ıguez, On real forms of JB*-triples, Manuscripta Math. 86, 311-335 (1995). [27] J.M. Isidro, A. Rodr´ıguez, Isometries of JB-algebras, Manuscripta Math. 86, 337-348 (1995). [28] J.M. Isidro, A.R. Palacios, On the definition of real W ∗-algebras, Proc. Amer. Math. Soc. 124, 3407-3410 (1996). [29] R. Jajte, A. Paszkiewicz, Vector measures on the closed subspaces of a Hilbert space, Studia Math. 63, no. 3, 229-251 (1978). [30] K. Jarosz, Automatic continuity of separating linear isomorphisms, Canad. Math. Bull. 33, no. 2, 139-144 (1990). [31] R.V. Kadison, Isometries of Operator Algebras, Ann. of Math. 54, 325-338 (1951). ORTHOGONAL FORMS AND ORTHOGONALITY PRESERVERS 21 [32] J.P. Kahane, W. Zelazko, A characterization of maximal ideals in commutative Banach al- gebras, Studia Math. 29, 339-343 (1968). [33] W. Kaup, A Riemann Mapping Theorem for bounded symmentric domains in complex Ba- nach spaces, Math. Z. 183, 503-529 (1983). [34] W. Kaup, On real Cartan factors, Manuscripta Math. 92, 191-222 (1997). [35] S.H. Kulkarni, Gleason-Kahane- Zelazko theorem for real Banach algebras, J. Math. Phys. Sci. 18, Special Issue, S19-S28 (1983/84). [36] S.H. Kulkarni, S. Arundhathi, Isometries of real function algebras, Comment. Math. Prace Mat. 30, no. 2, 343-356 (1991). [37] S.H. Kulkarni, B.V. Limaye, Real Function Algebras, Monographs and Textbooks in Pure and Applied Mathematics, 168. Marcel Dekker, 1992. MR1197884 (93m:46059) [38] C.-W. Leung, Ch.-W. Tsai, N.-Ch. Wong, Separating linear maps of continuous fields of Banach spaces, Asian-European J. Math. 2, 3, 445-452 (2009). [39] B.R. Li, Real operator algebras, World Scientific Publishing (Singapore), 2003. [40] T. Oikhberg, A.M. Peralta, M. Ram´ırez, Automatic continuity of M-norms on C*-algebras, J. Math. Anal. Appl., 381, 799-811 (2011). [41] T. Oikhberg, A.M. Peralta, D. Puglisi, Automatic continuity of L-norms on the predual of a von Neumann algebra, to appear in Revista Matem´atica Complutense (Springer). [42] A.L.T. Paterson, A.M. Sinclair, Characterisation of isometries between C∗-algebras, J. Lon- don Math. Soc. 5, 755-761 (1972). [43] A. Pe lczy´nski, Banach spaces on which every unconditionally converging operator is weakly compact. B ull. Acad. Polon. Sci. S´er. Sci. Math. Astronom. Phys., 10 641-648 (1962). [44] A.M. Peralta, I. Villanueva, J. D. M. Wright, K. Ylinen, Weakly compact operators and the strong∗ topology for a Banach space, Proc. R. Soc. Edinb. 140A, 1249-1267 (2010). [45] D. P´erez, I. Villanueva, Orthogonally additive polynomials on spaces of continuous functions, J. Math. Anal. Appl. 306, 97-105, (2005). [46] H. Pfitzner, Weak compactness in the dual of a C*-algebra is determined commutatively, Math. Ann. 298, no. 2, 349-371 (1994). [47] C.E. Rickart, General theory of Banach algebras, Kreiger, New York (1974). MR 22:5903. [48] Ch.-W. Tsai, The orthogonality structure determines a C ∗-algebra with continuous trace, Oper. Matrices 5, no. 3, 529-540 (2011). [49] M. Wolff, Disjointness preserving operators in C∗-algebras, Arch. Math. 62, 248-253 (1994). [50] J.D.M. Wright and M. Youngson, On isometries of Jordan algebras, J. London Math. Soc. 17, 339-344 (1978). [51] K. Ylinen, Fourier transforms of noncommutative analogues of vector measures and bimea- sures with applications to stochastic processes, Ann. Acad. Sci. Fenn. Ser. A I Math. 1, no. 2, 355-385 (1975). E-mail address: [email protected] Departamento de An´alisis Matem´atico, Facultad de Ciencias, Universidad de Granada, 18071 Granada, Spain. E-mail address: [email protected] Departamento de An´alisis Matem´atico, Facultad de Ciencias, Universidad de Granada, 18071 Granada, Spain.
1410.3846
2
1410
2016-12-20T20:19:32
Group actions on graphs and $C^*$-correspondences
[ "math.OA" ]
If $G$ acts on a $C^*$-correspondence ${\mathcal H}$, then by the universal property $G$ acts on the Cuntz-Pimsner algebra ${\mathcal O}_{\mathcal H}$ and we study the crossed product ${\mathcal O}_{\mathcal H}\rtimes G$ and the fixed point algebra ${\mathcal O}_{\mathcal H}^G$. Using intertwiners, we define the Doplicher-Roberts algebra ${\mathcal O}_\rho$ of a representation $\rho$ of a compact group $G$ on ${\mathcal H}$ and prove that ${\mathcal O}_{\mathcal H}^G$ is isomorphic to ${\mathcal O}_\rho$. When the action of $G$ commutes with the gauge action on ${\mathcal O}_{{\mathcal H}}$, then $G$ acts also on the core algebras ${\mathcal O}_{\mathcal H}^{\mathbb T}$, where $\mathbb T$ denotes the unit circle. We give applications for the action of a group $G$ on the $C^*$-correspondence ${\mathcal H}_E$ associated to a directed graph $E$. If $G$ is finite and $E$ is discrete and locally finite, we prove that the crossed product $C^*(E)\rtimes G$ is isomorphic to the $C^*$-algebra of a graph of $C^*$-correspondences and stably isomorphic to a locally finite graph algebra. If $C^*(E)$ is simple and purely infinite and the action of $G$ is outer, then $C^*(E)^G$ and $C^*(E)\rtimes G$ are also simple and purely infinite with the same $K$-theory groups. We illustrate with several examples.
math.OA
math
GROUP ACTIONS ON GRAPHS AND C ∗-CORRESPONDENCES VALENTIN DEACONU Abstract. If G acts on a C ∗-correspondence H over the C ∗- algebra A (see Definition 2.4), then by the universal property G acts on the Cuntz-Pimsner algebra OH and we study the crossed product OH ⋊ G and the fixed point algebra OG H. Using intertwin- ers, we define the Doplicher-Roberts algebra Oρ of a representation ρ of a compact group G on H and prove that under certain con- ditions OG H is isomorphic to Oρ. The action of G commutes with the gauge action on OH, therefore G acts also on the core alge- bras OT H, where T denotes the unit circle. We give applications for the action of a group G on the C ∗-correspondence HE asso- ciated to a topological graph E. If G is finite and E is discrete and locally finite, we prove that the crossed product C ∗(E) ⋊ G is isomorphic to the C ∗-algebra of a graph of C ∗-correspondences and stably isomorphic to a locally finite graph algebra. If C ∗(E) is simple and purely infinite and the action of G is outer, then C ∗(E)G and C ∗(E) ⋊ G are also simple and purely infinite with the same K-theory groups. We illustrate with several examples. 1. introduction Suppose the group G acts on a directed (topological) graph E. This means that G acts on the vertex space E0 and on the edge space E1, preserving incidences. By duality, we get an action of G on the C ∗- algebra C0(E0) and on the space Cc(E1), which extends to the C0(E0)− C0(E0) C ∗-correspondence HE. In particular, there is a homomorphism ρ : G → LC(HE) into the set of invertible C-linear operators on HE, called a representation of G on HE. By the universal property, this determines an action of G on the Cuntz-Pimsner algebra OHE , also called the graph C ∗-algebra and denoted C ∗(E). For example, if G finite acts on the graph with one vertex and n loops, then we get an n- dimensional representation ρ : G → L(Cn) and an action on the Cuntz algebra On. It is known that the fixed point algebra OG n is isomorphic 1991 Mathematics Subject Classification. Primary 46L05. Key words and phrases. C ∗-algebra; C ∗-correspondence; Group action; Group representation; Doplicher-Roberts algebra; Graph algebra; Cuntz-Pimsner algebra. 1 2 VALENTIN DEACONU to the Doplicher-Roberts algebra Oρ (denoted by OG in [6]), which in turn is a full corner in a Cuntz-Krieger algebra (see [26]). In a more general setting, given a group G acting on a A − A C ∗- correspondence H, our goal is to study the fixed point algebra OG H and the crossed product OH ⋊ G. We define the Doplicher-Roberts algebra Oρ associated to ρ : G → LC(H) from intertwiners (ρm, ρn), where ρn = ρ⊗n is the tensor power representation of G on the balanced tensor product H⊗n. We prove that in certain cases Oρ is isomorphic to OG H and strongly Morita equivalent to OH ⋊ G. If G is finite and it acts on a discrete and locally finite graph E, we prove that C ∗(E) ⋊ G is isomorphic to the C ∗-algebra of a graph of (minimal) C ∗-correspondences, constructed using the orbits in E0 and E1 and the characters of the stabilizer groups. In the proof we use results about the crossed product of a C ∗-correspondence by a group G. As a consequence, C ∗(E) ⋊ G is strongly Morita equivalent to a graph algebra, so its K-theory can be computed in terms of the incidence matrix. Since the action of G commutes with the gauge action of T on C ∗(E), the group G also acts on the core AF-algebra C ∗(E)T and C ∗(E)T ⋊ G ∼= (C ∗(E) ⋊ G)T is an AF-algebra. We recover some examples of group actions on AF-algebras considered by Handelman and Rossmann, see [14]. The paper is organized as follows. In the first section we define group actions on topological graphs and on C ∗-correspondences, and we extend these actions to the associated Cuntz-Pimsner algebras. In the next section we define the Doplicher-Roberts algebra associated to a group action on a C ∗-correspondence. We continue with general results about crossed products of C ∗-correspondences and graphs of C ∗-correspondences. The following section contains the main result about finite group actions on discrete graphs. We conclude with several examples of group actions on graphs and K-theory computations for the crossed product and the fixed point algebra. Acknowledgements. The author would like to thank Alex Kumjian and Bruce Blackadar for helpful and illuminating discussions. 2. Group actions on graphs and graph algebras Let E = (E0, E1, r, s) be a topological graph (see [4, 19]). Recall that E0, E1 are locally compact Hausdorff spaces and r, s : E1 → E0 are continuous with s a local homeomorphism. Denote by H = HE its C ∗-correspondence over A = C0(E0), obtained by completing Cc(E1) with the inner product GROUP ACTIONS ON GRAPHS AND CROSSED PRODUCTS 3 hξ, ηi(v) = Xs(e)=v ξ(e)η(e), ξ, η ∈ Cc(E1) and multiplications (ξ · f )(e) = ξ(e)f (s(e)), (f · ξ)(e) = f (r(e))ξ(e). The C ∗-algebra of a graph E is defined as the Cuntz-Pimsner algebra OH of the C ∗-correspondence H = HE, see [19]. Definition 2.1. Let E, F be two topological graphs. A graph mor- phism ϕ : E → F is a pair of continuous maps ϕ = (ϕ0, ϕ1) where ϕi : Ei → F i, i = 0, 1 such that ϕ0 ◦ r = r ◦ ϕ1 and ϕ0 ◦ s = s ◦ ϕ1, i.e. the diagram E0 ϕ0y F 0 ϕ1y ϕ0y s←−−− E1 r−−−→ E0 s←−−− F 1 r−−−→ F 0 is commutative. An isomorphism of topological graphs is a graph mor- phism ϕ = (ϕ0, ϕ1) such that ϕi is a homeomorphism for i = 0, 1. It follows that ϕ−1 = ((ϕ0)−1, (ϕ1)−1) is also a graph morphism. We denote by Aut(E) the group of automorphisms of a topological graph E. Definition 2.2. A locally compact group G acts on E if there are continuous maps αi : G × Ei → Ei, write αi(g, x) = αi g(x) for i = 0, 1 or just g · x, such that g 7→ αg = (α0 g) is a group homomorphism from G into Aut(E). This means that G acts on the vertex space E0 and on the edge space E1 such that the actions are compatible with the range and source maps r, s. This action can be extended to finite paths e1 · · · en ∈ En by g · (e1 · · · en) = (g · e1) · · · (g · en) and similarly to the set of infinite paths E∞. g, α1 Remark 2.3. The group action on the graph E determines a represen- tation ρ : G → LC(H) by invertible C-linear operators on H and an action of G on C0(E0) by ∗-automorphisms such that (ρ(g)ξ)(e) = ξ(g−1·e) for ξ ∈ Cc(E1) A routine verification shows that these actions are compatible with the inner product and the bimodule structure. and (g·a)(v) = a(g−1·v) for a ∈ C0(E0). Definition 2.4. We say that a locally compact group G acts on the C ∗-correspondence H over the C ∗-algebra A if G acts on H via a map ρ : G → LC(H) such that ρ(g) is a C-linear isomorphism and for all ξ ∈ 4 VALENTIN DEACONU H the map g 7→ ρ(g)ξ is norm continuous, G acts by ∗-automorphisms on A such that for all a ∈ A the map g 7→ g · a is norm continuous, and the following compatibility relations are satisfied hρ(g)ξ, ρ(g)ηi = g · hξ, ηi, ρ(g)(ξa) = (ρ(g)ξ)(g · a), ρ(g)(aξ) = (g · a)(ρ(g)ξ). The map ρ is called a representation of G on H. Remark 2.5. In particular, a group action on the graph E determines as above a group action on the C ∗-correspondence HE. Notice though that a group action on a C ∗-correspondence associated to a directed graph is not necessarily determined by an action on the graph, see Example 6.6. Theorem 2.6. An action of G on the C ∗-correspondence H determines in a natural way an action on KA(H), the C ∗-algebra generated by the finite rank operators, and an action on the Cuntz-Pimsner algebra OH. The action of G commutes with the gauge action, therefore we get an action of G on the core algebra OT H, the fixed point algebra under the gauge action. In particular, an action on a topological graph E determines an action on the graph algebra C ∗(E) by g · Se = Sg·e, where Se is a generator of C ∗(E) for e ∈ E1, and an action on the core algebra C ∗(E)T. Proof. Recall that KA(H) is generated by operators θξ,η where θξ,η(ζ) = ξhη, ζi and we define g · θξ,η = θρ(g)ξ,ρ(g)η. The first part follows from the universal property of OH. Recall that the gauge action γ on OH is defined on generators by γ(z)a = a, γ(z)ξ = zξ for z ∈ T and is extended to OH using the universal property. Since ρ(g) : H → H is C-linear, we have ρ(g)(zξ) = zρ(g)ξ, so we get an action of G on the core algebra OT H. (cid:3) Recall that a discrete graph (a topological graph where E0, E1 are at most countable) is row finite if each vertex receives finitely many edges, and is locally finite if in addition each vertex emits finitely many edges. For free actions on discrete graphs we have the following result: Theorem 2.7. (Kumjian and Pask, [24]) If G, E are discrete, the ac- tion of G on E is free and E is locally finite, then C ∗(E)G ∼= C ∗(E/G) and C ∗(E) ⋊ G ∼= C ∗(E/G) ⊗ K(ℓ2(G)), where E/G is the quotient graph. GROUP ACTIONS ON GRAPHS AND CROSSED PRODUCTS 5 This result is inspired from a theorem of Green about group actions on locally compact spaces, see [12]. A similar result was proved for free and proper actions of locally compact groups on topological graphs in [4], namely that C ∗(E) ⋊r G is strongly Morita equivalent to C ∗(E/G). In the same paper [24], Kumjian and Pask showed that if G is abelian and c : E1 → G is a cocycle, then this induces an action of G on C ∗(E) such that C ∗(E) ⋊ G is isomorphic to C ∗(E(c)), where E(c) is the skew product graph ( G × E0, G × E1, r, s) with r(χ, e) = (χc(e), r(e)), s(χ, e) = (χ, s(e)) for χ ∈ G. By diagonalization, the action of G on C ∗(E) is equivalent to the action α given by αg(Se) = hc(e), giSe, where Se are the generators of C ∗(E). Remark 2.8. If G abelian acts on the On-graph with E1 = {e1, e2, ..., en} and E0 = {v}, a cocycle c : E1 → G determines a representation ρ of G on H = span{ξ1, ξ2, ..., ξn}, where ρ(g)ξi = hc(ei), giξi. Conversely, an n-dimensional representation of the abelian group G determines a cocycle on the On-graph with values in G. Remark 2.9. Actions of Zl on k-graphs were studied by Farthing, Pask and Sims in [11]. In particular, K-theory computations were done for actions of Z on a row finite 1-graph with no sources such that the orbit of each vertex is finite and either K0(C ∗(E)) or K1(C ∗(E)) is trivial. 3. Doplicher-Roberts algebras The Doplicher-Roberts algebras (denoted by OG in [6]) were in- troduced to construct a new duality theory for compact Lie groups G ⊆ U(n) which strengthens the Tannaka-Krein duality. Let TG de- note the representation category whose objects are tensor powers of the n-dimensional representation ρ of G defined by the inclusion G ⊆ U(n) and whose arrows are the intertwiners. The C ∗-algebra OG is iden- tified in [6] with the fixed point algebra OG n , where On is the Cuntz algebra. If σG denotes the restriction to OG of the canonical endomor- phism of the Cuntz algebra, then TG can be reconstructed from the pair (OG, σG). Subsequently, Doplicher-Roberts algebras were associated to any object ρ in a strict tensor C ∗-category, see [7], [5]. Suppose that the group G acts on the C ∗-correspondence H over A via the representation ρ : G → LC(H). Inspired from [6], we consider the tensor power representation ρn : G → LC(H⊗n), where H⊗n is the balanced tensor product of n copies of H over A, and we define the set 6 VALENTIN DEACONU (ρm, ρn) of intertwining operators by (ρm, ρn) = {T ∈ LA(H⊗n, H⊗m) T ρn = ρmT }. By definition H⊗0 = A and ρ0 : G → LC(A) is the trivial representation ρ0(g)(a) = a. We identify (ρm, ρn) with a subset of (ρm+r, ρn+r) via T 7→ T ⊗ Ir, where Ir : H⊗r → H⊗r is the identity map. After this (ρm, ρn) has identification, it follows that the linear span 0Oρ of [m,n≥0 a natural multiplication given by composition: T ∈ (ρp, ρq), then the product ST is if S ∈ (ρm, ρn) and (S ⊗ Ip−n) ◦ T ∈ (ρm+p−n, ρq) if p ≥ n, or S ◦ (T ⊗ In−p) ∈ (ρm, ρq+n−p) if p < n. The adjoint of T ∈ (ρm, ρn) is T ∗ ∈ (ρn, ρm). We assume that kT k = sup{kπ(T )k : π is a ∗−representation of 0Oρ on a Hilbert space} is finite. Definition 3.1. Under this assumption, we define the Doplicher-Roberts algebra Oρ associated to the representation ρ : G → LC(H) as the C ∗- closure of the normed ∗-algebra 0Oρ with the above operations. Remark 3.2. The ∗-algebra 0Oρ has a natural Z-grading and tensoring with I on the left induces a ∗-endomorphism σ. Theorem 3.3. Let H be a full finite projective C ∗-correspondence over A (i.e. H is a direct summand of Ak for some k and the inner products generate A) and assume that the left multiplication A → L(H) is in- jective. If G is a compact group acting on H via ρ : G → LC(H), then the Doplicher-Roberts algebra Oρ is well defined and it is isomorphic to the fixed point algebra OG H. Proof. Since H is finite projective and the left multiplication is injec- tive, it is known that LA(H) ∼= KA(H) and that OH is isomorphic to KA(H⊗m, H⊗n) after we the C ∗-algebra generated by the span of [m,n≥0 identify T with T ⊗ I (see Proposition 2.5 in [17]). Note that G acts on KA(H⊗n, H⊗m) by (g · T )(ξ) = ρm(g)T (ρn(g−1)ξ) and the fixed point algebra is (ρm, ρn). It follows that 0Oρ ⊆ OH and that Oρ is isomorphic to OG H. (cid:3) GROUP ACTIONS ON GRAPHS AND CROSSED PRODUCTS 7 Corollary 3.4. Let E be a topological graph such that HE is full fi- nite projective and the left multiplication of C0(E0) is injective. If G is a compact group acting on E, and ρ : G → LC(HE) denotes the representation, then Oρ ∼= C ∗(E)G. Moreover, if C ∗(E) is simple and purely infinite, G is finite and the action on C ∗(E) is (pointwise) outer, then Oρ and C ∗(E) ⋊ G are simple purely infinite and have the same K-theory, therefore are stably isomorphic. Proof. The first part follows directly from the above theorem. The second part is a consequence of a result of A. Kishimoto and A. Kumjian (see Lemma 10 in [23] and Theorem 3.1 in [21]) : If A is simple and purely infinite, G is discrete and α : G → Aut(A) is an action such that αg is outer for all g ∈ G\{e}, then A⋊αr G is simple and purely infinite. ∼= C ∗(E)G is a full The stable isomorphism follows from the fact that Oρ corner in C ∗(E) ⋊ G and from classification results of simple separable purely infinite algebras satisfying UCT. (cid:3) Remark 3.5. The natural inclusions C ∗(E)G ⊆ C ∗(E) ⊆ C ∗(E) ⋊ G determine group homomorphisms K0(C ∗(E)G) → K0(C ∗(E)) → K0(C ∗(E) ⋊ G). Assuming C ∗(E) is unital, these homomorphisms give information on the class of the identity in K0(C ∗(E)G) and K0(C ∗(E) ⋊ G). Example 3.6. If γ is the gauge action of T on a C ∗-correspondence H over A, then Oγ ∼= OT H. Example 3.7. If the group G acts on a C ∗-algebra A and π : G → U(n) is a faithful unitary representation, then H = Cn ⊗ A has a natural structure of C ∗-correspondence over A such that G acts on H by ρ(g)(x ⊗ a) = π(g)x ⊗ g · a. It is easy to check that in this case Oρ is well defined and it is isomorphic to Oπ ⊗ AG, where Oπ is the (old) Doplicher-Roberts algebra associated to the representation π. Remark 3.8. Let the group G act on the C ∗-correspondence H over A. We have AG ⊆ (ρ, ρ), where AG denotes the fixed point algebra. Indeed, if a ∈ AG, then a(ρ(g)ξ) = (g · a)(ρ(g)ξ) = ρ(g)(aξ). Example 3.9. Consider a finite group G acting on the graph En with one vertex and n ≥ 2 edges. We denote by ρ the corresponding repre- sentation on H = HE = Cn. Let G denote the set of equivalence classes of irreducible unitary representations, and construct as in [26] a graph with the incidence 8 VALENTIN DEACONU matrix B = B(ρ), where B(v, w) is the multiplicity of w in v ⊗ ρ for v, w ∈ G. It is shown in [26] that Oρ is a full corner in the Cuntz- Krieger algebra OB. For G = Sn the symmetric group acting by permuting the edges of En, we get an outer action on the Cuntz algebra On such that On ⋊ Sn is simple and purely infinite, stably isomorphic to Oρ n . We also ∼= Mn∞ such that Mn∞ ⋊ Sn get an action of Sn on the core algebra OT n is AF. ∼= OSn For n = 3, using the character table of S3, it was calculated in [26] that which gives B =  1 0 1 0 1 1 1 1 2   , K0(O3 ⋊ S3) = K0(Oρ) = K0(OB) ∼= Z, K1(O3 ⋊ S3) = K1(Oρ) = K1(OB) ∼= Z, The inclusions OS3 3 K0(M3∞ ⋊ S3) ∼= lim−→(Z3, B). ֒→ O3 ֒→ O3 ⋊ S3 determine the K0-theory maps Z → Z2 → Z. In particular the action of S3 on O3 does not have the Rokhlin property, since the map Z → Z2 is not injective (see Theorem 3.13 in [16]) and OS3 3 , O3 ⋊ S3 are not isomorphic since the classes of the identity in their K0-groups do not coincide. Remark 3.10. If R(S3) ∼= K0(S3) is the representation ring of S3, then the matrix B above is determined by the map R(S3) → R(S3) given by multiplication with the character of the representation ρ (see [13, 14]). Remark 3.11. An action of a group G on a row-finite (discrete) graph E with no sources induces an action of G on the associated graph groupoid G = GE such that C ∗(E) ⋊ G ∼= C ∗(G ⋊ G), where G ⋊ G is the semidirect product groupoid with multiplication inverse operation (γ, g)(g−1 · γ′, g′) = (γγ′, gg′), (γ, g)−1 = (g−1 · γ−1, g−1) and range and source maps r(γ, g) = (r(γ), e), s(γ, g) = (g−1 · s(γ), e). The unit space of G ⋊ G is identified with G0. GROUP ACTIONS ON GRAPHS AND CROSSED PRODUCTS 9 In particular, for the Sn action above we get an action of Sn on the Cuntz groupoid Gn = {(x, p − q, y) ∈ X × Z × X : σpx = σqy}, where σ : X → X is the shift on the unit space X = {1, ..., n}N such that On ⋊ Sn ∼= C ∗(Gn ⋊ Sn). Example 3.12. Given a finite-dimensional unitary representation ρ of a compact group G, Kumjian, Pask, Raeburn and Renault (see [25]) realize the Doplicher-Roberts algebra Oρ as a corner in a graph C ∗- algebra and as a groupoid algebra. The graph has vertices G, the set of equivalence classes of irreducible representations, and the groupoid is the reduction of the graph groupoid to the set of infinite paths starting at the trivial representation. It turns out that if ρ takes values in SU(n) and is faithful, then the graph is irreducible and locally finite, in particular Oρ is simple. Moreover, if n ≥ 2, G is an infinite compact Lie group and βρ denotes the endomorphism of the representation ring R(G) given by tensoring with ρ, then K0(Oρ) ∼= R(G)/im(1 − βρ) and K1(Oρ) = 0. This last result appeared also in A. Wassermann's thesis [34]. 4. Group actions on C ∗-correspondences and crossed products We will need to allow B -- A C ∗-correspondences where A and B are not necessarily the same C ∗-algebras, so we extend our notion of group action: Definition 4.1. Given C ∗-algebras A, B and a B -- A C ∗-correspondence H, an action of a locally compact group G on H is determined by a ho- momorphism ρ : G → LC(H) such that ρ(g) is a C-linear isomorphism and g 7→ ρ(g)ξ is continuous and continuous actions of G on A and B by ∗-automorphisms with compatibility relations hρ(g)ξ, ρ(g)ηi = g · hξ, ηi, ρ(g)(ξa) = (ρ(g)ξ)(g · a), ρ(g)(bξ) = (g · b)(ρ(g)ξ), where ξ ∈ H, a ∈ A, b ∈ B. As we mentioned before, an action of G on H determines an action of G on K(H) given by g · θξ,η = θρ(g)ξ,ρ(g)η, where θξ,η(ζ) = ξhη, ζi. Recall that if A = B, an action of G on H determines an action on the Cuntz-Pimsner algebra OH (called quasi-free) and, since the action commutes with the gauge action, an action on the core algebra OT H. 10 VALENTIN DEACONU Definition 4.2. Suppose the group G acts on the B -- A C ∗-correspondence H. The crossed product C ∗-correspondence H⋊G is defined as H⋊G = H ⊗ϕ (A ⋊ G), where ϕ : A → L(A ⋊ G) is the embedding of A in the multiplier algebra of A ⋊ G, regarded as a Hilbert module over itself. Remark 4.3. The crossed product H ⋊ G becomes a B ⋊ G -- A ⋊ G C ∗- correspondence after the completion of Cc(G, H) using the operations hξ, ηi(t) =ZG (ξ · f )(t) =ZG (h · ξ)(t) =ZG s−1 · hξ(s), η(st)ids, ξ(s)(s · (f (s−1t)))ds, h(s) · (s · ξ(s−1t))ds, where ξ, η ∈ Cc(G, H), f ∈ Cc(G, A), h ∈ Cc(G, B). Note that the right and left multiplications are given by convolution, and the inner product formula could be also expressed as hξ ⊗ f, η ⊗ f ′i = f ∗hξ, ηif ′, where this time ξ, η ∈ H, f, f ′ ∈ Cc(G, A) and f ∗(t) = t · f (t−1)∗. Theorem 4.4. (G. Hao, C.-K. Ng, [15]) Let H be a C ∗-correspondence over A and let the amenable locally compact group G act on H. Then OH⋊G ∼= OH ⋊ G. Corollary 4.5. For G amenable acting on a C ∗-correspondence H we have OT H⋊G ∼= OT H ⋊ G. Corollary 4.6. Let G be a compact group acting on a topological graph E with C ∗-correspondence HE. Then C ∗(E) ⋊ G ∼= OHE ⋊G. Example 4.7. Let G be a compact group and let E be a Hermitian vector bundle over a locally compact space X such that G acts on both E and X in a compatible way (see [1], section 1.6). Such a vector bundle is called a G-vector bundle, generalizing both ordinary vector bundles (when G is trivial) and G-modules (when X reduces to a point). The set of sections Γ(E) becomes in the usual way a C ∗-correspondence over C0(X), and the group G acts on Γ(E). In particular, G acts on its Cuntz-Pimsner algebra, which is a continuous field of Cuntz algebras, see [33]. GROUP ACTIONS ON GRAPHS AND CROSSED PRODUCTS 11 Example 4.8. Let G be compact and let ρ : G → U(n) be a unitary ∼= On where H = representation. This determines an action of G on OH ∼= Mn∞(see Cn and a product type action Mn ∞O1 ∞O1 Adρ on OT H ∼= [13]). We obtain the isomorphisms OH⋊G ∼= On ⋊ G, OT H⋊G ∼= Mn∞ ⋊ G. Remark 4.9. Let G be a compact group and let E be a finite graph with C ∗-correspondence HE. If G acts on HE, using the universal property we obtain an action of G on C ∗(E). Since the action commutes with the gauge action, we get an action on the core algebra C ∗(E)T ∼= lim−→ An, where An have dimension mn. For a locally representable action as in [13, 14], K0(C ∗(E)T ⋊ G) is the inductive limit of K0(An ⋊ G) ∼= K0(G)mn, where the inclusion maps are determined by matrices with entries in the representation ring K0(G) ∼= R(G). Remark 4.10. Note that some actions which permute vertices in a graph with more than one vertex may not induce locally representable actions on the core algebra (see Example 6.3). 5. Graphs of C ∗-correspondences and applications to finite groups actions on discrete graphs Given a discrete graph E = (E0, E1, r, s), associate to each vertex v ∈ E0 a C ∗-algebra Av and to each edge e ∈ E1 a nondegenerate Ar(e) -- As(e) C ∗-correspondence He. This way we obtain an E-system of C ∗-correspondences or a graph of C ∗-correspondences. The C ∗- algebra associated to this graph of C ∗-correspondences is OH, where H =Le∈E 1 He becomes a C ∗-correspondence over A =Lv∈E 0 Av in a natural way. For more information, see [3], where we discuss systems of C ∗-correspondences over k-graphs Λ and we construct a Fell bundle over the path groupoid GΛ such that its reduced cross-sectional algebra is isomorphic to the C ∗-algebra of the Λ-system. Unlike in [3], here we allow graphs with sources and C ∗-correspondences which are not full. Example 5.1. Given a discrete graph E, associate to each vertex the C ∗-algebra C and to each edge the C ∗-correspondence C. This is a graph of C ∗-correspondences with associated C ∗-algebra isomorphic to C ∗(E). Example 5.2. Let E have one vertex v and one loop e, and let Av = C, He = Cn. Then the C ∗-algebra of this graph of C ∗-correspondences is On. If Av = A is any C ∗-algebra and He = H is a C ∗-correspondence over A, then we get OH. 12 VALENTIN DEACONU Example 5.3. Consider a C ∗-correspondence H over a unital C ∗-algebra A such that A decomposes into a direct sum A1 ⊕ A2 ⊕ · · · ⊕ An. If pj piHpj is the identity of Aj, then H decomposes into a direct sum Mi,j and we can construct a graph of C ∗-correspondences with n vertices {v1, ..., vn}, by assigning the C ∗-algebra Ai at vi and the Ai -- Aj C ∗- correspondence piHpj 6= 0 at an edge joining vj with vi. If some of these correspondences are trivial, there is no edge between the corresponding vertices. Recall that if a finite group G acts on a finite or countable set X, then C0(X) ⋊ G decomposes as a direct sum of crossed products C(Gx) ⋊ G over the orbit space X/G. Since the action on each orbit Gx is transitive, this orbit can be identified with the homogeneous space G/Gx, where Gx is the stabilizer group and G acts on G/Gx by left translation. Moreover, C(G/Gx) ⋊ G ∼= MGx ⊗ C ∗(Gx), which is isomorphic to a finite direct sum of matrix algebras. Corollary 5.4. If a finite group G acts on a discrete graph E, then C0(E0) ⋊ G ∼= ME 0/G M(Gv), where M(Gv) is a finite direct sum of matrix algebras. In particular, C0(E0) ⋊ G is strongly Morita equivalent (SME) to a direct sum of finite dimensional abelian C ∗-algebras. To describe the crossed product C ∗(E) ⋊ G, we first consider the case when C0(E0) ⋊ G is abelian. Proposition 5.5. Suppose C0(E0)⋊G = C0(V ) with V finite or count- able, and denote by {pt}t∈V the minimal projections in C0(V ). The isomorphism classes of separable nondegenerate C ∗-correspondences H over C0(V ) with ∗-homomorphism ϕ : C0(V ) → L(H) correspond to matrices (ast)s,t∈V where ast are nonnegative integer entries or ast = ∞. More precisely, ast = dim ϕ(ps)Hpt. Proof. See Theorem 1.1 in [18]. (cid:3) Corollary 5.6. If C0(E0) ⋊ G = C0(V ) is abelian, then C ∗(E) ⋊ G is the graph algebra with incidence matrix (ast)s,t∈V . It is also the C ∗- algebra of a graph of C ∗-correspondences where the vertex algebras are C (one for each vertex t ∈ V ) and the C ∗-correspondences are Hilbert spaces of dimension ast. GROUP ACTIONS ON GRAPHS AND CROSSED PRODUCTS 13 Proposition 5.7. Suppose A and B are SME C ∗-algebras with A -- B imprimitivity bimodule X . If H is a C ∗-correspondence over A, then H′ = X ∗ ⊗A H ⊗A X is a C ∗-correspondence over B such that OH and OH′ are SME. Proof. Let R = H ⊗A X and let S = X ∗. Then R ⊗B S ∼= H, S ⊗A R ∼= H′, so by a theorem in [27] (see also [28]), we get that OH and OH′ are SME. (cid:3) Corollary 5.8. Given a discrete locally finite graph E = (E0, E1, r, s) and a finite group G acting on E, the crossed product C ∗(E)⋊G is SME to a locally finite graph C ∗-algebra, where the number of vertices is the cardinality of the spectrum of C0(E0) ⋊ G. In particular, the K-theory of C ∗(E) ⋊ G and of C ∗(E)T ⋊ G can be computed if we determine the incidence matrix of the graph. Proof. We apply the Proposition with A = C0(E0)⋊G and B = C0(V ). (cid:3) Theorem 5.9. Given a discrete locally finite graph E = (E0, E1, r, s) and a finite group G acting on E, the crossed product C ∗(E) ⋊G is iso- morphic to the C ∗-algebra of a graph of (minimal) C ∗-correspondences, where at each vertex v we associate a matrix algebra Mn(v) and at each edge joining w and v we associate Mn(v),n(w), the set of rectangular matrices with n(v) rows and n(w) columns. Proof. Since the group is finite, the orbits in E0 and E1 are finite. We decompose the C ∗-correspondence HE ⋊ G over the C ∗-algebra C0(E0) ⋊ G. This decomposition is obtained in two stages, from the orbits in E0 and from the characters of the stabilizer groups. For the first stage, we consider the quotient graph E/G and at each vertex [v] ∈ E0/G we associate the C ∗-algebra C(Gv) ⋊ G, where Gv is the orbit of v in E0 and at each edge [e] ∈ E1/G we associate the C(Gr(e)) ⋊ G -- C(Gs(e)) ⋊ G C ∗-correspondence C(Ge) ⋊ G of the orbit Ge in E1. For the second stage, we decompose each C(Gv) ⋊ G ∼= MGv ⊗ C ∗(Gv) into simple components. mMi=1 Let C0(E0) ⋊ G ∼= Mn(i), where m ∈ N ∪ {∞} and Mn(i) denotes the set of n(i) × n(i) matrix algebras. Consider now the graph with m vertices and at each vertex vi we assign the C ∗-algebra Mn(i). If pi is the unit in Mn(i), whenever pi(HE ⋊ G)pj 6= 0, we decompose this as a direct sum of minimal Mn(i) -- Mn(j) C ∗-correspondences. A minimal C ∗- correspondence is of the form Mn(i),n(j), the set of rectangular matrices with n(i) rows and n(j) columns, with the obvious bimodule structure 14 VALENTIN DEACONU and inner product. Of course, Mn,n = Mn and Mn,1 = Cn. This decomposition determines the number of edges between vj and vi. By construction, it follows that C ∗(E) ⋊ G is isomorphic to the C ∗-algebra of this graph of C ∗-correspondences. (cid:3) Remark 5.10. Given a compact group G, denote by R(G) its repre- If G acts on a C ∗-algebra A, recall that K0(A ⋊ G) sentation ring. has a structure of R(G)-module. Indeed, given M a finite dimensional G-module with character χ and N a finitely generated projective A⋊G- module, then M ⊗N has a structure of A⋊G-module and we can define the product [N]·χ as [M ⊗N]. In particular, given a finite group G act- ing on a finite graph E, the groups K0(C ∗(E)⋊G) and K0(C ∗(E)T ⋊G) have a structure of R(G)-modules. crete groups G on the set of finite words X ∗ = S∞ Remark 5.11. Nekrashevych (see [29]) studied faithful actions of dis- k=0 X k over a finite alphabet X, which are self-similar in the sense that for all g ∈ G and x ∈ X there exist unique y ∈ X and h ∈ G such that g · (xw) = y(h · w) where w ∈ X ∗. A self-similar action determines an action of G on the rooted tree TX with root ∅ and edges from w ∈ X ∗ to wx for x ∈ X. He constructed a C ∗-correspondence M = Lx∈X C ∗(G) over the C ∗- algebra C ∗(G), where the left action is the integrated form of a unitary representation of G in L(M), defined using the self-similarity condition. Since in our notion of representation ρ : G → LC(H) the operator ρ(g) is not A-linear, the C ∗-correspondence M is not the C ∗-correspondence associated with the (infinite) tree TX, and the Cuntz-Pimsner algebra OM , denoted also O(G,X) in [29], is not isomorphic to the crossed prod- uct C ∗(TX ) ⋊G. Observe though that OM contains a copy of the Cuntz algebra On, the tree TX is the universal covering of the graph En with one vertex and n = X edges, and M = Cn ⊗ C ∗(G). Exel and Pardo in [10], inspired from the Nekrashevych construction, associate a C ∗-algebra OG,E from a countable discrete group G acting on a finite graph E and a one-cocycle ϕ : G×E1 → G which determines an action of G on the space of finite paths E∗ such that g · (αβ) = (g · α)(ϕ(g, α) · β). This C ∗-algebra is defined as the Cuntz-Pimsner algebra of a C ∗-correspondence M over C(E0) ⋊ G and contains a copy of the graph algebra C ∗(E). In particular for G = Z acting on a graph E with N × N incidence matrix A by fixing the vertices and permuting the edges in a way determined by another N × N integer matrix B (which also determines the cocycle), they prove that OG,E is isomorphic to Katsura's algebra OA,B, see [20], used to model all Kirchberg algebras. Note that in this case C(E0) ⋊ G is isomorphic to the direct sum of N copies of C(T) ∼= C ∗(Z). GROUP ACTIONS ON GRAPHS AND CROSSED PRODUCTS 15 The relationship between the C ∗-algebras studied by these authors and our crossed products C ∗(E) ⋊ G remains to be explored. 6. Examples Example 6.1. Let G = Zn act freely on its cyclic Cayley graph E with n vertices and n edges. We get a representation ρ of Zn on H = Cn and an action of Zn on A = Cn which permutes cyclically the basis. In this case LA(H) = {T ∈ Mn : T (ξa) = T (ξ)a} ∼= Cn (diagonal matrices). ∼= C(T), since the quotient graph E/G Moreover, (ρ, ρ) ∼= CI and Oρ has one vertex and one loop. The crossed product H ⋊ Zn ∼= Mn becomes a C ∗-correspondence over A ⋊ G ∼= Mn, and C ∗(E) ⋊ G ∼= Mn ⊗ C(T). The graph of C ∗-correspondences has one vertex and one loop with Mn attached to each. Example 6.2. Let E be the graph with three vertices v1, v2, v3 and edges connecting each vi with vj for i 6= j. v1 v2 v3 The permutation group G = S3 acts (non-freely) on E by permuting the vertices. The action on edges is uniquely determined. There is a single orbit in E0 and E1, and the stabilizer group for each vertex is ∼= C6, the representation ρ of S3 is 6-dimensional and Z2. Since HE ∼= C ∗(E)G ∼= O2. We have C ∗(E) ∼= M3(O2) ∼= O2, so we get Oρ ∼= M3(C ∗(Z2)) ∼= an action of S3 on O2. Here A = C3 and A ⋊ S3 M3 ⊕ M3. The C ∗-correspondence HE ⋊ G over M3 ⊕ M3 decomposes into M3 ⊕ M3 ⊕ M3 ⊕ M3. The graph of C ∗-correspondences is M3 M3 M3 M3 M3 M3 and C ∗(E) ⋊ G is stably isomorphic to O2. 16 VALENTIN DEACONU Example 6.3. Consider the graph E with three vertices v, v1, v2 and four edges e1, e2, f1, f2 like in the figure. v e1 e2 f1 f2 v2 v1 The group G = Z2 acts on E by fixing v and interchanging v1 and v2. This action takes the edge e1 into e2 and the edge f1 into f2. Since the set {v} is hereditary and saturated, C ∗(E) has an ideal isomorphic to K, the C ∗-algebra of compact operators such that C ∗(E)/K ∼= M2(C(T)). Since E has sources, the K-theory of C ∗(E) is computed using Theorem 3.2 in [32] and K0(C ∗(E)) = Z ⊕ Z2, K1(C ∗(E)) = 0. We have A = C3, H = C4, A ⋊ Z2 ∼= C8 which decomposes as C2 ⊕ C2 ⊕ M2. The crossed product C ∗(E) ⋊ Z2 is the C ∗-algebra of the following graph of C ∗-correspondences ∼= C ⊕ C ⊕ M2 and H ⋊ Z2 C C C2 C2 M2 M2 If E′ denotes the subjacent graph, there is an extension 0 → K → C ∗(E′) → T → 0, where T is the Toeplitz algebra. We have K0(C ∗(E)⋊Z2) = K0(C ∗(E′)) = Z2, K1(C ∗(E)⋊Z2) = K1(C ∗(E′)) = 0. GROUP ACTIONS ON GRAPHS AND CROSSED PRODUCTS 17 Example 6.4. Let S3 act on the graph E with one vertex and three loops by permuting the loops. We get a 3-dimensional representation ρ of S3 and a non-free action on O3. We already know (see [26]) that Oρ is a full corner in a graph algebra with incidence matrix ∼= OS3 3   1 0 1 0 1 1 1 1 2   . ∼= (O3 ⋊ Z3) ⋊ Z2. We will describe Since S3 the graph of C ∗-correspondences using this iterated crossed product. ∼= Z3 ⋊ Z2, we have O3 ⋊S3 It follows that O3 ⋊ Z3 is isomorphic to C ∗(E(c)), where c : E1 → cZ3 is a cocycle and E(c) is the graph with three vertices v1, v2, v3 and nine edges connecting each vi with vj. v1 v2 v3 The group Z2 acts on E(c) by fixing v1 and interchanging v2 with v3. The action on edges is uniquely determined. If π is the corresponding ∼= C ∗(E(c))Z2. The representation of Z2 on C9, it follows that Oπ quotient graph E(c)/Z2 has two vertices u1, u2 corresponding to the two orbits in E(c)0 and five edges corresponding to the orbits in E(c)1: one loop at u1, two loops at u2, one edge from u1 to u2 and one edge 18 VALENTIN DEACONU from u2 to u1. u1 u2 ∼= C ⊕ C ⊕ M2 We have C ∗(E(c)) ∼= O3, so we get a non-free action of Z2 on O3. ∼= C ∗(S3). Moreover, O3 ⋊ Here A = C3 and A ⋊ Z2 ∼= C ∗(E(c)) ⋊ Z2 is the C ∗-algebra of the following graph of C ∗- S3 correspondences. For the vertex u1 the C ∗-algebra is C ∗(Z2) ∼= C2 ∼= M2. For the loop at u1 the and for u2 the C ∗-algebra is C2 ⋊ Z2 C ∗-correspondence is C2. For each of the two loops at u2 we have a copy of M2. Finally, for each of the remaining two edges in E(c)/Z2 we have C2 ⊕ C2. Since the vertex u1 splits in two, the C ∗-correspondence C9 ⋊ Z2 over C2 ⊕ M2 decomposes further as C ⊕ C ⊕ M2 ⊕ M2 ⊕ C2 ⊕ C2 ⊕ C2 ⊕ C2 and we get the following graph of minimal C ∗- correspondences: there are three vertices with C ∗-algebras C, C and M2 respectively. The incidence matrix of the graph is   1 0 1 0 1 1 1 1 2   GROUP ACTIONS ON GRAPHS AND CROSSED PRODUCTS 19 and the C ∗-correspondences are as in the figure C C C C C2 C2 C2 C2 M2 M2 M2 The group S3 also acts on the core M3∞ of O3. Since O3 ⋊S3 is strongly Morita equivalent to a graph algebra, it follows that its core is strongly Morita equivalent to M3∞ ⋊ S3. ∼= Z4 ⋊ Z2 act in the natural Example 6.5. Let the dihedral group D4 way on the "cross" graph E with five vertices and four edges by fixing the central vertex. ∼= We get an action of D4 on C ∗(E) ∼= M5. It is known that C ∗(E)⋊D4 M5 ⊗ C ∗(D4). The vertex space has two orbits and the edge space one ∼= C4 ⊕ M2 ⊕ M4 ⊕ M4 orbit. We have C5 ⋊ D4 and C4 ⋊ D4 decomposes as C4 ⊕ C4 ⊕ C4 ⊕ C4 ⊕ M2,4 ⊕ M2,4, obtaining ∼= C ∗(D4) ⊕ C4 ⋊ D4 20 VALENTIN DEACONU the following graph of C ∗-correspondences: C C4 M4 C4 C M2,4 M2,4 M2 C C4 M4 C4 C B =  0 1 0 1 1 1 0 1 0   . Example 6.6. Let the symmetric group S3 = hτ, σi act on O2 by τ (s1) = s2, τ (s2) = s1, σ(s1) = ωs1, σ(s2) = ω2s2, where τ = (12), σ = (123) and ω2 + ω + 1 = 0. This action corresponds to the two-dimensional irreducible representation ρ of S3, and it commutes with the gauge action. Then OS3 2 and O2 ⋊ S3 have the K-theory of O3, since (see [26]) Oρ is a full corner in the algebra of the graph with three vertices and incidence matrix In particular, we get an action of S3 on the CAR algebra M2∞ and K0(M2∞ ⋊ S3) ∼= lim−→(Z3, B). References [1] M.F. Atiyah, K-Theory, Benjamin, New York, 1967. [2] J. Cuntz, D.E. Evans, Some remarks on the C ∗-algebras associated with certain topological Markov chains, Math. Scand. 48 (1981), no. 2, 235 -- 240. [3] V. Deaconu, A. Kumjian, D. Pask, A. Sims, Graphs of C ∗-correspondences and Fell bundles, Indiana Univ. Math. J. 59 (2010), no. 5, 1687 -- 1735. [4] V. Deaconu, A. Kumjian, J. Quigg, Group actions on topological graphs, Ergod. Th. Dyn. Sys. 32 (2012), no. 5, 1527 -- 1566. [5] S. Doplicher, C. Pinzari, R. Zuccante, The C ∗-algebra of a Hilbert bimodule, Boll. Unione Mat. Ital. Sez. B Artic. Ric. Mat. (8) 1 (1998), no. 2, 263 -- 281. [6] S. Doplicher, J.E. Roberts, Duals of compact Lie groups realized in the Cuntz Algebras and their actions on C*-algebras, J. of Funct. Anal. 74(1987) 96 -- 120. [7] S. Doplicher, J.E. Roberts, A new duality theory for compact groups, Invent. Math. 98 (1989) 157 -- 218. [8] M. Enomoto, M. Fujii, H. Takehana, Y. Watatani, Automorphisms on Cuntz algebras II, Math. Japon. 24 (1979/80), no. 4, 463 -- 468. [9] M. Enomoto, H. Takehana, Y. Watatani, Automorphisms on Cuntz algebras, Math. Japon. 24 (1979/80), no. 2, 231 -- 234. GROUP ACTIONS ON GRAPHS AND CROSSED PRODUCTS 21 [10] R. Exel, E. Pardo, Self-Similar graphs: a unified treatment of Katsura and Nekrashevych C ∗-algebras, arxiv preprint 1409.1107 v2. [11] C. Farthing, D. Pask, A. Sims, Crossed products of k-graphs C ∗-algebras by Zl, Houston J. Math 35 (2009) no. 3, 903 -- 933. [12] P. Green, C ∗-algebras of transformation groups with smooth orbit space, Pa- cific J. Math 72 (1977), no. 1, 71 -- 97. [13] D. Handelman, W. Rossmann, Product type actions of finite and compact groups, Indiana Univ. Math. J. 33 (1984), no.4 , 480 -- 509. [14] D. Handelman, W. Rossmann, Actions of compact groups on AF C*- algebras, Illinois J. Math. 29 (1985), no. 1, 51 -- 95. [15] G. Hao, C.-K. Ng, Crossed products of C*-correspondences by amenable group actions, J. Math. Anal. Appl. 345 (2008), no. 2, 702 -- 707. [16] M. Izumi, Finite group actions on C ∗-algebras with the Rohlin property I, Duke Math. J. 122 (2004), no. 2, 233 -- 280. [17] T. Kajiwara, C. Pinzari, Y. Watatani, Ideal structure and simplicity of the C ∗-algebras generated by Hilbert bimodules, J. Funct. Anal. 159 (1998), no. 2, 295 -- 322. [18] S. Kaliszewski, N. Patani, J.Quigg, Characterizing correspondences, Houston J. Math. 38 (2012), no. 3, 751 -- 759. graph C ∗- [19] T. Katsura, A class of C ∗-algebras generalizing both graph algebras and homeomorphism C ∗-algebras. I. Fundamental results, Trans. Amer. Math. Soc. 356(11) (2004), 4287 -- 4322. [20] T. Katsura, A construction of actions on Kirchberg algebras which induce given actions on their K-groups, J. reine angew. Math., 617 (2008), 27 -- 65. [21] A. Kishimoto, Outer automorphisms and reduced crossed products of simple C*-algebras, Comm. Math. Phys. 81 (1981), no. 3, 429 -- 435. [22] A. Kishimoto, Actions of finite groups on certain inductive limit C ∗-algebras, Internat. J. Math. 1 (1990), no. 3, 267 -- 292. [23] A. Kishimoto, A. Kumjian, Crossed products of Cuntz algebras by quasi-free automorphisms, Fields Inst. Comm., 13 (1997), 173 -- 192. [24] A. Kumjian, D. Pask, C ∗-algebras of directed graphs and group actions, Ergod. Th. Dyn. Sys. 19 (1999) 1503 -- 1519. [25] A. Kumjian, D. Pask, I. Raeburn, J. Renault, Graphs, Groupoids and Cuntz- Krieger algebras, J. of Funct. Anal. 144 (1997) No. 2, 505 -- 541. [26] M.H.Mann, I. Raeburn, C.E. Sutherland, Representations of finite groups and Cuntz-Krieger algebras, Bull. Austral. Math. Soc. 46 (1992), 225 -- 243. [27] P. Muhly, D. Pask, M. Tomforde, Strong shift equivalence of C ∗- correspondences, Israel J. Math. 167 (2008), 315 -- 346. [28] P. Muhly, B. Solel, On the Morita equivalence of tensor algebras Proc. Lon- don Math. Soc. (3) 81 (2000), no. 1, 113 -- 168. [29] V. Nekrashevych, C ∗-algebras and self-similar groups, J. Reine Angew. Math. 630 (2009) 59 -- 123. [30] M. V. Pimsner, Lectures on KK-theory (1989) (unpublished). [31] M. V. Pimsner, A class of C ∗-algebras generalizing both Cuntz-Krieger al- gebras and crossed products by Z, Fields Inst. Comm. 12 (1997), 189 -- 212. [32] I. Raeburn, W. Szyma´nski, Cuntz-Krieger algebras of infinite graphs and matrices, Trans. AMS 356 (2004), 39 -- 59. 22 VALENTIN DEACONU [33] E. Vasselli, The C ∗-algebra of a vector bundle and fields of Cuntz algebras, J. Funct. Anal. 222 (2005), no. 2, 491 -- 502. [34] A. Wassermann, Automorphic actions of compact groups on operator alge- bras, Ph.D. thesis, 1981. Received March 26, 2015 Valentin Deaconu, Department of Math & Stat (084), University of Nevada, Reno NV 89557-0084, USA E-mail address: [email protected]
1903.02911
1
1903
2019-03-07T13:54:07
Tight and cover-to-join representations of semilattices and inverse semigroups
[ "math.OA" ]
We discuss the relationship between tight and cover-to-join representations of semilattices and inverse semigroups, showing that a slight extension of the former, together with an appropriate selection of co-domains, makes the two notions equivalent. As a consequence, when constructing universal objects based on them, one is allowed to substitute cover-to-join for tight and vice-versa.
math.OA
math
TIGHT AND COVER-TO-JOIN REPRESENTATIONS OF SEMILATTICES AND INVERSE SEMIGROUPS R. Exel∗ We discuss the relationship between tight and cover-to-join representations of semilattices and inverse semi- groups, showing that a slight extension of the former, together with an appropriate selection of co-domains, makes the two notions equivalent. As a consequence, when constructing universal objects based on them, one is allowed to substitute cover-to-join for tight and vice-versa. 1. Introduction. Exactly twelve years ago, to be precise on March 7, 2007, I posted a paper on the arXiv [3] describing the notion of tight representations of semilattices and inverse semigroups, which turned out to have many applications and in particular proved to be useful to give a unified perspective to a significant number of C*-algebras containing a preferred generating set of partial isometries ([1], [2], [4], [6], [8], [9], [14], [15]). The notion of tight representations (described below for the convenience of the reader) is slightly in- volving as it depends on the analysis of certain pairs of finite sets X and Y , but it becomes much simplified when X is a singleton and Y is empty (see [4: Proposition 11.8]). In this simplified form it has been re- discovered and used in many subsequent works (e.g. [2], [10], [11], [12]) under the name of cover-to-join representations. The notion of cover-to-join representations, requiring a smaller set of conditions, is consequently weaker and, as it turns out, strictly weaker, than the original notion of tightness. Nevertheless, besides being easier to formulate, the notion of cover-to-join representations has the advantage of being applicable to representations taking values in generalized Boolean algebras, that is, Boolean algebras without a unit. Explicitly mentioning the operation of complementation, tight representations only make sense for unital Boolean algebras. The goal of this note is to describe an attempt to reconcile the notions of tight and cover-to-join representations: slightly extending the former, and adjusting for the appropriate co-domains, we show that, after all, the two notions coincide. One of the main practical consequences of this fact is that the difference between the two notions becomes irrelevant for the purpose of constructing universal objects based on them, such as the completion of an inverse semigroup recently introduced in [12]. We are moreover able to fix a slight imprecision in the proof of [2: Theorem 2.2], at least as far as its consequence that the universal C*-algebras for tight vs. cover-to-join representations are isomorphic. 2. Generalized Boolean algebras. We begin by recalling the well known notion of generalized Boolean algebras. 2.1. Definition. A generalized Boolean algebra [16: Definition 5] is a set B equipped with binary operations ∧ and ∨, and containing an element 0, such that for every a, b and c in B, one has that (i) (commutativity) a ∨ b = b ∨ a, and a ∧ b = b ∧ a, (ii) (associativity) (a ∧ b) ∧ c = a ∧ (b ∧ c), (iii) (distributivity) a ∧ (b ∨ c) = (a ∧ b) ∨ (a ∧ c), (iv) a ∨ 0 = a, (v) (relative complement) if a = a ∧ b, there is an element x in B, such that x ∨ a = b, and x ∧ a = 0, (vi) a ∨ a = a = a ∧ a. ∗ Universidade Federal de Santa Catarina and University of Nebraska at Lincoln. Date: March 7, 2019. 1 It follows that (ii) and (iii) also hold with ∨ and ∧ interchanged, meaning that ∨ is associative [16: Theorems 55 & 14], and that ∨ distributes over ∧ [16: Theorems 55 & 11]. When a = a ∧ b, as in (v), one writes a ≤ b. It is then easy to see that ≤ is a partial order on B. The element x referred to in (v) is called the relative complement of a in b, and it is usually denoted b \ a. 2.2. Definition. (cf. [16: Theorem 56]) A generalized Boolean algebra B is called a Boolean algebra if there exists an element 1 in B, such that a ∧ 1 = a, for every a in B. For Boolean algebras, the complement of an element a relative to 1 is often denoted ¬ a. Recall that an ideal of a generalized Boolean algebra B is any nonempty subset C of B which is closed under ∨, and such that a ≤ b ∈ C ⇒ a ∈ C. Such an ideal is evidently also closed under ∧ and under relative complements, so it is a generalized Boolean algebra in itself. Given any nonempty subset S of B, notice that the subset C defined by C = (cid:8)a ∈ B : a ≤ Wz∈Z z, for some finite subset Z ⊆ S(cid:9), is an ideal of B and it is clearly the smallest ideal containing S, so we shall call it the ideal generated by S, and we shall denote it by (cid:10)S(cid:11). 3. Tight and cover-to-join representations of semilattices. ◮ From now on let us fix a semilattice E (always assumed to have a zero element). 3.1. Definition. A representation of E in a generalized Boolean algebra B is any map π : E → B, such that (i) π(0) = 0, and (ii) π(x ∧ y) = π(x) ∧ π(y), for every x and y in E. In order to spell out the definition of the notion of tight representations, introduced in [4], let F be any subset of E. We then say that a given subset Z ⊆ F is a cover for F , if for every nonzero x in F , there exists some z in Z, such that z ∧ x 6= 0. Furthermore, if X and Y are finite subsets of E, we let EX,Y = {z ∈ E : z ≤ x, ∀x ∈ X, and z ⊥ y, ∀y ∈ Y }. 3.2. Definition. (cf. [4: Definition 11.6]) A representation π of E in a Boolean algebra B is said to be tight if, for any finite subsets X and Y of E, and for any finite cover Z for EX,Y , one has that _ π(z) = ^ π(x) ∧ ^ ¬ π(y). z∈Z x∈X y∈Y Observe that if Y is empty and X is a singleton, say X = {x}, then EX,Y = E{x},∅ = {z ∈ E : z ≤ x}, and if Z is a cover for this set, then (3.2.1) reads π(z) = π(x). _ z∈Z 2 (3.2.1) (3.3) To check that a given representation is tight, it is not enough to verify (3.3), as it is readily seen by considering the example in which E = {0, 1} and B is any Boolean algebra containing an element x 6= 1. Indeed, the map π : E → B given by π(0) = 0, and π(1) = x, satisfies all instances of (3.3) even thought it is not tight. The reader might wonder if the fact that π fails to preserve the unit is playing a part in this counter-example, but it is also easy to find examples of cover-to-join representations of non-unital semilattices which are not tight. Representations π satisfying (3.3) whenever Z is a cover for E{x},∅ have been considered in [4: Propo- sition 11.8], and they have been called cover-to-join representations in [2]. It is a trivial matter to prove that a cover-to-join representation satisfies (3.2.1) whenever X is nonempty (see the proof of [4: Lemma 11.7]), so the question of whether a cover-to-join representation is indeed tight rests on verifying (3.2.1) when X is empty. In this case, and assuming that Z is a cover for E∅,Y , it is easy to see that Z ∪ Y is a cover for the whole of E. Should we be dealing with a semilattice not admitting any finite cover, this situation will therefore never occur, that is, one will never be required to check (3.2.1) for an empty set X, hence every cover-to-join representation is automatically tight. This has in fact already been observed in [4: Proposition 11.8], which says that every cover-to-join representation is tight in case E does not admit any finite cover, as we have just discussed, but also if E contains a finite set X such that π(x) = 1. _ x∈X (3.4) The latter condition is useful for dealing with characters, i.e. with representations of E in the Boolean algebra {0, 1}, because the requirement that a character be nonzero immediately implies (3.4), so again cover-to-join suffices to prove tightness. On the other hand, an advantage of the notion of cover-to-join representations is that it makes sense for representations in generalized Boolean algebras, while the reference to the unary operation ¬ in (3.2.1) precludes it from being applied when the target algebra lacks a unit, that is, for a representation into a generalized Boolean algebra. Again referring to the occurrence of ¬ in (3.2.1), observe that if X is nonempty, then the right hand side of (3.2.1) lies in the ideal of B generated by the range of π. This is because, even though ¬ π(y) is not necessarily in (cid:10)π(E)(cid:11), this term will appear besides π(x), for some x in X, and hence π(x) ∧ ¬ π(y) = π(x) \ (cid:0)π(x) ∧ π(y)(cid:1) ∈ (cid:10)π(E)(cid:11). This means that: 3.5. Proposition. If E is a semilattice not admitting any finite cover then, whenever X and Y are finite subsets of E, and Z is a finite cover of EX,Y , the right hand side of (3.2.1) lies in (cid:10)π(E)(cid:11). As a consequence we see that definition (3.2) may be safely applied to a representation of E in a generalized Boolean algebra, as long as E does not admit a finite cover: despite the occurrence of ¬ in (3.2.1), once its right hand side is expanded, it may always be expressed in terms of relative complements, hence avoiding the use of the missing unary operation ¬ . We may therefore consider the following slight generalization of the notion of tight representations: 3.6. Definition. A representation π of E in a generalized Boolean algebra B is said to be tight if, either B is a Boolean algebra and π is tight in the sense of (3.2), or the following two conditions are verified: (i) E admits no finite cover, and (ii) (3.2.1) holds for any finite subsets X and Y of E, and for any finite cover Z for EX,Y . As already stressed, despite the occurrence of ¬ in (3.2.1), condition (3.6.ii) will always make sense in a generalized Boolean algebra. So here is a result that perhaps may be used to reconcile the notions of tightness and cover-to-join representations: 3 3.7. Theorem. Let π be a representation of the semilattice E in the generalized Boolean algebra B. Then (i) if π is tight then it is also cover-to-join, (ii) if π is cover-to-join then there exists an ideal B′ of B, containing the range of π, such that, once π is seen as a representation of E in B′, one has that π is tight. Proof. Point (i) being immediate, let us prove (ii). Under the assumption that E does not admit any finite cover, we have that π is tight as a representation into B′ = B, by [4: Proposition 11.8], or rather by its obvious adaptation to generalized Boolean algebras. It therefore remains to prove (ii) in case E does admit a finite cover, say Z. Setting we claim that e = _ π(z), z∈Z π(x) ≤ e, ∀ x ∈ E. (3.7.1) (3.7.2) To see this, pick x in E and notice that, since Z is a cover for E, we have in particular that the set is a cover for x, so the cover-to-join property of π implies that {z ∧ x : z ∈ Z} π(x) = _ π(z ∧ x) ≤ _ π(z) = e, z∈Z z∈Z proving (3.7.2). We therefore let B′ = {a ∈ B : a ≤ e}, which is evidently an ideal of B containing the range of π by (3.7.2). By (3.7.1) we then have that π satisfies [4: Lemma 11.7.(i)], as long as we see π as a representation of (cid:3) E in B′, whose unit is clearly e. The result then follows from [4: Proposition 11.8]. 4. Non-degenerate representations of semilattices. The following is perhaps the most obvious adaptation of the notion of non-degenerate representations ex- tensively used in the theory of operator algebras [17: Definition 9.3]. 4.1. Definition. We shall say that a representation π of a semilattice E in a generalized Boolean algebra B is non-degenerate if, for every a in B, there is a finite subset Z of E such that a ≤ Wz∈Z π(z). In other words, π is non-degenerate if and only if B coincides with the ideal generated by the range of π. Observe that, if both E and B have a unit, and if π is a unital map, then π is evidently non-degenerate. More generally, if π satisfies (3.4), then the same is also clearly true. The following result says that, by adjusting the co-domain of a representation, we can always make it non-degenerate. 4.2. Proposition. Let π be a representation of E in the generalized Boolean algebra B. Letting C be the ideal of B generated by the range of π, one has that π is a non-degenerate representation of E in C. Proof. Obvious. (cid:3) For non-degenerate representations we have the following streamlined version of (3.7): 4.3. Corollary. Let π be a non-degenerate representation of the semilattice E in the generalized Boolean algebra B. Then π is tight if and only if it is cover-to-join. Proof. The "only if" direction being trivial, we concentrate on the "if" part, so let us assume that π is cover-to-join. By (3.7) there exists an ideal B′ of B, containing the range of π, and such that π is tight as a representation in B′. Such an ideal will therefore contain the ideal generated by π(E), which coincides with B by hypothesis. Therefore B′ = B, and hence π is tight as a representation into its default co-domain B. (cid:3) 4 5. Representations of inverse semigroups. By its very nature, the concept of a tight representation pertains to the realm of semilattices and Boolean algebras. However, given the relevance of the study of semilattices in the theory of inverse semigroups, tight representations have had a strong impact on the latter. Recall that a Boolean inverse semigroup (see [5] but please observe that this notion is not equivalent to the homonym studied in [10] and [18]) is an inverse semigroup whose idempotent semilattice E(S) is a Boolean algebra. In accordance with what we have been discussing up to now, it is sensible to give the following: 5.1. Definition. (i) A generalized Boolean inverse semigroup is an inverse semigroup whose idempotent semilattice is a generalized Boolean algebra. (ii) (cf. [4: Definition 13.1] and [5: Proposition 6.2]) If S is any inverse semigroup1 and T is a generalized Boolean inverse semigroup, we say that a homomorphism π : S → T (always assumed to preserve zero) is tight if the restriction of π to E(S) is a tight representation into E(T ), in the sense of (3.6). (iii) If π is as above, we say that π is cover-to-join if the restriction of π to E(S) is cover-to-join. We then have the following version of (3.7) and (4.2): 5.2. Corollary. Let π be a representation of the inverse semigroup S in the generalized Boolean inverse semigroup T . Then (i) if π is tight then it is also cover-to-join, (ii) if π is cover-to-join then there exists a generalized Boolean inverse sub-semigroup T ′ of T , containing the range of π, such that, once π is seen as a representation of S in T ′, one has that π is tight. (iii) if π is cover-to-join, and if the restriction of π to E(S) is non-degenerate, then π is tight. Proof. The proof is essentially contained in the proofs of (3.7) and (4.2), except maybe for the proof of (ii) under the assumption that E(S) admits a finite cover, say Z. In this case, let e be as in (3.7.1) and put T ′ = {t ∈ T : t∗t ≤ e, tt∗ ≤ e}, observing that T ′ is clearly an inverse sub-semigroup of T , and that its idempotent semilattice is a Boolean algebra. Given any s in S, observe that s∗s lies in E(S) and π(s)∗π(s) = π(s∗s) ≤ e, where the last inequality above follows as in (3.7.2). By a similar reasoning one shows that also π(s)π(s)∗ ≤ e, so we see that π(s) lies in T ′, and we may then think of π as a representation of S in T ′. As in (3.7), one may now easily prove that π becomes a tight representation into T ′. (cid:3) 6. Conclusion. As a consequence of the above results, when defining universal objects (such as semigroups, algebras or C*-algebras) for a class of representations of inverse semigroups, one may safely substitute cover-to-join for tight and vice-versa. Given the widespread use of tight representations, there are many instances where the above principle applies. Below we spell out one such result to concretely illustrate our point, but similar results may be obtained as trivial reformulations of the following: 6.1. Theorem. Let S be an inverse semigroup and let C∗ 13.3] for tight Hilbert space representations of S [4: Definition 13.1]. Also let C∗ tight(S) be the universal C*-algebra [4: Theorem (S) be the universal cover- to-join C*-algebra for cover-to-join Hilbert space representations of S. Then C∗ tight(S) ≃ C∗ cover- to-join (S). 1 All inverse semigroups in this note are required to have a zero. 5 Proof. It suffices to prove that C∗ let tight(S) also has the universal property for cover-to-join representations. So π : S → B(H) be a cover-to-join representation of S on some Hilbert space H. Should the idempotent semilattice of S admit no finite covers, one has that π is tight so there is nothing to do. On the other hand, assuming that Z is a finite cover for E(S), let e be as in (3.7.1). Writing He for the range of e and letting K = H ⊥ e , we then obviously have that H = He ⊕ K. It then follows from (3.7.2) that each π(s) decomposes as a direct sum of operators π(s) = π′(s) ⊕ 0, thus defining a representation π′ of S on He which is clearly also cover-to-join. It is also clear that π′ is non-degenerate on E(S), so we have by (5.2.iii) that π′ is tight. Therefore the universal property provides a *-representation ϕ′ of C∗ tight(S). It then follows that ϕ := ϕ′ ⊕ 0 coincides with π on S, concluding the proof. (cid:3) tight(S) on B(He) coinciding with π′ on the canonical image of S within C∗ References [1] G. Boava, G. G. de Castro, F. de L. Mortari, "Inverse semigroups associated with labelled spaces and their tight spectra", Semigroup Forum, 94 (2017), no. 3, 582 -- 609. [2] A. P. Donsig and D. Milan, "Joins and covers in inverse semigroups and tight C*-algebras", Bull. Aust. Math. Soc., 90 (2014), no. 1, 121 -- 133. [3] R. Exel, "Inverse semigroups and combinatorial C*-algebras (preprint version)", arXiv:math/0703182v1 [math.OA], March 7, 2007. [4] R. Exel, "Inverse semigroups and combinatorial C*-algebras", Bull. Braz. Math. Soc. (N.S.), 39 (2008), 191 -- 313. [5] R. Exel, "Tight representations of semilattices and inverse semigroups", Semigroup Forum, 79 (2009), 159 -- 182. [6] R. Exel, D. Goncalves, C. Starling, "The tiling C*-algebra viewed as a tight inverse semigroup algebra", Semigroup Forum, 84 (2012), no. 2, 229 -- 240. [7] R. Exel, E. Pardo, "The tight groupoid of an inverse semigroup", Semigroup Forum, 92 (2016), no. 1, 274 -- 303. [8] R. Exel, E. Pardo, "Self-similar graphs, a unified treatment of Katsura and Nekrashevych C∗-algebras", Adv. Math., 306 (2017), 1046 -- 1129. [9] R. Exel, C. Starling, "Self-similar graph C ∗-algebras and partial crossed products", J. Operator Theory, 75 (2016), no. 2, 299 -- 317. [10] M. V. Lawson, "Non-commutative Stone duality: inverse semigroups, topological groupoids and C -algebras", Internat. J. Algebra Comput. 22 (2012), no. 6, 1250058, 47 pp. [11] M. V. Lawson and D. G. Jones, "Graph inverse semigroups: Their characterization and completion", J. of Algebra, 409 (2014), 444 -- 473. [12] M. V. Lawson and A. Vdovina, "The universal Boolean inverse semigroup presented by the abstract Cuntz-Krieger relations", arXiv:1902.02583v3 [math.OA], 2018. [13] D. Milan, B. Steinberg, "On inverse semigroup C ∗-algebras and crossed products", Groups Geom. Dyn., 8 (2014), no. 2, 485 -- 512. [14] C. Starling, "Boundary quotients of C∗-algebras of right LCM semigroups", J. Funct. Anal., 268 (2015), no. 11, 3326 -- 3356. [15] C. Starling, "Inverse semigroups associated to subshifts", J. Algebra, 463 (2016), 211 -- 233. [16] M. H. Stone, "Postulates for Boolean Algebras and Generalized Boolean Algebras", Amer. J. Math., 57 (1935), 703 -- 732. [17] M. Takesaki, "Theory of operator algebras. I", Springer-Verlag, New York-Heidelberg, 1979. vii+415 pp.. [18] F. Wehrung, "Refinement monoids, equidecomposability types, and Boolean inverse semigroups", Lecture Notes in Mathematics, 2188, Springer, 2017. 6
0808.4049
2
0808
2010-02-24T15:32:38
The Unitary Implementation of a Measured Quantum Groupoid action
[ "math.OA" ]
Mimicking the von Neumann version of Kustermans and Vaes' locally compact quantum groups, Franck Lesieur had introduced a notion of measured quantum groupoid, in the setting of von Neumann algebras. In a former article, the author had introduced the notion of actions, crossed-product, dual actions of a measured quantum groupoid: a biduality theorem for actions had been proved. This article continues that program : we prove the existence of a standard implementation for an action, and a bidulaity theorem for weights. We generalize this way results which were proved, for locally compact quantum groups by S. Vaes, and for measured groupoids by T. Yamanouchi.
math.OA
math
THE UNITARY IMPLEMENTATION OF A MEASURED QUANTUM GROUPOID ACTION MICHEL ENOCK Abstract. Mimicking the von Neumann version of Kustermans and Vaes' locally com- pact quantum groups, Franck Lesieur had introduced a notion of measured quantum groupoid, in the setting of von Neumann algebras. In a former article, the author had introduced the notions of actions, crossed-product, dual actions of a measured quantum groupoid; a biduality theorem for actions has been proved. This article continues that program : we prove the existence of a standard implementation for an action, and a bid- uality theorem for weights. We generalize this way results which were proved, for locally compact quantum groups by S. Vaes, and for measured groupoids by T. Yamanouchi. 0 1 0 2 b e F 4 2 ] . A O h t a m [ 2 v 9 4 0 4 . 8 0 8 0 : v i X r a 1 1. Introduction 1.1. In two articles ([Val1], [Val2]), J.-M. Vallin has introduced two notions (pseudo- multiplicative unitary, Hopf-bimodule), in order to generalize, up to the groupoid case, the classical notions of multiplicative unitary [BS] and of Hopf-von Neumann algebras [ES] which were introduced to describe and explain duality of groups, and leaded to appropriate notions of quantum groups ([ES], [W1], [W2], [BS], [MN], [W3], [KV1], [KV2], [MNW]). In another article [EV], J.-M. Vallin and the author have constructed, from a depth 2 inclusion of von Neumann algebras M0 ⊂ M1, with an operator-valued weight T1 verifying a regularity condition, a pseudo-multiplicative unitary, which leaded to two structures of Hopf bimodules, dual to each other. Moreover, we have then constructed an action of one of these structures on the algebra M1 such that M0 is the fixed point subalgebra, the algebra M2 given by the basic construction being then isomorphic to the crossed-product. We construct on M2 an action of the other structure, which can be considered as the dual action. If the inclusion M0 ⊂ M1 is irreducible, we recovered quantum groups, as proved and studied in former papers ([EN], [E2]). Therefore, this construction leads to a notion of "quantum groupoid", and a construction of a duality within "quantum groupoids". 1.2. In a finite-dimensional setting, this construction can be mostly simplified, and is studied in [NV1], [BSz1], [BSz2], [Sz],[Val3], [Val4], and examples are described. In [NV2], the link between these "finite quantum groupoids" and depth 2 inclusions of II1 factors is given. 1.3. F. Lesieur, in [L], starting from a Hopf-bimodule, as introduced in [Val1], when there exist a left-invariant operator-valued weight, and a right-invariant operator-valued weight, mimicking in that wider setting the technics of Kustermans and Vaes ([KV1], [KV2]), obtained a pseudo-multiplicative unitary, which, as in quantum group theory, "contains" all the information about the object (the von Neumann algebra, the coproduct) and allows to construct important data (an antipod, a co-inverse, etc.) Lesieur gave the name of "measured quantum groupoids" to these objects. A new set of axioms for these had been given in an appendix of [E5]. In [E4] had been shown that, with suitable conditions, the objects constructed from [EV] are "measured quantum groupoids" in the sense of Lesieur. In [E5] have been developped the notions of action (already introduced in [EV]), 1.4. crossed-product, etc, following what had been done for locally compact quantum groups in ([E1], [ES1], [V1]); a biduality theorem for actions had been obtained in ([E5], 11.6). Several points were left apart in [E5], namely the generalization of Vaes' theorem ([V1], 4.4) on the standard implementation of an action of a locally compact quantum group, which was the head light of [V1], and a biduality theorem for weights, as obtained in [Y3], [Y4] (in fact, we were much more inspired by the shorter proof given in an appendix of [BV]). We solve here these two problems when there exists a normal semi-finite faithful operator- valued weight from the von Neumann algebra on which the measured quantum groupoid is acting, onto the copy of the basis of this measured quantum groupoid which is put inside this algebra. In fact, these results appear much more as a biduality theorem of operator-valued weights rather than a biduality theorem on weights, which seems quite natural in the spirit of measured quantum groupoids, where, for instance, left-invariant 2 weight on a locally compact quantum group is replaced by a left-invariant operator-valued weight. The strategy for the proofs had been mostly inspired by [V1] and [BV]. 1.5. This article is organized as follows : In chapter 2, we recall very quickly all the notations and results needed in that article; we have tried to make these preliminaries as short as possible, and we emphazise that this article should be understood as the continuation of [E5]. In chapter 3, we follow ([V1], 4.1 to 4.4), and prove, for any dual action, the result on the standard implementation of an action. Chapter 4 is rather technical; let G = (N, M, α, β, Γ, T, T ′, ν) be a measured quantum groupoid, and let b be an injective ∗-anti-homomorphism from N into a von Neumann algebra A; let us suppose that there exists a normal semi-finite faithful operator-valued weight T from A onto b(N), and let us write ψ = νo ◦ b−1 ◦ T. Then, we can define on A b∗α L(H) a weight ψ, which will generalize the tensor product of ψ and T r b∆−1 (when N G is a locally compact quantum group, and therefore N = C). In chapter 5, using this auxilliary weight introduced in chapter 4, and the particular case of the dual actions studied in chapter 3, we calculate the standard implementation of an action, whenever there exists a normal semi-finite faithful operator-valued weight from A onto b(N). This condition is fulfilled trivially when the measured quantum groupoid is a locally compact quantum group, or is a measured groupoid; therefore, we recover in both cases the results already obtained. Chapter 6 is another technical chapter; we define conditions on a weight ψ defined on L(H) a weight ψδ which generalize the tensor A which allow us to construct on A b∗α N product of ψ and T r(δ b∆)−1(when G is a locally compact quantum group, and therefore N = C). In chapter 7 we use both auxilliary weights constructions made in chapters 4 and 6; then, when there exists a normal semi-finite faithful operator-valued weight T from A onto b(N) such that ψ = νo ◦ b−1 ◦ T, we can define a Radon-Nikodym derivative of the weight ψ with respect to the action, which will be a cocycle for this action. This condition is fulfilled trivially when the measured quantum groupoid is a locally compact quantum group, or is a measured groupoid, and, therefore, we recover in both cases the results already obtained. 2. Definitions and notations This article is the continuation of [E5]; preliminaries are to be found in [E5], and we just recall herafter the following definitions and notations : 2.1. Spatial theory; relative tensor products of Hilbert spaces and fiber prod- ucts of von Neumann algebras ([C1], [S], [T], [EV]). Let N a von Neumann algebra, ψ a normal semi-finite faithful weight on N; we shall denote by Hψ, Nψ, Sψ, Jψ, ∆ψ... the canonical objects of the Tomita-Takesaki theory associated to the weight ψ; let α be a non degenerate faithful representation of N on a Hilbert space H; the set of ψ-bounded elements of the left-module αH is : D(αH, ψ) = {ξ ∈ H; ∃C < ∞, kα(y)ξk ≤ CkΛψ(y)k, ∀y ∈ Nψ} Then, for any ξ in D(αH, ψ), there exists a bounded operator Rα,ψ(ξ) from Hψ to H, defined, for all y in Nψ by : Rα,ψ(ξ)Λψ(y) = α(y)ξ 3 which intertwines the actions of N. If ξ, η are bounded vectors, we define the operator product < ξ, η >α,ψ= Rα,ψ(η)∗Rα,ψ(ξ) belongs to πψ(N)′, which, thanks to Tomita-Takesaki theory, will be identified to the opposite von Neumann algebra N o. If now β is a non degenerate faithful antirepresentation of N on a Hilbert space K, the relative tensor product K β⊗α H is the completion of the algebraic tensor product ψ K ⊙ D(αH, ψ) by the scalar product defined, if ξ1, ξ2 are in K, η1, η2 are in D(αH, ψ), by the following formula : (ξ1 ⊙ η1ξ2 ⊙ η2) = (β(< η1, η2 >α,ψ)ξ1ξ2) If ξ ∈ K, η ∈ D(αH, ψ), we shall denote ξ β⊗α η the image of ξ ⊙ η into K β⊗α H, and, writing ρβ,α η (ξ) = ξ β⊗α ψ ψ η, we get a bounded linear operator from K into K β⊗α ψ H, which ν is equal to 1K ⊗ψ Rα,ψ(η). Changing the weight ψ will give a canonical isomorphic Hilbert space, but the isomor- phism will not exchange elementary tensors ! We shall denote σψ the relative flip, which is a unitary sending K β⊗α ψ H onto H α⊗β ψo K, defined, for any ξ in D(Kβ, ψo), η in D(αH, ψ), by : σψ(ξ β⊗α ψ η) = η α⊗β ψo ξ In x ∈ β(N)′, y ∈ α(N)′, it is possible to define an operator x β⊗α y on K β⊗α H, with natural values on the elementary tensors. As this operator does not depend upon the weight ψ, it will be denoted x β⊗α y. We can define a relative flip ςN at the level of ψ ψ operators such that ςN (x β⊗α N N y) = y α⊗β N o x. If P is a von Neumann algebra on H, with α(N) ⊂ P , and Q a von Neumann algebra on K, with β(N) ⊂ Q, then we define the fiber product Q β∗α Q. N y, x ∈ Q′, y ∈ P ′}′, and we get that ςN (Q β∗α N P ) = P α∗β N o P as {x β⊗α N Moreover, this von Neumann algebra can be defined independantly of the Hilbert spaces on which P and Q are represented; if (i = 1, 2), αi is a faithful non degenerate homomor- phism from N into Pi, βi is a faithful non degenerate antihomomorphism from N into Qi, and Φ (resp. Ψ) an homomorphism from P1 to P2 (resp. from Q1 to Q2) such that Φ ◦ α1 = α2 (resp. Ψ ◦ β1 = β2), then, it is possible to define an homomorphism Ψ β1∗α1 Φ from Q1 β1∗α1 P1 into Q2 β2∗α2 P2. N N N The operators θα,ψ(ξ, η) = Rα,ψ(ξ)Rα,ψ(η)∗, for all ξ, η in D(αH, ψ), generates a weakly dense ideal in α(N)′. Moreover, there exists a family (ei)i∈I of vectors in D(αH, ψ) such that the operators θα,ψ(ei, ei) are 2 by 2 orthogonal projections (θα,ψ(ei, ei) being then the projection on the closure of α(N)ei). Such a family is called an orthogonal (α, ψ)-basis of H. 2.2. Measured quantum groupoids ([L], [E5]). A measured quantum groupoid is an octuplet G = (N, M, α, β, Γ, T, T ′, ν) such that ([E5], 3.8) : (i) (N, M, α, β, Γ) is a Hopf-bimodule (as defined in [E5], 3.1), 4 (ii) T is a left-invariant normal, semi-finite, faithful operator valued weight T from M to α(N), (iii) T ′ is a right-invariant normal, semi-finite, faithful operator-valued weight T ′ from M to β(N), (iv) ν is normal semi-finite faitfull weight on N, which is relatively invariant with respect to T and T ′. We shall write Φ = ν ◦α−1 ◦T , and H = HΦ, J = JΦ, and, for all n ∈ N, β(n) = Jα(n∗)J, α(n) = Jβ(n∗)J. The weight Φ will be called the left-invariant weight on M. Then, G can be equipped with a pseudo-multiplicative unitary W from H β⊗α H onto H ([E5], 3.6), a co-inverse R, a scaling group τt, an antipod S, a modulus δ, ν H α⊗ β νo a scaling operator λ, a managing operator P , and a canonical one-parameter group γt of automorphisms on the basis N ([E5], 3.8). Instead of G, we shall mostly use (N, M, α, β, Γ, T, RT R, ν) which is another measured quantum groupoid, denoted G, which is equipped with the same data (W , R, ...) as G. Canonically associated to G, can be defined also the opposite measured quantum groupoid is Go = (N o, M, β, α, ςN Γ, RT R, T, νo) and the commutant measured quantum groupoid A dual measured quantum group bG, which is denoted (N,cM , α, β,bΓ, bT , bRbT bR, ν), can be constructed, and we have bbG = G. Gc = (N o, M ′, β, α, Γc, T c, RcT cRc, νo); we have (Go)o = (Gc)c = G, cGo = (bG)c, cGc = (bG)o, and Goc = Gco is canonically isomorphic to G ([E5], 3.12). The pseudo-multiplicative unitary of bG (resp. Go, Gc) will be denoted cW (resp. W o, W c). The left-invariant weight on bG (resp. Go, Gc) will be denoted bΦ (resp. Φo, Φc). Let aHb be a N − N-bimodule, i.e. an Hilbert space H equipped with a normal faithful non degenerate representation a of N on H and a normal faithful non degenerate anti- representation b on H, such that b(N) ⊂ a(N)′. A corepresentation of G on aHb is a unitary V from H a⊗β νo ν HΦ onto H b⊗α HΦ, satisfying, for all n ∈ N : V (b(n) a⊗β N o 1) = (1 b⊗α N β(n))V V (1 a⊗β N o α(x)) = (a(n) b⊗α N 1)V such that, for any ξ ∈ D(aH, ν) and η ∈ D(Hb, νo), the operator (ωξ,η ∗ id)(V ) belongs to M (then, it is possible to define (id ∗ θ)(V ), for any θ in M α,β ∗ which is the linear set generated by the ωξ, with ξ ∈ D(αH, ν) ∩ D(Hβ, νo)), and such that the application θ → (id ∗ θ)(V ) from M α,β into L(H) is multiplicative ([E5] 5.1, 5.5). ∗ 2.3. Action of a measured quantum groupoid ([E5]). An action ([E5], 6.1) of G on a von Neumann algebra A is a couple (b, a), where : (i) b is an injective ∗-antihomomorphism from N into A; (ii) a is an injective ∗-homomorphism from A into A b∗α N M; (iii) b and a are such that, for all n in N: (which allow us to define a b∗α N id from A b∗α N M into A b∗α N M β∗α N M) and such that : a(b(n)) = 1 b⊗α N β(n) (a b∗α N id)a = (id b∗α N Γ)a 5 The set of invariants is defined as the sub von Neumann algebra : Aa = {x ∈ A ∩ b(N)′, a(x) = x b⊗α N 1} If the von Neumann algebra acts on a Hilbert space H, and if there exists a representation a of N on H such that b(N) ⊂ A ⊂ a(N)′, a corepresentation V of G on the bimodule aHb 1)V ∗ , for all x ∈ A ([E5], will be called an implementation of a if we have a(x) = V (x a⊗b N o 6.6); we shall look at the following more precise situation : let ψ is a normal semi-finite faithful weight on A, and V an implementation of a on a(Hψ)b (with a(n) = Jψb(n∗)Jψ), such that : V ∗ = (Jψ α⊗β νo J bΦ)V (Jψ b⊗α ν J bΦ) Such an implementation had been constructed ([E5] 8.8) in the particular case when the weight ψ is called δ-invariant, which means that, for all η ∈ D(αHΦ, ν) ∩ D(δ1/2), such that δ1/2η belongs to D((HΦ)β, νo), and for all x ∈ Nψ, we have: δ1/2ηk2 ψ((id b∗α N ωη)a(x∗x)) = kΛψ(x) a⊗β νo and bears the density property, which means that the subset D((Hψ)b, νo) ∩ D(aHψ, ν) is dense in Hψ. This standard implementation is then given by the formula ([E5], 8.4) : Vψ(Λψ(x) a⊗β νo δ1/2η) =Xi Λψ((id b∗α N ωη,ei)a(x)) b⊗α ν ei for all x ∈ Nψ, η ∈ D(αH, ν) ∩ D(δ1/2) such that δ1/2η belongs to D(Hβ, νo), (ei)i∈I any orthonormal (α, ν)-basis of H. Moreover ([E5], 8.9), it is possible to define one parameter groups of unitaries ∆it , with natural values and ∆it δ−it∆−it bΦ ψ b⊗α N δ−it∆−it bΦ ψ a⊗β N o on elementary tensor, and we have : Vψ(∆it ψ a⊗β N o δ−it∆−it bΦ ) = (∆it ψ b⊗α N δ−it∆−it bΦ )Vψ and, therefore, for any x in A, t in R, we have : a(σψ t (x)) = (∆it ψ b⊗α N δ−it∆−it bΦ )a(x)(∆−it ψ b⊗α N δit∆it bΦ) 2.4. Crossed-product ([E5]). The crossed-product of A by G via the action a is the von Neumann algebra generated by a(A) and 1 b⊗α N cM ′ ([E5], 9.1) and is denoted A ⋊a G; α, a) of (bG)c on A ⋊a G. The biduality theorem ([E5], 11.6) says that the bicrossed-product (A ⋊a G) ⋊a bGc is then there exists ([E5], 9.3) an action (1 b⊗α N L(H); more precisely, this isomorphism is given by : canonically isomorphic to A b∗α N Θ(a b∗α N id)(A b∗α N L(H)) = (A ⋊a G) ⋊a bGc where Θ is the spatial isomorphism between L(H b⊗α H β⊗α H) and L(H b⊗α ν ν H α⊗β νo H) ν implemented by 1H b⊗α ν sends the action (1 b⊗α N σν W oσν; the biduality theorem says also that this isomorphism β, a) of G on A b∗α N L(H), by : L(H), defined, for any X ∈ A b∗α N σνoW σνo)∗ a(X) = (1 b⊗α N σνoW σνo)(id b∗α N ςN )(a b∗α N id)(X)(1 b⊗α N 6 on the bidual action (of Gco) on (A ⋊a G) ⋊a bGo. There exists a normal faithful semi-finite operator-valued weight Ta from A ⋊a G onto a(A); therefore, starting with a normal semi-finite weight ψ on A, we can construct a dual weight ψ on A ⋊a G by the formula ψ = ψ ◦ a−1 ◦ Ta ([E5] 13.2). These dual weights are exactly the δ−1-invariant weights on A ⋊a G bearing the density property ([E5] 13.3). Moreover ([E5] 13.3), the linear set generated by all the elements (1 b⊗α a)a(x), for N all x ∈ Nψ, a ∈ N bΦc ∩ N T c, is a core for Λ ψ, and it is possible to identify the GNS representation of A ⋊a G associated to the weight ψ with the natural representation on Hψ b⊗α HΦ by writing : ν Λψ(x) b⊗α ν Λ bΦc(a) = Λ ψ[(1 b⊗α N a)a(x)] which leads to the identification of H ψ with Hψ b⊗α the linear set generated by the elements of the form a(y∗)(Λψ(x) b⊗α Nψ, and a in N bΦc ∩ N T c ∩ N∗ T c is a core for S ψ, and we have : Λ bΦc(a)) = a(x∗)(Λψ(y) b⊗α ν bΦc ∩ N∗ S ψa(y∗)(Λψ(x) b⊗α ν ν ν Λ bΦc(a∗)) H. Moreover, using that identification, Λ bΦc(a)), for x, y in Then, the unitary U a ψ = J ψ(Jψ a⊗β N o J bΦ) from Hψ a⊗β νo HΦ onto Hψ b⊗α ν HΦ satisfies : and we have ([E5] 13.4) : (i) for all y ∈ A : U a ψ(Jψ b⊗α N J bΦ) = (Jψ b⊗α N J bΦ)(U a ψ)∗ a(y) = U a ψ(y a⊗β N o 1)(U a ψ)∗ (ii) for all b ∈ M : (iii) for all n ∈ N : (1 b⊗α N JΦbJΦ)U a ψ = U a ψ(1 a⊗β N o JΦbJΦ) U a ψ(b(n) a⊗β N o 1) = (1 b⊗α N β(n))U a ψ U a ψ(1 a⊗β N o α(n)) = (a(n) b⊗α N 1)U a ψ Therefore, we see that this unitary U a is a corepresentation. If it is, we shall say that it is a standard implemantation of a. ψ ◦ Θ ◦ (a b∗α We can define the bidual weight ψ "implements" a, but we do not know whether it id) ψ on (A ⋊a G) ⋊a bGo, and the weight L(H), that we shall denote ψa for simplification (or ψ if there is no ambiguity N on A b∗α N about the action). Then we get ([E5], 13.6) that the spatial derivative dψ dψo is equal to the modulus operator ∆ ψ. There exists a normal semi-finite faithful operator-valued weight Ta from A b∗α N L(H) onto A ⋊a G such that ψa = ψ ◦ Ta Using twice ([T] 4.22(ii)), we obtain, for any x ∈ A and t ∈ R, that σψa and if ψ1 and ψ2 are two normal semi-finite faithful weights on A, , we get : t (a(x)) = a(σψ t (x)); (Dψ1a : Dψ2a)t = (D ψ1 : D ψ2)t = a((Dψ1 : Dψ2)t) 7 2.5. Examples of measured quantum groupoids. Examples of measured quantum groupoids are the following : (i) locally compact quantum groups, as defined and studied by J. Kustermans and S. Vaes ([KV1], [KV2], [V1]); these are, trivially, the measured quantum groupoids with the basis N = C. (ii) measured groupoids, equipped with a left Haar system and a quasi-invariant measure on the set of units, as studied mostly by T. Yamanouchi ([Y1], [Y2], [Y3], [Y4]); it was proved in [E6] that these measured quantum groupoids are exactly those whose underly- ing von Neumann algebra is abelian. (iii) the finite dimensional case had been studied by D. Nikshych and L. Vainermann ([NV1], [NV2], [NV3]), J.-M. Vallin ([Val3], [Val4]) and M.-C. David ([D]); in that case, non trivial examples are given, for instance Temperley-Lieb algebras ([NV3], [D]), which had appeared in subfactor theory ([J]). . (iv) continuous fields of (C∗-version of) locally compact quantum groups, as studied by E. Blanchard in ([Bl1], [Bl2]); it was proved in [E6] that these measured quantum groupoids are exactly those whose basis is central in the underlying von Neumann algebras of both the measured quantum groupoid and its dual. (v) in [DC], K. De Commer proved that, in the case of a monoidal equivalence between two locally compact quantum groups (which means that these two locally compact quan- tum group have commuting ergodic and integrable actions on the same von Neumann algebra), it is possible to construct a measurable quantum groupoid of basis C2 which contains all the data. Moreover, this construction was usefull to prove new results on locally compact quantum groups, namely on the deformation of a locally compact quan- tum group by a unitary 2-cocycle; he proved that these measured quantum groupoids are exactly those whose basis C2 is central in the underlying von Neumann algebra of the measured quatum groupoid, but not in the underlying von Neumann algebra of the dual measured quantum groupoid. (vi) starting from a depth 2 inclusion M0 ⊂ M1 of von Neumann algebras, equipped with an operator-valued weight T1 from M1 onto M0, satisfying appropriate conditions, such that there exists a normal semi-finite faithful weight χ on the first relative com- mutant M ′ , it has been proved ([EV], [E4]) that it is possible to put on the second relative commutant M ′ 0 ∩ M2 (where M0 ⊂ M1 ⊂ M2 ⊂ M3... is Jones' tower associated to the inclusion M0 ⊂ M1) a canonical structure of a measured quantum groupoid; moreover, its dual is given then by the same construction associated to the inclusion M1 ⊂ M2, and this dual measured quantum groupoid acts canonically on the von Neumann algebra M1, in such a way that M0 is equal to the subalgebra of invariants, and the inclusion M1 ⊂ M2 is isomorphic to the inclusion of M1 into its crossed-product. This gives a "geometrical" construction of measured quantum groupoids; in another article in preparation ([E7]), in which is used the biduality theorem for weights proved in 7.3, had been proved that any measured quantum groupoid has an outer action on some von Neumann algebra, and can be, there- fore, obtained by this "geometrical construction". The same result for locally compact quantum groups relies upon [V2] and the corresponding result for measured quantum groupoids had been pointed out in [E5]. (vii) in [VV] and [BSV] was given a specific procedure for constructing locally compact quantum groups, starting from a locally compact group G, whose almost all elements be- long to the product G1G2 (where G1 and G2 are closed subgroups of G whose intersection is reduced to the unit element of G); such (G1, G2) is called a "matched pair" of locally compact groups (more precisely, in [VV], the set G1G2 is required to be open, and it is 0 ∩ M1, invariant under the modular automorphism group σT1 t 8 not the case in [BSV]).Then, G1 acts naturally on L∞(G2) (and vice versa), and the two crossed-products obtained bear the structure of two locally compact quantum groups in duality. In [Val5], J.-M. Vallin generalizes this constructions up to groupoids, and, then, obtains examples of measured quantum groupoids; more specific examples are then given by the action of a matched pair of groups on a locally compact space, and also more exotic examples. 3. The standard implementation of an action : the case of a dual action In this chapter, following [V1], we prove that the unitary U a ψ introduced in 2.4 is a standard implementation of a, for all normal semi-finite faithful weight ψ on A, whenever a is a dual action (3.4). For this purpose, we prove first that, if for some weight ψ1, the unitary U a ψ is a standard implementation (3.1). Second (3.2), we prove, for a δ-invariant weight ψ, that U a ψ is equal to the implementation Vψ constructed in ([E5] 8.8) and recalled in 2.3. Thanks to ([E5] 13.3), recalled in 2.4, we then get the result. ψ1 is a standard implementation, then, for any weight ψ, U a 3.1. Proposition. Let G be a measured quantum groupoid, and (b, a) an action of G on a von Neumann algebra A; let ψ1 and ψ2 be two normal faithful semi-finite weights on A and U a ψ2 the two unitaries constructed in 2.4; let u be the unitary from Hψ1 onto Hψ2 intertwining the representations πψ1 and πψ2; then : (i) the unitary u b⊗α N 1 intertwines the representations of A ⋊a G on Hψ1 b⊗α ψ1 and U a HΦ and on ν Hψ2 b⊗α HΦ; moreover, we have : ν (u b⊗α N 1)U a ψ1 = U a ψ2(u a1⊗β N o 1) where a1(n) = Jψ1πψ1(b(n∗))Jψ1, for all n ∈ N. (ii) if U a (iii) if U a a. ψ1 is a corepresentation of G on Hψ1, then U a ψ1 is a standard implementation of a, then U a ψ2 is a corepresentation of G on Hψ2. ψ2 is a standard implementation of Proof. Let us write J2,1 the relative modular conjugation, which is an antilinear surjective isometry from Hψ1 onto Hψ2. Then we have u = J2,1Jψ1 = Jψ2J2,1, by ([St] 3.16). Moreover, let us define, for x ∈ A, and t ∈ R σ2,1 t (x); then, by ([St], 3.15), for x ∈ Nψ1, y ∈ D(σ2,1 t (x) = [Dψ2 : Dψ1]tσψ1 −i/2), xy∗ belongs to Nψ2 and : Λψ2(xy∗) = J2,1πψ1(σ2,1 −i/2(y))Jψ1Λψ1(x) Therefore, if a ∈ N bΦc, (1 b⊗α N denotes the unitary from Hψi b⊗α HΦ onto H ψi defined in 2.4 : ν a)a(xy∗) belongs to N ψ2, and, we have, where Vi (i = (1, 2)) Λ ψ2[(1 b⊗α N a)a(xy∗)] = V2(Λψ2(xy∗) b⊗α ν Λ bΦc(a)) = V2J2,1πψ1(σ2,1 −i/2(y))Jψ1Λψ1(x) b⊗α Λ bΦc(a)) ν which is equal to : V2(J2,1πψ1(σ2,1 −i/2(y))Jψ1 b⊗α N 9 1)V ∗ 1 Λ ψ1[(1 b⊗α N a)a(x)] and, as the linear set generated by the elements of the form (1 b⊗α N Λ ψ1, we get, for any z ∈ N ψ1, that za(y∗) belongs to N ψ2, and that : a)a(x) is a core for Λ ψ2(za(y∗)) = V2(J2,1πψ1(σ2,1 −i/2(y))Jψ1 b⊗α N 1)V ∗ 1 Λ ψ1(z) Let us denote by J2,1 the relative modular conjugation constructed from the weights ψ1 and ψ2, and σ2,1 the one-parameter group of isometries of A ⋊a G constructed from these two weights by the formula, for any X ∈ A ⋊a G : t t (X) = [D ψ2 : D ψ1]tσ σ2,1 ψ1 t (X) Using ([St], 3.15) applied to these two weights, we get that a(y) belongs to D(σ2,1 that : −i/2) and J2,1π ψ1(σ2,1 −i/2(a(y)))J ψ1 = V2(J2,1πψ1(σ2,1 −i/2(y))Jψ1 b⊗α N 1)V ∗ 1 We easily get that σ2,1 t (a(y)) = a(σ2,1 −i/2(y)) = J2,1 ∗ π ψ1(a(σ2,1 t (y)) and, therefore, we have : V2(J2,1πψ1(σ2,1 −i/2(y))Jψ1 b⊗α N 1)V ∗ 1 J ψ1 As we have, using 2.4 : we get : (Jψ1 b⊗α N J bΦ)V ∗ 1 J ψ1 = U a ψ1 V ∗ 1 π ψ1(a(σ2,1 −i/2(y)) = J2,1 ∗ V2(J2,1 a1⊗β N o J bΦ)(πψ1(σ2,1 −i/2(y)) a1⊗β N o 1)U a ψ1V ∗ 1 and, therefore, using 2.4 : ∗ J2,1 V2(J2,1 a1⊗β N o J bΦ)(πψ1(σ2,1 −i/2(y)) a1⊗β N o 1) = π ψ1(a(σ2,1 −i/2(y))V1(U a ψ1)∗ = V1a(σ2,1 −i/2(y))(U a ψ1)∗ which, using 2.4, is equal to : By density, we get : V1U a ψ1(πψ1(σ2,1 −i/2(y)) a1⊗β N o 1) U a ψ1 = V ∗ 1 ∗ J2,1 V2(J2,1 a1⊗β N o J bΦ) and, therefore, using 2.4 again : 1Hψ1 b⊗α N 1HΦ = V ∗ 1 ∗ J2,1 = V ∗ 1 ∗ J2,1 J bΦ)(Jψ1 b⊗α N J bΦ)V ∗ 1 J ψ1 V1 V2(J2,1 a1⊗β N o 1)V ∗ V2(u b⊗α N 1 J ψ1 V1 which implies that : and : 1H ψ1 b⊗α N 1HΦ = J2,1 ∗ V2(u b⊗α N 1)V ∗ 1 J ψ1 V2(u b⊗α N 1)V ∗ 1 = J2,1J ψ1 10 But J2,1J ψ1 = J ψ2 from which we get the first result. This formula gives also, where a2(n) = Jψ2πψ2(b(n∗))Jψ2, for all n ∈ N : J2,1 is the unitary from H ψ1 onto H ψ2 which intertwines π ψ1 and π ψ2; U a ψ2 = V ∗ 2 J ψ2 V2(Jψ2 a2⊗β N o J2,1 1)V ∗ 1 J ψ1 J bΦ) ∗ J ψ2 J bΦ) V2(Jψ2 a2⊗β N o J bΦ) = (u b⊗α N = (u b⊗α N = (u b⊗α N = (u b⊗α N = (u b⊗α N = (u b⊗α N V2(Jψ2 a2⊗β N o J bΦ)V ∗ 1)V ∗ 1 ∗ J2,1 1)U a 1)U a 1)U a 1)U a ψ1(Jψ1 b⊗α N ψ1(Jψ1 b⊗α N ψ1(Jψ1 b⊗α N a2⊗β N o ψ1(u∗ 1 J ψ1 ∗ J2,1 V2(Jψ2 a2⊗β N o J bΦ) J bΦ)V ∗ 1 V1(u∗ b⊗α N 1)V ∗ 2 V2(Jψ2 a2⊗β N o J bΦ) J bΦ)(u∗ b⊗α N 1)(Jψ2 a2⊗β N o J bΦ) 1) from which we finish the proof of (i). Using the intertwining properties of u, (i) and ([E5] 5.2), we then get (ii). Using then (ii) and the properties of U a ψ ([E5] 13.4) recalled in 2.4, we get (iii). (cid:3) 3.2. Proposition. Let G be a measured quantum groupoid, and (b, a) an action of G on a von Neumann algebra A; let ψ be a δ-invariant weight on A, bearing the density condition, as defined in 2.3; then : (i) the unitary U a in 2.3. (ii) the dual weight satisfies ∆it ψ ψ constructed in 2.4 is equal to the implementation Vψ of a constructed (δ∆ bΦ)−it, where this last one-parameter group = ∆it ψ b⊗α N of unitaries had been defined in 2.3. Proof. Let ξ ∈ D(αHΦ, ν), x, y in Nψ ∩ N∗ ψ, a ∈ N T c ∩ N∗ T c ∩ N bΦc ∩ N∗ bΦc, such that Λ bΦc(a∗) belongs to the set cEτ introduced in ([E5]4.4). We have, using 2.4: Λ bΦc(a)) = (ρb,α ξ )∗S ψa(x∗)(Λψ(y) b⊗α ν ξ )∗a(y∗)(Λψ(x) b⊗α ν (ρb,α Λ bΦc(a∗)) and, as Λ bΦc(a∗) belongs to D(αHΦ, ν), thanks to ([E5]4.4) it is equal to : ωΛ bΦc (a∗),ξ)a(y∗)x) ωΛ bΦc (a∗),ξ)a(y∗)Λψ(x) = Λψ((id b∗α (id b∗α N N Let us suppose now that x is analytic with respect to ψ; as δ1/2Λ bΦc(a∗) belongs to D((HΦ)β, νo), thanks again to ([E5]4.4), we get, using ([E5] 8.4.(iii)), that it is equal to : JΨσψ −i/2(x∗)JψΛψ[(id b∗α N ωΛ bΦc (a∗),ξ)a(y∗)] = = JΨσψ −i/2(x∗)Jψ(id ∗ ωδ1/2Λ bΦc (a∗),ξ)(Vψ)Λψ(y∗) = (ρb,α ξ )∗(JΨσψ −i/2(x∗)Jψ b⊗α ν 1)Vψ(Λψ(y∗) a⊗β νo δ1/2Λ bΦc(a∗)) from which we get that : S ψa(x∗)(Λψ(y) b⊗α ν Λ bΦc(a)) = (JΨσψ −i/2(x∗)Jψ b⊗α ν 1)Vψ(Λψ(y∗) a⊗β νo δ1/2Λ bΦc(a∗)) 11 and, taking a bounded net xi strongly converging to 1, such that σψ verging to 1, and using the fact that S ψ is closed, we get : −i/2(x∗ i ) is also con- S ψ(Λψ(y) b⊗α ν Λ bΦc(a)) = Vψ[Jψ∆1/2 ψ Λψ(y) a⊗β νo J bΦ(δ∆ bΦ)−1/2Λ bΦc(a)] from which we deduce that : Vψ(Jψ b⊗α N J bΦ)(∆1/2 ψ b⊗α N δ∆ bΦ −1/2 ) ⊂ S ψ where ∆1/2 ψ b⊗α N unitaries ∆it ψ b⊗α N have, using 2.4 : (δ∆ bΦ)−1/2 is the infinitesimal generator of the one-parameter group of δ−it∆−it introduced in 2.3. But, on the other hand, for all t ∈ R, we bΦ (∆it ψ b⊗α N δ−it∆−it bΦ )S ψa(x∗)(Λψ(y) b⊗α ν Λ bΦc(a)) = = (∆it ψ b⊗α N δ−it∆−it bΦ )a(y∗)(Λψ(x) b⊗α ν Λ bΦc(a∗)) which, using 2.3, is equal to : a(σψ t (y∗))(Λψ(σψ t (x)) b⊗α ν S bΦcδ−it∆−it bΦ Λ bΦc(a)) = a(σψ t (y∗))(Λψ(σψ t (x)) b⊗α ν S bΦcδ−it∆−it bΦ Λ bΦc(a)) which is equal, using again 2.4, to : S ψa(σψ t (x∗))(Λψ(σψ t (y)) b⊗α ν δ−it∆−it bΦ Λ bΦc(a)) Taking again a family xi converging to 1, and using the closedness of S ψ, we get that : (∆it ψ b⊗α N δ−it∆−it bΦ )S ψ(Λψ(y) b⊗α ν Λ bΦc(a)) = S ψ(Λψ(σψ t (y)) b⊗α ν δ−it∆−it bΦ Λ bΦc(a)) = S ψ(∆it ψ b⊗α N δ−it∆−it bΦ )(Λψ(y) b⊗α ν Λ bΦc(a)) from which, using 2.4, we deduce that (∆it ψ b⊗α N δ−it∆−it bΦ )S ψ = S ψ(∆it ψ b⊗α N δ−it∆−it bΦ ) and, therefore, we have : Vψ(Jψ b⊗α N J bΦ)(∆1/2 ψ b⊗α N δ∆ bΦ −1/2 ) = S ψ and, by polar decomposition, we have : J ψ = Vψ(Jψ b⊗α N J bΦ) which, by definition of U a We also get : ψ, leads to (i). which leads to (ii). ∆1/2 ψ = ∆1/2 ψ b⊗α N δ∆ bΦ −1/2 12 (cid:3) 3.3. Corollary. Let G be a measured quantum groupoid, and (b, a) an action of G on a von Neumann algebra A; let us suppose that there exists on A a δ-invariant weight on A, bearing the density condition, as defined in 2.3; then, for any normal semi-finite faithful weight ψ on A, the unitary U a ψ constructed in 2.4 is a standard implementation of a as defined in 2.4. Proof. If ψ is a δ-invariant weight on A, bearing the density condition, as defined in 2.3, we have the result using 3.2; for another weight, using 3.1(iii), we get the result. (cid:3) 3.4. Corollary. Let G be a measured quantum groupoid, and (b, a) an action of G on a von Neumann algebra A; let us suppose that A is isomorphic to a crossed-product B ⋊bbGo where b is an action of bGo on a von Neumann algebra B, and that this isomorphism sends a on b. Then, for any normal semi-finite faithful weight ψ on A, the unitary U a ψ constructed in 2.4 is a standard implementation of a as defined in 2.4. Proof. We have recalled in 2.4 that any dual weight on B ⋊b bGo is a δ-invariant weight on B ⋊b bGo, bearing the density condition; therefore, using 3.3, we get the result. 3.5. Corollary. Let G be a measured quantum groupoid, and (b, a) an action of G on L(H), a von Neumann algebra A; let us consider the action (1 b⊗α N (cid:3) β, a) of G on A b∗α N L(H), the introduced in 2.4; then, for any normal semi-finite faithful weight ψ on A b∗α N unitary U a ψ is a standard implementation of the action a. Proof. This is just a corollary of 3.4 and of the biduality theorem, recalled in 2.4. (cid:3) 3.6. Corollary. Let G be a measured quantum groupoid, and (b, a) an action of G on a von Neumann algebra A; let ψ be a δ-invariant weight on A, bearing the density condition, as defined in 2.3; then, for any x ∈ cM ′, t ∈ R, we have : ψ t (1 b⊗α σ N x) = 1 b⊗α N ∆it Φx∆−it Φ Proof. Using 3.2(ii), we get that : ψ t (1 b⊗α σ N x) = 1 b⊗α N (δ∆ bΦ)−itx(δ∆ bΦ)it But, using ([E5]3.11(ii)), we know that (δ∆ bΦ)it = (δ∆Φ)−it; as δ is affiliated to cM , we get the result. (cid:3) 3.7. Corollary. Let G be a measured quantum groupoid, and (b, a) an action of G on a von Neumann algebra A; let ψ be a normal semi-finite faithful weight on A; then, for any x in M ′, t ∈ R, we have : ψ t (1 b⊗α σ N 1 α⊗β N o x) = 1 b⊗α N 1 α⊗β N o ∆−it bΦ x∆it bΦ Proof. Let's apply 3.6 to the dual action (1 b⊗α N ψ, and we get the result. 13 α, a) of Gc on A ⋊a G, and the dual weight (cid:3) 3.8. Corollary. Let G be a measured quantum groupoid, and (b, a) an action of G on a β, a) von Neumann algebra A; let ψ be a normal semi-finite faithful weight on A; let (1 b⊗α N be the action of G on A b∗α N L(H) obtained by transporting on A b∗α N L(H) the bidual action and ψa be the normal semi-finite faithful weight on A b∗α N the bidual weight. Then, for any x in M ′, t ∈ R, we have : L(H) obtained by transporting σψa t (1 b⊗α N x) = 1 b⊗α N ∆−it bΦ x∆it bΦ Proof. The canonical isomorphism between A b∗α N x (cf. [E5] 11.2). So, the result is a straightforward L(H) and (A ⋊a G) ⋊a bGc sends, for all x ∈ M ′, 1 b⊗α N x on 1 b⊗α N consequence of 3.7. 1 α⊗β N o (cid:3) 4. An auxilliary weight ψ. If b is a normal faithful non degenerate anti-homomorphism from N into a von Neumann algebra A, such that there exists a normal faithful semi-finite operator-valued weight T from A on b(N), we associate to the weight ψ = νo ◦ b−1 ◦ T a weight ψ on A b∗α L(H) N (4.4); we calculate its modular automorphism group (4.8), and the GNS representation of A b∗α N L(H) given by this weight (4.10). t (b(n)) = b(σν 4.1. Definitions. Let b be an injective ∗-antihomorphism from a von Neumann algebra N into a von Neumann algebra A; we shall then say that (N, b, A) (or simply A) is a faithful right von Neumann N-module. If there exists a normal semi-finite faithful operator-valued weight T from A onto b(N), we shall say that this faithful right N- module is weighted. Let then ψ be a normal faithful semi-finite weight on A; if, for all t in R, n in N, we have σψ −t(n)), then there exists a normal semi-finite faithful operator-valued weight T from A onto b(N) such that ψ = νo ◦ b−1 ◦ T; such a weight ψ on A will be said lifted from νo by T (or, simply, a lifted weight). If ψ is a normal semi-finite faithful weight on A, lifted from νo by T, then the weight ψ bears the density property introduced in ([E5], 8.1), recalled in 2.3. Namely, using ([E5], 2.2.1), one gets that D(aHψ, ν) ∩ D((Hψ)b, νo) contains all the vectors of the form Λψ(x), T ∩ Nψ ∩ N∗ where x ∈ NT ∩ N∗ ψ is analytical with respect to ψ, and such that, for any ψ; therefore D(aHψ, ν) ∩ D((Hψ)b, νo) is dense z ∈ C, σz(x) belongs to NT ∩ N∗ T ∩ Nψ ∩ N∗ in Hψ, which is the density property. If (b, a) is an action of a measured quantum goupoid G = (N, M, α, β, Γ, T, T ′, ν) on a von Neumann algebra A, we shall say that this action is weighted if the faithful right N-module (N, b, A) is weighted. 4.2. Lemma. Let (N, b, A) be a faithful weighted right von Neumann N-module, and let T be a normal semi-finite faithful operator-valued weight from A onto b(N); let α be a nomal faithful representation of N on a Hilbert space H and ν a normal semi-finite faithful weight on N; then, it is possible to define a canonical normal semi-finite faithful α(N)′ (which is equal to operator-valued weight (T b∗α N L(H) onto 1 b⊗α N id) from A b∗α N L(H), by ([E5], 2.4)), such that, if ψ denotes the weight on A lifted from νo by b(N) b∗α N 14 T, we get, for any X ∈ (A b∗α N L(H))+, that (T b∗α N id)(X) = 1 b⊗α N (ψ b∗α ν id)(X), where T b∗α N id and ψ b∗α ν id are slice maps introduced in [E4] and recalled in ([E5], 2.5). Proof. Let us represent A on a Hilbert space H; using Haagerup's theorem ([T], 4.24), there exists a canonical normal semi-finite faithful operator valued weight T−1 from b(N)′ onto A′; considering the representation of b(N)′ on H b⊗α H, and using again Haagerup's theorem, we obtain another normal semi-finite faithful operator-valued weight (T−1)−1 from the commutant of A′ on H b⊗α L(H)) onto the commutant ν H (which is A b∗α N ν of b(N)′ on H b⊗α H (which is b(N) b∗α N ν L(H)). As both T and (T−1)−1 are obtained by taking the commutants, within two different representations, of the same operator- valued weight T−1, a closer look at this construction leads ([EN], 10.2) to the fact that (T−1)−1 = (T b∗α id) is recalled in ([E5] N id). The link between (T b∗α N id) and (ψ b∗α ν 2.5). (cid:3) 4.3. Proposition. Let (N, b, A) be a von Neumann faithful right N-module, and let α be a normal faithful non degenerate representation of N on a Hilbert space H and ν a normal semi-finite faithful weight on N; then : (i) let's represent A on a Hilbert space K; the linear set generated by all operators on ξ2 )∗, with a in A and ξ1, ξ2 in D(αH, ν), is a ∗-algebra, K β⊗α L(H). H, of the form ρβ,α ξ1 a(ρβ,α ν which is weakly dense in A b∗α N (ii) let ψ be a normal faithful semi-finite weight on A, and let's represent A on Hψ; H, ψo) then, for any a in Nψ and ξ in D(αH, ν), Λψ(a) b⊗α ν 1 of Ao = JψAJψ), and we have (where we deal with the representation x 7→ x b⊗α N θψo(Λψ(a) b⊗α ν (iii) for all n ∈ N, let us define a(n) = Jψb(n∗)Jψ; let G = (N, M, α, β, Γ, T, T ′, ν) be a measured quantum groupoid; then, the representation of A b∗α H N ξ belongs to D(Hψ b⊗α L(H) on H β⊗a ξ aa∗(ρb,α ξ, Λψ(a) b⊗α ξ) = ρb,α Hψ b⊗α ξ )∗. ν ν ν ν x is standard, when we equip the Hilbert space with the antilinear defined by x 7→ 1 β⊗a N involutive isometry J defined, for any ξ, η in D(αH, ν), ζ in Hψ, by : J(J bΦη β⊗a ν ζ b⊗α ν ξ) = J bΦξ β⊗a ν Jψζ b⊗α ν η and with the closed cone P generated by all elements of the form J bΦξ β⊗a ξ is in D(αH, ν), and ζ ′ in the cone Pψ given by the Tomita-Takesaki theory associated to the weight ψ. (iv) let ϕ be a normal semi-finite faithful weight on A b∗α N ξ )∗ belongs to M+ dψo )1/2) if and only if ρb,α L(H); then Λψ(a) b⊗α ϕ , and then : ξ aa∗(ρb,α to D(( dϕ ξ belongs ξ, when b⊗α ν ν ν ζ ′ ϕ(ρb,α ξ aa∗(ρb,α ξ )∗) = k( dϕ dψo )1/2(Λψ(a) b⊗α ν ξ)k2 15 Moreover, if Λψ(a) b⊗α N ξ belongs to D(( dϕ dψo )1/2(Λψ(a) b⊗α N H, ϕ), and the canonical isomorphism between Hϕ and H β⊗a dψo )1/2), the vector ( dϕ to D(Hψ b⊗α ν sends Rϕ(( dϕ dψo )1/2(Λψ(a) b⊗α ν ξ))∗(ζ b⊗α ν η) on J bΦξ β⊗a ν JψaJψζ b⊗α ν η. ν ξ) belongs Hψ b⊗α ν H Proof. Using 2.1, we get, for a1, a2 in A, and ξ1, ξ2, ξ3, ξ4 in D(αH, ν), that : ρβ,α ξ1 a1(ρβ,α ξ2 )∗ρβ,α ξ3 a2(ρβ,α ξ4 )∗ = ρβ,α ξ1 a1b(< ξ3, ξ2 >α,ν)a2(ρβ,α ξ4 )∗ from which we see that this linear set is indeed an algebra; moreover, it is clear that it is invariant under taking the adjoint. Let's take c ∈ A′; we have : ρβ,α ξ1 a(ρβ,γ ξ2 )∗(c β⊗α N 1) = ρβ,α ξ1 ac(ρβ,α ξ2 )∗ = ρβ,α ξ1 ca(ρβ,α = (c β⊗α N ξ2 )∗ 1)ρβ,α ξ1 a(ρβ,α ξ2 )∗ from which we get that ρβ,α ξ1 a(ρβ,α ξ2 )∗ belongs to A β∗α N L(H). Let now X ∈ A β∗α N L(H), and let (ei)i∈I be a (α, ν)-orthogonal basis of H; we get that (id β∗α ωei,ej )(X) belongs to A, and we have, when we take the weak limits over the finite subsets J, J ′ of I : ν X = limJ,J ′ Xi∈J,j∈J ′ = limJ,J ′ Xi∈J,j∈J ′ (1 β⊗α N θα,ν(ei, ei))X(1 β⊗α N θα,ν(ej, ej)) ρβ,α ei (id β∗α ν ωei,ej )(X)(ρβ,α ej )∗ which proves (i). Let a ∈ Nψ, ξ ∈ D(αH, ν); then, for all x ∈ Nψ, we have : JψxJψΛψ(a) b⊗α ν ξ = aJψΛψ(x) b⊗α ξ ν ξ aJψΛψ(x) = ρb,α Therefore, Λψ(a) b⊗α we get that θψo(Λψ(a) b⊗α ν By [S](3.1), we know that A b∗α N ν ξ belongs to D(Hψ b⊗α ξ, Λψ(a) b⊗α ξ) = ρb,α ν H, ψo), and Rψo(Λψ(a) b⊗α ν ξ aa∗(ρb,α ξ )∗, which is (ii). ξ) = ρb,α ξ a. So, L(H) has a standard representation x 7→ x ⊗ψ 1 on the ν ν ν Hilbert space (Hψ b⊗α H) ⊗ψ (Hψ b⊗α H). Using then (ii) and [S](0.3.1), we get that this Hilbert space is isomorphic to H β⊗a ν Hψ b⊗α ν H, and that this isomorphism sends, for b ∈ Nψ, η ∈ D(αH, ν) : a) the vector (Λψ(a) b⊗α ν ξ) ⊗ψ (Λψ(b) b⊗α ν η) on J bΦη β⊗a ν JψbJψΛψ(a) b⊗α ν ξ, b) the standard representation x 7→ x ⊗ψ 1 on the representation x 7→ 1 β⊗a x, c) the antilinear involutive isometry which sends (Λψ(a) b⊗α ν (Λψ(b) b⊗α ν η) ⊗ψ (Λψ(a) b⊗α ν ξ) on J, N ξ) ⊗ψ (Λψ(b) b⊗α ν η) to d) the cone generated by all elements of the form (Λψ(a) b⊗α ν ξ) ⊗ψ (Λψ(a) b⊗α ν ξ) on P, 16 which gives (iii). Using (ii), we get that : ϕ(ρb,α ξ aa∗(ρb,α ξ )∗) = ϕ(θψo (Λψ(a) b⊗α ν ξ, Λψ(a) b⊗α ν ξ)) and, by definition of the spatial derivative, we know that, if Λψ(a) b⊗α N D(( dϕ dψo )1/2), we have : ξ belongs to ϕ(ρb,α ξ aa∗(ρb,α ξ )∗) = ϕ(θψo (Λψ(a) b⊗α ν ξ, Λψ(a) b⊗α ν ξ) = k( and, if Λψ(a) b⊗α ξ does not belong to D(( dϕ ξ )∗) = +∞. So, we have the first part of (iv). Then, the second part of (iv) is given by [S](3.2) and (iii). (cid:3) dψo )1/2), we know that ϕ(ρb,α ν dϕ dψo )1/2(Λψ(a) b⊗α ξ aa∗(ρb,α ν ξ)k2 4.4. Proposition. Let G = (N, M, α, β, Γ, T, T ′, ν) be a measured quantum groupoid; let (N, b, A) be a faithful weighted right von Neumann N-module, and let ψ be a normal semi-finite faithful weight on A lifted from νo in the sense of 4.1. Then : (i) it possible to define a one-parameter group of unitaries on ∆it on Hψ b⊗α ψ b⊗α ν ∆−it bΦ ν H, with natural values on elementary tensors. This one-parameter group of unitaries implements on A b∗α N L(H) the one-parameter group of automorphisms σψ Ad∆−it bΦ t b∗α N . (ii) there exists a normal semi-finite faithful weight ψ on A b∗α N L(H) such that the spatial derivative dψ dψo is equal to the generator ∆ψ b⊗α ν ∆−1 bΦ of the one-parameter group of unitaries ψ constructed in (i); the modular automorphism group σ t group σψ constructed in (i). t b∗α N Ad∆−it bΦ (iii) for any a in Nψ ∩ N∗ D(αH, ν), we have : ψ, and ξ ∈ D(αH, ν) ∩ D(∆−1/2 bΦ is equal to the automorphism ), such that ∆−1/2 bΦ ξ belongs to ψ(ρb,α ξ aa∗(ρb,α ξ )∗) = k∆1/2 ψ Λψ(a) b⊗α ν ∆−1/2 bΦ ξk2 Proof. If ξ ∈ D(αH, ν), we get, for all t ∈ R and n ∈ Nν : α(n)∆−it bΦ ξ = ∆−it bΦ = ∆−it bΦ = ∆−it bΦ bΦ t (α(n))ξ σ α(σν t (n))ξ Rα,ν(ξ)∆it ν Λν(n) from which we get that ∆−it bΦ ξ, ∆−it Therefore, we have < ∆−it bΦ bΦ ξk2 = (b(< ∆−it bΦ α,ν= σν ψη b⊗α ∆−it bΦ k∆it ξ >o ν ξ belongs to D(αH, ν), and Rα,ν(∆−it bΦ Rα,ν(ξ)∆it ν . α,ν). Taking now η ∈ Hψ, we get : ξ) = ∆−it bΦ −t(< ξ, ξ >o ξ >o α,ν)∆it ψη∆it ψη) α,ν))∆it α,ν))∆it ψη∆it ψη) ψη∆it ψη) ξ, ∆−it bΦ −t(< ξ, ξ >o t (b(< ξ, ξ >o = (b(σν = (σψ = (b(< ξ, ξ >o ξk2 = kη b⊗α α,ν)ηη) ν 17 from which we get the existence of the one-parameter group of unitaries. It is then easy to finish the proof of (i). As (∆it 1, we obtain ([T], 3.11) 1)(∆−it ) = Jψσψ ψ b⊗α t (x)Jψ b⊗α N that there exists a normal faithful semi-finite weight ψ on A b∗α N )(JψxJψ b⊗α N ψ b⊗α L(H) such that : ∆−it bΦ ∆it bΦ ν ν dψ dψo = ∆ψ b⊗α ν ∆−1 bΦ ψ Moreover, the modular automorphism group σ t tomorphism group σψ t b∗α N Ad∆−it bΦ is then equal to the one-parameter au- , constructed in (i), which finishes the proof of (ii). So, using now 4.3(iv) applied to ψ, we get that ρb,α ξ aa∗(ρb,α ξ )∗ belongs to M+ ψ if and only if Λψ(a) b⊗α ν ξ belongs to D(∆1/2 ψ b⊗α ν ∆−1/2 bΦ ), and then, we have : ψ(ρb,α ξ aa∗(ρb,α ξ )∗) = k∆1/2 ψ Λψ(a) b⊗α ν ∆−1/2 bΦ ξk2 from which we get (iii). (cid:3) 4.5. Corollary. Let G = (N, M, α, β, Γ, T, T ′, ν) be a measured quantum groupoid; let (N, b, A) be a faithful weighted right von Neumann N-module, ψ1 (resp. ψ2) be a normal faithful semi-finite weight on A lifted from νo and ψ1 (resp. ψ2) be the normal semi-finite faithful weight on A b∗α N L(H) constructed in 4.4(ii); then : (i) the cocycle (Dψ1 : Dψ2)t belongs to A ∩ b(N)′; (ii) we have : (Dψ1 : Dψ2)t = (Dψ1 : Dψ2)t b⊗α 1. N Proof. As ψ1 and ψ2 are lifted weights, (i) is well known ([T], 4.22. (iii)). Let (H, π, J, P) be a standard representation of the von Neumann algebra A; then Ao is represented on H by JAJ; for any normal semi-finite faithful weight ψ on A, we have dψo = ∆1/2 dψ ψ ; moreover, we have then : ( dψ1 dψo 1 )it(Dψo 1 : Dψo 2)t( dψo 2 dψ2 )it = ( )it( dψo 2 dψ1 )−it( dψo 2 dψ2 )it )it( dψ1 dψo 1 dψ1 dψo 2 dψo 1 dψ1 dψ2 dψo 2 = (Dψ1 : Dψ2)t )it( = ( )−it and, therefore (Dψo H b⊗α H : ν 1 : Dψo 2)t = ∆−it ψ1 (Dψ1 : Dψ2)t∆it ψ2. By similar arguments, we have on (Dψ1 : Dψ2)t = ( = ( dψ1 dψo 1 dψ1 dψo 1 )it( dψo 1 dψ2 )it )it(Dψo 1 : Dψo 2)t( dψ2 dψo 2 )−it As (Dψo 1 : Dψo 2)t belongs to JAJ b⊗α N 1H and is therefore equal to : ∆−it ψ1 (Dψ1 : Dψ2)t∆it ψ2 b⊗α N 1H 18 we obtain, using 4.4(ii), that (Dψ1 : Dψ2)t is equal to : (∆it ψ1 b⊗α N ∆−it bΦ )(∆−it ψ1 (Dψ1 : Dψ2)t∆it ψ2 b⊗α N 1H)(∆−it ψ2 b⊗α N ∆it bΦ) from which we get the result. (cid:3) 4.6. Remarks. Let us consider the trivial action (id, β) of G on N o ([E5], 6.2); it is clearly a weighted action (with the identity of N o as operator-valued weight); the crossed product is then cM ′, and the dual action is equal to bΓc ([E5], 9.4); the operator-valued weight from cM ′ onto β(N) is then T c, and, therefore, the dual weight νo of the weight νo on N o is the Haar weight bΦc; by the biduality theorem (2.3), we get that the crossed- product cM ′ ⋊bΓc bGc is isomorphic to N o ∗ L(H) = α(N)′; transporting the bidual weight νo on cM ′ ⋊bΓc bGc by this isomorphism, we obtain the weight νo thanks to 2.4, for all ξ ∈ D(αH, ν) ∩ D(∆−1/2 β on α(N)′, which verifies, ) : bΦ β(θα,ν(ξ, ξ)) = k∆−1/2 νo bΦ ξk2 and, for all t ∈ R, x ∈ α(N)′, σ On the other hand, for any y ∈ N T c ∩ N bΦc, z ∈ NT oc ∩ NΦoc, we have, by construction of νo β : . x∆it bΦ νo t (x) = ∆−it β bΦ νo β(y∗z∗zy) = cΦc(y∗T oc(z∗z)y) = kΛ bΦc(y) α⊗β νo ΛΦoc(z)k2 Let now (b, a) be an action of G on a von Neumann algebra A, and ψ a normal semi-finite faithful weight on A; by construction of ψa, we have, for any x ∈ Nψ : ψa(a(x∗)(1 b⊗α N y∗z∗zy)a(x)) = kΛψ(x) b⊗α ν Λ bΦc(y) α⊗β νo ΛΦoc(z)k2 and, by applying ([E5],13.3) to the weight νo, we get, for any X ∈ Nνo belongs to D(αHνo , ν) : β β such that Λνo β (X) ψa(a(x∗)(1 b⊗α N X ∗X)a(x)) = kΛψ(x) b⊗α ν (X)k2 Λνo β 4.7. Lemma. Let G = (N, M, α, β, Γ, T, T ′, ν) be a measured quantum groupoid; (N, b, A) be a faithful weighted right von Neumann right N-module; then : (i) if ξ, η are in D(αH, ν) ∩ D(∆−1/2 < ∆−1/2 (ii) there exists an (α, ν)-orthogonal basis of H such that, for all i ∈ I, ei belongs to D(αH, ν) ∩ D(∆−1/2 (iii) for any such basis, the weight νo ), such that ∆−1/2 −i/2(< ∆−1/2 β defined in 4.6 satisfies, for all x ∈ α(N)′+ : bΦ α,ν) =< ξ, ∆−1/2 ξ and ∆−1/2 ξ, η >o bΦ −i/2) and σν ei belongs to D(αH, ν); α,ν belongs to D(σν η belong to D(αH, ν), ) and ∆−1/2 ξ, η >o η >o α,ν. let bΦ bΦ bΦ bΦ bΦ bΦ νo β(x) =Xi (x∆−1/2 bΦ ei∆−1/2 bΦ ei) Proof. We get, for any n ∈ Nν, analytic with respect to ν : Rα,ν(∆−1/2 bΦ ξ)Λν(n) = α(n)∆−1/2 bΦ ξ = ∆−1/2 bΦ 19 α(σν −i/2(n))ξ = ∆−1/2 bΦ Rα,ν(ξ)∆1/2 ν Λν(n) and, using ([C2], 1.5) : Λν(< η, ∆−1/2 bΦ from which we get (i). ξ >o bΦ ν Λν(< ∆−1/2 α,ν) = Jν∆1/2 ν Rα,ν(∆−1/2 = Jν∆1/2 bΦ = JνRα,ν(ξ)∗∆−1/2 = JνΛν(< ξ, ∆−1/2 η bΦ η >α,ν) bΦ ξ, η >o α,ν) ξ)∗η Applying ([E3]2.10) to the inclusion α(N) ⊂ cM and the operator-valued weight T , we get that it is possible to construct an orthogonal (α, ν)-basis (ei)i∈I such that ei = J bΦΛ bΦ(xi), ei = J bΦΛ bΦ(x∗ with xi ∈ N bΦ ∩ N∗ i ) bΦ which belongs to D(αH, ν); which is (ii). Using (ii) and (i), we have : ; so, ei belongs to D(∆−1/2 ), and ∆−1/2 ∩ N T ∩ N∗ T bΦ bΦ ei) = kRα,ν(ξ)∗∆−1/2 = kΛν(< ei, ∆−1/2 bΦ = ν((Rα,ν(∆−1/2 bΦ eik2 ξ >o bΦ α,ν)k2 ξ)∗θα,ν(ei, ei)Rα,ν(∆−1/2 ξ)o) bΦ (θα,ν(ξ, ξ)∆−1/2 bΦ ei∆−1/2 bΦ and we get, using (i) and 4.6 : ei∆−1/2 (θα,ν(ξ, ξ)∆−1/2 Xi from which we get that Pi ω bΦ bΦ ei) = ν(< ∆−1/2 bΦ ξ, ∆−1/2 bΦ ξ >o α,ν) = k∆−1/2 bΦ ξk2 = νo β(θα,ν(ξ, ξ)) unicity of the spatial derivative, we get this weight is equal to νo β. (cid:3) is a normal semi-finite weight on α(N)′, and, by ei ∆ −1/2 bΦ 4.8. Theorem. Let G = (N, M, α, β, Γ, T, T ′, ν) be a measured quantum groupoid; let (ei)i∈I be an (α, ν)-orthogonal basis of H such that, for all i ∈ I, ei belongs to D(αH, ν) ∩ D(∆−1/2 ei belongs to D(αH, ν); let (N, b, A) be a faithful weighted right von Neumann right N-module, and let ψ be a normal semi-finite faithful weight on A lifted from νo; then, we have, with the notations of 4.4, 4.2 and 4.6 : ) and ∆−1/2 bΦ bΦ ψ =Xi ψ b∗α ν ω∆−1/2 bΦ ei = νo β ◦ (ψ b∗α ν id) Proof. Let X ∈ (A b∗α N L(H))+; we have : Xi ψ b∗α ν ω∆−1/2 bΦ ei (X) = Xi ν ◦ b−1 b∗α ν ω∆−1/2 bΦ ei (T b∗α N id)(X) = (ν ◦ b−1 b∗α ν νo β)(T b∗α N id)(X) = νo β(ν ◦ b−1 = νo β ◦ (ψ b∗α ν id)(T b∗α N b∗α ν id)(X) id)(X) which is the second equality, and proves therefore thatPi ψ b∗α ν semi-finite faithful weight on A b∗α N 20 L(H), which does not depend on the choice of the ω∆−1/2 bΦ ei defines a normal (α, ν)-orthogonal basis (ei)i∈I. Let us denote ψ0 that weight. We get : ψ0(ρb,α ξ aa∗(ρb,α ξ )∗) =Xi ψ(b(< ∆−1/2 bΦ ei, ξ >o α,ν)∗aa∗b(< ∆−1/2 bΦ ei, ξ >o α,ν)) Applying 4.7(i), if ξ belongs to D(αH, ν) ∩ D(∆−1/2 D(αH, ν), we get that b(< ∆−1/2 ei, ξ >o bΦ α,ν)∗) belongs to D(σψ ), and is such that ∆−1/2 −i/2) and that : bΦ bΦ ξ belongs to −i/2(b(< ∆−1/2 σψ bΦ ei, ξ >o α,ν)∗) = b(σν i/2(< ξ, ∆−1/2 bΦ ei >α,ν)o) = b(< ∆−1/2 bΦ ξ, ei >o α,ν) So, with such an hypothesis on ξ, and if a belongs to Nψ ∩ N∗ ψ, we get that : ψ0(ρb,α ξ aa∗(ρb,α ξ )∗) = Xi = Xi = Xi kJψb(< ei, ∆−1/2 bΦ ξ >o α,ν)JψΛψ(a∗)k2 kb(< ∆−1/2 bΦ ξ, ei >o α,ν)∆1/2 ψ Λψ(a)k2 (b(< ∆−1/2 bΦ ξ, ei >o α,ν< ei, ∆−1/2 bΦ ξ >o α,ν)∆1/2 ψ Λψ(a)∆1/2 ψ Λψ(a)) = (b(< ∆−1/2 = k∆1/2 bΦ ψ Λψ(a) b⊗α ξ, ∆−1/2 ξ >o ∆−1/2 α,ν)∆1/2 ξk2 bΦ bΦ ν ψ Λψ(a)∆1/2 ψ Λψ(a)) bΦ bΦ bΦ ξ ab(ρb,α ξ aa∗(ρb,α η )∗) = ψ0(ρb,α ξ )∗) = ψ0(ρb,α ), such that ∆−1/2 ξ )∗), for all a in Nψ ∩ N∗ ), and is such that ∆−1/2 ξ aa∗(ρb,α Using 4.4(iii), we get that ψ(ρb,α ψ and ξ ∈ D(αH, ν) ∩ D(∆−1/2 ξ belongs to D(αH, ν). By polarisation, we get ψ(ρb,α ψ and ξ, η in D(αH, ν) ∩ D(∆−1/2 η belong to D(αH, ν). The linear set generated by such elements is an involutive algebra, whose weak closure contains, using ([E5] 2.2.1) and the semi-finiteness of ψ, all operators of the form ρb,α ξ2 )∗, for any ξ1, ξ2 in D(αH, ν) ξ1 and c in A; therefore, using 4.3, we get that ψ and ψ0 are equal on a dense involutive algebra. We easily get that ψ0 ◦ σ η )∗), for all a, b in Nψ ∩ N∗ ξ ab(ρb,α ξ and ∆−1/2 t = ψ0 ◦ (σψ c(ρb,α ; the ω bΦ bΦ ψ t b∗α N Ad∆−it bΦ ) is equal to Pi ψ b∗α ν family ∆−it ei is another (α, ν)-orthogonal basis of H, which bears the same properties as bΦ (ei)i∈I. As, using (i), we know that the definition of ψ0 does not depend on the choice of ψ t , and, therefore ψ = ψ0, the orthogonal (α, ν)-basis, we get that ψ0 is invariant under σ which finishes the proof. (cid:3) ∆−1/2 ∆−it ei bΦ bΦ 4.9. Example. Looking again at the particular example given in 4.6, we get, using 4.8, that νo = νo β. bΦ ) and ∆−1/2 4.10. Theorem. Let G = (N, M, α, β, Γ, T, T ′, ν) be a measured quantum groupoid; let (ei)i∈I be an (α, ν)-orthogonal basis of H such that, for all i ∈ I, ei belongs to D(αH, ν) ∩ D(∆−1/2 ei belongs to D(αH, ν); let (N, b, A) be a faithful weighted right von Neumann right N-module, and let T be a normal semi-finite faithful operator-valued weight from A onto b(N); let us write ψ = νo ◦ b−1 ◦ T and ψ the normal semi-finite faith- L(H) constructed in 4.4; for n ∈ N, let us define a(n) = Jψb(n∗)Jψ. ful weight on A b∗α N Let ξ be in D(αH, ν) ∩ D(∆−1/2 ξ belongs to D(αH, ν); let η, ξ1 be in ), such that ∆−1/2 bΦ bΦ bΦ 21 D(αH, ν), and ξ2 ∈ D(Hβ, νo); let z be in Nψ, ζ be in Hψ, X be in A b∗α N (i) the operator ρb,α η z(ρb,α ξ )∗ belongs to Nψ, and we have : L(H). Then : Λψ(ρb,α η z(ρb,α ξ )∗) = J bΦ∆−1/2 bΦ ξ β⊗a ν Λψ(z) b⊗α ν η Moreover, the linear set generated by the operators ρb,α D(αH, ν), and ξ is in D(αH, ν) ∩ D(∆−1/2 core for Λψ. (ii) we have : Jψ(ξ2 β⊗a η z(ρb,α ), such that ∆−1/2 Jψζ b⊗α ζ b⊗α bΦ bΦ J bΦξ2. ξ1) = J bΦξ1 β⊗a X. ν ν (iii) we have : πψ(X) = 1 β⊗a ν ν N ξ )∗, where z is in Nψ, η is in ξ belongs to D(αH, ν), is a (iv) it is possible to define a one parameter group of unitaries ∆−it bΦ β⊗a ν H with natural values on elementary tensors, and ∆1/2 ψ on H β⊗a Hψ b⊗α ∆it ψ b⊗α ν ∆−it bΦ is equal to its ν ν generator ∆−1/2 bΦ β⊗a ν ∆1/2 ψ b⊗α ν ∆−1/2 bΦ . η z(ρb,α ξ )∗)∗ρb,α Proof. We have (ρb,α Mψ, by 4.4(iii). Let a in Nψ ∩ N∗ 4.3(iv) applied to the weight ψ, and 4.4(ii) : η z(ρb,α ξ )∗ = ρb,α ξ z∗b(< η, η >o α,ν)z(ρb,α ξ )∗, which belongs to ψ; let us take η1 satisfying the same hypothesis as ξ. We have, using J bΦη1 β⊗a ν JψaJψζ b⊗α ν and, therefore : η1 = Rψ(∆1/2 ψ Λψ(a) b⊗α ν ∆−1/2 bΦ η1)∗(ζ b⊗α ν ξ1) (Λψ(ρb,α η z(ρb,α ξ )∗)J bΦη1 β⊗a ν JψaJψζ b⊗α ν ξ1) = which, using 4.7(i), and the definition of ψ, is equal to : (ρb,α η z(ρb,α ξ )∗(∆1/2 ψ Λψ(a) b⊗α ν ∆−1/2 bΦ η1)ζ b⊗α ν ξ1) (ρb,α η zb(< ∆−1/2 bΦ η1, ξ >o α,ν)∆1/2 ψ Λψ(a)ζ b⊗α ν ξ1) = (zb(< ∆−1/2 bΦ η1, ξ >o α,ν)∆1/2 ψ Λψ(a) b⊗α ν ηζ b⊗α ν ξ1) = (zb(σν −i/2(< η1, ∆−1/2 bΦ ξ >o α,ν))∆1/2 ψ Λψ(a) b⊗α ν ηζ b⊗α ν ξ1) = Let us suppose that z belongs to D(σψ i/2); we get that : (z∆1/2 ψ b(< η1, ∆−1/2 bΦ ξ >o α,ν)Λψ(a) b⊗α ν ηζ b⊗α ν ξ1) z∆1/2 ψ b(< η1, ∆−1/2 bΦ ξ >o α,ν)Λψ(a) = ∆1/2 ψ σψ i/2(z)b(< η1, ∆−1/2 ξ >o α,ν)Λψ(a) bΦ ξ, η1 >o = JψΛψ((a∗b(< ∆−1/2 = Jψa∗b(< ∆−1/2 = Jψa∗Jψa(< η1, ∆−1/2 = Jψa∗Jψa(< J bΦ∆−1/2 bΦ bΦ bΦ 22 bΦ ξ, η1 >o α,ν)σ−i/2(z∗)) α,ν)JψΛψ(z) ξ >o α,ν)Λψ(z) ξ, J bΦη1 >β,νo)Λψ(z) which remains true for all z ∈ Nψ; therefore, we then get that : (Λψ(ρb,α η z(ρb,α ξ )∗)J bΦη1 β⊗a ν JψaJψζ b⊗α ν ξ1) = (Jψa∗Jψa(< J bΦ∆−1/2 bΦ ξ, J bΦη1 >β,νo)Λψ(z) b⊗α ν ηζ b⊗α ν ξ1) = (a(< J bΦ∆−1/2 bΦ ξ, J bΦη1 >β,νo)Λψ(z) b⊗α ν ηJψaJψζ b⊗α ν ξ1) = (J bΦ∆−1/2 bΦ ξ β⊗a ν Λψ(z) b⊗α ν ηJ bΦη1 β⊗a ν JψaJψζ b⊗α ν ξ1) from which, by density, we get the first result of (i). σψ t (z)(ρb,α Using 4.4(ii), we get that σ ∆−it ψ t (ρb,α η z(ρb,α ξ )∗) = ρb,α ∆−it η z(ρb,α ξ )∗, where z belongs to Nψ ∩ N∗ bΦ bΦ η ξ erated by the operators ρb,α in D(αH, ν) ∩ D(∆−1/2 bΦ subalgebra of Nψ ∩ N∗ ), such that ∆−1/2 ξ (resp. ∆−1/2 ψ t . It is possible to put on the image of this algebra under Λψ a structure of left-Hilbert algebra, L(H), equal to ψ which, in turn, leads to a faithful normal semi-finite weight ψ0 on A b∗α N bΦ L(H) by 4.3, and globally invariant under σ ψ, dense in A b∗α N bΦ )∗; so, the linear set gen- ψ, and ξ (resp. η) is η) belongs to D(αH, ν) is a ∗- on this subalgebra, and invariant under σ of (i). On the other hand, let's apply 4.3(iii) to the standard representation of A b∗α N ψ t . So, we get ψ0 = ψ, which finishes the proof L(H) given by the weight ψ, and we get (ii) and (iii). Let now ξ ∈ D(Hβ, νo); we have, for all t ∈ R, n ∈ Nν : β(n∗)∆−it bΦ ξ = ∆−it bΦ = ∆−it bΦ = ∆−it bΦ = ∆−it bΦ τt(β(n∗))ξ β(σν t (n∗))ξ Rβ,νo Rβ,νo (ξ)JνΛν(σν (ξ)Jν∆it t (n)) ν Λν(n) and, therefore, ∆−it bΦ < ∆−it bΦ we get : ξ, ∆−it bΦ ξ belongs to D(Hβ, νo), and Rβ,νo(∆−it bΦ ν , and −t(< ξ, ξ >β,νo). Therefore, if ξ ′ belongs to D(αH, ν), η ∈ Hψ, ξ) = ∆−it bΦ Rβ,νo(ξ)∆it ξ >β,νo= σν k∆−it bΦ ξ β⊗a ν ∆it ψη b⊗α ν ∆−it bΦ ξ ′k2 = (b(< ∆−it bΦ ξ ′, ∆−it bΦ ξ ′ >o α,ν)a(< ∆−it bΦ ξ, ∆−it bΦ ξ >β,νo)∆it ψη∆it ψη) −t(< ξ ′, ξ ′ >o t (b(< ξ ′, ξ ′ >o = (b(σνo = (σψ = (b(< ξ ′, ξ ′ >o = (b(< ξ ′, ξ ′ >o = (b(< ξ ′, ξ ′ >o = kξ β⊗a ν η b⊗α ν 23 α,ν))a(σν α,ν))Jψb(σν ψ Jψσψ −t(< ξ, ξ >β,νo))∆it ψη∆it −t(< ξ, ξ >β,νo))Jψ∆it t (b(< ξ, ξ >β,νo))Jψ∆it ψη) ψη∆it ψηη) ψη) α,ν)∆−it α,ν)Jψb(< ξ, ξ >β,νo)Jψηη) α,ν)a(< ξ, ξ >β,νo)ηη) ξ ′k2 Now, from (i) and (ii), we get that the infinitesimal generator ∆−1/2 bΦ β⊗a ν ∆1/2 ψ b⊗α ν ∆−1/2 bΦ of this one-parameter of unitaries is included in ∆1/2 we get (iv). ψ ; these operators being self-adjoint, (cid:3) 4.11. Corollary. Let G = (N, M, α, β, Γ, T, T ′, ν) be a measured quantum groupoid, (b, a) an action of G on a von Neumann algebra A, ψ a normal semi-finite faithful weight on A, ψa the normal semi-finite faithful weight constructed on A b∗α L(H) by transporting N the bidual weight. Then, for any x ∈ Nψ, ξ ∈ D(αH, ν), η ∈ D(αH, ν) ∩ D(∆−1/2 that ∆−1/2 ) such θα,ν(ξ, η))a(x) belongs to Nψa, and η belongs to D(Hβ, νo), the operator (1 b⊗α N bΦ bΦ we have: ψa(a(x∗)((1 b⊗α N θα,ν(ξ, η)∗θα,ν(ξ, η))a(x)) = kΛψ(x) b⊗α ν J bΦ∆−1/2 bΦ η β⊗α ξk2 ν Proof. Using 4.10 applied to νo, we get that Λνo(θα,ν(ξ, η)) = J bΦ∆−1/2 belongs to D(αHνo, ν); so, using 4.6 and 4.9, we get the result. bΦ η β⊗α ξ, which ν (cid:3) 5. Standard implementation: using the weight ψ. In that section, we calculate (5.5) the dual weight f(ψ) of ψ, with respect to the action a (5.2(ii)); this will allow us to calculate Jg(ψ) (5.5), and then, to obtain a formula linking ψ and U a U a ψ is a corepresentation (and, therefore, a standard implementation) whenever it is possible to construct ψ (5.8). ψ is a corepresentation by 3.5, we obtain then that U a ψ (5.6). As U a 5.1. Proposition. Let G = (N, M, α, β, Γ, T, T ′, ν) be a measured quantum groupoid; let A be a von Neumann algebra acting on a Hilbert space H, (b, a) be an action of G on A, (1 b⊗α L(H) introduced in 2.4; then, let us write, for N β, a) be the action of G on A b∗α N any Y in L(H b⊗α ν H β⊗α H), ν Θ(Y ) = (1 b⊗α N W )∗(id b∗α N ςN )(Y )(1 b⊗α N W ) which belongs to L(H b⊗α H β⊗α H); then, we have : (i) for any X ∈ A b∗α N ν ν L(H), Θ(a(X)) = (a b∗α N id)(X) and : Θ((A b∗α N L(H)) ⋊a G) = (A ⋊a G) β∗α N L(H) (ii) (1 b⊗α N α, (id b∗α N ςN )(a β∗α N id)) is an action of Gc on (A ⋊a G) β∗α N L(H), and : ( Θ α∗β N o id)f(a) = (id b∗α N ςN o)(a α∗β N o id) Θ where f(a) is the dual action of a (it is therefore an action of bGc on (A b∗α 24 N L(H)) ⋊a G). Proof. By the definition of a, we get the first formula of (i). The second formula of (i) was already proved in ([E5] 11.4). Moreover, using (i), we have, for all X ∈ A b∗α L(H) : N ( Θ α∗β N o id)f(a)(a(X)) = ( Θ α∗β N o = (a b∗α N 1) id)(a(X) α⊗β N o 1 id)(X) α⊗β N o ςN o)(a α∗β N o ςN o)(a α∗β N o = (id b∗α N = (id b∗α N id)(a b∗α N id)(X) id) Θ(a(X)) and, for all z ∈ cM ′, we have : ( Θ α∗β N o id)f(a)(1 b⊗α N 1 α⊗β N o z) = ( Θ α∗β N o id)(1 b⊗α N bΓc(z)) which, thanks again to ([E5] 11.4), is equal to : (1 b⊗α N 1 β⊗α N JΦJ bΦ β⊗α N 1)(bΓoc α∗β N o id)bΓc(z)(1 b⊗α N 1 β⊗α N J bΦJΦ α⊗β N o 1) and we have : (id b∗α N ςN o)(a α∗β N o id) Θ(1 b⊗α N 1 α⊗β N o z) = (id b∗α N ςN o)(bΓc β∗α N id)[(1 β⊗α N JΦJ bΦ)bΓoc(z)(1 β⊗α N J bΦJΦ)] from which we deduce that : ( Θ α∗β N o id)f(a)(1 b⊗α N 1 α⊗β N o z) = (id b∗α N ςN o)(a α∗β N o id) Θ(1 b⊗α N 1 α⊗β N o z) and we get that : ( Θ α∗β N o id)f(a) = (id b∗α N ςN o)(a α∗β N o id) Θ from which we deduce that (1 b⊗α N L(H), which finishes the proof. α, (id b∗α N ςN )(a β∗α N Neumann algebra (A ⋊a G) β∗α N id)) is an action of Go on the von (cid:3) 5.2. Corollary. Let G = (N, M, α, β, Γ, T, T ′, ν) be a measured quantum groupoid; let A be a von Neumann algebra acting on a Hilbert space H, (b, a) an action of G on A, L(H) introduced in 2.4 and Θ the isomorphism (1 b⊗α N β, a) be the action of G on A b∗α N L(H)) ⋊a G onto (A ⋊a G) β∗α N L(H); then, we have introduced in 5.1 which sends (A b∗α N Θ ◦ T f(a) = (Ta β∗α id) Θ. N 25 Proof. Using 5.1(ii), we get : Θ ◦ Ta = (id b∗α N = (id b∗α N id α∗β id α∗β N N bΦc)( Θ α∗β N bΦc)(id b∗α N id) Θ id)a ςN o)(a α∗β N o id) Θ = ((id α∗β = (Ta β∗α N N bΦc)a β∗α N id) Θ which is the result. (cid:3) 5.3. Theorem. Let G = (N, M, α, β, Γ, T, T ′, ν) be a measured quantum groupoid; let A be a von Neumann algebra acting on a Hilbert space H, (b, a) a weighted action of G on L(H) introduced in 2.4 and Θ the isomorphism A, (1 b⊗α N β, a) be the action of G on A b∗α N introduced in 5.1 which sends (A b∗α N L(H)) ⋊a G onto (A ⋊a G) β∗α N L(H); then : (i) (N, 1 b⊗α β, A ⋊a G) is a von Neumann faithful right N-module; let ψ be a lifted weight N on A, then ψ is a lifted weight on A ⋊a G. Let's denote then ψ and ( ψ) the weights constructed by 4.4 applied to ψ and ψ. (ii) we have ( ψ) ◦ Θ = f(ψ) and, for all t ∈ R, σ (iii) moreover, ψa is a lifted weight on A b∗α N faithful weight (ψa) on Ab∗α N L(H)β∗α N semi-finite faithful weight (ψ)a on A b∗α N L(H) β∗α N L(H), and we can define a normal semifinite L(H) . On the other hand, we can define the normal L(H). Then, we have (ψa) ◦ Θ = (ψ)a. ( ψ) t g(ψ) ◦ Θ = Θ ◦ σ t . Proof. Let T be a normal faithful semi-finite operator valued weight from A into b(N); then a ◦ T ◦ a−1 is a normal faithful semi-finite operator valued weight from a(A) into β(N), and a ◦ T ◦ a−1 ◦ Ta is a normal faithful semi-finite operator-valued weight 1 b⊗α N β(N); then, if we write ψ = νo ◦ b−1 ◦ T, the dual weight ψ can β)−1 ◦ (a ◦ T ◦ a−1 ◦ Ta), which finishes the proof of (i). from A ⋊a G into 1 b⊗α N be written as νo ◦ (1 b⊗α N We have then, using the notations of 4.4, and results 5.2 and 5.1(i) : ( ψ) ◦ Θ =Xi =Xi =Xi ( ψ β∗α ν ω∆−1/2 bΦ ei ) ◦ Θ (ψ ◦ a−1 ◦ Ta β∗α ν ω ∆ −1/2 bΦ ei ) ◦ Θ (ψ b∗α ν ω ∆−1/2 bΦ ei ) ◦ (a b∗α N id)−1 ◦ (Ta β∗α N id) ◦ Θ = ψ ◦ (a b∗α N id)−1 ◦ Θ ◦ Ta = ψ ◦ (a)−1 ◦ Ta = f(ψ) 26 which finishes the proof of (ii). We have : ψa = νo ◦ (1 b⊗α N β)−1 ◦ (a ◦ T ◦ a−1 ◦ Ta) ◦ Ta So, by composition of operator-valued weights, we get that ψa is a lifted weight on the β), and, applying 4.4, we can construct the faithful right N-module (N, A b∗α N L(H), 1 b⊗α N normal semi-finite faithful weight (ψa) on A b∗α N L(H) β∗α N L(H). On the other hand, as ψ is a normal semi-finite faithful weight on A b∗α N L(H), and as (1 b⊗α N β, a) (2.4) is an action of G on A b∗α N L(H), we can define (2.4) a weight (ψ)a L(H). As Θ is an isomorphism from A b∗α N L(H) β∗α N L(H), we can define then another normal semi-finite faithful weight L(H) β∗α N L(H) onto on A b∗α N L(H) β∗α N A b∗α N (ψa) ◦ Θ on A b∗α N L(H) β∗α N L(H). Let's represent A on Hψ and consider the isomorphism Θ from L(Hψ b⊗α ν H β⊗α H) onto ν L(Hψ b⊗α H β⊗α H). The commutant of A b∗α N L(H) β∗α N L(H) on the Hilbert space 1H, which is isomorphic to Ao. Let us consider the on Hψ b⊗α H β⊗α ν H. As, for x ∈ A′, Θ sends x b⊗α N 1H β⊗α N 1H ν Hψ b⊗α ν ν H β⊗α ν H is A′ ν spatial derivative 1H β⊗α N b⊗α N d(ψ)a ◦ Θ−1 dψo on x b⊗α N 1H β⊗α 1H, we get that : N d(ψ)a ◦ Θ−1 dψo = Θ( d(ψ)a dψo ) where the spatial derivative d(ψ)a dψo (using [St] 12.11), we get that : is taken on the Hilbert space Hψ b⊗α ν H β⊗α H. But, ν d(ψ)a dψo = df(ψ) ◦ Ta dψo = df(ψ) d ψo where we write, for simplification, ψo for the weight taken on (A b∗α N image by Θ is, thanks to (i), equal to (A ⋊a G)′ β⊗α 1H. L(H) ⋊a G)′, whose Therefore, using (ii), we get that : N Θ( d(ψ)a dψo ) = df(ψ) ◦ Θ−1 d ψo = d( ψ) d ψo = d(ψa) dψo which gives the result. (cid:3) 5.4. Lemma. Let G = (N, M, α, β, Γ, T, T ′, ν) be a measured quantum groupoid, W its pseudo-multiplicative unitary, (ei)i∈I an orthogonal (α, ν)-basis of H; then, we have, for all a ∈ N bΦc ∩ N T c, ζ ∈ D(αH, ν) ∩ D(H β, νo) : Xi Λ bΦc(ωJ bΦJΦζ,J bΦJΦei α∗β N o id)bΓc(a) β⊗α ν 27 ei = W ∗(Λ bΦc(a) α⊗ β νo ζ) Proof. Let us first remark that JΦJ bΦζ and JΦJ bΦei belong to D( αH, ν), and that Λ bΦc(a) belongs, thanks to ([E5] 2.2), to D(αH, ν). Applying then the definition ([E5] 3.6 (i)) of the pseudo-multiplicative unitary W c of the measured quantum group bGc, we get that : id)bΓc(a)) = (ωJ bΦJΦζ,J bΦJΦei ∗ id)(cW c∗)Λ bΦc(a) Λ bΦc((ωJ bΦJΦζ,J bΦJΦei α∗β N o As cW c∗ = (cW o)∗ = σW oσ, we get : and, using [E5] 3.12 (v) and 3.11(iii), we get : (ωJ bΦJΦζ,J bΦJΦei ∗ id)(cW c∗) = (id ∗ ωJ bΦJΦζ,J bΦJΦei)(W o) (id ∗ ωJ bΦJΦζ,J bΦJΦ ei)(W o) = J bΦ(id ∗ ωJΦζ,JΦei)(W )J bΦ = (id ∗ ωζ,ei)(W ∗) from which we get the result. (cid:3) 5.5. Proposition. Let G = (N, M, α, β, Γ, T, T ′, ν) be a measured quantum groupoid; let β, a) be the A be a von Neumann algebra, (b, a) a weighted action of G on A, (1 b⊗α N L(H) introduced in 2.4 and Θ the isomorphism introduced in 5.1 action of G on A b∗α N L(H)) ⋊a G onto (A ⋊a G) β∗α N which sends (A b∗α N L(H); let ψ be a lifted weight on A, and ψ be the normal semi-finite faithful weight on A b∗α N L(H)) ⋊a G; let ( ψ) be the normal semi-finite faithful weight on L(H) introduced in 4.4, and f(ψ) its dual weight on (A b∗α N L(H) introduced by applying 4.4 to the weight ψ on A ⋊a G. Then : (A ⋊a G) β∗α N (i) for any X ∈ Ng(ψ), Θ(X) belongs to N( ψ), and : Λ( ψ)( Θ(X)) = (1H β⊗a N 1Hψ b⊗α N W ∗σν)Λg(ψ)(X) (ii) we have : J( ψ)(1H β⊗a N 1Hψ b⊗α N W ∗σν) = (1H β⊗a N 1Hψ b⊗α N W ∗σν)Jg(ψ). Proof. The fact that Θ(X) belongs to N( ψ) is a straightforward corollary of 5.2(ii). Let us take x in Nψ, ξ in D(αH, ν) ∩ D(∆−1/2 ζ belongs to D(αH, ν), η in D(αH, ν), and a in N bΦc ∩ N T c. Then, by 4.10(i), we get that ρb,α ξ )∗ belongs to Nψ, ξ )∗) belongs to Ng(ψ). Moreover, we have, where and, by ([E5] 13.3), (1 b⊗α N η x(ρb,α (ei)i∈I is an orthogonal (α, ν)-basis of H : ), such that ∆−1/2 η x(ρb,α a)a(ρb,α bΦ bΦ Λ( ψ)( Θ((1 b⊗α N a)a(ρb,α η x(ρb,α ξ )∗)) = Λ( ψ)( Θ(1 b⊗α N a) Θa(ρb,α η x(ρb,α ξ )∗)) = Λ( ψ)( Θ(1 b⊗α N = Λ( ψ)( Θ(1 b⊗α a)(a b∗α N a)ρβ,α N η a(x)(ρβ,α ξ )∗) id)(ρb,α η x(ρb,α ξ )∗)) =Xi Λ( ψ)(ρβ,α ei (ρβ,α ei )∗ Θ(1 b⊗α N a)ρβ,α η a(x)(ρβ,α ξ )∗) Then, using 5.1, we get that Θ(1 b⊗α N a) = 1 b⊗α N therefore, that : (ρβ,α ei )∗ Θ(1 b⊗α N a)ρβ,α η = 1 b⊗α N 28 (id β∗ α N o J bΦJΦ), and, (1 β⊗α N o N o JΦJ bΦ)bΓoc(a)(1 β⊗α ωJ bΦJΦη,J bΦJΦei)bΓoc(a) and, we get then, applying 4.10(i) to the weight ψ, that : Λ( ψ)( Θ((1b⊗α N is equal to : Xi a)a(ρb,α η x(ρb,α ξ )∗)) =Xi Λ( ψ)(ρβ,α ei (1b⊗α N (idβ∗ α N o ωJ bΦJΦη,J bΦJΦei)bΓoc(a))a(x)(ρβ,α ξ )∗) J bΦ∆−1/2 bΦ ξ β⊗a ν Λ ψ((1 b⊗α N (id β∗ α N o ωJ bΦJΦη,J bΦJΦei)bΓoc(a))a(x)) β⊗α ν ei where, for all n ∈ N, we put a(n) = J ψ(1 b⊗α N a(n) = U a β(n∗))J ψ. We then get, by 2.4, that 1. And, therefore, using now ([E5] 13.3), we get α(n))(U a ψ(1 a⊗β N o ψ)∗ = a(n) b⊗α N η x(ρb,α ξ )∗)) is equal to : that Λ( ψ)( Θ((1 b⊗α a)a(ρb,α N Xi J bΦ∆−1/2 bΦ ξ β⊗a ν Λψ(x) b⊗α ν Λ bΦc((id β∗ α N o which, thanks to 5.4 is equal to : ωJ bΦJΦη,J bΦJΦei)bΓoc(a)) β⊗α ν ei (1H β⊗a N 1Hψ b⊗α N W ∗σν)(J bΦ∆−1/2 bΦ ξ b⊗a ν Λψ(x) b⊗α ν ζ β⊗α Λ bΦc(a)) ν which, using 4.10(i) again, is equal to : (1H β⊗a N 1Hψ b⊗α N and, by ([E5] 13.3) again, to : W ∗σν)(Λψ((ρb,α η x(ρb,α ξ )∗)) β⊗α Λ bΦc(a)) ν (1H β⊗a N 1Hψ b⊗α N W ∗σν)Λg(ψ)((1 b⊗α N a)a(ρb,α η x(ρb,α ξ )∗)) Using now 4.10(i), we get that, for any a in N bΦc ∩ N T c and Y in Nψ : Λ( ψ)( Θ((1 b⊗α N a)a(Y )) = (1H β⊗a N 1Hψ b⊗α N W ∗σν)Λg(ψ)((1 b⊗α N a)a(Y )) and, using now ([E5] 13.3), we finish the proof of (i). g(ψ) Let's suppose now that X is analytic with respect to f(ψ), such that σ −i/2(X ∗) belongs to Ng(ψ). Then, using 5.2(ii) and (i), we get that Θ(X) is analytic with respect to ( ψ), and that σ ( ψ) −i/2( Θ(X ∗)) belongs to N( ψ). More precisely, we then get : J( ψ)(1H β⊗a N 1Hψ b⊗α N W ∗σν)Λg(ψ)(X) = J( ψ)Λ( ψ)( Θ(X)) = Λ( ψ)(σ ( ψ) −i/2( Θ(X ∗))) = Λ( ψ)( Θ(σ g(ψ) −i/2(X ∗))) = (1H β⊗a N = (1H β⊗a N 1Hψ b⊗α N 1Hψ b⊗α N W ∗σν)Λg(ψ)(σ g(ψ) −i/2(X ∗))) W ∗σν)Jg(ψ)Λg(ψ)(X) which, by density, gives (ii). (cid:3) 29 5.6. Proposition. Let G = (N, M, α, β, Γ, T, T ′, ν) be a measured quantum groupoid; let β, a) be the A be a von Neumann algebra, (b, a) a weighted action of G on A, and (1 b⊗α N action of G on A b∗α N L(H) introduced in 2.4; let ψ be a lifted weight on A, and ψ be the L(H) introduced in 4.4. Then, the unitary U a ψ normal semi-finite faithful weight on A b∗α N satisfies : U a ψ = (1H β⊗a N 1Hψ b⊗α N σW σ)(1H β⊗α N (ida∗β N o ςN )(U a ψ a⊗β N o 1H))σβ,α 1 (W o∗ β⊗a N 1Hψ b⊗α N 1H)(σβ, α 1 )∗ where σβ,α 1 and σβ, α 1 is the flip from (H α⊗β νo H β⊗a is the flip from H β⊗α ν ν H) β⊗a Hψ b⊗α H onto H β⊗α ((Hψ b⊗α H), ν ν Hψ b⊗α ν H onto (H β⊗a ν ν Hψ b⊗α ν ν H) α⊗β νo H) a⊗β νo H. Proof. Let us recall (2.4) that (1 b⊗α N β, a) is an action of G on A b∗α N representation of N on Hψ defined, for all n ∈ N by : L(H). Let a be the a(n) = Jψπψ(1 b⊗α N β(n∗))Jψ Using 4.10 (iii) and (ii), we get that : β(n∗)J bΦ β⊗a a(n) = J bΦ N 1Hψ b⊗α N 1H = α(n) β⊗a N 1Hψ b⊗α N 1H and, therefore, U a ψ is a unitary from (H β⊗a Hψ b⊗α ν H) α⊗β νo H onto H β⊗a ν Hψ b⊗α ν H β⊗α H ν ν given by the formula : We have, using 5.5 : (1H β⊗a N 1Hψ b⊗α N U a ψ = Jg(ψ)(Jψ α⊗β νo J bΦ) W ∗σν)U a ψσβ, α 1 = J( ψ)(1H β⊗a N 1Hψ b⊗α N W ∗σν)(Jψ α⊗β νo J bΦ)σβ, α 1 Let ξ1, ξ2 in D(Hβ, νo), ξ3 in D(αH, ν), η ∈ Hψ; then J bΦξ1 belongs to D(αH, ν), and let us define ζi ∈ D(Hβ, νo) and ζ ′ i ∈ D(αH, ν) such that : W ∗(J bΦξ1 α⊗ β νo J bΦξ2) = limJXi∈J (ζi β⊗α ν ζ ′ i) the limit being taken on the filter of finite subsets J ⊂ I. Let us look at the image of the vector ξ1 β⊗α ξ3 under the unitary : ξ2 β⊗a η b⊗α ν ν ν J( ψ)(1H β⊗a N 1Hψ b⊗α N W ∗σν)(Jψ α⊗β νo J bΦ)σβ, α 1 This vector is first sent by σβ, α 1 on (ξ2 β⊗a η b⊗α J bΦξ3 β⊗a ν Jψη b⊗α ν J bΦξ2 β⊗α ν ν J bΦξ1, then (1H β⊗a N ν ξ3) α⊗β νo 1Hψ b⊗α N ξ1, then (Jψ α⊗β νo W ∗σν) sends it on J bΦ) sends it on J bΦξ3 β⊗a ν Jψη b⊗α ν W ∗(J bΦξ1 α⊗ β νo and J( ψ) sends it then on : J bΦξ2) = limJXi∈J (J bΦξ3 β⊗a ν Jψη b⊗α ν ζi β⊗α ν ζ ′ i) limJXi∈J (J bΦζ ′ i β⊗a ν J ψ(Jψη b⊗α ν ζi) β⊗α ν ξ3) = limJXi∈J 30 (J bΦζ ′ i β⊗a ν U a ψ(η a⊗β νo J bΦζi) β⊗α ν ξ3) which is equal to : (1H β⊗a N U a ψ b⊗α N 1H)(1H β⊗α N σν b⊗α N 1H)(σW o∗ β⊗a N 1Hψ b⊗α N 1H)(ξ1 β⊗α ξ2 β⊗a ν ν η b⊗α ν ξ3) from which, using again the density of finite sums of elementary tensors in the relative Hilbert tensor product, we get that : (1H β⊗a N 1Hψ b⊗α N W ∗σν)U a ψσβ, α 1 = (1H β⊗a N U a ψ b⊗α N 1H)(1H β⊗α 1H)(σW o∗ 1H) = σν b⊗α N ςN )(U a N (id a∗β N o β⊗a N 1Hψ b⊗α N (W o∗ ψ b⊗α N 1H))σβ,α 1 (1H β⊗a N from which we get the result. β⊗a N 1Hψ b⊗α N 1H) (cid:3) 5.7. Proposition. Let G = (N, M, α, β, Γ, T, T ′, ν) be a measured quantum groupoid; let A be a von Neumann algebra, and let (b, a) be a weighted action of G on A; let ψ be a lifted weight on A; then, the unitary U a ψ introduced in 2.4 is a copresentation of G. Proof. With the notations of 5.6, we get, using 5.6, that : 1H β⊗a N (id a∗β N o ςN )(U a ψ b⊗α N 1H) = (1H β⊗a N 1Hψ b⊗α N σW ∗σ)U a ψσβ, α 1 (W o β⊗a N 1Hψ b⊗α N 1H)(σβ,α 1 )∗ which we shall write, for simplification, with the usual leg numbering notation : ψ is a corepresentation of G by 3.5, and ψ(W o)4,1 (U a ψ)2,4 = cW3,4U a But cW is a corepresentation of Go ([E5], 5.6), U a ψ)2,4,5 = cW3,5cW3,4(U a (id ∗ Γ)(U a σW oσ is a corepresentation of Go by ([E5], 5.6 and 5.3). So, we get : ψ)1,2,3,4(U a ψ)1,2,3,5W o 5,1W o 4,1 ψ)1,2,3,5W o ψ)1,2,3,5W o 5,1W o 4,1 5,1W o 4,1 ψ)2,4W o∗ 4,1(U a = cW3,5(U a = (U a = (U a = (U a ψ)2,4W o∗ ψ)2,4W o∗ ψ)2,4(U a 4,1cW3,5(U a ψ)2,5W o 4,1 4,1(U a ψ)2,5 which shows that U a we leave to the reader. ψ is a corepresentation. A more complete proof is a painful exercise (cid:3) 5.8. Theorem. Let G = (N, M, α, β, Γ, T, T ′, ν) be a measured quantum groupoid; let A be a von Neumann algebra, and let (b, a) be a weighted action of G on A; then, for any normal semi-finite faithful weight ψ on A, the unitary U a ψ introduced in 2.4 is a standard implementation of a, in the sense of 2.3. Proof. As the action is weighted, there exists a normal semi-finite faithful weight ψ on A which is lifted from νo; we get that U a ψ is a corepresentation by 5.7, and is therefore a standard implementation. Using now 3.1, we easily get that it remains true for any normal semi-finite faithful weight on A, which is the result. (cid:3) 31 5.9. Remark. In 5.8, we had obtained that U a ψ is a standard implementation of a, if there exists a normal-semi-finite faithful operator-valued weight from A onto b(N); this is true in particular in the following cases : (i) G is a locally compact quantum group (N = C); this result was obtained in ([V1] 4.4); (ii) if N is abelian and b(N) ⊂ Z(A); in particular, if G is a measured groupoid; we shall discuss this particular case in 5.10. More general, if G is a continuous field of locally compact quantum groups (2.5 (iv)), or is De Commer's example (2.5 (v)). (iii) A is a type I factor; if we write A = L(H), starting from any normal semi-finite weight on b(N)′, we get a normal faithful semi-finite operator-valued weight from A to b(N). More generally, this remains true if A is a sum of type I factors; (iv) N is a sum of type I factors (in particular, if N is a finite dimensional algebra, which is the case, in particular if G is a finite dimensional quantum groupoid); (v) N and A are semi-finite. In 3.2, the result was proved if a is a dual action. 5.10. Example. Let G be a measured groupoid, with G(0) as its set of units, r and s its range and source application, (λu)u∈G(0) its Haar system, and ν a quasi-invariant measure; let µ = RG(0) λudν ; let us consider the von Neumann algebra L∞(G, µ), which is a L∞(G(0))-bimodule, thanks to the two homomorphisms rG and sG defined, for f in L∞(G(0)) by rG(f ) = f ◦ r and sG(f ) = f ◦ s. We have shown in ([E5], 3.1, 3.4 and 3.17) how it is possible to put a measured quantum groupoid structure on this von Neumann bimodule. An action (b, a) of this measured quantum groupoid on a von Neumann algebra A verifies G(0) Axdν(x) G(0) Axdν(x). Then ψ is a lifted weight; more precisely, there exists a measurable field ψx of normal semi- G(0) ψxdν(x) in the sense of ([T] 4.6), and Hψ = that b(L∞(G(0))) ⊂ Z(A), and, therefore, A can be decomposed as A = R ⊕ ([E5], 6.1); moreover, let ψ be a normal semi-finite faithful on A = R ⊕ finite faithful weights, such that ψ = R ⊕ R ⊕ G(0) Hψxdν(x). On the other hand, the action a is ([E5], 6.3) an action of G in the sense of ([Y3], 3.1), i.e. for all g ∈ G, there exists a family of ∗-isomorphisms ag from As(g) onto Ar(g), such that, if (g1, g2) ∈ G(2), we have ag1g2 = ag1ag2, and such that, for any normal positive functional G(0) yxdν(x), the function g 7→ ωr(g)(ag(ys(g))) is µ- measurable. These ∗-isomorphisms have standard implementations ug : Hψs(g) → Hψr(g) such that ag(ys(g)) = ugys(g)u∗ More precisely, the Hilbert space Hψ b⊗rG G(0) ωxdν(x), and any y = R ⊕ g. if (g1, g2) ∈ G(2), we have ug1g2 = ug1ug2. ω = R ⊕ L2(G, µ) can be identified with R ⊕ ν G Hψr(g)dµ(g). We then get : a(Z ⊕ G(0) yxdν(x)) =Z ⊕ G ag(ys(g))dµ(g) In [Y1] and [Y2] is given a construction of the crossed product of A by G; using ([Y3] 2.14), we see ([E5], 9.2) that this crossed-product is isomorphic to the definition given in ([E5], 9.1). Moreover, we get the same notion of dual action ([E5], 9.6) and of dual weight ([E5], 13.1). As b is central, we have a = b, and the Hilbert space Hψ a⊗sG L2(G, µ) can be identified with R ⊕ unitary from R ⊕ νo G Hψs(g)dµ(g). Using then [Y3], 2.6, we get that U a G Hψs(g)dµ(g) onto R ⊕ G Hψr(g)dµ(g). 32 ψ = R ⊕ G ugdµ(g), which is a 6. The (b, γ) property for weights If G = (N, M, α, β, Γ, T, T ′, ν) be a measured quantum groupoid, and if b is a normal faithful non degenerate anti-homomorphism from N into a von Neumann algebra A, we define the (b, γ) property for normal faithful semi-finite weights on A (6.1). We define then, for such a weight, a normal semi-finite faithful weight ψδ on A b∗α L(H) (6.4). We N obtain then several technical results (6.6, 6.7, 6.8) which will be used in chapter 7. 6.1. Definition. Let G = (N, M, α, β, Γ, T, T ′, ν) be a measured quantum groupoid, and let b be a normal faithful non degenerate anti-homomorphism from N into a von Neumann algebra A; we shall say that a normal faithful semi-finite weight ψ on A satisfies the (b, γ) property if, for all n ∈ N and t ∈ R, we have σψ t (b(n)) = b(γt(n)), where γt is the one-parameter automorphism group of N defined by σT t (β(n)) = β(γt(n)) ([E5], 3.8 (v)). 6.2. Example. Let G = (N, M, α, β, Γ, T, T ′, ν) be a measured quantum groupoid, A a von Neumann algebra, (b, a) an action of G on A, ψ a δ-invariant normal faithful semi- finite weight on A bearing the density property, as defined in ([E5]) and recalled in 2.3. Then, ψ satisfies the (b, γ) property. Namely, for any x ∈ A, t ∈ R, we have : a(σψ t (x)) = (∆it ψ b⊗α ν δ−it∆−it bΦ )a(x)(∆−it ψ b⊗α ν δit∆it bΦ) and, therefore, for any n ∈ N, we get, using ([E5], 3.8(ii)) : a(σψ t (b(n))) = (∆it ψ b⊗α β(n))(∆−it ψ b⊗α ν δit∆it bΦ) δ−it∆−it )(1 b⊗α bΦ N β(n)δit∆it bΦ ν δ−it∆−it bΦ t σΦ◦R σΦ −t τ−t(β(n)) σΦ t (β(n)) β(γt(n)) = 1 b⊗α N = 1 b⊗α N = 1 b⊗α N = 1 b⊗α N from which we get the property, by the injectivity of a. = a(b(γt(n))) 6.3. Example. Let G = (N, M, α, β, Γ, T, T ′, ν) be a measured quantum groupoid, and let (N, b, A) be a faithful weighted right von Neumann right-module, in the sense of 4.1; let ψ be a normal faithful semi-finite weight on A, lifted from νo, and let ψ be the normal β, γ) faithful semi-finite weight on A b∗α N property. Namely, using 4.8 and ([E5] 3.10 (vii)), we get : L(H) defined in 4.4. Then, ψ satisfies the (1 b⊗α N ψ t (1 b⊗α σ N β(n)) = 1 b⊗α N ∆−it bΦ β(n)∆it bΦ = 1 b⊗α N bΦ −t( β(n)) = 1 b⊗α σ N β(γ−t(n)) = 1 b⊗α N β(γt(n)) 6.4. Theorem. Let G = (N, M, α, β, Γ, T, T ′, ν) be a measured quantum groupoid, and let b be a normal faithful non degenerate anti-homomorphism from N into a von Neumann algebra A; let ψ be a normal faithful semi-finite weight on A satisfying the (b, γ) property; then : (i) it is possible to define a one-parameter group of unitaries ∆it (δ∆ bΦ)−it on Hψ b⊗α ψ b⊗α H, ν ν 33 with natural values on elementary tensors. We shall denote ∆1/2 generator. (ii) there exists a normal semi-finite faithful weight ψ ν ψ b⊗α (δ∆ bΦ)−1/2 its analytic on A b∗α N δ L(H) such that : dψ δ dψo = ∆1/2 ψ b⊗α ν (δ∆ bΦ)−1/2 (iii) for any a in Nψ ∩ N∗ belongs to D(αH, ν), we have : ψ, and ξ ∈ D(αH, ν) ∩ D((δ∆ bΦ)−1/2), such that (δ∆ bΦ)−1/2ξ ψ δ (ρb,α ξ aa∗(ρb,α ξ )∗) = k∆1/2 ψ Λψ(a) b⊗α ν (δ∆ bΦ)−1/2ξk2 Proof. Let η ∈ D(αH, ν), n ∈ Nν; then, we get : α(n)(δ∆ bΦ)−itη = (δ∆ bΦ)−itσ bΦ t σΦ◦R σΦ −t(α(n))η t t γ−tτ−t(n))η = (δ∆ bΦ)−itα(σν = (δ∆ bΦ)−itRα,ν(η)Λν(γ−t(n)) There exists a positive self-adjoint non singular operator h on Hν such that : Λν(γt(n)) = hitΛν(n) We then get that : α(n)(δ∆ bΦ)−itη = (δ∆ bΦ)−itRα,ν(η)h−itΛν(n) from which we get that (δ∆ bΦ)−itη belongs to D(αH, ν), and that : Rα,ν((δ∆ bΦ)−itη) = (δ∆ bΦ)−itRα,ν(η)h−it from which we get that : < (δ∆ bΦ)−itη, (δ∆ bΦ)−itη >α,ν= hit < η, η >α,ν h−it As we have, for all m ∈ N, γt(m) = hitmh−it, we therefore get that : < (δ∆ bΦ)−itη, (δ∆ bΦ)−itη >o α,ν= γt(< η, η >o α,ν) and, therefore, for all ξ ∈ Hψ : k∆it ψξ b⊗α ν (δ∆ bΦ)−itηk2 = (b(γt(< η, η >o α,ν))∆it ψξ∆it ψξ) t (b(< η, η >o α,ν))∆it ψξ∆it ψξ) = (σψ = kξ b⊗α ν ηk2 which is (i). As (∆it ψ b⊗α ν (δ∆ bΦ)−it)(JψxJψ b⊗α N 1)(∆it ψ b⊗α ν (δ∆ bΦ)−it) = Jψσψ t (x)Jψ b⊗α N 1, we get (ii). Result (iii) is just a corollary of (ii) and 4.3(iv). (cid:3) 6.5. Corollary. Let G be a measured quantum groupoid, and (b, a) an action of G on a von Neumann algebra A; let ψ be a δ-invariant weight on A, bearing the density condition, as defined in 2.3, and ψa the weight constructed on A b∗α L(H) by transporting the bidual N weight (2.4) of ψ. Using 6.2, we can use 6.4 and define the weight ψ on A b∗α N δ L(H) Then, we have : ψa = ψ . δ 34 Proof. We have, in general, dψa result. dψo = ∆1/2 ψ (2.4). So, using 3.2(ii) and 6.4, we get the (cid:3) 6.6. Corollary. Let G = (N, M, α, β, Γ, T, T ′, ν) be a measured quantum groupoid, and let b be a normal faithful non degenerate anti-homomorphism from N into a von Neumann algebra A; let ψ1 (resp. ψ2) be a normal faithful semi-finite weight on A satisfying the (b, γ) property; then : (i) the cocycle (Dψ1 : Dψ2)t belongs to A ∩ b(N)′; (ii) we have : (Dψ1δ )t = (Dψ1 : Dψ2)t b⊗α N : Dψ2δ 1. Proof. For any x ∈ A, we have : and, therefore : t (x) = (Dψ1 : Dψ2)tσψ2 σψ1 t (x)(Dψ1 : Dψ2)∗ t t ◦ σψ2 σψ1 −t(x) = (Dψ1 : Dψ2)tx(Dψ1 : Dψ2)∗ t In particular, we get, for any n ∈ N : b(n) = (Dψ1 : Dψ2)tb(n)(Dψ1 : Dψ2)∗ t from which we get (i). Let (H, π, J, P) be a standard representation of the von Neumann algebra A; then Ao is represented on H by JAJ; for any normal semi-finite faithful weight ψ on A, we have dψ ψ ; moreover, we have then : dψo = ∆1/2 ( dψ1 dψo 1 )it(Dψo 1 : Dψo 2)t( dψo 2 dψ2 )it = ( )it( dψo 2 dψ1 )−it( dψo 2 dψ2 )it )it( dψ1 dψo 1 dψ1 dψo 2 dψo 1 dψ1 dψ2 dψo 2 = (Dψ1 : Dψ2)t )it( = ( )−it and, therefore (Dψo H b⊗α H : ν 1 : Dψo 2)t = ∆−it ψ1 (Dψ1 : Dψ2)t∆it ψ2. By similar arguments, we have on (Dψ1δ : Dψ2δ )t = ( = ( dψ1δ dψo 1 dψ1δ dψo 1 )it( dψo 1 dψ2δ )it )it(Dψo 1 : Dψo 2)t( dψ2δ dψo 2 )−it As (Dψo 1 : Dψo 2)t belongs to JAJ b⊗α N 1H and is therefore equal to : ∆−it ψ1 (Dψ1 : Dψ2)t∆it ψ2 b⊗α N 1H we obtain, using 6.4(ii), that (Dψ1δ : Dψ2δ )t is equal to : (∆it ψ1 b⊗α ν (δ∆ bΦ)−it)(∆−it ψ1 (Dψ1 : Dψ2)t∆it ψ2 b⊗α N 1H)(∆−it ψ2 b⊗α ν (δ∆ bΦ)it) from which we get the result. (cid:3) 35 6.7. Proposition. Let G = (N, M, α, β, Γ, T, T ′, ν) be a measured quantum groupoid; it ∆it is possible to define one parameter groups of unitaries ∆it , bΦ and (δ∆ bΦ)it bΦ β⊗α ∆it bΦ ν α⊗ β νo with natural values on elementary tensors, and we have : W (∆it bΦ β⊗α ν ∆it bΦ)W ∗ = (δ∆ bΦ)it ∆it bΦ α⊗ β νo Proof. From ([E5] 3.10 (vi)), we get that ∆ bΦ is the closure of P JΦδ−1JΦ, where P is the managing operator of the pseudo-multiplicative unitary W , and δ the modulus of G; in ([E5] 3.8 (vii)), we had got that it is possible to define one parameter groups of unitaries P it P it, with natural values on elementary tensors, and that : P it and P it β⊗α ν α⊗ β νo W (P it β⊗α P it) = (P it ν P it)W α⊗ β νo On the other hand, it is possible ([E5], 3.8 (vi)) to define a one parameter group of unitaries δit δit, with natural values on elementary tensors, and that : β⊗α ν δit β⊗α ν δit = Γ(δit) = W ∗(1 α⊗ β N o δit)W Moreover, we know, from ([E5], 3.11 (iii)), that : W (J bΦ α⊗ β νo JΦ) = (J bΦ α⊗ β νo JΦ)W ∗ and from ([E5] 3.8 (vi)) that J bΦδ−itJ bΦ = R(δit) = δ−it. With all these data, we get that it is possible to define ∆it ∆it bΦ β⊗α ν ∆it bΦ = (P it β⊗α ν P it)(JΦδitJΦ β⊗α N ∆it bΦ as : bΦ β⊗α ∆it bΦ as : ν JΦδitJΦ) and (δ∆ bΦ)it α⊗ β νo α⊗ β νo (δ∆ bΦ)it ∆it bΦ = (P it α⊗ β νo P it)(J bΦ β⊗α ν JΦ)(δit β⊗α ν δit)(J bΦ α⊗ β νo JΦ)(JΦδitJΦ α⊗ β N o 1) and to verify that : bΦ β⊗α W (∆it ν ∆it bΦ)W ∗ = W (P it β⊗α P it)(JΦδitJΦ β⊗α ν N P it)W (JΦδitJΦ β⊗α N JΦδitJΦ)W ∗ JΦδitJΦ)W ∗ P it)(JΦδitJΦ α⊗ β N o 1)W (1 β⊗α JΦδitJΦ)W ∗ N = (P it = (P it α⊗ β νo α⊗ β νo which is equal to : (P it α⊗ β νo P it)(JΦδitJΦ α⊗ β N o 1)(J bΦ β⊗α ν JΦ)W ∗(1 α⊗ β N o δit)W (J bΦ α⊗ β νo JΦ) and, therefore, to : (P it α⊗ β νo P it)(JΦδitJΦ α⊗ β N o 1)(J bΦ β⊗α ν JΦ)(δit β⊗α ν δit)(J bΦ α⊗ β νo JΦ) or to : (P it α⊗ β νo P it)(JΦδitJΦ α⊗ β N o 1)(δit α⊗ β νo J bΦδitJ bΦ) = (δ∆ bΦ)it ∆it bΦ α⊗ β νo which finishes the proof. (cid:3) 36 6.8. Proposition. Let G = (N, M, α, β, Γ, T, T ′, ν) be a measured quantum groupoid, (b, a) a weighted action of G on a von Neumann algebra A, and ψ a normal semi-finite faithful weight on A, lifted from νo; then the von Neumann algebra A b∗α L(H) is a faithful N right N-module in two different ways, using 1 b⊗α N ψ constructed in 4.4 is a lifted weight from ν, using 1 b⊗α N β, and 1 b⊗α N β, and, on the other hand, β; moreover, the weight β, γ) property ; therefore, we can define a normal semi-finite faithful satisfies the (1 b⊗α N weight ψ on A b∗α N L(H) β∗α N on A b∗α N L(H) β∗α N L(H), and another normal semi-finite faithful weight (ψ) δ L(H). As in 5.1, let us write, for any Y in L(H b⊗α ν H β⊗α H), ν which belongs to L(H b⊗α ν Θ(Y ) = (1 b⊗α N H). Then, we have : W )∗(id b∗α N H β⊗α ςN )(Y )(1 b⊗α N W ) ν ψ ◦ Θ = (ψ) δ Proof. By definition, the weight ψ is defined on A b∗α N Hψ β⊗α H the spatial derivative : ν L(H) β∗α N L(H) by considering on and, using 4.10, we therefore get, on H β⊗a ν H β⊗α H, that : ν ν dψ d(ψ)o = ∆ψ β⊗α Hψ b⊗α ν ∆−1 bΦ dψ d(ψ)o = ∆−1 bΦ β⊗a ν ∆ψ b⊗α ν ∆−1 bΦ β⊗α ν ∆−1 bΦ On the other hand, the weight (ψ) Hψ β⊗α ν H = H β⊗a ν Hψ b⊗α ν δ H β⊗α ν is defined on A b∗α N L(H) β∗α N H the spatial derivative : L(H) by considering on d(ψ) d(ψ)o = ∆ψ β⊗α δ ν (δ∆ bΦ)−1 = ∆−1 bΦ β⊗a ν ∆ψ b⊗α ν ∆−1 bΦ β⊗α (δ∆ bΦ)−1 ν from which we get, using 6.7 and the definition of Θ that : dψ d(ψ)o = (id β∗a N Θ)( d(ψ) d(ψ)o ) δ The weight (ψ)o is defined on Jψπψ(A b∗α N L(H))Jψ, which, using again 4.10, is equal Θ) sends to L(H) β∗a N A′ b⊗α N 1H; we see, therefore, for X ∈ L(H) β∗a N A′, that (id β∗a N 1H, and leaves (ψ)o invariant. From which we deduce 1H β⊗α N 1H on X b⊗α N 1H β⊗α N X b⊗α N that : from which we get the result. dψ ◦ Θ d(ψ)o = d(ψ) δ d(ψ)o 37 (cid:3) 7. Biduality of weights In that chapter, following what had been done for locally compact quantum groups in [Y4], [Y5], and [BV], starting from an action a of a measured quantum groupoid on a von Neumann algebra A, we define the Radon-Nikodym derivative of a lifted weight on A with respect to this action (7.2); this operator is an a-cocycle (7.3), which measures, in a certain sense, how the weight ψ behaves towards the action. In particular, we prove that this cocycle is equal to 1 if and only if the weight is invariant by the action (7.7, 7.9). 7.1. Theorem. Let G = (N, M, α, β, Γ, T, T ′, ν) be a measured quantum groupoid, (b, a) a weighted action of G on a von Neumann algebra A, ψ a normal semi-finite faithful weight on A lifted from νo; let ψ be the dual weight on the crossed-product A ⋊a G, and let ψa be the normal semi-finite faithful weight on A b∗α L(H) obtained from the bidual N weight ψ and the isomorphism between A b∗α be normal semi-finite faithful weight on A b∗α N L(H) and the double crossed-product; let ψ L(H) constructed in 4.4. We have then : N (Dψa : Dψ)t = ∆it ψ(∆−it ψ b⊗α ∆it bΦ) ν Moreover, the unitaries ∆it ψ (∆−it ψ b⊗α ν ∆it bΦ ) belong to A b∗α N (M ∩ β(N)′). Proof. We have (Dψa : Dψ)t = ( dψa 2.4 and 4.8. So, we get that the unitaries ∆it dψo )it( dψo )−it, from which we get the first result, by L(H); let's dψ ) belong to A b∗α N take x ∈ M ′; using 3.8, we have σψa that σ 1 b⊗α N x) = 1 b⊗α N ψ t (1 b⊗α N x, and, therefore, belongs to A b∗α N ∆−it bΦ x∆it bΦ M. Let n ∈ N; we have : ψ b⊗α ψ(∆−it ν x) = 1 b⊗α N ∆it bΦ ∆−it bΦ t (1 b⊗α N x∆it bΦ dψo )it( ; therefore, we get that ( dψa dψ dψo )−it commutes with , and, using 4.8, we get ψ t (1 b⊗α σ N β(n)) = 1 b⊗α N ∆−it bΦ β(n)∆it bΦ = 1 b⊗α N τ−t(β(n)) = 1 b⊗α N β(σν −t(n)) and, on the other hand : σψa t (1 b⊗α N β(n)) = σψa t (a(b(n))) = a(σψ t (b(n))) = a(b(σν −t(n)) = 1 b⊗α N β(σν −t(n)) which proves that both ψ and ψa are lifted weights from the weight νo, and, therefore, that (Dψa : Dψ)t belongs to A b∗α (cid:3) N β(N)′, which finishes the proof. 7.2. Definition. Let G = (N, M, α, β, Γ, T, T ′, ν) be a measured quantum groupoid, (b, a) a weighted action of G on a von Neumann algebra A, ψ a normal semi-finite faithful weight on A lifted from νo; we shall call the unitaries (Dψa : Dψ)t ∈ A b∗α (M ∩β(N)′) the N Radon-Nikodym derivative of the weight ψ with respect to the action (b, a), and denote it, for simplification, (Dψ ◦ a : Dψ)t, following the notations of ([BV], 10.2). 38 7.3. Theorem. Let G = (N, M, α, β, Γ, T, T ′, ν) be a measured quantum groupoid, (b, a) a weighted action of G on a von Neumann algebra A, ψ a normal semi-finite faithful weight on A lifted from νo; the Radon-Nikodym derivative (Dψ ◦ a : Dψ)t introduced in 7.2 is a a-cocycle, i.e., we have : (id b∗α N Γ)((Dψ ◦ a : Dψ)t) = (a b∗α N id)((Dψ ◦ a : Dψ)t)((Dψ ◦ a : Dψ)t) β⊗α 1) N Proof. For all t ∈ R, (a b∗α N id)((Dψ ◦ a : Dψ)t) belongs to A b∗α N M β∗α N M, and the operator a((Dψ ◦ a : Dψ)t) = Θ−1(a b∗α N id)((Dψ ◦ a : Dψ)t) belongs to A b∗α N M β∗α N M (where Θ had been defined in 6.8) . We have, using successively 2.4, 6.5 and 5.1(iii) : a((Dψ ◦ a : Dψ)t) = a((Dψa : Dψ)t) = (D(ψa)a : D(ψ)a)t = (D(ψa) : D(ψa) ◦ Θ)t δ On the other hand, using successively 4.5(ii) and 6.8 : Θ−1((Dψ ◦ a : Dψ)t) β⊗α N 1) = Θ−1((Dψa : Dψ)t β⊗α N 1) = Θ−1(Dψa : Dψ)t) = (Dψa ◦ Θ : Dψ ◦ Θ)t = (Dψa ◦ Θ : D(ψ) )t δ and, therefore, we get that : Θ−1[(a b∗α N id)((Dψ ◦ a : Dψ)t)((Dψ ◦ a : Dψ)t) β⊗α 1)] N = a((Dψ ◦ a : Dψ)t) Θ−1((Dψ ◦ a : Dψ)t) β⊗α N 1) is equal, using 6.6(ii), to : (D(ψa) δ : D(ψa) ◦ Θ)t(D(ψa) ◦ Θ : D(ψ) δ )t = (D(ψa) : D(ψ) )t δ δ = (Dψa : Dψ)t β⊗α N 1 from which we get that : (a b∗α N id)((Dψ ◦ a : Dψ)t)((Dψ ◦ a : Dψ)t) β⊗α N = (Dψ ◦ a : Dψ)t β⊗α N 1 1) = Θ((Dψ ◦ a : Dψ)t β⊗α 1) N = (id b∗α N Γ)((Dψ ◦ a : Dψ)t) which is the result. (cid:3) 7.4. Example. Let G be a locally compact quantum group, and a an action of G on a von Neumann algebra A; then this result had been obtained in ([Y4], 4.8 and [Y5], 3.7 and [BV], 10.3). 39 7.5. Example. Let G be a measured groupoid; let us use all the notations introduced G(0) Axdν(x), and G(0) ψxdν(x) a normal semi-finite faithful weight on A. Then, the Radon-Nikodym in 5.10. Let (a)g∈G be an action of G on a von Neumann algebra A = R ⊕ ψ = R ⊕ derivative of ψ with respect to the action a, is, using ([Y3], 2.6), given by : (Dψ ◦ a : Dψ)t =Z ⊕ G (Dψr(g) : Dψs(g) ◦ ag−1)tdν(g) which is acting on R ⊕ G Hψr(g)dµ(g) = Hψ b⊗rG L2(G, µ). ν 7.6. Definition. Let (b, a) an action of a measured quantum groupoid G on a von Neu- mann algebra A. A normal semi-finite faithful weight ψ on A will be said invariant by a if, for all η ∈ D(αH, ν) ∩ D(Hβ, νo) and x ∈ Nψ, we have : ψ[(id b∗α N ωη)a(x∗x)] = kΛψ(x) a⊗β νo ηk2 We shall always suppose that such weights bear the density property, defined in 2.3, as for δ-invariant weights. 7.7. Theorem. Let (b, a) an action of a measured groupoid G on a von Neumann algebra A, ψ a normal semi-finite faithful weight on A, invariant by a in the sense of 7.6, and bearing the density property, as defined in 2.3. Then, let (ei)i∈I be an (α, ν)-orthogonal basis of H, x ∈ Nψ, η ∈ D(αH, ν) ∩ D(Hβ, νo) : (i) for any ξ ∈ D(αH, ν), (id b∗α N ωη,ξ)a(x) belongs to Nψ; ωη,ei)a(x) b⊗α ν ei is strongly converging; its limit does not depend (ii) the sumPi Λψ((id b∗α N upon the choice of the (α, ν)-othogonal basis of H, and allow us to define an isometry V ′ ψ from Hψ a⊗β νo H to Hψ b⊗α H such that : ν V ′ ψ(Λψ(x) a⊗β νo η) =Xi Λψ((id b∗α N ωη,ei)a(x) b⊗α ν ei (iii) we have : Λψ((id b∗α N ωη,ξ)a(x)) = (id ∗ ωη,ξ)(V ′ ψ)λψ(x) (iv) for any y ∈ A, z ∈ M ′, n ∈ N, we have : a(y)V ′ ψ = V ′ ψ(y a⊗β N o 1) (1 b⊗α N (a(n) b⊗α N z)V ′ ψ = V ′ 1)V ′ ψ = V ′ ψ(1 a⊗β N o ψ(1 a⊗β N o z) α(n)) (1 b⊗α N (1 b⊗α N β(n))V ′ ψ = V ′ β(n))V ′ ψ = V ′ 1) ψ(b(n) a⊗β N o β(n)) ψ(1 a⊗β N o (v) the operator V ′ implements a; (vi) we have : ψ is a unitary; moreover, it is a copresentation of G on a(Hψ)b which ψ(∆it V ′ ψ a⊗β N o ∆−it bΦ ) = (∆it ψ b⊗α N ∆−it bΦ )V ′ ψ 40 Moreover, the weight ψ is lifted from νo; more precisely, there exists a normal faithful semi-finite operator-valued weight T from A onto b(N) such that ψ = νo ◦ b−1 ◦ T, and, for all x ∈ NT ∩ Nψ, we have : (T b∗α N id)a(x∗x) = 1 b⊗α N β ◦ b−1T(x∗x) = a(T(x∗x)) (ψ b∗α ν id)a(x∗x) = β ◦ b−1T(x∗x) (vii) we have : a(σψ t (y)) = (σψ t b∗α N τt)a(y) (viii) the standard implementation U a ψ = ∆it (ix) the dual weight satisfies ∆it ψ is equal to V ′ ψ; ψ b⊗α N ∆−it bΦ ; (x) the Radon-Nikodym derivative (Dψ ◦ a : Dψ)t is equal to 1. Proof. Result (i) is identical to ([E5], 8.3(i)), and (ii) is similar to ([E5], 8.3(ii) and 8.4(i)); the proof of (iii) is similar to the proof of ([E5], 8.4(ii) and (iii)), and the proof of (iv) is similar (and somehow simpler) to the proof of ([E5], 8.4(iv) and (v)). Now result (v) is obtained in a similar way to ([E5], 8.5 and 8.6); by similar calculations to ([E5], 8.7 and 8.8(i)), we obtain that, for all t ∈ R, we have σψ −t(n)), which gives the existence of a normal faithful semi-finite operator-valued weight T from A onto b(N) such that ψ = νo ◦ b−1 ◦ T. For any x ∈ Nψ ∩ NT, the vector Λψ(x) belongs to D(αH, ν), and we have, for any η ∈ H : t (b(n)) = b(σν kΛψ(x) a⊗β νo ηk2 = (β ◦ b−1T(x∗x)ηη) So, using the density property and 7.6, we get, for all x ∈ Nψ ∩ NT, that : (ψ b∗α ν id)a(x∗x) = β ◦ b−1T(x∗x) and, therefore, that : (T b∗α N id)a(x∗x) = 1 b⊗α N β ◦ b−1T(x∗x) = a(T(x∗x) we finish the proof of (vi) in a similar way to ([E5], 8.8(ii)). Then (vii) is a straightforward corollary of (vi) and (v), and (viii) and (ix) are obtained in a similar way to 3.2(i) and dψ (ii). As ∆ ψ = dψa dψo ([E5] 13.6) and ∆ψ b⊗α dψo by 4.4(ii), we infer from (ix) that N ψa = ψ, which, by 7.2, finishes the proof. ∆−1 bΦ = (cid:3) 7.8. Corollary. Let (b, a) be an action of a measured quantum groupoid G on a von Neumann algebra A; let ψ1, ψ2 be two invariant normal faithful semi-finite weights on A, as defined in 7.6, and let us suppose that both ψ1 and ψ2 bear the density property, as defined in 2.3. Then, for all t ∈ R, (Dψ1 : Dψ2)t belongs to Aa. Proof. The proof is similar to ([E5], 8.11). (cid:3) 7.9. Theorem. Let (b, a) be a weighted action of a measured quantum groupoid G on a von Neumann algebra A, and ψ a normal semi-finite faithful weight on A, lifted from νo. If the Radon-Nikodym derivative (Dψ ◦ a : Dψ)t is equal to 1, then the weight ψ is invariant by a in the sense of 7.6. 41 Proof. Let ξ ∈ D(αH, ν) ∩ D(Hβ, νo) ∩ D(∆−1/2 bΦ let us remark first that if y belongs to N bΦ ∩ N∗ bΦ bΦ ∩ N T N∗ t , and such that σz(x) belongs to N bΦ ∩ N∗ σ T bΦ all those conditions, and this gives that the set of such elements ξ is dense in H. Let η be in D(αH, ν) ∩ D(∆−1/2 η belongs to D(αH, ν), and x ∈ Nψ, analytic with respect to ψ, such that σ−i/2(x∗) belongs to Nψ. Then, we have, using 4.10(i) applied to νo : ξ belongs to D(αH, ν); , and is analytic with respect to , for all z ∈ C, then Λ bΦ(z) satisfies ) such that ∆−1/2 ∩ N T ∩ N∗ T ) such that ∆−1/2 bΦ bΦ bΦ ((ψ b∗α ν id)a(x∗x)ξ α⊗β N o ((ψ b∗α ν J bφ∆−1/2 bΦ ηξ α⊗β N o J bφ∆−1/2 bΦ η) = id)a(x∗x)Λνo(θα,ν(ξ, η))Λνo(θα,ν(ξ, η)) = νo(θα,ν(ξ, η)∗(ψ b∗α ν id)a(x∗x)θα,ν(ξ, η) which is equal, using 4.8 and 4.9, to : ψ(1 b⊗α N θα,ν(ξ, η))∗a(x∗x)(1 b⊗α N θα,ν(ξ, η))) By hypothesis, as ψa = ψ by 7.2, we get, using 2.3 that σ Moreover, we can write, thanks to the hypothesis and to 4.10 applied to νo : t (a(x)) = σψa t (a(x)) = a(σψ t (x)). ψ JνoΛνo(θα,ν(ξ, η)) = J bΦξ β⊗α ν ∆−1/2 bΦ η = Λνo(θα,ν(∆−1/2 bΦ η, ∆1/2 bΦ ξ)) from which we get that [a(x)(1 b⊗α N θα,ν(ξ, η))]∗ belongs to D(σ ψ −i/2), and, therefore, that: ((ψ b∗α ν id)a(x∗x)ξ α⊗β νo J bφ∆−1/2 bΦ ηξ α⊗β νo J bφ∆−1/2 bΦ η) is equal to : kΛψ(σ ψ −i/2([a(x)(1 b⊗α N θα,ν(ξ, η))]∗)k2 = kΛψ((1 b⊗α N θα,ν(∆−1/2 bΦ η, ∆1/2 bΦ ξ))a(σψ −i/2(x∗)))k2 which, thanks again to the hypothesis and to 4.11, is equal to : ηk2 = kJψΛψ(x) b⊗α ν −i/2(x∗)) b⊗α ν J bΦξ β⊗α kΛψ(σψ ∆−1/2 bΦ ν J bΦξ β⊗α ∆−1/2 ηk2 ν bΦ J bΦ∆−1/2 bΦ ηk2 = kΛψ(x) a⊗β νo ξ α⊗β νo So, finally, we get the equality : ((ψ b∗α ν id)a(x∗x)ξ α⊗β νo J bφ∆−1/2 bΦ ηξ α⊗β νo J bφ∆−1/2 bΦ η) = kΛψ(x) a⊗β νo ξ α⊗β νo J bΦ∆−1/2 bΦ ηk2 which, by continuity, remains true for any x ∈ Nψ and ξ ∈ D(αH, ν) ∩ D(Hβ, νo); from which we infer that : ((ψ b∗α ν id)a(x∗x)α(< J bΦ∆−1/2 bΦ η, J bΦ∆−1/2 bΦ η >β,νo)ξξ) = (Λψ(x) a⊗β νo α(< J bΦ∆−1/2 bΦ η, J bΦ∆−1/2 bΦ η >β,νo)ξΛψ(x) a⊗β νo ξ) from which, by density of the elements of the form < J bΦ∆−1/2 we get, for any n ∈ N + : bΦ η, J bΦ∆−1/2 bΦ η >β,νo in N +, ((ψ b∗α ν id)a(x∗x)α(n)ξξ) = (Λψ(x) a⊗β νo α(n)ξΛψ(x) a⊗β νo ξ) 42 from which we get the result, by density of D(αH, ν) ∩ D(Hβ, νo). (cid:3) 7.10. Proposition. Let G be a measured quantum groupoid, (b, a) a weighted action of G on a von Neumann algebra A, ψ1 and ψ2 two normal semi-finite faithful weights on A, lifted from νo, and (Dψ1 ◦ a : Dψ1)t, (Dψ2 ◦ a : Dψ2)t their Radon-Nikodym derivatives with respect to the action (b, a), as defined in 7.2. Then, the Radon-Nikodym derivative (Dψ1 : Dψ2)t belongs to A ∩ b(N)′, and we have, for all t ∈ R : (Dψ2 ◦ a : Dψ2)t = a((Dψ2 : Dψ1)t)(Dψ1 ◦ a : Dψ1)t((Dψ2 : Dψ1)∗ t b⊗α N 1) Proof. As ψ1 and ψ2 are lifted weights from ν, we get that (Dψ1 : Dψ2)t belongs to A ∩ b(N)′ by ([T], 4.22(iii)); moreover, we have : (Dψ2a : Dψ2)t = (Dψ2a : Dψ1a)t(Dψ1a : Dψ1)t(Dψ1 : Dψ2)t from which we get the result, using 2.3, 7.2 and 4.5(ii). (cid:3) 7.11. Corollary. Let G be a measured quantum groupoid, (b, a) a weighted action of G on a von Neumann algebra A; then, are equivalent : (i) there exists a normal semi-finite faithful weight on A, which is invariant and bears the density condition; (ii) there exists a normal semi-finite faithful weight ψ on A, lifted from νo, and a σψ t - cocycle ut on A ∩ b(N)′ such that (Dψ ◦ a : Dψ)t = a(u∗ (iii) for any normal semi-finite faithful weight ψ on A, lifted from νo, there exists a σψ t -cocycle ut on A ∩ b(N)′ such that (Dψ ◦ a : Dψ)t = a(u∗ t )(ut b⊗α N 1); t )(ut b⊗α N 1). Proof. Let suppose (i), and let ϕ be an invariant weight on A, bearing the density con- dition; then, by 7.7(vi), the weight is lifted, and, if ψ is any another lifted weight on A, ut = (Dϕ : Dψ)t is a σψ t -cocycle in A ∩ b(N)′ by ([T], 4.22(iii)); moreover, using 7.10, we get (iii). Conversely, if we suppose (ii), there exists a normal semi-finite faithful weight ϕ on A such that ut = (Dϕ : Dψ)t; as ψ is lifted, and ut belongs to A ∩ b(N)′, we know, using ([T], 4.22(iii)), that ϕ is lifted, too. Using now 7.10, we get that (Dϕ ◦ a : Dϕ)t = 1, which, thanks to 7.9, gives the result. (cid:3) References [BS] S. Baaj and G. Skandalis : Unitaires multiplicatifs et dualit´e pour les produits crois´es de C∗- alg`ebres, Ann. Sci. ENS, 26 (1993), 425-488. 2 [BSV] S. Baaj, G. Skandalis and S. Vaes : Non-semi-regular quantum groups coming from number theory, Comm. Math. Phys., 235 (2003), 139-167. 8, 9 [BV] S. Baaj and S. Vaes : Double crossed products of locally compact quantum groups, J. Inst. Math. Jussieu, 4 (2005), 135-173. 2, 3, 38, 39 [Bl1] E. Blanchard : Tensor products of C(X)-algebras over C(X), Ast´etrisque 232 (1995), 81-92. 8 [Bl2] E. Blanchard : D´eformations de C∗-alg`ebres de Hopf, Bull. Soc. Math. France, 24 (1996), 141-215. 8 [BSz1] G. Bohm and K. Szlach´anyi : A Coassociative C∗-Quantum group with Non Integral Dimensions, Lett. Math. Phys., 38 (1996), 437-456. 2 [BSz2] G. Bohm and K. Szlach´anyi : Weak C∗-Hopf Algebras : the coassociative symmetry of non- integral dimensions, in Quantum Groups and Quantum spaces Banach Center Publications, 40 (1997), 9-19. 2 [C1] A. Connes: On the spatial theory of von Neumann algebras, J. Funct. Analysis, 35 (1980), 153-164. 3 [C2] A. Connes: Non commutative Geometry, Academic Press, 1994 20 43 [D] M.-C. David : C∗-groupoıdes quantiques et inclusions de facteurs ; structure sym´etrique et auto- dualit´e, action sur le facteur hyperfini I I1, J. Operator Theory, 54 (2005), 27-68. 8 [DC] K. De Commer : Monoidal equivalence for locally compact quantum groups, mathOA/0804.2405, to appear in J. Operator Theory 8 [E1] M. Enock : Produit crois´e d'une alg`ebre de von Neumann par une alg`ebre de Kac, J. Funct. Analysis, 26 (1977), 16-46 2 [E2] M. Enock : Inclusions irr´eductibles de facteurs et unitaires multiplicatifs II, J. Funct. Analysis, 137 (1996), 466-543. 2 [E3] M. Enock : Quantum groupoids of compact type, J. Inst. Math. Jussieu, 4 (2005), 29-133. 20 [E4] M. Enock : Inclusions of von Neumann algebras and quantum groupoids III, J. Funct. Analysis, 223 (2005), 311-364. 2, 8, 15 [E5] M. Enock : Measured Quantum Groupoids in action, M´emoires de la SMF , 114 (2008), 1-150. 2, 3, 4, 5, 6, 7, 8, 9, 11, 13, 14, 15, 19, 21, 25, 28, 29, 31, 32, 33, 36, 41 [E6] M. Enock : Measured Quantum Groupoids with a central basis, mathOA/0808.4052 8 [E7] M. Enock : Outer actions of measured quantum groupoids, mathOA/0909.1206. 8 [EN] M. Enock, R. Nest : Inclusions of factors, multiplicative unitaries and Kac algebras, J. Funct. Analysis, 137 (1996), 466-543. 2, 15 [ES1] M. Enock, J.-M. Schwartz : Produit crois´e d'une alg`ebre de von Neumann par une alg`ebre de Kac II, Publ. RIMS Kyoto, 16 (1980), 189-232. 2 [ES] M. Enock, J.-M. Schwartz : Kac Algebras and Duality of locally compact Groups, Springer-Verlag, Berlin, 1989. 2 [EV] M. Enock, J.-M. Vallin : Inclusions of von Neumann algebras and quantum groupoids, J. Funct. Analysis, 172 (2000), 249-300. 2, 3, 8 [J] V. Jones : Index for subfactors, Invent. Math., 72 (1983), 1-25. 8 [KV1] J. Kustermans and S. Vaes : Locally compact quantum groups, Ann. Sci. ENS, 33 (2000), 837- 934. 2, 8 [KV2] J. Kustermans and S. Vaes, Locally compact quantum groups in the von Neumann algebraic setting, Math. Scand., 92 (2003), 68-92. 2, 8 [L] F. Lesieur : Measured Quantum Groupoids, M´emoires de la SMF, 109 (2007), 1-158. 2, 4 [MN] T. Masuda and Y. Nakagami : A von Neumann Algebra framework for the duality of the quantum groups, Publ. RIMS Kyoto, 30 (1994), 799-850. 2 [MNW] T. Masuda, Y. Nakagami and S.L. Woronowicz : A C∗-algebraic framework for quantum groups, Internat. J. math. , 14 (2003), 903-1001. 2 [NV1] D. Nikshych, L. Vaınerman : Algebraic versions of a finite dimensional quantum groupoid, in Lecture Notes in Pure and Applied Mathematics, Marcel Dekker, 2000. 2, 8 [NV2] D. Nikshych, L. Vaınerman : A characterization of depth 2 subfactors of I I1 factors, J. Funct. Analysis, 171 (2000), 278-307. 2, 8 [NV3] D. Nikshych, L. Vaınerman : Finite quantum groupoids and their applications, in New Directions in Hopf algebras, S. Montgomery and H.-J. Schneider editors, MSRI Publ. 43, 211-262, Cambridge University Press (2002). 8 [S] J.-L. Sauvageot : Sur le produit tensoriel relatif d'espaces de Hilbert, J. Operator Theory, 9 (1983), 237-352. 3, 16, 17 [St] S¸. Stratila : Modular theory in Operator Algebras, Abacus Press, Turnbridge Wells, England, 1981 9, 10, 27 [Sz] K. Szlach´anyi : Weak Hopf algebras, in Operators Algebras and Quantum Field Theory, S. Doplicher, R. Longo, J.E. Roberts, L. Zsido editors, International Press, 1996. 2 [T] M. Takesaki : Theory of Operator Algebras II, Springer, Berlin, 2003. 3, 7, 15, 18, 32, 43 [V1] S. Vaes : The unitary implementation of a locally compact Quantum Group action, J. Funct. Analysis, 180 (2001), 426-480. 2, 3, 8, 9, 32 [V2] S. Vaes : Strictly outer actions of groups and quantum groups, J. reine angew. Math., 578 (2005), 147-184. 8 [VV] S. Vaes, L. Vaınerman : Extensions of locally compact quantum groups and the bicrossed product construction, Advances in Mathematics, 175 (2003), 1-101. 8 [Val1] J.-M. Vallin : Bimodules de Hopf et Poids op´eratoriels de Haar, J. Operator theory, 35 (1996), 39-65 2 [Val2] J.-M. Vallin : Unitaire pseudo-multiplicatif associ´e `a un groupoıde; applications `a la moyennabilit´e, J. Operator theory, 44 (2000), 347-368. 2 44 [Val3] J.-M. Vallin : Groupoıdes quantiques finis, J. Algebra, 239 (2001), 215-261. 2, 8 [Val4] J.-M. Vallin : Multiplicative partial isometries and finite quantum groupoids, in Locally Compact Quantum Groups and Groupoids, IRMA Lectures in Mathematics and Theoretical Physics 2, V. Turaev, L. Vainerman editors, de Gruyter, 2002. 2, 8 [Val5] J.-M. Vallin : Measured quantum groupoids associated with matched pairs of locally compact groupoids, mathOA/0906.5247. 9 [W1] S.L. Woronowicz : Tannaka-Krein duality for compact matrix pseudogroups. Twisted SU (N ) group. Invent. Math., 93 (1988), 35-76. 2 [W2] S.L. Woronowicz : Compact quantum group, in "Sym´etries quantiques" (Les Houches, 1995), North-Holland, Amsterdam (1998), 845-884. 2 [W3] S.L. Woronowicz : From multiplicative unitaries to quantum groups, Int. J. Math., 7 (1996), 127-149. 2 [Y1] T. Yamanouchi : Crossed product by groupoid actions and their smooth flows of weights, Publ. RIMS, 28 (1992), 535-578. 8, 32 [Y2] T. Yamanouchi : Dual weights on crossed products by groupoid actions, Publ. RIMS, 28 (1992), 653-678. 8, 32 [Y3] T. Yamanouchi : Duality for actions and coactions of measured Groupoids on von Neumann Alge- bras, Memoirs of the A.M.S., 101 (1993), 1-109. 2, 8, 32, 40 [Y4] T. Yamanouchi : Takesaki duality for weights on locally compact quantum group covariant systems, J. Operator Theory, 50 (2003), 53-66. 2, 8, 38, 39 [Y5] T. Yamanouchi : Canonical extension of actions of locally compact quantum groups, J. Funct. Analysis, 201 (2003), 522-560. 38, 39 Institut de Math´ematiques de Jussieu, Unit´e Mixte Paris 6 / Paris 7 / CNRS de Recherche 7586, 175, rue du Chevaleret, Plateau 7E, F-75013 Paris E-mail address: [email protected] 45
1710.08406
1
1710
2017-10-23T17:55:27
On dualities of actions and inclusions
[ "math.OA", "math.DS" ]
Following the results known in the case of a finite abelian group action on $C\sp*$-algebras we prove the following two theorems; 1. an inclusion $P\subset A$ of (Watatani) index-finite type has the Rokhlin property (is approximately representable) if and only if the dual inclusion is approximately representable (has the Rokhlin property). 2. an inclusion $P\subset A$ of (Watatani) index-finite type has the tracial Rokhlin property (is tracially approximately representable) if and only if the dual inclusion is tracially approximately representable (has the tracial Rokhlin property). Moreover, we provide an alternate proof of Phillips' theorem about the relations between tracial Rokhlin action and tracially approximate representable dual action using a new conceptual framework suggested by authors.
math.OA
math
ON DUALITIES OF ACTIONS AND INCLUSIONS HYUN HO LEE AND HIROYUKI OSAKA Abstract. Following the results known in the case of a finite abelian group action on C ∗-algebras we prove the following two theorems; • an inclusion P ⊂ A of (Watatani) index-finite type has the Rokhlin property (is approximately representable) if and only if the dual inclusion is approximately representable (has the Rokhlin property). • an inclusion P ⊂ A of (Watatani) index-finite type has the tracial Rokhlin property (is tracially approximately representable) if and only if the dual in- clusion is tracially approximately representable (has the tracial Rokhlin prop- erty). Moreover, we provide an alternate proof of Phillips' theorem about the relations between tracial Rokhlin action and tracially approximate representable dual action using a new conceptual framework suggested by authors. 7 1 0 2 t c O 3 2 ] . A O h t a m [ 1 v 6 0 4 8 0 . 0 1 7 1 : v i X r a 1. Introduction In [7] M.Izumi introduced the Rokhlin property for a finite group action on C ∗- algebras. In addition, he showed that a finite abelian group action has the Rokhlin property if and only if the dual action is approximately representable which means that the action is strongly approximately inner. This observation is sometimes useful and easier to verify rather than the Rokhlin property itself. Based on Izumi's works, the second named author, Kodaka, and Teruya extend notions of the Rokhlin property and the approximate representability for inclusions of unital C ∗-algebras. In this note we show that an inclusion has the Rokhlin property if and only if its dual inclusion is approximately representable. On the other hand, a great success of classifying nuclear simple tracially AF C ∗- algebras satisfying the UCT [9, 10] provided a conceptual revolution in the Eliiott program; no inductive limit structure is assumed. The added flexibility to allow "small" tracial error in the local approximation led to important advances in the theory of C ∗-algebras. It is thus desirable to expect "tracial" versions of other C ∗- algebra concepts, see [5],[3] for instance. But also in part due to the fact that often the Rokhlin property imposes several restrictions on the K-theory of the original algebra and the K-theory of the crossed product algebra it is expected to have a less restrictive or tracial version of the Rokhlin property, and such a notion was suggested 2000 Mathematics Subject Classification. Primary:46L35. Secondary:47C15. Key words and phrases. Tracially sequentially-split map, (tracial) Rokhlin property, (tracial) Approximate representability, Inclusion of C ∗-algebras. The first author's research was supported by Basic Science Research Program through the National Research Foundation of Korea(NRF) funded by the Ministry of Education(NRF- 2015R1D1A1A01057489) The second author's research was partially supported by the JSPS grant for Scientific Research No. 17K05285. 1 2 HYUN HO LEE AND HIROYUKI OSAKA by N.C. Phillips, which is called the tracial Rokhlin property. He then extended many statements appeared in [7] when a finite group action on the infinite dimensional simple separable unital C ∗-algebra has the tracial Rokhlin property. In particular, he shows that a finite abelian group action has the tracial Rokhlin property if and only if the dual group action is tracially approximately representable (see Theorem 4.14 or [15, Theorem 3.11]). In this note we are going to define the tracial versions of the Rokhlin property and the approximate representability for inclusions of unital C ∗-algebras and show a duality between them. One might think that it is not difficult to find analogous notions while extending original ones. But in our case, though the pair of the crossed product algebra and the original algebra is a standard model for inclusions of C ∗-algebras, it is not ob- vious which part of the definition is made flexible; the tracial Rokhlin property is straightforward but not at all for the tracial approximate representability. Thus we wish to point out a philosophical guiding principle; a regularity property involving two objects can be expressed as a certain property of a map between two objects. More precisely, Barlak and Szab´o [1] provide an unified conceptual framework to deal with the permanence of various regularity properties from the target algebra to the domain algebra in the name of sequentially split ∗-homomorphisms between C ∗-algebras. This concept is actually originated from Toms and Winter's charac- terization of D-stability, where D is a unital strongly self-absorbing C ∗-algebra, but nicely fits into the Rokhlin property as well. A tracial version of the concept has been suggested by the authors in [11] and turned out to work nicely with tracial versions of regularity properties, for instance tracial Z-stability and tracial Rokhlin property. In this note we solidify our guiding principle by exhibiting that the Rokhlin property and the approximate representability of an inclusion of unital C ∗-algebras P ⊂ A could be characterized by the existence of a map from A to the sequence algebra of P and a map from the C ∗-basic construction to the sequence algebra of A (see Proposition 3.9 and Proposition 3.6). Accordingly, we characterize the tracial version of the Rokhlin property and the approximate representability of the action of a finite abelian group G on a unital separable C ∗-algebra A by the existence of a map from C(G) the algebra of continuous functions on G to the central sequence algebra of A and a map from the crossed product algebra to the sequence algebra of A as is expected. Then we provide an alternate proof of the duality result of Phillips using these characterizations. Moreover, we show an interplay between group actions and inclusions of C ∗-algebras based on our duality results for both the strict case and the tracial case. 2. Tracially sequentially-split homomorphism between C ∗-algebras In this section we briefly review the definition of tracially sequentially-split map between separable C ∗-algebras from [11] and introduce notations which will be used throughout the note. For a C ∗-algebra A, we set the C ∗-algebra of bounded sequence over N with values in A and the ideal of sequences converging to zero as follows; l∞(N, A) = {(an) {kank} bounded} c0(N, A) = {(an) lim n→∞kank = 0}. A TRACIALLY SEQUENTIALLY-SPLIT ∗-HOMOMORPHISMS BETWEEN C ∗-ALGEBRAS II 3 Then we denote by A∞ = l∞(N, A)/c0(N, A) the sequence algebra of A with the norm of a given by lim supn kank, where (an)n is a representing sequence of a. We can embed A into A∞ as a constant sequence, and we denote the central sequence algebra of A by A∞ ∩ A′. For an automorphism of α on A, we also denote by α∞ the induced automorphism on A∞ and A∞ ∩ A′ without confusion. We save the notation . for the Cuntz subequivalence of two positive elements; for two positive elements a, b in A a . b if there is a sequence (xn) in A such that kxnbx∗ n − ak → 0 as n → ∞. Often when p is a projection, we see that p . a if and only if there is a projection in the hereditary C ∗-subalgebra generated by a which is Murray-von Neumann equivalent to p. For more details, we refer [2, 16, 17] for example. Definition 2.1. Let A and B be (unital) separable C ∗-algebras. A ∗-homomorphism φ : A → B is called tracially sequentially-split, if for every positive nonzero element z ∈ A∞ there exist a ∗-homomorphism ψ : B → A∞ and a nonzero projection g ∈ A∞ ∩ A′ such that (1) ψ(φ(a)) = ag for each a ∈ A, (2) 1A∞ − g is Murray-von Neumann equivalent to a projection in a hereditary C ∗-subalgebra zA∞z in A∞. We also consider the following alternative stronger condition to replace (2) in the above definition; for any δ > 0 there exist a ∗-homomorphism ψ : B → A∞ and a projection g ∈ A∞ ∩ A′ such that (1) ψ(φ(a)) = ag for each a ∈ A, (2)' τ (1 − g) < δ for all τ ∈ T (A∞) the tracial states of A∞. Since (ψ◦ φ)(a) = a−a(1A∞ −g), we can view ψ◦ φ equal to ι up to "tracially small error". If A and B are unital C ∗-algebras and φ is unit preserving, then g = ψ(1B). Moreover, if g = 1A∞, then φ is called a (strictly) sequentially split ∗-homomorphism following Barlak and Szabo [1] since the second condition is automatic . The ψ in the above definition is called a tracial approximate left-inverse of φ. Although the diagram below is not commutative, we still use it to symbolize the definition of a tracially sequentially-split map φ and its tracial approximate left inverse ψ; (1) A / A∞ ι B ❄❄❄❄❄❄❄❄ φ =⑤⑤⑤⑤⑤⑤⑤⑤ ψ tracially(cid:9) Definition 2.2. A C ∗-algebra A has the property (SP) if any nonzero hereditary C ∗-subalgebra of A has a nonzero projection. Proposition 2.3. Let φ : A → B be a tracially sequentially-split ∗-homomorphism. Then A has the property (SP) or φ is a (strictly) sequentially split ∗-homomorphism. Proof. Suppose that A has no property (SP), then A∞ has no property (SP). Then there is a positive nonzero element x in A∞ which generates a hereditary subalgebra  / = 4 HYUN HO LEE AND HIROYUKI OSAKA that contains no nonzero projections. Then since φ : A → B is tracially sequentially- split, there are a projection g ∈ A∞ ∩ A′ and a ∗-homomorphism ψ : B → A∞ such that (1) ψ(φ(a)) = ag for all a ∈ A, (2) 1 − g is Murray-von Neumann equivalent to a projection in xA∞x. The second condition implies 1 − g = 0 so that ψ(φ(a)) = a, thus φ is strictly sequentially split. (cid:3) The following proposition which is a tracial version of [1, Lemma 2.4] will be used in Section 4. Proposition 2.4. Let C and A be unital C ∗-algebras. Suppose that for any nonzero positive element z ∈ A∞ there exists a map φ : C → A∞ ∩ A′ such that 1 − φ(1C) is Murray-von Neumann equivalent to a projection in zA∞z. Then the first factor embedding idA ⊗1C : A → A⊗C is tracially sequentially-split. Moreover, the converse is also true. Proof. Let us denote by m the map from A ⊗ (A∞ ⊗ A′) to A∞ which is defined by a ⊗ [(an)] → [(aan)]. For any nonzero positive element x in A∞ we take a map φ : C → A∞ as above and define ψ as the composition of two maps m and idA ⊗φ, i.e., ψ = m ◦ (idA ⊗φ). It follows immediately that ψ(1A ⊗ 1C) = φ(1C). Then it is easily checked that (1) (ψ ◦ (idA ⊗1C)(a) = ψ(a ⊗ 1C) = aφ(1C) = aψ(1A ⊗ 1C) for any a ∈ A, (2) 1− ψ(1A⊗ 1C) = 1− φ(1C) is Murray-von Neumann equivalent to a projection in xA∞x. Conversely, for any positive nonzero element x in A∞ we consider a tracial approx- imate inverse ψ for idA ⊗1C. Let φ(c) = ψ(1A ⊗ c). Obviously, φ is a ∗-homomorphism from C to A∞. Moreover, φ(c)a = ψ(1A ⊗ c)a = ψ(1A ⊗ c)ψ(1A ⊗ 1C)a = ψ(1A ⊗ c)ψ(a ⊗ 1C) = ψ(a ⊗ 1C)ψ(1A ⊗ c) = aψ(1A ⊗ 1C)ψ(1A ⊗ c) = aφ(c) Therefore φ(C) ⊂ A∞ ∩ A′. Finally, 1 − φ(1C) = 1 − ψ(1A ⊗ 1C) is Murray-von Neumann equivalent to a projection in xA∞x. (cid:3) When A and B are equipped with group actions, instead of ordinary ∗-homomorphism we consider equivariant ones to define the equivariant version of a tracially sequentially- split map. Definition 2.5. Let A and B be separable unital C ∗-algebras and G a finite group. Let α : G y A and β : G y B be two actions. An equivariant ∗-homomorphism φ : (A, α) → (B, β) is called G-tracially sequentially-split if for every nonzero positive element x in A∞ there exist an equivariant ∗-homomorphism ψ : (B, β) → (A∞, α∞) and a projection g in A∞ ∩ A′ such that A TRACIALLY SEQUENTIALLY-SPLIT ∗-HOMOMORPHISMS BETWEEN C ∗-ALGEBRAS II 5 (1) ψ(φ(a)) = ga = ag, (2) 1 − g is Murray-von Neumann equivalent to a projection in xA∞x. We shall use the following diagram to describe the equivariant case of tracially sequentially-split map φ and its tracial approximate left inverse ψ; (2) (A, α) /❴❴❴❴❴❴❴❴❴ (A∞, α∞) ι $■■■■■■■■■ φ tracially(cid:9) (B, β) 9rrrrrrrrrr ψ The following is a straightforward generalization of Proposition 2.4 to the equivariant case. We remark that a group G could be compact for this statement but we restrict ourselves to finite groups with applications in mind. Proposition 2.6. Let C and A be unital C ∗-algebras. Let α : G y A and β : G y C be two action of a finite group G. Suppose that for any nonzero positive element x ∈ A∞ there exists an equivariant ∗-homomorphism φ : (C, β) → (A∞∩A′, α∞) such that 1− φ(1C) is Murray-von Neumann equivalent to a projection in zA∞z. Then the first factor embedding idA ⊗1C : (A, α) → (A⊗ C, α⊗ β) is tracially sequentially-split. Moreover, the converse is also true. Proof. The proof is almost same, the only thing to be careful is that the map m : (A∞ ∩ A′) ⊗ A → A∞ in Proposition 2.4 is equivariant with respect to actions and this is easily checked. (cid:3) 3. Dualities of actions and inclusions: The strict case Definition 3.1 (Watatani[20]). Let P ⊂ A be an inclusion of unital C ∗-algebras and E : A → P a conditional expectation. Then we way that E has a quasi-basis if there exist elements {(uk, vk)} for k = 1, . . . , n such that for any x ∈ A In this case, we define the Watatani index of E as x = ujE(vjx) = E(xuj)vj. nXj=1 Index E = ujvj. nXj=1 nXj=1 In other words, we say that E has a finite index if there exist a quasi-basis. 1), . . . , (un, u∗ It is proved that once we know the existence of a quasi-basis then a quasi-basis can be chosen as {(u1, u∗ n)} so that Index E is a nonzero positive element in A commuting with A. Thus if A is simple, it is a nonzero positive scalar. We also recall Watatani's notion of the C ∗-basic construction of the above triple (P, A, E : A → P ); since we only consider the case that the conditional expectation E : A → P is of index-finite type , we do not distinguish the reduced construction and maximal one. $ / 9 6 HYUN HO LEE AND HIROYUKI OSAKA Definition 3.2. Let P ⊂ A be an inclusion of unital C ∗-algebras and E : A → P a conditional expectation. Now we assume E is faithful. Let EE be the Hilbert P - module completion of A by the norm given by a P -valued Hermitian bilinear form hx, yiP = E(x∗y) for x, y ∈ A. As usual L(EE) will be the algebra of adjointable bounded operators on EE. There are an injective ∗-homomorphism λ : A → L(EE) defined by a left multiplication and the natural inclusion map ηE from A to EE. The the Jones projection eP is defined by Then the C ∗-basic construction is the C ∗-algebra given by ep(ηE(x)) = ηE(E(x)). C ∗hA, ePi = { nXi=1 λ(xi)eP λ(yi) xi, yi ∈ A, n ∈ N}. C ∗hA, ePi onto A such that for x, y ∈ A When E is of index-finite type, there is a dual conditional expectation bE from Moreover, bE is also of index-finite type and faithful. bE(λ(x)epλ(y)) = (Index E)−1xy. From now on, otherwise stated, we only consider a faithful conditional expectation. Definition 3.3 (Osaka, Kodaka, and Teruya). A conditional expectation E : A → P of index-finite type has the Rokhlin property if there is a projection e ∈ A∞ ∩ A′ such that E∞(e) = (Index E)−1 and the map A ∋ x 7→ xe is injective. We call e a Rokhlin projection. Definition 3.4. (Osaka and Teruya) A conditional expectation E : A → P of index- finite type is approximately representable if there exists a projection e ∈ P∞ ∩ P ′ and a finite set of elements {ui} ⊂ A such that (1) exe = E(x)e for every x ∈ A, (3) the map x 7→ xe is injective for x ∈ P . (2) Pi uieu∗ i = 1 Remark 3.5. The condition (2) in the above didn't appear in the original definition of approximate representability [13], but we include it for the definition. The following proposition gives us a characterization of the approximate repre- sentability for an inclusion of C ∗-algebras reflecting the whole theme of our note. Proposition 3.6. A conditional expectation E : A → P of index-finite type is ap- proximately representable if and only if there is a unital injective ∗-homomorphism ψ : C ∗hA, epi → A∞ such that ψ(ep) ∈ P∞ ∩ P ′, ψ(x) = x for every x ∈ A Proof. Suppose that E : A → P is approximately representable. We note that condi- tion (2) implies that the C ∗-algebra generated by {xey x, y ∈ A} is nondegenerate. It follows that (ι, e) is a nondegenerate covariant representation where ι is the natural embedding of A into A∞ by a constant sequence. By [20, Proposition 2.2.11], we have i ae uieu∗ uiE(u∗ ae =Xi =Xi i a)ep). By the injectivity of ψ, aep =Pi uiE(u∗ i a)e Then we have ψ(aep) = ψ(Pi uiE(u∗ It follows that a =Pi uiE(u∗ is a quasi-basis, and ψ is a unital map. i a). Similarly, we have a =Pi E(aui)u∗ i a)ep. i )} Conversely, if we have a unital injective map ψ : ChA, epi → A∞ such that ψ(ep) ∈ P∞ ∩ P and ψ(a) = a for all a ∈ A, then from epaep = E(a)ep for all a ∈ A we also have eae = E(a)e for all a ∈ A. In addition, pe = 0 implies that ψ(pep) = 0. Thus pep = 0 by the injectivity of ψ. If {(ui, u∗ i = 1 in C ∗hA, epi. Thus i . Thus {(ui, u∗ i )} is a quasi-basis for E, thenPi uiepu∗ i ) =Xi i = 1. ψ(Xi uiepu∗ uieu∗ A TRACIALLY SEQUENTIALLY-SPLIT ∗-HOMOMORPHISMS BETWEEN C ∗-ALGEBRAS II 7 a map ψ from the C ∗-basis construction C ∗hA, epi to AeA such that ψ(xepy) = xey for all x, y ∈ A. Moreover the condition (3) implies that ψ is injective. Note that (cid:3) Remark 3.7. If a conditional expectation E : A → P of index-finite type is approxi- i )} is a quasi-basis for E. mately representable, and {ui} ⊂ A is the set satisfyingPi uieu∗ i = 1, then {(ui, u∗ Corollary 3.8. If η : A → C ∗hA, epi the natural embedding for the inclusion P ⊂ A of index-finite type, and the conditional expectation E : A → P is approximately representable, then η is a sequentially split ∗-homomorphism. Similarly we characterize an inclusion P ⊂ A with the Rokhlin property using a map from A to P∞ in the same spirit of Proposition 3.6. Proposition 3.9. Let P ⊂ A be an inclusion of unital C ∗-algebras and E : A → P be a conditional expectation of index-finite type. Suppose further A is simple. Then E has the Rokhlin property if and only if there exist a map β : A → P∞ and a projection e ∈ A∞ ∩ A′ such that (1) ae = β(a)e for all a ∈ A, (2) (Index E)eepe = e, (3) ye = ze implies that y = z for all y, z ∈ P∞, Proof. For "only if" part, see [13, Theorem 5.7]. In fact, take e a Rokhlin projection and it follows that (2) is true with the fact that (Index E)E∞(e) = 1, and for x ∈ A∞ xe = (Index E)bE∞(epxe) = (Index E)2bE∞(epxeepe) = (Index E)2bE∞(E∞(xe)epe) = (Index E)E∞(xe)e. 8 HYUN HO LEE AND HIROYUKI OSAKA Thus we define β(x) = (Index E)E∞(xe). Finally, so that if y and z satisfy ye = ze, then y =y(Index E)E∞(e) = (Index E)E∞(ye) =(Index E)E∞(ze) =z(Index E)E∞(e) = z Conversely, suppose that we have a map β and a projection e satisfying (1)-(3). Then (2) implies that ae = (Index E)E∞(ae)e as shown above. Then the conditions (1), (3) imply that β(a) = (Index E)E∞(ae). Then (1) and (3) again imply that β(1) = 1 so that (Index E)E∞(e) = 1. Thus e is a Rokhlin projection and the inclusion P ⊂ A has the Rokhlin property. (cid:3) Whenever we have a finite group action α : G y A, then we have an inclusion Aα ⊂ A where Aα is the fixed point algebra with a faithful conditional expectation E : A → Aα defined by E(a) = It is not always true that E is of index-finite type. But the following is known. GXg αg(a). 1 Theorem 3.10. [8, Theorem 4.1] Let α : G y A be an action of a finite group G on A. Then α is saturated if and only of E : A → Aα defined by E(a) = of index-finite type. GXg αg(a)is 1 Definition 3.11 (M. Izumi). Let α : G y A be an action of a finite group G on A. We say α has the Rokhlin property if there is a partition of unity {eg}g∈G of projections in A∞ ∩ A′ such that for all g, h ∈ G α∞,h(eg) = ehg where α∞ is the induced action. If α has the Rokhlin property, then α is outer, thus saturated. By Theorem 3.10 E : A → Aα defined by E(a) = is a characterization of the Rokhlin property of an action in term of inclusion of C ∗-algebras due to Osaka, Kodaka, and Teruya. αg(a) is of finite index. Then the following 1 GXg Proposition 3.12. [13, Proposition 3.2] Let G be a finite group, α : G y A be an action of a finite group G on a simple unital C ∗-algebra A, and E the conditional expectation defined by E(a) = αg(a). Then α has the Rokhlin property if and only if E has the Rokhlin property. 1 GXg In the following we present a parallel result for the approximately representable action α : G y A. In other words, we would like to characterize the approximate representability of α in terms of the conditional expectation E from A onto the fixed point algebra Aα. Definition 3.13. (M. Izumi) Let α : G y A be an action of a finite group G on A. When G is abelian, then the action α is called approximately representable if there A TRACIALLY SEQUENTIALLY-SPLIT ∗-HOMOMORPHISMS BETWEEN C ∗-ALGEBRAS II 9 is a unitary representation ω : G → (Aα)∞ such that for all a ∈ A α(a) = ωgaω∗ g. A tracial version the following theorem will be proved later, but at this moment we need it as a tool for our goal. Theorem 3.14. [7, Lemma 3.8] Let α : G y A be an action of a finite abelian group product algebra. G on a unital C ∗-algebra A and bα y A ⋊α G the dual action of bG on the crossed (1) α has the Rokhlin property if and only if bα is approximately representable. (2) bα has the Rokhlin property if and only if α is approximately representable. Proof. For the proof, we rather recommend [1, Theorem 4.27] which covers even for a second countable compact abelian group with its dual as a discrete countable abelian group. (cid:3) Although we don't use the following observation later, we include it since it provides a clue to what we want to obtain. Proposition 3.15. Let G be a finite abelian group and α : G y A approximately representable. Then the conditional expectation E : A → Aα defined by E(a) = 1 αg(a) satisfies a covariant relation (see (1) of Definition 3.4). i.e. there is a GXg projection e ∈ (Aα)∞ ∩ (Aα)′ such that for all a ∈ A eae = E(a)e. Proof. Let e = 1 GPg ωg. Then e∗ = 1 ωh 1 ωg 1 1 GXh GPg ωg−1 = e and GXg G2Xg,h GXg Ph ωgh GXg G e = e. ωgh 1 1 e2 = = = = Thus e is a projection. Moreover, for a ∈ Aα ea = 1 GXg ωga = 1 GXg αg(a)ωg = a 1 GXg ωg = ae. 10 HYUN HO LEE AND HIROYUKI OSAKA Hence e ∈ (Aα)∞ ∩ (Aα)′. Finally, for x ∈ A ωg! x 1 GXh ωh! 1 1 1 = exe = 1 GXg G2Xg,h G2Xh,s G2Xs,h GXs GXs = = 1 1 = = αg(x)ωgh αh−1s(x)ωs αsh−1(x)ωs αs(cid:18)Ph αh−1(x) G E(x)ωs = E(x)e (cid:19) ωs (cid:3) We need the following theorem as a final piece of our puzzle, which is of independent interest in the sense that it is an extension of Theorem 3.14. We remark that the following statement was claimed in [13] but the proof was incomplete in view of Definition 3.4. Theorem 3.16. Let P ⊂ A be an inclusion of unital C ∗-algebras and E a conditional expectation from A onto P with a finite index. Let B = C ∗hA, ePi be the basic construction and bE the dual conditional expectation of E from B onto A. Then (1) E has the Rokhlin property if and only if bE is approximately representable; (2) E is approximately representable if and only if bE has the Rokhlin property. Proof. (1): Let e be the Rokhlin projection in A∞ ∩ A′. We will show that there exists a finite set {ui} ⊂ B such that Let {(vi, v∗ i )} a quasi-basis for E. Then put ui = √Index Eviep. Then uieu∗ Xi Xi i =Xi =Xi =Xi uieu∗ i = 1. Index E(vieP eeP v∗ i ) Index E(viE∞(e)eP v∗ i ) vieP v∗ i = 1 Conversely, if bE is approximately representable, then we have a projection e ∈ A∞∩A and a finite set {ui} ∈ B such that (3) eze = bE(z)e ∀z ∈ B, A TRACIALLY SEQUENTIALLY-SPLIT ∗-HOMOMORPHISMS BETWEEN C ∗-ALGEBRAS II 11 (4) uieu∗ i = 1 Xi Write ui =Pj aijeP bij for some aij, bij ∈ A. Let wi =Pj E(aij)bij. Note that E∞(wiew∗ Xi i )eP =Xi =Xi =Xi = eP . E∞ Xj eP Xj eP Xj E(aij)bij! e Xk E(aij)bij! e Xk aijeP bij! e Xk ikE(a∗ b∗ ik)!! eP ik)!! eP ik! eP b∗ ikE(a∗ b∗ ikeP a∗ i ) = 1. If x ∈ P∞ such that xe = 0, then It follows thatPi E(wiew∗ Therefore, x = xE∞(ePi wiw∗ x → xe is injective for x ∈ P∞. Note that eeP e = bE(eP )e = (Index E)−1e. Then wiw∗ i i ) = 0. In other words, the map 0 = xe = xeXi i ) = E∞(xePi wiw∗ e = (Index E)bE∞(eP e) = (Index E)2bE∞(eP eeP e) = (Index E)2bE∞(E∞(e)eP e) = (Index E)E∞(e)e. Since P∞ ∋ x → xe injective, (Index E)E∞(e) = 1. (2): Suppose that E is approximately representable with a projection e ∈ P∞ ∩ P ′ satisfying that exe = E(x)e for any x in A. Let {u1, . . . , un} be a finite set of A such that (5) Define an element f in B∞ by uieu∗ i = 1. Xi f = nXi uieeP u∗ i . Since {(ui, u∗ Moreover, f commutes with eP . Therefore, f ∈ B∞ ∩ B ′. Since i )} is a quasi-basis for E, by [13, Remark 2.2 (4)] it commutes with A. bE∞(f ) = (Index E)−1 nXi=1 uieu∗ i = (Index E)−1, 12 HYUN HO LEE AND HIROYUKI OSAKA it remains to show that the map B ∋ x → xf is injective. Suppose that xf = 0. We may assume x is of the form aeP b for some a, b in A. Then i i . eP buieP eu∗ i i = 0. Then xf = aeP bX uieeP u∗ = aXi =X aE(bui)eP eu∗ By taking bE∞, we havePi aE(bui)eu∗ 0 =Xi =Xi = xXi i ) =Xi ψ(Xi aE(bui)eP u∗ i aE(bui)eP u∗ aeP buieP u∗ i uieP u∗ i = x. where ψ is the map in Proposition 3.6. By the injectivity of ψ, aE(bui)eu∗ i = 0 Conversely, suppose that bE has the Rokhlin property with a Rokhlin projection f in B∞ ∩ B ′. Define an element e in A∞ by e = (Index E)bE∞(f eP ). Then we can check that e is a projection in P∞ ∩ P ′ and satisfy the covariance relation for E. Also, it is easy to show that P ∋ x → xe is injective. For details, see [13, Proposition 3.4]. Now take a quasi-basis {(ui, u∗ uieu∗ i )} for E. Then Xi i i =Xi ui(Index E)bE∞(f eP )u∗ = (Index E)XbE∞(uif eP u∗ = (Index E)bE∞(f (Xi = (Index E)bE∞(f ) = 1. uieP u∗ i ) i )) (cid:3) Finally we are ready to give a reformulation of approximate representability in term of inclusion of C ∗-algebras. Proposition 3.17. Let G be a finite abelian group and A a unital simple C ∗-algebra. Suppose that α : G y A be an action such that the crossed product A ⋊α G is simple. Then α is approximately representable if and only if the conditional expectation E : A → Aα defined by E(a) = αg(a) is approximately representable. 1 GXg∈G Proof. Note that Aα is stably isomorphic to the cross product A⋊α G, where the latter is simple. It follows that α is saturated, thus E is of finite index and Index E = G. Since the kernel of the map Aα ∋ x → xe is trivial, it is injective. Now if α is A TRACIALLY SEQUENTIALLY-SPLIT ∗-HOMOMORPHISMS BETWEEN C ∗-ALGEBRAS II 13 approximately representable, then bα has the Rokhlin property by Theorem 3.14. Thus the conditional expectation from A ⋊α G to A has the Rokhlin property by Proposition 3.12. But this is a dual inclusion of Aα ⊂ A so that the inclusion Aα ⊂ A is approximately representable by Theorem 3.16. The converse follows by the same reverse argument. (cid:3) 4. Dualities of actions and inclusions:The tracial case In this section we recall tracial versions of Rokhlin property and approximate rep- resentability for finite group actions and presents tracial versions of such notions for inclusions of C ∗-algebras with a finite index and establish a duality result between two such notions. 4.1. Actions on C ∗-algebras. N.C. Phillips defined a tracial version of the Rokhlin property of a finite group action in [15]. Definition 4.1. (Phillips[15]) Let G be a finite group and A be an infinite dimensional separable unital C ∗-algebra. We say that α : G y A has the tracial Rokhlin property if for every finite set F ⊂ A, every ǫ > 0, any nonzero positive element x ∈ A there exist {eg}g∈G mutually orthogonal projections such that (1) kαg(eh) − eghk ≤ ǫ, (2) kega − aegk ≤ ǫ, ∀g, h ∈ g, ∀g ∈ G, ∀a ∈ F , (3) Write e =Pg eg, and 1− e is Murray-von Neumann equivalent to a projection in xAx. Then we can reformulate the tracial Rokhlin property of α : G y A in terms of exact relations using the central sequence algebra. Theorem 4.2. Let G be a finite group and A be a separable unital C ∗-algebra. Then α : G y A has the tracial Rokhlin property if and only if for any nonzero positive element x ∈ A∞ there exist a mutually orthogonal projections eg's in A∞ ∩ A′ such that (1) α∞,g(eh) = egh, (2) 1 −Pg eg is Murray-von Neumann equivalent to a projection in xA∞x. We recall that the strict Rokhlin property of α : G y A even for a compact group ∀g, h ∈ G where α∞ : G y A∞ is the induced action, G can be rephrased as follows. Definition 4.3. (cf. [1], [6]) Let A be a separable unital C ∗-algebra and G a second countable, compact group. Let σ : G y C(G) denote the canonical G-shift, that is, σg(f ) = f (g−1·) for all f ∈ C(G) and g ∈ G. A continuous action α : G y A is said to have the Rokhlin property if there exists a unital equivariant ∗-homomorphism (C(G), σ) → (A∞ ∩ A′, α∞). In the following we demonstrate the same perspective still holds in the case of tracial Rokhlin action of a finite group. Theorem 4.4. Let G be a finite group and A a separable unital infinite dimen- sional C ∗-algebra. Then α has the tracial Rokhlin property if and only if for every nonzero positive element x in A∞ there exists an equivariant ∗-homomorphism φ from 14 HYUN HO LEE AND HIROYUKI OSAKA (C(G), σ) to (A∞∩ A′, α∞) such that 1− φ(1C(G)) is Murray-von Neumann equivalent to a projection in a hereditary C ∗-subalgebra generated by x in A∞. Proof. "=⇒": Based on Theorem 4.2, for any nonzero positive x ∈ A∞ we can take mutually orthogonal projections eg's in A∞ ∩ A′ such that 1 −P eg - x. Then we define φ(f ) =Pg f (g)eg for f ∈ C(G). Then 1 − φ(1C(G)) = 1 −Pg eg - x . Using the condition (1) in Theorem 4.2, it is easily shown that φ is equivariant. "⇐=": Let x be a nonzero positive element in A∞ and suppose that we have an equivariant ∗-homomorphism φ : (C(G), σ) → (A∞ ∩ A′, α∞). Let χg be the char- acteristic function on a singleton g. Then eg = φ(χg) is a projection in A∞ ∩ A′ α∞,g(eh) = α∞,g(φ(χh)) = φ(σg(χh)) = φ(χgh) = egh, so we are done. (cid:3) such that eg ⊥ eh for g 6= h and 1 −Pg eg = 1 − φ(1C(G)) - x. Moreover, Lemma 4.5. Let G be a finite group and both C and A unital C ∗-algebras. Let β : G y C and α : G y A be actions of G on C and A respectively. Suppose that for any positive nonzero element x in A∞ there exists an equivarinat ∗-homomorphism φ : (C, β) → (A∞ ∩ A′, α∞) such that 1 − φ(1C) is Murray-von Neumann equivalent to a projection in xA∞x. Then idA ⊗1C : (A, α) → (A ⊗ C, α ⊗ β) is G-tracially sequentially-split. Moreover, the converse is true. Proof. The proof is a straightforward generalization from the nonequivariant case Proposition 2.4. (cid:3) Corollary 4.6. Let G be a finite group and α : G y A an action of G on a simple unital infinite dimensional C ∗-algebra A. Then α has the tracial Rokhlin property if and only if the map 1C(G)⊗idA : (A, α) → (C(G)⊗A, σ⊗α) is G-tracially sequentially split. Proof. It follows from Theorem 4.4 and Lemma 4.5. (cid:3) representation for the action α so that we write an element of A ⋊α G asPg∈G agλα We denote by φ ⋊ G a map from A ⋊α G to B ⋊β G as a natural extension of an equivariant ∗-homomorphism φ : (A, α) → (B, β) where α : G y A and β : G y B. In the following, we denote by λα : G → U(A ⋊α G) the implementing unitary g . The embedding of A into A ⋊α G is defined by a 7→ aλα e or just a without confusion. Lemma 4.7. [15, Proposition 1.12] Let A be an infinite dimensional simple unital C ∗-algebra with the property (SP), and α : G y A be an action of a finite group G on A such that A ⋊α G is also simple. Let B ⊂ A ⋊α G be a nonzero hereditary C ∗-subalgebra. Then there exists a nonzero projection p ∈ A which is Murray-von Neumann equivalent to a projection in B in A ⋊α G. Definition 4.8. Let G be a finite abelian group and A be an infinite dimensional unital separable simple C ∗-algebra. We say α : G y A is tracially approximately representable if for every positive nonzero element z in A∞, there are a projection e in A∞ ∩ A′ and a unitary representation ω : G → eA∞e such that (1) ag(eae) = ωg(eae)ω∗ (2) α∞,g(ωh) = ωh for all g, h ∈ G, (3) 1 − e is Murray-von Neumann equivalent to a projection zA∞z. g in A∞, A TRACIALLY SEQUENTIALLY-SPLIT ∗-HOMOMORPHISMS BETWEEN C ∗-ALGEBRAS II 15 Remark 4.9. Of course, in this case also we have a dichotomy that A has the property (SP) or α is approximately representable. Proposition 4.10. Let G be a finite abelian group and A an infinite dimensional unital separable simple C ∗-algebra. Then α : G y A is tracially approximately repre- sentable if and only if for every nonzero positive element z in A∞ there are a projection e ∈ A∞ ∩ A′ and an equivariant ∗-homomorphism ψ : (A ⋊α G, Ad λα) → (A∞, α∞) such that g ) = ωg, (1) ψ(ιA(a)) = ae for all a ∈ A, ψ(λα (2) 1−ψ(1A) = 1−e is Murray-von Neumann equivalent to a projection in zA∞z. Proof. Suppose that α : G y A is tracially approximately representable. Choose a nonzero positive element z in A∞. Then we have a projection e ∈ A∞ ∩ A′ and a unitary representation ω : G → eA∞e satisfying two conditions as in Definition 4.17. We consider the map φ : A ∋ x → xe ∈ A∞. Then it is easily checked that (φ, ω) gives a covariant pair for (A, α) so that it induces a map ψ from A ⋊α G to A∞ such that ψ(λα g ) = ωg and ψ(ιA(a)) = ae for all a in A. Since 1−e is Murray-von Neumann equivalent to a projection in zA∞z, so is 1 − φ(1A). The equivariantness of ψ follows from α∞,h(cid:0)ψ(aλα g )(cid:1) = α∞,h(aωg) = αh(a)ωg = ψ(Ad(λα Conversely, for any nonzero positive element z in A∞ we have a projection e ∈ A∞∩A′ and an equivariant map ψ : (A ⋊α G, Ad λα) → (A∞, α∞) such that (1) ψ(ιA(a)) = ae for all a ∈ A, ψ(λα (2) 1−ψ(1A) = 1−e is Murray-von Neumann equivalent to a projection in zA∞z. h) for h ∈ G. Note that ωh ∈ U(eA∞e). The equivariantness of ψ (cid:3) We put ωh = ψ(λα implies that αg(eae) = ωgaeω∗ g ) = ωg, h)(aλα g )). g for any a ∈ A. Corollary 4.11. Let G be a finite abelian group and A be a unital separable simple C ∗-algebra. Then α : G y A is tracially approximately representable if and only if ιA : (A, α) → (A ⋊α G, Ad(λα)) is G-tracially sequentially-split. Lemma 4.12. Let G be a finite group and α : G y A and β : G y B two actions on unital separable simple infinite dimensional C ∗-algebras A and B respectively. Then φ : (A, α) → (B, β) is G-tracially sequentially-split if and only if φ ⊗ idMn : (A⊗ Mn, α⊗ ρ) → (B ⊗ Mn, β ⊗ ρ) is G-tracially sequentially-split for n = G. Here, in fact, K(l2(G)) = Mn. Proof. Since the strict case follows from [1, Proposition 3.14], we may assume that A has the property (SP). Note that A∞ ⊗ Mn ∼= (A⊗ Mn)∞ by the map [(an)n]⊗ eij 7→ [(an ⊗ eij)n]. Suppose that φ : (A, α) → (B, β) is G-tracially sequentially-split. Consider a nonzero positive element z in (A ⊗ Mn)∞. Since (A ⊗ Mn) is also simple and has the property (SP), there is a projection p′ of the form p⊗ 1 in A∞ ⊗ Mn such that p′ is Murray-von Neumann equivalent to a projection in z(A ⊗ Mn)∞z (see [15, Lemma 1.11]). Then we take a tracial approximate left inverse ψ : (B, β) → (A∞, α∞) and a projection e ∈ A∞ ∩ A′ such that (1) ψ(φ(a)) = ae, (2) 1 − e is Murray-von Neumann equivalent to a projection p′A∞p′. 16 HYUN HO LEE AND HIROYUKI OSAKA Then 1A∞⊗Mn − (e ⊗ 1) = (1 − e) ⊗ 1 . p′ . z. Note that e ⊗ 1 is in (A ⊗ Mn)∞ ∩ (A⊗ Mn)′ and 1A∞⊗Mn − ψ⊗ idMn(1) is Murray-von Neumann equivalent a projection in z(A ⊗ Mn)∞z. Also, ψ ⊗ idMn(φ ⊗ idMn(a ⊗ eij)) = ae ⊗ eij = (a ⊗ eij)(e ⊗ 1). So φ ⊗ idMn is G-tracially sequentially-split. Conversely, suppose that φ ⊗ idMn is G-tracially sequentially-split. Take any nonzero positive element z ∈ A∞, and consider a tracial approximate left inverse eψ cor- responding to z ⊗ e11. Then we define ψ : B → A∞ by the restriction of eψ to It follows that eψ(1) = g ⊗ 1 where g ∈ A∞ ∩ A′. Since 1 − eψ(1) = (1 − g) ⊗ 1 is Murray-von Neumann equivalent B ⊗ 1. Since (A∞ ⊗ Mn) ∩ (A ⊗ Mn)′ ⊂ (A∞ ⊗ Mn) ∩ (1 ⊗ Mn)′ = A∞ ⊗ 1, (A∞ ⊗ Mn) ∩ (A ⊗ Mn)′ = (A∞ ∩ A′) ⊗ 1. to a projection in zA∞z ⊗ e11Mne11, 1 − g is Murray-von Neumann equivalent to a projection in zA∞z. Also, we see that by viewing A = A ⊗ 1 Thus, ψ is a G-tracial approximate left inverse for φ corresponding to z. ψ(φ(a)) = ψ(φ(a) ⊗ 1) = eψ((φ(a) ⊗ 1) = (a ⊗ 1)(g ⊗ 1) = ag. We are ready to prove the following duality result for G-tracially sequentially-split maps in the case of G a finite abelian group as one of our main results. Theorem 4.13. Let G be a finite abelian group and A and B infinite dimensional unital separable simple C ∗-algebras where α and β acts on respectively. Further we assume that α : G y A is an action such that A ⋊α G is simple, in particular an outer action. Then the equivariant ∗-homomorphism φ : (A, α) → (B, β) is G- tracially sequentially-split if and only if bφ = φ ⋊ G : (A ⋊α G,bα) → (B ⋊β G,bβ) is bG-tracially sequentially-split. Proof. Since the strict case follows from [1, Proposition 3.14], we may assume that A has the property (SP). Suppose that an equivariant map φ : (A, α) → (B, β) is G- tracially sequentially-split and consider a nonzero positive element z in (A⋊αG)∞. By Lemma 4.21 there is a projection p in A∞ which is Murray-von Neumann equivalent to a projection r in z(A ⋊α G)∞z. Then we can take a tracial approximate left inverse ψ such that 1 − ψ(1) is Murray-von Neumann equivalent to a projection in pA∞p. Let us denote e by ψ(1). Then 1 − e . p ∼ r ∈ z(A ⋊α G)∞z. Thus 1 − e is Murray-von Neumann equivalent to a projection in z(A ⋊α G)∞z. From the equivariantness of ψ, e is invariant under the action of α∞ so that (cid:3) (ψ ⋊ G)(φ(a)λβ g ) = ψ(φ(a))λα∞ Moreover, via A∞ ֒→ A∞ ⋊α∞ G ֒→ (A ⋊α G)∞ g = (ae)λα∞ g = aλα∞ g e. 1 − (ψ ⋊ G)(1) = 1 − ψ(1) = 1 − e. Conversely, suppose that the equivariant map φ ⋊ G : (A ⋊α G,bα) → (B ⋊β G,bβ) is bG-tracially sequentially-split. Then by the above proof we know thatbbφ : (A ⋊α G ⋊ bα bG,bbα) → (B ⋊β G ⋊ bβ bG,bbβ) is G-tracially sequentially-spilt. Note that by Takai duality A TRACIALLY SEQUENTIALLY-SPLIT ∗-HOMOMORPHISMS BETWEEN C ∗-ALGEBRAS II 17 [18] there are equivariant isomorphisms, where ρ is the G-action on the algebra of compact operators K(l2(G)) = Mn induced by the right-regular representation, such that the following commutative diagram is commutative: κA : (A ⋊α G ⋊ bα bG,bbα) ∼= (A ⊗ Mn, α ⊗ ρ) κB : (B ⋊β G ⋊ bβ bG,bbβ) ∼= (B ⊗ Mn, β ⊗ ρ) / (B ⋊β G ⋊ bβ bG,bbβ) (A ⋊α G ⋊ bα bG,bbα) κB κA bbφ φ⊗idMn (A ⊗ Mn, α ⊗ ρ) / (B ⊗ Mn, β ⊗ ρ) Note that the isomorphism κA induces the isomorphism denoted by (κA)∞ between z in (A ⊗ Mn)∞ consider (κA)−1 ((A⋊αG⋊ bαbG)∞, (bbα)∞) and ((A⊗Mn)∞, (α⊗ρ)∞). Now for a positive nonzero element ∞ (z) = z in (A ⋊α G ⋊ bα bG)∞. Since bbφ is G-tracially sequentially-split, we have a tracial approximate left inverse ψ from (B ⋊β G ⋊ bβ bG,bbβ) to ((A ⋊α G ⋊ bα bG)∞, (bbα)∞) such that (1) ψ(bbφ(x)) = xg for x ∈ A ⋊α G ⋊ bαbG where g = ψ(1) ∈ (A ⋊α G ⋊ bαbG)∞ ∩ (A ⋊α G ⋊ bα bG)′, (2) 1−g is Murray-von Neumann equivalent to a projection r in z(A ⋊α G ⋊ bα bG)∞ z. Now consider the map ψ = (κA)∞ ◦ ψ ◦ (κB)−1. Then it is equivariant since all three maps are. For any a ∈ A, ( ψ ◦ (φ ⊗ idMn))(a ⊗ eij) = ((κA)∞ ◦ ψ ◦ κ−1 B ◦ (φ ⊗ idMn))(a ⊗ eeij) = ((κA)∞ ◦ ψ ◦bbφ ◦ κ−1 = (κA)∞(κ−1 A (a ⊗ eij)g) = (a ⊗ eij)((κA)∞(g)). A )(a ⊗ eij) Moreover, 1− (κA)∞(g) = (κA)∞(1− g) is Murray-von Neumann equivalent to a pro- jection (κA)∞(r) in z(A ⊗ Mn)∞z. It follows that φ⊗ idMn is G-tracially sequentially- split. By Lemma 4.12 we conclude that φ is G-tracially sequentially-split. (cid:3) Then we provide an alternative proof based on Theorem 4.13 and Takai duality for the following result of N.C. Phillips which is a tracial version of Izumi's result. Theorem 4.14 (N.C.Phillips). Let A be an infinite dimensional simple separable unital C*- algebra, and let α : G y A be an action of a finite abelian group G on A such that A ⋊α G is also simple. Then representable. (1) α has the tracial Rokhlin property if and only if bα is tracially approximately (2) α is tracially approximately representable if and only if bα has the tracial Rokhlin property.   /   / 18 HYUN HO LEE AND HIROYUKI OSAKA Proof. (1): Suppose that α has the tracial Rokhlin property. Then by Corollary 4.6 the map 1C(G)⊗idA : (A, α) → (C(G)⊗A, σ⊗α) is G-tracially sequentially-split. Thus z in (A ⋊α G)∞ there are a projection g in the sequence algebra of A ⋊α G and a corresponding equivariant tracial approximate inverse ψ from ((C(G) ⊗ A) ⋊σ⊗α Note that we have the following commutative diagram of equivariant maps (see [1, Theorem 4.13 implies that (1C(G)⊗idA)⋊G : (A⋊αG,bα) → ((C(G)⊗A)⋊σ⊗αG,\σ ⊗ α) is bG-tracially sequentially-split. This means that for every nonzero positive element G,\σ ⊗ α) to ((A ⋊α G)∞, (bα)∞). (A ⋊α G,bα) / ((C(G) ⊗ A) ⋊σ⊗α G,\σ ⊗ α) 3❤❤❤❤❤❤❤❤❤❤❤❤❤❤❤❤❤❤❤ Proposition 4.25]); (6) )❙❙❙❙❙❙❙❙❙❙❙❙❙❙❙ (1C(G) ⊗idA)⋊G ιA⋊αG φ ((A ⋊α G) ⋊ bα bG, Ad(λ bα)) Here φ is an isomorphism which comes from Takai-duality. Now consider a map eψ = ψ ◦ φ. Then for any b ∈ A ⋊α G (eψ ◦ ιA⋊αG)(b) = (ψ ◦ φ ◦ ιA⋊αG)(b) = (ψ ◦ (1C(G) ⊗ idA) ⋊ G)(b) = bg tracially approximately representable. Moreover, 1 − eψ(1) = 1 − ψ(1) = 1 − g is Murray-von Neumann equivalent to a projection in z(A ⋊α G)∞z. Therefore, we have shown that ιA⋊αG : (A ⋊α G,bα) → ((A ⋊α G) ⋊ bα bG, Ad(λ bα) is bG-tracially sequentially-split. Then by Corollary 4.11bα is Conversely, let B = A ⋊α G. If bα is tracially approximately representable, then ιB : (B,bα) → (B ⋊ bα bG, Ad(λ bα)) is bG-tracially sequentially-split. We also employ the (A ⋊α G,bα) / ((A ⋊α G) ⋊ bα bG, Ad(λ bα)) diagram (6) with κ = φ−1 as follows; (7) 3❤❤❤❤❤❤❤❤❤❤❤❤❤❤❤❤❤❤❤ *❚❚❚❚❚❚❚❚❚❚❚❚❚❚❚❚ (1C(G) ⊗idA)⋊G ιA⋊αG κ ((C(G) ⊗ A) ⋊σ⊗α G,\σ ⊗ α) Then using the same argument as before we can show that (1C(G) ⊗ idA) ⋊ G is (C(G) ⊗ A, σ ⊗ α) is G-tracially sequentially-split. Thus α has the tracial Rokhlin property by Corollary 4.6. (2): Suppose that α is the tracially approximately representable. Then we have the bG-tracially sequentially-split. By Theorem 4.13 the map 1C(G) ⊗ idA : (A, α) → ) / 3 * / 3 A TRACIALLY SEQUENTIALLY-SPLIT ∗-HOMOMORPHISMS BETWEEN C ∗-ALGEBRAS II 19 following diagram (8) (A, α) By the duality, /❴❴❴❴❴❴❴❴❴❴❴❴❴❴❴ (A∞, α∞) 'PPPPPPPPPPPP ιA tracially(cid:9) 6♠♠♠♠♠♠♠♠♠♠♠♠♠ ψ (A ⋊α G, Ad(λα)) ι ι /❴❴❴❴❴❴❴❴❴❴❴❴❴❴❴❴❴❴❴❴ (9) (A ⋊α G,bα) *❚❚❚❚❚❚❚❚❚❚❚❚❚❚❚ ιA⋊G 4✐✐✐✐✐✐✐✐✐✐✐✐✐✐✐✐ ψ⋊G ((A ⋊α G)∞, (bα)∞) tracially(cid:9) ((A ⋊α G) ⋊Ad(λα) G, \Ad λα) Then we also have the following diagram by [1, Proposition 4.26]; (A ⋊α G,bα) 1C( bG) ⊗idA⋊αG /❴❴❴❴❴❴❴❴❴❴❴❴❴❴❴❴❴❴❴❴❴❴❴❴ ι ιA⋊G +❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲ 3❤❤❤❤❤❤❤❤❤❤❤❤❤❤❤❤❤❤❤ φ tracially(cid:9) ((A ⋊α G) ⋊Ad(λα) G, \Ad λα) 4✐✐✐✐✐✐✐✐✐✐✐✐✐✐✐✐ ψ⋊G ((A ⋊α G)∞, (bα)∞) (C(bG) ⊗ (A ⋊α G), σ ⊗bα) This means that the second factor embedding 1C( bG)⊗idA⋊αG : (A⋊α G,bα) → (C(bG)⊗ (A ⋊α G), σ ⊗bα) is bG-tracially sequentially-split. By Corollary 4.6 bα has the tracial Conversely ifbα has the tracial Rokhlin property, then we have the following diagram; /❴❴❴❴❴❴❴❴❴❴❴❴❴❴❴❴❴❴❴ Rokhlin property. (10) ι (A ⋊α G,bα) )❚❚❚❚❚❚❚❚❚❚❚❚❚❚ ⊗idA⋊α G 1C( bG) tracially(cid:9) 4✐✐✐✐✐✐✐✐✐✐✐✐✐✐✐✐ ψ ((A ⋊α G)∞, (bα)∞) Then again (A ⋊α G,bα) ιA⋊G /❴❴❴❴❴❴❴❴❴❴❴❴❴❴❴❴❴❴❴❴❴❴❴❴ (C(bG) ⊗ (A ⋊α G), σ ⊗bα) +❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲❲ 3❤❤❤❤❤❤❤❤❤❤❤❤❤❤❤❤❤❤❤ ⊗idA⋊α G 1C( bG) φ−1 ι tracially(cid:9) (C(bG) ⊗ (A ⋊α G), σ ⊗bα) 4✐✐✐✐✐✐✐✐✐✐✐✐✐✐✐✐ ψ ((A ⋊α G)∞, (bα)∞) ((A ⋊α G) ⋊Ad(λα) G, \Ad λα), This means that the map ιA ⋊ G : (A ⋊α G,bα) → ((A ⋊α G) ⋊Ad(λα) G, \Ad λα) is bG-tracially sequentially-split. Then the duality implies that ιA : (A, α) → (A ⋊α G, Ad(λα)) is G-tracially sequentially-split. It follows from Corollary 4.11 that α is tracially approximately representable. (cid:3) ' / 6 * / 4   + / 4 3 ) / 4   + / 4 3 20 HYUN HO LEE AND HIROYUKI OSAKA 4.2. Inclusion of C ∗-algebras. Now we turn to consider inclusions of unital C ∗- algebras. Definition 4.15 (Osaka and Teruya [14]). Let P ⊂ A be an inclusion of unital C ∗- algebras such that a conditional expectation E : A → P has a finite index. We say E has the tracial Rokhlin property if for every positive element z ∈ P∞ there is a Rokhlin projection e ∈ A∞ ∩ A so that subalgebra of A∞ generated by z in A∞, (1) (Index E)E∞(e) = g is a projection, (2) 1 − g is Murray-von Neumann equivalent to a projection in the hereditary (3) A ∋ x → xe ∈ A∞ is injective. As we notice, the third condition is automatically satisfied when A is simple. As in the case of action with the tracial Rokhlin property, if P ⊂ A is an inclusion of C ∗- algebras and a conditional expectation E : A → P of index-finite type has the tracial Rokhlin property, or shortly P ⊂ A an inclusion with the tracial Rokhlin property, then either A has property(SP) or E has the Rokhlin property (see [14, Lemma 4.3]). Like in the strict case the following observation was obtained by the second author and T. Teruya in [14]. Proposition 4.16. [14, Proposition 4.6] Let G be a finite group, α an action of G on an infinite dimensional finite simple separable unital C ∗-algebra A, and E the conditional expectation defined by E(a) = αg(a). Then α has the tracial 1 GXg Rokhlin property if and only if E has the tracial Rokhlin property. We note that in this case Aα is strongly Morita equivalent to A ⋊α G, thus if an approximation property is preserved by the strong Morita equivalence, and if the inclusion Aα ⊂ A of finite index is tracially sequentially-split, then such an approximation property can be transferred to A ⋊α G from A when α has the tracial Rokhlin property. Definition 4.17. Let P ⊂ A be an inclusion of unital C ∗-algebras and E : A → P be a conditional expectation of index-finite type. A conditional expectation E is said to be tracially approximately representable if for every nonzero positive element z ∈ A∞ there exist a projection e ∈ P∞ ∩ P ′, a projection r ∈ A∞ ∩ A′, and a finite set {ui} ⊂ A such that (2) Pi uieu∗ (1) eae = E(a)e for all a ∈ A, i = r, and re = e = er, (3) the map P ∋ x 7→ xe is injective, (4) 1 − r is Murray-von Neumann equivalent to a projection in zA∞z in A∞. Proposition 4.18. Let P ⊂ A be an inclusion of unital C ∗-algebras and E : A → P be a conditional expectation of index-finite type. Suppose that E is tracially approx- imately representable. Then A has the property (SP) or E is approximately repre- sentable. Proof. If A does not have Property (SP), neither does A∞. Thus there is a nonzero positive element z in A∞ such that a hereditary subalgebra of A∞ generated by z does not have any nonzero projection. However, by the assumption there is a projection A TRACIALLY SEQUENTIALLY-SPLIT ∗-HOMOMORPHISMS BETWEEN C ∗-ALGEBRAS II 21 r ∈ A∞ ∩ A′ such that 1 − r is Murray-von Neumann equivalent to a projection in zA∞z. i = 1, so E is approximately representable. (cid:3) It follows that 1 − r = 0. Thus Pi uieu∗ We need some preparations to prove a main result of this section. Lemma 4.19. [11, Lemma 3.12] Let p, q be two projections in P∞ and e ∈ A∞ ∩ A′ be a projection such that (Index E)E∞(e) is a projection in P∞ ∩ P ′. If pe = ep and q . pe in A∞, then q . p in P∞ Lemma 4.20. [11, Theorem 3.13] Let P ⊂ A be inclusion of C ∗-algebras of index- finite type and A separable. Suppose E : A → P has the tracial Rokhlin property. Then for any nonzero positive element z ∈ P∞, there exists a projection e in a central sequence algebra of A such that (Index E)E∞(e) = g is a projection such that 1 − g is Murray-von Neumann equivalent to a projection in zP∞z in P∞. The following lemma is crucial as an analogous result of Lemma 4.7 in the case of inclusion of C ∗-algebras. Lemma 4.21. Let P ⊂ A be an inclusion of unital C ∗-algebras and E : A → P be a conditional expectation of index-finite type and of finite depth. Suppose A has the (SP)-property. Then for any nonzero projection p ∈ A there is a projection q in P such that q . p. Moreover, every non-zero hereditary C ∗-subalgebra of A has a projection which is Murray-von Neumann equivalent to some projection in P . Proof. In fact, the assumption satisfies the outer condition in the sense of Kishimoto; for any non-zero positive element x in A and an arbitrary positive number ǫ there is an element y in P such that ky∗(x − E(x))yk < ǫ, ky∗E(x)yk ≥ kE(x)k − ǫ. More precisely, we say E : A → P is outer if for any element x ∈ A with E(x) = 0 and any nonzero hereditary C ∗-subalgebra C of A, inf{kcxck c ∈ C +,kck = 1} = 0. See the proof of Theorem 2.1 in [12] for more details. (cid:3) Then we first show characterizations of the tracial Rokhlin property and tracial approximate representability for inclusions of unital C ∗-algebras as we have done in Section 3. Proposition 4.22. Let P ⊂ A be an inclusion of unital C ∗-algebras and E : A → P be a conditional expectation of index-finite type and of finite depth. Suppose further A is simple. Then E has the tracial Rokhlin property if and only if for every nonzero positive element z in P∞ there are a projection e ∈ A∞ ∩ A′ and an injective map β : A → P∞ such that (1) ae = β(a)e for all a ∈ A, (2) (Index E)eeP e = e, (3) ye = ze implies that y = z for all y, z ∈ P∞, (4) 1 − β(1) is Murray-von Neumann equivalent to a projection in zP∞z in P∞. 22 HYUN HO LEE AND HIROYUKI OSAKA Proof. Suppose that E has the tracial Rokhlin property. By Lemma 4.20, for any nonzero positive element z in P∞ we can take a projection e in A∞ ∩ A′ such that (Index E)E∞(e) = g is a projection in P∞ ∩ P ′ and 1 − g is equivalent to a projection in zP∞z. Then we define β(a) = (Index E)E∞(ae). Thus we see immediately that 1−β(1) = 1−g is Murray-von Neumann equivalent to a projection in zP∞z. The other conditions are verified as in the proof of Proposition 3.9. Conversely, consider a nonzero positive element z in A∞. Since we assume A has Property (SP), so does A∞. Then by Lemma 4.21 we have a projection q in P∞ which is Murray-von Neumann equivalent to a projection in the hereditary C ∗-subalgebra generated by z of A∞. Now for this q there are a projection e ∈ A∞ ∩ A′ and an injective map β : A → P∞ with properties (1)-(3) as above such that 1 − β(1) . q. By the similar arguments in Proposition 3.9, we see that β(a) = (Index E)E∞(ae) and 1 − (Index E)E∞(e) = 1 − β(1) . q . z in A∞. So we are done. (cid:3) Proposition 4.23. Let P ⊂ A be an inclusion of unital C ∗-algebras and E : A → P be a conditional expectation of index-finite type. E is tracially approximately repre- sentable if and only if for every nonzero positive element z ∈ A∞ there are an injective ∗-homomorphism from C ∗hA, ePi to A∞ and a projection r ∈ A∞ ∩ A′ such that (1) ψ(x) = xr for any x ∈ A, (2) ψ(eP ) ∈ P∞ ∩ P ′ and ψ(eP )r = rψ(eP ) = ψ(eP ), (3) 1 − r is Murray-von Neumann equivalent to a projection in zA∞z. Proof. Suppose that E is tracially approximately representable and consider a nonzero positive element z in A∞. Since there exists a projection e ∈ P∞ ∩ P ′ such that eae = E(a)e and the map x 7→ xe is injective for x ∈ P , the universal property of C ∗hA, ePi induces an injective ∗-homomorphism ψ from C ∗hA, ePi to A∞, in fact the image is generated by A and e, such that ψ(xeP y) = xey. If we consider {ui} such i = r is a projection in A∞ ∩ A′ and re = e = er, then for any a ∈ A uieu∗ i )ae =Xi uiE(u∗ i a)e. uiE(u∗ i a)eP ) thatPi uieu∗ Therefore ae = are = a(Xi uieu∗ i )e = (Xi ψ(aeP ) = ψ(Xi i a). Similarly, we can show that a =Pi E(aui)u∗ which implies that a =Pi uiE(u∗ 1 − ψ(Pi uieP u∗ i )} is a quasi-basis for E, and thus Pi uieP u∗ i . It follows that {(ui, u∗ i = 1. Then i ) = 1 − r is Murray-von Neumann equivalent to a projection in zA∞z, so we are done. Conversely, if for every nonzero positive element z in A∞ there are an injective map from ChA, ePi to A∞ and a projection r in A∞ ∩ A′ satisfying three conditions in the above. We let ψ(eP ) = e. Then for a quasi-basis {(ui, u∗ uireru∗ ueP u∗ i )} uieu∗ i = r. ψ(Xi i ) = (X uireu∗ i r) =Xi i =Xi A TRACIALLY SEQUENTIALLY-SPLIT ∗-HOMOMORPHISMS BETWEEN C ∗-ALGEBRAS II 23 Therefore 1−r is Murray-von Neumann equivalent to a projection in zA∞z. Moreover, since eP aeP = E(a)eP for any a ∈ A we have ψ(eP aeP ) = ψ(E(a)eP ). It follows that eae = E(a)e. (cid:3) We now use the above technical lemmas to prove our main result; we also derive a consequence of it. Theorem 4.24. Let P ⊂ A be an inclusion of unital C ∗-algebras and E : A → P be a conditional expectation of index-finite type and of finite depth. If we denote by B representable. Rokhlin property. Suppose that A is simple. Then the basic construction for E, then we have a dual conditional expectation bE : B → A. (1) E has the tracial Rokhlin property if and only if bE is tracially approximately (2) E is tracially approximately representable if and only if bE has the tracial Proof. (1): Assume that E has the tracial Rokhlin property and A has the property (SP). Let z be a nonzero positive element in B∞. By Lemma ??, there is a projection p in P∞ which is Murray-von Neumann equivalent to a projection in zB∞z. Let {(vi, v∗ i ) i = 1, . . . , n} be a quasi-basis for E. Then we take mutually equivalent orthogonal projections r1, . . . , rn in pP∞p since P has the property (SP). For one of such projections, we can take a Rokhlin projection e ∈ A∞ ∩ A′ such that 1 − (Index E)E∞(e) . ri for i = 1, . . . , n. Now let ui = √Index EvieP and g = (Index E)E∞(e). Then we can easily see that nXi=1 uieu∗ i =Xi vigepv∗ i . Thus for any x ∈ A Xi i x)e =Xi uibE(u∗ Similarly, we can show thatPi ebE(xui)u∗ Let r =Pi uieu∗ (Index E)viePbE(eP v∗ i . Since geP = eP g and g ∈ P∞ ∩ P ′, vjgepv∗ j vigeP v∗ i = ex. i x)e = Xi vieP v∗ i! xe = xe. iXj vigeP (v∗ i vj)eP gv∗ j vigE(v∗ i vj)eP gv∗ j viE(v∗ i vj)eP gv∗ j vjeP gv∗ j = r r2 =Xi =Xi,j =Xi,j =Xj Xi =Xj 24 HYUN HO LEE AND HIROYUKI OSAKA We verify that r ∈ B∞ ∩ B ′. Let a ∈ A. k! k! E(v∗ i avk)v∗ E(v∗ i avk)geP v∗ viE(v∗ i avk)geP v∗ k vigeP v∗ i a ra =Xi vigeP Xk =Xi vi Xk =Xi =Xk Xi =Xk reP =Xi =Xi eP r =Xi =Xi geP E(vi)v∗ eP vieP gv∗ avkgeP v∗ vigeP v∗ k = ar i eP =Xi i =Xi i = geP viE(v∗ i )eP g = geP (Index E)eeP e = e. vigE(v∗ i )eP E(vi)eP gv∗ i Using the same argument in [13, Proposition 3.4], we can show that It follows that for x, y ∈ A Now denote by {eij}n e(xeP y)e = xeeP ey = x(Index E)−1ey = (Index E)−1xye = bE(xeP y)e. i,j=1 the matrix units in Mn. In B∞ ⊗ Mn v∗ j ⊗ ej1] vi ⊗ e1i][Xk (1 − g)eP ⊗ ekk][Xj (1 − g)eP ⊗ ekk][Xj j ⊗ ej1][Xi v∗ vi ⊗ e1i][Xk (1 − g)eP ⊗ ekk] (1 − r) ⊗ e11 = [Xi ∼ [Xk .Xk .Xk ∼ (Xk (1 − g)eP ⊗ ekk rkeP ⊗ ekk rkep) ⊗ e11 . peP p ⊗ e11 . p ⊗ e11. Hence we conclude that 1 − r . p . z in B∞. A TRACIALLY SEQUENTIALLY-SPLIT ∗-HOMOMORPHISMS BETWEEN C ∗-ALGEBRAS II 25 Conversely, suppose that bE is tracially approximately representable. Take a posi- tive nonzero element z in A∞(⊂ B∞). By Lemma 4.21 there is a projection p in P∞ which is Murray-von Neumann equivalent to a projection zA∞z. For peP p ∈ B∞, we have a projection e ∈ A∞ ∩ A′, a projection r ∈ B∞ ∩ B ′, and a finite set {ui} ⊂ B such that (11) (12) (13) 1 − r is Murray-von Neumann equivalent to a projection in peppB∞peP p. From (11) we have (Index E)eeP e = e. Let wi = (Index E)bE(eP ui) ∈ A. Then (Index E)2eeP uieu∗ wiew∗ i eP e Xi i! eP e uieu∗ eze = bE(z)e ∀z ∈ B, i = r, re = e = er, uieu∗ Xi i! =Xi = (Index E)2eeP Xi = (Index E)2eeP reP e = (Index E)2ereP e = (Index E)2eeP e = (Index E)e. It follows thatPi ww∗ i ) = (Index E)E∞(e) which will be denoted by g. Using the argument in the proof of 3.16 we can show that geP = reP . i = Index E since A ∋ x → xe is injective. Thus E∞(Pi wiew∗ By (13), there is a partial isometry v in B∞ such that vv∗ = 1 − r and vv∗ = q ∈ peppB∞peP p. Set w = vep. Then w∗w = (1 − r)eP and ww∗ = q0 ≤ q. Note that q0 ∈ pP∞eP p. Now let u = q0w(1 − r)eP . It is easily checked that u∗u = (1 − g)eP and uu∗ = q0. Note that u = pceP for some c ∈ P∞. Then consider bu∗bueP = c∗ppceP = eP c∗pceP = u∗u = (1 − g)eP . bu = (Index E)bE∞(u) ∈ A∞. This implies that bu is a partial isometry in A∞ such that bu∗bu = 1 − g and bubu∗ ≤ p. Since p is Murray-von Neumann equivalent to a projection to some projection in xA∞x, so is 1 − g. (2): Suppose E is tracially approximately representable. Take a nonzero positive element z in B∞. By Lemma 4.21, we have a projection p in A∞ such that p is Murray- von Neumann equivalent to a projection in zB∞z. For this p there is a projection e in P∞ ∩ P ′ such that (14) for any x ∈ A, a projection g in A∞ ∩ A′, and a finite set {ui} in A such that exe = E(x)e (15) Xi i = g,Xi uieu∗ uiE(u∗ i x)e = xe,Xi eE(xui)u∗ i = ex ∀x ∈ A 26 HYUN HO LEE AND HIROYUKI OSAKA (In fact, ge = e implies that gxe = (Pi uieu∗ Similarly, eg = e implies thatPi eE(xui)u∗ i )xe = Pi uiE(u∗ i = ex.) (16) 1 − g is Murry-von Neumann equivalent to a projection in pA∞p i x)e = xe for x ∈ A. Define f in B∞ by Then f is a projection since Note that for any a ∈ A uieeP u∗ f 2 =Xi =Xi,j =Xj Xi =Xj uiepeu∗ uieeP u∗ i a f a =Xi =Xi uieP Xj =Xi =Xj Xi =Xj aujeeP u∗ f =Xi iXj uieeP u∗ i . ujeeP u∗ j uiepeu∗ i ujeeP u∗ j uiE(u∗ j by (14) i uj)e! eP u∗ ujeeP u∗ j = f by (15). i a (eP eeP = E∞(e)EP = eeP ) j! by (15) by (15) eE(u∗ i auj)u∗ uiE(u∗ i auj)e! eP u∗ j j = af. Moreover, f eP = Xi =Xi =Xi = eeP uieeP u∗ i! eP uieE(u∗ i )eP (uiE(u∗ i )e) ep by (15) A TRACIALLY SEQUENTIALLY-SPLIT ∗-HOMOMORPHISMS BETWEEN C ∗-ALGEBRAS II 27 and uieP eu∗ i! eP uieP eu∗ i E(ui)eP eu∗ i eP eE(ui)u∗ i eP f = eP Xi =Xi =Xi =Xi = eP e by (15) Thus we showed that f ∈ B∞ ∩ B ′. Finally bE∞(f ) = (Index E)−1 Xi i! . uieu∗ Therefore (Index E)bE∞(f ) = g is a projection such that 1− g is Murry-von Neumann Conversely, suppose that bE has the tracial Rokhlin property. For any positive equivalent to a projection in zB∞z since 1 − g . p . z. nonzero element z ∈ A∞ we consider a hereditary subalgebra zA∞z. Since A has the property (SP), so is A∞. Thus there is a projection p in zA∞z which is also contained in zB∞z. Then we consider mutually orthogonal nonzero projections r1, r2 in pA∞p such that r1 . r2. Then we can think of a Rokhlin projection f ′ for r2. For f ′r1 we can take a Rokhlin projection f ∈ B∞ ∩ B ′ such that (Index E)bE∞(f ) is a projection such that 1− (Index E)bE∞(f ) is Cuntz subequivalent to f ′r1. Then define an element e = (Index E)bE∞(f eP ) in A∞.Using the fact that f eP = epf = (Index E)bE∞(f eP )eP , we can show that e is a projection. By the same argument in [13, Proposition 3.4], we can show that e is in P∞∩ P ′ and eae = E(a)e for any a ∈ A. Then take a quasi-basis {(ui, u∗ i )} for E. Then uieu∗ Xi i uibE∞(f eP )u∗ uif eP u∗ i ) i = (Index E)Xi = (Index E)bE∞(Xi = (Index E)bE∞(f ) Thus g =Pi uieu∗ i is a projection in A∞ ∩ A′. By applying Lemma 4.19 to f ,1 − g, and f ′r1, we have 1 − g is Cuntz subequivalent to r2 in A∞. So 1 − g . r2 . p . z in A∞ . (cid:3) Corollary 4.25. Let G be a finite abelian group, α an outer action of G on an infinite dimensional simple separable unital C ∗-algebra A such that A ⋊α G is simple, and E as in Proposition 3.12. Then α is tracially approximately representable if and only if E is tracially approximate representable. 28 HYUN HO LEE AND HIROYUKI OSAKA Proof. The proof goes exactly same as the proof of Proposition 3.17 only replacing Theorem 3.14, Proposition 3.12, Theorem 3.16 with Theorem 4.14, Proposition 4.16, Theorem 4.24 respectively. (cid:3) 5. Acknowledgement This research was carried out during the first author's stay at KIAS and his visit to Ritsumeikan University. He would like to appreciate both institutions for excel- lent supports. The second author also would like to appreciate KfAS for the worm hospitality when he visited it. References 99(2016), [1] S. Barlak, G. Sazab´o, Sequentially split ∗-homomorphisms between C ∗-algebras, Int. J. Math. [2] J. Cuntz Dimension functions on simple C ∗-algebras, Math. Ann. 233(1978), 145 -- 153., [3] G. A. Elliott and Z. Niu, On tracial approximation, J. Funct. Anal. 254(2008), 396 - 440. [4] G. A. Elliott and A. Toms, Regularity properties in the classification program for separable amenable C ∗-algebras, Bull. Amer. Math. Soc. 45(2008), no. 2., 229 -- 245. [5] I. Hirshberg and J. Orovitz, Tracially Z-absorbing C ∗-algebras, J. Funct. Anal. 265(2013), 765 [6] I. Hirshberg, W. Winter Rokhlin actions and self-absorbing C-algebras, Pacific J. Math. 233 -- 785. (2007), no. 1, pp. 125143. [7] M. Izumi, Finite group actions on C-algebras with the Rokhlin property I, Duke Math. J. 122 (2004), no. 2, 233 -- 280. [8] J. Jeong, K. Park Saturated actions by finite-dimensional Hopf *-algebras on C ∗-algebras, In- ternational J. Math. 19(2008), no.2, 125 -- 144 [9] H. Lin, Tracial AF C*-algebras, Trans. Amer. Math. Soc. 353(2001), 693 -- 722. [10] H. Lin, Classification of simple C ∗-algebras of tracial topological rank zero Duke Math. J. 125(2004), 91 -- 119. arXiv:1707.07377 [11] H. Lee, H. Osaka, Tracially sequentially-split ∗-homomorphisms between C ∗-algebras preprint [12] H. Osaka (SP)-property for a pair of C ∗-algebras J. Operator Theory46(2001), 159 -- 171 [13] H. Osaka, K. Kodaka, T. Teruya, Rokhlin property for inclusion of C ∗-algebras with a finite Watatani index Operator structures and dynamical systems, Contemp.Math. 503(2009), Amer. Math. Soc. 177 -- 195 [14] H. Osaka, T. Teruya, The Jiang-Su absorption for inclusions of C ∗-algebras, to appear in Canad. J. Math, arXiv: 1404.7663. [15] N. C. Phillips, The tracial Rokhlin property for actions of finite groups on C ∗-algebras Amer. J. Math. 133(2011), no. 3, 581 -- 636, arXiv:math.OA/0609782. [16] M. Rørdam, On the structure of simple C*-algebras tensored with a UHF-algebra, J. Funct. Anal. 100(1991), 1 -- 17. [17] M. Rørdam, On the structure of simple C*-algebras tensored with a UHF-algebra II, J. Funct. Anal. 107(1992), 255 - 269. [18] H. Takai On a duality for crossed products of C-algebras, J. Funct. Anal. 19 (1975), 2539. [19] A. Toms, W. Winter, Strongly self-absorbing C ∗-algebras, Trans. A. M. S. 359(2007), no.8, 3999-4029. [20] Y. Watatani, Index for C ∗-subalgebras, Mem. Amer. Math. Soc. 424, Amer. Math. Soc., Prov- idence, R. I., (1990). Department of Mathematics, University of Ulsan, Ulsan, South Korea 44610, and, School of Mathematics, Korea Institute of Advanced Study, Seoul, South Korea, 130-772 E-mail address: [email protected] A TRACIALLY SEQUENTIALLY-SPLIT ∗-HOMOMORPHISMS BETWEEN C ∗-ALGEBRAS II 29 Department of Mathematical Sciences, Ritsumeikan University, Kusatsu, Shiga 525-8577, Japan E-mail address: [email protected]
1207.2560
2
1207
2013-05-28T10:25:23
Deformation of operator algebras by Borel cocycles
[ "math.OA", "math.QA" ]
Assume that we are given a coaction \delta of a locally compact group G on a C*-algebra A and a T-valued Borel 2-cocycle \omega on G. Motivated by the approach of Kasprzak to Rieffel's deformation we define a deformation A_\omega of A. Among other properties of A_\omega we show that A_\omega\otimes K(L^2(G)) is canonically isomorphic to A\rtimes_\delta\hat G\rtimes_{\hat\delta,\omega}G. This, together with a slight extension of a result of Echterhoff et al., implies that for groups satisfying the Baum-Connes conjecture the K-theory of A_\omega remains invariant under homotopies of omega.
math.OA
math
DEFORMATION OF OPERATOR ALGEBRAS BY BOREL COCYCLES JYOTISHMAN BHOWMICK, SERGEY NESHVEYEV, AND AMANDIP SANGHA Abstract. Assume that we are given a coaction δ of a locally compact group G on a C∗-algebra A and a T-valued Borel 2-cocycle ω on G. Motivated by the approach of Kasprzak to Rieffel's deforma- tion we define a deformation Aω of A. Among other properties of Aω we show that Aω ⊗ K(L2(G)) is canonically isomorphic to A ⋊δ G ⋊δ,ω G. This, together with a slight extension of a result of Echterhoff et al., implies that for groups satisfying the Baum-Connes conjecture the K-theory of Aω remains invariant under homotopies of ω. Introduction Assume G is a discrete group and A = ⊕g∈GAg is a G-graded algebra. Given a C∗-valued 2- cocycle ω on G we can define a new product on A by the formula ag · ah = ω(g, h)agah for ag ∈ Ag and ah ∈ Ah. Some of the well-known examples of C∗-algebras, such as irrational rotation algebras and, more generally, twisted group C∗-algebras or twisted crossed products, are operator algebraic variants of this construction. Nevertheless the question what this construction means for a general C∗-algebra A and a locally compact group G has no obvious answer. A natural replacement of a G-grading is a coaction of G on A. But then the subspaces Ag are often trivial for non-discrete G and it is not clear how to define the new product. In [9] Rieffel succeeded in defining the product in the case G = Rd using oscillatory integrals. A few years ago Kasprzak [4] proposed an alternative approach that works for any locally compact group G and a continuous T-valued 2-cocycle ω. In fact, he considered only abelian groups and, correspondingly, actions of G rather than coactions of G, but it is easy to see that his construction makes sense for arbitrary G. It should also be mentioned that for discrete groups a different, but equivalent, approach has been recently suggested by Yamashita [14]. Kasprzak's idea is as follows. Given a coaction δ of G on A, consider the dual action δ of G on A ⋊δ G. Using the cocycle ω we can deform this action to a new action δω. Then by general results on crossed products it turns out that A ⋊δ G has another crossed product decomposition Aω ⋊δω G such that δω becomes dual to δω. The C∗-algebra Aω is the deformation of A we are looking for. The goal of this note is to define Aω for arbitrary Borel cocycles ω. For abelian groups, restricting to continuous cocycles is not a serious omission, essentially since Borel cocycles correspond to Borel bicharacters and these are automatically continuous. But for general groups the class of continuous cocycles is too small and the right class is that of Borel cocycles [6, 7]. Given a Borel cocycle ω, there are no obvious reasons for the twisted dual action δω to be well-defined on A ⋊δ G. What started this work is the observation that δω is well-defined for dual coactions. Since any coaction is stably exterior equivalent to a dual coaction, and it is natural to expect that exterior equivalent coactions produce strongly Morita equivalent deformations, this suggested that Aω could be defined for arbitrary δ. In the end, though, we found it easier to relate Aω to twisted crossed products rather than to use dual coactions. This simplifies proofs, but the fundamental reasons for why Aω is well-defined become somewhat hidden. Our deformed algebras Aω enjoy a number of expected properties. In particular, they come with canonical coactions δω. However, the isomorphism A ⋊δ G ∼= Aω ⋊δω G, which played an important Date: July 11, 2012; minor corrections April 19, 2013. Supported by the Research Council of Norway. 1 2 J. BHOWMICK, S. NESHVEYEV, AND A. SANGHA role in [4] and [14], is no longer available for general cocycles. isomorphism Aω ⊗ K(L2(G)) ∼= A ⋊δ G ⋊δ,ω G, which is equally well suited for studying Aω. Instead we construct an explicit Let us finally say a few words about sources of examples of coactions. The easiest is, of course, to take the dual coaction on a crossed product A = B ⋊α G. In this case the deformation produces the twisted crossed product B ⋊α,ω G, as expected. But even if we start with dual coactions, we can get new coactions by taking e.g. free products. Given a corepresentation of the dual quantum group G, we can also consider infinite tensor products, as well as free Araki-Woods factors, see [12] and references therein. Acknowledgement. It is our pleasure to thank Pawe l Kasprzak and Makoto Yamashita for the inspiring correspondence. 1. Actions, coactions and crossed products In this preliminary section we fix our notation and list a number of facts that we will freely use later. Let G be a second countable locally compact group. Fix a left-invariant Haar measure on G. Denote by λ and ρ the left and right regular representations on G. We will usually identify the reduced group C∗-algebra C ∗ r (G) with its image under λ. Similarly, we will usually identify C0(G) with the algebra of operators of multiplication by functions on L2(G). Denote by K the algebra of compact operators on L2(G). Denote by ∆ : C0(G) → M (C0(G) ⊗ C0(G)) = Cb(G × G) and ∆ : C ∗ r (G)) r (G) → M (C ∗ r (G) ⊗ C ∗ the standard comultiplications, so ∆(f )(g, h) = f (gh), ∆(λg) = λg ⊗ λg. Let W ∈ M (C0(G) ⊗ C ∗ r (G)) be the fundamental unitary, defined by (W ξ)(s, t) = ξ(s, s−1t) for ξ ∈ L2(G × G). In other words, if we identify M (C0(G) ⊗ C ∗ maps G → M (C ∗ r (G)), then W (g) = λg. We have r (G)) with the algebra of bounded strictly continuous W ∗(1 ⊗ f )W = ∆(f ) for f ∈ C0(G), We will also use the unitary V = (ρ ⊗ ι)(W21) ∈ M (ρ(C ∗ W (λg ⊗ 1)W ∗ = ∆(λg) and W ∗(ρg ⊗ 1)W = ρg ⊗ λg for g ∈ G. r (G)) ⊗ C0(G)). We have V (f ⊗ 1)V ∗ = ∆(f ) for f ∈ C0(G), V ∗(1 ⊗ ρg)V = ρg ⊗ ρg and V (1 ⊗ λg)V ∗ = ρg ⊗ λg for g ∈ G. Assume now that α : G → Aut(B) is a (continuous) action of G on a C∗-algebra B. We consider α as a homomorphism α : B → M (B ⊗ C0(G)), so that α(b)(g) = αg(b). Then (α ⊗ ι)α = (ι ⊗ ∆)α. We define the reduced crossed product B ⋊α G by B ⋊α G = α(B)(1 ⊗ ρ(C ∗ r (G))) ⊂ M (B ⊗ K). This is equivalent to the standard definition. Since we consider only reduced crossed products in this paper, we omit r in the notation. By a coaction of G on a C∗-algebra A we mean a non-degenerate injective homomorphism δ : A → r (G)) is a dense subspace r (G)) such that (δ ⊗ ι)δ = (ι ⊗ ∆)δ and the space δ(A)(1 ⊗ C ∗ r (G). The crossed product is then defined by M (A ⊗ C ∗ of A ⊗ C ∗ A ⋊δ G = δ(A)(1 ⊗ C0(G)) ⊂ M (A ⊗ K). DEFORMATION OF OPERATOR ALGEBRAS BY BOREL COCYCLES 3 The algebra A ⋊δ G is equipped with the dual action δ of G defined by δg = Ad(1 ⊗ ρg). Thinking of δ as a homomorphism A ⋊δ G → M ((A ⋊δ G) ⊗ C0(G)), we have δ(δ(a)) = δ(a) ⊗ 1, δ(1 ⊗ f ) = 1 ⊗ ∆(f ). It follows that δ(x) = V23(x ⊗ 1)V ∗ 23 for x ∈ A ⋊δ G ⊂ M (A ⊗ K). Similarly, starting with an action α of G on B we get a dual coaction α of G on B ⋊α G such that Therefore α(α(b)) = α(b) ⊗ 1, α(1 ⊗ ρg) = 1 ⊗ ρg ⊗ λg. α(x) = W ∗ 23(x ⊗ 1)W23 for x ∈ B ⋊α G ⊂ M (B ⊗ K). A 1-cocycle for an action α of G on B is a strictly continuous family U = {ug}g∈G of unitaries in M (B) such that ugh = ugαg(uh). Given such a cocycle, we can define a new action αU of G on B by αU,g = ugαg(·)u∗ g. The actions α and αU are called exterior equivalent. We have an isomorphism B ⋊α G ∼= B ⋊αU G respecting the dual coactions, defined by α(b) 7→ αU (b), 1 ⊗ ρg 7→ αU (u∗ g)(1 ⊗ ρg). If we think of U as an element of M (B ⊗ C0(G)), then this isomorphism is implemented by the inner automorphism Ad U of M (B ⊗ K). Similarly, a 1-cocycle for a coaction δ of G on A is a unitary U ∈ M (A ⊗ C ∗ r (G)) such that (ι ⊗ ∆)(U ) = (U ⊗ 1)(δ ⊗ ι)(U ). Given such a cocycle, we can define a new coaction δU by δU (a) = U δ(a)U ∗. The coactions δ and δU are called exterior equivalent. The inner automorphism Ad U of M (A ⊗ K) defines an isomorphism of A ⋊δ G onto A ⋊δU G respecting the dual actions, see [5, Theorem 2.9]. In particular, given a coaction δ of G on A we can consider the coaction a ⊗ T 7→ δ(a)13(1 ⊗ T ⊗ 1) of G on A ⊗ K, then take the 1-cocycle 1 ⊗ W ∗ for this coaction (the cocycle identity means that (ι ⊗ ∆)(W ) = W13W12) and get a new coaction on A ⊗ K. In order to lighten the notation we will denote this new coaction by δW ∗. Then the Takesaki-Takai(-Katayama-Baaj-Skandalis) duality states that Explicitly, the isomorphism is given by (A ⋊δ G ⋊δ G, δ) ∼= (A ⊗ K, δW ∗ ). δ(δ(a)) = δ(a) ⊗ 1 7→ δ(a), δ(1 ⊗ f ) = 1 ⊗ ∆(f ) 7→ 1 ⊗ f, 1 ⊗ 1 ⊗ ρg 7→ 1 ⊗ ρg. If we identify A ⊗ K with δ(A) ⊗ K ⊂ M (A ⊗ K ⊗ K), then this isomorphism is simply Ad W23. We finish this section by discussing how to recover A from A ⋊δ G for a coaction δ. Consider the homomorphism η : A ⋊δ G → M ((A ⋊δ G) ⊗ K) ⊂ M (A ⊗ K ⊗ K) defined by η(x) = W23δ(x)W ∗ with the Takesaki-Takai duality isomorphism A ⋊δ G ⋊δ G ∼= δ(A) ⊗ K. Explicitly, 23. In other words, η is the composition of δ : A ⋊δ G → M (A ⋊δ G ⋊δ G) η(δ(a)) = (δ ⊗ ι)δ(a), η(1 ⊗ f ) = 1 ⊗ f ∈ M ((A ⋊δ G) ⊗ K) From this we see that δ(A) ⊂ M (A ⋊δ G) is the closed linear span of elements of the form (ι ⊗ ϕ)η(x) with x ∈ A ⋊δ G and ϕ ∈ K ∗. More generally, assume we are given an action α of G on a C∗-algebra B and a nondegenerate homomorphism π : C0(G) → M (B) such that α(π(f )) = (π ⊗ ι)∆(f ). Put X = (π ⊗ ι)(W ) and consider the homomorphism η : B → M (B ⊗ K), η(x) = Xα(x)X ∗. Then by a Landstad-type result of Quigg [11, Theorem 3.3] and, more generally, Vaes [13, Theo- rem 6.7], the closed linear span A ⊂ M (B) of elements of the form (ι ⊗ ϕ)η(x), with x ∈ B and 4 J. BHOWMICK, S. NESHVEYEV, AND A. SANGHA ϕ ∈ K ∗, is a C∗-algebra, the formula δ(a) = X(a⊗ 1)X ∗ defines a coaction of G on A, and η becomes an isomorphism B ∼= A ⋊δ G that intertwines α with δ. Denote by Z 2(G; T) the set of T-valued Borel 2-cocycles on G, so ω ∈ Z 2(G; T) is a Borel function 2. Deformation of algebras G × G → T such that ω(g, h)ω(gh, k) = ω(g, hk)ω(h, k). For every cocycle ω consider also the cocycles ω and ¯ω defined by ω(g, h) = ω(h−1, g−1) and ¯ω(g, h) = ω(g, h). Define operators λω g and ρω g on L2(G) by ∗ Then λω g = ω(g−1, ·)λg, ρω g = ω(·, g)ρg. g λω λω h = ω(g, h)λω gh, ρω g ρω h = ω(g, h)ρω gh and [λω g , ρω h ] = 0 for all g, h ∈ G. Fix now a cocycle ω ∈ Z 2(G; T) and consider a coaction δ of G on a C∗-algebra A. Assume first that the cocycle ω is continuous. In this case the functions ω(·, g) belong to the multiplier algebra of C0(G), so we can define a new twisted dual action δω on A ⋊δ G by letting δω g ). In other words, if we consider ω as a multiplier of C0(G) ⊗ C0(G), then g = Ad(1 ⊗ ρω δω(x) = ω23δ(x)ω∗ 23 = ω23V23(x ⊗ 1)V ∗ 23 ω∗ 23 ∈ M (A ⊗ K ⊗ K). For f ∈ C0(G) we obviously have δω(1 ⊗ f ) = δ(1 ⊗ f ) = 1 ⊗ ∆(f ). By the Landstad-type duality result of Quigg and Vaes it follows that δω is the dual action on a crossed product Aω ⋊δω G for some C∗-subalgebra Aω ⊂ M (A ⋊δ G) ⊂ M (A ⊗ K) and a coaction δω of G, and this subalgebra is defined using slice maps applied to the image of A ⋊δ G under the homomorphism ηω : A ⋊δ G → M (A ⊗ K ⊗ K), ηω(x) = W23 ω23δ(x)ω∗ 23W ∗ 23. If the cocycle ω is only assumed to be Borel, it is not clear whether the action δω is well-defined. Nevertheless, the homomorphism ηω : A ⋊δ G → M (A ⊗ K ⊗ K) defined above still makes sense. Therefore we can give the following definition. Definition 2.1. The ω-deformation of a C∗-algebra A equipped with a coaction δ of G is the C∗-subalgebra Aω ⊂ M (A ⊗ K) generated by all elements of the form (ι ⊗ ι ⊗ ϕ)ηωδ(a) = (ι ⊗ ι ⊗ ϕ) Ad(W23 ω23)(δ(a) ⊗ 1), where a ∈ A and ϕ ∈ K ∗. In case we want to stress that the deformation is defined using the coaction δ, we will write Aδ,ω instead of Aω. Note that if we considered elements of the form (ι ⊗ ι ⊗ ϕ)ηω(x) for all x ∈ A ⋊δ G, this would not change the algebra Aω, since ηω(1 ⊗ f ) = 1 ⊗ 1 ⊗ f . In order to get an idea about the structure of Aω consider the C∗-algebra C ∗ r (G, ω) generated by operators of the form λω f = ZG f (g)λω g dg, f ∈ L1(G). g and ρ ω ∗The operators λω g are more commonly defined by λω g = ρgω(·, g−1) = ω(·g, g−1)ρg. With our definition some of the formulas will look better. If the cocycle ω satisfies ω(g, e) = ω(e, g) = ω(g, g−1) = 1 for all g ∈ G, then the two definitions coincide, that is to say ω(h−1, g) = ω(g, g−1h), which follows by applying the cocycle identity for ω to the triple (h−1, g, g−1h). Any cocycle is cohomologous to a cocycle satisfying the above normalization conditions, so in principle we could consider only such cocycles. g = λgω(g, ·) = ω(g, g−1 ·)λg and ρ ω DEFORMATION OF OPERATOR ALGEBRAS BY BOREL COCYCLES 5 When necessary we denote by λω the identity representation of C ∗ computation shows that r (G, ω) on L2(G). A simple The map g 7→ λω to the regular representation, so it defines a representation of C ∗ by λω ⊠ λ¯ω. We can then write (2.1) g therefore defines a representation of G on L2(G × G) that is quasi-equivalent r (G). Denote this representation W ω(λg ⊗ 1)ω∗W ∗ = λω g ⊗ λ¯ω g . g ⊗ λ¯ω ηωδ(a) = (ι ⊗ (λω ⊠ λ¯ω))δ(a) for a ∈ A. Since the image of C ∗ that Aω ⊂ M (A ⊗ C ∗ r (G) under λω ⊠λ¯ω is contained in M (C ∗ r (G, ω)). r (G, ω)⊗C ∗ r (G, ¯ω)), we see in particular Example 2.2. Assume the group G is discrete. Denote by Ag ⊂ A the spectral subspace corresponding to g ∈ G, so Ag consists of all elements a ∈ A such that δ(a) = a ⊗ λg. The spaces Ag, g ∈ G, span a dense ∗-subalgebra A ⊂ A. By (2.1), if a ∈ Ag then ηωδ(a) = a ⊗ λω g . Thus the linear span of elements (ι ⊗ ι ⊗ ϕ)ηωδ(a), with a ∈ A and ϕ ∈ K ∗, coincides with the linear span Aω of elements a ⊗ λω g , with a ∈ Ag and g ∈ G. The space Aω is already a ∗-algebra and Aω is the closure of Aω in A ⊗ C ∗ r (G, ω). In particular, we see that for discrete groups our definition of ω-deformation is equivalent to that of Yamashita, see [14, Proposition 2]. ♦ g ⊗ λ¯ω The following theorem is the first principal result of this section. Theorem 2.3. The C∗-algebra Aω ⊂ M (A ⊗ K) coincides with the norm closure of the linear span of elements of the form (ι ⊗ ι ⊗ ϕ)ηωδ(a), where a ∈ A and ϕ ∈ K ∗. While proving this theorem we will simultaneously obtain a description of Aω ⊗ K. We need to introduce more notation in order to formulate the result. In addition to λω we have another equivalent representation ρω of C ∗ λω g ∈ M (C ∗ Given an action α of G on a C∗-algebra B, the reduced twisted crossed product is defined by r (G, ω) on L2(G) that maps r (G, ω)) into ρω g . B ⋊α,ω G = α(B)(1 ⊗ ρω(C ∗ r (G, ω))) ⊂ M (B ⊗ K). The reduced twisted crossed product has a dual coaction, which we again denote by α, defined by α(x) = W ∗ 23(x ⊗ 1)W23, so α(α(b)) = α(b) ⊗ 1, α(1 ⊗ ρω g ) = 1 ⊗ ρω g ⊗ λg. The last ingredient that we need is the well-known fact that the cocycles ω and ¯ω are cohomologous. Explicitly, ω(g, h) = ¯ω(g, h)v(g)v(h)v(gh)−1 , where v(g) = ω(g−1, g)ω(e, e). This follows from the cocycle identities ω(h−1, g−1)ω(h−1g−1, gh) = ω(h−1, h)ω(g−1, gh), ω(g−1, gh)ω(g, h) = ω(g−1, g)ω(e, h); recall also that ω(e, h) = ω(e, e) for all h, which follows from the cocycle identity applied to (e, e, h). We can now formulate our second principal result. Theorem 2.4. Put u(g) = ω(g−1, g)ω(e, e). Then the map Ad((1 ⊗ W ω)(1 ⊗ 1 ⊗ u)) : A ⋊δ G ⋊δ,ω G → M (A ⊗ K ⊗ K) defines an isomorphism A ⋊δ G ⋊δ,ω G ∼= Aω ⊗ K. For discrete groups the fact that the C∗-algebras Aω and A ⋊δ G ⋊δ,ω G are strongly Morita equivalent was observed by Yamashita [14, Corollary 15]. 6 J. BHOWMICK, S. NESHVEYEV, AND A. SANGHA Proof of Theorems 2.3 and 2.4. Denote by θ the map in the formulation of Theorem 2.4. In order to compute its image, observe first that since ¯ω(h, g) = ω(h, g)u(h)u(g)u(hg)−1 , we have uρω g u∗ = u(g)ρ ¯ω g . Next, it is straightforward to check that W ω commutes with 1 ⊗ ρ¯ω δ(a) ⊗ 1 7→ ηωδ(a), 1 ⊗ ∆(f ) 7→ 1 ⊗ 1 ⊗ f, 1 ⊗ 1 ⊗ ρω g . We thus see that θ acts as g 7→ 1 ⊗ 1 ⊗ uρω g u∗. In particular, we see that the image of the C∗-subalgebra (1 ⊗ ∆(C0(G)))(1 ⊗ 1 ⊗ ρω(C ∗ r (G, ω))) ∼= C0(G) ⋊Ad ρ,ω G of M (A ⋊δ G ⋊δ,ω G) is Therefore 1 ⊗ 1 ⊗ K is a nondegenerate C∗-subalgebra of M (θ(A ⋊δ G ⋊δ,ω G)) ⊂ M (A ⊗ K ⊗ K). It follows that there exists a uniquely defined C∗-subalgebra A1 ⊂ M (A ⊗ K) such that 1 ⊗ 1 ⊗ uC0(G)C ∗ r (G, ω)u∗ = 1 ⊗ 1 ⊗ K. θ(A ⋊δ G ⋊δ,ω G) = A1 ⊗ K. By definition of crossed products and the above computation of θ we then have A1 ⊗ K = ηωδ(A)(1 ⊗ 1 ⊗ K). Applying the slice maps ι ⊗ ι ⊗ ϕ we conclude that the closed linear span of elements of the form (ι ⊗ ι ⊗ ϕ)ηωδ(a) coincides with the C∗-algebra A1. This finishes the proof of both theorems. (cid:3) Theorem 2.4 essentially reduces the study of ω-deformations to that of (twisted) crossed prod- ucts. As a simple illustration let us prove the following result that refines and generalizes [14, Proposition 14]. Proposition 2.5. Assume we are given two exterior equivalent coactions δ and δX of G on a C∗- algebra A. Then Aδ,ω ⊗ K ∼= AδX ,ω ⊗ K. Proof. Since δ and δX are exterior equivalent, we have (A ⋊δ G, δ) ∼= (A ⋊δX A ⋊δ G ⋊δ,ω G ∼= A ⋊δX G, δX ), and hence (cid:3) G ⋊δX ,ω G. Note that for continuous cocycles this result is also a consequence of the following useful fact combined with the Takesaki-Takai duality. Proposition 2.6. If the cocycle ω is continuous, then any two exterior equivalent coactions have exterior equivalent twisted dual actions. More precisely, assume X ∈ M (A ⊗ C ∗ r (G)) is a 1-cocycle for a coaction δ of G on A. Then the element U = X12 ω23X ∗ 23 ∈ M (A⊗K ⊗C0(G)) is a 1-cocycle for the action δω G intertwines δω with (δω Proof. Denote by Ψ the isomorphism Ad X : A ⋊δ G → A ⋊δX G, and the isomorphism Ad X : A ⋊δ G → A ⋊δX X of G on A ⋊δX G and put X )U . 12 ω∗ Y = 1 ⊗ ω ∈ M (1 ⊗ C0(G) ⊗ C0(G)) ⊂ M ((A ⋊δ G) ⊗ C0(G)) ∩ M ((A ⋊δX G) ⊗ C0(G)). Then U = (Ψ ⊗ ι)(Y )Y ∗ ∈ M ((A ⋊δX observe first that G) ⊗ C0(G)). In order to show that U is a 1-cocycle for δω X , (2.2) which is simply the cocycle identity for ω. We also have the same identity for δ. Furthermore, since Ψ intertwines δ with δX , we also get (Y ⊗ 1)(δX ⊗ ι)(Y ) = (ι ⊗ ι ⊗ ∆)(Y )ω34, ((Ψ ⊗ ι)(Y ) ⊗ 1)(δX ⊗ ι)(Ψ ⊗ ι)(Y ) = (ι ⊗ ι ⊗ ∆)(Ψ ⊗ ι)(Y )ω34. Multiplying this identity by the adjoint of (2.2) we obtain ((Ψ ⊗ ι)(Y ) ⊗ 1)(δX ⊗ ι)(U )(Y ∗ ⊗ 1) = (ι ⊗ ι ⊗ ∆)(U ). DEFORMATION OF OPERATOR ALGEBRAS BY BOREL COCYCLES 7 Since δω X = Y δX (·)Y ∗, this is exactly the cocycle identity (U ⊗ 1)(δX ⊗ ι)(U ) = (ι ⊗ ι ⊗ ∆)(U ). Since δω = Y δ(·)Y ∗, δω X = Y δX (·)Y ∗ and Ψ intertwines δ with δX , we immediately see that Ψ intertwines δω with (Ψ ⊗ ι)(Y )δX (·)(Ψ ⊗ ι)(Y )∗ = U δω X (·)U ∗. We finish the section with the following simple observation. (cid:3) Proposition 2.7. Assume ω1, ω2 ∈ Z 2(G; T) are cohomologous cocycles. Then Aω1 ∼= Aω2 . Proof. By assumption there exists a Borel function v : G → T such that ω1(g, h) = ω2(g, h)v(g)v(h)v(gh)−1 , that is, ω1 = ω2(v ⊗ v)∆(v)∗. Note that then λω1 and that W commutes with v ⊗ 1, for any operator x on L2(G) we get g = v(g−1)vλω2 g v∗. Using that W ∆(v)W ∗ = 1 ⊗ v W ω1(x ⊗ 1)ω∗ 1W ∗ = (v ⊗ v∗)W ω2(x ⊗ 1)ω∗ 2W ∗(v∗ ⊗ v). This shows that ηω1 = Ad(1 ⊗ v ⊗ v∗)ηω2, which in turn gives Aω1 = Ad(1 ⊗ v)(Aω2). (cid:3) 3. Canonical and dual coactions By the Landstad-type result of Quigg and Vaes the twisted dual action δω, when it is defined, is dual to some coaction. The action δω is apparently not always well-defined on A ⋊δ G. Nevertheless the new coaction on Aω always makes sense. Theorem 3.1. For any cocycle ω ∈ Z 2(G; T) and a coaction δ of G on a C∗-algebra A we have: (i) the formula δω(x) = W23(x ⊗ 1)W ∗ (ii) if the twisted dual action δω is well-defined on A ⋊δ G, then A ⋊δ G = Aω(1 ⊗ C0(G)) and the map ηω : A ⋊δ G → M (A ⊗ K ⊗ K) gives an isomorphism A ⋊δ G ∼= Aω ⋊δω G that intertwines the twisted dual action δω on A ⋊δ G with the dual action to δω on Aω ⋊δω G. Proof. (i) We repeat the computations of Vaes in the proof [13, Theorem 6.7]. Since 23 defines a coaction of G on Aω; W13W12 = (ι ⊗ ∆)(W ) = W23W12W ∗ 23, for x = (ι ⊗ ι ⊗ ϕ)ηω(y), y ∈ A ⋊δ G, we have δω(x) = (ι ⊗ ι ⊗ ϕ ⊗ ι)(W24(ηω(y) ⊗ 1)W ∗ 24) = (ι ⊗ ι ⊗ ϕ ⊗ ι)(W24W23 ω23(δ(y) ⊗ 1)ω∗ = (ι ⊗ ι ⊗ ϕ ⊗ ι)(W34W23W ∗ = (ι ⊗ ι ⊗ ϕ ⊗ ι)(W34(ηω(y) ⊗ 1)W ∗ 23W ∗ 34 ω23(δ(y) ⊗ 1)ω∗ 34). 23W ∗ 24) 23W34W ∗ 23W ∗ 34) From this one can easily see that the closure of δω(Aω)(1 ⊗ 1 ⊗ C ∗ r (G), because (K ⊗ 1)W (1 ⊗ C ∗ r (G). Since 1 ⊗ W is a 1- cocycle for the trivial coaction on A⊗K (so (ι⊗ ∆)(W ) = W12W13), the identity (ι⊗ ∆)δω = (δω ⊗ι)δω follows. r (G)) coincides with Aω ⊗ C ∗ r (G) and W ∗(K ⊗ C ∗ r (G)) = K ⊗ C ∗ r (G)) = K ⊗ C ∗ (ii) This is [13, Theorem 6.7] applied to the action δω. (cid:3) The twisted dual action is well-defined for continuous cocycles, but as the following result shows it can also be well-defined even if the cocycle is only Borel. Proposition 3.2. If δ is a dual coaction, then the twisted dual action δω of G on A ⋊δ G is well- defined for any ω ∈ Z 2(G; T). 8 J. BHOWMICK, S. NESHVEYEV, AND A. SANGHA Proof. By assumption we have A = B ⋊αG and δ = α for some B and α. Then A⋊δ G = B ⋊αG⋊ α G is the closure of (α(B) ⊗ 1)(1 ⊗ (ρ ⊗ λ) ∆(C ∗ r (G)))(1 ⊗ 1 ⊗ C0(G)) ⊂ M (B ⊗ K ⊗ K). We have to check that the inner automorphisms Ad(1 ⊗ 1 ⊗ ρω g ) of B ⊗ K ⊗ K define a (continuous) action of G on this closure. Since these automorphisms act trivially on α(B) ⊗ 1, we just have to check that the automorphisms Ad(1 ⊗ ρω g ) of K ⊗ K define an action on the C∗-algebra r (G))(1 ⊗ C0(G)) ∼= C ∗ r (G) ⋊ G. (ρ ⊗ λ) ∆(C ∗ The operator V commutes with 1 ⊗ ω(·, g), and Ad V ∗ maps the above algebra onto 1 ⊗ K. Hence Ad(1⊗ ω(·, g)), and therefore also Ad(1⊗ρω g ), is a well-defined automorphism of that algebra. Finally, the continuity of the action holds, since any Borel homomorphism of G into a Polish group, such as the group Aut(K), is automatically continuous. (cid:3) For dual coactions it is, however, straightforward to describe the deformed algebra, see [14, Ex- In order to formulate the result, define a unitary W ω on ample 8] for the discrete group case. L2(G × G) by (W ωξ)(g, h) = ω(g−1, h)ξ(g, g−1h). In other words, if we let W ∗(G, ω) = C ∗ and W ω(g) = λω g . r (G, ω)′′, then W ω ∈ L∞(G) ¯⊗W ∗(G, ω) = L∞(G; W ∗(G, ω)) Proposition 3.3. Assume α is an action of G on a C∗-algebra B. Consider the dual coaction δ on A = B ⋊α G. Then for any ω ∈ Z 2(G; T) the map B ⋊α,ω G 7→ M (B ⊗ K ⊗ K), x 7→ W ω∗ 23 (x ⊗ 1)W ω 23, defines an isomorphism (B ⋊α,ω G, α) ∼= (Aω, δω). Proof. First of all observe that by (2.1) we have This implies that Aω is the closed linear span of elements of the form ηω(δ(1 ⊗ ρg)) = 1 ⊗ ρg ⊗ λω g ⊗ λ¯ω g . (δ(b) ⊗ 1)ZG f (g)(1 ⊗ ρg ⊗ λω g )dg, where b ∈ B and f ∈ L1(G). Using the easily verifiable identity g ⊗ 1)W ω = ρg ⊗ λω g , W ω∗(ρω we get the required isomorphism α(B)(1 ⊗ ρω(C ∗ r (G, ω))) → Aω, x 7→ W ω∗ 23 (x ⊗ 1)W ω 23. In order to see that this isomorphism respects the coactions, we just have to check that δω(1 ⊗ ρg ⊗ λω g ) = 1 ⊗ ρg ⊗ λω g ⊗ λg, that is, W (λω since λω g ⊗ 1)W ∗ = λω g is λg multiplied by a function that automatically commutes with the first leg of W . g ⊗ λg. But this follows immediately from W (λg ⊗ 1)W ∗ = λg ⊗ λg, (cid:3) Consider now an arbitrary coaction δ of G on a C∗-algebra A and choose two cocycles ω, ν ∈ Z 2(G; T). Using the coaction δω on Aω we can define the ν-deformation (Aω)ν of Aω. Proposition 3.4. The map Aων → M (A ⊗ K ⊗ K), x 7→ W23 ν∗ 23(x ⊗ 1)ν23W ∗ 23, defines an isomorphism Aων ∼= (Aω)ν . In particular, the map ηωδ : A → M (A ⊗ K ⊗ K) defines an isomorphism A ∼= (Aω)¯ω. DEFORMATION OF OPERATOR ALGEBRAS BY BOREL COCYCLES 9 Proof. For a ∈ A and ϕ ∈ K ∗ consider the element x = (ι ⊗ ι ⊗ ϕ)ηωδ(a) = (ι ⊗ ι ⊗ ϕ)(ι ⊗ (λω ⊠ λ¯ω))δ(a) ∈ Aω. Recall that λω ⊠ λ¯ω denotes the representation of C ∗ r (G) defined by λg 7→ λω g ⊗ λ¯ω g . Then δω(x) = W23(x ⊗ 1)W ∗ g ⊗ 1)W ∗ = λω Since W (λω above identity can be written as 23 = (ι ⊗ ι ⊗ ϕ ⊗ ι)(W24((ι ⊗ (λω ⊠ λ¯ω))δ(a) ⊗ 1)W ∗ 24). g ⊗ λg, as was already used in the proof of the previous proposition, the δω(x) = (ι ⊗ ι ⊗ ϕ ⊗ ι)(ι ⊗ ((λω ⊠ λ¯ω) ⊠ λ))δ(a). It follows that ην δω(x) = (ι ⊗ ι ⊗ ϕ ⊗ ι ⊗ ι)(ι ⊗ ((λω ⊠ λ¯ω) ⊠ (λν ⊠ λ¯ν)))δ(a). Therefore (Aω)ν is the closed linear span of elements of the form (ι ⊗ ι ⊗ ϕ ⊗ ι ⊗ ψ)(ι ⊗ ((λω ⊠ λ¯ω) ⊠ (λν ⊠ λ¯ν )))δ(a), where a ∈ A and ϕ, ψ ∈ K ∗. Observe next that W ν∗(λων g ⊗ 1)νW ∗ = λω g ⊗ λν g , which is simply identity (2.1) for the cocycle ¯ν multiplied on the left by ω(g−1, ·)ν(g−1, ·) ⊗ 1. It follows that the unitary Σ23(νW ∗ ⊗ ν∗W ∗)Σ23 on L2(G)⊗4, where Σ is the flip, intertwines the representation (λω ⊠ λ¯ω) ⊠ (λν ⊠ λ¯ν ) of C ∗ sentation (λων ⊠ λ¯ω¯ν) ⊗ 1 ⊗ 1. Furthermore, for any y ∈ C ∗ r (G) we have r (G) with the repre- (Ad νW ∗)(ι ⊗ ϕ ⊗ ι ⊗ ψ)((λω ⊠ λ¯ω) ⊠ (λν ⊠ λ¯ν ))(y) = 24((λων ⊠ λ¯ω ¯ν)(y) ⊗ 1 ⊗ 1), where = (ϕ ⊗ ψ)(Ad W ν) ∈ (K ⊗ K)∗. Therefore for any a ∈ A we get (Ad ν23W ∗ 23)(ι ⊗ ι ⊗ ϕ ⊗ ι ⊗ ψ)(ι ⊗ ((λω ⊠ λ¯ω) ⊠ (λν ⊠ λ¯ν)))δ(a) = 35(ηων δ(a) ⊗ 1 ⊗ 1). This shows that Ad(ν23W ∗ 23) maps the algebra (Aω)ν onto Aων ⊗ 1, which proves the first part of the proposition. Then the second part also follows, since the deformation of A by the trivial cocycle is equal to δ(A). (cid:3) 4. K-theory We say that two cocycles ω0, ω1 ∈ Z 2(G; T) are homotopic if there exists a C([0, 1]; T)-valued Borel 2-cocycle Ω on G such that ωi = Ω(·, ·)(i) for i = 0, 1. Our goal is to show that under certain assumptions on G the deformed algebras Aω0 and Aω1 have isomorphic K-theory. For this we will use the following slight generalization of invariance under homotopy of cocycles of K-theory of reduced twisted group C∗-algebras, proved in [1]. Theorem 4.1. Assume G satisfies the Baum-Connes conjecture with coefficients. Then for any action α of G on a C∗-algebra B and any two homotopic cocycles ω0, ω1 ∈ Z 2(G; T), for the corre- sponding reduced twisted crossed products we have K∗(B ⋊α,ω0 G) ∼= K∗(B ⋊α,ω1 G). The proof follows the same lines as that of [1, Theorem 1.9]. The starting point is the isomorphism K ⊗ (B ⋊α,ω G) ∼= (K ⊗ B) ⋊Ad ρ ¯ω⊗α G, x 7→ ω∗ 13V13xV ∗ 13ω13, g ⊗ 1 ⊗ ρω which maps ρ¯ω g into 1 ⊗ 1 ⊗ ρg. This is a particular case of the Packer-Raeburn stabi- lization trick, see [8, Section 3]. Therefore instead of twisted crossed products we can consider (K ⊗ B) ⋊Ad ρ ¯ω⊗α G. Now, given a homotopy Ω of cocycles, consider the action Ad ρ ¯Ω of G on C[0, 1] ⊗ K defined, upon identifying C[0, 1] ⊗ K with C([0, 1]; K), by (Ad ρ ¯Ω g )(f )(t) = (Ad ρ¯ωt g )(f (t)), where ωt = Ω(·, ·)(t). 10 J. BHOWMICK, S. NESHVEYEV, AND A. SANGHA Lemma 4.2 (cf. Proposition 1.5 in [1]). For any compact subgroup H ⊂ G and any t ∈ [0, 1], the restrictions of the actions Ad ρ ¯Ω and id ⊗ Ad ρ¯ωt to H are exterior equivalent. Note that this is easy to see for homotopies of the form ωt = ω0eitc usually considered in applica- tions, where c is an R-valued Borel 2-cocycle. Indeed, by [6, Theorem 2.3] the second cohomology of a compact group with coefficients in R is trivial, so there exists a Borel function b : H → R such that c(h′, h) = b(h′) + b(h) − b(h′h). Extend b to a function on G as follows. Choose a Borel section s : G/H → G of the quotient map G → G/H, g 7→ g, such that s( e) = e. Then put b(g) = b(s( g)−1g) − c(s( g), s( g)−1g) + b(e). A simple computation shows that c(g, h) = b(g) + b(h) − b(gh) for all g ∈ G and h ∈ H. Then the unitaries uh ∈ M (C[0, 1] ⊗ K) defined by uh(t) = eit(b−b(·h)) form a 1-cocycle for the action (Ad ρ ¯Ω)H such that Ad(uhρ ¯Ω h ) = id ⊗ Ad ρ¯ω0 h . Proof of Theorem 4.1. For every t ∈ [0, 1] consider the evaluation map evt : C[0, 1]⊗K ⊗B → K ⊗B. It is G-equivariant with respect to the action Ad ρ ¯Ω ⊗ α of G on C[0, 1] ⊗ K ⊗ B and the action Ad ρ¯ωt ⊗ α of G on K ⊗ B. We claim that it induces an isomorphism (evt ⋊G)∗ : K∗((C[0, 1] ⊗ K ⊗ B) ⋊ Ad ρ ¯Ω⊗α G) → K∗((K ⊗ B) ⋊Ad ρ ¯ωt ⊗α G). By [1, Proposition 1.6] in order to show this it suffices to check that for every compact subgroup H of G the map evt induces an isomorphism (evt ⋊H)∗ : K∗((C[0, 1] ⊗ K ⊗ B) ⋊ Ad ρ ¯Ω⊗α H) → K∗((K ⊗ B) ⋊Ad ρ ¯ωt ⊗α H). By Lemma 4.2 the action Ad ρ ¯Ω ⊗ α of H on C[0, 1] ⊗ K ⊗ B is exterior equivalent to the action id ⊗ Ad ρ¯ωt ⊗ α, so that (C[0, 1] ⊗ K ⊗ B) ⋊ Ad ρ ¯Ω⊗α H ∼= C[0, 1] ⊗ ((K ⊗ B) ⋊Ad ρ ¯ωt ⊗α H). If the cocycle U = {uh}h∈H defining the exterior equivalence is chosen such that uh(t) = 1 for all h ∈ H, then the corresponding homomorphism C[0, 1] ⊗ ((K ⊗ B) ⋊Ad ρ ¯ωt ⊗α H) → (K ⊗ B) ⋊Ad ρ ¯ωt ⊗α H is simply the evaluation at t. Obviously, it defines an isomorphism in K-theory. (cid:3) Combining Theorems 2.4 and 4.1 we get the following result that generalizes several earlier results in the literature [10, 14]. Corollary 4.3. Assume G satisfies the Baum-Connes conjecture with coefficients. Then for any coaction δ of G on a C∗-algebra A and any two homotopic cocycles ω0, ω1 ∈ Z 2(G; T), we have an isomorphism K∗(Aω0) ∼= K∗(Aω1 ). We finish by noting that for some groups it is possible to prove a stronger result. For example, generalizing Rieffel's result for Rd [10] we have the following. Proposition 4.4. If G is a simply connected solvable Lie group, then for any coaction δ of G on a C∗-algebra A and any cocycle ω ∈ Z 2(G; T) we have K∗(Aω) ∼= K∗(A). Proof. By the stabilization trick and Connes' Thom isomorphism we have Ki(A ⋊δ G ⋊δ,ω G) ∼= Ki+dim G(A ⋊δ G) ∼= Ki(A ⋊δ G ⋊δ G) ∼= Ki(A). (cid:3) DEFORMATION OF OPERATOR ALGEBRAS BY BOREL COCYCLES 11 Appendix A. Rieffel's deformation It was stated by Kasprzak [4] that for G = Rd his approach to deformation, which our construction extends, is equivalent to that of Rieffel [9], but no proof of this was given. A sketch of a possible proof was then proposed by Hannabuss and Mathai [2], but in our opinion it is not easy to obtain a complete proof following the suggested strategy. The goal of this appendix is to give a different rigorous proof using completely positive maps constructed by Kaschek, Neumaier and Waldmann [3]. We will use the conventions in [3] that are slightly different from those of Rieffel. Assume V is a 2n-dimensional Euclidean space with scalar product h·, ·i, and J is a complex structure on V , so J is an orthogonal transformation and J 2 = −1. Fix a deformation parameter h > 0. Assume we are given an action α of V on a C∗-algebra A. Denote by A∞ the subalgebra of smooth vectors for this action. It is a Fr´echet algebra equipped with differential norms k · kk, k ≥ 1. Rieffel defines a new product ∗h on A∞ by a ∗h b = 1 (πh)2n ZV ×V αx(a)αy(b)e− 2i h hx,Jyidx dy, where the integral is understood as an oscillatory integral. Denote by Ax the spectral subspace of A∞ corresponding to x ∈ V , so Ax consists of elements a ∈ A such that αy(a) = eihx,yia for all y ∈ V . Then for a ∈ Ax and b ∈ Ay we have a ∗h b = e ih 2 hx,Jyiab. Note that the spectral subspaces are often trivial, so this formula by no means determines ∗h. Nev- 2 hx,Jyi. The Rieffel deformation ertheless it indicates that the cocycle of deformation is ω(x, y) = e of A is a certain C∗-algebraic completion of A∞ equipped with the product ∗h and with the involution inherited from A, see [9] for details. We denote it by Aω. ih The action α can be viewed as a coaction δ of V on A. Namely, define the Fourier transform F : L2(V ) → L2(V ) by (Ff )(x) = (2π)n ZV Then Ad F defines an isomorphism of C0(V ) onto C ∗ r (V ), and by letting δ = Ad(1 ⊗ F)α we get a coaction of V on A. Note that a ∈ A lies in the spectral subspace Ax if and only if δ(a) = a ⊗ λx, in agreement with our previous notation. We can then consider the ω-deformation Aω of A. Our aim is to construct an isomorphism between Aω and Aω. f (y)e−ihx,yidy. 1 Following [3] define a map Φ : A → A by Φ(a) = 1 (πh)n ZV e− 1 h kxk2 αx(a)dx. We have Φ(a) = e− h 4 kxk2 a for a ∈ Ax. (A.1) The image of Φ is contained in A∞. So we can consider Φ as a map T : A → Aω. Identifying A with Rieffel's deformation of Aω corresponding to the complex structure −J, we also get a similarly defined map S : Aω → A, so the restriction of S to A∞ coincides with the restriction of Φ to A∞. Since Φ considered as a map (A, k · k) → (A∞, k · kk) is bounded for any k, the map T : A → Aω is bounded by standard estimates for the operator norm on Aω, see [9, Proposition 4.10]. By symmetry the map S is also bounded. The main result in [3] states that the maps T and S are completely positive. We will reprove this a bit later. We want to define analogues of the maps T and S for Aω. For this, define a unit vector ξ0 ∈ L2(V ) by ξ0(x) = (cid:18) h 2π(cid:19)n/2 e− h 4 kxk2 . 12 J. BHOWMICK, S. NESHVEYEV, AND A. SANGHA Consider the normal state ϕ0 = (· ξ0, ξ0) on B(L2(V )). We have 4 kxk2 ϕ0(λω x ) = ϕ0(λ¯ω x ) = e− h . Note that this means that on the C∗-algebra generated by the operators λω x , which is the algebra of canonical commutation relations for the space V equipped with the Hermitian scalar product hhx, yi + ihhx, Jyi, the state ϕ0 is simply the vacuum state. Define T : A → Aω and S : Aω → A by T (a) = (ι ⊗ ι ⊗ ϕ0)ηωδ(a), S(b) = (ι ⊗ ϕ0)(b). Using that δ(A)(1 ⊗ C ∗ r (V ) it is not difficult to see that the image of S is indeed contained in A rather than in M (A). This will also become clear from the proof of Lemma A.2 below. r (V )) ⊂ A ⊗ C ∗ The maps T and S are completely positive. Using that ηωδ(a) = a ⊗ λω x ⊗ λ¯ω x for a ∈ Ax, we get T (a) = e− h 4 kxk2 a ⊗ λω x and S(a ⊗ λω x ) = e− h 4 kxk2 a for a ∈ Ax. (A.2) Lemma A.1. For any n ≥ 1 and a1, . . . , an ∈ A we have S( T (a1) . . . T (an)) = S(T (a1) . . . T (an)). Proof. If for every j the element aj lies in a spectral subspace Axj , then the identity in the formulation follows immediately from (A.1) and (A.2). We will show that this is enough to conclude that it holds for arbitrary elements. We claim that there exists a von Neumann algebra M containing A such that the action α of V on A extends to a continuous (in the von Neumann algebraic sense) action of V on M and such that M is generated as a von Neumann algebra by the spectral subspaces of this action. Indeed, first represent the crossed product A ⋊α V faithfully on some Hilbert space H and consider the von Neumann algebra N ⊂ B(H) generated by A. The action α of V on A extends to an action β of V V in the von Neumann algebraic sense. on N . Consider the double crossed product M = N ⋊β V ⋊ β β) ∼= (N ¯⊗B(L2(V )), β ⊗ Ad ρ). This gives us an equivariant By the Takesaki duality we have (M, β. It is also clear that M is generated by the embedding of A ⊂ N into M equipped with the action spectral subspaces of the action, so our claim is proved. We continue to denote by α the action of V on M . Denote by M ⊂ M the set of elements a ∈ M such that the map x 7→ αx(a) is norm-continuous. This is an ultrastrongly operator dense C∗-subalgebra of M . We continue to denote by T, S, T , S the maps defined for the C∗-algebra M in place of A. The maps T and S have obvious extensions to normal maps between the von Neumann algebras generated by M and Mω. On the other hand, the map Φ, Φ(a) = 1 (πh)n ZV e− 1 h kxk2 αx(a)dx, is still well-defined on M , but now the integral should be taken with respect to the ultrastrong operator topology. The image of M under Φ is contained in M∞. It therefore makes sense to ask whether the identity Φ(Φ(a1) ∗h · · · ∗h Φ(an)) = S(T (a1) . . . T (an)) holds for all a1, . . . , an ∈ M , which would imply the assertion of the lemma. Since this identity holds for a1, . . . , an lying in spectral subspaces of M , it suffices to show that both sides of the identity are normal maps in every variable aj running through the unit ball M1 of M . This is clearly the case for the right hand side. In order to prove the same for the left hand side it suffices to show that for any b, c ∈ M∞ the map is continuous in the ultrastrong operator topology. M1 → M, a 7→ Φ(b ∗h Φ(a) ∗h c), DEFORMATION OF OPERATOR ALGEBRAS BY BOREL COCYCLES 13 Using basic estimates for oscillatory integrals, see [9, Chapter 1], and the fact that the map Φ is bounded as a map (M, k · k) → (M∞, k · kk) for every k, it is easy to check that Φ(b ∗h Φ(a) ∗h c) can be approximated in norm uniformly in a ∈ M1 by integrals of the form ZV 3 ψ(x, y, z)αx(b)αy(a)αz(c)dx dy dz, where ψ is a smooth compactly supported function and the integral is taken with respect to the ultrastrong operator topology. Since such integrals are clearly continuous in a ∈ M1 in this topology, this finishes the proof of the lemma. (cid:3) We will need the above lemma only for n = 1, 2, 3. Lemma A.2. The maps T and S are completely positive, and all four maps T, S, T , S are injective and their images are dense. Proof. We begin by proving that T and S are injective. It suffices to consider T . Assume T (a) = 0. Then 1 (πh)n ZV e− 1 h hx−y,x−yiαx(a)dx = αy(Φ(a)) = 0 (A.3) for all y ∈ V , hence, by analyticity, for all y in the complexification VC of V . This implies that the Fourier transform of the A-valued function x 7→ e− 1 αx(a) is zero, whence a = 0. h kxk2 Next we will show that the images of T and S are dense. It suffices to consider S, and then it is enough to show that the image of Φ is dense. It is well-known, and is easy to check using e.g. Wiener's Tauberian theorem, that the translations of the function e− 1 span a dense subspace of L1(V ). Using the first equality in (A.3) and that αy(Φ(a)) = Φ(αy(a)), we conclude that the closure of the image of Φ contains all elements of the form RV f (x)αx(a)dx with f ∈ L1(V ). Hence Let us show now that T and S are completely positive. Again, it is enough to consider S. Since this closure coincides with A. h kxk2 by Lemma A.1 we have S( T (a)∗ T (a)) = S(T (a)∗T (a)) ≥ 0, and the image of T is dense, we see that S is positive. Passing to deformations of matrix algebras over A we conclude that S is completely positive. This finishes the proof of the lemma for T and S. Turning to T and S, by Lemma A.1 we have ST = S T = Φ2. Since the map Φ is injective and its image is dense, it follows that the map T is injective and the image of S is dense. Consider the maps T ′ : Aω → (Aω)¯ω and S′ : (Aω)¯ω → Aω defined by (Aω, δω) in the same way as T and S were defined by (A, δ). Then T ′ is injective and the image of S′ is dense. By Proposition 3.4 the map ηωδ defines an isomorphism A ∼= (Aω)¯ω. By definition of T and S′ we immediately get T = S′ηωδ. Hence the image of T is dense. We also have ηωδS = T ′. Indeed, a simple computation similar to the ones used in the proof of Proposition 3.4 shows that for b = (ι ⊗ ι ⊗ ϕ)ηωδ(a) ∈ Aω we have ηωδS(b) = (ι ⊗ ι ⊗ ι ⊗ ϕ0 ⊗ ϕ)(ι ⊗ ((λω ⊠ λ¯ω) ⊠ (λω ⊠ λ¯ω)))δ(a) and T ′(b) = (ι ⊗ ι ⊗ ϕ ⊗ ι ⊗ ϕ0)(ι ⊗ ((λω ⊠ λ¯ω) ⊠ (λ¯ω ⊠ λω)))δ(a). Alternatively, the identity ηωδS = T ′ is immediate on elements of the form a ⊗ λx, where a ∈ Ax, hence it holds on arbitrary elements by an argument similar to the one used in the proof of Lemma A.1. It follows that the map S is injective. (cid:3) Note that instead of injectivity we will only need to know that S and S are faithful. While it is obvious that S is faithful, this is not the case for S, since the state ϕ0 is very far from being faithful on the von Neumann algebra generated by C ∗ r (V, ω). This von Neumann algebra is a factor of type I∞ and ϕ0 is a normal pure state on it, as can be shown by recalling that ϕ0 defines the vacuum state on the algebra of canonical commutation relations generated by the operators λω x . 14 J. BHOWMICK, S. NESHVEYEV, AND A. SANGHA Theorem A.3. There is a unique isomorphism Aω ∼= Aω that maps T (a) into T (a) for every a ∈ A. Proof. Assume first that there exists a faithful state ψ on A. Consider the positive linear functionals ψω = ψS and ψω = ψ S on Aω and Aω. Since the positive maps S and S are faithful, these functionals are faithful. Consider the faithful GNS-representation of Aω on H with cyclic vector ξ defining ψω, and the faithful GNS-representation of Aω on H with cyclic vector ξ defining ψω. By Lemma A.1 for n = 2 we have (T (a)ξ, T (b)ξ) = ( T (a)ξ, T (b)ξ). Since the images of T and T are dense, it follows that there exists a unitary operator U : H → H such that U T (a)ξ = T (a)ξ. By Lemma A.1 for n = 3 we have (T (a)T (b)ξ, T (c)ξ) = ( T (a) T (b)ξ, T (c)ξ), that is, (U T (a)T (b)ξ, U T (c)ξ) = ( T (a)U T (b)ξ, U T (c)ξ). Therefore U T (a) = T (a)U , so Ad U defines the required isomorphism. In the general case the proof is basically the same, but instead of one state ψ we have to choose a faithful family of states on A and consider direct sums of the GNS-representations defined by the corresponding positive linear functionals on Aω and Aω. (cid:3) References [1] S. Echterhoff, W. Luck, N.C. Phillips and S. Walters, The structure of crossed products of irrational rotation algebras by finite subgroups of SL2(Z), J. Reine Angew. Math. 639 (2010), 173 -- 221. [2] K.C. Hannabuss and V. Mathai, Noncommutative principal torus bundles via parametrised strict deformation quantization, in: "Superstrings, geometry, topology, and C ∗-algebras", 133 -- 147, Proc. Sympos. Pure Math., 81, Amer. Math. Soc., Providence, RI, 2010. [3] D. Kaschek, N. Neumaier and S. Waldmann, Complete positivity of Rieffel's deformation quantization by actions of Rd, J. Noncommut. Geom. 3 (2009), no. 3, 361 -- 375. [4] P. Kasprzak, Rieffel deformation via crossed products, J. Funct. Anal. 257 (2009), no. 5, 1288 -- 1332. [5] M.B. Landstad, J. Phillips, I. Raeburn and C.E. Sutherland, Representations of crossed products by coactions and principal bundles, Trans. Amer. Math. Soc. 299 (1987), no. 2, 747 -- 784. [6] C.C. Moore, Extensions and low dimensional cohomology theory of locally compact groups. I, Trans. Amer. Math. Soc. 113 (1964), 40 -- 63. [7] C.C. Moore, Group extensions and cohomology for locally compact groups. III, Trans. Amer. Math. Soc. 221 (1976), no. 1, 1 -- 33. [8] J.A. Packer and I. Raeburn, Twisted crossed products of C ∗-algebras, Math. Proc. Cambridge Philos. Soc. 106 (1989), no. 2, 293 -- 311. [9] M.A. Rieffel, Deformation quantization for actions of Rd, Mem. Amer. Math. Soc. 106 (1993), no. 506. [10] M.A. Rieffel, K-groups of C ∗-algebras deformed by actions of Rd, J. Funct. Anal. 116 (1993), no. 1, 199 -- 214. [11] J.C. Quigg, Landstad duality for C ∗-coactions, Math. Scand. 71 (1992), no. 2, 277 -- 294. [12] S. Vaes, Strictly outer actions of groups and quantum groups, J. Reine Angew. Math. 578 (2005), 147 -- 184. [13] S. Vaes, A new approach to induction and imprimitivity results, J. Funct. Anal. 229 (2005), no. 2, 317 -- 374. [14] M. Yamashita, Deformation of algebras associated to group cocycles, preprint arXiv:1107.2512v1 [math.OA]. E-mail address: [email protected] E-mail address: [email protected] E-mail address: [email protected] Department of Mathematics, University of Oslo, P.O. Box 1053 Blindern, NO-0316 Oslo, Norway
1607.08769
1
1607
2016-07-29T11:29:49
A no-go theorem for the continuum limit of a periodic quantum spin chain
[ "math.OA", "math-ph", "math.GR", "math.GT", "math-ph" ]
We show that the Hilbert space formed from a block spin renormalization construction of a cyclic quantum spin chain (based on the Temperley-Lieb algebra) does not support a chiral conformal field theory whose Hamiltonian generates translation on the circle as a continuous limit of the rotations on the lattice.
math.OA
math
A NO-GO THEOREM FOR THE CONTINUUM LIMIT OF A PERIODIC QUANTUM SPIN CHAIN. VAUGHAN F. R. JONES Abstract. We show that the Hilbert space formed from a block spin renormal- ization construction of a cyclic quantum spin chain (based on the Temperley-Lieb algebra) does not support a chiral conformal field theory whose Hamiltonian gen- erates translation on the circle as a continuous limit of the rotations on the lattice. 1. Introduction This paper is part of an ongoing effort to construct a conformal field theory for every finite index subfactor in such a way that the standard invariant of the subfactor, or at least its quantum double, can be recovered from the CFT. In [12] an infinite dimensional Hilbert space, envisioned as a limit of the Hilbert spaces of finite quantum spin chains on the circle, was constructed using the following data: 1) A (positive definite) "planar algebra" P ([8]) together with an affine unitary representation of it ([11],[13]). (An affine unitary representation is N-graded and the nth graded component is thought of as the Hilbert space of a period quantum spin chain with n spins. For the simplest planar algebra, the nth graded component is just ⊗n(C2) so it is literally the Hilbert space of a quantum spin chain. In [9] it is argued that planar algebras are indeed physically meaningful generalisations of ordinary spin chains.) 2) An element R of P4 with the normalization property Figure 1.0.1. (See the appendix for an explanation of planar algebra. But one does not need to know planar algebras to understand the constructions. Just interpret the R's inside V.J. is supported by the NSF under Grant No. DMS-0301173 and grant DP140100732, Symme- tries of subfactors. 1 *=RR 2 VAUGHAN F. R. JONES the pictures as tensors with indices on the strings and the picture as giving a scheme for contracting indices-[17]. This is already common usage in the physics literature.) The element R ∈ P4 serves as a way of embedding the Hilbert space of a quantum spin chain with n spins into the Hilbert space of a spin chain with 3n spins. As a result of conversation with Tobias Osborne and Guifre Vidal we shall call these limit Hilbert spaces "semicontinuous limits" of Hilbert spaces for the quantum spin chains. We will begin in the next section by giving a simplified and more general version of the construction of the semicontinuous limit of [12]. For the circular version this will give us unitary representations of Thompson's group T which acts by local scale transformations. This representation was hoped to tend to a representation of Dif f +(S1) by taking limits of elements of T on the semicontinuous limit. In particular the rotation group Rot(S1) was hoped to arise as the closure of the dyadic rotations is T . This approach is somewhat naive and very open to criticism on physical grounds, and in this paper we show that this possibility fails as dramatically as possible, at least for one example of a semicontinuous limit H. We show in fact that in this case for any two vectors ξ, η ∈ H, n→∞(cid:104)ρ 1 lim 2n ξ, η(cid:105) = 0 where ρx is unitary on H representing rotation of the circle R/Z by a dyadic rational x. Thus even in the weak topology the rotations by dyadic rationals are discontnuous (though we do not show that limr→0(cid:104)ρrξ, η(cid:105) = 0). Faced with this failure there are two possibilities. The first is to abandon the semicontinuous limit and look for other ways to obtain the Hilbert space of the conformal field theory. One idea which is relatively close to our approach is to replace our embeddings of quantum spin chains one in another by Evenbly and Vidal's MERA (see [6]) which introduces more local interactions between the spins. Vidal's numerical evidence could be interpreted as saying that the Hilbert space obtained by the MERA embeddings should naturally support a CFT. We have not made any progress along these lines. See also [3]. Another possibility is to redefine the goal. After all, the direct limit approach does produce states of a quantum spin chain that transform according to local scale transformations of the lattice. Perhaps this semicontinuous limit is of value in the analysis of critical behaviour of lattice quantum spin chains. The mathematics is completely different from that of CFT but the structure of the nogo theorem certainly ξ, η(cid:105) tends to yields numerical data that could be relevant, e.g. the rate at which (cid:104)ρ 1 zero. And the appearance of the transfer matrix in the proof is oddly dual to the role of the transfer matrix in models which, if [16] is to be believed, should have CFT as a scaling limit. In these solvable models the Hamiltonian-the infinitesimal generator of time evolution-is obtained as the logarithmic derivative of the transfer matrix with respect to the spectral parameter. In our case the infinitesimal behaviour of time (=space) evolution is governed by the transfer matrix. 2n A NO-GO THEOREM FOR THE CONTINUUM LIMIT OF A PERIODIC QUANTUM SPIN CHAIN.3 In a future paper we will investigate scale invariant Hamiltonians and transfer matrices on the semicontinuous limit. In this paper all planar algebras will be unshaded and all representations will have a positive definite invariant inner product unless otherwise specified. (See the appendix for the meaning of this terminology.) 2. A categorical construction of the Thompson groups. 2.1. A group of fractions for certain categories. The following construction of groups is well known and goes back at least as far as a 1931 result on semigroups of Ore. (See also the work [4] in the category context.) The use of direct limits and functors to construct representations of groups of fractions is probably also well known, but less so as the corresponding representations of Thompson's groups seem to have appeared first in [12]. For this reason we give a self-contained exposition of the whole business. The extension from group of fractions to groupoid of fractions is clear. Let K be a small category with the following 3 properties. (i) (Unit) There is an element 1 ∈ Ob(K) with M orK(1, a) (cid:54)= ∅ for all a ∈ Ob(K). (ii) (Stabilisation) Let D = ∪ M orK(1, a). Then for each f, g ∈ D there are a∈Ob(K) morphisms p and q with pf = qg. If pf = qf for f ∈ D then p = q. (iii) (Cancellation) Proposition 2.1.1. If we define (cid:22) on D by f (cid:22) g iff g = pf for some morphism p then D becomes a directed set. Moreover given a functor F from K to some category C then the sets Af , for f ∈ D, together with the maps ιg Af = M orC(Φ(1), Φ(target(f ))) f : Af → Ag when f (cid:22) g and g = pf given by f (v) = Φ(p) ◦ v ιg form a direct system denoted A(Φ). Proof. The proof is just verification of the axioms of directed set and direct system (Note that (cid:22) is not necessarily a partial order, just a from the properties of K. (cid:3) preorder.) For instance the directed set property follows from stabilisation. We will explore the direct limit lim→ A(Φ)f . Recall that the direct limit lim→ Ai of a direct system is by definition the disjoint union of the Ai (which we will call P) ∼= y ∈ Aj ⇐⇒ ∃k with modulo the equivalence relation ∼= defined by x ∈ Ai j (y) . If the ι's are injections then each Φ(S) is naturally i (cid:22) k, j (cid:22) k and ιk identified with a subset of lim→ Φ(S). First we take the functor Φ to be the identity functor I from K to itself. By definition then the direct limit lim→ A(I)f is the quotient of the set P of all ordered i (x) = ιk 4 VAUGHAN F. R. JONES pairs (f, g) with f, g ∈ D by the equivalence relation (f1, g1) ∼= (f2, g2) ⇐⇒ ∃p, q ∈ K such that (pf1, pg1) = (qf2, qg2): lim→ A(I)f = P/ ∼= Now given two elements (f1, g1) and (f2, g2) in P we can choose by stabilisation morphisms p, q ∈ K with pg1 = qf2. Then define (f1, g1)(p,q)(f2, g2) = (pf1, qg2). Proposition 2.1.2. The map from P × P → lim→ A(I)f taking ((f1, g1), (f2, g2)) to [(f1, g1)(p,q)(f2, g2)] depends neither on the choice of (p, q) nor on the choices of (f1, g1) and (f2, g2) in their ∼= equivalence classes. The resulting operation makes lim→ A(I)f into a group. Proof. The proof follows in a relatively routine manner from stabilization and can- cellation. The identity element is [(1, 1)] and the inverse of [(f, g)] is [(g, f )]. We will have to do all the details of well-definedness again to prove the next result so we (cid:3) leave the rest of the proof to be checked then. Definition 2.1.1. The group defined by the previous proposition will be called the group of fractions GK of K. If Φ is not the identity functor we obtain an action of GK on lim→ A(Φ)f . The direct limit lim→ A(Φ)f is the quotient of the set Q of all ordered pairs (f, g) with f ∈ D and g ∈ M orC(Φ(1), target(Φ(f ))) by the equivalence relation (f1, g1) ∼= (f2, g2) ⇐⇒ ∃p, q ∈ K such that (pf1, Φ(p) ◦ g1) = (qf2, Φ(q) ◦ g2): lim→ A(Φ)f = Q/ ∼= Now given an element (f1, g1) ∈ P (as in proposition 2.1) and (f2, g2) in Q we can choose by stabilisation morphisms p, q ∈ K with pg1 = qf2. Then define (f1, g1)(p,q)(f2, g2) = (pf1, Φ(q) ◦ g2). Proposition 2.1.3. The map from P × Q → lim→ A(Φ)f taking ((f1, g1), (f2, g2)) to [(f1, g1)(p,q)(f2, g2)] depends neither on the choice of (p, q) nor on the choices of (f1, g1) and (f2, g2) in their ∼= equivalence classes. The resulting operation defines an action of GK on lim→ A(Φ)f . (f1, g1)((f2, g2)) = (pf1, Φ(q) ◦ g2). If the category C is linear, the action of GK is linear and if moreover the Hom spaces of C are Hilbert spaces and the Φ(f ) are isometries then lim→ A(Φ)f is a pre-Hilbert space and the action of GK is unitary. Each individual Hilbert space (f, Φ(target(f ))) is a Hilbert subspace of lim→ A(Φ)f and hence its Hilbert space completion. A NO-GO THEOREM FOR THE CONTINUUM LIMIT OF A PERIODIC QUANTUM SPIN CHAIN.5 Proof. First suppose p and q are changed to p(cid:48) and q(cid:48). Then by stabilisation there are r and s such that sp(cid:48)g1 = rpg1. So by cancellation sp(cid:48) = rp. Moreover rpg1 = rqf2 and sp(cid:48)g1 = sq(cid:48)f2 hence rqf2 = sq(cid:48)f2 and by cancellatioin rq = sq(cid:48) Thus (pf1, Φ(q)g2) ∼= (rpf1, Φ(rq)g2) = (sp(cid:48)f1, Φ(sq(cid:48))g2) ∼= (p(cid:48)f1, Φ(q(cid:48))g2). To see the action property (or associativity of the group operation), let (f1, g1), (f2, g2) ∈ P and (f3, g3) ∈ Q be given. Choose r, s ∈ K with rg2 = sf3 and p, q such that pg1 = qrf2. Then to calculate [(f1, g1)]([(f2, g2)]([(f3, g3)])) and ([(f1, g1)][(f2, g2)])([(f3, g3)]) we can, by well-definedness, use (pf1, pg1), (qrf2, qrg2) and (qsf3, Φ(qs)g3) instead. Then both expressions yield [(pf1, Φ(qs)g3)]. The assertions about linearity and (cid:3) unitarity are trivial. Examples of groups and representations constructed in this way are more or less interesting depending on how "small" the category is compared to the group it pro- duces. We list a few examples below where our point of view brings nothing new. The first example shows that the construction is universal in some sense but of no interest at all in this case. W = ∩ n∈NT nV on which T acts surjectively hence invertibly by what we call T∞. The isomor- phism takes a pair (p, v) in lim→ Φ to T −n∞ v provided n is sufficiently large that T nv ∈ W . (ii) (Fundamental group) (i) (All groups) If G is a group, consider it as a small category G with one el- ement. It trivially satisfies the conditions of 2.1 and of course GG = G. The representations obtained are just the usual group actions. If X is a path-connected space and G its homotopy groupoid then if we choose 1 to be any point of X the properties of 2.1 are trivially satisfied and one obtains π1(X) from the construction. If the target category C for Φ is Vect and we are given a flat connection on X then Φ can be constructed by parallel transport and one obtains the holonomy representation of π1(X). If we take N ∪ {0} with addition we obtain Z. No new representations will be obtained in this way. A functor Φ to sets is given by the image of 1 which is simply a transformation T of the set (image under Φ of the object of N). If T is invertible then the map (n, x) (cid:55)→ T −n(x) defines a Z-equivariant bijection from lim→ Φ to X. If T is not invertible things are more complicated. (For instance for the identity functor.) We leave it to the reader to work out the answer in general but we observe that if T is a linear transformation of a finite dimensional vector space V then lim→ Φ is (iii) (Integers and rationals) VAUGHAN F. R. JONES 6 (iv) (Braids) See also [4]. If we take the category Bn consisting of positive braids on n strings, with one object, then Garside theory shows that any braid a is of the form ∆kb for some positive braid b and k ≤ 0 where ∆ is the positive half twist braid. So if p, q ∈ Bn then q = qp−1p. Writing qp−1 as ∆kb we see bp = ∆−kq. This shows that stabilisation holds in Bn. Cancellation is obvious and it is clear we get GBn = Bn. If Bn is represented by invertible matrices they define a functor to Vect and we get a representation of Bn. We have not fully analyzed the situation when the matrices representing Bn are not all invertible. We now turn to examples of categories F which we will use to obtain genuinely interesting representations of GF. 2.2. The category of planar forests and Thompson's group F . By "forest" we will mean a planar binary forest whose roots lie on a horizontal line and whose leaves lie on another horizontal line above the roots. Two such forests will be con- sidered the same if they can be isotoped one to another in the obvious way. Here is an example of a forest: Forests form a category F whose objects are N and whose morphisms from m to n are the forests with m roots and n leaves. Obviously M orF(m, n) is only non-empty if m ≤ n. Composition of morphisms is just the obvious stacking of planar forests. Clearly M orF(1, n) is the set of all planar binary rooted trees with n leaves. So ∪nM orF(1, n) is the directed set T of all such trees. It is obvious that, for T ,S ∈ T there is at most one morphism F ∈ F with T ◦ F = S. Moreover any binary planar tree can be completed to a full binary tree with 2m leaves for some large m. Thus the category F satisfies the conditions of 2.1 and we may form the group GF. Proposition 2.2.1. The group GF is isomorphic to Thompson's group F of piecewise linear homeomorphisms of [0, 1]. Proof. See [2] for an explanation of how elements of F can be represented by pairs of binary trees, up to a certain equivalence relation. Check that this is the same as (cid:3) our definition of GF. The construction generalizes immediately to the categories Fm of planar rooted forests all of whose vertices are n + 1-valent. The groups GFm are the Thompson groups Fn (not free groups!) where 2 is replaced by n in the definitions. A NO-GO THEOREM FOR THE CONTINUUM LIMIT OF A PERIODIC QUANTUM SPIN CHAIN.7 2.3. The category of annular forests and Thompson's group T . Any elegance this treatment has derives from the paper [7] of Graham and Lehrer. Definition 2.3.1. We define a rooted, affine binary forest Fm,n to be a planar binary forest which can be drawn in the strip R × [0, 1] ⊂ R2 as a diagram with m roots in the open interval (0, 1) and which is invariant under horizontal translation by Z. The subforest connected to the roots in (0, 1) is to have its n leaves on R × {1} we may suppose none of the leaves has an integral x-coordinate. Here is a picture of an element F3,7: Where the diagram is continued to the left and right by periodicity. The forest below in M orAF(n,n) will be called ρn (illustrated with n = 4): And we will set τn = ρn n. By planarity and peridocity the leaves of the subforest connected to the roots in (0, 1) lie in an interval of length 1. There are exactly m roots in any interval of length 1 on the x-axis and exactly n leaves in any interval of length 1 on the line R × {1}. Rooted affine binary forests may be stacked by lining up the leaf of the bottom forest with smallest positive x coordinate with the root on the top forest with smallest positive x coordinate. Planarity dictates how all the other roots match up with leaves. (Alternatively one could insist that all the roots and leaves lie on specific points so they will line up automatically.) Thus the set AF of all rooted affine binary forests forms a category. Note that M orAF(m, m) is a group isomorphic to Z generated by ρn and every Fm,n can be composed with an element of this group so that it becomes a forest inside [0, 1] × [0, 1] extended to the whole strip by periodicity. With this observation it is clear that AF satisfies the conditions of 2.1 with the number 1 as the object 1. We see moreover that the data of an element Fm,n ∈ AF is the same n ◦ F = Fm,n. (In the as a planar forest F as in 2.2 and a unique integer k so that ρk 2............012............01 8 VAUGHAN F. R. JONES example of a forest we have given above, k = 3 but nothing is to stop it being bigger than n.) The group GAF is a natural central extension T of Thompson's group F which may be defined as piecewise linear periodic foliations of the strip which are smooth except at dyadic rationals and whose lines in the smooth parts have slope a power of 2. To obtain Thompson's group T on the nose from a category construction we take quotient of AF by the action of Z which acts on M orAF(m, n) by composing with powers of τn. That this action is compatible with composition follows from the simple relation τn ◦ Fm,n = Fm,n ◦ τ−1 m So one obtains a category T with the same objects as AF and finitely many morphisms for each n, one for each pair (an element of F with n leaves,an element of Z/nZ). (such a pair obviously represents an orbit under the action of ρn). Thus an element of GT is an equivalence class of pairs of rooted binary trees with the same number n of leaves, one of them marked for each tree. Using a power of ρn, one of the marks can be taken to be the leftmost leaf. Comparing with [2] we see we have obtained Thompson's group T . As for F , one can replace 2 by any larger integer to get affine categories whose groups of fractions are the Thompson groups Tn. 2.4. Thompson's group V and the braided Thompson groups. Thompson's group V is a larger group than F which allows discontinuous piecewise linear maps of the circle that swap the intervals on which an element is linear. Thus any element is given by a pair of binary planar rooted trees together with a permutation of the leaves of one of them which determines how the intervals are to be identified. We can capture this group with our category method by letting V be the category whose objects are N and whose morphisms are pairs consisting (F, π) where F ∈ F and π is a permutation of the leaves of F. The permutations themselves are morphisms in V and the key observation is that for each π ∈ M orV(m, m) and each F ∈ M orF(m, m), there is a σ ∈ M orV(n, n) with σ ◦ F = F ◦ π. We leave it to the reader to make sense of this and how it yields a well defined category structure on V whose group of fractions is V . For the braided Thompson BF group the situation is very similar, the category BF consists of pairs (F, α) where α is an n − string braid where n is the number of leaves of F. See [5] for the definitions of braided Thompson groups. 3. Review and development of the action of the Thompson groups on the semicontinuous limit. 3.1. How to obtain representations, unitary and otherwise. The previous section would be no more than a curiosity were it not for the fact that we can A NO-GO THEOREM FOR THE CONTINUUM LIMIT OF A PERIODIC QUANTUM SPIN CHAIN.9 now mechanically and uniformly construct actions of all the Thompson groups using functors. We will use the language of planar algebras (see the appendix) but we would like to make it clear once again that all the essential ideas and many interesting examples are exhibited in the tensor planar algebra so one needs to understand no more about planar algebras than how diagrams specify ways to contract tensors. The common ingredient is a *-planar algebra P = (Pn) and an element R of Pn+1 where we are dealing with Thompson groups relevant to rationals of the form a nk with a and k integers. The element R must satisfy the "unitarity" condition: Where there are n strings joining the discs containing R and R∗. Such an R will give rise to representations of the various forest categories of section 2.1. We fix P and R and treat each case individually as there are some caveats. (i) Thompson's group F . Here R is in P3. Let V = (Vn) be a representation of the rectangular category R of P (In the case of tensors, the objects are just the tensor powers of a fixed vector space , with morphisms being tensors mapping between the different tensor powers). A morphism in M orR(m, n) is just a rectangle with m marked points on the bottom and n on the top, filled with elements of P in discs connected by strings among themselves and to the marked points on the boundary. Definition 3.1.1. Let F be a rooted planar forest in M orF(m, n). Define ΦR(F) to be the element of M orR(m, n) obtained by replacing every vertex in F by a disc containing a copy of R as follows: F = Φ(F) = =RR*..............RRRRR 10 VAUGHAN F. R. JONES Proposition 3.1.1. The map ΦR defines a functor from F to the category V ect of vector spaces and linear maps. Proof. This is trivial, ΦR takes the object n ∈ F to Vn and the functor property (cid:3) follows from stacking of diagrams in planar algebras. (If one prefers, one could take an appropriate tensor category C with a fixed object V , let Vn = ⊗nV and choose an element R ∈ Hom(V, V ⊗V ). The previ- ous proposition is then just the well-known pictorial composition of morphisms in a tensor category.) We now come to the main object of interest in this paper. Definition 3.1.2. Let P, V and R ∈ P3 as above be given and form the functor ΦR. By section 2 we then have a direct system f (cid:55)→ A(ΦR)f of vector spaces on the directed set T of binary planar rooted trees. The vector space VR = lim→ A(ΦR)f will be called the semicontinous limit vector space for R. V contains all of the spaces Vn (of V) embedded one in the other by the maps Φ(F) defined above. If the planar algebra has positivity, e.g. a subfactor planar algebra, and R satisfies unitarity ,the inclusion maps in the direct limit are isometries so the semicontinuous limit V has a preHilbert space structure . The Hilbert space completion of TR will be denoted HR and called the semicontinous limit Hilbert space. Note that the pre-Hilbert space structure on VR is preserved by the (linear) action of F . Thus this action extends to a unitary representation πR on HR. As an exercise, let us calculate a coefficient of πR. Suppose Vn = Pn and that dimP1 = 1. Choose a unit vector Ω ∈ P0 (the "vacuum"). Let g ∈ F be given by the pair of trees (T1, T2). We want to calculate (cid:104)πR(g)Ω, Ω(cid:105). Let ι be the tree with one vertex and no edges. By definition πR(g)Ω is (T1, T2)((ι, Ω)). See proposition 2.1.3 from which we see that this can also be written (T1, Φ(T2)(Ω)). But also Ω = (T1, Φ(T1)(Ω). We see that (cid:104)πR(g)Ω, Ω(cid:105) = (cid:104)Φ(T2)(Ω), Φ(T1)(Ω)(cid:105), the inner product being taken in the planar algebra. For instance if T1 = and T2 = then A NO-GO THEOREM FOR THE CONTINUUM LIMIT OF A PERIODIC QUANTUM SPIN CHAIN.11 = (cid:104)πR(g)Ω, Ω(cid:105) so that if δ is the loop parameter of the planar algebra then (cid:104)πR(g)Ω, Ω(cid:105) = 1 δ (cid:40) Ri,j,k = There are many interesting choices of R. For instance if the planar algebra is the tensor planar algebra on a vector space of dimension 3 with orthogonal basis {1, 2, 3} then we may define R to be the 3-tensor 0 if i = j or j = k or i = k 1 otherwise . Then (cid:104)πR(g)Ω, Ω(cid:105) is equal to the num- ber of ways of 3-colouring the edges of the three valent graph underlying the diagram for this inner product in such a way that the 3 colours at any ver- tex are distinct. The positivity of these coefficients for all g ∈ F is known to be equivalent to the 4-colour theorem! We are grateful to Roland Bacher for pointing this out-see [18]. Or, if the planar algebra is the version of the tensor planar algebra in which the n indices sit in the regions and (cid:40) 0 if i = j or j = k or i = k 1 otherwise . = then n(cid:104)πR(g)Ω, Ω(cid:105) is the number of ways of n-colouring the map defined by the *RRR**RRR*RRR*R*RRRkij 12 VAUGHAN F. R. JONES diagram for this inner product. If n = 3 the map can be coloured in 6 ways or not at all so we find that the set of all g for which (cid:104)πR(g)Ω, Ω(cid:105) = 2 is a subgroup of F . Yunxiang Ren has shown that it is isomorphic to the Thompson group F4. (ii) Thompson's group T. Here things work almost exactly as they do for F . One takes the same kind of R as before and an affine representation of the planar algebra. Replacing the vertices of morphisms in AF by discs containing R gives a functor from AF to Vect, taking n to the n vector space of the affine represen- tation. We thus get a representation of T . If the representation is irreducible and the rotation acts as a scalar then one obtains a projective representation of T . The unitary affine representations of the Temperley-Lieb algebra are well understood for all values of the loop parameter - see [7],[11] and [13]. The same examples of R as for F yield similar interpretations of coefficients. Definition 3.1.3. The semicontinuous limit vector space and Hilbert space VR and HR are defined in exactly the same way in this annular context they were in 3.1.2 for rectangular representations of the planar algebra. Note that T acts unitarily on HR if the planar algebra has positivity and R satisfies unitarity. It is clear that the projective representation of T will be an ordinary repre- sentation if the affine representation is in fact annular (see B.0.7). (iii) Thompson's group V. The representations are easiest to describe if we use the tensor planar algebra based on an underlying vector space V . We can choose any tensor R with three indices satisfying the unitarity condition. A permutation π in Sn defines a linear map ⊗nπ on ⊗nV by permuting coordinates so for (F, π) ∈ M orV(m, n) we may define Φ((F, π)) by filling in F's vertices with discs containing R as before then composing the corresponding linear map from ⊗mV to ⊗nV with ⊗nπ. It is easy to check that this Φ is a functor and hence defines a unitary representation of V . Note that there is a purely diagrammatic way to represent the category V by drawing permutations as strings connecting permuted points. So if one could find a planar algebra (Pn) with an element of P4 satisfying the obvious relations of a transposition: such that and then one will get a representation of V provide the R ∈ P3 and the crossing in P4 satisfy: == A NO-GO THEOREM FOR THE CONTINUUM LIMIT OF A PERIODIC QUANTUM SPIN CHAIN.13 This condition is automatic if the crossing is the transposition acting on V ⊗ V . (iv) Braided Thompson group. This works just like for V . There is a purely dia- grammatic representation of morphisms in BF which is just like the the one for V except that the transposition is allowed to be a crossing: satisfying the braid group relations and the following two relations with the vertices of the trees: = and = . Representations of BF are easy to come by in planar algebras/tensor cate- gories. Coeffiecients of the form (cid:104)gΩ, Ω(cid:105) are just the partition functions in the corresponding planar algebra. Let us make one curious remark. We saw that the braid group can be ob- tained as the group of fractions of the semigroup of positive braids. So it is with BF which we can make smaller by requiring that all the crossings be pos- itive. Then to obtain a representation of BF we only need the first of the two relations above between the crossing and the trivalent vertex. We have not investigated this. 3.2. The relation between these constructions and those of [12]. The repre- sentations of F obtained in [12] may be obtained by the construction of this paper by first embedding F = F2 in F3 by taking a pair of trees (T1, T2) ∈ F2 and adding strings to turn all the trivalent vertices into quadrivalent ones, obtaining the pair ( T1, T2) as illustrated below: (T1, T2) =(cid:0) (cid:1) → ( T1, T2) =(cid:0) (cid:1) Thus it was possible to use elements R ∈ P4 rather than R ∈ P3 to obtain representations of F2. Otherwise the construction of representations of F [12] was just a more clumsy version of what we have done in this paper in much greater generality. =RR,, 14 VAUGHAN F. R. JONES 4. The NoGo theorem. The Thompson group T contains the subgroup Rot of rotations of the circle by dyadic rationals. We will prove that the representation of Rot on the semicontinuous limit Hilbert space (from an affine representation of a positive definite planar algebra) is highly dicontinuous if we topologise Rot as a subgroup of R/Z. This is not at all surprising. The geometric structure underlying the semicontinuous limit is the full binary tree whose branches are dangling and do not feel the topology of the circle. The discontinuity result is true in great generality but we will only prove it for a single family of planar algebras (with positive definite inner product). We have chosen this family because there is, up to an irrelevant scalar, only one choice of R. To be precise, let Q = (Qn) be the planar algebra obtained from the TL planar algebra with loop parameter δ = 2 cos π/n for n = 6, 7, 8, 9,··· by cabling 2 strings and cutting down by the JW idempotent (this is also quantum SO(3) at a root of unity). See [15]. We can represent the JW idempotent in T L4 with 4 boundary points as a box , entirely defined by the relations , and . One checks that = δ2 − 1. In Q it is well known that Q3 is spanned by the single element (cid:114) δ R = δ2 − 2 , the normalisation guaranteeing unitarity. Q is obtained by combining the cabled strings to a single string. Thus R is an element of Q3 and in Q the loop parameter is d = δ2 − 1. Note that R is rotationally invariant so we will suppress it in all pictures, i.e. from now on will mean . Now let H be the semicontinuous limit Hilbert space constructed in 3.1.2 and 3.1.3 from the planar algebra Q and an annular representation V = (Vn) of it, using the element R defined above to construct the functor Φ. By section 3.1 we know that Thompson's group T acts unitarily on H. In particular for every dyadic rational 2n ∈ [0, 1) we have a unitary ρr on H representing the rotation by a in T . r = a We have proved the following for d = 4 cos2 π/n + 1 for n ∈ N, 7 ≤ n ≤ 20 and d = 3. It is false for n = 5 and n = 6 and surely true for all n ≥ 7. ξ, η(cid:105) = 0. Theorem 4.0.1. For any vectors ξ, η ∈ H, n→∞(cid:104)ρ 1 lim 2n =*== 0R A NO-GO THEOREM FOR THE CONTINUUM LIMIT OF A PERIODIC QUANTUM SPIN CHAIN.15 Proof. Note that since the representation is unitary we may suppose that ξ and η are actually in some space (T, V2k ) where T is the annular tree (shown for n = 4). The following diagram is (cid:104)ρ ξ, η(cid:105) which 1 2k+n+1 we illustrate here for k = 1 and n = 3. Note that we are applying periodic boundary conditions. Now all the regions in the blue dotted circles can be isotoped to look like so if we call x this element of Q4 the picture becomes: ηξ 16 VAUGHAN F. R. JONES (The positions of the $ signs in the picture are obvious, so suppressed.) We recognise the transfer matrix (see appendix B) T2n+k (x) ! Definition 4.0.1. We define the bilinear map B : Q4 × Q4 → Q4 by B(x, y) = and the renormalisa- tion map R(x) = B(x, x) Observe that B makes Q4 into a commutative non-associative algebra for which R is the squaring operation. We see the inner product formula becomes (if y = R(x)): ηxxxxxxxxxxxxxxξxxyx A NO-GO THEOREM FOR THE CONTINUUM LIMIT OF A PERIODIC QUANTUM SPIN CHAIN.17 Continuing in this way we see that (cid:104)ρ 1 2k+n+1 ξ, η(cid:105) = (cid:104)T2k (Rn(x))ξ, η(cid:105) We thus have to understand the iterates of the renormalisation transformation R : Q4 → Q4. We begin by calculating R explicitly. For this we use the basis { } of Q4 and write an arbitrary element of Q4 as , , a = p + q + r . Since B is bilinear it is easy to expand and compute R(a) using the skein relations in Q. A sufficient set of relations is the following (see [15]): = 0, = d−2 d−1 ( ) , and of course unitarity, . (A quick way to deduce the second picture-both sides are eigenvalues for the rotation of π/2 with eigenvalue −1. But, modulo the TL subspace, the rotation has eigenvalue +1 by looking at a spanning set of TL diagrams reduced by the JW. Thus the two sides of the equation are proportional and the constant can be obtained by capping and using unitarity.) With these relations it is not hard to show that: R(a) = { d2 − 5d + 7 pr + q2 + r2} d − 2 (d − 1)2 p2 + 2pq + 2 d − 1 − { (2pq + q2)} (d − 1)3 p2 + d − 1 +{ d2 − 3d + 3 1 (d − 1)3 p2 + d − 1 1 1 (2pq + q2)} . ηyyyyyyyyξ 18 VAUGHAN F. R. JONES Completing some squares we get R(a) = {(p + q)2 + (r + 1 (d − 1) − { +{(p + q)2 + (r + (p + q)2 − d(d − 2) d − 2 d − 1 p2} (d − 1)2 p)2 − (d + 1)(d − 2) (d − 1)2 p2} p)2 − (d + 1)(d − 2) d − 2 d − 1 (d − 1)2 + q Now define the norm − 1 on Q4 by p p2} + r . 1 = p + q + r . Then the above shows that R(a)1 ≤ d + 1 d − 1 (p + q)2 + (r + d − 2 d − 1 p)2 + d(d + 1)(d − 2) (d − 1)3 p2 By convexity the maximum of the right hand side on the − 1 unit ball is M = d + 1 d − 1 + ( d − 2 d − 1 )2 + d(d + 1)(d − 2) (d − 1)3 . Hence R(a)1 ≤ Ma2 1 and if there is an n for which Rn( then Rn+1( )1 < KMRn( )1 < K for some K > 0 with M K < 1 )1 and limn→∞ Rn( ) = 0. the values of d mentioned before the statement of the theorem. Computer calculations show that such an n exists (indeed is rather small) for all Now consider the following element Y ∈ Hom(V2k+2, V2k+2), with y = Rn( ) (illustrated for n = 3): Y = , and the following elements ξ and η of V2k+2: ξ = η = (Remember that we have imposed periodic horizontal boundary conditions.) Then a picture shows that (cid:104)Y ξ, η(cid:105) = (cid:104)T2k (Rn(x))ξ, η(cid:105) But we can now easily estimate Y for it is a composition y1y2 ··· yn where yi is the element of Hom(V2k+2, V2k+2) with a copy of y between the (i + 1)th. and yyyyyyyyξη A NO-GO THEOREM FOR THE CONTINUUM LIMIT OF A PERIODIC QUANTUM SPIN CHAIN.19 (i + 2)th. boundary points as illustrated below: yi = But the norms of the yi are all equal to the norm of y as an element of the C∗- algebra Q4. And, all norms being equivalent, we have shown that Y → 0 as n → ∞. This proves the theorem. (cid:3) Appendix A. Some notions of planar algebra. In this paper a planar algebra P will be a graded vector space Pn, graded by N∪{0} and admitting multilinear operations indexed by planar tangles T which are subsets of the plane consisting of a large (output) circle containing smaller (input circles). There are also non-intersecting smooth curves called strings whose end points, if they have any, lie on the circles where they are called marked points. Elements of P are "inserted" into the input circles with an element of Pn going into a disc with n marked points, and the result of the operation specified by the tangle is in Pk where there are k marked points on the output circle. In order to resolve cyclic ambiguities, each of the circles of T comes with a privileged interval between marked points which we will denote in pictures by putting a $ sign near that interval. The $ signs are used to define an obvious notion of gluing of one tangle inside an internal disc of another. Here is an example of a planar tangle: The result of the operation indexed by T on elements v1, v2,··· , vn of P is denoted ZT (v1, v2,··· , vn) where there are n input discs. See [8] for details. The operation ZT depends only on T up to smooth planar isotopy so one has a lot of freedom drawing the tangles, in particular the circles may be replaced by rectangles when it is convenient. The operations ZT are compatible with the gluing of tangles. Tangles may also be "labelled" by actually writing appropriately graded elements of P inside some of the internal circles. ....i+1 i+2....y$Output circleInput circlesStrings$$$Marked points 20 VAUGHAN F. R. JONES It is a very useful convention to shrink the input discs in a planar tangle to points so that the boundary intervals of the circle become the regions adjacent to the points. And for labelled tangles one places the label in the region corresponding to the $ sign. Thus is represented by the picture We will also often omit the output disc and/or dollar signs provided they are obvious in context. Definition A.0.2. Given a planar tangle T all of whose internal circles are labelled by v1, v2,··· , vn we call ZT (v1, v2,··· , vn) the element of Pk which it defines. If k = 0 and the dimension of P0 is one, this may be identified with a scalar using the rule that Z(emptytangle) = 1. Planar tangles can be glued in an obvious way along input circles and the opera- tions ZT are by definition compatible with the gluing. For connections with physics and von Neumann algebras, planar algebras will have more structure, namely an antilinear involution ∗ on each Pn compatible with orientation reversing diffeomorphisms acting on tangles. If, moreover, dimP0 = 0 we get a sesquilinear inner product (cid:104)S, R(cid:105) on each Pn given by . A planar algebra will be called positive definite if this inner product is. Our planar algebras will all have a parameter δ which is the value of a closed string which may be removed from any tangle with multiplication by the scalar δ. Two examples of planar algebras should be mentioned. The first is the Temperley- Lieb algebra T L (which has its origins in [19] though its appearance here should properly be attributed to [14], via [1]-see also [10]). A basis of T Ln consists of all isotopy classes of systems of non-crossing strings joining 2n points inside the disc. In particular T Ln is zero if n is odd. The planar algebra operations are the obvious gluing ones with the rule that any closed strings that may be formed in the gluing process are discarded but each one counts for a multiplicative factor of δ, called the "loop parameter". The * structure is given by complex conjugation on basis diagrams, extended by conjugate linearity. This planar algebra is positive definite iff δ ≥ 2. If δ = 2cosπ/n for n = 3, 4, 5,··· T L admits a quotient planar algebra which is positive definite. The second examples of planar algebras which we will use are the tensor planar algebras. For fixed integer k ≥ 2 one considers A Hilbert space V of dimension k with a basis so that elements of the tensor power ⊗nV may be represented as tensors with n indices, each index running from 1 to n. The planar algebra P⊗ is then defined $PQR$$$$PQRRS*** 0 = C, and for n ≥ 1, P ⊗ A NO-GO THEOREM FOR THE CONTINUUM LIMIT OF A PERIODIC QUANTUM SPIN CHAIN.21 by P ⊗ n = ⊗nV . The action of planar tangles on tensors is nothing but contraction of tensors along the indices connected by the the strings of the tangle, together with the rule that indices have to be constant along the strings. The tensor planar algebras P⊗ are positive definite when given the *-structure R∗ i1,i2,··· ,in = ¯Rin,in−1,··· ,i1. Appendix B. The affine category of a planar algebra. Definition B.0.3. The affine category Af f (P) is the (linear) category whose objects are sets ¯m of m points on the unit circle in C ,and whose vector space of morphisms from ¯m to ¯n is the set of linear combinations of labelled tangles (with marked boundary points ¯m∪ ¯n) between the unit circle and a circle of larger radius modulo any relations in P which occur within contractible discs between the unit circle and the larger circle. Composition of morphisms comes from rescaling and gluing the larger circle of the first morphism to the smaller circle of the second. If P is positive definite the morphism spaces of Af f (P) admit an adjoint x (cid:55)→ x∗ obtained by reflecting a labelled annular tangle about a circle between the inner and outer circles of the tangle and taking the ∗'s of the labels. Use of ¯m adds to clutter so we will abuse notation by using just m for an object of Af f (P) with m points. We could also just suppose that the boundary points are always just the roots of unity. One needs to be careful with this definition (see [11],[7]). In a representation of Af f (P), morphisms may be changed by planar isotopies without affecting the action, but the isotopies are required to be the identity on the inner and outer circles. Thus the tangle of rotation by 360 degrees does not necessarily act by the identity in a representation of Af f (P). The representations we will consider of Af f (P) are called lowest weight modules and may be defined as in [11] by taking a representation W of the algebra M or(n, n) for some n (the "lowest weight") and inducing it in the obvious way. This may cause problems with positive definiteness but it is known that subfactor planar algebras possess a host of such representations. The vector spaces Vk of such a lowest weight representation are zero if k ≤ n and spanned by diagrams consisting of a vector w ∈ W inside a disc with n marked points, surrounded by a labelled planar tangle of P with k marked points on the output circle. lowest weight space V2: and the action of a morphism in Af f (P) on it: Here is a vector w in a V6 created by the action of an affine morphism on v in the w=vRS 22 VAUGHAN F. R. JONES and here is a diagram illustrating the result of acting on the above vector w with a morphism in M or(6, 4): The planar algebra itself defines an affine representation simply by applying an- nular labelled tangles to elements of P. This representation is irreducible and plays the role of the trivial representation. In the TL case which is what we will mostly consider, irreducible lowest weight representations are parametrized by their lowest weight (the smallest n for which Vn is non-zero), and a complex number of absolute value one which is the eigenvalue for the rotation tangle. The case n = 0 is exceptional and the rotation is replaced by the tangle which surrounds an element v of V0 by a circular string. If v is an eigenvector for this tangle and there are some restrictions on the eigenvalue µ-see [11],[13]. The case where µ = δ is precisely the trivial representation. In this case the vector v ∈ V0 is the empty diagram so it never features in pictures. Definition B.0.4. An affine representation V = Vk of a positive definite planar algebra will be called a Hilbert representation if each Vk is equipped with a Hilbert space inner product which is invariant in the sense that (cid:104)aξ, η(cid:105) = (cid:104)ξ, a∗η(cid:105) where a ∈ M or(m, n), ξ ∈ Vn, η ∈ Vm and ∗ is the structure we defined earlier on the affine category. There are two particularly important affine tangles which play a big role in our main theorem. Definition B.0.5. If S ∈ P4 we define the "transfer matrix" Tn(S) to be the element of M or(n, n) defined by the following annular tangle: which it will be more convenient to draw horizontally with implcit periodic bound- ary conditions thus: RvRSRRRSSSSSS A NO-GO THEOREM FOR THE CONTINUUM LIMIT OF A PERIODIC QUANTUM SPIN CHAIN.23 (Here we have illustrated with n = 6.) The second important tangle is the rotation Definition B.0.6. The rotation ρn ∈ M or(n, n) is given by: (illustrated for n = 8). Definition B.0.7. An affine representation V = Vk of a planar algebra will be called an annular representation if the rotation by 2π (= ρn n) acts by the identity. ρn generates a copy of Z inside M or(n, n). References [1] Baxter, R. J. (1982). Exactly solved models in statistical mechanics. Academic Press, New York. [2] Cannon, J.W., Floyd,W.J. and Parry, W.R.(1996) Introductory notes on Richard Thompson's groups. L'Enseignement Mathématique 42 215–256 [3] Cirac, J. I. and Verstraete, F. (2009) Renormalization and tensor product states in spin chains and lattices. JOURNAL OF PHYSICS A-MATHEMATICAL AND THEORETICAL 42 (50) [4] P. Dehornoy, F. Digne, E. Godelle, D. Krammer, J. Michel, Foundations of Garside Theory. arXiv:1309.0796 354 Ð 409 [5] P.Dehornoy (2006), The group of parenthesized braids, Advances in Mathematics 205 (2006) [6] G. Evenbly, G. Vidal, Tensor Network Renormalization, arXiv:1412.0732 [7] J. J. Graham and G.I. Lehrer, The representation theory of affine Temperley Lieb algebras, L'Enseignement Mathématique 44 (1998), 1–44. [8] V.F.R. Jones, Planar Algebras I, preprint. math/9909027 [9] Jones, V. F. R. In and around the origin of quantum groups. Prospects in mathematical physics. Contemp. Math., 437 Amer. Math. Soc. (2007) 101–126. math.OA/0309199. [10] Jones, V. F. R. (1989). On knot invariants related to some statistical mechanical models. Pacific [11] V.F.R. Jones, The annular structure of subfactors, in "Essays on geometry and related topics", Journal of Mathematics, 137, 311–334. Monogr. Enseign. Math. 38 (2001), 401–463. [12] V.F.R. Jones (2014) Some unitary representations of Thompson's groups F and T, [13] Jones, V. and Reznikoff, S. (2006) Hilbert Space representations of the annular Temperley-Lieb algebra. Pacific Math Journal, 228, 219–250 [14] Kauffman, L. (1987). State models and the Jones polynomial. Topology, 26, 395–407. [15] S. Morrison, E.Peters, N. Snyder, (2015) Categories generated by a trivalent vertex. arXiv:1412.7740 arXiv:1501.06869 .....SSSSSS..... 24 VAUGHAN F. R. JONES [16] V Pasquier, H Saleur, Common structures between finite systems and conformal field theories through quantum groups Nuclear Physics B 330 (2), 523-556 [17] Penrose, R. (1971). Applications of negative dimensional tensors. Applications of Combinatorial Mathematics, Academic Press, 221–244 [18] R.Thomas , (1998) An Update on the Four-Color Theorem, Notices of the AMS (45) 848-859 [19] Temperley, H. N. V. and Lieb. E. H. (1971). Relations between the "percolation" and "colouring" problem and other graph-theoretical problems associated with regular planar lattices: some exact results for the "percolation" problem. Proceedings of the Royal Society A, 322, 251–280.
1604.02539
1
1604
2016-04-09T08:41:22
Topologically irreducible representations of the Banach *-algebra associated with a dynamical system
[ "math.OA" ]
We describe (infinite-dimensional) irreducible representations of the crossed product C$^*$-algebra associated with a topological dynamical system (based on $Z$) and we show that their restrictions to the underling $\ell^1$-Banach $*$-algebra are not algebraically irreducible under mild conditions on the dynamical system. The above description of irreducible representations has two ingredients, ergodic measures on the space and ergodic extensions for the tensor product with type I factors; the latter which may not have been explicitly taken up before will be explored by examples. A new class of ergodic measures is also constructed for irrational rotations on the circle.
math.OA
math
Topologically irreducible representations of the Banach ∗-algebra associated with a dynamical system Aki Kishimoto∗and Jun Tomiyama† April 2016 Abstract We describe (infinite-dimensional) irreducible representations of the crossed prod- uct C∗-algebra associated with a topological dynamical system (based on Z) and we show that their restrictions to the underling ℓ1-Banach ∗-algebra are not alge- braically irreducible under mild conditions on the dynamical system. The above description of irreducible representations has two ingredients, ergodic measures on the space and ergodic extensions for the tensor product with type I factors; the latter which may not have been explicitly taken up before will be explored by ex- amples. A new class of ergodic measures is also constructed for irrational rotations on the circle. sentation, ergodic measure, ergodic extension Keywords: dynamical system, Banach ∗-algebra, C∗-algebra, irreducible repre- Mathematics Subject Classification: 46H15, 37A05, 46L99 1 Introduction Let X be a compact metrizable space and σ a homeomorphism of X, which forms a classical dynamical system Σ = (X, σ). The corresponding C∗-dynamical system is (C(X), α) where α is the automorphism of the continuous functions C(X) on X defined by α(b)(x) = bσ−1(x), x ∈ X for b ∈ C(X). We denote by ℓ1(Z, C(X)) the Banach space of ℓ1 functions from Z into C(X), which is a Banach ∗-algebra denoted by ℓ1(Σ) when equipped with a product and a ∗-involution as follows: (f g)(n) =Xk f (k)αk(g(n − k)) and ∗E-mail: [email protected] †E-mail: [email protected] f ∗(n) = αn(f (−n))∗. 1 We denote by δn ∈ ℓ1(Σ) for n ∈ Z the function δn(k) = δn,k on Z and by f ∈ C(X) the function f δ0 : Z → C(X). Thus δ∗ n = δ−n and δ1f δ−1 = α(f ) for f ∈ C(X). We denote by C ∗(Σ) the enveloping C∗-algebra of ℓ1(Σ), also identified with the crossed product C∗- algebra of C(X) by α. Note that any topologically irreducible representation of ℓ1(Σ) (on a Hilbert space) extends to an irreducible representation of C ∗(Σ) and that the universal C ∗-norm on ℓ1(Σ), by which C ∗(Σ) is defined, is determined by these representations. Each of C ∗(Σ) and ℓ1(Σ) (as a norm-closed algebra generated by C(X) and δ±1) en- ables us to recover Σ and so is as good as the other in this sense. Though ℓ1(Σ) looks more tamable with its explicit definition, a close examination on ℓ1(Σ) reveals complex- ity or irregularity as an algebra which C ∗(Σ) glosses over in exchange of adopting a representation-friendly intangible norm. A difference between the two objects seems to most manifestly appear in the case of the simplest example with X a singleton, C(T) and ℓ1(Z), where T = R/Z is the dual of Z. Then the convolution algebra ℓ1(Z) is known to have a non-self-adjoint closed ideal while the closed ideals of C(T) are all self-adjoint. (This fact is directly translated into a general Σ if σ has a finite orbit.) Another difference may be found on the lines of Kadison's result: If a representation of a C∗-algebra is topo- logically irreducible, then it is automatically algebraically irreducible ([1]; see also [7, 6]). We naturally expect that a topologically irreducible, infinite-dimensional, representation of ℓ1(Σ) is not algebraically irreducible. Indeed this is shown for irreducible represen- tations induced from aperiodic orbits in X ([5]). There must be other properties which exhibit a stark difference between these two objects, deserving thorough investigation but beyond the scope of our present research. Thus we are here confined to the problem of irreducible representations. (See [2, 3, 4, 5, 8, 9, 10] for the ideal structures and some irreducible representations). We will show this algebraic non-irreducibility for all infinite-dimensional irreducible representations of ℓ1(Σ) if σ preserves a metric on X which induces the right topology and will give a sufficient condition in other cases. For this purpose we first give a procedure for constructing irreducible representations of C ∗(Σ) in terms of ergodic σ-quasi-invariant probability measures on X and some ergodic extensions of the transformation induced by σ (Proposition 2.2). Then we give the aforementioned result on algebraic non-irreducibility of representations of ℓ1(Σ) (Theorem 3.4). In Section 4 we elucidate how ergodic exten- sions may be possible by examples. Specifically given an ergodic transformation σ on a probability space L∞(X) we ask a question of whether σ can be extended to an ergodic transformation on L∞(X)⊗Mn when n < ∞. We manage to give a positive answer in the case of Bernoulli shifts (Proposition 4.5) and irrational rotations on the circle (Proposition 4.11) by specifying a certain form of unitaries in L∞(X, µ) ⊗ Mn for this extension. We also work on unitary equivalence among those ergodic extensions (Propositions 4.7, 4.8, 4.13 and 4.14). But we leave the problem unanswered for general ergodic transformations. Finally we construct a new class of ergodic quasi-invariant probability measures on the circle for an irrational rotation, which is neither atomic nor Lesbegue, where the condition of ergodicity seems to require a specific proof (Proposition 5.4). 2 2 Irreducible representations Let π be an irreducible representation of C ∗(Σ) and let µ be a probability measure on X such that πC(X) extends to an isomorphism from L∞(X, µ) onto π(C(X))′′. Let U = π(δ1), a unitary satisfying Ad Uπ = πα on C(X), which implies that µ must be σ-quasi-invariant. Since π(C(X))′′ ∩ U ′ ⊂ π(C(X))′ ∩ U ′ = π(C ∗(Σ))′ = C1, we conclude that Ad U acts on π(C(X))′′ ergodically; thus µ is ergodic. Lemma 2.1 Let π be an irreducible representation of C ∗(Σ) on a Hilbert space Hπ. Then there is an ergodic σ-quasi-invariant probability measure µ on X and a Hilbert space H such that Hπ is identified with L2(X, µ) ⊗ H and π(f ) = Mf ⊗ 1 for f ∈ C(X), where Mf denote the multiplication of f on L2(X, µ). Proof. Note that Hπ is separable. The commutant π(C(X))′ is isomorphic to L∞(X, µ)En ⊗ B(Hn) Mn where En is a projection in L∞(X, µ) and Hn is an n-dimensional Hilbert space with n including infinity. Since Ad U acts on π(C(X))′ ergodically, we conclude that En must be zero or 1 and there is only one direct summand. Hence π(C(X))′ is unitarily equivalent to L∞(X, µ) ⊗ B(H) for some Hilbert space H where π(f ) corresponds to Mf ⊗ 1. QED Define a unitary V on L2(X, µ) by (V ξ)(x) = (dµσ−1/dµ)1/2(x)ξσ−1(x). Then for f ∈ C(X) (V Mf ξ)(x) =(cid:0) dµσ−1 dµ (x)(cid:1)1/2 (Mf ξ)(σ−1(x)) = f (σ−1(x))(V ξ)(x), which implies that V Mf V ∗ = Mα(f ). Ad(U(V ⊗ 1)∗) defines an automorphism of Z = L∞(X, µ)⊗B(H) which acts trivially on its center. Hence there is a unitary W ∈ Z such that Ad(U(V ⊗1)∗) = Ad W on Z, i.e., U(V ⊗ 1)∗W ∗ ∈ Z ′ ⊂ Z (Proposition 8.9.2 of [6]). We may suppose that U = W (V ⊗ 1) by further modifying W by a central unitary of Z if necessary. Proposition 2.2 All the irreducible representations of C ∗(Σ) are constructed as follows: Choose an ergodic σ-quasi-invariant probability measure µ on X and find a unitary W ∈ L∞(X, µ) ⊗ B(H) for some Hilbert space H such that Ad(W (V ⊗ 1)) acts ergodically on L∞(X, µ) ⊗ B(H) where V is the unitary induced by σ as above. Then one can define an irreducible representation π of C ∗(Σ) on L2(X, µ) ⊗ H by π(f ) = Mf ⊗ 1, f ∈ C(X), π(δ1) = W (V ⊗ 1). Let us denote the above representation by π(µ,H,W ). Then πi = π(µi,Hi,Wi), i = 1, 2 are unitarily equivalent with each other if and only if µ1 and µ2 are absolutely continuous with each other and dim(H1) = dim(H2) and there is a unitary operator ζ from L2(X, µ2)⊗H2 onto L2(X, µ1) ⊗ H1 such that π1(f ) = ζπ2(f )ζ ∗, f ∈ C(X) and W1 = ζW2 ¯α(ζ)∗, where ¯α(ζ) = (V2 ⊗ 1)ζ(V1 ⊗ 1)∗ and Vi is the V defined for µ = µi. 3 Proof. The first half is proved before this proposition. The unitary equivalence is by definition the existence of ζ above. The other conditions are redundant but follow from this. QED Proposition 2.3 If π = π(µ,H,W ) with dim(H) > 1 then U = π(δ1) has no eigenvectors. Moreover U does not satisfy the equality Uξ = Y ξ for any unit vector ξ ∈ L2(X, µ) ⊗ H and any unitary Y ∈ L∞(X, µ) ⊗ 1. Proof. Suppose that U = π(δ1) has a generalized eigenvector, say Uξ = Y ξ for some unit vector ξ ∈ L2(X, µ) ⊗ H and some unitary Y ∈ L∞(X, µ) ⊗ 1. Since (dµσ−1/dµ)1/2(x)ξ(σ−1(x)) = W (x)∗Y (x)ξ(x) a.e. we deduce that the set of x with kξ(x)k = 0 is σ-invariant. Hence ξ(x) 6= 0 a.e. Let e1(x) = ξ(x)/kξ(x)k, x ∈ X, which forms a vector e1 in L2(X, µ) ⊗ H. There is a family ei = ei(x), i = 2, 3, . . . of vectors in L2(X, µ) ⊗ H such that (ei(x))i≥1 is a complete orthnormal system in H for almost all x (3.3 of [7]). Hence the projection onto the closed subspace π(C(X))ξ is a proper projection in the commutant of π(C ∗(Σ)), which contradicts the irreducibility of π. Thus U does not have a generalized eigenvector. QED If µ is σ-invariant and dim(H) = 1 in π = π(µ,H,W ) then U = π(δ1) = W V satisfies U1 = W 1 where 1 is regarded as a function in L2(X, µ). Lemma 2.4 Suppose that µ is an ergodic σ-quasi-invariant probability measure on X. Then µ is either atomic in which case there is x ∈ X such that {σn(x) x ∈ Z} has measure 1, or completely non-atomic. Proof. Note that µ is the sum of an atomic part and a completely non-atomic part and that the decomposition into these parts is unique. Since µ is ergodic, one of them must be zero. If it is atomic then µ must be supported by an orbit as it is ergodic. QED Proposition 2.5 Let π = π(µ,H,W ) be an irreducible representation of C ∗(Σ) and suppose that µ is atomic. Then H = C. Moreover if µ has infinite support then π is unitarily equivalent to π(µ,C,1) and if µ consists of k atoms then π is unitarily equivalent to π(µ,C,λ) where λ ∈ {e2πiθ ∈ C 0 ≤ θ < 1/k}. Proof. Suppose that L2(X, µ) ∼= ℓ2(Z) and V is the unitary induced by the shift σ : n 7→ n + 1. We identify W with the sequence (Wn)n∈Z where Wn is a unitary on H. Define a sequence (ζn) of unitaries on H as follows: ζ0 = 1, ζn = ζn−1W ∗ n for n > 0, and ζn = ζn+1Wn+1 for n < 0, and set ζ = (ζn) ∈ ℓ∞(Z) ⊗ B(H). Then ζnWnζ ∗ n−1 = 1 for all n, which implies that ζW ¯α(ζ)∗ = 1, where ¯α = Ad(V ⊗ 1) is the shift on ℓ∞(Z) ⊗ B(H). Thus the fixed point algebra of ℓ∞(Z) ⊗ B(H) under Ad(W (V ⊗ 1)) is Adζ(1 ⊗ B(H)) 4 as ζW (V ⊗ 1)ζ ∗ = V ⊗ 1. Hence it follows that H ∼= C and π is unitarily equivalent to π(µ,C,1). Suppose that L2(X, µ) ∼= ℓ2(Z/kZ) with Z/kZ identified with {0, 1, . . . , k − 1} and V is the unitary induced by the shift. Let Z be a unitary in B(H) such that Z k = WkWk−1 · · · W1 where W = (Wn)n with Wk = W0. Let ζ0 = 1 and ζn = Zζn−1W ∗ n for n = 1, 2, . . . , k − 1. Then ζnWnζ ∗ k−1 = W0Wk−1 · · · W1Z −k+1 = Z for n = 0). With ζ = (ζn) the fixed point algebra of ℓ2(Z/kZ)⊗B(H) under Ad(W (V ⊗ 1)) is Adζ(1 ⊗ B(H) ∩ Z ′). Hence it follows that H ∼= C. We may assume that Z is a constant as in the statement. QED n−1 = Z (e.g., ζ0W0ζ ∗ Remark 2.6 The irreducible representations presented in the above proposition have been already explored in [8]. In particular, in the latter case, π(µ,C,ei2πθ), which is equivalent to π(µ,C,Wθ) with Wθ = (ei2πkθ, 1, . . . , 1), are mutually disjoint for 0 ≤ θ < 1/k. 3 Topological versus algebraic Let π be an irreducible representation of C ∗(Σ). We assume that π = π(µ,H,W ) as in Proposition 2.2. We will show that πℓ1(Σ) is not algebraically irreducible if L2(X, µ) is infinite-dimensional under some condition on the quasi-invariance of µ. Lemma 3.1 Let π = π(µ,H,W ) be as above. Define a bounded linear map TΦ of ℓ1(Σ) into L2(X, µ) ⊗ H for a unit vector Φ ∈ L2(X, Σ) ⊗ H by TΦ(b) = πµ(b)Φ. Let LΦ be the kernel of TΦ, which is a closed left ideal of ℓ1(Σ). If TΦ is surjective, then there is a constant KΦ > 0 such that kb + LΦk1 ≤ KΦkT (b)k. Proof. This follows from the closed graph theorem. QED Define Fk ∈ L2(X, µ) for k ∈ Z by Fk(x) = ( dµσ−k dµ (x))1/2 and note that V satisfies that (V kξ)(x) = Fk(x)ξ(σ−k(x)), ξ ∈ L2(X, µ). Lemma 3.2 Let η be a unit vector of H. Let A be a measurable subset of X with µ(A) > 0 and let S =Pk akδk ∈ ℓ1(Σ) be such that πµ(S)1 ⊗ η = µ(A)1/2 ⊗ η, χA where χA is the characteristic function of A. Then it follows that 1 ≤Pk kakkµσ−k(A)1/2 ≤ kSk1 supk µ(σk(A))1/2. 5 Proof. Note that π(δk) = (W (V ⊗ 1))k = Wk(V k ⊗ 1) where W0 = 1, Wk = W ¯α(Wk−1) for k > 0 and Wk = ¯α−1(W ∗Wk+1) for k < 0, and ¯α = Ad(V ⊗ 1); Wk are all unitaries in L∞(X, µ) ⊗ B(H). Let ξ = χA/µ(A)1/2, a unit vector. Then we compute: 1 =Xk hξ ⊗ η, (ak ⊗ 1)Wk(Fk ⊗ η)i =Xk hξ ⊗ η, (ak ⊗ 1)Wk(χAFk) ⊗ ηi which is at most Pk kakkkχAFk ⊗ ηk =Pk kakkµσ−k(A)1/2. QED Lemma 3.3 Suppose that µ is non-atomic and that there is a metric d on X such that d induces the topology on X and σ preserves d, i.e., d(x, y) = d(σ(x), σ(y)) for x, y ∈ X. Then for any ǫ > 0 there is a measurable subset A of X such that 0 < supk µ(σk(A)) < ǫ. Proof. Suppose, on the contrary, that there is an ǫ > 0 such that supk µ(σk(A)) > ǫ for any A with µ(A) > 0. Let x ∈ X be such that any open neighborhood of x has positive measure. Let Un = {y ∈ X d(x, y) < 1/n} for n ∈ N. Then there is a kn ∈ Z such that µ(σkn(Un)) > ǫ. Since X is compact there is a subsequence in (σkn(x)) converging, say to z ∈ X. Then it follows that any open neighborhood of z contains σkn(Un) for some n and hence has measure greater than ǫ, which implies µ({z}) ≥ ǫ. Hence µ is atomic, a contradiction. QED Let Λ = {supk µ(σk(A)) A measurable with µ(A) > 0}, which is a subset of (0, 1]. If µ is σ-invariant and is non-atomic, then Λ = (0, 1]. If µ is atomic, then Λ ⊂ [λ0, 1] where λ0 = supk µσk(A) > 0 with A an atom. Theorem 3.4 Let π = π(µ,H,W ) be as above and suppose that µ is non-atomic. Assume that for any ǫ > 0 there is a measurable subset A of X such that 0 < supk µσk(A) < ǫ (which follows if µ is σ-invariant or σ preserves a metric on X which induces the topology). Then πℓ1(Σ) is not algebraically irreducible. Proof. The parenthesized statement follows from the previous lemma and the remark before this theorem. Suppose that πℓ1(Σ) is algebraically irreducible. Let η be a unit vector of H. Then the map T1⊗η from ℓ1(Σ) into L2(X, µ) ⊗ H is surjective. Hence Lemma 3.1 gives a constant K > 0 satisfying: For any unit vector Ψ ∈ L2(X, µ) there is an S ∈ ℓ1(Σ) such that π(S)1 ⊗ η = Ψ and kSk1 ≤ K. Lemma 3.2 leads us to a contradiction under the hypothesis by taking Ψ = χA/µ(A)1/2 ⊗ η for A with small supk µσk(A). QED Corollary 3.5 Let Σ = (X, σ) and suppose that X is a metric space and σ preserves the metric on X and has no periodic points. Then πℓ1(Σ) is not algebraically irreducible for any irreducible representation π of C ∗(Σ). 6 Proof. Under the hypothesis all irreducible representations are infinite-dimensional. If µ is completely non-atomic then this follows from Theorem 3.4. If µ is atomic, then L2(X, µ) ∼= ℓ2(Z) and this is proved in [5]. Let us repeat the proof in the atomic case, which seems subtler, but simpler, than In this case we may work in ℓ2(Z) with σ the shift on Z. the one of Theorem 3.4. Denote by ξn the function in ℓ2(Z) defined by ξn(k) = δn,k. Suppose that there is S = Pk akδk ∈ ℓ1(Σ) such that π(S)ξ0 = P∞ k=1 k−1ξk. Since π(S)ξ0 = Pk π(ak)ξk it follows that π(ak)ξk = k−1ξk for k = 1, 2, . . ., which implies that kakk ≥ 1/k for k ≥ 1. This contradicts Pk kakk < ∞. QED 4 Ergodic extensions The observation on irreducible representations of C ∗(Σ) in Proposition 2.2 gives rise to a problem of whether given an ergodic transformation σ on (X, µ) there is a unitary W ∈ Z = L∞(X, µ) ⊗ B(H) such that γ = Ad(W (V ⊗ 1)) acts on Z ergodically for a given Hilbert space H. Here V is the unitary on L2(X, µ) defined by (V ξ)(x) = (dµσ−1/dµ)1/2(x)ξ(σ−1(x)), which implements α on L∞(X, µ) (where α(f ) = f σ−1, f ∈ L∞(X, µ) as before). We have shown that if µ is atomic and dim(H) > 1 then there is no such W . Hence we assume that µ is non-atomic and call this the problem of ergodic extensions. We shall write V in place of V ⊗ 1 from now on. Let W = R ⊕ x ∈ X (as W is a unitary). Let T = R ⊕ almost everywhere. In particular kT σ−1(x)k = kT (x)k a.e., which implies that kT (x)k = kTk almost everywhere. If H is finite-dimensional it then easily follows: Lemma 4.1 When dimH = n < ∞, Z ∩ (W V )′ is isomorphic to a ∗-subalgebra of B(H) for any W . X W (x)dµ(x) ∈ Z where we assume that W (x) is a unitary on H for all X T (x)dµ(x) ∈ Z. Then T ∈ Z γ = Z ∩ (W V )′ if T σ−1(x) = W (x)∗T (x)W (x) and only if Proof. Let T ∈ Z ∩ (W V )′. Then it follows that det(T (x) − λ1), as a polynomial of order n in λ, is almost constant. Hence T has at most n eigenvalues, which implies that Z ∩ (W V )′ is finite-dimensional. Hence one can show that there is an x ∈ X such that Z ∩ (W V )′ ∋ T 7→ T (x) is an injective homomorphism. QED The problem we cannot answer in general is whether there is a unitary W ∈ Z = L∞(X, µ)⊗B(H) for H 6= C1 such that Z ∩ (W V )′ = C1 when µ is non-atomic, let alone how to classify those W modulo unitary equivalence. In what follows we restrict ourselves to two specific examples of Σ = (X, σ); Bernoulli If Σ is a Bernoulli shift then C ∗(Σ) is not simple but shifts and irrational rotations. primitive and if Σ is an irrational rotation on X = T then C ∗(Σ) is simple (see [6, 7 8]). Among many ergodic measures on X we shall choose specific invariant probability measures on X and discuss ergodic extensions. Let Λ be a finite set of more than one elements. Let X = ΛZ = {(xn) xn ∈ Λ} and σ be the shift on X to the right and define an automorphism α on C(X) by α(f )(x) = f σ−1(x). Then the fixed point algebra C(X)α of C(X) under α is C1. This follows because there is an x ∈ X whose σ-orbit is dense in X. First we consider a C(X)-version instead of L∞(X, µ), which is considerably simpler, i.e., we assert that for any integer n > 1 there is an automorphism β of C(X) ⊗ Mn such that β(f ⊗ 1) = α(f ) ⊗ 1 for f ∈ C(X) and (C(X) ⊗ Mn)β = C1. We shall prove this assertion. Define a diagonal unitary u by u = 1⊕ω⊕ω2⊕· · ·⊕ωn−1 with ω = e2πi/n and a shift unitary v ∈ Mn such that vuv∗ = ωu. Note that Mn∩{u, v}′ = C1. Let C1 be a non-empty proper subset of Λ and let C = {x ∈ X x0 ∈ C1}, a closed and open subset of X. Let D = X \ C = {x ∈ X x0 ∈ Λ \ C1}. Define a unitary W ∈ C(X) ⊗ Mn by W = χC ⊗ u∗ + χD ⊗ v∗ and an automorphism β of C(X) ⊗ Mn by β = Ad W ◦ (α ⊗ id). Note that for f ∈ C(X) ⊗ Mn, β(f )(x) = W (x)f (σ−1(x))W (x)∗ =(Ad u∗(f (σ−1(x))) x ∈ C Ad v∗(f (σ−1(x))) x ∈ D. Hence if β(f ) = f then f σ−1(x) = Ad W (x)∗(f (x)). Lemma 4.2 Let T ∈ (C(X) ⊗ Mn)β. Then there is a T0 ∈ Mn such that T takes all the values in Γ = {Ad(upvq)(T0); p, q = 0, 1, 2, . . . , n − 1}. Proof. Since un = 1 = vn and Ad uAd v = Ad vAd u, the finite subset Γ of Mn is invariant under Ad u and Ad v for any T0 ∈ Mn. Let T ∈ Z ∩ (W V )′ and x0 ∈ X be such that the orbit {σk(x0) k ∈ Z} is dense in X. Let T0 = T (x0). Then T σ−1(x0) = Ad u(T0) or Ad v(T0) depending on x0 ∈ C or x0 ∈ D; thus T σ−1(x0) ∈ Γ. Repeating this process it follows that T σk(x0) ∈ Γ for all k. Since T is continuous on X we deduce that T (x) ∈ Γ for all x ∈ X. Let c(x, k) = #{i xi ∈ C1, 0 ≤ i < k} for x ∈ X and k ∈ N and d(x, k) = k − c(x, k), where #B denote the number of points in a set B. Then we obtain that T (σ−k(x)) = Ad(uc(x,k)vd(x,k))(T (x)) for k ∈ N. Let S be the cylinder subset of X consisting of x with x0, x1, . . . , xn2−1 specified as follows: xi ∈ C1 for 0 ≤ i < n, xn 6∈ C1, xn+1+i ∈ C1 for 0 ≤ i < n−1, x2n 6∈ C1, x2n+1+i ∈ 8 C1 for 0 ≤ i < n− 1, x3n 6∈ C1, . . . , x(n−1)n 6∈ C1, x(n−1)n+1+i ∈ C1 for 0 ≤ i < n− 1. Then for x ∈ S the set of pairs (c(x, k) + nZ, d(x, k) + nZ) with k = 0, 1, 2, . . . , n2 − 1 exhausts the whole Z/nZ × Z/nZ. This shows that T (σk(x)), k = 0, 1, 2, . . . , n2 − 1 exhausts the whole Γ. QED By the above lemma T ∈ (C(X) ⊗ Mn)β takes a finite number of values, say Ti, i = 1, 2, . . . , m, where m divides n2. If m = 1 this implies that Ad(upvq)(T0) = T0 for all p, q, i.e., T0 ∈ C1 or T ∈ 1 ⊗ C1. So we assume that m > 1. Note that Fi = T −1(Ti) is a closed and open subset of X. Then Fi is a cylinder subset. (Given z ∈ Fi there is a (open) cylinder subset U(z) such that z ∈ U(z) ⊂ Fi. Since Fi is compact one can find a finite number of U(z) whose union equals Fi.) Hence there is an N ∈ N such that all Fi's are determined by subsets of QN −1 k=−N Λ. Let Si be a subset of QN −1 k=−N Λ such that Fi = {x ∈ X (x−N , x−N +1, . . . , xN −1) ∈ Si}. Let y ∈ S1 and z ∈ S2. We shall construct an element x ∈ X containing y, z as Let m1 = #{i yi ∈ C1, 0 ≤ i ≤ N − 1} and m2 = #{i zi ∈ C1,−N ≤ i ≤ −1} and let ℓ1 = N − m1 and ℓ2 = N − m2. Let a, b be integers between 0 and n − 1 such that m1 + m2 + a = 0, ℓ1 + ℓ2 + b = 0 modulo n. We define x ∈ X as an element satisfying the following conditions: segments whose existence gives a contradiction. xi = yi, xi ∈ C1, xi 6∈ C1, xi = zi−2N −a−b, −N ≤ i ≤ N − 1, N ≤ i ≤ N + a − 1, N + a ≤ i ≤ N + a + b − 1, N + a + b ≤ i ≤ 3N + a + b − 1. Then x ∈ F1 and σ−(2N +a+b)(x) ∈ F2 (as σ−(2N +a+b)(x)i = x2N +a+b+i = zi for −N ≤ i ≤ N−1). Since c(x, 2N +a+b) = m1+a+m2 = 0 (mod n) and d(x, 2N +a+b) = ℓ1+b+ℓ2 = 0 (mod n), it follows that T (σ−(2N +a+b)x) = Ad(uc(x,2N +a+b)vd(x,2N +a+b))(T1) = T1, which contradicts that T (σ−(2N +a+b)(x)) = T2 following from σ−(2N +a+b)(x) ∈ F2. Thus one can conclude that m = 1. Proposition 4.3 Let X = ΛZ and σ the shift on X as above. If α is the automorphism of C(X) induced by σ then C(X)α = C1. If n is an integer greater than 1 and β = Ad W (α⊗ 1) is an automorphism of C(X)⊗Mn with W as above, it follows that (C(X)⊗Mn)β = C1. We will now prove the L∞-version of the above result. Let X = ΛZ and σ the shift on X as above. Let µ1 be a probability measure on Λ such that µ1({λ}) > 0 for all λ ∈ Λ and define a measure µ on X as the infinite product of copies of µ1. Then µ is a σ-invariant probability measure on X. Define a unitary V on L2(X, µ) as the unitary induced by σ as before. Then Ad V acts on L∞(X, µ) ergodically. 9 This is shown in a standard way. For any pair A, B of cylinder subsets of X we obtain that µ(A ∩ σk(B)) → µ(A)µ(B) as k → ∞. It then follows that this is true for any measurable subsets A, B. If A is a σ-invariant subset, i.e., µ(A\ σ(A)) = 0 = µ(σ(A)\ A) then it follows that µ(A) = µ(A∩ σk(A)) = µ(A)2, i.e., µ(A) = 0 or 1. Hence σ is ergodic. Next we will prove: For any n > 1 there is a unitary W in Z = L∞(X, µ) ⊗ Mn such that Ad(W V ) acts on Z ergodically. We have defined the unitaries u, v ∈ Mn and define a unitary W ∈ L∞(X, µ) ⊗ Mn as before: W = χC ⊗ u∗ + χD ⊗ v∗, where C = {x ∈ X x0 ∈ C1} and D = X \ C. Lemma 4.4 Let T ∈ L∞(X, µ) ⊗ Mn ∩ (W V )′. Then there is a T0 ∈ Mn such that T takes values in Γ = {Ad(upvq)(T0) p, q = 0, 1, 2, . . . , n − 1}. Moreover T takes all the values in Γ. Proof. As we have remarked, the subset Γ of Mn is invariant under Ad u and Ad v for any T0 ∈ Mn. Let T ∈ Z∩(W V )′. Let T0 ∈ Mn be such that kT0k = kTk and {x ∈ X kT (x)−T0k < ǫ} has positive measure for any ǫ > 0. Let Bǫ = {S ∈ Mn kS − T0k < ǫ}. Then it follows that T (x) ∈ Fǫ = Sp,q Ad(upvq)(Bǫ) for almost all x. (If T (x) ∈ Fǫ then T σ−1(x) ∈ Fǫ since T σ−1(x) = Ad u(T (x)) or Ad v(T (x)). Repeating this it follows that T σk(x) ∈ Fǫ for all k.) By taking the intersection of Fǫ with ǫ > 0 we conclude that {x ∈ X T (x) ∈ Γ} has full measure. Let Xp,q = {x ∈ X T (x) = Ad(upvq)(T0)} for p, q = 0, 1, . . . , n − 1. Let S be the cylinder subset defined in the proof of Lemma 4.2. There is a pair p, q such that S ∩ Xp,q is not a null set. Then it follows from the property of S that σ−k(S ∩ Xp,q), k = 0, 1, 2, . . . , n2 − 1 visit all Xi,j. (For example σ−k(S ∩ Xp,q) ⊂ Xp+k,q for k = 0, 1, . . . , n − 1, σ−n(S ∩ Xp,q) ⊂ Xp+n−1,q+1, etc.) This proves the last statement. QED If Γ is a singleton, then T0 ∈ C1 because u, v generate the whole Mn. Suppose that Γ includes at least two points; say T0 and T1 = Ad(up′vq′)(T0) 6= T0 for some p′, q′. Let Fi = {x ∈ X T (x) = Ti} for i = 0, 1 and let η = min{µ(Fi) i = 0, 1} > 0. Let Ki be a compact subset such that Ki ⊂ Fi and µ(Ki) > η/2. Let ǫ > 0 be very small and choose a cylinder subset Oi such that Ki ⊂ Oi and µ(Oi \ Ki) < ǫ. (We will specify ǫ > 0 later.) We choose N ∈ N such that Oi can be regarded as a subset of QN −1 −N Λ for i = 0, 1. We assume that N is a multiple of n. 10 Let N −1 Gp ={x ∈ X Hp ={x ∈ X χC1(xi) = p mod n}, χC1(xi) = p mod n}, −1 Xi=0 Xi=−N Xi=N N +n−1 Lp ={x ∈ X χC1(xi) = p mod n} for p = 0, 1, . . . , n− 1. Regarding O0 ∩ Gp as a subset ofQN −1 of QN +n−1 the right), we construct a cylinder set F (corresponding to a subset of Q3N +n−1 Λ and O1 ∩ Hp as a subset of Q3N +n−1 i=−N Λ and Ln−p−q as a subset i=N +n Λ (after being translated 2N + n to i=−N Λ) by concatenating triple finite sequences in i=N O0 ∩ Gp × Ln−p−q × O1 ∩ Hq. [p,q Then F ⊂ O0 and σ−2N −n(F ) ⊂ O1 (because O1∩ Hq to be concatenated has been shifted to the right by 2N + n). Since Sn−1 µ(O0 ∩ Gp)Xq p=0 Gp = X etc., we estimate µ(Ln−p−q)µ(O1 ∩ Hq) ≥ η0µ(O0)µ(O1) ≥ η0η2/4 µ(F ) =Xp where η0 = minp µ(Lp) > 0. We assume that ǫ < η0η2/8 (as η0 depends only on µ and n); then it follows that µ(F ∩ K0 ∩ σ2N +n(K1)) is positive because it is bounded below by µ(F ) − µ(F \ K0) − µ(F \ σ2N +n(K1)) ≥ µ(F ) − 2ǫ > 0. i=0 Let x ∈ F ∩ K0∩ σ2N +n(K1). Then T (σ−2N −n(x)) = T (x) becauseP2N +n−1 χC1(xi) = 0 mod n (by the construction of F ) and 2N +n = 0 mod n. This is a contradiction because T (σ−2N −n(x)) = T1 6= T0 = T (x) follows from σ−2N −n(x) ∈ K1 ⊂ F1 and x ∈ K0 ⊂ F0. Thus Γ must be a singleton. Proposition 4.5 Let X = ΛZ and σ the shift on X and µ = QZ µ1 a probability mea- sure on X as above. Then the automorphism α on L∞(X, µ) induced by σ satisfies L∞(X, µ)α = C1. If n is an integer greater than 1 and β = Ad W (α ⊗ 1) is an automor- phism of L∞(X, µ) ⊗ Mn with W as above then it follows that (L∞(X, µ) ⊗ Mn)β = C1. We have defined the subset C by specifying C1 ⊂ Λ: C = {x ∈ X x0 ∈ C1} with C1 6= ∅, Λ. Let C ′ 1} and D′ = X \ C ′. Let W ′ ∈ L∞(X, µ)⊗ Mn be the corresponding unitary χC ′ ⊗ u∗ + χD′ ⊗ v∗. We consider the problem of when W V and W ′V are unitarily equivalent. Suppose that there is a unitary ζ ∈ L∞(X, µ) ⊗ Mn such that W V = ζW ′V ζ ∗, i.e., W (x) = ζ(x)W ′(x)ζ(σ−1(x))∗ for almost all x or ζσ−1(x) = W (x)∗ζ(x)W ′(x). Then depending on x0 ∈ Λ we have the following cases: 1 be anther subset of Λ and define C ′ = {x ∈ X x0 ∈ C ′ 11 (1) If x ∈ C ∩ C ′, ζ(σ−1(x)) = uζ(x)u∗, (2) If x ∈ D ∩ D′, ζ(σ−1(x)) = vζ(x)v∗, (3) If x ∈ C ∩ D′, ζ(σ−1(x)) = uζ(x)v∗, (4) If x ∈ D ∩ C ′, ζ(σ−1(x)) = vζ(x)u∗. Hence if ζ(x) is defined ζ(σ−1(x)) is obtained by applying one of the four maps on ζ(x): φ1 = LuRu∗, φ2 = LvRv∗, φ3 = LuRv∗, φ4 = LvRu∗ depending on x0, where Lb denotes the left multiplication of b ∈ Mn etc. They satisfy φn i = id for i = 1, 2, 3, 4, φ1φ2 = φ2φ1, and φ1φ3 = ωφ3φ1, φ2φ3 = ωφ3φ2, φ3φ4 = ω−2φ4φ3, φ1φ4 = ω−1φ4φ1, φ2φ4 = ω−1φ4φ2, by the commutation relation vu = ωuv. Note also that φk φk 4(ζ) = Ad vk(ζ)vku−k. 3(ζ) = Ad uk(ζ)ukv−k and Lemma 4.6 In the above situation there is a unitary ζ0 ∈ Mn such that ζ takes values in Γ = {Ad(upvq)(ζ0)ωk(uv∗)ℓ p, q, k, ℓ = 0, 1, 2 . . . , n − 1} almost everywhere and {x ∈ X ζ(x) = ζ0} is not a null set. Proof. Since Γ is invariant under φi, i = 1, 2, 3, 4 this can be proved in the same way as Lemma 4.4. QED Suppose that ζ takes at least two values, ζ0 above and ζ1 ∈ Γ. Let Fi = {x ∈ X ζ(x) = ζi} for i = 0, 1 and η = min{µ(F0), µ(F1)} > 0. Let Ki ⊂ Fi be a compact subset such that µ(Ki) > η/2. For any ǫ > 0 there are cylinder subsets Oi such that Ki ⊂ Oi and µ(Oi \ Ki) < ǫ. There is an N ∈ N such that each Oi is determined by a subset of QN −1 We have assumed C 6= C ′: There are six cases depending on which C ∩ C ′, D∩ D′, C ∩ D′,D∩ C ′ are empty. (i) C ∩ C ′ = D ∩ D′ = ∅ (ii) only C ∩ C ′ = ∅, (iii) only C ∩ D′ = ∅, (iv) only D ∩ C ′ = ∅, (v) only D ∩ D′ = ∅, (vi) none of the above intersections are empty. We will consider each case separately. Suppose (i), i.e., C = D′, D = C ′. In this case only φ3 and φ4 appear when we express i=−N Λ. We may assume that N is a multiple of n. ζσ−1(x) in terms of ζ(x). We furthermore assume that n is even. Let x ∈ O0. Then ζσ−N (x) can be uniquely expressed as ω2kφℓ 4 (ζ(x)), depending on whether each of x0, x1, . . . , xN −1 falls into C or D, with 0 ≤ k ≤ n/2 − 1, 1 ≤ ℓ ≤ n, and n ≤ m ≤ 2n − 1. Then, since n + 1 ≤ ℓ + m ≤ 3n − 1 and ℓ + m = 0 (mod n), it follows that ℓ + m = 2n. Hence we can also express this as ζσ−N (x) = φℓ−1 (ζ(x)). We denote by O0(k, ℓ, m) the 3φm 4 3 φk 4φ3φm−k 12 Let y ∈ O1. Then in the same way ζ(y) = φℓ′−1 set of x ∈ O0 with x0, x1, . . . , xN −1 giving this expression on ζσ−N (x), ζ(x). The union of O0(k, ℓ, m) with all possible k, ℓ, m equals O0. (ζσN (y)) depending on y−N , y−N +1, . . . , y−1, for k′, ℓ′, m′ with 0 ≤ k′ ≤ n/2− 1, 1 ≤ ℓ′ ≤ n, and n ≤ m′ ≤ 2n + 1. We denote by O1(k′, ℓ′, m′) the set of y ∈ O1 with y−N , y−N +1, . . . , y−1 giving rise to this relation on ζ(y), ζσN(y). Let L(k, ℓ, m) = C m−1 × Dk × C × Dℓ−k ⊂ Λ2n. We define F to be the cylinder set of 4 φ3φm′−k′ 3 φk′ 4 X corresponding to [k,ℓ,m,k′,ℓ′,m′ as a subset of Q3N +4n−1 i=−N O0(k, l, m) × L(k, ℓ, m) × L(k′, ℓ′, m′) × O2(k′, ℓ′, m′) Λ. Then we deduce for x ∈ F that ζ(σ−2N −4n(x)) = ζ(x) since ζ(σ−N −2n(x)) = φℓ−k 4 φ3φk 4φm−1 3 φℓ−1 3 φk 4φ3φm−k 4 (ζ(x)) = ζ(x) (because φm−1 3 φℓ−1 3 = φ−2 3 , φ3φk 4 = ω−2kφk 4φ3, φk 4φ3 = ω2kφ3φk 4 etc.) and ζ(σ−2N −4n(x)) = φℓ′−1 3 φk′ 4 φ3φm′−k′ 4 φℓ′−k′ 4 φ3φk′ 4 φm′−1 3 (ζ(σ−N −2n(x))) = ζ(σ−N −2n(x)). Note that µ(F ) ≥ µ(O0)µ(O1)η2 0 > η0η2/4 where η0 = min{µ(L(k, ℓ, m))} > 0 is indepen- dent of the choice of Ki etc. If ǫ < (η0η)2/8 then K0 ∩ σ2N +4n(K1) ∩ F is not a null set, which contradicts ζ(σ−2N −4n(x)) = ζ(x) on F as shown as before. Hence ζ(x) takes just one value ζ0 almost everywhere. Thus we conclude that ζ0 = uζ0v∗ and ζ0 = vζ0u∗ or u = ζ0vζ ∗ 0 , v = ζ0uζ ∗ 0 . So the pair (u, v) maps to (v, u) under Ad ζ0; this happens when and only when n = 2. In the case n = 2 we may take ζ0 = 1 1 √2(cid:18)1 1 −1(cid:19) . 4 3φm 3 φk 4φ3φm−k 4 (ζ(x)) = φℓ−1 We now assume that n is odd. Given x ∈ O0 the value ζσ−N (x) can be uniquely expressed as ω2kφℓ (ζ(x)) with 0 ≤ k ≤ n − 1, 1 ≤ ℓ ≤ n, and n ≤ m ≤ 2n − 1. (Then ℓ + m = 2n.) We denote by O0(k, ℓ, m) the set of x ∈ O0 with x0, x1, . . . , xN −1 giving this expression on ζσ−N (x), ζ(x). Given y ∈ O1 we have the unique expression ζ(y) = ω2k′φℓ′ (ζσN (y)) with 0 ≤ k′ ≤ n − 1, 1 ≤ ℓ′ ≤ n − 1 and n ≤ m′ ≤ 2n − 1 depending on y−N , y−N +1, . . . , y−1. We define O1(k′, ℓ′, m′) as above and then proceed as before. Since n ≥ 3 there is no solution for ζ in this case. Suppose (ii), i.e., C = C ∩ D′, D ∩ D′, C ′ = D ∩ C ′ are non-empty. In this case only φ2, φ3, φ4 appear when we express ζσ−1(x) in terms of ζ(x). Let x ∈ O0. Then, depending on x0, x1, . . . , xN −1, we obtain a unique expression 4 (ζσN (y)) = φℓ′−1 4 φ3φm′−k′ 3 φk′ 3 φm′ 4 ζσ−N (x) = ωjφk 2φℓ 3φm 4 (ζ(x)) = φk−j 2 φ3φj 2φℓ−1 3 φm 4 (ζ(x)) 13 with 0 ≤ j ≤ n − 1, n − 1 ≤ k ≤ 2n − 2, and 1 ≤ ℓ, m ≤ n. Then k + ℓ + m = 2n or k + ℓ + m = 3n (because n + 1 ≤ k + ℓ + m ≤ 4n − 2 and k + ℓ + m = 0 modulo n). In the same way as above we define O0(j, k, ℓ, m) as the subset of O0 consisting of x with x0, x1, . . . , xN −1 giving this relation on ζσ−N (x), ζ(x). Similarly O1(j, k, ℓ, m) is defined as the subset of O1 consisting of y with y−N , y−N +1, . . . , y−1 giving the relation ζ(y) = ωjφk 2φℓ 3φm 4 (ζσN (y)) = φk−j 2 φ3φj 2φℓ−1 3 φm 4 (ζσN (y)). When k + ℓ + m = 2n let L(j, k, ℓ, m) = (D ∩ D′)2n−k+j × (C ∩ D′)n−1 × (D ∩ D′)n−j × (C ∩ D′)n−ℓ+1 × (D ∩ C ′)n−m as a subset of Λ4n, and when k + ℓ + m = 3n let L(j, k, ℓ, m) = (D ∩ D′)3n−k+j × (C ∩ D′)n−1× (D∩ D′)n−j × (C ∩ D′)n−ℓ+1× (D∩ C ′)n−m as a subset of Λ4n where the exponent of D ∩ D′ is increased by n. Let F denote the cylinder subset of X determined by [ O0(j, k, ℓ, m) × L(j, k, ℓ, m) × L(j′, k′, ℓ′, m′) × O1(j′, k′, ℓ′, m′) as a subset of Q4N +8n null set for a sufficiently small ǫ > 0 and ζ(σ−2N −8n(x)) = ζ(x), x ∈ F (e.g., i=−N Λ. Then one shows as before that K0 ∩ σ2N +8n(K1) ∩ F is not a ζσ−N −4n(x) = φn−m 4 φn−ℓ+1 3 φn−j 2 φn−1 3 φ2n−k+j 2 φk−j 2 φ3φj 2φℓ−1 3 φm 4 (ζ(x)) = ζ(x) where the first five φ's are derived from L(j, k, ℓ, m)). Since this is a contradiction we conclude that ζ(x) = ζ0 almost everywhere. This implies that ζ0 = vζ0v∗, ζ0 = uζ0v∗, ζ0 = vζ0u∗, entailing uv∗ = 1, a contradiction. Thus there is no such ζ. Suppose (iii), i.e., ζσ−1(x) is one of φ1(ζ(x)), φ2(ζ(x)), φ4(ζ(x)). If x ∈ O0 we deduce that ζσ−N (x) is uniquely expressed as ωjφk 4 (ζ(x)) with 0 ≤ j ≤ n−1, 1 ≤ k ≤ n, n− 1 ≤ ℓ ≤ 2n−2, and 1 ≤ m ≤ n, depending on x0, x1, . . . , xN −1. Then k+ℓ+m = 2n or 3n. Note 2φm 1φℓ 1φℓ 1φℓ−j 2 φ4φj 2φm ζσ−N (x) = ωjφk 4 (ζ(x)) = φk Define O0(j, k, ℓ, m) and O1(j, k, ℓ, m) as before and L(j, k, ℓ, m) = (C ∩ C ′)n−k × (D ∩ D′)2n−ℓ+j × (D ∩ C ′)n−1 × (D ∩ D′)n−j × (D ∩ C ′)n−m+1 as a subset of Λ4n when k + ℓ + m = 2 and L(j, k, ℓ, m) by the same product as above with the first factor replaced by (C ∩ C ′)2n−k when k + ℓ + m = 3n. We can then proceed as before. 2φm−1 (ζ(x)). 4 We can treat the cases (iv), (v), and (vi) similarly; so we omit the details. Proposition 4.7 Let X = Qk∈Z Λ, σ, µ, and u, v ∈ Mn with n > 1 be as above. Let 1 be non-empty proper subsets of Λ and let C, C ′ be the corresponding cylinder subsets C1, C ′ of X determined at 0 ∈ Z. Define W = χC ⊗ u∗ + χD ⊗ v∗ and W ′ = χC ′ ⊗ u∗ + χD′ ⊗ v∗ in L∞(X, µ) ⊗ Mn with D = X \ C and D′ = X \ C ′ for C 6= C ′. Then W V is unitarily equivalent to W ′V if and only if n = 2 and C = D′ and D = C ′. 14 For λ = (λ1, λ2) ∈ T2 we define W (λ) = χC ⊗ λ1u∗ + χD ⊗ λ2v∗. Since Ad(W (λ)V ) = Ad(W V ) with W = W (1, 1) as above, we may ask when W (λ)V and W (λ′)V are unitarily equivalent for λ, λ′ ∈ T2. Then we deduce Suppose that there is a unitary ζ ∈ L∞(X, µ) ⊗ Mn such that W (λ)V = ζW (λ′)V ζ ∗. W (λ)(x) = ζ(x)W (λ′)(x)ζ(σ−1(x))∗. 2 χD(x). Setting f (x) = det(ζ(x)) and 2 )nχD(x), as measurable functions on X of modulus one, 1λ−1 2λ−1 1 χC(x) + λn 1 )nχC(x) + (λ′ Note that (−1)n−1 det(W (λ)(x)) = λn h(x) = (λ′ we obtain f (σ−1(x)) = h(x)f (x). Hence for any N ∈ N Yi=0 h(σi(x)) · f (σN (x)), Yi=1 f (x) = f (x) = N −1 N h(σ−i(x)) · f (σ−N (x)). Since h(σk(x)) depends only on x−k the first equality implies that f is a function measur- i=−∞ Λ and i=0 Λ. (For example i=−N Λ in the i=1 hσi · gσN i=−∞ Λ.) Since f is both measurable with respect to i=0 Λ we conclude that f (x) is a constant, which implies that h(x) = 1, able with respect to (the Borel sets generated by cylinder sets coming from)Q−1 the second implies that f is a function measurable with respect to Q∞ if we approximate f by a cylinder function g measurable with respect to QN −1 i=1 hσi · gσNk1 < ǫ and QN sense that kf − gk1 < ǫ then it follows that kf −QN is measurable with respect to Q−1 i=−∞ Λ and Q∞ Q−1 Proposition 4.8 In the situation of Proposition 4.7 define W (λ) = χC⊗λ1u∗+χD⊗λ2v∗ for λ ∈ T2 and let λ, λ′ ∈ T2. Then W (λ)V is unitarily equivalent to W (λ′)V if and only if λn 1 )n = 1 = (λ′ i.e., (λ′ 2 )n. 1λ−1 2λ−1 1)n and λn 2 = (λ′ 1 = (λ′ 2)n. 1 = (λ′ Proof. If λn taking uℓv−k for ζ it follows that W (λ′ Thus the 'if' part is obvious. The 'only if' part is shown before this proposition. QED 2)V are unitarily equivalent to W (ωkλ′ 2 for some k, ℓ ∈ Z. By 2)V . 2)n then λ1 = ωkλ′ 1 and λ2 = ωℓλ′ 1)n and λn 2 = (λ′ 1, ωℓλ′ 1, λ′ The other example is based on a dynamical system on X = T = R/Z. Let θ ∈ (0, 1/2) be an irrational number and denote by σ the translation by θ: x 7→ x + θ on T. If α denotes the automorphism of C(T) defined by α(f )(x) = f σ−1(x), then the C∗-algebra crossed product of C(T) by α is a so-called irrational rotation algebra. Let µ be the Lebesgue measure on T, which is an ergodic σ-invariant probability measure and the only σ-invariant probability measure. Note that µ is also invariant under the action of T by translations, which is the fact we will use later. There are many singular continuous probability measures on T which are ergodic σ-quasi-invariant; we will construct such measures in the next section but we do not know if there are ergodic extensions for such measures. In this case we do not have any results for a C∗-version of ergodic extensions since our choice of W , similar to the one in the previous case, is not continuous on T. 15 Let n be an integer greater than 1 and let u, v be unitaries in Mn as above. Let C ⊂ T be a measurable subset of T such that 0 < µ(C) < 1 and let D = T \ C. Define a unitary W ∈ Mn by W = χC ⊗ u∗ + χD ⊗ v∗ and define an automorphism β = Ad(W V ) on L∞(T) ⊗ Mn (where V is the unitary on L2(T) implementing α). Lemma 4.9 If T ∈ (L∞(T) ⊗ Mn)β then there is a T0 ∈ Mn such that T takes values in ∆ = {Ad(upv−p)(T0) 0 ≤ p < n} for almost all x ∈ T. Proof. It follows from the proof of Lemma 4.4 that T takes values in Γ = {Ad(upvq)(T0) 0 ≤ p, q < n} for some T0 ∈ Mn. We may suppose that {x ∈ T T (x) = T0} has positive measure and let A = {x ∈ T T (x) ∈ ∆}. Then it follows that T σ−n(x) ∈ ∆ for almost all x ∈ A (by repeating T (σ−1(x)) = Ad u(T (x)) or Ad v(T (x)) n times). Since A is σn-invariant and σn is ergodic it follows that A has full measure. QED Let Ti, i = 0, 1, . . . , m be all the distinct elements in the (essential) range of T . If m = 0 then T0 ∈ C1 (as T0 = T (σ−1(x)) = Ad u(T0) for x ∈ C and = Ad v(T0) for x ∈ D) and this is what we wanted to prove. Let Fi = T −1(Ti), which is a non-null measurable subset such that Si Fi has full measure. Let (kr/mr) be the sequence of rational numbers obtained from the continued fraction of θ, which is given as (cid:18) kr mr(cid:19) =(cid:18) kr−1 for r ≥ 0 where (br)r≥1 is some sequence of natural numbers and b0 = 0, k−2 = 0, k−1 = 1, m−2 = 1, m−1 = 0. Note that for r ≥ 1 mr−1 mr−2(cid:19)(cid:18)br 1(cid:19) kr−2 θ − kr/mr < 1/m2 r, krmr−1 − kr−1mr = (−1)r+1. Since mrθ + Z converges to 0 ∈ T, the sum Pm−1 r → ∞. Let (r(j)) be a subsequence such that i=0 µ(Fi ∩ (Fi + mrθ)) converges to 1 as µ(Fi ∩ (Fi + mr(j)θ) ∩ (Fi + mr(j)+1θ))(cid:1) < ∞ Xj (cid:0)1 −Xi ℓ=1T∞ and let X0 =S∞ j=ℓ Gj with Gj =Si(Fi ∩ (Fi + mr(j)θ) ∩ (Fi + mr(j)+1θ)), which has full measure. Since if x ∈ Gj then x, x − mr(j)θ, x − mr(j)+1θ ∈ Fi for some i we deduce that T (σ−mr(j)(x)) = T (σ−mr(j)+1(x)) = T (x). If x ∈ X0 the equalities hold for all large j. T (σ−mr (x)) = Ad(uc(x,r)vd(x,r))(T (x)) for almost all x. Let c(x, r) = #{i x − iθ ∈ C, 0 ≤ i < mr} and d(x, r) = mr − c(x, r). Then Now we assume that C is an interval [0, θ) of T. Then it follows that µ({x ∈ T c(x, r) = kr}) = 1 − mrθ − kr (since applying σ to x ∈ T mr times results in 16 rotating x around the circle T almost kr times; see the lemma below for details). Hence it follows that T (σ−mr(x)) = Ad(ukrvmr−kr)(T (x)) except for x in a subset of measure mrθ − kr. Suppose that Ad(ukrvmr−kr )(T (x)) = Ad(ukr+1vmr+1−kr+1)(T (x)) = T (x), which holds for all large r = r(j) for almost all x ∈ X0. Since kr mr+1 mr(cid:19) (cid:18) kr+1 has determinant 1 or -1 it follows that there is an inverse matrix consisting of integers, say c d(cid:19) . (cid:18)a b (Actually a = (−1)r+1mr, b = −(−1)r+1kr, c = −(−1)r+1mr+1, d = (−1)r+1kr+1.) Since (ukrvmr−kr)c(ukr+1vmr+1−kr+1)a is proportional to uv∗ (as the exponent of u is krc+kr+1a = 1 and the exponent of v is mrc−krc+mr+1a−kr+1a = −1), it follows that Ad(uv∗)(T (x)) = T (x). Since T (x) = Ad(upv−p)(T0) for some p, this implies that Ad(uv∗)(T0) = T0, i.e., m = 0 and T (x) = T0 for almost all x. Since T σ−1(x) = Ad u(T (x)) or Ad v(T (x)) depending on x, this shows that T0 ∈ C1. Thus we conclude that (L∞(T) ⊗ Mn)β = C1. Lemma 4.10 Let m ∈ N and let [mθ] be the largest integer satisfying [mθ] ≤ mθ. Then µ({x ∈ T c′(x, m) = [mθ]}) = [mθ] + 1 − mθ, µ({x ∈ T c′(x, m) = [mθ] + 1}) = mθ − [mθ], where c′(x, m) = #{i x − iθ ∈ C, 0 ≤ i < m}. If mθ ≈ k then µ({x ∈ T c′(x, m) = k}) = 1 − mθ − k. Proof. We have assumed that C = [0, θ). Since C is an interval of length θ < 1/2, if x−iθ ∈ C then x−(i−1)θ 6∈ C and x−(i+1)θ 6∈ C and if two consecutive points x−iθ, x−(i+1)θ in orbit passes the middle point θ/2 of C then one and only one of them falls into C. If m′ is the smallest positive integer with [m′θ] = [mθ] then c′(x, m′) ≥ [mθ] (as at least [mθ] of x, x− θ, . . . , x− (m′ − 1)θ belong to C) and c′(x, m′) = [mθ] + 1 occurs only when (m′ − 1)θ, x ∈ C. If c′(x, m′) = [mθ] and one of x − m′θ, x − (m′ + 1)θ . . . , x − (m − 1)θ belongs to C then c′(x, m) = [mθ] + 1. If [(m − 1)θ] = [mθ] then m′ ≤ m − 1. Note that x lies on the arc of length θ between (m′− 1)θ and m′θ. Then it follows that if one of x, x− m′θ, x− (m′ + 1)θ, . . . , x− (m− 1)θ belongs to C then c′(x, m) = [mθ] + 1; otherwise c′(x, m) = [mθ]. The set of x with 17 c(x, m) = [mθ] + 1 is given by the condition 0 ≤ x < θ, m′θ − [mθ] ≤ x < (m′ + 1)θ − [mθ], . . . , or (m − 1)θ − [mθ] ≤ x < mθ − [mθ], i.e., 0 ≤ x < mθ − [mθ]. If [(m − 1)θ] < [mθ] (i.e., [(m − 1)θ] < (m − 1)θ < [mθ] < mθ) then m′ = m and only when both x, x − (m − 1)θ fall into C we have that c′(x, m) = [mθ] + 1. Since 0 ≤ x < mθ − [mθ] if and only if 0 < [mθ] − (m − 1)θ ≤ x + [mθ] − (m − 1)θ < θ, the set of x with c(x, m) = [mθ] + 1 is 0 ≤ x < mθ − [mθ]. If k > mθ > [mθ] then k = [mθ] + 1; so mθ − [mθ] = 1 − (k − mθ) and if mθ > k then k = [mθ]; [mθ] + 1 − mθ = 1 − (mθ − k). Thus we derive the last statement. QED We have shown the following: Proposition 4.11 Let σ denote the map x 7→ x + θ on T = R/Z where θ ∈ (0, 1/2) is irrational and let µ be the Lesbegue measure on T. Let V denote the the unitary on L2(T) induced by σ. Let n be an integer greater than 1 and let W = χC ⊗ u∗ + χD ⊗ v∗ where C is an interval [0, θ) ⊂ T and D = T \ C where u, v are the canonical pair of unitaries in Mn defined before. Then the automorphism Ad(W V ) acts on L∞(T) ⊗ Mn ergodically. In the case n = 1 we may choose a scaler for W and ask when λV is unitarily equivalent to λ′V for λ, λ′ ∈ T = {z ∈ C z = 1}. That is, when is there a unitary ζ ∈ L∞(T) such that λV = ζλ′V ζ ∗, or λλ′ = ζ(x)ζ(x − θ), x ∈ T? Let λλ′ = e2πiη with η ∈ R. Since e2πiqη = ζ(x)ζ(x − qθ) for q ∈ Z and ζ(· )ζ(· − qθ) converges to 1 as qθ converges to 0 in T, we deduce the hypothesis of the following lemma. Lemma 4.12 Let θ, η ∈ R. Suppose that θ ∈ (0, 1/2) is an irrational number and that if qkθ converges to 0 in T = R/Z then qkη converges to 0 in T for any sequence (qk) in Z. Then η = mθ for some m ∈ Z. Proof. Define a map φ of θZ/Z ⊂ T = R/Z into T by qθ + Z 7→ qη + Z, q ∈ Z. This is well-defined because θ is irrational. If (qk) is a sequence in Z and (qkθ + Z) is Cauchy then (qkη + Z) is Cauchy too. (If (qkη + Z) is not Cauchy then there are subsequences k(ℓ), k′(ℓ) such that (qk(ℓ) − qk′(ℓ))η + Z does not converges to zero, which contradicts that (qk(ℓ) − qk′(ℓ))θ + Z converges to zero.) Hence φ extends to a continuous map of T into T. Since φ(T) is a connected compact subset of T, η is either 0 or an irrational. If µ is irrational then φ is onto. If x = qθ and y = q′θ with q, q′ ∈ Z, then it follows that φ(x + y) = qη + q′η, mod Z. Hence we deduce that φ(x + y) = φ(x) + φ(y) in T for all x, y ∈ T and that φ(rx) = rφ(x), x ∈ T for all rational r. Since φ is continuous we have φ(tx) = tφ(x) for all t ∈ R. Since φ(0) = 0 there is a neighborhood U of 0 ∈ T (identified with (−1/2, 1/2]) such that U ⊂ (−1/2, 1/2) and φ(U) ⊂ (−1/2, 1/2). Let x ∈ U \ {0}. Then φ(tx) = tφ(x) = mtx with m = φ(x)/x, or φ(t) = mt. Since φ is a map from T onto T, it follows that m is an integer. QED Proposition 4.13 In the situation of Proposition 4.11 suppose that n = 1 (and W = 1). The following conditions are equivalent for λ, λ′ ∈ T. 18 (1) λV and λ′V are unitarily equivalent. (2) λ = e2πimθλ′ for some m ∈ Z. Proof. Define Y ∈ C(T) ⊂ L∞(T) by Y (x) = e2πix. Then Y V Y ∗ = e2πiθV . Hence (2) ⇒(1). The other implication follows from Lemma 4.12 and its preceding remark. QED The above proposition for n = 1 is perhaps known. We present a version for n > 1 in the situation of Proposition 4.11. Let λ = (λ1, λ2) ∈ T2 and let W (λ) = χC ⊗ λ1u∗ + χD ⊗ λ2v∗. Then it follows that Ad(W (λ)V ) = Ad(W (λ′)V ) on L∞(T) ⊗ Mn for all λ′ ∈ T2 (as W (λ) = (χC ⊗ λ11 + χD ⊗ λ21)W and χC ⊗ λ11 + χD ⊗ λ21 is in L∞(T) ⊗ C1); so we could ask when W (λ)V and W (λ′)V are unitarily equivalent, i.e., there is a unitary ζ ∈ L∞(T) ⊗ Mn such that W (λ)V = ζW (λ′)V ζ ∗. By taking ζ = 1 ⊗ ukvℓ it follows that W (λ)V is unitarily equivalent to W (ωℓλ1, ωkλ2) for all k, ℓ with ω = e2πi/n. Similarly by taking ζ = Ua ⊗ 1 with Ua(x) = e2πiax, x ∈ T for some a ∈ R it follows that W (λ)V is unitarily equivalent to W (e2πia(1−θ)λ1, e2πiaθλ2)V . Proposition 4.14 In the situation of Proposition 4.11 the following conditions are equiv- alent for λ, λ′ ∈ T2. (1) W (λ)V and W (λ′)V are unitarily equivalent, i.e., there is a unitary ζ ∈ L∞(T)⊗Mn such that W (λ) = ζW (λ′) ¯α(ζ)∗ where ¯α = Ad V . (2) λn 1 = e2πia(θ−1)(λ′ 1)n and λn 2 = e2πiaθ(λ′ 2)n for some a ∈ R. Proof. Suppose (2). Then λ1 = e2πia(θ−1)/nωkλ′ Since W (λ′ a/n by a too, that λ1 = e2πia(θ−1)λ′ and W (λ′)V are unitarily equivalent just before this proposition. 2)V is unitarily equivalent to W (ωkλ′ 1 and λ2 = e2πiaθλ′ 2 for some k, ℓ ∈ Z. 1, ωℓλ′ 2)V we may suppose, replacing 2. We have shown in this case W (λ)V 1 and λ2 = e2πiaθ/nωℓλ′ 1, λ′ Suppose (1). Then it follows that W (λ)(x) = ζ(x)W (λ′)(x)ζ(x − θ)∗ for almost all 2 (−1)n−1χD (as det(u) = det(v) = 1 (−1)n−1χC(x) + λn x ∈ T. Since det(W (λ)(x)) = λn (−1)n−1) we deduce that λn 1 = (λ′ λn 2 = (λ′ 1)n det(ζ(x))det(ζ(x − θ)), x ∈ C, 2)n det(ζ(x))det(ζ(x − θ)), x ∈ D. Define ηi ∈ (−π, π] by e2πiηi = (λiλ′ i)n for i = 1, 2. Setting f (x) = det(ζ(x)) and h(x) = e2πiη1χC(x) + e2πiη2χD(x) as measurable functions of modulus one on T, this amounts to f (x)f (x − θ) = h(x). 19 If qnθ − pn converges to zero with qn, pn ∈ Z then f (x)f (x − qnθ) → 1 in L1. Since f (x)f (x − qnθ) = qn−1 Yk=0 h(x − kθ), whose right-hand side is e2πi(pnη1+(qn−pn)η2) outside a subset of small measure, one con- cludes that pnη1 + (qn − pn)η2 converges to zero in T. As in the proof of Lemma 4.12 there is a continuous map φ of R into T such that φ(qθ − p) = pη1 + (q − p)η2 = η2q + (η1 − η2)p. Since φ satisfies that φ(x + y) = φ(x) + φ(y), x, y ∈ R, we conclude that φ(x) = ax (mod Z) where a is a constant. If a = 0 then ηi = 0. Suppose that a 6= 0. Since a(qθ − p) = η2q + (η1 − η2)p (mod Z) we obtain that aθ = η2 and a = η2 − η1 or η1 = −a(1 − θ) and η2 = aθ (mod Z). This concludes the proof. QED If we restrict ourselves to the case λ1 = λ2 in the above proposition, then it follows that a must be an integer, i.e., λW V is unitarily equivalent to λ′W V if and only if λn = e2πimθ(λ′)n for some m ∈ Z (cf. Proposition 4.13). We have to leave many problems unanswered. For example we did not explore all possible ergodic extensions in Propositions 4.5 and 4.11 in the case of Mn = B(Cn), let alone the case of B(H) with H an infinite-dimensional Hilbert space. We did not attempt to solve the problem for general (X, σ). 5 Quasi-invariant measures Let θ ∈ (0, 1) be an irrational number and let σ denote the homeomorphism on T = R/Z defined by x 7→ x + θ (mod 1). In this case we have noted that there are at least two kinds of ergodic σ-quasi-invariant probability measures on T: the Lebesgue measure on T (which is σ-invariant) and an atomic measure on each orbit {x + mθ m ∈ Z}. We shall construct ergodic σ-quasi-invariant singular continuous probability measures on T. In the following we denote by (x) the representative in (−1/2, 1/2] of x + Z ∈ T for x ∈ R. Let P = Q∞ i=1{0, 1}, the infinite direct product of copies of {0, 1} with the product topology. We define a continuous map Φ : P → T as follows: Let (mi) be an increasing sequence in N such that (m1θ) < 1/3 and (miθ) < (mi−1θ)/3 for i > 1. With such a sequence (mi) let ∞ Then Φ is well-defined and continuous. Xi=1 Φ(x) = xi(miθ), x = (xi) ∈ P. Lemma 5.1 Let λi ∈ {−1, 0, 1} for i = 1, 2, . . .. If t = Pi λi(miθ) then t < 1/2 and (λi) is uniquely determined by t. 20 i(miθ) for another sequence (λ′ Proof. If t =Pi λi(miθ) then t ≤Pi (miθ) < 1/2. If t =Pi λ′ i) in {−1, 0, 1} different from (λi) then there i for i < N and λN 6= λ′ is N ∈ N such that λi = λ′ i)(miθ) = 0 it follows that (mN θ) ≤ 2P∞ i=N +1 (miθ). But from the assumption on (mi) we deduce that Xi=N +1 N . Since P∞ (mN θ)/3i = (mN θ)/2, i=N (λi − λ′ (miθ) < ∞ Xi=1 ∞ which is a contradiction. QED The above lemma, in particular, implies that Φ is injective. Hence we conclude that Φ(P ) is a compact subset of T and Φ is a homeomorphism of P onto Φ(P ). Let aN denote the sum of (miθ) < 0 with i > N and bN the sum of (miθ) > 0 with i > N. (It follows i=N +1 (miθ) < (mN θ)/2.) Since Φ(P ) is from the above calculation that bN − aN = P∞ contained in the intersection of the decreasing sequence XS⊂{1,2,...,N }Xi∈S (miθ) + [aN , bN ] of closed subsets of T (as each closed intervalPi∈S(miθ)+[aN , bN ] shrinks into two disjoint intervals Pi∈S(miθ) + [aN +1, bN +1] and its translate by (mN +1θ) at the next stage and bN − aN = (mN +1θ) + bN +1 − aN +1 > 3(bN +1 − aN +1)), Φ(P ) is a Cantor set. Thus it follows that Φ(P ) is a null set with respect to the Lebesgue measure. Let ν0 denote the product measure on P given byQ∞ i=1{ai, 1−ai} for some sequence (ai) with 0 < ai < 1 such that sup ai < 1 and inf ai > 0. Note that ν0 is a non-atomic measure (even when restricted to the subfield generated by the co-ordinates xn1, xn2, . . . for any subsequence (ni)). We define a probability measure ν′ 0 : A 7→ ν0Φ−1(A∩ Φ(P )) and then a probability measure ν on T as follows: 0 on T by ν′ ∞ ν(A) = γ1+kν′ 0σk(A), Xk=−∞ where γ = √2 − 1. Let B = Sk σk(Φ(P )), an Fσ subset of T. Then ν(B) = 1 and B has Lesbegue measure 0. Since ν′ 0 is non-atomic, so is ν. Thus ν is a non-atomic measure singular from the Lebesgue measure, i.e., ν is singular continuous. Since γν(A) ≤ νσ(A) ≤ γ−1ν(A) for all Borel sets A we conclude that ν is σ-quasi-invariant. Note that ν is not σ-invariant and is not equivalent to a σ-invariant probability measure. (If it is σ-invariant and φ is the state on C(T) defined by ν and Y ∈ C(T) is defined by Y (t) = e2πit, then one can show that φ(Y k) = φ(N −1PN −1 k=0 αk(Y )), which is valid for all N yielding φ(Y k) = 0 for k 6= 0 as α(Y ) = Y σ−1 = e−2πiθY . Thus it follows that φ(f ) =R f dt for f ∈ C(T), a contradiction.) Lemma 5.2 Let t ∈ R. Then Φ(P ) ∩ (Φ(P ) + t) is a null set with respect to ν′ only if t does not belong to {Pi λi(miθ) λi = −1, 0, +1} modulo Z (where λi = 0 except for a finite number of i). 0 if and 21 Proof. If Φ(P ) ∩ (Φ(P ) + t) 6= ∅ then there are x, y ∈ P such that Φ(x) = Φ(y) + t, or t = Pi(xi − yi)(miθ) + k for some k ∈ Z. From the previous lemma it follows that t − k determines λi = xi − yi ∈ {−1, 0, 1}. If λi = 0 except for a finite number of i then t =Pi λimiθ modulo Z and Φ(P ) ∩ (Φ(P ) + t) is equal to the set of Φ(x) satisfying xi = 1 if λi = 1, xi = 0 if λi = −1, and xi is arbitrary if λi = 0, which implies that Φ(P ) ∩ (Φ(P ) + t) has positive measure. Otherwise Φ(P )∩ (Φ(P ) + t) is a null set. QED 0σk(Φ(P )) is non-zero. 0σkΦ(P ) is absolutely continuous with respect to ν′ Lemma 5.3 νσk(Φ(P )) is mutually absolutely continuous with respect to ν′ for all k ∈ Z. Proof. We may suppose that k = 1 (as ν is a kind of average of ν0σk over k). It is obvious that ν′ 0Φ(P ) is absolutely continuous with respect to νΦ(P ). What we have to show is that ν′ 0Φ(P ) for any k. It follows 0σkΦ(P ) from the previous lemma that k must be Pi λimi for some (λi) in order that ν′ In this case Φ(P ) ∩ σk(Φ(P )) is the cylinder set of P determined by xi = 1 for i ∈ I1 = {i λi = 1} and xi = 0 for i ∈ I0 = {i λi = −1}. (Here and henceforth we 0 with ν0.) The inverse image of Φ(P ) ∩ σk(Φ(P )) under σk identify Φ(P ) with P and ν′ is the cylinder set Q of P determined by xi = 0 for i ∈ I1 and xi = 1 for i ∈ I0. Note (1 − ai) and ν0σk(Q) = Qi∈I1 that ν0(Q) = Qi∈I1 ai. Then the definition of ν0 implies ν0σkQ = ckν0Q where ck = Yi∈I1 ai/(1 − ai). (1 − ai)/ai ·Yi∈I0 ai ·Qi∈I0 (1 − ai) ·Qi∈I0 Hence we conclude that ν0σkΦ(P ) ≤ ckν0Φ(P ). QED Proposition 5.4 Let ν be a probability measure on T constructed from ν0 on P and Φ : P → T as above. Then ν is an ergodic σ-quasi-invariant singular continuous measure on T. Proof. What remains to show is that ν is ergodic with respect to σ. Suppose that A is a σ-invariant measurable subset of T with 0 < ν(A). Let A0 = A ∩ Φ(P ), which has positive measure since σk(A0) = A∩σk(Φ(P )) and A =Sk A∩σk(Φ(P )) (modulo null sets). We regard A0 as a measurable subset of P . If x ∈ A0 and y ∈ P satisfies xi = yi for all large i then y ∈ A0 (by Lemma 5.2), i.e., A0, as a measurable subset of P , does not depend on the first N-coordinates for any N ∈ N. This implies that ν0(A0 ∩ C) = ν0(A0)ν0(C) for any cylinder set C of P and hence for any measurable set C. Thus we conclude that ν0(A0) = ν0(A0)2, i.e., ν0(A0) = 1 or ν(Φ(P ) \ A) = 0, which implies that ν(A) = 1. Hence ν is ergodic. QED Let ai = a for all i and denote by ν0a the corresponding probability measure ν0 on P . If a, b ∈ (0, 1) are different then ν0a and ν0b are mutually singular. By using the same 22 Φ : P → T we construct a probability measure νa on T from ν0a. They are all ergodic σ-quasi-invariant singular continuous probability measures on T. Corollary 5.5 The above νa, 0 < a < 1 are mutually singular. Proof. Let a, b ∈ (0, 1) with a 6= b. Since νaP ≃ ν0aP and νbP ≃ ν0bP and ν0a and ν0b are mutually singular we deduce that νaP and νbP are mutually singular; in particular νaP 6= νbP . Since νa and νb are both ergodic, we conclude that νa and νb are mutually singular. QED Let Σ = (T, σ) and let ν be an ergodic σ-quasi-invariant probability measure on T. Then π(ν,C,1) is a simplest kind of irreducible representations of C ∗(Σ). Let V denote the unitary on L2(T, ν) defined by (V ξ)(x) = ξσ−1(x)(dνσ−1/dν)(x)1/2. Then the spectrum of V is T and there is a probability measure ν1 on T such that the isomorphism Y k 7→ V k, k ∈ Z of C(T) into B(L2(T, ν)) extends to the one of L∞(T, ν1) into B(L2(T, ν)), where Y : x 7→ e2πix. Since V Y V ∗ = e−2πiθY or Y V Y ∗ = e2πiθV , we deduce that ν1 is quasi-invariant under σ. Let V1 denote the unitary on L2(T, ν1) defined by (V1ξ)(x) = ξσ(x)(dν1σ/dν)(x)1/2 (where we have used σ instead of σ−1). Then by Proposition 2.2 we conclude that L2(T, ν) ∼= L2(T, ν1) ⊗ H for some Hilbert space H where V (resp. Y ) corresponds to Y ⊗ 1 (resp. W (V1 ⊗ 1)) for some unitary W in L∞(T, ν1) ⊗ B(H). That is, exchanging the roles of Y and V we deduce that π(ν,C,1) is equivalent to π(ν1,H,W ), an irreducible representation for (T, σ−1). Suppose that ν is the Lebesgue measure; in this case V has a complete set of eigen- vectors. Then ν1 must be atomic and ergodic. Then by Proposition 2.5 we obtain H ∼= C If dim(H) > 1 then W (V1 ⊗ 1) has no eigenvalues by Proposition 2.3, which contradicts that Y is diagonal. Thus H = C and hence L2(T, ν) ∼= L2(T, ν1). Hence W V1 has an eigenvector, say W V1ξ = λξ for a unit vector ξ ∈ L2(T, ν1) and a complex number λ of modulus 1. Then it follows that Suppose that ν is atomic, i.e., Y has a complete set of eigenvectors. and can assume that W = 1. The converse also follows. W (x)ξσ(x)( (x))1/2 = λξ(x). dν1σ dν1 Hence we deduce that ξ(x)2dν1(x) is a σ-invariant probability measure, which must be the Lebesgue measure on T. (In this case V1 is diagonal and hence W must be a constant.) Thus we have: Proposition 5.6 In the above situation ν is mutually absolutely continuous with respect to the Lebesgue measure if and only if ν1 is an ergodic σ-quasi-invariant atomic probability measure. Furthermore ν is atomic if and only if ν1 is mutually absolutely continuous with respect to the Lebesgue measure and H = C. In these cases L2(T, ν) ∼= L2(T, ν1). We do not know if the case dim(H) > 1 can actually occur when we start from π = π(ν,C,1) or if V ′′ can fail to be maximal abelian (when π(Y )′′ is maximal abelian). If ν is singular continuous then ν1 is either singular continuous or mutually absolutely continuous with respect to the Lebesgue measure with dim(H) > 1. 23 References [1] R.V. Kadison, Irreducible operator algebras, Proc. Nat. Acad. Sci. U. S. A. 43 (1957), 273 -- 276. [2] M. de Jeu, C. Svensson, and J. Tomiyama, On the Banach ∗-algebra crossed product associated with a toplogical dynamical system, J. Funct. Anal. 262 (2012), 4746 -- 4765. [3] M. de Jeu and J. Tomiyama, Maximal abelian subalgebras and projections in two Banach algebras associated with a topological dynamical system, Studia Math. 208 (2012), 47 -- 75. [4] M. de Jeu and J. Tomiyama, Noncommutative spectral sysnthesis for the involutive Banach algebra associated with a topological dynamical system, Banach J. of Math. Anal. 7 (2013), 103 -- 135. [5] M. de Jeu and J. Tomiyama, Algebraically irreducible representations and structure space of the Banach algebra associated with a topological dynamical system, preprint. [6] G.K. Pedersen, C∗-algebras and their automorphism groups, Academic Press, 1979. [7] S. Sakai, C∗-algebras and W∗-algebras, Springer, 1971. [8] J. Tomiyama, Invitation to C∗-algebras and topological dynamics, World Scientific, 1987. [9] J. Tomiyama, The interplay between topological dynamics and theory of C∗-algebras, Lec- ture Note 2, Res. Inst. Math., Seoul 1992. [10] J. Tomiyama and M. Cho, Note on the structure of non-commutative ℓ1-algebras associated with topological dynamical system, preprint. 24
1502.02093
1
1502
2015-02-07T03:48:25
C*-algebras generated by multiplication operators and composition operators with rational symbol
[ "math.OA", "math.FA" ]
Let $R$ be a rational function of degree at least two, let $J_R$ be the Julia set of $R$ and let $\mu^L$ be the Lyubich measure of $R$. We study the C$^*$-algebra $\mathcal{MC}_R$ generated by all multiplication operators by continuous functions in $C(J_R)$ and the composition operator $C_R$ induced by $R$ on $L^2(J_R, \mu^L)$. We show that the C$^*$-algebra $\mathcal{MC}_R$ is isomorphic to the C$^*$-algebra $\mathcal{O}_R (J_R)$ associated with the complex dynamical system $\{R^{\circ n} \}_{n=1} ^\infty$.
math.OA
math
C∗-ALGEBRAS GENERATED BY MULTIPLICATION OPERATORS AND COMPOSITION OPERATORS WITH RATIONAL SYMBOL HIROYASU HAMADA Abstract. Let R be a rational function of degree at least two, let JR be the Julia set of R and let µL be the Lyubich measure of R. We study the C∗- algebra MCR generated by all multiplication operators by continuous functions in C(JR) and the composition operator CR induced by R on L2(JR, µL). We show that the C∗-algebra MCR is isomorphic to the C∗-algebra OR(JR) associated with the complex dynamical system {R◦n}∞ n=1. 1. Introduction Let D be the open unit disk in the complex plane and H 2(D) the Hardy space of analytic functions whose power series have square-summable coefficients. For an analytic self-map ϕ on the unit disk D, the composition operator Cϕ on the Hardy space H 2(D) is defined by Cϕg = g ◦ ϕ for g ∈ H 2(D). Let T be the unit circle in the complex plane and L2(T) the square integrable measurable functions on T with respect to the normalized Lebesgue measure. The Hardy space H 2(D) can be identified as the closed subspace of L2(T) consisting of the functions whose negative Fourier coefficients vanish. Let PH 2 be the projection from L2(T) onto the Hardy space H 2(D). For a ∈ L∞(T), the Toeplitz operator Ta on the Hardy space H 2(D) is defined by Taf = PH 2 af for f ∈ H 2(D). Recently several authors considered C∗-algebras generated by composition operators (and Toeplitz operators). Most of their studies have focused on composition operators induced by linear fractional maps ([6, 7, 13, 14, 15, 18, 20, 21, 22]). There are some studies about C∗-algebras generated by composition operators and Toeplitz operators for finite Blaschke products. Finite Blaschke products are examples of rational functions. For an analytic self-map ϕ on the unit disk D, we denote by T Cϕ the Toeplitz-composition C∗-algebra generated by both the compo- sition operator Cϕ and the Toeplitz operator Tz. Its quotient algebra by the ideal K of the compact operators is denoted by OCϕ. Let R be a finite Blaschke product of degree at least two with R(0) = 0. Watatani and the author [5] proved that the quotient algebra OCR is isomorphic to the C∗-algebra OR(JR) associated with the complex dynamical system introduced in [11]. In [4] we extend this result for general finite Blaschke products. Let R be a finite Blaschke product R of degree at least two. We showed that the quotient algebra OCR is isomorphic to a certain Cuntz-Pimsner algebra and there is a relation between the quotient algebra OCR 2010 Mathematics Subject Classification. Primary 46L55, 47B33; Secondary 37F10, 46L08. Key words and phrases. composition operator, multiplication operator, Frobenius-Perron op- erator, C∗-algebra, complex dynamical system. 1 2 HIROYASU HAMADA and the C∗-algebra OR(JR). slightly different. In general, two C∗-algebras OCR and OR(JR) are In this paper we give a relation between a C∗-algebra containing a composition operator and the C∗-algebra OR(JR) for a general rational function R of degree at least two. In the above studies we deal with composition operators on the Hardy space H 2(D), while we consider composition operators on L2 spaces in this case. Composition operators on L2 spaces has been studied by many authors (see for example [23]). Let (Ω, F , µ) be a measure space and let ϕ a non-singular transfor- mation on Ω. We define a measurable function by Cϕf = f ◦ ϕ for f ∈ L2(Ω, F , µ). If Cϕ is bounded operator on L2(Ω, F , µ), we call Cϕ the composition operator with ϕ. Let R be a rational function of degree at least two. We consider the Julia set JR of R, the Borel σ-algebra B(JR) on JR and the Lyubich measure µL of R. Let us denote by MCR the C∗-algebra generated by multiplication operators Ma for a ∈ C(JR) and the composition operator CR on L2(JR, B(JR), µL). We regard the C∗-algebra MCR and multiplication operators as replacements of Toeplitz-composition C∗- algbras and Toeplitz operators, respectively. We prove that the C∗-algebra MCR is isomorphic to the C∗-algebra OR(JR) associated with the complex dynamical system. There are two important points to prove this theorem. First one is to analyze operators of the form C ∗ RMaCR for a ∈ C(JR). We now consider a more general case. Let (Ω, F , µ) be a finite measure space and ϕ is a non-singular transformation. If Cϕ is bounded, then we have C ∗ ϕMaCϕ = MLϕ(a) for a ∈ L∞(Ω, F , µ), where Lϕ is the Frobenius-Perron operator for ϕ. This is an extension of covariant relations considered by Exel and Vershik [2]. Moreover similar relations have been studied on the Hardy space H 2(D). Let ϕ be an inner function on D. Jury showed a covariant ϕTaCϕ = TAϕ(a) for a ∈ L∞(T), where Aϕ is the Aleksandrov operator. relation C ∗ Second important point is an anaysis based on bases of Hilbert bimodules. In [4] and [5], a Toeplitz-composition C∗-algebra for a finite Blaschke product R is isomorphic to a certain Cuntz-Pimsner algebra of a Hilbert bimodule XR, using a finite basis of XR. Let R be a rational function of degree at least two. The C∗- algebra OR(JR) associated with complex dynamical system is defined as a Cuntz- Pimsner algebra of a Hilbert bimodule Y . Unlike the cases of [4] and [5], the Hilbert bimodule Y does not always have a finite basis. Kajiwara [9], however, constructed a concrete countable basis of Y . Thanks to this basis, we can prove the desired theorem. 2. Covariant relations Let (Ω, F , µ) be a measure space and let ϕ : Ω → Ω be a measurable transfor- mation. Set ϕ∗µ(E) = µ(ϕ−1(E)) for E ∈ F . Then ϕ∗µ is a measure on Ω. The measurable transformation ϕ : Ω → Ω is said to be non-singular if ϕ∗µ(E) = 0 whenever µ(E) = 0 for E ∈ F . If ϕ is non-singular, then ϕ∗µ is absolutely contin- uous with respect to µ. When µ is σ-finite, we denote by hϕ the Radon-Nikodym derivative dϕ∗µ dµ . Let 1 ≤ p ≤ ∞. We shall define the composition operator on Lp(Ω, F , µ). Every non-singular transformation ϕ : Ω → Ω induces a linear operator Cϕ from LF p(Ω, F , µ) to the linear space of all measurable functions on (Ω, F , µ) defined as Cϕf = f ◦ ϕ for f ∈ Lp(Ω, F , µ). If Cϕ : Lp(Ω, F , µ) → Lp(Ω, F , µ) is bounded, C∗-ALGEBRAS GENERATED BY COMPOSITION OPERATORS 3 it is called a composition operator on Lp(Ω, F , µ) induced by ϕ. Let (Ω, F , µ) be σ-finite. For 1 ≤ p < ∞, Cϕ is bounded on Lp(Ω, F , µ) if and only if the Radon- Nikodym derivative hϕ is bounded (see for example [23, Theorem 2.1.1]). If Cϕ is bounded on Lp(Ω, F , µ) for some 1 ≤ p < ∞, then Cϕ is bounded on Lp(Ω, F , µ) for any 1 ≤ p < ∞ since hϕ is independent of p. For p = ∞, Cϕ is bounded on L∞(Ω, F , µ) for any non-singular transformation. Definition. Let (Ω, F , µ) be a σ-finite measure space, let ϕ : Ω → Ω be a non- singular transformation and let f ∈ L1(Ω, F , µ). We define νϕ,f by νϕ,f (E) =Zϕ−1(E) f dµ for E ∈ F . Then νϕ,f is an absolutely continuous measure with respect to µ. By the Radon-Nikodym theorem, there exists Lϕ(f ) ∈ L1(Ω, F , µ) such that ZE Lϕ(f )dµ =Zϕ−1(E) f dµ for E ∈ F . We can regard Lϕ as a bounded operator on L1(Ω, F , µ) (see for example [16, Proposition 3.1.1]). We call Lϕ the Frobenius-Perron operator. Lemma 2.1. Let (Ω, F , µ) be a finite measure space and let ϕ : Ω → Ω be a non- singular transformation. Suppose that Cϕ : L1(Ω, F , µ) → L1(Ω, F , µ) is bounded. Then the restriction LϕL∞(Ω,F ,µ) is a bounded operator on L∞(Ω, F , µ) and C ∗ ϕ = LϕL∞(Ω,F ,µ). Proof. Let f ∈ L∞(Ω, F , µ). First we shall show Lϕ(f ) ∈ L∞(Ω, F , µ). There exists M > 0 such that f ≤ M . It follows from [16, Proposition 3.1.1] that Lϕ(f ) ≤ Lϕ(f ) ≤ M Lϕ(1). Since Lϕ(1) = hϕ and Cϕ is bounded on L1(Ω, F , µ), we have Lϕ(1) ∈ L∞(Ω, F , µ). Hence Lϕ(f ) ∈ L∞(Ω, F , µ). By the definition of Lϕ, we have ZΩ χELϕ(f )dµ =ZΩ χϕ−1(E)f dµ =ZΩ (CϕχE)f dµ for E ∈ F , where χE and χϕ−1(E) are characteristic functions on E and ϕ−1(E) respectively. Since Cϕ is bounded on L1(Ω, F , µ) and the set of integrable simple functions is dense in L1(Ω, F , µ), the restriction map LϕL∞(Ω,F ,µ) is bounded on L∞(Ω, F , µ) and C ∗ (cid:3) ϕ = LϕL∞(Ω,F ,µ). Let (Ω, F , µ) be a finite measure space and ϕ : Ω → Ω a non-singular transfor- mation. We consider the restriction of Lϕ to L∞(Ω, F , µ). From now on, we use the same notation Lϕ if no confusion can arise. For a ∈ L∞(Ω, F , µ), we define the multiplication operator Ma on L2(Ω, F , µ) by Maf = af for f ∈ L2(Ω, F , µ). We show the following covariant relation. Proposition 2.2. Let (Ω, F , µ) be a finite measure space and let ϕ : Ω → Ω be a non-singular transformation. If Cϕ : L2(Ω, F , µ) → L2(Ω, F , µ) is bounded, then we have for a ∈ L∞(Ω, F , µ). C ∗ ϕMaCϕ = MLϕ(a) 4 HIROYASU HAMADA Proof. For f, g ∈ L2(Ω, F , µ), we have hC ∗ ϕMaCϕf, gi = hMaCϕf, Cϕgi =ZΩ aCϕ(f g)dµ =ZΩ =ZΩ = hMLϕ(a)f, gi a(f ◦ ϕ)(g ◦ ϕ)dµ Lϕ(a)f gdµ by Lemma 2.1, where Cϕ is also regarded as the composition operator on L1(Ω, F , µ). (cid:3) 3. C∗-algebras associated with complex dynamical systems of X is called a countable basis of X if ξ = P∞ We recall the construction of Cuntz-Pimsner algebras [19] (see also [12]). Let A be a C∗-algebra and let X be a right Hilbert A-module. A sequence {ui}∞ i=1 i=1 uihui, ξiA for ξ ∈ X, where the right hand side converges in norm. We denote by L(X) the C∗-algebra of the adjointable bounded operators on X. For ξ, η ∈ X, the operator θξ,η is defined by θξ,η(ζ) = ξhη, ζiA for ζ ∈ X. The closure of the linear span of these operators is denoted by K(X). We say that X is a Hilbert bimodule (or C∗-correspondence) over A if X is a right Hilbert A-module with a ∗-homomorphism φ : A → L(X). We always assume that φ is injective. A representation of the Hilbert bimodule X over A on a C∗-algebra D is a pair (ρ, V ) constituted by a ∗-homomorphism ρ : A → D and a linear map V : X → D satisfying ρ(a)Vξ = Vφ(a)ξ, V ∗ ξ Vη = ρ(hξ, ηiA) for a ∈ A and ξ, η ∈ X. It is known that Vξρ(b) = Vξb follows automatically (see for example [12]). We define a ∗-homomorphism ψV : K(X) → D by ψV (θξ,η) = VξV ∗ η for ξ, η ∈ X (see for example [10, Lemma 2.2]). A representation (ρ, V ) is said to be covariant if ρ(a) = ψV (φ(a)) for all a ∈ J(X) := φ−1(K(X)). Suppose the Hilbert bimodule X has a countable basis {ui}∞ i=1 and (ρ, V ) is a representation of ui − ρ(a)k → 0 as n → ∞ i=1 ρ(a)Vui V ∗ n=1 is an approximate unit for K(X). X. Then (ρ, V ) is covariant if and only if kPn for a ∈ J(X), since {Pn i=1 θui,ui }∞ Let (i, S) be the representation of X which is universal for all covariant represen- tations. The Cuntz-Pimsner algebra OX is the C∗-algebra generated by i(a) with a ∈ A and Sξ with ξ ∈ X. We note that i is known to be injective [19] (see also [12, Proposition 4.11]). We usually identify i(a) with a in A. Let R be a rational function of degree at least two. We recall the definition of the C∗-algebra OR(JR). Since the Julia set JR is completely invariant under R, that is, R(JR) = JR = R−1(JR), we can consider the restriction RJR : JR → JR. Let A = C(JR) and Y = C(graph RJR), where graph RJR = {(z, w) ∈ JR × JR w = R(z)} is the graph of RJR. We denote by eR(z) the branch index of R at z. Then Y is an A-A bimodule over A by (a · f · b)(z, w) = a(z)f (z, w)b(w), a, b ∈ A, f ∈ Y. We define an A-valued inner product h , iA on Y by hf, giA(w) = Xz∈R−1(w) eR(z)f (z, w)g(z, w), f, g ∈ Y, w ∈ JR. C∗-ALGEBRAS GENERATED BY COMPOSITION OPERATORS 5 Then Y is a Hilbert bimodule over A. The C∗-algebra OR(JR) is defined as the Cuntz-Pimsner algebra of the Hilbert bimodule Y = C(graph RJR) over A = C(JR). 4. Main theorem Let R be a rational function. We define the backward orbit O−(w) of w ∈ C by O−(w) = {z ∈ C R◦m(z) = w for some non-negative integer m}. A point w in C is an exceptional point for R if the backward orbit O−(w) of w is finite. We denote by ER the set of exceptional points. Definition (Freire-Lopes-Man´e [3], Lyubich [17]). Let R be a rational function and n = degR. Let δz be the Dirac measure at z ∈ C. For w ∈ C r ER and m ∈ N, we define a probability measure µw m on C by µw m = eR◦m(z)δz. 1 nm Xz∈(R◦m)−1(w) The sequence {µw m=1 converges weakly to a probability measure µL, which is called the Lyubich measure of R. The measure µL is independent of the choice of w ∈ C r ER. m}∞ Let R be a rational function of degree at least two. We will denote by B(JR) the Borel σ-algebra on the Julia set JR. In this section we consider the finite measure space (JR, B(JR), µL). It is known that the support of the Lyubich measure µL is the Julia set JR. Moreover the Lyubich measure µL is regular on the Julia set JR and a invariant measure with respect to R, that is, µL(E) = µL(R−1(E)) for E ∈ B(JR). Thus the composition operator CR on L2(JR, B(JR), µL) is an isometry. Definition. For a rational function R of degree at least two, we denote by MCR the C∗-algebra generated by all multiplication operators by continuous functions in C(JR) and the composition operator CR on L2(JR, B(JR), µL). Let a rational function R of degree at least two. In this section we shall show that the C∗-algebra MCR is isomorphic to the C∗-algebra OR(JR). First we give a concrete expression of the restriction of LR to C(JR). This result immediately follows from [17] and Lemma 2.1. Proposition 4.1 (Lyubich [17, Lemma, p.366]). Let R be a rational function of degree n at least two. Then LR : C(JR) → C(JR) and (LR(a))(w) = 1 n Xz∈R−1(w) eR(z)a(z), w ∈ JR for a ∈ C(JR). Let X = C(JR) and n = deg R. Then X is an A-A bimodule over A by (a · ξ · b)(z) = a(z)ξ(z)b(R(z)) a, b ∈ A, ξ ∈ X. We define an A-valued inner product h , iA on X by hξ, ηiA(w) = 1 n Xz∈R−1(w) eR(z)ξ(z)η(z) (cid:0) = (LR(ξη))(w)(cid:1) , ξ, η ∈ X. 6 HIROYASU HAMADA Then X is a Hilbert bimodule over A. Put kξk2 = khξ, ξiAk1/2 ∞ for ξ ∈ X, where k k∞ is the sup norm on JR. It is easy to see that X is isomorphic to Y as Hilbert bimodules over A. Hence the C∗-algebra OR(JR) is isomorphic to the Cuntz-Pimsner algebra OX constructed from X. We need some analyses based on bases of the Hilbert bimodule X to show an equation containing the composition operator CR and multiplication operators. Lemma 4.2. Let u1, . . . , uN ∈ X. Then N Xi=1 for a ∈ A. MuiCRC ∗ RM ∗ uia = ui · hui, aiA N Xi=1 Proof. Since a = MaCR1, we have MuiCRC ∗ RM ∗ uia = N Xi=1 = = = = = which completes the proof. N N N N Xi=1 Xi=1 Xi=1 Xi=1 Xi=1 Xi=1 N N MuiCRC ∗ RM ∗ ui MaCR1 MuiCRC ∗ RMuiaCR1 MuiCRMLR(uia)1 by Proposition 2.2 MuiMLR(uia)◦RCR1 uiLR(uia) ◦ R ui · hui, aiA, (cid:3) Lemma 4.3. Let {ui}∞ i=1 be a countable basis of X. Then N MuiCRC ∗ RM ∗ ui ≤ I. 0 ≤ Xi=1 ui . It is clear that TN is a positive operator. RM ∗ We shall show TN ≤ I. By Lemma 4.2, i=1 MuiCRC ∗ Proof. Set TN := PN hTN f, f i =ZJR PN and k k∞ are equivalent (see the proof of [11, Proposition 2.2]), PN i=1 is a countable basis of X, for f ∈ C(JR), we have i=1 ui · hui, f iA → f with respect to k k2 as N → ∞. Since the two norms k k2 i=1 ui · hui, f iA (TN f )(z)f (z)dµL(z) =ZJR N Xi=1 ui · hui, f iA! (z)f (z)dµL(z) converges to f with respect to k k∞. Thus for f ∈ C(JR). Since {ui}∞ f (z)f (z)dµL(z) = hf, f i as N → ∞ hTN f, f i →ZJR C∗-ALGEBRAS GENERATED BY COMPOSITION OPERATORS 7 for f ∈ C(JR). Therefore hTN f, f i ≤ hf, f i for f ∈ C(JR). Since the Lyubich measure µL on the Julia set JR is regular, C(JR) is dense in L2(JR, B(JR), µL). Hence we have TN ≤ I. This completes the proof. (cid:3) Let B(R) be the set of branched points of a rational function R. We now recall a description of the ideal J(X) of A. By [11, Proposition 2.5], we can write J(X) = {a ∈ A a vanishes on B(R)}. We define a subset J(X)0 of J(X) by J(X)0 = {a ∈ A a vanishes on B(R) and has compact support on JR r B(R)}. Since B(R) is a finite set ([1, Corollary 2.7.2]), J(X)0 is dense in J(X). Lemma 4.4. There exists a countable basis {ui}∞ i=1 of X such that ∞ Xi=1 for a ∈ J(X). MaMui CRC ∗ RM ∗ ui = Ma Proof. By [9, Subsection 3.1], there exists a countable basis {ui}∞ i=1 of X satisfying the following property. For any b ∈ J(X)0, there exists M > 0 such that supp b ∩ supp um = ∅ for m ≥ M . Since J(X)0 is dense in J(X), for any a ∈ A and any ε > 0, there exists b ∈ J(X)0 such that ka − bk < ε/2. Let m ≥ M . Then by Lemma 4.2 and bui = 0 for i ≥ m, it follows that m m ∞ MbMui CRC ∗ RM ∗ ui f = bui · hui, f iA = bui · hui, f iA = bf = Mbf Xi=1 Xi=1 for f ∈ C(JR). Since C(JR) is dense in L2(JR, B(JR), µL), we have MbMuiCRC ∗ RM ∗ ui = Mb. From Lemma 4.3 it follows that MaMui CRC ∗ RM ∗ MaMuiCRC ∗ RM ∗ ui − MbMui CRC ∗ RM ∗ ui(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) m m ≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xi=1 +(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xi=1 ≤ kMa − Mbk (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xi=1 = ε, + < m ε 2 ε 2 m Xi=1 ui − Mb(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ui(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) RM ∗ MbMuiCRC ∗ RM ∗ + kMb − Mak MuiCRC ∗ + kMa − Mbk Xi=1 m (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xi=1 m Xi=1 ui − Ma(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) which completes the proof. (cid:3) The following theorem is the main result of the paper. Theorem 4.5. Let R be a rational function of degree at least two. Then MCR is isomorphic to OR(JR). Proof. Put ρ(a) = Ma and Vξ = MξCR for a ∈ A and ξ ∈ X. Then we have ρ(a)Vξ = MaMξCR = MaξCR = Va·ξ and V ∗ ξ Vη = C ∗ RM ∗ ξ MηCR = C ∗ RMξηCR = MLR(ξη) = ρ(LR(ξη)) = ρ(hξ, ηiA) 8 HIROYASU HAMADA for a ∈ A and ξ, η ∈ X by Proposition 2.2. Let {ui}∞ Then, applying Lemma 4.4, i=1 be a countable basis of X. ∞ ∞ ρ(a)Vui V ∗ ui = Xi=1 MaMui CRC ∗ RM ∗ ui = Ma = ρ(a) Xi=1 for a ∈ J(X). Since the support of the Lyubich measure µL is the Julia set JR, the ∗-homomorphism ρ is injective. By the universality and the simplicity of OR(JR) ([11, Theorem 3.8]), the C∗-algebra MCR is isomorphic to OR(JR). (cid:3) Acknowledgement. The author wishes to express his thanks to Professor Hiroyuki Takagi for several helpful comments concerning to composition operators. References [1] A. F. Beardon, Iteration of rational functions, Complex analytic dynamical systems, Grad- uate Texts in Mathematics, 132, Springer-Verlag, New York, 1991. [2] R. Exel and A. Vershik, C∗-algebras of irreversible dynamical systems, Canad. J. Math. 58 (2006), 39 -- 63. [3] A. Freire, A. Lopes and R. Man´e, An invariant measure for rational maps, Bol. Soc. Brasil. Mat. 14 (1983), 45 -- 62. [4] H. Hamada, Quotient algebras of Toeplitz-composition C∗-algebras for finite Blaschke prod- ucts, Complex Anal. Oper. Theory 8 (2014), 843 -- 862. [5] H. Hamada and Y. Watatani, Toeplitz-composition C∗-algebras for certain finite Blaschke products, Proc. Amer. Math. Soc. 138 (2010), 2113 -- 2123. [6] M. T. Jury, The Fredholm index for elements of Toeplitz-composition C∗-algebras, Integral Equations Operator Theory 58 (2007), 341 -- 362. [7] M. T. Jury, C∗-algebras generated by groups of composition operators, Indiana Univ. Math. J. 56 (2007), 3171 -- 3192. [8] M. T. Jury, Completely positive maps induced by composition operators, preprint. [9] T, Kajiwara, Countable bases for Hilbert C∗-modules and classification of KMS states, Op- erator structures and dynamical systems, 73 -- 91, Contemp. Math., 503, Amer. Math. Soc., Providence, RI, 2009. [10] T. Kajiwara, C. Pinzari and Y. Watatani, Ideal structure and simplicity of the C∗-algebras generated by Hilbert bimodules, J. Funct. Anal. 159 (1998), 295 -- 322. [11] T. Kajiwara and Y. Watatani, C∗-algebras associated with complex dynamical systems, In- diana Math. J. 54 (2005), 755 -- 778. [12] T. Katsura, On C∗-algebras associated with C∗-correspondences, J. Funct. Anal. 217 (2004), 366 -- 401. [13] T. L. Kriete, B. D. MacCluer and J. L. Moorhouse, Toeplitz-composition C∗-algebras, J. Operator Theory 58 (2007), 135 -- 156. [14] T. L. Kriete, B. D. MacCluer and J. L. Moorhouse, Spectral theory for algebraic combinations of Toeplitz and composition operator, J. Funct. Anal. 257 (2009), 2378 -- 2409. [15] T. L. Kriete, B. D. MacCluer and J. L. Moorhouse, Composition operators within singly generated composition C∗-algebras, Israel J. Math. 179 (2010), 449 -- 477. [16] A. Lasota and M. C. Mackey, Chaos, fractals, and noise, Stochastic aspects of dynamics, Second edition, Applied Mathematical Sciences, 97, Springer-Verlag, New York. [17] M. J. Lyubich, Entropy properties of rational endomorphisms of the Riemann sphere, Ergodic Theory Dynam. Systems 3 (1983), 351 -- 385. [18] E. Park, Toeplitz algebras and extensions of irrational rotation algebras, Canad. Math. Bull. 48 (2005), 607 -- 613. [19] M. V. Pimsner, A class of C∗-algebras generating both Cuntz-Krieger algebras and crossed product by Z , Free Probability Theory, Fields Inst. Commun., Vol 12, Amer. Math. Soc., Providence, RI, pp. 189 -- 212. [20] K. S. Quertermous, A semigroup composition C∗-algebra, J. Operator Theory 67 (2012), 581 -- 604. [21] K. S. Quertermous, Fixed point composition and Toeplitz-composition C∗-algebras, J. Funct. Anal. 265 (2013), 743 -- 764. C∗-ALGEBRAS GENERATED BY COMPOSITION OPERATORS 9 [22] M. K. Sarvestani and M. Amini, The C∗-algebra generated by irreducible Toeplitz and com- position operators, arXiv:1408.1057. [23] R. K. Singh and J. S. Manhas, Composition operators on function spaces, North-Holland Mathematics Studies, 179, North-Holland Publishing Co., Amsterdam, 1993. National Institute of Technology, Sasebo College, Okishin, Sasebo, Nagasaki, 857- 1193, Japan. E-mail address: [email protected]